PLACE N RETURN BOXtomnovoflibchockomnom you: record. To AVOID FINES Mom on o: More duo duo. DfiEE DUE DATE DUE DATE DUE MSU Is An Mmuflvo ActiorVEqnl Oppommy Institution WM i E g E F . LIBRARY Michigan State University This is to certify that the thesis entitled Experimental Observations of Composting Dairy Manure Solids , ' presented by i ' Stephen Earl Ferns has been accepted towards fulfillment of the requirements for Masters degreein Agric. Engr. Dmt_flovember 25, 1987 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution EXPERIMENTAL OBSERVATIONS OF COMPOSTING DAIRY MANURE SOLIDS BY Stephen Earl Ferns A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE in Agricultural Engineering Department of Agricultural Engineering 1987 ABSTRACT EXPERIMENTAL OBSERVATIONS OF COMPOSTING DAIRY MANURE SOLIDS BY Stephen Earl Ferns Dairy manure solids (DMS) were composted in a batch windrow process of 3 weeks duration with weekly turnings. Diffusion or natural convection provided heat and mass transfer. Temperature, gas concentration (02 & C02), ambient conditions, insitu physical properties and pile size were monitored. A temperature and gas probe and a rotary corer developed show promise for future insitu sampling in composting systems of this type. Sensitivity analyses were performed on the experimental methods and suggestions for improvements were made. Results show that the assumption of spatially constant physical and thermal properties used in a previous model of distributed heat and mass transfer in compost windrows is not justified in all cases. Significant changes in physical properties over time were not observed. DMS compost behaved similarly to other compost substrates. However, DMS composted at higher moisture contents and lower maximum temperatures were observed than with other compost substrates. Distinct color zones that _. #_.~—_fi.._ correlated with time-temperature patterns were observed in the windrow cross-sections. Analysis of time-temperature patterns indicated areas where either significant thermal death or regrowth of mastitis-causing coliform organisms could occur. Several methods were discussed for managing the DMS composting process in order to produce acceptable Mam Major Professor a AKA Major Professor Department Chairman dairy bedding material. . _. _—«.__ __. _ To my wife Lori. iv ACKNOWLEDGMENTS Many individuals have contributed to the success of my MS studies and research at MSU. My Co-Major Professors, Dr. John Gerrish and Dr. Ted Loudon provided direction, assistance and support. Dr. Gerrish was an inspiration in his ability to creatively solve research problems. Dr. Loudon provided pactical research advice and was generous in his sharing of knowledge on soil and water engineering. Dr. Fred Bakker-Arkema served on my committee and provided insight into heat and mass transfer processes and vigorously tried to move me along in my program. Dr. Larry Segerlind graciously discussed applying Finite Element modeling to DMS composting and served on my committee. One of my biggest disappointments was not being able to actually use this in the thesis. Dr. John Speicher and Dr. Roger Mellenberger provided invaluable insights into practical dairy management and the nature of the coliform mastitis problem. Bob Gardner, a fellow MS student from the Animal Science Department, offered invaluable moral and technical support: there was at least one other person in the university crazy enough to play in the DMS. The Kellogg Biological Station dairy manager, Rob Ashley and his staff cheerfully put up with many somewhat V unusual requests and created a supportive research environment. Several individuals very generously shared their laboratory space and time. These people included Dr. Mackenzie Davis, of the Environmental Engineering Department; Dr. David Dilley and Dr. Steve Sargent of the Horticultural Department; and Dr. James Tiedje and Peter Groffman of the Crop and Soil Science Department. vi TABLE OF CONTENTS Page List Of TablesO O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O OXiii List Of FigureSO O O O O O O O O O O O O O O O O O O O O O O O O O O O O‘O O O O O O O O OXVi WER 1: IntrOductionO O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O 1 1.1: Criteria for Use of Dairy Manure Solids as a Bedding MaterialOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO 1.2: Effects of Mastitis on the Dairy Industry........ 10201: PhYSical EffeCtSOOO O OOOO OOO O O O OO O O O O O O O O O O 1O 2 O 2: Economic EffeCtSo O O O O O O O O O O O O O O O O O O O O O O O O O 1.3: Micro-organisms Associated with Mastitis......... 1.3.1: General Organisms......................... 10302: COliform organismSOOOOOOOOOOOOOOOOOOOOOOOO 1.4: Treatment of Dairy Manure Solids for Bedding..... 1.4.2: Treatment by Composting................... 1.5 The Composting Process........................... Definition................................ Parameters Affecting Composting........... Spatial Variability....................... Need for Distributed Heat and Mass Transfer Composting Model................. Modeling Composting Heat and Mass Transfer Processes........................ : Recent Advances in Composting. ..... . .... .1 2 2 3 3 3 3 5 1.4.1: General Treatment Methods.................5 5 6 6 7 7 7 8 9 CHAPTER 2: Objectives................................ll CHAPTER 3: ‘Literature ReVieWOOOOOOOOOOOOOOOOOOOOOOOOOlz 3.1: Description of Composting.......................12 vii 3.5: Analogous Systems...............................13 MicrObiOIOQYOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO14 30301 30302 GeneralOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO14 Types of Micro-organisms.................lS l: Bacteria and Actinomycetes........15 2: FungiOOOOOOOOOOOOOOOOOOOOOOOOOOOOO16 Compost Microbial Ecology........ ...... ..l7 . : General...........................l7 . : Temperature.......................18 . : Moisture Content and Water Activity..........................27 : pH................................29 : Redox Potential...................31 : Compost Microbial Ecology Studies...........................31 BiOChemistrYOOOOOOOOOOOOOOOOOOOOOO OOOOOO O OOOOO OO35 3.4.1: 3.4.2: StOiChiometrYOOOOOOOOOOOOO OOOOOOO OOOOOOOO35 compost StUdleSOOOOOOOOOOOOOOOOOOOOOOOOOO36 Organic Substrates................36 Nitrogen Transformations..........38 Effect of Carbon Source...........41 Effect of Inorganic Materials.....4l tthH mpirical StUdieSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO43 UIUIU'IUIU‘ o. coo 4»wa“ oo oo o MDWNH Temperature..............................43 Oxygen and Carbon Dioxide................43 M01sture Content.........................43 Windrow Size.............................47 Windrow Zones............................48 Kinetic Modeling................................49 Microbial Growth.........................49 Microbial Death..........................51 Regrowth................... ..... .........55 Combined Equations.......................56 Physical Properties..... ............ ............58 U'IPU-DNH Particle Size......... ........ ...........58 Particle Density.........................59 Bulk Density.............................59 Derived Quantities.......................59 Compressibility... ................. ......60 viii 3.7.6: Settlement Behavior......................60 3.8: Heat Transfer...................................61 3080l: DifoSionOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO61 3.8.1.1: Thermal Conductivity... ..... ......61 3080102: SpeCifiC HeatOOOOOOOOOOOOOOOOOOOOO66 3.8.1.3: Thermal Diffusivity...............67 3.8.2: Natural convectionOOOOOOOOOOOOOOOOOOOOOOO67 309: Mass TranSferOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO67 3.9.1: Windrow Void Space Transfer...... ...... ..68 3.9.1.1: DifoSionOOOOOOOOOOOOOOOOOOOOOOOOO68 3 O 9 O 1 O 2: Evaporation. O O O O O O O O O O O O O O O O O O O O O O 69 3.9.1.3: LiquidSOOOOOOOOOOOOOOOOOOOOOOOOOOO70 3.9.2: Transport in Liquid Films................71 3.10: Complete Heat and Mass Transfer Models.........73 CHAPTER 4: Experimental Methods......................75 4.1: Data Requirements...............................75 4.2: General Experimental Design and Methods.........75 4.3: Probe Design and Placement................ ..... .78 4.4: Temperature Measurement.........................83 4.4.1: Equipment and Method.....................83 4.4.2: Error analysis and calibration...........85 4.5: Gas Sampling and Analysis.......................87 : Equipment and Method.....................87 : Error AnalYSiSOOOOOOOOOOOOOOOOOOOOOOOOOOO90 4.5.2.1: Effect of Air Drawn into Sample from Outside of Sampling Area.....91 4.5.2.2: Effect of Withdrawing Successive Samples on Changes in Gas Concentration....... .............. 94 4.5.2.3: Effect of Leaks into Gas Sampling System........ ........... 96 4.5.2.4: Effect of Diffusion and/or Gas Flow into Syringe and Gas Storage Containers.......... ...... 97 4.5.2.5: Gas Chromatograph Calibration ix 4.6: 4.10: 4.11: CHAPTER 5: 5.1: 5.2: 5.3: 5.4: 5.5: 5.6: 5.7: 5.8: 5.9: and ErrorSOOOOOOOOOOOOOOOOOOOOOOO104 Core Sampling Method...........................106 Review of Sampling Methods and Corer DeSignOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO106 Comparison of Core Sampling Methods.....109 Description of Experimental Core Sampling Method.........................112 Moisture Content and Volatile Solids...........115 Methods and Ca1cu1ations................115 Sensitivity Analysis....................118 4.7.1: 4.7.2: 31.11]!DenSitY.....o.............................122 calcu1ationSOOOOOOOOOOOOOOOOOOOOOOOOOOOO122 senSitiVity AnalYSiSOOOOOOOOOOOOOOOOOOOO122 4.8.1: 4.8.2: Porosity and Free Air Space....................127 calCUIationSOOOOOOOOOOOOOOOOOOOOOOOOOOOO127 Sensitivity Analysis....................129 4.9.1: 4.9.2: Additional Physical Properties................135 Windrow Size ChangeSOOOOOOOOOOOOOOOOOOOOOOOOOO138 Experimental Results..... ......139 OOOOOOOOOOOOOOOOO139 General Observations.. Windrow Size and ShapeOOOOOOOOOOOOOOOOOOOOOOOOO139 TemperatureOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO146 Temperature Variability.................146 Temperature Patterns....................149 Time-Temperature Relationships..........157 UIU'IU'I .3.1: .3.2: .3.3: Gas concentrationSOOOOOOOOOOOOOOOOOOOOOOOOOOOOO157 Moisture Content................... ............ 173 VOIatile SOlidSOOOOOOOO OOOO17'7 Wet Bulk Density ....... ........ POtOSitYooooo-ooo Free Air Space......... ..... . .................. 191 5.10: Additional Moisture and Physical Properties...197 5.11: Windrow COlor ChangeSOOOOOOOOOOOOOOOOOOOOOOOOO197 CHAPTER 6: DiSCUSSiOnOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO212 6.1: Experimental Methods...........................212 6.1.1: Probe Design............................212 6.1.2: Temperature.............................212 6.1.3: Gas Concentration.......................213 6.1.4: Gravimetric Moisture Content............215 6.1.5: Volatile Solids.........................215 6.1.6: Dairy Manure Solids Sampling and Bulk Density.................................216 6.1.7: Porosity and Free Air Space.............216 6.1.8: Dairy Manure Solids P1acement...........217 6.1.9: Windrow Size Measurements...............217 6.2: Experimental Results..... ....... ...............217 6.2.1: Temperature.............................217 6.2.2: Gas Concentration.......................218 6.2.3: Volatile Solids.........................223 6.2.4: Moisture Content........................225 6.2.5: Bulk Density............................226 6.2.6: Additional Derived Physical Parameters..227 6.2.7: Thermal Properties......................229 6.2.8: Windrow Size Changes....................232 6.3: Reaction and Transport Characteristics.........232 60301: Type Of ReactionOOOOOOOOOOOOOOOOOOOOOOOO232 6.3.2: Unreacted Core Model....................233 6.3.3: Pore Space Heat and Mass Transfer.......235 6.4: Pathogen surVivalOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO237 6.5: Application of Composting to Dairy Manure Solids for Production of Bedding Material......249 6.6: Preliminary Assessment of Heat and Mass Transfer Models for Composting Dairy Manure Solids....................... ..... . ..... 250 CHAPTER 7: Conclusions...... ......... .. ........ .....252 CHAPTER 8: Directions for Future Research ....... ....255 8.1: Experimental Methods...... ...... . ........ . ..... 255 xi 8.2: Windrow 8.3: Compost APPENDICES Appendix Appendix Appendix Appendix Appendix Appendix Appendix Appendix Compost Processes........... ........... 257 Heat and Mass Transfer Modeling........258 Temperature Teasurement Error Calculations.......................260 Sample Thermocouple Water Bath Calibration Data........ ..... ......264 Effect of Sequential Gas Samples and Location on Carbon Dioxide and Oxygen Concentration...........269 ' Gas Concentration Correction Factor Calculations................270 Volumetric Gas Standards Error Calculations.......................283 Preliminary Bulk Sampler Experiments........................286 Derivation of Error Propagation Equations..........................293 Windrow size, temperature, gas concentration, and physical property data and sample location availablity........................296 ReferenceSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO297 xii Table Table Table Table Table Table Table Table Table Table Table Table Table Table Table 3.1: LIST OF TABLES Growth temperatures for micro- organismSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOlS 910 values for garbage composts.......19 Decimal reduction times for selected micro-organisms...........................23 2 values for selected micro-organisms.....23 Approximate minimum levels of water activity permitting growth of micro- organisms at temperatures near optimal....29 The limits of pH allowing initiation of growth by selected micro-organisms.....30 Confidence levels that all material will obtain a temperature equal to or greater than a particular temperature for a desired number of days..............55 Experimental values of diffusion coefficients in gases at one atmosphereOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO68 Diffusion coefficients at infinite dilution in water at 25°C.............73 Sampling methods and schedule.............76 Dairy characteristics.....................77 Temperature measurement errors. ....... ....86 Water bath calibration of 32 thermocouples................... .......... 87 Potential sources of error in gas sampling, storage and analysis...... ...... 90 Internal volume of sampling tubes and diffusion chamber...... .. ...... 95 xiii Table Table Table Table Table Table Table Table Table Table Table Table Table Table Table Table Table 4.7: 4.8: 4.9: 6.2: 6.3: Gas concentration standards used in gas Chromatograph calibration................105 Comparison of moisture contents expressed on a wet and dry solids baSiSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOll7 Comparison of volatile solids content expressed on a dry and ash solids baSiSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO118 Sensitivity analysis for moisture content expressed on a dry Selids baSiSOOOOOOOOOOOOOOOOOOOOOOOOOOOO121 Sensitivity analysis for volatile solids expressed on a dry ash and salids baSiSOOOOOOOOOOOOOOOOOOOOOOOOOOOO121 Values of constants A and B from dairy manure solids specific gravity measurementSOOOOOOOOOOOOOOOOOOOOOOOOOOOO128 Partial derivatives of porosity calCUIationOOOOOOOOOOOOOOOOOOOOOOOOOOOOO130 Error levels of contributing factors to porosity and free air space varianceOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO13]- Partial derivatives of free air space calcu1ationOOOOOOOOOOOOOOOOOOOOOOOOOOOOO134 Summary of windrow weight and size.......l41 Time-temperature data for monitoring locations in Windrow l...................158 Time-temperature data for monitoring locations in Windrow 2...................159 Windrow color changes....................209 Windrow penetration depths...............211 Effect of windrow physical property variability on the calculated thermal properties in Windrow lB.................23l Length of time spent in selected temperature ranges by location and windrow........ 239 Temperature penetration time as a xiv ———— Table Table Table Table Table Table Table Table Table Table Table Table Table Table 6.4: A.1: A.2 B.l C.1: D.4: D.5: D.6: E.1: F.l: F.2: function of clump diameter and thermal properties...............................246 Effect of initial decimal reduction, fraction of lethal temperature and thermal death coefficent on the number of windrow turns required....... ......... 247 Limits of error for thermocouple wire (Reference Junction at 0°C...........261 Summary of sources of error in thermocouple measurements................263 Water bath thermocouple calibration --Run lOOOOOOOOOOOOOOOOOOOOOOOOOOOOO OOOOO 264 Effect of sequential gas sample and location on carbon dioxide and oxygen concentrations....................269 Gas Chromatograph trace length as a function of withdrawal number............272 Oxygen concentration as a function of withdrawal number........................273 Effect of initial pressure, sample container volume, and sample size on sample container and syringe internal pressure after sampling.........273 Low error intermediate calculations......275 Final gas correction calculations........277 Gas concentration calculations...........280 Concentration and syringe use data.......284 Effect of sampler type and sample depth on calculated bulk density, porosity and free air space......... ..... 286 Statistical analysis of bulk density and free air space data .................. 292 K V Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure LIST OF FIGURES Hypothetical bacterial growth curve (From Olson and Nottingham, 1980)........l4 Effect of temperature on the generation time of a typical mesophile Escherichia coli) (Adapted from Olson and Nottingham, 1980)....................19 Generalized types of survivor curves observed in studies of heat inactivation of microorganisms (From Hang, 1980).coco-coo00.000000000000000.0.25 Observed oxygen consumption rates for various composting mixtures and reactor types as a function of temperature. Each curve represents the best fit of observed data (From Haug, l980)..........44 Effect of moisture content on oxygen consumption rates of various composting substrates (From Haug, 1980).............46 Section through a mushroom compost heap illustrating the typical differences in temperature and aeration (From Lambert, 1941)OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO48 "Apparent” thermal conductivity of various materials at different densities. A, kapock; B, baggasse; C, cork board; D, slag wool; E, mineral wool (From Pratt, 1969).......63 Effective conductivity of porous material due to latent heat transfered by movement of water vapor (From Pratt 1969)....................................66 Conceptual illustration of mass transport of major components during composting (From Haug, 1980).............72 Probe for sampling temperatures and xvi Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 4.11: 4.12: gas showing alternating wooden and copper sections. Probes had one to four perforated copper diffusion chambers depending on location...........82 Typical temperature and gas sampling locationSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO83 Orthogonal projection of test windrow showing temperature and oxygen probe placement. Windrow height and base dimensions are typical...................84 Gas sampling apparatus. Valve is in position to allow sample to be drawn from diffusion chamber in pile...........88 Cross-section of probe insertion hole showin annular void space created by probe 1nsertion..........................92 Sampling points for sequential gas samplingOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO95 Sequential sample gas concentration, Location lOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO96 Effect of cummulative gas withdrawal on Oxygen concentrationOOOOOOOOOOOOOOOOOOOOO99 Effect of sample pressure (calculated) on oxygen concentration..................99 Percent influx as a function of internal pressure......................100 Rotary core sampler....................113 Core sample locations. Cylindrical in shape, the center of the sample was located at approximately the same point in the y-z plane as the temperature/gas sampling point.........114 Effect of length errors on total variance as a percent of total VOIUmEOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO124‘ Effect of changes in radius variance on total variance as a percent of total v01umeOOOOOOOOOOOOOOOOOOOOOOOOOOO124 Effect of two different radius errors on total variance as a percent of bulk xvii Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure density measurements...................126 Relative contribution of various sources of error to total variance as a percent of bulk density measurements...........................126 Relative contribution of radius and height errors to total variance as a percent of bulk density percent measurement............................127 Total porosity variance for 3 levels Of factor varianceOOOOOOOOOOOOOOOOOOOOO132 Contributions to total porosity variance due to high factor variance...133 Total free air space variance for 3 levels of factor variance..............136 Contributions to total free air space variance due to high factor variance...136 Ambient weather conditions..............140 Cross-sections of Windrows 1A and 2A....141 Initial elevations for front half of Windrow 18 (plan view)..................143 Elevation changes in front half Windrow 18 after one week...............l44 Cross section of Windrow lB elevation ChangeSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO14S Cross section of windrow showing distances from a monitoring point to windrow surface along the probe (Dp) and the nearest surface point (Ds)......145 Measurements of temperature variability in Windrow 2................147 Temperature profiles for Locations 1, 2’ and 3' Windrow1....OOOOOOOOOOOOOOOOOlSO Temperature profiles for Locations 4, 5' ands, Windrowloo-00000000000000....150 Temperature profiles for Locations 8, 9, 10 and Slab, Windrow 1..............151 xviii Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 5.11: 5.12: 5.13: 5.14: 5.15: 5.16: 5.17: 5.18: 5.19: 5.20: 5.21: 5.22: 5.23: 5.24: 5.25: 5.26: for Windrow at noon......152 profiles on Day 1 Temperature (°C) 1A cross-section for Windrow at noon......153 profiles on Day 3 Temperature (°C) 1A cross-section for Windrow at noon......154 profiles on Day 5 Temperature (°C) 1A cross-section Final temperature (°C) profiles for Windrow 1A cross-section on Day 7......155 Final temperature (°C) profile for Windrow 2COOOOOOOOOOOOOOOOOOOOOOOOOOOOO156 Time spent in each temperature class by location in Windrow 1A. Total time for windrow was 6.96 days.........160 Time spent in each temperature class by location in Windrow 13. Total t1me for windrow was 6.89 days.........161 Time spent in each temperature class by location in Windrow 1C. Total time for windrow was 6.81 days.........162 Time spent in each temperature class by location in Windrow 2A. Total t1me for windrow was 5.80 days.........l63 Time spent in each temperature class by location in Windrow 2B. Total time for windrow was 6.84 days.........164 Time spent in each temperature class by location in Windrow 2C. Total t1me for windrow was 6.88 days.........165 Gas concentrations at Location 10 showing suspect high oxygen levels.....166 Difference between original and corrected oxygen concentration for selected dataOOOOOOOOOOOOOOOOOOOOOOOOOO167 Oxygen concentration data for Windrow lOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO168 Oxygen concentration data for Windrow 2000OOOOOOOOOOOOOOOOOOOOOOOOOOO169 Carbon dioxide concentration data for Windrow 1OOOOOOOOOOOOOOOOOOOOOOOOOO170 xix Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 5.27: 5.28: 5.29: 5.30: 5.31: 5.32: 5.33: 5.34: 5.35: 5.36: 5.37: 5.38: 5.39: 5.40: 5.41: 5.42: Carbon dioxide concentration data for Windrow 2OOOOOOOOOOOOOOOOOOOOOOOOOO171 Gas concentrations by location, Windrow 13' Day 3000OOOOOOOOOOOOOOOOOOO172 Gas concentrations over time, Location 1. Lines indicate general trends in data, not intermediate values..........172 Gas concentrations over time, Location 3. Lines indicate general trends in data, not intermediate values..........l74 Gas concentrations over time, Location 10. Lines indicate general trends in data, not intermediate values..........174 Gravimetric moisture content (% wb) by location for Windrow 1..............175 Gravimetric moisture content (% wb) by location for Windrow 2..............l76 Average gravimetric moisture content (% wb) by location and turning for Windrows l and 2.......................177 Average volatile solids content (% db) by location and turning for Windrows land 20.000000.00000000000000000000000178 Volatile solids content (% db) by location for Windrow 1.................l79 Volatile solids content (% db) by location for Windrow 2.................180 Wet bulk density (kg/m3) by location forWindrow1OOOOOOOOOOOOOOOOOOOOOOOOOO183 Wet bulk density (kg/m3) by location for Windrow 2000OOOOOOOOOOOOOOOOOOOOOOO184 Average wet bulk density (kg/m3) by location and turning for w1ndrow51and 2000OOOOOOOOOOOOOOOOOOOOlBS Wet bulk density changes over time.....187 Wet bulk density changes over time by locationOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO187 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 5.43: 5.44: 5.45: 5.46: 5.47: 5.48: 5.49: 5.50: 5.51: 5.52: 5.53: 5.54: 5.55: 5.56: 5.57: 5.58: 5.59: 5.60: 5.61: 5.62: Wet bulk density changes over time greater than 11 % coefficent of variationOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO188 Porosity by location for Windrow 1.....189 Porosity by location for Windrow 2.....190 Average porosity by location and turning for Windrows l and 2...........191 Porosity changes over time.............l92 Porosity changes over time by locationOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO0192 Free air space by location for Windrow1000OOOOOOOOOOOOOOOOOOOOOOOOOOO193 Free air space by location for Windrow 2000OOOOOOOOOOOOOOOOOOOOOOOOOOO194 Average free air space by location and turning for Windrows l and 2.......l95 Free air space changes over time.......l96 Free air space changes over time by locationOOOOOOOOOOOOOOOOOOOOOOOOOOOO196 Free air space changes over time greater than 7 % coefficent of variationOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO197 Volumetric moisture content by location for Windrow l.................198 Volumetric moisture content by location for Windrow 2000OOOOOOOOOOOOOO199 Fractional solids content by location for Windrow l.................200 Fractional solids content by location for Windrow 2.................201 Void ratio by location for Windrow 1...202 Void ratio by location for Windrow 2...203 Degree of saturation by location for Windrow 1..... .204 Degree of saturation by location for xxi Figure 5.63 Figure Figure Figure 6.2 Figure 6.3: Figure 6.4 Figure Windrow 2000OOOOOOOOOOOOOOOOOOOOOOOOOOO205 Windrow 1 cross-section showing color zoneSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO206 Windrow 2 cross-section showing color zoneSOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO207 Summary of assessment of time- temperature effects on microbial surVival and growt'hOOOOOOOOOOOOOOOOOOOOO240 Windrow 1B temperature and 210918 2, 3 reduction at Locations ...... 42 Windrow lB temperature and 1091 reduction at Locations 4, 5, an 6......242 Windrow lB temperature and 1091 reduction at Locations 8, 9, an 10.....243 Decimal reductions in Windrow 18 by locationOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO244 xxii CHAPTER 1 INTRODUCTION In the last decade the use of separated dairy manure solids as a bedding material for dairy cattle has received increasing attention and use. Depending on herd size and the availability and cost of other bedding materials, the separated dairy manure solids can provide an economic alternative (Keys et a1., 1977; Allen et al., 1979). The direct economic benefits in dairies using liquid flush manure handling systems can be augmented by savings from decreased labor in the manure flushing system, the construction of smaller waste storage lagoons and less frequent removal of manure solids (Hermanson, 1985; Lindemann, 1985). Some studies have shown that cows prefer separated manure solids to other types of bedding (Keys et a1, 1977). 1.1: Criteria for Use 95 Dairy Manure Solids as a Bedding MateriaI In order to be a satisfactory bedding material, the dairy manure solids must meet several criteria. Moisture content affects general cow cleanliness and the ease of handling the bedding material. More importantly, high levels of certain micro-organisms in bedding material have been associated with mastitis in dairy cows (Bramely and Neave, 1975). 1.2: Effects of Mastitis on the Dairy Industry 1.2.1: Physical Effects Mastitis is an infection of the mammary gland. In dairy cows it is characterized by swelling, heat, redness, pain and disturbed function. Clinical signs of the disease are a result of the cow's defense mechanism attempting to destroy invading micro-organisms, return milk production to normal and repair the damaged gland (Jarrett, 1981). Mastitis can begin with injury to the teat or the entry of pathogenic bacteria into the teat (Bramley, 1974; Carroll, 1977). The entry of pathogenic bacteria can be affected by the milking routine, disease prevention procedures, faulty milking machines and the sanitary quality of the cow's environment, ie. the bedding and water supply. The severity of mastitis attacks can range from chronic, low-level infections which affect the quality and quantity of the milk produced, to peracute in which all signs of inflammation are present, together with fever, depression, shivering, loss of appetite and rapid weight loss (Jain, 1979). Peracute mastitis can lead to the cows' death within hours; cows that survive may not recover their milk production and must be culled. Mastitis severity is related to the type of bacteria and the resistance of the host (Jarrett, 1981). The pathogenicity of coliform organism serotypes, for instance, varies widely (Carroll et al., 1973). Dairy cows are 3 particularly susceptible to coliform mastitis just prior to parturition and in the early stages of lactation (Bramely, 1974; Eberhart & Buckelew, 1972). 3 1.2.2: Economic Effects Economic losses from bovine mastitis can be serious. ‘Estimates of the United States' annual losses due to mastitis during 1976 were $1,294 million (Blosser, 1979). Reduced production and discarded milk accounted for $897 (69.3 %) and $142 (11.0 %) million, respectively. Increased replacement costs due to death and early culling caused a $103 (8.0 %) million estimated loss. 1.3: Micro-organisms Associated with Mastitis 1.3.1: General Organisms A wide variety of micro—organisms has been associated with mastitis. Streptococcal and staphylococcal organisms account for many cases of chronic, low level mastitis that reduce the quantity and quality of milk production (Nat. Masitis Coun., 1978). Coliform organisms have been implicated in outbreaks of peracute mastitis and the subsequent death of cows (Nat. Mastitis Coun.,ll978). Bovine mastitis produced by Cornebacteria (Packer, 1977a), Psuedomonas aeruginosa (Packer, 1977b), mycoplasma (Jasper, 1977) and yeast-like fungi (Farnsworth, 1977) have also been reported. 1.3.2: Coliform Organisms Mastitis caused by coliform organisms is of particular concern when dairy manure solids are used as bedding. 4 "Coliform" loosely refers to lactose-fermenting organisms of the family Enterobactericeae, including the genera Escherichia, Klebsiella, and Enterobacter (Eberhart, 1977). Coliform organisms are prevalant in the dairy environment (Brander, 1973). Levels of coliform organisms greater than 1x106 have been associated with outbreaks of mastitis (Bramely, 1974; Bramely and Neave, 1975). This level of coliform organisms occurs in many types of fresh and used bedding in dairy environments (Bramely and Neave, 1975; Neave and Oliver, 1962; Jasper et al., 1975, Carrol and Jasper, 1979; Bishop et al., 1980; Rendos et al., 1975). High-producing cows tend to spend longer times in contact with bedding materials than low producers (Francis, 1981); this leads to high temperatures and increased coliform growth in bedding used by high producers (Francis, 1981). The existance of high coliform organism levels does not not necessarily lead to coliform mastitis, however. Coliform mastitis is considered a disease of glands not affected by other pathogens (Jain, 1979). Colonies of Staphylococcus aureus and Streptococcus agalactia adhere much more easily to the teat than do coliform organisms .(Frost, 1977). In the absence of specific therapy, colonies of S. aureus and S. agalactia grow into the cow's teats and produce low levels of chronic mastitis (Jain, 1979). Coliform infections are difficult to superimpose upon other infections. 500,000 leukocytes/m1 of foremilk from any irritant, will usually prevent the establishment of 5 experimental innoculums of coliforms (Schalm et al., 1971). Significant progress has been made in the development of programs to reduce Streptococcus agalactiae and Staphylococcus aureus (Carroll, 1977). These programs for teat dipping and therapy of non—lactating cows are not effective against coliform organisms, however (Carroll, 1977; Eberhart, 1977). The removal of the staphylococcal and streptococcal infections opens the way for infection by coliform organisms. 1.4: Treatment of Dairy Manure Solids for Bedding 1.4.1: General Treatment Methods One means of reducing the incidence of coliform mastitis is to treat the bedding material so that it will not support high levels of the organisms. Early work with paraformaldehyde indicated that its application would initially reduce levels of coliform organisms (Bramely and Neave, 1975). Recovery of coliform numbers was swift, however, and the economics of frequent chemical applications was judged to be prohibitive (Bramely and Neave, 1975). The ' use of lime additions to raise pH levels has been used with some success (Smith, 1985). 1.4.2: Treatment by Composting The application of composting to reduce coliform numbers and produce a relatively inert bedding material from separated dairy manure solids has also been attempted (Carroll and Jasper, 1979; Bishop et al., 1980; Allen et al., 1980). Composting has been shown to reduce the numbers 6 of mastitis-causing pathogens below the critical level of 1x106 (Carroll and Jasper, 1978; Bishop et al., 1980; Allen et al., 1980). While regrowth of mastitis causing organsism has been noticed in compost used as bedding (Bishop et al., 1981), similar levels of regrowth have also been shown in other bedding materials (Rendos et al., 1975). Furthermore, the compost used in these experiments was not subjected to an intensely managed compost process such as a force aeration system. Coliform control in composting dairy manure solids can be achieved in two manners. Temperatures inside compost piles of refuse and sewage sludge reach 80°C (Wiley, 1957; Shultze, 1962; Willson et al., 1980). Dairy manure solid compost temperatures up to 55°C have been reported (Allen et al., 1980). Bramely and Neave (1975) found that maintaining a temperature of 50°C in sawdust bedding samples killed the organisms, while 30 and 40°C temperatures led to increased numbers. Secondly, if the pile is correctly managed, the substrate should be reduced, thus inhibiting regrowth. 1.5: The Composting Process 1.5.1: Definition Composting is a microbial self-heating process in which heat given off during aerobic respiration, in conjunction with reduced local oxygen concentration, depleted substrate, or extreme moisture level, tends to limit microbial growth (Finstein et al., 1980). Heat transfer and mass transfer of 7 oxygen, water and carbon dioxide is governed by diffusion, natural convection, forced convection or by a combination of processes. 1.5.2: Parameters Affecting Composting Research on composting has focused on empirical descriptions of the overall composting process and the impact of selected parameters on the desired objectives (Snell, 1957; Schulze, 1962; Jeris and Regan, l973a,b,c). Temperature, oxygen level, moisture content, free air space and substrate properties, such as the amount of volatile solids present and the carbon/nitrogen ratio, have been found to be important parameters. 1.5.3: Spatial Variability The distribution of temperature and oxygen in a compost pile has been found to be non-uniform (Lambert, 1941). Furthermore, Carrol and Jasper (1978) have observed that the distribution of mastitis-causing organisms within a composting pile varies with time and temperature. 1.5.4: Need for Distributed Heat and Mass Transfer Composting Modelf Because of the spatial variability and the difficulty in obtaining time-temperature relationships, it is difficult to predict the overall level of mastitis-causing organisms in a compost pile based on grab samples. This is particularly true when different parts of a compost pile are mixed together and used as bedding material. For example, if 90 % of the compost has coliform levels of 1x103 per gram of bedding, and only 10 % has levels of 1x107, the I. thoroughly mixed pile will still have dangerous levels of 1x106 per gram of bedding. Extensive experimentation is necessary to determine the correct pile size and shape to achieve uniform reductions in mastitis-causing organisms. A model that predicts spatial and temporal changes in temperature would show where in the pile treated bedding material could be obtained. 1.5.5: Modeling Composting Heat and Mass Transfer Processes Little work has been done to model composting. Comprehensive simulations of continuous-feed completely- mixed and batch models of the composting process have been developed by Haug (1980). Finger (1975) developed a distributed heat-and mass-transfer model for windrow composting that is based on constant property thermal heat conduction and the diffusion of oxygen into the compost pile. Modeling spatially-distributed heat-and-mass— transfer processes in a compost pile can be improved. Haug's models, while sound in theory, are not distributed models. Finger's model, while a distributed model, has four potentially serious shortcomings: (1) it does not take into account the reduced respiration that occurs at high temperatures; (2) it ignores the role that water could play in the heat-and-mass-transfer process, (3) it is based on constant physical and thermal properties, and, (4) it can be applied only to rectangular geometries. 9 1.5.6: Recent Advances in Composting Recent research has indicated the nature of the shortcomings of Finger's model. McKinley and Vestal (1984) found that the maximal respiration for sewage sludge composting occurs in the range of 25°C to 45°C. Other studies (Finstein et al., 1980) have noted that the temperature optimum for composting is in the range of 50°C to 55°C. Moisture content has been found to affect the thermal and physical properties of compost. Bonhoff, et al. (1984) found that for separated dairy manure, the specific heat, thermal conductivity, thermal diffusivity, and free air space are functions of the percent moisture content. The thermal diffusivity is also a function of temperature and bulk density. Mears et a1. (1975) found that the specific heat and thermal conductivity of swine manure compost is a function of moisture content. Oxygen diffusion rates in ground garbage compost material are linearly related to the free inter-connected air space and inversely affected by increases in moisture content (Shell, 1955). Water can play a large role in heat and mass transfer within a compost pile. Finstein et a1. (1983) estimated that when vaporization is used as a strategy in forced aeration systems for temperature control, almost 88 % of the heat removal is through vaporization, while only 10 % is through dry air convection and 2 % by conduction. Bulk density increases and pile volume decreases with time in long—term undisturbed swine manure compost windrows (Mears et al., 10 1975). Reliance on experience with other types of composting may not lead to an accurate understanding of the problems related to dairy manure solids composting, however. Creation of adequate free air space (air filled porosity) is an important concern with sewage sludge composting because of its small particle size (Haug, 1980). Mote and Griffis (1980) found that moisture management is of crucial importance with composting cotton gin trash. Composts made from different carbon sources exhibited very different temperature profiles when carbon/nitrogen ratio and bulk density were held constant (Mote and Griffis, 1980). Since the few studies of composting dairy manure solids (Carrol and Jasper, 1978; Bishop, et al., 1980; Allen et al., 1980) do not focus on the actual composting process, it would be desirable to observe the composting of this substrate as part of an overall modeling effort. CHAPTER 2 OBJECTIVES The objectives of this research are (l) to conduct experimental studies of a partially enclosed windrow dairy manure solids (DMS) composting systems aimed at assessing the importance of variables affecting composting heat and mass transfer; (2) to assess the accuracy of the data collection effort; and (3) to make suggestions for future modeling of heat and mass transfer in DMS composting. ll CHAPTER 3 LITERATURE REVIEW 3.1: Description pf Composting Composting is a member of a class of processes that exhibit self-heating behavior (Cooney and Emerson, 1964). There is general agreement about the sequence of events leading to self-heating and, potentially, the spontaneous combustion of materials. ...If the material is reasonably fresh, the initial warming may be ascribed in part to the enzymatic activities of the still living plant material itself. However, it is generally conceded that the major; buildup of heat results from the metabolism of the mixed saprophytic microbial flora that promptly develops. The temperature rises quickly,...often in a matter of hours if the decomposing material is finely divided and rich in readily available nutrient (such as fresh grass, chopped guayule, or manure), and soon passes the optimum for mesophilic forms. Here the thermophiles become of critical importance; they multiply rapidly and raise the temperature to the peak that can be reached by microbial activity--about 70°C or slightly higher. Subsequent heating results then from the autocatalytic chemical processes which can begin to operate at this temperature. The processes cause further heating, marked chemical changes in the heating mass, and ultimate ignition if the necessary conditions are maintained (Cooney and Emerson, 1964). Composting differs from this general course of events in two manners. Since most composting processes involve dead or waste materials, the inital temperature rise is primarily caused by mesophilic saprophytic microorgansims 12 13 instead of the respiration of live plant cells. Secondly, temperatures do not generally reach the autocatalytic stage. Instead, temperatures are maintained at levels close to the maximum temperature for thermophilic growth until either the microbial population is so debilitated that it cannot support the high temperatures (Finstein, 1980), moisture content is reduced to inhibitory levels, or the substrate is exhausted (Finstein and Morris, 1975). At this point the temperatures begin to fall. 3.2: Analogous Systems Insight into the underlying principles of dairy manure solid composting can be gained from a number of sources. Work in the late 19th and early 20th century led to a basic understanding of the self-heating and spontaneous combustion of a variety of agricultural products (Cohn, 1889; Miehe, 1905, 1930, Waksman et al.,l939). Some of the products investigated include hay, manure, barley, cornmeal, cracked corn, oats, cotton, hemp and hops (Cooney and Emerson, 1964). Self-heating caused large losses from fires and overheating and was estimated to be over $20 million per year in the 1920's (Browne, 1929). Losses due to overheating and spontaneous combustion have also been noted in the peat (Finnish Peatland Society, 1982) and forest products (Hajny, 1966) industries. On the positive side, microbial thermogenesis has been used to process a number of important products including cacao (Chatt, 1953), tobacco (Garner, 1946), horticultural l4 container media (Hoitink and Poole, 1980) and composts for mushroom production (Haiser and Sinden, 1953) The application of composting to waste processing and treatment has also been extensively studied. Substrates examined include municipal wastes (Jeris and Reagan, 1973), leaves (Strom et. al., 1980), sewage sludge (Haug, 1979), septage (Lombardo, 1977), fruit and vegetable cannery wastes (Rose et al., 1965), cotton gin trash (Griffis and Mote, 1978), grain dust (Chang et al., 1980), and horse (Pizer, 1950) swine (Martin et al., 1974), poultry litter (Bell and P05, 1971), and dairy manure (Hummel and Willson, 1975). 3.3: Microbiology 3.3.1: General The growth of a micro-organism in a favorable environment procedes as shown in Figure 3.1. There is an initial period of adjustment to the environment, called the lag phase. Growth begins and accelerates into a period of rapid and constant exponential growth, called the LOGARITHM OF NUMBERS OF BACTERIA l 0 4 8 12 15 20 24 28 32 36 40 UNITS OF TIME Figure 3.1: Hypothetical bacterial growth curve (From Olson and Nottingham, 1980) 15 exponential or logarithmic growth phase. This growth is eventually slowed and then stops due to the accumulation of toxic materials or depletion of nutrients. At this point, the stationary phase, cell division and death are in approximate balance. The final phase occurs when cell death becomes greater than new cell production (Olson and Nottingham, 1980). The temperature ranges for growth can be used to distinguish between three major groups of micro-organisms. These groups are shown in the Table 3.1. Within each group there are organisms that can survive into the temperature ranges of the adjacent groups (Olson and Nottingham, 1980). Table 3.1: Growth temperatures for micro-organisms. Temperature (°C) Group Minimum Optimum Maximum Psycrophiles -5 to,+5 12 to 15 15 to 20 Mesophiles 5 to 15 30 to 45 35 to 47 Thermophiles 40 to 45 55 to 75 60 to 90 3.3.2: Types 93 Micro-organisms A wide variety of micro-organisms have been identified in composts. Bacteria (including actinomycetes) and fungi are the predominant organisms found. 3.3.2.1: Bacteria and Actinomycetes Bacteria have a number of morphological forms and are typically unicellular, but multicellular associations are 16 common. These associations can be true multicellular states or accumulations due to cell division. Actinomycetes have an elaborate multicellular structure that arise from a single reproductive cell. Extensive chains and branches develop into mycelia which structurally resemble molds. Actinomycete filaments are one tenth to one fourth the diameter of fungi; actinomycetes have a procaryotic structure compared with eucaryotic fungi, as well as biochemical and sexual differences (Haug, 1980). The most common mode of bacterial reproduction is by binary fission, although sexual reproduction or budding can also occur in some cases. When unfavorable environmental conditions are encountered, some bacteria are capable of forming dormant cells which are more resistant to harsh environmental conditions. These forms include endospores, cysts, and exospores. Endospores are the most stable of the three forms (Haug, 1980). 3.3.2.2: §u_gi Fungi are eucaryotic and heterotrOphic spore-bearing organisms that lack chlorophyll (Atlas and Bartha, 1981). Slime molds and true fungi make up the two broad divisions of fungi. The true fungi can be further divided into molds and yeasts. Both aerobic and anaerobic metabolism is observed among yeasts, while molds are aerobic (Haug, 1980). Fungi can tolerate low moisture conditions, have a larger optimum pH range and often have a lower nitrogen requirement than bacteria (Haug,l980). Bacteria and fungi rely on 17 similar organic substrates in composting systems and they are often in direct competition. 3.3.3: Compost Microbial Ecology 3.3.3.1: General Environmental conditions determine which organisms can be present on a sustained basis in a given ecosystem. From an ecological standpoint the interaction between a biological system and the abiotic environment is best described by Oden's Combined Law, which states ...the presence and success of an organism or a group of organisms depends upon a complex set of conditions. Any condition which approaches or exceeds the limits of tolerance is said to be a limiting condition or a limiting factor (Atlas and Bartha, 1981). Atlas and Bartha (1981) discuss the wide variety of environmental factors that can affect micro-organism growth: nutrients, temperature, water activity, pH, redox potential, pressure, and inhibitors. Tolerance ranges for any one parameter are interactive with other parameters. Nutrients or inhibitors must not only be present in the micro- organism's environment, but they must also be in a form that is available to the micro-organism. The microhabitat of a micro-organism exerts the most direct effect on microbial growth and survival. The ability of micro-organisms to alter their environment can have a positive, negative or neutral effect on what microbial populations will be present. 18 3.3.3.2: Temperature The metabolic activity of an organism is affected by temperature. Higher, but non-lethal temperatures increase metabolism. The change in enzyme activity caused by a 10°C rise in temperature is called the Q10 value (Atlas and Bartha, 1981). Experimentally determined Q10 values for forced aeration garbage compost are shown in Table 3.2 (Moore, 1958; Wiley and Pierce, 1955). Temperature levels affect the length of time for a new generation of micro-organisms to be produced, or the generation time. Figure 3.2 shows a curve for a typical meSOphile (E. coli). Initially, as temperatures increase, growth also increases and generation times decrease. In the optimal growth range, generation times are short and fairly constant. Generation times go to infinity at a temperature only slightly greater than the optimal growth maximum (Olson and Nottingham, 1980). Despite the general increases in biochemical activity with temperature, thermophiles have longer generation times than mesophiles and have lower cell yields in proportion to the amount of substrate utilized. This occurs because thermophillic organisms spend a great deal of time and energy repairing heat damage (Atlas and Bartha, 1981). Micro-organisms make a number of adaptations to high temperature conditions (Atlas and Bartha, 1981). These adaptations include increases in the proportions of saturated lipids in membranes, synthesis of heat resistant 19 Table 3.2: Q10 values for garbage composts. Temperature Q10 Range (?C) Moore (1957) Wiley and Pierce (1955) 30 to 40 2.6 1.70 40 to 50 1.80 1.65 50 to 60 1.77 1.60 _~1soo~ ‘23: E 51000- g g E 66/! L”; 5 mm» 5 l - , 10 20 30 ‘30 50 TEMPERATURE (’C) Figure 3.2: Effect of temperature on the generation time of a typical meSOphile (Escherichia coli) (Adapted from Olson and Nottingham, 1980) 20 enzymes, and the inactivation of different enzyme systems at different temperatures. In addition, thermophilic organisms exhibit amino-acid and growth factor requirements at high temperatures that are not apparent at optimal temperatures. The maximum heat tolerence of vegetative thermophiles is correlated with the heat stability of their ribosomes. When temperatures exceed the optimum for each organsim, injury or death can occur. Mild time/temperature exposures above the optimum temperature can cause stress which leads to injury. While an injured cell may remain viable, it may be unable to reproduce until the injury is healed (Olson and Nottingham, 1980; Busta, 1978). The type of organism can affect survival rates. Spores of bacteria are resistant to high temperatures. Survival for minutes at 120°C or for hours at 100°C is not uncommon. Vegetative cells are killed after brief exposures to 70 to 80°C whether they are spore-formers or not (Olson and Nottingham, 1980). A number of possible mechanisms have been proposed to explain the lethal influences of temperature on vegetative bacteria, including (1) coagulation of protein; (2) inactivation of enzymes; (3) disruption of cellular lipids; (4) damage to the genetic apparatus; (5) breakdown of RNA (Allwood and Russell, 1970). Adams (1978) reviewed the heat injury of bacterial spores. Spore injury manifestation can fall into four major classes: 21 .....(1) requirements by survivors for non-nutrient germination stimulants, (2) modified optimum temperatures for the enumeration of survivors, (3) increased sensitivity of survivors to inhibitors, and (4) altered nutritional requirements by survivors. Others, such as changes in the influence of pH, or Eh, and the choice of the recovery medium, have not been studied sufficiently to place them in specific classes (Adams, 1978). Spores must germinate and complete outgrowth before vegetative cell growth can begin (Adams, 1978). During this period, the resistance of a spore to environmental stresses caused by heat, radiation, chemicals and extreme pH decreases. Spore germination is a multistage process that is mediated by a variety of agents (Adams, 1978). Injury that is expressed during outgrowth can involve a number of spore structures and metabolic activities. Actinomycete temperature optima are not well defined (Finstein and Morris, 1975). Some researchers report temperature maxima in the range of 55°C, while others have found actinomycetes at temperatures.up to 75°C. Data on temperature ranges for thermophilic fungi have been presented (Cooney and Emerson, 1964). The minimum growth temperature is between 20 and 30°C; the maximum temperatures are 55 to 60°C. The relationship between temperature and growth of 6 fungi isolated from municipal waste compost was studied by Kane and Mullins (1973). Growth of fungal colony diameters was greatest in a range of 35 to 50°C. Between 55 and 60°C the growth dropped to zero. A study of the temperature and pH optima for 21 species of both thermophilic and thermotolerant fungi was conducted by 22 Rosenburg (1975). His results agreed well with those of Kane and Mullins. No correlation was found between temperature optimum and pH optimum among members of the group tested. An organism's resistance to these higher temperatures is expressed as the thermal death time. The thermal death time is the amount of time at a particular temperature that is required to kill a given number of organisms (Atlas and Bartha, 1981). By plotting the numbers of micro—organisms that survive a given temperature after a certain time on a semi-logaritmic plot, the decimal reduction value (Dr)' or time to give a ten-fold reduction, can be calculated (Atlas and Bartha, 1981). Some common decimal reduction times are given in Table 3.3. The decimal reduction time for a given organism will change with temperature. By plotting the Dr values for each temperature on semilog paper, a straight line will generally be formed if death is first-order. The slope of this line, 2, is the temperature interval required for a log reduction in the Dr value (Atlas and Bartha, 1981). Values of z for selected microorganisms are given in Table 3.4. Values of z are not constant over the entire temperature range because the value becomes infinite at some point (Hansen and Riemann, 1963). Differences in the 2 values between species occur. High heat resistance of an organism and high 2 values may be correlated (Hansen and Riemann, 1963). Despite the simplicity of logarithmic survivor curves, Table 3.3: 23 organisms. Organism Time Temperature (min.) (°C) Escherichia coli 20-30 57a Pseudomonas aeruginosa 2 55b Staphlococcus aureus 19 60a a Source: Atlas and Bartha (1981) b Source: Olson and Nottingham (1980) Table 3.4: 2 values for selected micro-organisms. Organism 2 (°C) Yeasts 3 - 5 Most non-sporing bacteria 4 - 6 Source: Hansen and Riemann (1963). Decimal reduction times for selected micro- 24 curves of other shapes are frequently observed (Figure 3.3). Curve A is the standard logarithmic curve. Curve B, with its lag in deaths, is frequently observed with clumped cells (Hansen and Riemann, 1963). Curve C is of particular concern because it seems to indicate that there is some temperature where the lethal effect of temperature is reduced (Haug, 1980). Two theories have been advanced to explain the deviations from the logarithmic survivor curves. The ”multiple target" theory was used by Moats (1971) to explain for thermal injury and death in bacteria. This theory could account for the lag period in curve C but not curve B. Wei and Chang (1975) advanced a theory that random collisions between disinfectant molecules and micro- organisms were the cause of microbial death. The probability for this collision was modeled with a Poisson distribution. Clumping of varying numbers of micro- organisms lead to a multi-Poisson distribution model. The result was that organism clumping was the major reason for the different survival curves. The lag period of curve C was due to a high proportion of the population being present in clumps. Curve B was explained by a large proportion of non-clumping cells combined with a few large clumps. The heat resistance of spores and vegetative bacterial cells increases with an increase in their growth temperature (Hansen and Riemann, 1963; Olson and Nottingham, 1980). Slow growing cells exhibit greater heat resistance than fast 25 LOG POPULATION TIME Figure 3.3: Generalized types of survior curves observed in studies of heat inactivation of microorganisms (From Haug, 1980) 26 growing cells (Hansen and Riemann, 1963). If large numbers of cells are initially present, longer times will be required to reduce the number of survivors to a given level (Olson and Nottingham, 1980). Heated organisms may show an increased lag period before beginning their exponential growth phase (Lembke, 1937). This occurs even if the heat treatment did not kill large numbers of organisms (Kaufmann et al., 1959). Injured micro-organisms lose resistance to selective chemical agents including salts and antibiotics. Repair of sublethal damage to microorganims occurs more readily in simple than more complex media (Olson and Nottingham, 1980). Bacteria, including E. coli, exhibit the lowest heat resistance during the exponential growth phase. The initial lag phase and the stationary phase have higher heat resistance (Hansen and Riemann, 1963; Olson and Nottingham, 1980). Bacterial levels can continue to decline even after heat treatment stops and optimal growth conditions are restored (Jackson and Woodbine, 1962). The heat resistance of microbial cells increases with decreasing humidity. Dry air at 140 to 150°C has less killing effect than wet steam at 100°C (Hansen and Riemann, 1963). Heat resistance of vegetative cells and spores can be increased in substances with relatively reduced water activity (Christian, 1980). This must be tested for each microbe and substrate combination, however (Olson and 27 Nottingham, 1980). 3.3.3.3: Moisture Content and Water Activity Water must be available for micro-organisms to grow. A useful measurement of the availability is water activity (a Water activity can be changed by either reducing the w). amount of water or by adding solutes (Christian, 1980). Water activity of a substance is the ratio of the water vapor pressure of the substance (p) to that of pure water (p0) at the same temperature: aw = p / po (3.1) If a solution becomes more concentrated, the vapor pressure decreases and the aw falls from the pure water value of l. aw values can be related to the equilibrium relative humidity (ERH) and osmotic pressure. Ideally, water activity can be related to solute concentration using Raoult's law in the form aw = p / po = n2 / (n1 + n2) (3.2) where n1 = moles of solute n2 = moles of solvent Non—ideal behavior can result, however, either from reductions due to interaction between solutes or increases because of dissociation. Experimentally determined molal osmotic coefficients are used to compensate for this nonideal behavior (Christian, 1980). aw values can be calculated from loge a = -vm ¢ / 55.51 (3.3) V where v = number of ions generated by each molecule of solute (v = l for nonelectrolytes) 28 molal concentration of solute molal osmotic coefficent m ¢ Sources of information on the molalites for various solutes at a range of aw values are available (Scott, 1957). Despite the predictive value of the preceding equation for relatively simple solutions, for complex substrates it is more accurate to directly measure the relationship between water activity and moisture content. A number of factors that can influence vapor pressure and thus water activity include adsorption of water molecules onto surfaces and capillary forces, as well as solution concentration effects (Christian, 1980). Hysteresis is also present between drying and wetting moisture isotherms. At high aw, micro-organism growth may be higher on the desorption isotherm than on the adsorption isotherm at the same aw (Acott and Labuza, 1975). The effect of water activity on the growth of many species is largely independent of the solute that controls aw (Scott, 1957). Micro-organism growth is most rapid at a levels w ranging from 0.995 to 0.980 (Christian, 1980; Atlas and Bartha, 1981). At lower aw, growth rates and the stationary population decreases and the lag phase length increases. With sufficently low aw, the lag phase becomes infinite (Scott, 1957). Values of the limiting aw for several selected organisms are shown in Table 3.5. Gram-negative bacteria, including Pseudomonas spp. and the Enterobacteriaceae only grow well above 0.96 and 0.93 aw, respectively. Gram- 29 positive non-spore-forming bacteria such as Staphylococcus aureus have a lower limit of 0.86 aw. Spore-forming bacteria can grow at aw levels of 0.94 to 0.89, with the Table 3.5: Approximate minimum levels of water activity permiting growth of micro—organisms at temperatures near optimal. Organism aw Molds Aspergillus fumigatus - 0.82 Bacteria Bacillus cereus 0.95 B. stearothermophilus 0.93 Enterobacter aerogenes 0.94 Escherichia coli 0.95 Psuedomonas fluorescens 0.97 Salmonella sp. 0.95 Staphylococcus aureus 0.86 Source: Troller and Christian (1978). common limit being 0.90 to 0.91 aw. Fungi grow at much lower aw than bacteria; fungal growth at high aw levels is much slower than bacteria (Christian, 1980). 3.3.3.4: pH The pH of a substance is an important factor in determining the survival and growth of bacteria in that substance. pH effects are difficult to separate from those of other effects such as the concentration of undissociated weak acids that are affected by pH (Corlett and Brown, 1980). The relative rate of proton leakage into the cell vs. the proton-rejecting capacity of the cell determines if 30 an environment is inhibitory due to the activity of weak acids (Freese et al., 1973). pH growth limits can vary among micro-organisms but most have optimum growth near pH 7 with ranges between pH 5 and 8. Minimum and maximum pH values for several selected micro-organisms are shown in Table 3.6. Table 3.6: The limits of pH allowing initiation of growth by selected micro-organisms. Organism Minimum Maximum PH PH Gram-negative bacteria Escherichia coli 4 4 Klebsiella pneumoniae (aerogenes) 4.4 Pseudomonas aeruginosa 5.6 Salmonella spp. 4.0 Gram-positive bacteria Bacillus cereus 4. B. stearothermophilus 5. Staphylococcus aureus 4. Streptococcus pyogenes 6. Source: Corlett and Brown (1980). The heat resistance of bacteria is decreased by acid or alkaline conditions. The heat resistance generally has a rather narrow pH range outside of which resistance falls off quite rapidly. The optimum heat resistance typically occurs at pH between 6 and 8 (Hansen and Riemann, 1963; Corlett and Brown, 1980). 31 3.3.3.5: Redox Potential Enzymatic reactions of micro-organisms are often oxidation-reduction reactions. Whether an organism can carry these redox reactions out depends on the redox state of the environment (Atlas and Bartha, 1981). Micro-organism such as strict aerobes or anaerobes which have only one terminal metabolic system are limited to a relatively narrow range of environmental redox potentials (Atlas and Bartha, 1981; Brown and Emberger, 1980). Facultive anaerobes have alternate systems which can be switched by either the environmental redox potential or oxygen concentration (Brown and Emberger, 1980). Rapid decreases in redox potential have been associated with the early logarithmic phase of micro-organism growth (Hewitt, 1950), spore germination, and regrowth (Douglas et al., 1973). 3.3.3.6: Compost Microbial Ecology Studies Finstein and Morris (1975) reviewed the microbiology of composting as it applied to solid waste. The density and succession of various micro-organisms as a function of temperature during composting has been studied by several authors (Waksman et al., 1939; Chang and Hudson, 1967; Stanek, 1972). Bacteria are present in all stages of composting (Waksman et al., 1939; Chang and Hudson, 1967; Stanek, 1972; Finstein and Morris, 1975). Bacterial diversity decreases as temperature increases to inhibitory temperatures in the 32 thermophilic zone. Actinomycete populations were slower to colonize fresh substrates than either bacteria or fungi (Lacey, 1973; Chang and Hudson, 1967). This was attributed to the ability of actinomycetes to degrade more complex substrates including cellulose, hemicellulose, chitin, and perhaps lignin (Lacey, 1973). Actinomycetes have been found to be more visible on dry vs. wet particles (Schultze, 1962), restricted to within 6 inches of the compost surface (Erickson, 1952), and to not grow well in poorly aerated masses of compost (Anon., 1953). Explicit studies of Actinomycete sucession during the temperature ascent are lacking (Finstein and Morris, 1975). Different zones of temperature were noticed in the piles and fungi were active in the cool, dry exterior regions of the piles as compared with the hot interior (Chang and Hudson, 1967). High temperatures, acidity and anaerobic conditions were felt to limit fungal growth to the exterior of the compost pile (Kane and Mullins, 1973). Low pH values in stable manure composts were associated with slow decomposition rates and anaerobic conditions (Lambert and Davis, 1934). Initially, large mesophilic and smaller thermophilic fungal populations were present (Chang and Hudson, 1967). Both populations fell, with the thermophilic fungal population disappearing after 5 days, and the mesophilic population disappearing after 5 to 8 days. A thermophilic population began to appear by day 8 and recovered to levels 33 of 1x107 by approximately day 16. The mesophilic fungal population reappeared 20 to 34 days into the compost period and only recovered to levels of 1x103 to 1x105. The ratio of thermophilic to mesophilic fungi rose during composting, and high counts of thermophiles remained in areas that had cooled (Chang and Hudson, 1967). After studying the biochemical changes that occurred during the experiment and the ability of fungi to utilize cellulose as a carbon source, Chang (1967) concluded that the ability to use complex carbon sources and the ability to thrive at high temperatures are the two important characteristics of successful colonizers of composts. Waksman et a1. (1939) studied substrate and microbial changes in stable manure held at four different constant temperatures for 47 days. The micro-organisms involved in the decomposition of the manure were found to have the following characteristics: At 75°C, the animal population and the fungi were completely repressed. Actinomycetes appeared only seldom, at the surface of the compost. Only certain types of bacteria were active, belonging largely to the spore-forming, hemicellulose-decomposing types... At 65°C, the bacteria and actinomycetes were chiefly concerned in the decomposition process. Fungi appeared only seldom, and animal forms were absent. The first two groups were represented by a number of characteristic thermophilic groups. After a certain period, the bacteria were gradually reduced and the actinomycetes became the predominant organisms. The thermophilic actinomycetes are limited to very few species, but comprised several genera. At 50°C, certain thermophilic fungi were very active, in addition to the bacteria and actinomycetes. This selective population, in which fungi and actinomycetes played the predominant role, was 34 responsible for the most rapid decomposition of the manure...The actinomycetes were similar to those developing at 65°C. Lower temperatures, as typified by 28°C, gave rise to a highly heterogeneous population. Bacteria, fungi, actinomycetes, protozoa, and nematodes were well represented by a great variety of forms. A few days elapsed before certain active types became established, a fact which accounts for the delay in the rapidity of the decomposition process at this temperature. Stanek (1972) summarized the knowledge about microbial succession of mushroom compost as follows: 1. Number of micro-organisms growing at 25°C decreases during the fermentation process in compost; number of thermophilic and thermotolerant micro-organisms increases. 2. At first number of bacteria and then of actinomycetes and thermotolerant fungi increases. The dominant types of micro-organisms are: (a) at the start of the fermentation process: mesophilic and thermotolerant spore-yielding and non-spore-yielding bacteria and quickl growing fungi (Phycomycetes--genus Mucor, etc.). (b) 'during the period of the peak—fermentation process in the pile and during pasteurization: thermophilic actinomycetes (Streptomyces, Thermonospora spp.) and non-sporulating bacteria (Pseudomonas spp.), (c) at the end of pasteurization: thermotolerent fungi (Humicola spp.etc.). (3) The compost piles are colonized in various layers (Lambert and Davis, 1934) by various micro-organisms. This phenomenon is particularly striking before the first turning of piles: on the surface there appear bacteria and fungi (especially Phycomycetes) growing at a temperature of 25°C; in the layer where intensive aerobic processes occur grow thermophilic bacteria and actinomycetes; the centre of the pile is mostly colonized by anaerobic spore-forming bacteria (Clostridium spp. etc.). After turning the pile differences decrease by the homogenization of material. (4) Various groups of micro-organisms take part in the decomposition of various substances. Some thermophilic actinomycetes and thermotolerant fungi decompose 35 cellulose, some of the thermotolerant bacteria, actinomycetes and fungi decompose pectin etc. (5) The occurrence of individual kinds and groups of micro-organisms and their activity depend on the momentary conditions (humidity, aeration, etc.) and expecially on the quantity of accessible nutrients in the compost (C:N ratio, etc.: After adding glycides the number of bacteria had increased and the amount of the ammonium had decreased; by changing sources of nitrogen the ability of thermophilic actinomycetes and thermotolerant fungi to decompose cellulose also changed. 3.4: Biochemistry 3.4.1: Stoichiometry Haug (1980) presented information on the general chemical composition of a variety of organic materials. Assuming an average compositon of sludge organics of C10H1903N, Haug determined the stoichiometric oxygen requirement. C10H1903N + 12.5 02 + 10 C02 + 8 H20 + NH3 Elevated temperatures and pH > 7.0 would lead to the volatilization of ammonia and Haug felt that nitrification oxygen demands would not normally need to be considered. Based on this assumption, about 2.0 g 02 would be needed per gram of organic matter oxidized. Finger (1975) took a slightly different approach using the following overall equation for compost reactions. (C6H1205)n + w 02 + d NH3 + CaHbOCNd + Y coz + 2 H20 where a = 2.82n W = 3.03 n moles OZ/mole subst. b = 4.69n Y = 3.18 n moles COZ/mole subst. c = 1.40n Z = 4.31 n moles HZO/mole subst. d = 0.436n Finger based his stoichiometric equation on the assumptions that the substrate could be represented as a carbohydrate 36 macromolecule, that the elemental composition of the microbe is similar to yeast, and that the conversion efficiency of the substrate into cell material is 0.4 by weight. The respiratory quotient (RQ) is the ratio of carbon dioxide produced to the oxygen utilized by microbes. The theoretical RQ for the complete oxidation of carbohydrates, protein and fats was given as 1.0, 0.8 and 0.7, respectively (Braithwaite, 1956). The RQ for forced aeration ground garbage composting was determined to average about 0.9 (Moore, 1958). Braithwaite (1956) found that the RQ was 1.0 or greater for the first few days and progressively lower for the remainder of the composting period, eventually approaching 0.6. 3.4.2: Compost Studies Changes in a number of the biochemical constituents of compost or compost-like material have been studied by many authors. Gregory et a1. (1963) carried out an extensive investigation on microbial and biochemical changes in moldy timothy and fescue grass hay. Gerrits et al. (1965) and Muller (1965) studied changes in biochemical constituents during the preparation of synthetic mushroom composts. Higgins et a1. (1982) compared the organic composition of aerObic, anaerobic, and compost-stabilized sludges. 3.4.2.1: Organic Substrates Chang (1967) analyzed changes in hemicellulose, cellulose, lignin, diastase and ethanol soluble fractions, total N and ammonium and nitrate N in wheat straw composts. 37 The straw lost over half of its dry weight after 60 days of composting; almost all of the loss occured in the first 34 days. The first 5 days had the greatest rate of loss with an average loss of 2.66 % per day as opposed to an average of only 1.3 % per day for the next 30 days. The ethanol soluble fraction, which contained sugars, glucosides, and essential oils showed a slight increase. Starch and glycogen were hydrolysed by diastase. This remained essentially constant over the composting period. Losses of hemicellulose and cellulose acounted for most of the total dry weight lost. Hemicellulose decreased from 35.6 to 16.9 percent. Cellulose changed from 45.3 to 13.2 percent. The rate of hemicellulose decomposition was uniform through the composting but cellulose decompositon rates varied. High cellulose decomposition rates were associated with thermophilic temperatures. Lignin did not change appreciably from approximately 10 percent. Waksman et al. (1939) studied the influence of temperature upon the microbiological population and decomposition processes in horse manure composts. Manure was heated to four constant temperature levels for 47 days. Samples were withdrawn on the 9th, 19th, 33rd and 47th days for analysis of microbial populations, moisture, ash, ammonia, hemicellulose, cellulose, lignin, water soluble organic matter and water insoluble protein. Total decomposition at 50°C was always greater than that at 28°C. The 75°C samples by far had the smallest 38 total decomposition. 65°C total decomposition was the highest at day 9 but by the 47th day it was approximately the same as the 28°C sample. The rate of cellulose decomposition was initially steeper than that of hemicellulose; the decomposition rates and levels became very similar by the end of the experiment. The effect of temperature on changes in cellulose content paralleled the order of the total decomposition rates. Almost no cellulose degradation occured at 75°C, however. .Hemicellulose degradation was highest at 50°C and lowest at 75°C. Degradation of hemicellulose at 65°C was greater than that at 28°C. Increases in lignin contents were in the order of 28 > 50 > 65 >> 75°C. Protein content increases were ordered as 50 >> 65 = 28 >> 75°C. The results of Gerrits et a1. (1965) largely confirm Waksman et a1. (1939) results. 3.4.2.2: Nitrogen Transformations Initial accumulation of ammonia was inversely related to the rate of decomposition in fresh horse manure (Waksman et al., 1939). Whenever decomposition was delayed in the initial composting period, large nitrogen losses occurred. After 19 days the only traces of ammonia were found at 28°C and 50°C. Ammonia was present at 75°C for the entire experimental period, despite active volatilization due to high temperatures and alkaline conditions. Nitrate nitrogen began to appear in the 28°C and 50°C samples in 33 and 61 days, respectively. The 65°C samples had large accumulations of ammonia by the end of the experiment. Over the 61 day 39 experimental period, nitrate formation was greatest at 50°C followed by 28°C. Ammonia formation was greatest at 65 and 75°C. Burrows (1951b) analyzed the changes in nitrogen, phosphorus, carbon/nitrogen ratios, and pH during composting and mushroom cropping. In a companion paper (Burrows (1951a), he described his methods of analysis and conducted an analysis of errors due to sampling and variance in measurements of total ash, acid insoluble ash, moisture, total nitrogen and calcium in the compost. Burrows (1951b) found an increasing loss of nitrogen with nitrogen level without an increase in organic matter loss. Carbon/nitrogen ratios therefore narrowed during composting. Changes in pH occurring during the same time period indicated that ammonia volatilization was probably the cause of the nitrogen losses. Temperature was found to have a marked independent effect on nitrogen losses (Burrows, 1951b). Nitrogen losses plotted against the mean maximum temperature between turns indicate that stack losses are quite high at 70°C but almost negligible at 60°C. The retention of several nitrogen sources in a compost pile was evaluated (Burrows, 1951b). Differences between sources was highly significant with urea showing the greatest losses. Burrows speculated on the results of this experiment: ...Previous work, confirmed by the present 40 investigation, shows that a higher nitrogen content tends to increase the rate of formation of ammonia but a higher carbon-nitrogen ratio enable the ammonia formed to be reassimilated rapidly by organisms growing on the abundance of carbohydrate. The temperature effect may operate in two ways: first, by affecting the rate of bacterial transformation of proteins into ammonia. According to Waksman the rate appears to increase with rising temperature until an optimum in the region of 65°C is attained, when the rate decreases, no doubt owing to the adverse effect of the higher temperature upon the bacterial population. Secondly, a rise in temperature will cause a more rapid removal of ammonia, by physical agencies, from the neighbourhood of the growing bacteria in the compost heap, thus increasing nitrogen losses as the temperature rises. Low nitrogen composts had very high initial immobilization of ammonia (Beckwith and Parsons, 1980). Two thirds of the ammonia and half of the nitrate was mineralized and immobilized, respectively, in low nitrogen composts. All inorganic nitrogen was immobilized by day 10. The maximum incorporation of nitrogen into the organic nitrogen fraction occurred by the 20th day. Ammino acids showed similar patterns to the organic nitrogen values and were essentially constant after day 10. Amino sugars first appeared on day 10; maximum levels occurred on day 20 and 30 for the low and high nitrogen composts. Recovery of added fertilizer was highest in the low nitrogen (94 %) as opposed to the high nitrogen (88 %) compost (Beckwith and Parsons, 1980). In the initial stages of decomposition, there was a large increase in biomass. As activity declined, dead cells and cell contents provided a carbon source. Suzuki and Kumada (1977) concluded that nitrogen 41 transformation during the rotting process of rice straw compost involved ammonification followed by the simultaneous occurrence of nitrification and denitrification and finally nitrification. A spatial pattern of nitrogen transformation was also evident: the outermost layer underwent nitrification, the interior ammonification, and the intermediate layer had both processes occurring. 3.4.2.3: Effect pf Carbon Source The availability of the carbon source to micro- organisms in the composting process is as important as the nitrogen source. Mote and Griffis (1980) composted three different mixtures of materials with different carbon sources. Bulk densities and amounts of nitrogen, water, and carbon were held the same. Each compost exhibited different combinations of three composting rates. Gerrits et a1. (1965) found that the addition of amendments to horse manure compost that had high levels of readily available carbon caused rapid increases in compost temperatures. Amendments with the same nitrogen levels but less readily available carbon showed smaller increases. 3.4.2.4: Effect 93 Inorganic Materials Bretzloff and Fluegel (1962) studied the changes in several inorganic chemical constituents in a mushroom compost pile made of manure, corn cobs, hay and other supplements over a 30 day period. Moisture content, conductivity, pH, total nitrogen, Kjeldahl nitrogen, phosphorus, potassium, calcium, magnesium, sodium and 42 several ash fractions were determined from 9 locations along a compost windrow. Temperature and oxygen contents were measured at the same locations. The pile was turned 5 times during the 30 day composting period. Moisture content started at 52 % wb and increased to 72 % wb in 10 days, after which it leveled off for the remainder of the composting period (Bretzloff and Fluegel, 1962). Total ash content showed an initial jump from 20 to 25 % after the first 8 days and a linear increase to 35 % by the end of the 30th day. Acid-insoluble ash increased linearly from 10 % to 18 % over the composting period. pH showed an initial increase from 7.0 to 7.6 and an approximately linear decrease to 7.0 by the end of the composting period. The nitrogen, total phosphorus, calcium and magnesium increased during composting. Potassium and sodium contents were quite variable but showed small increases. Mineral amendments (bentonite, kaolinite and finely ground calcined aluminium oxide and ferric oxide) had little effect on the rate of decomposition in low nitrogen synthetic composts (Beckwith and Parsons, 1980). Kaolinite and bentonite produced a small accumulation of organic material on high nitrogen synthetic composts. The addition of ground gypsum improves aeration, water movement and drainage in composting due to flocculating action on the colloidal compost materials (Pizer, 1950). 43 3.5: Empirical Studies 3.5.1: Temperature Bartholomew and Norman (1953) carried out a series of experiments to study the influence of initial temperature on the rate of heat evolution. Decomposing straw was tested under adiabatic conditions at three separate initial temperatures. Lower initial temperatures had prolonged incubation periods. Maximum hourly temperature increments occurred when the temperature had risen 4 to 8°C above the starting temperature. McKinley and Vestal (1984) found that the maximal respiration for sewage sludge composting occurred in the range of 35 C to 45°C. Other studies have noted the temperature optimum for composting was in the range of 50 to 55°C (Finstein et al., 1980). 3.5.2: Oxygen and Carbon Dioxide Oxygen and carbon dioxide concentrations were measured in mushroom compost windrows (Lambert and Davis, 1934). Haug (1980) summarized data on oxygen uptake as a function of temperature (Figure 3.4). 3.5.3: Moisture Content Vigorous heating of a variety of substrates up to temperatures of 55 to 60°C when gravimetric moisture contents are between 40 to 60‘% wb has been reported by James et a1. (1928). Schultze (1961), Jeris and Regan (1973b) and Snell (1957) studied the relationship between oxygen uptake and the wet basis moisture content in 44 50 JERIS. et al CONTINUOUS. HIXED REFUSE. pH - 8.0 to 8.3 SCHULTZ 1: CONTINUOUS. 3 GARBAGE AND SLUDGE g SNELL ,- ‘ ' BATCH / \ m I > caouuo GARBAGE\/ \ e 10- / 3 I N 0 % SD— ~ CLARK ,° BATCH. __ WILEY SYNTHETIC REFUSE :: BATCH, O MIXED GARBAGE AND :- REF 5 JERIS. et a! a U E CONTINUOUS. 2 mxao aims: 8 Z O 0 top 0 CHELL :mo son: 2 CSNTI‘IUOUS. 3 an SLuacE >- 05— x O u 0 (U .— ( \ JERIS. er. JI \ JERIS. ct JI 0C 3ATCH. dATCH, CCHPOSTED MIXED 2EFUSE MiHSPRINT 0.1 l 1 1 4 ' 1 10 20 30 40 50 60 70 ‘ 80 TEMPERATURE. “C Figure 3.4: Observed oxygen consumption rates for various composting mixtures and reactor types as a funciton of temperature. Each curve represents the best fit of observed data (From Haug, 1980) 45 substrates of rewetted compost, refuse and ground garbage, respectively. This was summarized by Haug (1980) as shown in Figure 3.5. Below about 50 % wb moisture content \0 negatively affected oxygen uptake to the point where at 20 o wb almost no biological activity occurred (Glathe, 1960). Between 50 and 70 % wb, oxygen uptake rates were at their maximum; rates began to decrease at higher moisture contents. Haug (1980) indicates that low moisture contents can lower the rates of reaction because the bacteria that cause the composting action require an aqueous environment, and mass transport limitations for soluble components may be encountered under low—moisture conditions. Haug (1980) also notes the difficulty in isolating effects of moisture alone because of the relationships between moisture, bulk weight and FAS. Bartholomew and Norman (1953) studied the effect of initial moisture content on the temperature rise of decomposing straw under adiabatic conditions using forced aeration. The experiments were started at 25°C and a range of moisture contents (% db) from 75 to 275 % were examined. In the mesophilic range, the time-temperature profiles were similar; deviations were noticed in the thermophilic range. Microbiological activity was highest at higher moisture contents, as measured by heat production and decomposition. Increases in average moisture content in a naturally ventilated horse manure compost windrow lead to increases in 46 Aomma .momm EoLmv mmumuumnsm mcmumOQEou moofium> mo moumu cofiuQEDmcoo comaxo co ucmucoo muoummoe uo uommmm Ewommd .hzmpzoo $55.02 "m.m whomfim om on ok om on o4 on ON 0— n d 1 4 1 d 4 w ‘ 1 4 owu U 4. C 4 o D D n 4 4 4 4 4 ‘ 4 o 4 . 4 4 D “323 9323 Adz... D 4 4 4 one: .235. 92 254 4 O 4 4 4 52:8 Steam £525... 0 4 :4 o 0— ON on 0? .0m 00 ON on 00 00.. BLVH 3>W1dfl NEUAXO I’IDI'IIXVI'I :JO 1I~133t13d 47 the carbon dioxide content at locations in the windrow (Gerrits, 1972). Very wet composts using the same substrate exhibited slower temperature rises and higher maximum temperatures than drier mixtures. These trends were attributed to delays in microbial activity and the larger thermal mass present in the wetter compost, respectively. Moisture contents of between 70 to 72 % wb were associated with aerobic (5-20 % 02) and micro-aerophilic (0-5 % 02) conditions in the largest part of the heap. Observations of lower moisture content (68 %) composts indicated that they were heavily aerated with cold air. Temperatures were observed to be lower and decreased quickly after maximum temperatures were reached. Gerrits (1972) concluded that approximately 75 % of the heat generated through the combustion of dry matter by micro-organisms is used for evaporation. Some substrates, such as cotton gin trash dry out very quickly and need intensive water management if decomposition is to proceed in a rapid manner (Mote and Griffis, 1980). 3.5.4: Windrow Size Size reduction in a domestic refuse windrow from 10 ft x 20 ft to 5 ft x 13 ft improved natural aeration at the expense of increased heat loss (Horstmann and Ehgelhorn, 1969). Strom et al. (1980) recommended a leaf composting strategy that utilized small piles in the fall to promote aeration. Two of the small piles were combined into one larger windrow in the winter to retain heat. Proportions of 48 the compost subject to acid, anaerobic and non-reactive zones increased with the height and width of the windrow (Lambert and Davis, 1934). 3.5.5: Windrow Zones Lambert (1941) presented a detailed description of different zones commonly encountered in mushroom compost heaps (Figure 3.6). The patterns in physical conditions were dependent on the size, shape and compactness of the windrow and changed over time. Zone A is an outside layer Figure 3.6: Section through a mushroom compost heap illustrating the typical differences in temperature and aeration (From Lambert, 1941). that has temperatures varying from ambient temperatures to 43°C. This zone is well aerated. The sides can become very dry due to excessive aeration and the top wet in wet or cool weather. Zone B is moderately moist, well aerated and has temperatures from 43 to 60°C. The third zone, C, extends like a huge doughnut around the windrow 2 to 4 feet from the 49 sides and 3 feet from the top. The highest temperatures are found in this zone: 60 to 82°C. The fourth zone (D) is anaerobic and occupies the entire lower central part of the windrow. 3.6: Kinetic Modeling 3.6.1: Microbial Growth Whang and Meenaghan (1980) developed a kinetic model of the composting process. They assumed that enzyme kinetic concepts were applicable to the development of the composting model, that an intermediate complex of micro- organisms and substrate was formed under a quasi-equilibrium state and that endogenous reaction is irreversible. An equation similar to the Michaelis-Menten equation was derived: R = K2 (C) / (K1 + (C)) (3.4) where R = consumption of substrate Kl = (k_1 + k2) / kl = su strate XT = total micro-organism concentration = X + CX* X = free micro-organism CX* = activated substrate-organism complex kl, k-l' k2 = specific reaction rates Values of K1 and K2 were determined from an experimental compost reactor using humidified forced aeration with daily turning in a batch mode. Haug (1980) discussed the general sequence of events in the catabolism of solid substrate by composting micro— organisms. Some of the steps included: 1. Release of extracellular hydrolytic enzymes by the 50 cell and transport of the enzymes to the surface of the substrate; 2. Hydrolysis of substrate molecules into lower molecular weight, soluble fractions; 3. Diffusion transport of solubilized substrate molecules to the cell; 4. Diffusion transport of substrate into the microbial cell, floc or mycelia; 5. Bulk transport of oxygen (usually in air) through voids between particles; 6. Transport of oxygen across the gas-liquid interface and the unmixed regions which lie on either side of such an interface; 7. Diffusion transport of oxygen through the liquid region; 8. Diffusion transport of oxygen into the microbial cell, floc or mycelia; and 9. Aerobic oxidation of the substrate by biochemical reaction within the organism. Haug (1980) presented a derivation of the substrate consumption and microbial growth equations based on single substrate limitations. A form of the Monod equation was used. Microbial growth can be related to substrate use by the equation: dx / dt = Ym(-dS/dt) - kex (3.5) where dX/dt = net growth rate of microbes, mass/volume- time Ym = growth yield coefficient, mass of microbes/mass of substrate ke = endogenous respiration coefficient, l/time or mass of microbes respired/mass of of microbes-time X = concentration of microbes, mass/volume maximum utilization coefficient, maximum rate of substrate utilization at high substrate concentration, mass substrate/mass microbes-day Under substrate limiting conditions the combined growth equation can be derived: dX/dt = Y (kmSX/(Ks + 5)) - k x (3.6) m 01' (dX/dt)/X YmkmS/(KS + S) - ke (3.7) where (dX/dt)/X = net specific growth rate 51 k maximum net specific growth rate half velocity concentration, mass/volume (DI/)5 concentration of rate limiting substrate, mass/volume The four kinetic coefficients, Ym, km, Ks and ke need to be known for a specific substrate and microbe combination. Haug (1980) presents data on a range of biological processes. Sinclair and Ryder (1975) compared the ability of two different growth models to explain continuous culture data under both carbon and oxygen limiting conditions. The interacting substrate model postulates an interaction between two substrates and is based on double enzyme kinetics. Only one of two substrates would enter into the model at any one time in the alternate substrate model. Both Monod/Monod and Monod/Contois expressions worked well in both models. The models were able to satisfactorily explain experimental results. 3.6.2: Microbial Death Haug (1980) developed a first-order decay model for inactivation kinetics which is often refered to as "Chick’s Law." The most common form is given as dn / dt = -kdn (3.8) viable cell population thermal inactivation coefficient where n kc1 If kd is a constant, integration of the above equation yields nt = noexp(-kdt) (3.9) where nO = initial cell population 52 nt = later population at time t The temperature effect on kd is typically modeled by the Arrhenius form kd = C exp(-Ed/RTk) (3.10) where C = constant Ed = inactivation energy, kcal/mole Tk = temperature, K Inactivation energies for many spores and vegetative cells range between 50 and 100 kcal/mol (Bailey and Ollis, 1977). By taking the logarithim of the above equation, a plot of the log of kd vs. l/Tk from survivor plot data can be used to find C and Ed (Haug, 1980). Since the temperature of the compost changes with time, kd will not be constant. Combining Equations (3.8) and (3.10) gives dn / dt = -c n expi-Ed/RTk(t)]n (3.11) where Tk is a function of time, Tk(t). This expression can be used to evaluate the kill resulting from various time- temperature profiles by separating the variables and integrating from the intial to the final conditions: 1'. 1n(no/nf) = f c exp [-Ed/RTk(t)] dt (3.12) t o where the subscript 0 indicates initial and f final conditions. Haug (1980) noted that integrations for hyperbolic, exponential and linear time-temperature functions have been performed, in particular by Bailey and Ollis (1977). In composting situations, the time- temperature curves may not follow profiles for which 53 analytical solutions are available. Graphical (Haug, 1980) or numerical procedures could be used to calculate microbial death due to temperatures in those cases. Large clumps of compost particles have been observed to form during composting. Haug (1980) examined the effect of this solids clumping on microbial destruction. He assumed a spherical homogeneous ball of compost material heated from the outside and thermal and physical properties based on those for water and compost. Using these assumptions he calculated the length of time that it would take for the interior of clumps of various sizes to heat to 0.90 times the ambient temperature. Heating times for particles of between 1 and 10 cm radius were found to be negligible; only when clumps were greater than 20 cm radius did heating times become significant. Nonuniform temperature distributions can affect the overall thermal deactivation of pathogens. The effects of these nonuniformities have been examined for windrow (Haug, 1980) and static pile systems (Haug, 1980; Burge et al., 1978). For windrow systems, thermal inactivation can be described as: and nt = noifl + fhexp(-det)]N (3.13) f1 + fh = 1 (3.14) number of organisms surviving where nt number of organisms initially present 0 . . . . fl fraction of composting material in the low- temperature, sublethal zone fh = fraction of composting material in the high- temperature zone At = time interval between pile turnings 54 thermal death coefficient Rd N number of pile turnings Two different fl/fh values and several det values were evaluated. The fraction of surviving organisms was much greater than that Calculated by examining a specific time- temperature curve. Haug (1980) notes that this indicates that the exposure of all portions of a windrow to lethal temperatures is as important as achieving time-temperature curves at any one location in order to increase the average ‘thermal inactivation. Furthermore, the above analysis is based on the assumption that all particles are randomly mixed during the turning process. As Haug notes, this may be true if commercial turning devices are used. If a front-end loader is used, however, complete mixing will probably not occur. The loader can be used, however, to place portions of the windrow that need additional heat treatment in zones that should receive higher temperatures. If only a small portion of material on the outside of the pile comprising 1 % of the total compost mass escapes mixing, the average thermal inactivation would be limited to no more than 3 Dr (Haug, 1980). Burge et al. (1978) measured the temperature distributions in 15 static piles of sewage sludge and wood chips using forced aeration and used this data to predict the confidence levels for achieving a particular time- temperature relationship. The number of days that different 55 Table 3.7: Confidence levels that all material will obtain a temperature equal to or greater than a particular temperature for a desired number of days. Confidence Levels (%) Temperature 95 99 99.9 (°C) (Days) (Days) (Days) >50 13.8 13.3 12.6 >55 10.6 10.1 9.4 >60 7.3 6.8 6.3 >65 4.3 3.9 3.4 >70 1.2 1.0 0.8 temperatures could be reached are shown in Table 3.7. Burge et a1. (1978) also examined the mean and standard deviation of temperature in the toe area of the 15 test piles. Examining the data, Haug (1980) notes that Significant reduction of bacteriophage f2 would be predicted not only on the basis of the mean temperature but also the temperature obtained by subtracting the standard deviation from the mean. The confidence levels shown in Table 3.7 apply only to the 15 test piles as they will vary with the substrate material, bulking agent and manner of operation (Haug, 1980). 3.6.3: Regrowth Coliform and fecal streptococcus bacteria have been observed to regrow in sterile liquid sludge heated to 35°C (Brandon, 1878). The most dramatic regrowth of fecal streptococcus bacteria was observed in sterilized material that was then recontaminated (Ward and Brandon, 1977). Occaisional regrowth of coliform organisms was observed during windrow composting of sewage sludge, particularly in 56 wet winter (California) conditions (Selna and Smith, 1976). 3.6.4: Combined Equations Haug (1980) proposed a net rate coefficient equation based on empirical expressions from experimental data to describe the overall kinetics of sewage sludge composting: kd = (Fl)(F2)(F02)kdm (3.15) where Fl = moisture content correction F2 = free air space correction F02 = oxygen content correction kdm = maximum rate coefficient determined by substrate and temperature conditions Haug assumed that the oxidation of biodegradable volatile solids (BVS) is first-order with respect to BVS quantity: d(BVS) / dt = -kd(BVS) (3.16) The distinction between BVS and nonbiodegradable volatile solids is made to distiguish between materials that are readily degradable and those that take longer to degrade (Haug, 1980). Haug assumed that the rate constant kd was a function of temperature only. Based on studies of garbage composting by Schultze (1962) that determined oxygen consumption, Haug (1980) developed the following empirical equation: . kd = 0.00632 (1.066)’1"’20 (3.17) Where T is in degrees Celsius and kd (g BVS oxidized/g TVS- day). Heat inactivation due to increased temperatures was accounted for in Haug's models using an expression developed 57 by Andrews and Kubhu (1973) to describe similar effects during aerobic digestion of liquid wastes: _ (T-TRl) - (T-TRZ) kd - de1 [C1 C2 ] (3.18) where kd 1 = rate constant at temperature TRl' day“l Cl,82 = temperature coefficients TRl'TRZ reference temperatures, C Equations 3.17 and 3.18 were combined and reasonable values were assumed to arrive at the effect of temperature on kd (Haug,l980): kdm = 0.0126 [1.066(T‘20) - 1.21‘T‘50)] (3.19) The effect of moisture content on oxygen consumption rates and rate of BVS oxidation was estimated with data from Schultze (1962), Jeris and Regan (1973b) and Snell (1957). An "S" shaped curve was fit to the data by two equations for different moisture content ranges. The effect of low free air space (FAS) on oxygen transport was modeled by Haug (1980) using data from Jeris and Regan (1973b) for refuse compost. Thirty percent was the optimum FAS in terms of oxygen uptake rates. Two separate equations covering the range of FAS values were fit to the data. FAS effects were considered to be important only for those composting systems that did not use bulking agents. As long as some FAS was present, Haug (1980) felt that it would have a negligible effect on reaction kinetics. Knowledge of the bulk weight of the compost material was necessary to calculate the FAS. Oxygen content effects on compost reaction kinetics were felt to be very complex by Haug. The effect of free 58 air space oxygen concentration was assumed to follow a Monod-type expression: F02 = VOLP02 / (VOLP02 + 1.0) (3.20) where VOLP02 is the volume percent oxygen in the FAS. The effect of oxygen concentrations above 5 % were minimized by the assumption of a half-velocity constant of 1.0. 3.7: Physical Properties 3.7.1: Particle Size Chang and Rible (1975) analyzed fresh, deposited and composted livestock wastes to determine their particle size distribution and to characterize each size fraction in terms of its value as fertilzer, feed suplement and fuel. Composted dairy wastes showed a higher concentration of large fibrous materials than fresh waste. Moisture contents of larger size fractions remained the same while smaller sized particles showed large decreases. Crude fiber and protein decreased by nearly 50 % in most size fractions. Fat content decreased in all particle sizes. Ash contents increased by 75 to 100 percent. Mears et al. (1975) analyzed the particle size of a number of different compost mixtures. They had problems with measuring the particle size of material with greater than 45 % wb. The method of drying affected the particle size determinations for this material. Particle size measurement of material that was initially drier than 45 % wb was not affected by drying method. Particle size followed a log normal distribution. Geometric mean diameter was 59 determined for each windrow on a weekly basis and decreased over time. The rate of decrease varied between the substrate mixtures and the reported work did not consist of sufficent experimental windrows to draw conclusions about any one substrate. 3.7.2: Particle Density Chen (1982) reported mean values of beef cattle manure dried solids density of 1524 kg/m3. The average mean particle density of dairy manure solids obtained from three separate farms was 1551 kg/m3 (Bohnhoff and Converse, 1986a). Regardless of the source, composted material had mean particle densities that were significantly greater than the mean particle density of freshly separated solids. Bohnhoff et a1. (1984) and Bohnhoff and Converse (1986a) present two equations for determining particle density. The earlier equation is a totally empirical linear regression on volatile solids. The later equation is derived from the definition of mean dry particle density with terms for the densities of volatile and fixed solids obtained by nonlinear least squares regression. 3.7.3: Bulk Density Mixtures of swine waste and refuse were initially between 275 and 500 kg/m3 and increased to between 650 and 850 kg/m3 after 27 days of composting (Mears et al., 1975). 3.7.4: Derived quantities When bulk density, particle density and gravimetric moisture content are known a number of other quantities can 60 be calculated (Hillel, 1982). These include the porosity ( f ), void ratio ( e ), volumetric moisture content ( 8 ), degree of saturation ( s ) and air filled porosity ( fa ) or free air space (FAS). 3.7.5: Compressibility Based on a method developed for evaluating silage materials, Mears et a1. (1975) determined the compressibility of swine waste based compost materials. They found that a relationship between bulk density, Oh, and applied axial stress, a, adequately described the compressive properties of the composted materials: 1n 0 = 1n 00 + C 1n 0 (3.21) where 1n 00 and C are regression coefficients formed by the statistical analysis of the data. 3.7.6: Settlement Behavior Stentiford et a1. (1984) reported changes in 1.7 m tall triangular windrow cross-sections for mixed refuse composting by the static pile method. The maximum change in the pile height came in the first 5 days with 60 percent of the 36 cm total decrease. Mears et a1. (1975) measured windrow volume reductions in swine waste based windrow composts. Reduction in windrows composed entirely of swine waste ranged from 8 to 16 % after 10 days and 23 to 45 % in 40 days. A mixture of swine waste and straw had a reduction of 28 % after 50 days. The time-settlement behavior of milled urban refuse under saturated conditions has been studied by Chen et al. 61 (1977). A mathematical model for solid waste settlement was developed by Zimmerman et al. (1977). It consists of two simultaneous equations, one of which is non-linear. The effects of finite strain, biological and chemical activity, and the time variation of saturation have been included. 3.8: Heat Transfer 3.8.1: Diffusion 3.8.1.1: Thermal Conductivity Thermal conductivity is a basic heat transfer property of a material. Materials can be considered conductors (high conductivity, high heat transfer) or insulators (low conductivity, low heat transfer). The material in question can consist of a single phase (ie. gas, liquid or solid) or can be a combination of phases. Each phase can consist of a mixture of materials, as well. Heat conduction in gases, vapors and liquids depends largely on the molecular transfer of the kinetic energy of molecular movement (Karak and Yener, 1979). Temperature and pressure have an important influence on the thermal conductivity of these substances. Liquids with their closer molecular spacing have much higher thermal conductivities than gases. Thermal conductivties of mixed state materials are affected by the properties of each state and the way in which they are combined. Based on the structure of the solid state, materials can be classified as fibrous (textiles, fiberglass wool, straw), granular (powders, coal, 62 grain) or cellular (cork, foam insulating boards). Factors affecting thermal conductivity of such material include the (l) conductivity of the component material, (2) pressure of interstitial fluid, (3) temperature, (4) bulk density, (5) particle size, (6) particle size distribution, (7) void space and porosity, and (8) moisture content (Chun-Yung Chen, 1969). Simple models of thermal conductivity relate the porosity and conductivity of the two phases as separate resistances arranged in series and parallel (Pratt, 1969). Other theoretical models are based on the porous material having a continuous solid or continuous air phase (Hillel, 1981). Pratt (1969) presented experimental data showing the variation in effective thermal conductivity due to the presence of different gases in a fibrous insulation. Helium containing insulation had five time the thermal conductivity of air while the thermal conductivity of insulation filled with C02 was only 72 % of air. The thermal conductivity of the solids in the matrix has a direct effect on the effective thermal conductivity. Temperature changes effect the apparent conductivity through two mechanisms: the thermal conductivities of the components and the contribution of radiation (Chen-Yung Chen, 1969). Thermal conductivities of gases increase with temperature. Some liquids, such as water, show the same trend. In addition to the purely conductive effects, actual 63 heat transfer in heterogenous materials can be by convection or radiation as well as by conduction. Depending on the method or scale of measurement, these transfer modes are sometimes included in an apparent or effective thermal conductivity. At typical composting temperatures, the radiation effects are minute. Increases of fibrous material bulk density increase thermal conductivity. Pratt (1969) observed that many researchers have shown that the thermal conductivity in low bulk density fabrics with porosities of about 90 percent are almost independent of the component fiber. Allcut (1951) performed a series of measurements of apparent thermal conductivity for various material and densities. A characteristic hook shape is shown in all thermal conductivity curves (Figure 3.7). The cause for this shape is heat transmission by convection. Increasing densities Jud, :JLJ I" -‘ Thermal (cnduclmly I04 I .‘J (m.| act)" Hm 3....21 U'U LIIIH‘ f" O 5’.) ICU E.) Cox) Li. J C8PSIT‘,‘. h ;""3 Figure 3.7: "Apparent" thermal conductivity of various materials at different densities. A, kapock; B, baggasse; C, cork board; D, slag wool; E, mineral wool (From Pratt, 1969). 64 reduce the convection heat transfer contribution to the apparent thermal conductivity by dividing the air layers into smaller layers, thus increasing the medium's tortuosity and increasing it's resistance to fluid flow. After a certain density is reached, however, the apparent thermal conductivity begins to increase. The exact density varies between materials and is dependent on, among other things, the ratio KS/Kg of the solid to the gas thermal conductivity. Deissler and Boegli (1958) found that the effective conductivity of a void was strongly influenced by material arrangement for high KS/Kg. Chun-Yung Chen (1969) visualized this as being due to the contact point between the solid particles. Gas is a relative insulator in high KS/Kg systems, so most heat flow takes place near the contact points. For values of Ks/Kg of approximately 1000, nearly all the heat transfer takes place near the point of contact (Deissler and Boegli, 1958). Changes in bulk density increase the number of contact points, leading to greater contributions of conductive heat transfer to the overall apparent thermal conductivity. A bulk material contains two types of gas spaces: intra- and inter-particle. If the gas has a lower thermal conductivity than that of the solid, intra-partical gas lowers the particle effective conductivity. The cells can be large or small, closed or open. Closed cell material voids are essentially air and vapor tight, while open cell voids are interconnecting and permit free movement of air 65 and vapor through the material (Pratt, 1969). The effect of inter-particle gas space on the effective thermal conductivity depends on the characteristics of the system. At very low bulk densities, with large porosities and pore sizes, convective heat transfer is large and important. As overall porosity and pore size decrease, convective heat transfer decreases. This leads to a decrease in the effective thermal conductivity. It is important to note that decreases in pores sized can also lower thermal conductivity by increasing the tortuosity and resistance to convective heat transfer (Verschoor and Greeber, 1952). The influence of latent heat transfer by water vapor in air-filled pores can be significant (Hillel, 1981; Pratt, 1969). Latent heat transfer effects can be taken into account by the additon of an apparent conductivity due to the evaporation, transport, and condensation of water vapor to the thermal conductiviy of air (Hillel, 1981). This value is strongly temperature dependent and rises rapidly with increasing temperature (Figure 3.8). At about 55 to 60°C it is approximately the same value as the thermal conductivity of liquid water (Pratt, 1969). Mears et a1. (1975) calculated thermal conductivity for composts consisting primarily of swine wastes from previously determined values of thermal diffusivity, bulk density and specific heat: 66 I I l I I 7 4‘“ $6- 7 E g - «- 2g._e..flk "g 4‘. ‘\-A°+er p'va or A. 2‘ IE I} 3 ‘8 8 2' 2 :- ‘0 . 1 1 I L I IO 20 30 4O 5O 60 TO 80 Ternpevotuve.°C 0 Figure 3.8: Effective conductivity of porous material due to latent heat transfered by movement of water vapor (From Pratt 1969). K = 0.1163 a p C (3.22) P Statistical analysis of the data indicated that thermal conductivity varied linearly with moisture content. Thermal conductivity at any given moisture content increased significantly as the compost matured. Houkom et a1. (1974) and Chen (1983) measured the thermal conductivity of beef cattle manure. Chen developed three separate regression equations for different manures that related thermal conductivity to bulk density, total solids and porosity. Bohnhoff and Converse (1986a) measured the thermal conductivity of dairy manure solids at 2 different moisture contents, three temperatures and three bulk densities. A four parameter equation relating thermal conductivity to temperature and volumetric moisture content was selected. 67 3.8.1.2: Specific Heat Specific heat has been found to vary linearly with the wet basis moisture content for swine waste (Mears et a1, 1975) beef cattle manure (Houkom et al., 1972; Chen ,1982) and separated dairy manure solids (Bohnhoff and Converse, 1986a). 3.8.1.3: Thermal Diffusivity The thermal diffusivity of a swine waste compost was determined by Mears et a1. (1975) using the transient method. Thermal diffusivity of dairy manure solids was calculated by Bohnhoff and Converse (1986a). 3.8.2: Natural Convection Principles of convection through porous media are discussed by Bejan (1984). Haug (1980) developed a simple natural convection model to assess the ability of this process to supply oxygen at stoichiometric rates. 3221 Mass Transfer The spatial scale in a composting system is associated with different types of mass transfer processes. Gas and vapor transfer in windrow voids could occur by diffusion or natural convection. Liquid transfer in void spaces could occur due to saturated or unsaturated flow. Gas transfer from the free air space across the liquid film to the substrate-microbe complex is by diffusion. 68 3.9.1: Windrow Void Space Transfer 3.9.1.1: Diffusion Experimentally determined values of diffusion coefficients in gases at one atmosphere are shown in Table 3.8. The binary diffusion of gases can be predicted for given temperatures and pressures using the Chapman-Enskog kinetic theory or empirical correlations (Cussler, 1984). The Chapman-Enskog theory assumes nonpolar gases and this excludes water and ammonia (Cussler, 1984). Diffusion of gases vary with the 1.5 to 1.8 power of temperature (Cussler, 1984). Table 3.8: Experimental values of diffusion coefficients in gases at one atmosphere. Diffusion Gas Pair Temperature Coeficient (°K) (cm2 sec") Air--C02 276.2 0.142 Air--02 273.0 0.1775 Air--H20 289.1 0.282 298.2 0.260 312.6 0.277 333.2 0.3050 Source: Cussler (1984). Shell (1955) studied the diffusion rate of oxygen through a ground garbage compost. The relative diffusion rates were shown to be linearly related to the free air space of the material. Increases in moisture content decreased the relative oxygen diffusion rate. Increases in 69 bulk density decrease the relative diffusion rate. Hillel (1981) discussed a number of experimental efforts to relate the diffusion of gases in soils to the diffusion rate in air and derived a general equation for transient diffusion in the soil. Equimolar counterdiffusion occurs when the same number of moles of two gases in separate reservoirs diffuse towards one another (Geankopolis, 1983). Finger (1975) assumed that the equimolar counterdiffusion of oxygen and carbon dioxide accounted for mass transfer in his distributed heat and mass transfer model of a compost pile. Diffusion in many systems involves more than one component. In composting systems, observations of steaming indicate that water should be considered in addition to ongen and carbon dioxide. Multicomponent diffusion is estimated by converting the problem to a binary problem, solving it and then converting back to a multicomponent solution (Cussler, 1984). Diffusion of gases in porous solids and capillaries is discussed by Geankopolis (1983) under Fickian, Knudsen and transition regimes. Isothermal diffusivity of water vapor with respect to soil volumetric moisture content was derived by Jackson (1964). 3.9.1.2: Evaporation The drying rate of biological products with initial moisture contents above 70 to 75 % wb can be constant if external drying parameters such as air velocity, air 7O temperature and air humidity are constant (Brooker et al., 1981). Constant rate drying will be observed under constant external conditions when the internal resistance to moisture transport is much less than the external resistance to water vapor removal from the surface of the product (Brooker et al., 1981). Brooker et a1. (1981) presents an expression for the constant rate moisture loss for biological products. The surface area and either the heat or mass transfer coefficents must be known for the constant drying rate to be calculated. Bohnhoff and Converse (1986b) developed desorption isotherms for water desorption equilibria at five temperature levels. Three isotherm models were fit to the data; the relationship between moisture content, relative humidity and temperature was best described by the four parameter Chen-Clayton equation. Equations for estimating the isosteric and integral heats of desorption over a temperature range of 0 to 70°C were also developed (Bohnhoff and Converse, 1986b). 3.9.1.3: Liquids Hillel (1981) discusses the concept of hydraulic diffusivity to describe the convective transport of water in soil pores. Hydraulic diffusivity is the ratio of hydraulic conductivity to the specific water capacity. All three terms can be written as functions of the volumetric moisture content. An equation for the simultaneous transfer of both 71 liquid and vapor can be derived (Hillel, 1981). 3.9.2: Transport in Liquid Films Haug (1980) has presented a conceptual illustration of mass transport and reaction in compost particle water films. This is shown in Figure 3.9. Consumption of oxygen by microbes causes a concentration gradient and oxygen diffuses from the free air space into the substrate-water-microbe matrix. Aerobic microbial metabolism consUmes the substrate and oxygen. Microbial mass is synthesized and carbon dioxide, heat, water, and ammonia are produced. The metabolic end products are at elevated concentrations in the liquid phase and will diffuse toward the airspace. Diffusion coefficients in a gas are about 105 times greater than in a liquid. Due to higher liquid vs. gas concentrations, however, the flux in a gas is not that much greater, being only about 100 times faster. (Geankopolis, 1983). The slowness of diffusion in liquids often limits the overall rate of processes that occur in liquids (Cussler, 1984). Some experimentally determined diffusion coefficients are presented in Table 3.9. Cussler (1984) presents a number of methods for calculating diffusion in liquids and compares them for diffusion of oxygen in water at 25°C. Diffusivities in liquids are often dependent on the concentration of the diffusing components. Oxygen diffusion in bacterial slime layers can be significantly lower than in water, sometimes as low as 0.04 x 105 (Bailey and Ollis, 72 H- H20 In Iced Suburuo Ralouohon Endoconouu Baton-lion On flow In In. an lane. 02 NM‘HCOJ NO; 0.. PthI <-— ——- Subllnu. WIIor, Microbe Mull: Figure 3.9: Conceptual illustration of mass transport of major components during composting (From Haug, 1980). 73 Table 3.9: Diffusion coefficients at infinite dilution in water at 25°C. Solute D (-10’5 cmZ/sec) Carbon dioxide 1.92 Oxygen 2.10 Ammonia 1.64 Source: Cussler, 1984. 1977). Haug (1980) used a simplified model of gas transfer in a saturated matrix of solid substrate and microbes to estimate the effect of particle size on oxygen flux and the time required to satisfy the Stoichiometric oxygen requirement. He concluded that diffusion can match the oxygen consumption rate if the particle size is sufficiently small. Particles thicker than 1.0 cm would have large diffusional resistances that could dominate the process kinetics. Particles of about 0.10 cm would be small enough for diffusion to meet demand. Oxygen diffusion would no longer exert control over the overall composting rate if particles are less than 0.05 cm in diameter. 3.10: Complete Heat and Mass Transfer Models The method of volume averaging was used to derive the governing equations for heat and mass transport in a rigid medium by Ryan et al. (1981). Excellent agreement was found between theory and experimental work using a spatially perodic model of a porous medium for conductive and 74 diffusive transport. Latif and Lissik (1986) developed a non-distributed respiration model for heat and gases released during grain storage. The rate equation was a function of initial temperature and moisture content, and grain damage. Chau et a1. (1984) presented a numerical model for heat and mass transfer in spherical products. The effects of respiration, transpiration, conduction, convection and evaporative cooling were included. Haug (1980) developed comprehensive simulations of continuous-feed completely mixed and batch models of the composting process. Finger (1975) developed a distributed heat and mass transfer model for windrow composting that was based on thermal heat conduction and the diffusion of oxygen into the compost pile. CHAPTER 4 EXPERIMENTAL METHODS 4.1: Data Requirements A windrow system without enhanced natural convection or forced aeration was selected for study. This system was selected because it closely resembled current dairy manure solids (DMS) handling practices: simply piling and turning the separated DMS before using it as a bedding material. Based on a review of the literature, the data which must be used to validate a time-dependent spatially- distributed finite-element model capable of predicting the time/temperature histories in a compost windrow include the following: temperature, gas concentrations (02 &/or C02), moisture content, volatile solids, bulk density, porosity, air-filled porosity, and windrow size change. The last four items were measured in core samples; the first two were measured using a probe, the design of which is described below. Table 4.1 gives a summary of the sampling methods and schedule for each variable. 4.2: General Experimental Design and Methods The project was located at the dairy facility of the Kellogg Biological Station of Michigan State University. The facility, constructed in 1985, is a research and 75 Table 4.1: 76 Sampling methods and schedule. Variable Method Sampling Date 1 2 3 4 5 6 7 Temperature: Windrow: Slab: Air: Rel. Humidity: Gas Concentration: Core Samples: Windrow size & Shape: * Campbell Scientific CR21X datalogger and AM32 multiplexer; 9 locations: Windrow 1-3 reps. averaged, Windrow 2-3 reps. recorded; 1 minute sampling interval, averaged each 15 minutes. * Same instrumentation; 1 location, 3 reps. averaged; same sampling interval. * Same instrumentation; 1 location; same sampling interval. * Same instrumentation; 1 location; same sampling interval. * Probe sampling, evacuated cylinder storage, analysis for OZ and C02 with gas Chromatograph; 10 locations 3 replications. * Rotary corer; moisture content, volatile solids, coliform number, bulk density, porosity, free air space; 9 locations. * Windrow height, 30 to 36 locations; windrow height at probes, 12 locations; see text for method details. XXXXXXXXXXXXX XXXXXXXXXXXXX XXXXXXXXXXXXX XXXXXXXXXXXXX X X X X X X X X 77 demonstration dairy farm that includes a 192 cow free stall dairy and milking parlor and a variety of support buildings. Four times each day, manure and spent bedding were flushed from free-stall alleys into a settling tank. The tank contents were agitated daily and the slurry was pumped over an inclined separator. The manure solids that came from the separator were then conveyed to a series of 4.5 m x 4.5 m bins inside a naturally ventilated building. The separated dairy manure solids (DMS) were allowed to accumulate for a week. Table 4.2 summarizes the physical characteristics of the dairy. Table 4.2: Dairy characteristics. Number of cows: 192 Type: Holstein-Friesian Housed in free stalls Four stall alleys Four feed alleys Waste system: Flushed 4 times per day with tip and gated tanks 5 % slope Well water and recycled lagoon water Stationary screen separator Conveyer Dual lagoon Bedding: 1/3 composted DMS--l/2 skid steer bucket per week per stall 2/3 chopped straw When an experimental windrow was constructed, a batch of accumulated DMS was removed by a "skid steer" bucket loader, placed in a silage mixer, agitated, weighed and dumped on the ground. Samples for analysis of moisture content, volatile solids and coliform count were taken from 78 the mixed windrow. The experimental windrow was constructed across a bin using a skid steer to place the DMS. Great care was taken to assure uniform placement. The skid steer operator was directed where to put each bucket load. Each bucket was placed so that the material fell from a uniform height of 46 cm (18 inches) above the windrow surface. The skid steer bucket was not used to shape the windrow: all final shaping was done manually with a shovel. All windrows had a trapezoidal prismatic shape. The windrow was taken apart weekly, mixed in a silage mixer, weighed and formed into a new windrow using the methods described above. Exact uniformity of windrow size and shape was not possible between turnings because of losses due to composting and the difficulty of windrow construction. Each batch of solids was treated for 3 weeks. Two separate batches were monitored: one during the summer and the other during the fall of 1986. 4.3: Probe Design and Placement Methods of measuring temperature and gas concentration with probes in natural convection/diffusion based composting systems must meet several criteria. Accurate measurement must be possible. Probe placement and operation should disturb the windrow as little as possible. The probe should be sufficently rugged to withstand rough handling as well as the corrosive compost environment. Accurate measurement of temperature levels and gas concentration are covered more extensively in the 79 appropriate sections. In general, temperature and gas samples must accurately reflect the actual temperature and gas concentration at the sampling location. The thermal conductivity of the probe, for instance, must be small enough that heat from neighboring high temperature zones is not conducted along the probe to lower temperature locations, leading to an inaccurate, high temperature measurement. Similarly, the gas sampling process should not draw gas from areas other than the intended sample area. The method of probe placement can affect physical properties such as the bulk density and porosity in the area around the probe. This can in turn affect the heat and mass transfer which is being studied. If the probe is forced into the windrow, the surrounding DMS are compacted. If a hole for the probe is drilled into the windrow, the less dense material may settle away from the probe, exposing the probe to ambient conditions; this can be a serious problem with long term windrows (Mote, 1986). Placement of probes while the windrow is being made could lead to areas of the windrow "downstream" from the probe having different bulk densities than those on the "upstream" side. The timing of probe placement must also be considered. Probes can either be inserted into the windrow whenever a measurement is made or they can be left in place for the duration of the study. Measurement on demand reduces both the number of probes that are needed, as well as their complexity. Large homogeneous windrows are required, 80 however, to permit probe insertion at a new location at each time without affecting the process. In-place probes allow many measurements to be made over time in the same location. Operation of the measurement system should not affect the heat and mass transfer process in the windrow. This is particularly true for gas sampling. Suction of large volumes of gas draws outside air into the windrow. In forced aeration or suction composting systems this is not important. In natural convection or diffusion systems, however, this could greatly change the heat and mass transfer in the windrow, if only for a while. There have been many attempts to develop probes for sampling temperatures and gas concentrations. Finstein (1980) placed a thermocouple in a slotted wooden dowel to reduce resistance to the probe placement. While studying forced aeration composting systems, Singley et a1. (1982) used a probe that consisted of a perforated cylindrical diffusion chamber on the end of a pipe to measure both gas concentrations and temperatures at the same point in a windrow. A thermocouple was placed in the diffusion chamber. The probe was placed while the windrow was being built; air samples were removed as needed for analysis. Finger (1975) used a dissolved oxygen electrode mounted in a diffusion chamber at the end of a metal pipe. The probe was inserted into the compost windrow and the oxygen concentration measured at equilibrium. Minimization of windrow disturbance is of prime 81 importance in this study; probes are therefore left in place. The probes are rigid, are inserted into the windrow after the windrow is formed and then left in the windrow, and permit both measurement of temperatures and sampling of gasses in the windrow. The probe used in this study is shown in Figure 4.1.. It allows gas collection and measurement of temperature at several points along its length. It consists of alternating sections of slotted wooden dowels with perforated copper tubing. The wooden sections reduce conduction along the probe and are not structurally affected by the temperature differences in the compost windrow environment, as is the case with materials such as plastic tubing. The perforated copper pipe acts as a diffusion and gas storage chamber. Dowel and copper pipe outside diameters are matched to reduce resistance to probe insertion and prevent oversized holes from being formed. The tip of the probe is turned to a point. Depending on the number of measurement locations on a probe, between one and four slots are cut in each dowel section. Thermocouple leads and a polyethylene tube are laid in each of the slots; this pair continues down the probe until it reached the location to be measured. The tubing terminates inside the diffusion chamber while the thermocouple is extended out one of the holes in the copper tubing so that it senses the compost temperature outside the probe. Probes were constructed and placed so that 10 locations 82 .cowumoofi co ocwpcmdmp mumoEmco commowwmp Lmaaou omumLOMLmQ Lao“ ob oco pm: monoLm .mcomuoom Loadou 0cm cocoo3 mcwumcumuam mcm3ocm mom can mmuoumLmQEmu mamaQEmm Low mooLm "H.a muomwm umamm .2262 Eu ma (HP) Eu ma . Eu ma )b) Eu ma III III III III. who: u>ooLu cm uumz a ocmnae A Uh. 330° oooz / 9.3:... cu cucunw\m ouuoLOuuum 83 in a cross-section of the compost windrow could be monitored. Triplicate sets of probes were constructed to allow for three cross-sections (0.4 meters apart) to be studied for a total of 30 points or 10 locations with 3 repetitions. Figure 4.2 is a cross sectional view of the windrow showing typical probe placement. Location 7 was not Sampling location 7, l/ . 21 Z). I 3 ' l L0n1 "10L 1 3 6T $ Dimensions typical '4--1.22m 1.22m 1.22m Figure 4.2: Typical temperature and gas sampling locations. monitored because the windrows were not large enough. Figure 4.3 shows an orthogonal projection of the windrow and a collection system. 4.4: Temperature Measurement 4.4.1: Equipment and Method Copper constantan thermocouples were connected to a Campbell Scientific CR21X datalogger through a 32 channel multiplexer. The datalogger provides reference temperatures and the calibration curves for voltage—to-temperature conversion. A different method of connecting the A. Bulkrnatefialsanuflinglocafion D. Datahagger B. Reiativehumiditysensor E. Multiplexer C. Gas sampling/ Figure 4.3: F. Gas sampling ports thermocouple line Orthogonal projection of test windrow showing temperature and oxygen probe placement. Windrow height and base dimensions are typical“ 85 thermocouples to the datalogger was used in each of the two experimental windrows. In Windrow l, the thermocouples from the corresponding locations on different probes were connected in parallel and then run to the datalogger. This allowed an average temperature for each of the 10 locations to be obtained while using up fewer input ports on the datalogger. In Windrow 2, the thermocouples from each probe location were run directly to the datalogger so that the temperatures at all 30 locations were separately monitored. Temperature in the concrete slab under the compost, ambient temperature and ambient relative humidity were also monitored. The temperatures and relative humidities were sampled every minute and averaged every 15 minutes. 4.4.2: Error Analysis and Calibration Table 4.3 summarizes the sources of possible error with the Campbell CR21X datalogger and AM32 multiplexer used with copper-constantan thermocouples. According to the Campbell literature (Campbell Scientific, 1985) the maximum error is :1.5°C. A more detailed explanation of these calculations is included in Appendix A. The largest source of error shown in Table 4.3 is due to the difference in thermocouple output. Accordingly, the thermocouples were subjected to two tests. In the first the complete measurement system was tested by inserting thermocouples into an ice bath and then heating the water until it boiled. Temperatures were recorded every 5 seconds and averaged every minute using the same datalogger used in 86 Table 4.3: Temperature measurement errors. ERROR SOURCE TEMPERATURE ( C) Reference junction temperature 2 0.5 Thermocouple output 2 1.0 Thermocouple voltage measurement 1 0.05 Reference linearization t 0.001 Output linearization i 0.001 Total error i 1.5 the field. At the icepoint the thermocouples had an average temperature of 0.42°C with a standard deviation of 0.l3°C. At boiling, the average and standard deviation were 99.0°C and 0.2°C, respectively. The second test involved inserting the 33 thermocouples that were actually used in the experiment into a constant temperature water bath. Temperatures were recorded every 5 seconds and averaged over a minute. The water was agitated with a propeller attached to a 3/4 hp drill. The water temperature was varied between 27°C and 69°C, with four different temperature plateaus being held. The results of this test are shown in Table 4.4, with the average and standard deviation of all thermocouples at a certain time being given. This method was chosen because it was hard to maintain the water bath temperature exactly constant over time. Sample data from this experiment are given in Appendix B. As can be seen from the data in Table 4.4, the 87 Table 4.4: Water bath calibration of 32 thermocouples. Number of Average Average Data Temperature Standard Points (°C) Deviation 30 27.556 0.004 30 35.554 0.271 25 42.118 0.001 25 55.742 0.008 36 68.119 0.054 variability between thermocouples was very low. Only the second temperature step showed an average standard deviation greater than 0.l°C. This is probably because heat was first applied in this step and the induced convection currents had not stabilized at the time of monitoring. 4.5: Gas Sampling and Analysis 4.5.1: Equipment and Method Thin walled plastic tubing was run from each probe diffusion chamber in slots cut in the wooden dowel to the top of the probe (Figure 4.1). A connection was made to a flexible plastic tubing. The total length of tubing to each location was 4.9 m. The other end of the tube was connected to a 3-way valve (Figure 4.4). A 3 cc syringe body plugged with a rubber septum was attached to one of the two remaining valve outlets. The third was open to the atmosphere. All joints were sealed with silicon caulk and were inspected before a sample was drawn. Joint integrity was tested weekly when the windrow was turned and the probes were removed. Other pieces of the apparatus consisted of a 88 Gcc syringe with 1 1/2" - 229 needle \ Rubber septum \ .W‘ I I I 3ccs rin ebod .1 Y 9 Y L5. 3-Way valve [ O O J air—@— /4 lasti t b' / Silicon caulk and tape p C U ”‘9 l 3 way pipette bulb TO probe in compost Glass eyedrop tube Figure 4.4: Gas sampling apparatus. Valve is in position to allow sample to be drawn from diffusion chamber in pile., 89 50 cc rubber pipette bulb, a 6 cc syringe and needle, and evacuated test tubes with rubber septums. Samples taken from Windrow 2 had a plastic collar around the probe to discourage gas flow along the probe during sampling. Gas samples were drawn at four times during each week of treatment: on the lst, 3rd, 5th and 7th days from the start of each turned windrow. Samples from locations closer to the surface were taken first. The same order of sampling was maintained throughout the experiment. A total of thirty windrow and 3 ambient samples were drawn each day. When gas samples were taken, the long sampling tube was evacuated by attaching the compressed pipette bulb to the open end of the 3-way valve, opening the valve, and allowing the bulb to expand to its normal shape. This removed the gas present in the tube since the previous sample. Gas from the diffusion chamber was sucked up to the valve. This gas had previously reached equilibrium concentration with the gas outside the diffusion chamber. The valve was then turned to open the passage between the diffusion chamber and the 3 cc syringe body. A 6 cc sample was drawn from this and discarded. Six cc was again withdrawn, approximately 0.2 cc was ejected to clear the needle, and the sample was immediately injected into the evacuated test tube. The ‘volume of the test tube was 4 cc; gas samples were stored at (about 1.5 atmospheres. Two cc of gas at pressures greater ‘than.one atmospheres was available for gas analysis. The (Evacuated test tubes are commercially available and haVe 90 been successfully used to store samples for 2 weeks (Grofman, 1986). Storage tube septums were sealed with silicon caulk after the sample was injected into the test tube to protect the sample for longer periods of time. Gas samples were analyzed for carbon dioxide and oxygen on a Carle Model 8700 Basic Gas Chromatograph. A silica-gel column was used to analyze the carbon dioxide. A 5A molecular sieve with a 60/80 mesh column in series with the gel column separated and analysed oxygen and nitrogen. Two- ~to-four standard gases were used for developing calibration curves over a range of 1 to 20 percent for both carbon dioxide and oxygen. Concentrations were calculated based on peak heights. A Hewlett-Packard model 3390-A reporting integrator recorded areas under the peaks as back-up data. 4.5.2: Error Analysis There are several potential sources of error in the gas sampling and analysis system. The sources of error are shown in Table 4.5. Table 4.5: Potential sources of error in gas sampling, storage, and analysis. Gas drawn into sample from outside sample area. Leaks into gas sampling tubes before sampling. Diffusion through needle into sample before injection into the sample tube and/or G.C. . Flow of sample out of sample storage tube and diffusion of gases (primarily oxygen) into sample storage tube. Withdrawal of sample at less than atmospheric pressure. Gas Chromatograph calibration. Gas Chromatograph problems. 4:- (AMI—a O O 0 NOW (J) o 91 4.5.2.1: Effect pf Air Being Drawn into Sample from Outside Sampling Area The error caused by gas being drawn into the sample from outside the sample area is indeterminate. A decrease in pressure at the inlet to the sampling tube caused by suction applied at the sampling end causes gas to move into the tube. Gas will be drawn from those volumes with the least resistance to flow. In an ideal situation, this gas comes from a spherical volume immediately surrounding the inlet. Changes in the cross-section of the probe will cause the hole cross-section around the sampling point to be larger than the probe itself (Figure 4.5). The air in this void space has less resistance to flow than air in the dense layer around the probe. Three factors could increase this potential error. First, if a portion of the probe has a blunt tip or a cross- section that is much larger than average, the hole in the windrow is larger than the probe. Second, even if the probe and hole is initially the same diameter, over time the hole near the surface can become larger due to windrow settlement. As mentioned earlier, this problem was encountered with the use of permanent probes left in the windrows for one month or longer (Mote, 1986). Both of these factors increase the likelihood that ambient air could ‘be sucked into the sample. Finally, the use of a vacuum pump to draw samples exacerbates this problem. In order to estimate and control this source of error, three approaches were used: calculations of potential 92 xLlelx " Pedorated copper tube I x f1 Compost \— Diffusion chamber \ \— Wooden dowel . *- (.a75 1(1) .. .4... Figure 4.5: Cross-section of probe insertion hole showing annular void space created by probe insertion. 93 error, probe design, and experimentation. Calculation of potential errors are made for two cases: (1) an air-filled void around the probe, and (2), no void. Case 1 assumes that air only comes from the annular void around the probe. The affected height is calculated with the equation 1a = 0 (r02 - riz) / V5 (4.1) length of air annulus affected by sample withdrawal ' outer radius of annulus where 13 r = o . . ri = inner radius of annulus VS = volume of the sample In Case 2, the air is assumed to come from a spherical volume around the sampling point. The affected radius is calculated with the equation r3 = (Vs / n FAS)1/3 (4.2) where r = radius of compost air affected FAS = free air space A range of annular ring cross-sections for Case 1 and free air space for Case 2 are assumed. Sample size varies for both cases. The estimated volume of the sampling apparatus is also considered. In the first Case 2 analysis, a range of large sample volumes from 100 to 1000 cc was looked at. Assuming that 6 inches is the largest acceptable radius where no sample contamination occurs, 300 cc samples are the largest that can be safely taken for all potential compost free air space. This leads to the rejection of two methods of gas analysis: a Bacharach Fyrite tester and a flow-through gas Chromatograph. 94 Analysis of Case 1 indicated that larger sample sizes and small void space led to samples being drawn from further away. Large sample size and large annular void spaces draw air from less distance. It is easier for outside air to get into the windrow and affect the long term composting process in this case, however. A Case 2 analysis with smaller sample volumes indicate that for most tube diameters and lengths, the radius of influence is less than 6 inches. The air sampling system incorporates two features to reduce the possibility of contaminated air being sucked into the system. Probes are designed and constructed with a constant cross-section as determined by the materials and the need for probe rigidity. Diffusion chambers are incorporated into the probe to provide a reservoir of gas equilibrated to compost gas concentrations. The volume of the sampling system is shown in Table 4.6. 4.5.2.2: Effect of Withdrawing Succesive Samples 23 Changes 12 Gas CEncentration An experiment was conducted to measure the changes in gas concentration as successive samples were withdrawn from a sample location. The probes were inserted into the windrow in the same manner and locations as for an actual experiment. Successive 5 cc samples were withdrawn until 75 cc was removed. Four different windrow locations were sampled to provide a range of windrow conditions: 1, 5, 8, and 10 (Figure 4.6). The results of one of the sampling runs from Location 1 are shown in Figure 4.7. The 95 Table 4.6: Internal volume of sampling tubes and diffusion chamber. Volume (cm3) Sampling system: Tube ............................ 13.5 Syringe ......................... 5.0 Diffusion chamber ............... 7.2 25.7 Sample Size Rubber Bulb... .................. 50.0 Discard....... ................ .. 6.0 6.0 Sample.......................... 0‘ N o O Sampling location Figure 4.6: Sampling points for sequential gas sampling. 96 22 20 - + 16 ‘ LOCATION 1 09 V 18 CI! C - O “ o T + o: + + ‘ + I. 2 1o- 0 U S B -( o a o 0 0 a U U a .i U) (U o ,4- 23 D O I I I V T I f 0 2O 40 BO 80 Sequential Volume (cc) 0 C02 '9 02 Figure 4.7: Sequential sample gas concentration, Location 1. concentration of carbon dioxide in the first 5 cc sample is low but rises close to its final value by the third sample. Oxygen concentrations follow the reverse pattern. The lower values of the 5 cc sample indicate that there might be some leakage into the system, but that this is flushed out by the adopted experimental method. The constant value at the 65 cc volume, which corresponds to the usual sampling point, indicates that no contamination of the air sample occurred from this source. Results of the other tests are shown in Appendix C. 4.5.2.3: Effect 93 Leaks into Gas Sampling System Leakage into the gas sampling lines was discouraged by sealing all joints with silicon caulk. Each day before 97 samples were taken, the integrity of the lines and silicon seals were checked and repaired if necessary. At the end of each weekly batch, the probes were removed from the windrow and each line was leak tested in a water bath. Blocked or leaky lines were replaced. 4.5.2.4: Effect pf Diffusion and/or Gas Flow into Syringe and Gas Storage Containers Solutions to the third potential error were incorporated into the sampling procedures. To discourage diffusion into or out of the syringe, the syringe was kept in the sampling apparatus for 6 to 10 seconds after the second 6 cc sample was drawn. This allowed pressures inside the sampling apparatus to approach atmospheric. After the removal of the syringe from the apparatus, the ejection of approximately 0.2 cc cleared the needle of any ambient air that might have entered. At the start of the experiment, the use of the evacuated test tubes for gas storage was assumed to be accurate based on the experience other researchers at MSU (Groffman, 1986): analysis within 2 weeks of sampling does not affect accuracy. Samples from the first windrow were analyzed according to that timetable. Due to problems with the gas chromatograph, however, analysis of the second set of gas samples was delayed beyond the two week limit. In addition, the oxygen levels in the center of the windrow :seem high. An error was suspected in the gas concentration .analysis procedure. Accordingly, two experiments were (nanducted to assess the integrity and durability of the 98 storage system. A possible source of error is the withdrawal of air samples at less-than-atmospheric pressure. This occurs if sample withdrawal caused sub-atmospheric pressures in the syringe and sample storage tube. To test this, ten evacuated tubes were filled with nitrogen. Successive 1 cc samples were withdrawn and analyzed on a gas Chromatograph. Figure 4.8 shows the effect of the sample volume withdrawn on the measured oxygen concentration. Syringe internal pressure was estimated based on the volume of the sample container, the volume and number of withdrawals and the initial pressure. Comparison (Figure 4.9) of the calculated syringe internal pressure to measured oxygen concentrations indicates two almost linear segments. Assuming that the initial and atmospheric oxygen concentrations are 0 and 20.95 percent, respectively, the influx as a function of sample number, sample volume, and calculated internal pressure was calculated. The relationship between influx and calculated internal pressure is shown in Figure 4.10. The preceding results needed to be converted from the 3 cc sample container size used in the calibration experiment to the 4 cc size used in windrow monitoring. To accomplish this, the influx values for the 3 cc container were converted to volume influxes and then percent influx for the 4 cc container was calculated. The calculated 3 cc container internal pressures that correspond to the volume 99 1.5 ATM, 3 CC BOTTLE 5 1 CC SAMPLES ) (3) 3 I Oxygen OdMUhUD‘IDO l I I T I 5 ...- U Cummulative Withdrawals (cc) 0 Rep 1 + Rep 2 0 Rep 3 A Rep 4 x Rep 5 v Rep 6 Figure 4.8: Effect of cummulative gas withdrawal on oxygen concentration. 16 15~ v 14- 1.5 ATM, 3 CC BOTTLE & 1 CC SAMPLES (3) a...“ o-‘NU J_lll Oxygen dMUI-U‘OUOID J 0 q .4 1 4 .. .4 q -1 .. 0.2 0.4 0.6 0.8 1 1.2 Pressure (atm) 0 Rep 1 + Rep 2 o Rep 3 A Rvp 4 x Rep 5 v Rep 6 Figure 4.9: Effect of sample pressure (calculated) on oxygen concentration. 100 100 so 1 30 1.5 ATM INITIAL PRESSURE, 3 cc BOTTLE 704 (1) BO- 50-4 40--1 Influx 30+ 20- 10--( O I wI I I ‘I ‘—I I I I I I o: os 07 03 L1 L3 L5 Internal Pressure (atm) Figure 4.10: Percent influx as a function of internal pressure. withdrawn were used with the 4 cc influx values in a linear regression. This equation was used to calculate the influx at internal pressures corresponding to those occurring after each successive 0.55 cc sample volume was withdrawn from the 4 cc container. Finally, a linear regression equation was run on sample withdrawal number vs influx data to arrive at an equation useful for transforming the experimental gas concentration data. The resulting equation is In4 0.0198 * S + 0.002 (r = 1.000) (4.3) 4 cc influx, decimal where In4 S Sample number Observations of sample containers that had their rubber septums covered with silicon caulk after the samples were injected indicated that there was some leakage out of the 101 sample tube. Since calculations of internal pressure indicated that changes in starting pressure influenced the speed at which a vacuum was produced in the sample container, it was necessary to get an estimate of the error due to variations in sample starting pressure. I assumed that the worst case would occur if enough gas had leaked out of the sample container so that the gas was at atmospheric pressure. In this case, all syringe sample withdrawals would be at less than atmospheric pressure and so subject to an influx of air due to the pressure differential. I assumed that this influx corresponded to the second and steeper slope in the data shown previously in Figure 4.10. In order to quantify this relationship a series of calculations similar to the previous case was performed. All calculations are shown in Appendix D. The resulting regression equation (r = 0.998) for sample number vs. influx is given by: In4 = 0.0851 * S + 0.0027 (4.4) The effect of the withdrawal of a sample of a given volume and measured gas concentration on actual gas concentration is described by Equations 4.5 for oxygen and Equation 4.6 for carbon dioxide. For a given sample volume the effect is linear and can be described as follows for each gas: 02: O = (0m - In * Oe) / F (4.5) c02: ca - cm / F (4.6) 102 where O = oxygen concentration, percent C = carbon dioxide concentration, percent a = actual m = measured e = ambient In = influx gas in syringe, decimal F = original gas in syringe, decimal l — In Equations 4.3 or 4.4 can be substituted into Equations 4.5 and 4.6 to achieve the desired correction. Four different correction calculations were considered. The first, the low influx method, involves the insertion of Equation 4.3 into Equations 4.5 and 4.6. The second method was accomplished by inserting the high influx equation 4.4 into Equations 4.5 and 4.6. The third method was achieved by calculating the average of Methods 1 and 2. The fourth method was based on two assumptions: the measured carbon dioxide was correct and the sum of carbon dioxide and oxygen concentrations was approximately 21 percent. With this method, measured carbon dioxide was subtracted from 21 % to achieve the oxygen concentration. Method 3, based on average values, was chosen as the most appropriate correction calculation. Method 4 does not make the desired corrections. Methods 1 and 2 are subject to the uncertainty of the initial sample pressure at the start of analysis. The average method is at best a compromise, because it is subject to the same uncertainty. The measurement error due to the correction factor is 11.0 % oxygen at low concentrations and £0.25 % at high concentrations. Measurement errors for carbon dioxide are 20.25 % at low concentrations and 11.0 % carbon dioxide at 103 high concentrations. The effects of storage system type and time were also examined. Leakage of gas through the hole in the septum due to initial overpressure followed by diffusion of oxygen into the sample storage tube could occur and affect measured results. Evacuated test tubes were filled with pure nitrogen to 1.5 atmospheres pressure. Half of these had uncovered rubber septum tops as in Windrow 1 samples. The rubber septum tops of the other half were covered with silicon caulk to block the injection point, similar to Windrow 2 samples. Five tubes from each treatment were analyzed each week for 5 weeks. The same four point calibration curve was used to calculate the percent oxygen. There were no observable oxygen peaks in weeks 1 to 3. Week 4 samples showed small oxygen concentrations that were statistically the same at the a = 0.05 level. Both covered and uncovered samples were significantly lower than the initial oxygen concentrations. Week 5 covered samples were again significantly lower than week 4 covered samples. The covered samples from week 5 showed a very large variance and the average was significantly higher than any other time and treatment. Two of the four samples from this group exhibited the same concentration as the covered samples but the other two were quite higher. The covered samples exhibited small bubbles in the silicon, indicating that some gas probably escaped from the uncovered tubes. This is also 104 supported by the larger variability in week 5 uncovered samples. The results of this experiment do not elucidate the effect of storage time and treatment although they seem to indicate that for the covered system there is no loss of gas and for the uncovered system that it was only after 5 weeks of storage that a loss of gas in some storage containers may have occurred. Tentatively, it can be concluded that storage time has no effect on the covered samples used in Windrow 2. While the uncovered samples do show greater variability in week 5, this was far beyond the actual storage time for Windrow 1 samples, and the time effect could be ignored. 4.5.2.5: Gas Chromatograph Calibration and Errors Some potential for error is present in the procedures used to calibrate the gas chromatograph. Initially, I thought that point calibration would be adequate based on the experience of other researchers using the same gas chromatograph (Dilley, 1986). It shortly became apparent that more points were needed and another bottled standard was used for two point calibration; most of the samples from Windrow l were measured using two-point calibration. Samples from Windrow 2 were analyzed using a four point calibration curve. The additional standards were made up volumetrically. The bottled standards were manufactured with a fixed, absolute error. The volumetric standards were mixed before each sample analysis session. The method of mixing involved 105 inserting septums into 160 cc sample bottles, alternately evacuating the bottles with a vacuum pump and filling them with nitrogen gas, and then filling the evacuated bottle with precalculated amounts of oxygen, carbon dioxide, and nitrogen to 2 atmospheres pressure. 30 and 60 cc syringes were used to fill the bottles. Table 4.7 summarizes the standards and variability used in this study. Error calculations for the volumetric standards are given in Appendix B. Table 4.7: Gas concentration standards used in gas chromatograph calibration. Gas Concentration Standard Level Std. Dev. (% Atmos.) (% Atmos.) coz 02 coz 02 1 5.079 20.524 $0.10 $0.61 2 0.983 4.97 $0.002 $0.15 3 10.0 10.0 $0.226 $0.226 4 1.0 20.0 $0.238 $0.167 The gas chromatograph itself presented a significant source of error at times. Two types of problems arose. When a sample had a significantly lower concentration than the preceding sample, there was a lag in the apparent concentration. This caused the first new subsample to appear to have a higher concentration than it actually did. The second problem was due to the erratic behavior of the gas chromatograph when analysis of Windrow 2 samples 106 were to begin. As the first day's samples were analyzed, the gas chromatograph became more erratic. Finally, not even standards would behave consistently. The problem was tracked down to a faulty injection valve. The gas chromatograph behaved correctly after replacement. Concentration data from two days were lost, however. 4.6: Core Sampling Method 4.6.1: Review pf Sampling Methods and Corer Design Moisture content, volatile solids, coliform numbers, bulk density, porosity and free air space are expected to vary spatially. Accurate samples must be taken to understand the heat and mass transfer process in DMS composting and to provide data for future modeling efforts. Sample location in the windrow must also be known accurately. The sampling process, however, must disturb the windrow as little as possible. Sample handling and storage are also important but will be covered in later sections. Moisture content and volatile solids samples are the least difficult to obtain. Sampling does not change either of these two variables, but it may disturb the windrow. Hoyle and Matingly (1954) used a hollow auger that was 61 cm long and 2.67 cm in diameter to take vertical samples from 61 x 61 x 61 cm composting bins. The material was analyzed for moisture content, ash and various forms of nitrogen. A grid was used to ensure that samples were not taken from the same location in the windrow. No mention was made of the effect of sampling on the composting process. Finstein et 107 a1. (1980) used a clamshell posthole digger (5 inch diameter) to obtain samples from 7 ton windrows. Great care must be taken with sampling for micro-organisms to prevent contamination. Sampling for bulk density, porosity, and free air space is more difficult. Bulk density determinations are based on both sample weight and initial volume; the act of sampling can change the volume of the sample. Disturbed samples can be used to get an estimate of bulk density. Several methods that use disturbed samples and the addition of water in calibrated containers have been developed (Singley et al., 1982). These methods do not adequately account for the effect of the overburden of the compost and the compaction that it causes. The value of the bulk density at a particular location over time is hard to obtain as well. Several methods of determining soil bulk density are described in the soil mechanics literature (ASTM, 1986). These methods were either quite destructive of the windrow or the required equipment was not available for use in this study. Core samplers are frequently used for soil bulk density determinations. Raper and Erbach (1985) note that samplers can either be hammered, driven in at a constant speed, augered or inserted into the soil by some combination of the above methods. Evidence that the length of the core sample is shorter than the sampling depth indicates that compaction may be a problem (Wells, 1959). In order to minimize the 108 effects of compaction, Baver (1956) recommends using at least a 7.62 cm diameter tube. Core sampling device design greatly affects sampling accuracy (Raper and Erbach, 1985). Compaction can be avoided by sampler point and interior sampler design. Shaping and sharpening the sampler point shears the soil instead of deforming it ahead of the sampler and also directs unwanted soil to the outside of the sampler. Providing a taper or interior sampler volume greater than the interior diameter of the cutting head allows the sampled material to expand and reduces the friction resisting the penetration of the probe. Coating the interior with teflon does not appear to make a significant difference in the sampled core bulk density (Raper and Erbach, 1985). The accuracy of standard soil corers was analyzed by Baranowski (1983) with forest soils. He found that the core method could reveal intertreatment densities to accuracies no greater than :50 kg/m3, even with favorable conditions, careful sampling and adequate replications. Under less favorable conditions, the error was as much as 5 times larger. Peat sampling provides a system analogous to compost sampling; both involve taking samples of a fibrous organic material. Peat sampling equipment includes the use of Macaulay samplers, core cutters and stationary piston samplers (Jarret, 1986). Bulk density from these samplers can be determined by the kerosene, paraffin, or cylinder 109 methods (Nat. Res. Counc. Can., 1979). The most interesting method required a cylindrical cutter attached to a core holder and base, powered by a hand drill. In this method, bulk density was determined by dividing the oven-dry weight of the peat core by the inner volume of the cylinder. Stanek (1980) compared the results of bulk density determinations made by the paraffin and cylinder method to the kerosene method which was considered to be the most accurate. He found that the cylinder method correlated well with the kerosene method (r 5 0.979). Based on this information a method was adapted for sampling a compost windrow. 4.6.2: Comparison of Core Sampling Methods The rotary corer required some skill to operate properly. Unless the core was held very steady, it would rotate off-course and take in material from a volume greater than the volume intended. When the corer was removed, some material from around the lip occasionally fell into the hole. The sides of the cylindrical hole did not appear to be excessively disturbed. A narrowing of the hole began within 2 minutes and the bottom appeared to rebound following removal of overburden pressure. The end of the core sample showed evidence of the rotation and usually broke off cleanly. The length of the core sample was usually 10 to 20 % less than the hole in the windrow, indicating that some compaction had occurred in the corer, windrow or both. When the corer was used to take samples of 110 low density material such as that near the windrow "toe," it was observed to push some material aside. This was not observed in the more dense portions of the windrow. Two different length and diameter corers were tested. The smaller diameter corer was harder to insert into dense portions of the windrows; it was unable to penetrate windrows composed of long straw. An experiment was conducted to estimate the accuracy of several core sampling methods for bulk density, porosity and free air space determinations. Three 55 gallon drums were filled with 1 week old separated manure solids that had been mixed in a silage mixer. After leveling the tops, three sampling methods were used: (1) a 20.32 cm diameter ring, (2) a short 7.78 cm diameter rotary corer and (3) a long 5.08 cm diameter rotary corer. The rotary corers were powered by a 3/4 hp electric drill. Three samples of approximately 30 cm were taken from each barrel by the two rotary corers. The ring sampler was then pushed and rotated into the surface and the sample was then removed by hand. Approximately 30 cm of material was removed from the barrel and the ring sampling procedure was repeated. This was repeated to obtain the third ring sample. The above steps were repeated in each of the barrels until 3 replications of sample depth for each method had been obtained. Samples were stored and analyzed for weight, moisture content and volatile solids. Bulk density, porosity and free air space were calculated. The methods of lll analysis and calculation are described below in the appropriate sections. Detailed results are given in Appendix F. Wet bulk densities ranged from 198 to 566 kg/m3 depending on the sampling method and depth. Precision was good with the average coefficient of variation for each method at all depths being 8.7, 10.9 and 10.0 % for the ring, short corer, and long corer methods, respectively. Using the two-tailed Student's t test, the three methods were significantly different in at least two of the three depths at the 90 % level. The short sampler always gave the largest bulk densities followed by the long sampler and the ring method. Free air space values in this experiment ranged from 0.850 to 0.471. Using the same statistical test, the three methods were significantly different in at least two of the three depths at the 90 % level. The average coefficient of variation for each method at all depths was 3.2, 7.5 and 3.0 % for the ring, short corer and long corer methods, respectively. The ring method was not chosen for use because it could not be used to sample windrow interior bulk densities. The long corer with the small diameter was not chosen because of the difficulty encountered with inserting it into some windrows made of long straw. The short wide sampler was selected because it would be able to penetrate a wider range of compost material and its greater diameter would be less 112 succeptible to errors in the radius determination than the longer, narrower sampler. 4.6.3: Description 9f Experimental Core Sampling Method Samples for determining moisture content, volatile solids, bulk density and free air space were obtained on the lst, 3rd, and 7th days after a windrow was turned. A 7.78 cm diameter rotary corer attached to a 3/4 hp electric drill was used to drill vertical sections out of the windrow. The corer is shown in Figure 4.11. The cutting edge is serrated to cut through fibrous material and slanted away from the path of the core opening to push away any material not directly in the path of the core opening. To relieve friction from the surrounding windrow, the cutting head exterior dimension is greater than the core body. Similarly, the interior diameter of the cutting head is smaller than that of the corer body interior to allow the compost material to expand and give reduced resistance to the entry of material at the cutting head opening. The corer allowed vertical cylindrical compost samples of variable length to be removed from the windrow. The vertical center of each approximately 30 cm core sample was at the same horizontal and vertical point as a temperature and gas measurement location (Figure 4.12). At least 9 samples were taken at each time. No replicates of these samples were taken because additional samples would have seriously disturbed the windrow. The sampling procedure involved first determining the 113 \ .1. T 1.27 cm ' a F (0.5 In.) I V l \ ‘ \N": 1.27 cm Q2 \ I (0.5 in.) N I N N N I N 9" \\ N . . 8.89 cm (3.5 In.) Diameter /’ \v///\\ 1/16"1TfickroHedsteel ( LA ‘1 . a A» N l N N N . N t 74.9 cm N 1 (29.5in.) , N 0 N I j N L‘ V V N N N | h N W N :1" T 0 108cn1 ¥ l \\ (4.25 in.) ($5775?) 1 1' ' ' 8.57 cm (3 3/8 in.) 0.0. V ¢ . 1 3/16" Thick pipe 7 78 3.8) cm .. _ D: ’ cm 1.5. . (3.06in.) (‘— ( m) Figure 4.11: Rotary core sampler. 114 Center of sampling location -1313; 0 O u Figure 4.12: Core sample locations. Cylindrical in shape, the center of the sample was located at approximately the same point in the yz plane as the temperature/gas sampling p01nt. location of each sampling point below the windrow surface. The beginning and ending locations of each sample were calculated. A catwalk placed directly above the sample location provided a place for the operator to stand while taking core samples. A colored scale on the side of the corer guided the operator to the proper sampling depth. The exact hole depth was measured immediately afterwards. The compost sample was removed from the corer and placed into either a ziplock bag for moisture content and volatile solids analysis or a "whirlpak" for coliform analysis. If a visible color difference was present in a sample, the sample was split and separate analyses performed. The holes created by the withdrawal of samples were filled with DMS compost material of a similar age, lightly compacted and marked. In the experimental method adopted, a single large sample from each location was taken at each time. The first 115 set of samples was taken from one side of the windrow after the windrow was one day old. The second set was taken from the opposite side on the third day and the third set was taken on the seventh day from the same side as the first set, approximately 1.5 feet away. This method diminished the effect of sampling on the transport processes, but the method could be criticized because the second samples were taken from an area that was different. 4.7: Moisture Content and Volatile Solids 4.7.1: Methods and Calculations Gravimetric moisture content (wet basis) and volatile solids (total solids basis) were determined using Standard Methods for the Examination of Water and Wastewater (AWWA, 1976). Gravimetric moisture content on a dry basis, volumetric moisture content and volatile solids on an ash basis were calculated. Samples were removed from the pile as described in the bulk sampler section, and were placed in ziplock bags and stored at 2°C. Samples were first mixed by hand kneading the bag with the sample still inside. Sub-samples were removed by tongs and placed in ceramic crucibles. The crucibles had previously been fired in a furnace at 550°C t 50°C for 1 hour, cooled in a dessicator and weighed. The wet samples and crucible were weighed and dried in an oven at 103°C i 2°C. After 24 hours, the sample were removed, cooled in a dessicator and reweighed. Volatile solids were determined by ashing the dry 116 matter in covered crucibles at 550°C 1 50°C for 1 hour. If black or red material remained in the crucible at this point, the crucibles were placed back in the furnace for 5 minutes, sans covers, to completely ash the samples. After removal and cooling in a dessicator, the crucible and ash were weighed. I A Precision Scientific model 1254 oven, a model FD204C Hoskins Electric furnace, and a Sartorius model 1265 MT scale were used in this procedure. The scale is accurate to t 0.001 g. The following calculation for gravimetric moisture content (wet basis) was used: ucwb = [1 - (MCS- Mc)/(Mcsw — MC)] (4.7) where Mcwb moisture content, wet basis, decimal ' M x Mass of x, g csw = crucible, solids and water cs = cruelble and solids c = crucible One of the problems with wet basis moisture content calculations is that changes in the mass of water change both the numerator and denominator. This makes it impossible to directly compare the moisture contents at any two points in the process. If moisture contents are expressed on a dry basis, however, only the numerator changes. The dry basis calculation is as follows: MCd = Mw / Md (4.8) moisture content, dry basis, decimal where MCdb w water 117 d = dry solids A comparison of wet (% wb) and dry basis (% db) moisture contents is presented in Table 4.8. Dry basis moisture contents are much more sensitive to changes in the weight of Table 4.8: Comparison of moisture contents expressed on a wet and dry solids basis. Method Moisture Content (%) MCw 90 80 70 60 50 40 3O 23 20 MCd 900 400 233 150 100 67 43 30 25 water that is present in a substance. Volatile solids were calculated with: vs 2 100 * [1 - (Mca - MC)/(MCS - MC)] (4.9) where VS a volatile solids, dry solids basis, decimal ca = crucible and ash Expression of volatile solids on a total solids basis suffers the same drawback as the wet basis moisture content measurement. Both the numerator and denominator change as the reaction progresses. This makes it impossible to directly compare the volatile solids at any two points in the process. Another way of looking at substrate consumption using the same data is to calculate the change in the volatilized solids based on the ash fraction that remains. vsa = MV / Ma (4.10) where VS Volatile solids ash fraction, decimal volatilized solids ash solids w <0: u H II 118 Since the ash fraction should not change unless additional inert material is added to the composting process, VSa should provide a stable base for evaluating substrate consumption. Table 4.9 illustrates the difference between the two methods of expressing substrate consumption. Ash solids expressions of substrate consumption indicate a much larger decrease in substrate than the volatile solids method. Table 4.9: Comparison of volatile solids expressed on a dry and ash solids basis. Method Substrate (%) VS 95 93 90 89 85 82 80 VSa 1900 1329 900 809 567 456 400 Three replications of each measurement were done initially in Windrow 1. Five were used initially in Windrow 2. Outlying data were rejected with the Q test (Davis, 1981). If the variance was greater than one percent, an additional 5 samples were run. 4.7.2: Sensitivity Analysis The sensitivity of moisture content and volatile solids determinations to errors in their measurement was determined using the theory of the propagation of errors (Parratt, 1961; Topping, 1951). The following paragraphs summarize the derivation presented in Appendix G, which was orginally given by Parratt (1961). 119 A result U, which is computed from the quantities X and Y has the following variance for indeterminant errors: 5,.2 = (an/ax)2 5x2 + (aU/ay)z 5Y2 (4.11) and sy2 = variances associated with measurement of X and Y. 2 where sx For addition or subtraction the variance of the calculated result is given by the sum of the variances: 2 + z 5 SY (4.12) If U is a result of multiplication or division such as U U = x3 ya (4.13) then the variance is given by = [(axa"1yb)2 5,,2 + (bxayb'l)2 sy2 1 (4.14) 2 5u The fractional variance is given as: suz/ u2 = [(sx/x)z + (sY/y)z] (4.15) The fractional variance can be put in terms of percentage by multiplying by 100. Fractional variance for the wet basis moisture content calculation is given by: SMCwbz/ MCwbz = [((scs2 + scz)/ Mdz ) + Use“2 + sc’)/ Mwsz)] * 100 (4.16) where ws = wet solid A similar sensitivity analysis can be run for the dry basis moisture content calculation. The equation is sMCdbz = (l / Md)2 SW52 + (MWS / Mdz)2 Sdz (4.17) Where SW52 = variance of wet solids measurement = 10.002 9 ’ssz = variance of dry solids measurement 10.002 9 120 The variance of volatile solids (dry solids basis ) is given by Z 2 2 + (sCS + sC )/ Md ] * 100 (4.18) where 5X2 = variance associated With measurement Despite the large observed VSa changes, VSa is still subject to the same errors present in the volatile solids determination. Errors in VSa determination were analyzed for selected data using a sensitivity analysis procedure similar to that used for VS. The equation developed was: SVSaz = (1 / Ma)2 sd2 + (Md / Mal)2 sa2 (4.19) Where ss2 = variance of dry solids measurement = 10.002 9 sa2 = variance of ash solids measurement = $0.002 g The average sMCwb for 75 moisture content (% wb) determinations with measurement error variance of i 0.001 g is 2.26 % wb moisture content. Table 4.10 summarizes the results for six measurements selected to cover the typical range of dry solids mass and moisture contents encountered in the experiment. The average standard deviation of volatile solids (total solids basis) was much higher than that for MCw: 7.0 percent volatile solids. Four place accuracy in weighing is necessary to assure the same level of accuracy in volatile solids measurements as is achieved with 3 place accuracy of moisture content determinations. 121 Table 4.10: Sensitivity analysis for moisture content expressed on an dry solid basis. Wet Solids Dry Solids Mw Md 52 s (g) (g) (%) (%) 1.447 0.394 78.6 586 29.4 54.2 2.345 0.701 77.0 334 7.7 27.8 0.881 0.307 74.2 287 34.6 58.8 2.232 0.659 77.1 339 9.3 30.5 1.141 0.316 78.3 361 44.5 66.7 1.623 0.481 77.1 337 17.4 41.7 Table 4.11 summarizes the results for six volatile solids content measurements selected to cover a typical Table 4.11: Sensitivity analysis for volatile solids expressed on a dry and ash solids basis. Dry Solids Ash Solids VS VSa 52 s (g) (g) (%) (%) 0.307 0.044 85.7 598 5130 716 0.461 0.083 82.0 456 925 304 0.672 0.153 77.2 339 374 193 0.415 0.087 79.4 385 628 250 0.481 0.073 84.8 559 1670 408 0.338 0.051 84.9 563 3450 588 range of ash mass and volatile solids. In some cases the standard deviation of the measurement is larger than the measurement itself. The use of scales accurate to :0.001 mg is necessary to achieve reasonable accuracy. 122 4.8 Bulk Density 4.8.1: Calculations Bulk density was calculated using the "cylinder" method. The equation used in this calculation was: 0b = Mnet F / n r2 h (4.20) where 8b = wet bulk density, kg/m3 Wnet = net mass of compost removed from corer r = corer cutting head interior radius, in h = sample height, inches F = unit conversion factor The method used to obtain quantities necessary for bulk density determination has been partially described in the core sampling section. The length of the sample was determined by measuring its furthest penetration depth, and subtracting that from the penetration depth of the previous sample. A colored scale on the corer was used to take initial length measurments. This was immediately confirmed with a finer scale. Values were accurate to t 1/8 inch. The net mass of the compost material was calculated as: M = M15 + M - Mb (4.21) net SS where ls = large sample weight, 9 ss = small sample weight, g b = bag weight, 9 4.8.2: Sensitivity Analysis The variance of bulk density due to measurement errors can be broken down into four areas: net weight of solids, moisture content, effective core height and effective core radius. Variance due to the net mass of the solids is additive: 5M2 = 51 2 + s 2 + sb2 (4.22) 123 where the terms are as defined previously. For the analytical balance used, all variances were i 0.001 and s 2 equals 1 0.003. Moisture content errors have been net analyzed previously. The effect of errors in length measurement on the percent volume variance is shown in Figure 4.13. This is based on the following calculation: st12 / Vol2 = [(sh h)2 + (sr 2r)2 (4.23) where Vol = volume of sample, m3 From these data it can be seen that height does not greatly affect volume. Small volumes are more sensitive to errors than larger volumes. Radius errors effected volume variance to a much greater degree than errors in height. Figure 4.14 shows the effect of changes in radius variance from t 1/8 to 1 5/16 inches. The previous equation for variance was used in this calculation as well. Turning to the bulk density measurement, the partial derivative of bulk density with respect to weight is given by Bob / awnet = (l / n r2 h) (4.24) where the terms are previously defined. This shows that the rate of change of bulk density with respect to weight is a constant. The partial derivative of bulk density with respect to sample height is 30b / 3h = - 0b / h2 (4.25) 124 5.5 A . SA“ RADIUS VARIANCE +/- 0.125 1n 5.3 - a. 5;:. “e’ a 3.1 - ° H O .4 > S « ASJ a ‘.a .- + A g 1&7- A Q A '2 4.6-1 0° A 0 ° AA >' (AS - D o A '4'. 0° A ‘4 t. o o A. . ‘1 9 ‘. o A 6 A 0 0° A 43~ ch°°un;"**332§§‘ "2 i I U T t“ V fir F I U T 1' V U I T U I O 0.2 O.‘ 0.8 0.8 1 1.2 1 4 1.8 1.8 2 Volume (liters) o t/- 0.125 in + t/. 0.25 in o t/- 0.375 in A f/- 0.5 in Figure 4.13: Effect of length errors on total variance as a percent of total volume. 14 13- 12a LENGTH VARIANCE: +/- 1/4 in 11... .5 mmeXWXXXXXXX x E 10- 3 0.. 9—4 3 J AAAAAAAAAAAAAAAAAAA A B a v 7.1 0000000000000000000 O o 5‘ U S 5" .,.. H*”+.‘+4++++++++ + 1.. ‘4‘ G > .3- 2- 1... o I Ti—T’W t I FTiI’ Y Y T 1 I Y 1 Y T r O 0.2 0.4 0.6 0.8; 1 1.2 1.4 1.8 1.8 2 Volume (liters) _ + +/- 1/8 in 0 t/- 3/16 in A :/- 1/4 1r1 X */- 5/16 in Figure 4.14: Effect of changes in radius variance on total variance as a percent of total volume. 125 Changes in bulk density due to height vary with the inverse of the square of the height. The partial derivative of bulk density with respect to sample radius is Bob / 3r = - 2 0b / r3 (4.26) Errors in bulk density with radius variations vary inversely as the cube of the radius. Figure 4.15 shows the effect of two different radius errors on bulk density. The fractional variance equation used for this calculation is spbz / ab: = W2 Swz + 4 st2 / r2 + shz / h2 (4.27) The relative contributions of various measuring errors on bulk density are shown in Figure 4.16. Figure 4.17 shows the total, radius and height percent variances. Percent variance due to radius measurement errors makes up most of the total percent variance. Errors due to weighing and moisture content are smaller than that for core height. While this analysis is generally correct, it does not account for errors from several sources. First, if the core does not travel in a straight line as it goes into the windrow, larger effective volumes may be sampled. Secondly, the effective radius may be larger or smaller than the interior core head radius because of compost particles from outside or inside this radius being pushed out or pulled into the sample. Vibration of the core may shift some compost material out of the sampler. The effect of core 126 9 VARIANCE: WEIGHT +/- 1 g, 5‘ HEIGHT +/- 1/4 in, MC Exper. 7- 3 v B '- I!) 4 + 4 + C1 3 ++ + * c '5 ¢ 4 : + .2 5‘ a‘ +0.. .‘ ¢ . ‘ ++ + E ,¢w€‘+ “r “o: «1 + "1 0:10 o 0 DO D O 3... Dub (9 00 [P O D D txiecpmg'o ° ufihfififi5aprcp5= ‘1 2 I V I V I I O 200 400 800 Wet Bulk Density (kq/m3) 0 +/- 1/16 in + *l- 1/8 in Figure 4.15: Effect of two different radius errors on total variance as a percent of bulk density measurements. 0.38 A VARIANCE: WEIGHT *I- l g, HEIGHT *l- 1/4 in, 0.3.. RADIUS 9/- 1/8 MC Exper. 0 4 + g. 0.25 d f A. 4+ + d' 0.2 - + 8 + 6 4 4 +- C * 4 CO 0.15 -‘ 4+ 4* 4.... 4. o: A + A A g A “A 6' o 4 OJ 4 g? ** 999-2» 3 4 * A -H.+ o . ‘*¥ + ‘ngmr9 f'tv 1 0.0: ~ at :3 a A o- o A A a 0 6 Q 35° . A ‘4 A A a A A A 0 , 5fliaEaE3flnaya4uLJL41_ca_4nfid_;p__hr_hu___n O 200 400 800 Figure 4.16: Wet Bulk Density' (kg/m3) 0 Weight # height 9 MC Relative contribution of various sources of error to total variance as a percent of bulk density measurements. ' 4 £332.“ x x X X x V i“ x )Q ‘( .'I X 7 ' smegma” z: .. . . . q 4... f: VARIANCE: WEIGHT +/" 1 g, HEIGHT t/- 1/4 in, 3" RADIUS */- 1/8 MC Exper. m u c m 'C 2— m > 1..( 9* 4: 444 4 4 o .Mn‘h9fln L4” + t *+ T * ’ I I Y I 1' o 200 400 800 Wet Bulk Density (kg/m3) 0 Height + Weight 0 Total BD Figure 4.17: Relative contribution of radius and height errors to total variance as a percent of bulk density percent measurement. - rotation speed on these errors is also not described. 4.9: Porosity and Free Air Space 4.9.1: Calculations Porosity and free air space can be calculated with the results of the moisture content, bulk density and volatile solids content. The general equation for these calculations is: fx = l - (0b / 05) (4.28) where fx = porosity or free air space 9b = wet bulk density, kg/m 05 = particle density, wet or dry, kg/m3 Based, in part, on work by Bohnhoff et al. (1984), the particle density used in porosity determinations is: 05d = (A - B vs ) / (100 — MCwb) (4.29) where 05d =' particle density, dry, kg/m3 128 A,B = factors associated with solid particle density A and B are the parameters of a linear regression of volatile solids content vs. DMS particle density from three dairies with different separators and rations. These values are given in Table 4.12. Table 4.12: Values of constants A and B from dairy manure solids specific gravity measurements. Source A B Farm 1 (Centrifugal) 232000 814 Farm 2 (Stationary screen) 235600 868 Farm 3 (Perforated drum/ compressing roller) 200600 473 Average 227800 773 Source: Bohnhoff et al. (1984) Bohnhoff et a1. (1984) presented the following formula for particle density to be used in determing free air space: + (105 / mcwb)] (4.30) where 9pm = particle density, moist, kg/m3 Equations 4.23 and 4.24 are substituted into Equation 4.22 for the calculation of porosity and free air space, respectively. I = 1 — 0b * (100 - MCwb) / (A - B vs) (4.31) fa = 1 — 0b * [(100 — MCwb) / (A — B vs) 129 + (Mcwb / 105)] (4.32) porosity, decimal free air space, dec1ma1 where f fa and all terms are as previously defined. 4.9.2: Sensitivity Analysis Variance due to measurement errors was analyzed in a manner similar to that of the previous sections. The effects of sample weight, sample radius and height, moisture content and volatile solids were considered. The effects of changes in the parameters A and B were also examined. The partial derivatives of porosity with respect to the various factors are shown in Table 4.13. The partial derivatives with respect to bulk density, weight, and moisture content are all linear. The partial derivatives with respect to radius and height are similar to those for bulk density, varying with the cube of radius and the square of height, respectively. The partial derivatives of porosity with respect to the volatile solids and factors A and B are not linear but change with the volatile solids, A or B being divided by the square of the term (A - B*VS). The equation used to estimate the total variance in porosity is given below: 2 sf [(C/Vol)z * 5M2 + (0b C/h)2 * shz + (2 8b C/r)z * sr2 + (ob/(A-B*vs))2 * sMCwb2 + (0b*A/(A-B*VS)Z)2 * 5A2 + (pb*B/(A-B*VS)2)Z) * 532 + (0b*V/(A-B*VS)2)Z * sv2 (4.33) where C = (100 - MCwb) / (A - B VS) 130 Table 4.13: Partial derivatives of porosity calculation. f = l - 0b * [100 - MCwb) = 1 - (M / n r2 h) * (C) A '- 3 VS if = - ( C ) 30b 8f = - ( C )/ n r2 h 8E 3f = ( C )/ w r2 h2 85 8f = 2 W ( C )/ n r3 h 8? 3f - oh (1 / A - B VS) EfiCwb 3f = -0b(-B / (A - B VS)2) 83 at - -ob(B / (A — a v5)2) 88 at = -pb(VS / (A - B vs)z) 3;" 131 The variances contributed by each of the factors of porosity variance are shown in Table 4.14. Three levels of variance for each factor were used; they were labeled Low, Medium and High. The above equation was applied to 75 porosity measurements. The porosity variances presented in the following figures are given in terms of actual porosity Table 4.14: Error levels of contributing factors to porosity and free air space variance. Error Error Level Source Low Medium High Net Mass (9) 10.001 10.01 10.1 Core Height (in) 10.125 10.250 10.375 Core Radius (in) 10.125 10.250 10.375 Moisture Content (% wb) 11 15 110 Factor A 110 1500 11000 Factor B . 11 150 1100 Volatile Solids (%) 11 15 110 values and not as a percent of the actual porosity. The total porosity variance for the different variance levels is shown in Figure 4.18. For all three error levels, 80 percent of all measurements had variances less than 2 percent porosity. The higher variances occurred with the lower porosity readings. These corresponded almost entirely to the deepest samples in the center of the windrow. One characteristic of these samples was that they were typically one-half the length of other samples due to the length of the core sampler. It is also interesting to note two linear regions in the data, one that corresponds to 132 A 0 00 Q >. 5‘ a .III 0 m 8 4... 2 ° . v o 0 .— u 3 + ° 0 C .2 . o° 1‘ o g 2-( 4- 4‘ 6 o 1' 2: + + .. In 1.. 1 1’ °o o 8 O... ‘Cpooox O o "4’a°~0.6 0. 4. ' 43333 O U o o m « £ 0 ' mGLYILanqfi ‘ 1 I 0.9 0.92 0.94 0.98 0.98 1 Porosity (1) a Low 1 Medium 0 high Figure 4.18: Total porosity variance for 3 levels of factor variance. porosities of 95 percent and greater and the other corresponding to porosities between 90 and 95 percent. Closer examination of the high error case shows the contributions of several factors to the total porosity (Figure 4.19). Variance in determination of Factor A contributes almost all of the variance in porosity with the radius variance coming in a distant second. The effect of errors in the measurement of Factor B, height, moisture content, volatile solids and weight are almost negligible. Factors A and B arise from a linear regression of volatile solids data to determine the specific gravity of the solids that was done by Bohnhoff et al. (1984). This measurement was not part of the current research. Given the sensitivity 133 :; V i U 5.1 n v o A 3 A. “ V v A U V _ A 3 3 A V v on A C A a” 2 a- .1 ' >4 5 9 t: A W: '0 Ab‘fivq. O 1 ‘- U A: V L. \ a x Email a. h + + A I I 0 q—a—p—I—fi—fiiifi-yt-Magmfieém— u—.———( 0.9 0.92 0.94 0.98 0.98 1 Porosity (1) 0 Height 0- Radius 0 MC A A x B v Total Figure 4.19: Contributions to total porosity variance due to high factor variance. ) of porosity to Factor A, this measurement should be performed in future work of this nature. The partial derivatives of the free air space with respect to the various factors are shown in Table 4.15. The results are similar to those for porosity with a slightly different product. The partial derivatives with respect to bulk density, weight, and moisture content are all linear. The partial derivatives with respect to radius and height vary with the cube of radius and the square of height, respectively. The partial derivatives of porosity with respect to the volatile solids and factors A and B are not linear but change with the volatile solids, A or B being divided by the square of the term (A - B VS). 134 Table 4.15: Partial derivatives of free air space calculation. fa = 1 - OB * [100 - MCwb + MCwb] = 1 - (W / n r2 h) * (D) A - B vs 105 8f = - ( D ) '56: afa = — ( 0 )/ n r2 h EH afa = ( 0 )/ n r2 h2 FE afa = 2 W ( D )/ n r3 h 8? 8f = o ( -1 + 1/105 ) mewb b A " B vs at = -p (-A / (A - B vs)2) 5A3 b at = -0 (B / (A - B vs)‘) 53a b at = -pb(VS / (A - B VS)2) a. 135 The equation used to estimate the total variance in free air space is given below: sfa2 = [(D/Vol)2 * 5M2 + (0b D/h)2 * shz + (2 9b D/r)2 * sr2 + (pb*[(1/(A-B*vs))+1/105])2 * sMCwa + (0b*A/(A-B*VS)2)2 * 5A2 + (ob*B/(A-B*VS)2)Z) * 582 + (pb*VS/(A-B*VS)2)2 * svs2 (4.34) where D = ((100 - MCwb) / (A - B*VS)) + MCwb / 105 Free air space values have a much greater range than porosity: 35 to 87 percent. Variance due to errors in measurement are also muCh greater than those for porosity, with the maximum variance for each error level being 19, 11.5 and 4.5 percent free air space for the high, medium and low levels, respectively. This is shown in Figure 4.20, free air space total variance. Contributions of various factors to free air space total variance are somewhat different with errors in radius measurement being larger than errors in the measurement of Factor A (Figure 4.21). Errors in the measurement of other factors are negligible. 4.10: Additional Physical Properties A number of other physical and moisture related parameters can be calculated from the results of the bulk density, porosity and free air space calculations. These parameters can provide additional insight into the internal state of the windrow. Gravimetric moisture content provides information on the weight of water present in a given windrow volume, which is important for calculating specific heat. The relative %) (FAS, FAS Variance 136 19 18 17- 18-( 134 14- 13- 12.. 11 - 10- " 0 DO 00 ° 6 ghb Figure 4.20: %) (FAS, FAS Variance 19 I I I 07 (t) 0.5 Free Air Space Low Medium High ‘“ Oinfiflgii E‘szzfiiggii O u . Total free air space variance for 3 levels of factor variance. 18 17.. 18 1.1 1“ 15~ 12- 11— 10- 9- ad 7— 8- 5.1 44 3... 2.: ‘— 0 0.3 , R—F—fl—I—ll—C 0.5 Figure 4.21: 0 Height Free Air Space + Radius 0 MC AA (3) X B v Total Contributions to total free air space variance due to high factor variance. 137 weight of water, as opposed to that for the dry DMS, can mask important differences in the relative volumes of air, water and solids. Volumetric moisture content presents information on the volume of water present in a unit area. The volumetric moisture content can be calculated by 0 = Mwb 8b / 0w (4.35) where 8 = Volumetric moisture content, decimal 9w = water density, kg/m If the values of porosity and free air space are known, 8 can be calculated as follows: 0 = f - fa (4.36) The volumetric solids content can similarly be calculated by f = l - fa (4.37) where fs = volumetric solids content, decimal The void ratio relates the volume of the void to the volume of solids. The definition is given as follows: 8 = (V air + Vw) / Vd = Vp / (Vt - vp) (4.38) where void ratio, decimal volume of x air water pores total #15421»: n Void ratio can also be calculated-from the relationship 8 = (1 - f) /.f (4.39) where f is the decimal porosity. Equation 4.35 will be used in this study. The degreee of saturation, that is, the volume of void space occupied by water, is calculated as follows: 138 s = vw / vp = vw / (vair + vw) (4.40) where s is the decimal degree of saturation. The degree of saturation can also be calculated by s = l - (fa / f) (4.41) Equation 4.37 will be used in this study. 4.11: Windrow Size Changes Windrow size and shape changes were measured by two methods. The main observations were made by measuring the distance from the windrow surface upward to a grid of equally spaced co—planar wires stretched horizontally over the windrow. Before each set of measurements was taken, the height of the wire at the center was standardized by adjusting wire tension. The distance between the grid and windrow surface was determined by holding a plumb bob at the grid location and measuring the distance between the wire. and the point on the windrow surface that the plumb bob touched. Thirty or 36 locations were measured in this manner on the lst, 3rd, 5th and 7th days after the windrow was turned. The distance from the top of each probe to the top of the surface next to the probe was also measured. All probes touched the floor. Measurement error was 1 1/8 inch. Only one side and the top of the windrow were measured by this method. CHAPTER 5 EXPERIMENTAL RESULTS 5.1: General Observations Two windrows were monitored: one in the summer and one in the fall of 1986. Windrow turnings are designated A, B, and C. Ambient conditions during the summer windrow (Windrow 1) were generally warm and humid. The fall windrow (Windrow 2) was preceded by cold and wet weather. Data on ambient weather conditions are presented in Figure 5.1. Ambient temperatures rose during the first week of Windrow 2 but the overall weather was colder and rainier than for Windrow 1. Due to difficulties in correctly programming the relative humidity sensor, one third of these data were not collected. 5.2: Windrow Size and Shape Windrow size and shape varies between windrows and turnings. Figure 5.2 shows the cross-sections of Windrows 1A and 2A. The changes between windrows were due to the amount of material in the windrow: Windrow 2 had half again as much material as Windrow l. A summary of windrow weight and sizes is given in Table 5.1. Differences that occur when a windrow is turned are due to difficulties in constructing successive windrows to exactly the same shape 139 .140 .mcowumpcoo umzumm3 ucmeE< "H.m muoomm .483 2.: ON or Nr 0 Q 0 D F D b D I P b D h D h .r hi i I b b b h F mmaw mam: acouu Hmwumcw ma aoupcwz "m.m ousmwm A50 .x. acoum soupcmz 0.09. 0.094 0.0mm. 0.00m. 0.0mm Odom II?) ”.1111 \l i H\\\mflHHHHHH HMWHHHHHHH k\\\HMHHWIM p IJHIIIIF 0.5 7.0530 30.3353 II“ I (M pug MOJPUIM (m3 'A) 144 .x003 wco 000mm momcmno cowum>mam mam: ucoum ma soups“: «¢.m wpsmwm .60 .x. accum 3000:“: oodmv 00.00». 00.09.. 00.00». 00.0mm oodom Godm— oo.oo_. oodm 7 — a _ — J. _ q 1 A . (1‘4 4 _ _ P _ L oodm 00.00— 05 23:00 30.853 H pus MOJPUIM ‘3) (mo 145 (cm) 8 I ma Windrow Height I I l I I I 140 1673 1:13 203 I I o I I I I I I I I I I O 20 40 60 80 103 Windrow Base (cm) Day 1 Day 3 Day 5 Day 7 I 120 Figure 5.5: Cross section of Windrow lB elevation changes. Op “‘4 Cross section of windrow showing distances from a monitoring point to windrow surface along the probe (Dp) and the nearest surface point (Ds). Figure 5.6: 146 temperature or gas concentration readings that reflected ambient conditions. .2111 Temperature 5.3.1: Temperature Variablity A key assumption of the experiment is that the windrow is long enough so that transport processes along the length of the windrow can be neglected. While this assumption was not directly tested, data from Windrow 2 can be used to assess whether the three replications of the 9 locations were measuring the same phenomena. Because of the small sample size (1 for each turning) it is hard to draw conclusions about variability in temperature by location. The variance of temperature between the three replications of each location is calculated for each 15 minute average temperature for the three turnings of Windrow 2 and are shown in Figure 5.7. Windrows 2A and 2C exhibited average variances within the experimental measurement variance of 1 1.5°C with the exception of 1 point in each windrow. Windrow 2B exhibits rather large average variances, with 6 of the 9 points having variances larger than the variance due to instrument error. An analysis of the patterns of variability proved instructive. If there is a longitudinal effect, one would expect the left (L) and right (R) replications to be closer together than L and the center (C) or R and C. Longitudinal effects are assumed to be present if the difference between Hindrov 24 57.6 50.0 43.1 47.6 46.1 35.6 32.3 36.6 56.5 I 0 31 0.99 0.05 0.23 0.70 2.85 0.32 0.08 1.43 0.47 1.09 1.15 3.41 1.45 0.84» u v u u v u n u N f Figure 5.7: 1147' Uindrov 28 Hindrov ZC 51.7 57.3 58.0 57.3 52.5 62.2 52.5 57.1 60.9 52.5 39.6 58.3 49.8 42.8 12.1 57.9 49.6 12.8 Average Ieaperature (’C) 1.66 0.19 .55 0.74 3.51 2.96 2.17 . 6 0.17 0.12 1.10 0.27 2.97 0.83» 7.15 1.03 0.78 0.70 Variance (’C) 2.03 0.74 1.26 0.33 3.63 2.18 2.59 1.68 .68 0.93 1.79 0.63 3.42 2"1. 4.99 1.55 1.69 I.72_ I line Sale Sign Y N Y Y Y N N N Y Y n v n v f u u u v f Sililar Shaped Curve Measurements of temperature variability in Windrow 2. 148 C & L and C & R showed the same time-temperature pattern and the percent of time that they were the same sign. Figure 5.7 shows these results. Windrow 2A is low by both measurements, with same sign differences present at only 46 percent of the time and only 2 of 9 locations having similar patterns. Windrow 23 has 4 locations with similar patterns and the differences have the same sign 79 percent of the time. Windrow 2C has the highest number of locations with similar patterns; 72 percent of the time the differences have the same sign. No pattern is obvious: Windrow 23 has high values of this parameter and variance while 2C has slightly higher values but low variance. This is also true for the number of locations where temperature differences were judged to have the same shape. Based on Windrow 2A and 2C, it appears that the longitudinal variability is negligible. Due to higher variability in Windrow 28 and the effect of ambient conditions on the composting process, however, this conclusion must be tested further. In addition to measurement errors, there is also the effect of the probe placement to be considered. While each probe was placed so that the locations were approximately the same in the y-z plane, it was not always possible to achieve this. Thus, differences in y-z plane location could affect the measured temperature variability. It is also possible that the higher variance in temperature is due to the presence of 149 different transport processes in the windrow, i.e. natural convection as opposed to diffusion. Nonetheless, for purposes of analysis, the replications will be considered to be averageable. 5.3.2: Temperature Patterns The maximum temperature in either windrow is 68°C; most of the windrow tend to be in the 45°C to 55°C range. Temperatures near 35°C are seen deep in the center and at the outer edge of the windrow early and late in the runs, respectively. Temperature profiles are shown in Figures 5.8 through 5.10, for Windrow l. Temperatures in locations near the surface, 1, 2, 4, and 8, are initally high and either rose to a near constant level (Location 8 in Windrow 1A, Figure 5.10) or peak and fall off (Locations 1, 2, and 3 in Windrow 18, Figure 5.8). Towards the end of a given turning, the surface temperatures tend to reflect ambient conditions. This could be due either to an actual reduction in respiration in those areas of the windrow or in the case of locations 8 and 4 (Figures 5.10 and 5.9), due to the thermocouples becoming exposed late in the runs. Locations deeper in the windrow such as 3, 5 and sometimes 9, show either the same pattern as the the surface temperatures, i.e. a rise, plateau and slight fall (Locations 3 or 5 in Windrow 1C), or develop in an "S" type pattern (Location 3 in Windrow 2A). Locations 6, 9, and 10 tend to exhibit mainly the "S" type temperature profile (°C) 4 Temperature Figure 5.8: (°C) Temperature Figure 5.9: 70 60 'so 4O .90 2O 70 00 40 20 150 - 3 3 1 2 2 ‘ 1 2 1 1 - 3 2 AME T o r” 4 ‘ a . 1'; V If to 20 Time (days) Temperature profiles for Locations 1, 2, and 3, Windrow l. 0 4 ll 13 16 1‘0 l Time (days) Temperature profiles for Locations 4, 5, and 6, Windrow l. 151 70 9 co - A 8 10 a? 50 "' 9 0 SLAB 65 w _ 10 g SLAB 8 “ 5° + ‘NW/VW s 0 {- 20 - MB '1‘ IO - ' o r V I I I Y 1' I I V I V fit I r V T r I O 2 4 B B 10 ‘ 12 14 16 1B 20 Time (days) Figure 5.10: Temperature profiles for Locations 8, 9, 10 and Slab, Windrow 1. (Location 6 in Windrow 1A). The temperature profile in the windrow at any given time can yield important information on the locations in the windrow that reach a given temperature. Cross-sections of Windrow 1A taken at noon on days 1, 3, 5, and 7 are shown in Figures 5.11 through 5.14 for illustration. Figure 5.15 presents the same information for day 7 of Windrow 2C. Since temperature was not measured on the windrow surface or in the front of the toe area and since the thermal properties of the compost and air are different, assumptions about the temperature in those areas had to be made. The five Figures (5.11 to 5.15) are good for illustrative purposes only and do not reflect the actual temperatures at 152 on. .cooc up A woo co cowuummummoLo 4H 3oupcwz L0m mmaflmoLd Auov mucumLmQEme AEU .>v mmmm 3oLccm3 "Ha.m musmwm OO— 3ufiiaH AOqugm 42) (mo 153 .cooc um m xmo co comuommnmmoLu 4H zochmz L0u mmawmoua Auov musumeQEmB "NH.m mnsmmm AEU .r. omen Jauucmz om— . oo_. _ . ovr_|\.1\I/\u OO— 3qfiian noxpugm 12) (m3 154 .coo: an m won :0 cowuummummoLu 4H Boupcwz LCM mwawmoLd Auov musumumdsme "ma.m mpsmmm AEU .rv ommm zohpcmz om_ . oo— oo— iqbgau noapugm Iz) (mo 155 .h woo co cowuowmummouu 4H 3ochfiz L0u mmamuoLQ Auov muaumumaEmu Hmcwm ”¢H.m wuzmflm .Eu .r. oop q wmmm zoLocmz oo— JQEEBH MOqugM 42) (mo 156 .UN 3ocpcfiz L0H mHMLOLQ Auov mnsumhwdsmu Hmcwm AEu .>V amen Jaupcmz Omp . COP "mH.m musmmm oo— aubxaH MOJPUIM (m3 '2) 157 the toe and surface. In general, the temperature profile of a given windrow tends to have the highest temperatures at some distance from the surface, with slightly lower temperatures at the surface and much lower temperatures deep in the interior of the windrow. With time, the zone of higher temperatures moves inward and the edges cool. 5.3.3: Time-Temperature Relationships A key parameter in predicting the thermal death of pathogenic micro-organisms is the length of time a microbial population remains at lethal temperatures. Table 5.2 summarizes the time-temperature data for locations monitored in Windrow 1. Similar data for Windrow 2 are presented in Table 5.3. Figures 5.16 through 5.21 show the length of time that a given location remains in a given temperature range for Windrows l and 2. 5.4: Gas Concentrations Concentrations of oxygen and carbon dioxide were initially calculated using the method described in Chapter 4, unmodified. Examination of the data, however, indicates higher oxygen concentrations than expected in the interior of the windrow (Location 10, Figure 5.22). Under forced air composting, oxygen depletion is very rapid and quickly falls to levels of l or 2 percent. While the DMS composting system is not expected to produce such a quick reduction in oxygen concentration, because of the length of time between turnings, oxygen levels are expected to be in the same Table 5.2: 158 Time-temperature data for monitoring locations in Windrow l. Length of Time (days) Location Temperature Range (°C) 30-35 35-40 40-45 45-50 50-55 55-60 60-65 65-70 Windrow 15 Total Time: 6.96 days 1 0.35 1.74 2.62 2.25 2 0.49 0.47 1.57 4.43 3 0. 88 0.77 1.61 3.71 4 3.15 2.56 1. 24 - 5 0. 65 0.44 0.70 1.12 4.09 6 1.74 l. 34 1.42 2.46 8 0.52 0.46 1.01 4.97 9 0.89 0.81 1.40 3.86 10 3.18 2.48 1.30 Windrow 12 Total Time: 6.88 days 1 1.44 2.28 2.70 0.47 2 0.30 0.28 2.99 3.31 3 0.39 1.09 2.38 3.03 4 1.50 2.11 2.27 5 0.65 0.56 0.72 1.33 3.62 6 2. 98 2.04 1.47 0.40 8 0. 35 6. 53 9 0. 35 0.99 2.79 2.75 10 1.69 2.15 2.74 0.30 Windrow 19 Total Time: 6.81 days 1 2.38 1.73 1.22 1.49 2 1.71 2.24 1.88 0.99 3 2.63 1.84 2.34 4 Probe exposed to air. 5 0.24 0.27 0.62 5.68 6 1.34 0.74 1.00 3.73 8 0.22 6.59 9 0.72 0.49 4.81 0.80 10 0.20 1.77 0.88 1.13 2.84 159 Table 5.3: Time-temperature data for monitoring locations in Windrow 2. (days) (°C) 40-45 45-50 50-55 55-60 60-65 65-70 Length of Time Location Temperature Range 25-30 30-35 35-40 Windrow 25 Total Time: 5.80 days 1 0.19 0.68 3.27 1.67 2 0.08 0.45 0.73 1.92 2.62 3 0.98 1.00 0.71 0.52 1.21 1.38 4 0.10 0.42 0.58 4.70 5 0.01 0.86 0.74 0.86 1.29 2.03 6 1.26 1.51 1.45 1.58 8 0.09 0.59 0.66 4.46 9 0.52 1.33 1.35 2.28 0.31 10 1.13 3.38 1.29 25-30 30-35 35-40 40-45 45-50 50-55 55-60 60-65 65-70 Windrow 2g Total Time: 6.842 days 1 0.01 5.12 1.20 0.51 2 0.08 0.06 1.83 2.22 2.00 0.65 3 0.26 0.15 0.12 2.99 3.32 4 0.06 0.18 2.79 3.72 5 0.14 0.56 0.67 1.06 1.01 3.41 6 0.06 2.29 1.28 1.67 1.54 8 0.21 0.32 0.59 5.71 9 0.15 1.34 1.18 1.51 2.38 10 2.36 2.85 1.62 25-30 30—35 35-40 40-45 45-50 50-55 55—60 60-65 65-70 Windrow 29 Total Time: 6.88 days 1 2.38 2.66 1.83 2 0.19 1.56 3.74 1.46 3 0.26 0.31 0.40 2.89 2.84 4 0.28 0.52 6.07 5 0.54 0.55 0.57 0.90 1.34 2.97 6 2.26 1.40 1.56 1.66 8 0.29 0.55 4.72 1.31 9 0.11 1.84 1.08 1.96 1.88 10 2.15 3.85 0.88 160 .mwmp mm.m mm3 zochL3 Lou memo Hmuoe .4H 30chw3 cw comumuoH up mmmau wusumuwdswu comm cm ucmdm mafia "mH.m mulmwm 33333993 on L a...333399331U 3333399330 fl 1. I a .n .n I In 6 rt 1' 1? xn 3 J. CO CO 3.6036 0.2533,.» npooflro 2833399336 E333399.33.a .K33339339ma v— r. rp .u .n Tn fin 3 I 1' 1' iv 6 J. J a 8:83 r o c883 6 n 8:83 5 2333399331. 3133339930? .. f Tn E r». Yn I. I. J. J. 30 CO nanoflro opooflf «UV 59:0: «005 839.065» Ounce-339090... _ 8.88.. 0.... IO .— 1" I (“09) "no 0! own .mxmp mm.m mm3 aoupcwa Lou mew» Hmuoe .mH soups“: cm cowumuofl sh mmmHu mLDumLmQEmu comm cm ucwdm mELB “5H.m munmwm ADV 59:01 :00 23960:...» or. no Do on On 01 Ov an o». or. Go 0.0 on ad 0v 0. R 9.0 we. no Do on 00 av at on one P. no on on on 9 9v 0.» o... x IIJHI—I_ .414. .4dIIIIIII. f. 1. T3 5 a .n x A I” .1 T l I .1 I 1.. I. .r. o. c2603 r 0 3.083 r n .3803 6 u covuva... r. onnoOonaOonvgnnOn £30033anan RaconnOnoJOvnnoao r1“ 1—14—1-1-16 .1-1-1116 I 1 l v. #3 r3 a c383 fl n 8.503 fi n 8283 U r 0 Ragaavgnng Raccoonoanvovnfl fl! (5 62 fi . 1'7 ’ JI‘FE JJTFK u 8.603 v 8.503 (1’09) "41) 0! our” 162 .mme Hm.m mm: 3ouncm3 no“ mama Hmuoe .UH zouccwz cw cowumuoH >9 mmmau muzumquEmu comm cm ucmam meme ”mH.m musmfim AUV IDS :65 £39509...» 2283399291.." P82fi399n.9no R3833??nnomu 838339338 v. _ a. I a 7.. E x 3 I fio 1v 3 I I I l 38803 F 0:383 r ncavnofl 6 .233?) f or.8333??gnawV RBfl-nmlonimv7220 2383399223 — I l -—_— f Tu fin E I in. 1n I xv ..v J. I» a «5.603 f ncovuofl F. «8.603 F omnwmwnoowgnnomo omfipaflponwownnoao ; r fn .n A ,n T? 1' I... I». neon—.03 J .832: 6 (nip?) "mg 0' own 163 £823 OrnoOoL £893 or c3603 0001000». a c2503 0001000» u 8:83 FIT}. ..o '1 r53}. .mxmv om.m mm3 30uocw3 no“ mew» Hmuoe OhnuOonuOnnqunOmO n Cannon... OhndOoonooflvannoh v Con-duo... .4N 30uvcm3 cm cowumuoH xn mmmau musumumaamu comm cm ucwam meme uhndOonnO‘uonvanu III'|1—|‘—‘l-“-Il.-I|_lvo w « n. COVUOB R38389332 -o 7 '- N FT}; u c333 "mH.m whammm “UV :5 :05 83806:...» anooflnonvoendnx. . Conn X1. 1‘ J .- 3‘7 ’71" (o’x-p) ON") (1' an” 164 07.833 39‘Ovnn R3333? 2883 o— 8.503 «£383 009.an Gonna,» 3.1” J—JJI On u 60:83 O JW‘v’stT. .mmmc «m.m mm3 sauce“; new memo Hmuoe .mm zopucwz cm cowumooH >9 mmmao monummmasmu comm cm ucmam memb ObnuOonnbnninIflnOn II: M o Covuufl 283nu0001730n n c0983 m Ragnngflvvn v COPBQJ é a 6 A a I 63 Ragnngvgnngo n c383 "om.m mmammm 1 Tu vn J In. rm oundooanuanrvflnOn _ — — ..p N CODOOS 0 fl F313 2 A3 «0003 gram «8.5 Bancroft.» 3300308. ” cone X}. fl _ r. H... U. N W 3 N. O I > w. (n (It. 165 Bacon... 0009092. 0. Cop—woo..— Ragoaonnvgnn 4.314 2923 on a £2503 01093 a $0.603 64.33 mmmau musumuwQEmu comm cm ucmam meme "Hm.m munmmm 9» H 02 I .- 6W1}: .mamc mm.m mm: 30uccmz pew mama HmuOB .Um zopocwz cw cowumuoH an RagnuRnVOOnnOm I: n covuofi O vw F. VIII RnooonnonnvOoflan —_.. n copuos RQSnnoonvowonon > v Spoon.— Razongflvn‘nn n. Congo...— anoonnquvgnng ‘ N C0303 ADV :5 .905 Kain...» fl 3 2 3 399304 _ €09.3sz i f! JIIA (“99) "1‘3 u' uugl 166 20 19 0 18 a 17 a 8 1B- 151 14-1 13- :3“ WINDROW 1, LOCATION 10 10- 9-4 a- 7.1 81: 5d 4-1 * 4. 3‘ + + 2-4 1% on a D a +4» an %) ( 1+? + (II Gas Concentration + ++ -D I 1 3 5 7 9 11 13 15 17 19 21 Sample Date a C02 + 02 Figure 5.22: Gas concentrations at Location 10 showing suspect high oxygen levels. range. The experimental methods used in the gas sampling and analysis were re-examined as described in Chapter 4. In the course of this examination, it was found that ambient gases could have entered the sampling syringe. This was probably due to gas flowing into the syringe needle because of either diffusion or a pressure differential between the syringe body and the ambient air. The effect of calculated internal sample bottle pressure on the percent influx into the syringe was studied and a calculation procedure was derived to correct for this, based on the number of samples that had been withdrawn from the sample bottle. The effect of leakage during several weeks' storage time was found to be negligible, although further work in this area is necessary. 167 The difference between original and corrected oxygen data is shown in Figure 5.23 for selected data. The gas concentration data are presented in Figures 5.24 through 5.27. These data represent the average values for each location. Several trends are present in this data. At any given time, locations near the surface (Locations 1, 2, 4, and 8) have higher oxygen and lower carbon dioxide concentrations than those in the center (Locations 6, 9, 10, and sometimes 5 and 3). This is shown in Figure 5.28. Over time the relationship between carbon dioxide and oxygen depends on the location. Near the surface carbon dioxide starts high and decreases over time, while oxygen concentrations do the inverse (Location 1, Figure 5.29). A 2 ,9 v x x .. X x c ‘ x x O ... x x u 2 o 1;x x § x xx “ .9 ..* O A )1 4A 0 W X+ A ‘ 404 c: -‘ " A 5 ° 0 ‘4‘ x x. A 0 o U ‘ o 0 AA“ A o o 1. ‘co ° :1 m ‘ .30 O I o . co 1: "‘ o o a) u e -5 I I I I ‘I I r I I I I I 1 Iii ‘5 :5 s 7 9 11 1:3 15 17 19 U . Measured Concentration (2) + Low 0 High 11 Average x C02 Correct Figure 5.23: Difference between original and corrected oxygen concentration for selected data. 168 .H saunas: "¢N.m mLomwm Lo“ womb comumLucmucoo cmmxxo - so.” sm.o _~.a_ o~._~ . Kn.m_ ~_.o_ a_.o~ a_._~ _o..- _~._ on.». ~m.¢_ ...w .a... -.o~ m_.m_ o..o_ oa.o~ n..o- Ls... o~.m_ ao.o~ mm._~ mm.¢_ no._ on.“ .n.~_ vn.m_ - .m.n o~.. ~m.. .n.¢_ ,mm.~ n~._ n~.~ vs."— «v.. ~n.u ma.¢_ oo.~ a~.n ~n.m_ mo.n g..~ om.m_ an... on.n_ mm.~_ ...n. n..°_ ~_.m. - a~.~ as." «n.~ No.”— -o_.v um." nn.o_ .a.w_ .~_.n o_.. av.” mm.o_ a~._ «..n n_.¢_ no.” .m.n mo.c_ oo.o. Ln.n we... n«.o_ aw.e_ aa.o ca.n_ aw.” Lo.~_ h gag m an: n >a= l ox.~ an.~ w~.~ ~¢.m_ an._ n_.. m..._ o_.¢_ om.a_ - a¢.. mm.” .o.m an.” ...N .~.~ ~_.o oa.a .~.n_ .a.¢ «1.. an.. em.” m..a ma.. a~.°L .°.a _~.¢. _ sac 169 Log womb .m soaocwz cowompucmocou comwxo "mm.m mLommm .9." a... we... ...oa 1.x... ...a ._.__ _n.._ ..N._ ._.N .... ”a... - ...N ..._ an.” no... no.“ a... _..._ n..~ mu.” .m.._ a... x... mm.m_ a... a.._ ...o. a... mm... a... a... ma.. .n.. a... m... -..._ ...o .n... ...ou (a... ...m .N... ...._ -m... ...n m... xx... -..._ .~._ N..~ m..m_ "a.“ a... a..._ °~.~ N... 5°... ¢_.n ...N 5.... ~_.N wa.~ ...o. a... a... .0.» x..n_ m... a... ~_.~ a... .n._ a... m... ...m. ...n a... a... _..._ ...o n... N... N... ~_.~ ...o 9..“ s... .o.~ a... 5 us: n ».a n >~= a xaa 170 MUMU COM umhuchCOU .H 3oLoc.3 Lo. wowaMU coanU .mm.m muommm -2... a... .N.. 2.... 3.. .... 8.. ...o .3. .2. .... .... ..... ....N ..N. .... ..... ...o. 8.. .... o... .3. s... .... .... N... a... 2.. .3. .... 3.. 2.. a... 3... ...N. .... ..N -..... 8... .N... 3.. ,3... ..... 2... .... -2... ...2 ..... ...2 ..... 2... o... ..... z... .... s... 2.... N... s... ..... 3.... 2.. .... ...2 .... .N... .... s... N... -...... ...2 ..N. N... ...o. 2.2 2.... .... -..... ....N NE. .3. -..... ..... ..... 2... ..... ..... N... .N..N ...2 3.. ...N. ..... 9... .9... ...2 N... s... 2.. a... 2... ..... o... ...2 .... s an: n ..a n ..a . ..a 1'7]. .N 30.U:.3 no. mumn coquHucmucou momxowo conumu unm.m whammm ~°._~ on... ow.m vn.o n... mm.a_ aw.m nw.o on..~ sv._~ .m.¢_ .a.o v_.n~ ow.m~ vo.o~ mn.~ mn._~ -.o_ mo.m ~_._~ gs... .o.n ~a.o~ o~.a_ o~.n ao.- ~v.- no.0— ~n.__ «m... .o.m_ ao.o_ e“... am.n~ cm.a_ ~m.o_ - on..~ ow.o~ mm.m mm.o . oo.o~ av.m. .m.~ oo.ou .m._~ on... _~.~_ a... v¢.m~ mo.m~ mo.o~ .5.— mo... oa.~_ _~.~ an... an... .m._ o~._~ m~.a. vo.c v«.v~ vm.- vo.o_ .a.n_ .c.o~ no... on.v av.a_ ~a.~_ v~.v~ on.o_ . .~.- ~o.- ov.o~ so." so._~ aw.n~ -._~ nw.. mm..~ ~n..~ av... n~._~ ao..~ .N..— ~o.o~ ow... .o.o~ .n.n. . ..a n ..o n ..a . ..a 172 22 21— E 20- + 19- U ‘8- U 17- a D 15' a 15' 4' 14- 13-4 12-1 + I 11- 1O‘ 9" D B- B (%) +4- DD 4. on o 040 0+ 3. s— 4.. Gas Concentration +4”!- 4. 2- 1.. + T I O 2 4 ca +4 Windrow Location 0 C02 4 02 Figure 5.28: Gas concentrations by location, Windrow 13, Day 3. 22 4» + J 20- . + ; ¢ A ‘84 '. + 0 1 ” 8 ‘3‘ o + '0 + o c O o 14- ° 3 a , WINDROW 1, LOCATION l m 12.: O L. + u 5 10-4 9 U + C o B- D D U 4- m 3" "' 8 B D + D + _. U ‘ .a a E 2... a a a T D I: O T r I ‘ 1 .1 I r I I I I I I Y I I T I 1 1 O 2 4 8 B 10 12 14 15 1B 20 Sample Date a C02 + 02 Figure 5.29: Gas concentrations over time, Location 1. Lines indicate general trends in data, not intermediate values. 173 There is a partial recovery towards the initial concentration after the windrow is turned but the pattern then repeats itself. Locations in the middle zone (3 and 5, Figure 5.30) Show much longer durations of high carbon dioxide or low oxygen than those near the surface. Locations 6, 9 and 10 consistently show high carbon dioxide and low oxygen for the duration of the composting (Figure 5.31). Some gas samples from Locations 9 and 10 taken during Windrows 23 and 2C were analyzed for methane concentration. Low concentrations of between 0.5 % and 1.5 % methane were found. This result suggests that future studies include methane analysis as a means of more accurately determining the windrow reaction and transport processes. 5.5: Moisture Content Drying of the windrow surfaces was evident within a few hours after windrow formation. The compost material turned light brown, almost white, as it dried. The greatest penetration of the dry layer was only 0.5 to 1.0 cm, however. Blooms of small mushrooms appeared between 3 to 5 days after each windrow turning. Figures 5.32 and 5.33 present the wet basis moisture contents (% wb) by location for Windrows l and 2, respectively. Figure 5.34 shows the average moisture content by location for each turning. Moisture content remains essentially constant over the 3 week duration of each windrow, decreasing from 8) Gas Concentration Figure 5.30: (t) Gas Concentration Figure 5.31: 22 174 20- 18- 18- 14.- 12- 10- a— O WINDROW 1, LOCATION 3 + 'f-\ + 3. data, Gas concentrations over time, 12 Sample Date , a C02 +-02 Location Lines indicate general trends in not intermediate values. 22 0 D D O 20- ;///5‘\\\\ ////o"—__““a B D 0 a a 18- a c: g \\\\ 1 18- fl 0 ‘4‘ waoRow 1, LOCATION 10 ‘2- 10-1 8- a- * - \+ + " A- + + '0 \‘° * /., \ + + } 2.. Y\____* + + o I I I l I I I l I l I Y I I I I I Y I I O 2 4 8 8 1O 12 14 18 18 20 Sample Date cacoz . 02 Gas concentrations over time, Location 10. data, Lines indicate general trends in not intermediate values. 175 .... .... .... .... .... .... .... .... -.... .... .... ....11 .... .... .... .... .... .... ...... .... .... .... .... .... .... .... .... o... . ... .H aochMz Lo. :oMumooM xb ..bzw. .cmucou muoumMoe uMLumEM>mLU .... .... .... .... .... .... .... .... .... g. so..=.= -.... .... .... o... .... ..... .... .... .... .... 2 305.2: .... .... .... .... o... .... .... o... .... .... .. ....... . ... .mm.m musmMm ,o... .... .... .... .... .... .... .... .... .... .... o... .... .... .... .... .... .... .... -..o. .... .... .... .... .... .... o... .... . ... 176 .N aoupcMz Lo. coMumqu xn .n3w. acmucou muoumMoe uMLumEM>mLu .mm.m mLDmMm -.... .... .... .... -.... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .. ....... -.... .... .... .... -.... o... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .. ....... o... .... .... .... ..... .... .... .... .... .... N... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .... .. ....... . ... . ... . ... 177 Hindrov 1A Hindrov 2A 80.4 79.6 81.6 79.5 78.8 79.9 79.7 79.1 81.1 79.4 77.1 80.4 79.7 79.4 79.6_ 53.5 85.2 78.6 76.3 Hindrov 18 Hindrov 28 77.9 78.1 78.2 77.4 77.8 78.1 77.8 78.0 78.3 77.0 78.8 78.9 78.2 77.9 78.4 78.1 77.5 77 3 77.9 Hindrov It Hindrov ZC 51.5 77.0 77.9 78.1 77.4 77.3 77.4 77.4 77.9 77.9 77.5 78.0 77.8 77.6 77.3 77.6 77.8 77.9 Figure 5.34: Average gravimetric moisture content (%wb) by location and turning for Windrows l and 2. only from 81 to 77 % in both windrows. Average standard deviations of the measurements are 2.2 percent. Moisture content sampled after day l of each turning tends to have the smallest differences between locations. Samples taken on days 3 and 7 showed a greater range of moisture content. Locations near the surface and base (1 and 2) tended to be the driest in Windrow 1. This pattern is not present in Windrow 2, perhaps due to the accidental flooding with rain. 5.6: Volatile Solids Volatile solids measurements show small decreases within each one—week turning and over the entire three week composting period for each windrow. Figure 5.35 presents the average by location for each turning in Windrows l and 2. Figures 5.36 and 5.37 show the volatile solids content Hindrov 1A 94 95 94 95 95 96 95 95 96 96_ Hindrov 18 94 95 94 94 95 94 94 95 95 64 windrow 1C 62 94 93 94 94 93 93 94 94 Figure 5.35: 1778 Hindrov 2A 89 85 87 85 82 84 86 86 84, Hindrov 28 . 83 86 83 85 86 84 85 87 86 _ windrow 2C 82 82 82 83 84 83 83 84 84 Average volatile solids content (%db) by location and turning for windrows l and 2. 179 .. .. .. .. .. .. .. .. . .. m. .. ..1. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. n»: .M zoupcMz LCM coMumqu .n .no.. .cmucou we..om m...m.o> .. .. .. .. .. .. .. .. .. .. ....... .. .. .. .. .. .. .. .. .. .. .. ....... , .. .. .. .. .. .. .. .. .. .. .. ....... n .2. "mm.m musoMm .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. l .. .. .. .. .. .. .. .. .. . ... 180 ...m aoupch Lo. coMumqu >2 .npw. ucmucou mpMMom mMMumMo> .mm.m mLDmMm - .. .. o. .. -.. .. .. .. -.. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. ....._. . .. .. .. .. -.. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. ....._. . .. ... .. .. . -.. .. .. .. .. .. .. .. o. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. ....... . ... . ... . ... 181 by location for Windrows l and 2, respectively. The magnitude of the decreases are 2 % and 4 % VS for Windrows l and 2, respectively. Due to the volatile solids standard deviation of 7 % VS, however, this is not significant. Differences between the maximum and minimum values of volatile solids in each sampling period increase from day l to day 7, but due to the lack of replication and relatively high measurement error, these values are not significant. 5.7: Wet Bulk Density There are several difficulties with sampling for bulk density and comparing the results. Windrow bulk density can have large inherent variation. Too few samples can lead to results that reflect this variability rather than the underlying bulk densities. Taking too many samples can affect the transport processes in the windrow. Direct comparison of bulk densities is hampered because successive samples are rarely taken from exactly the same depths. Therefore, even if the variation of bulk density with depth is the same, differences in sampling depth lead to apparently different bulk densities for locations that are nominally the same. The data are best compared when used in conjunction with a model of bulk density. This has not been done, so the comparisons are somewhat crude. Initial experimentation to evaluate the method used to sample for bulk density indicated that the coefficient of variation is approximately 11 percent (See Chapter 4). 182 Figure 5.38 presents bulk densities by location for Windrow 1. Similar data from Windrow 2 are shown in Figure 5.39. The availability of data on the position of each sampling location in the windrow is given in Appendix H. The average bulk density by location in each windrow turning is shown in Figure 5.40. Comparison of the bulk densities for Windrow 18 with the overall bulk density calculated from the weight of the windrow and its measured dimensions provides another check on the bulk density sampling method. The gross bulk density of 442 kg/m’ compares favorably with the bulk densities shown in Figure 5.38. Even though the gross bulk density is greater than the bulk density at most locations, this is not a problem. The bulk density sampling method is not able to sample deep in the interior of the windrow where the greatest bulk densities are located. Experimental wet bulk densities range from 150 to 250 Kg/m’ near the surface to 300 to 650 Kg/m3 in the center of the windrow. Windrow bulk density shows the expected trend of increasing with windrow depth. Only 5 of the 82 sampling pairs (6 %) do not show this trend. Location 6 exhibited the greatest incidence of decreasing bulk density with depth with 3 of the 5 sampling pairs. Changes in density with respect to time are inconsistent. Figure 5.41 presents information on whether bulk density at each location increases from the beginning to the end of a windrow turning. With one exception,' 183 .H zochMs Lo. coMumqu an Mme\mx. qumch stn umz - ... ... ... ... ... ... ... ... -... ... ... ... ... ... ... ... ... ... -... ... ... ... ...w ... ... ... ... ... . ... ... ... ... ... ... ... ... ... ... .. ......3 -... ... ... ... ... ... ... ... ... ... .. ....... -... N.. ... ... ... ... ... ... ... ... .. ....... . ... “mm.m mLDmMm ... ... ... ... ... ... ... ... ..N -... ... ... ... ... ... ... ... ... ... .... ... ... ... ... ... ... ... N.. . ... 184 .m soupcMz Lo. :oMumqu so .mE\mxv qumch stn umz ... ... ... ... -... ... ... ... ... ... ... ... ... ... ... ... ..N .. ....... - ... ... ... .... ... ... .N. ... ... ... ... ..N ... ... ... ... ... .. ....... - ... ... ... ... -... ... ... ... ... ... ... ... ... ... ... ... ... .. ....... . ... n ... .mm.m musmMm -... ... ... ... ... ... ... ... ... , ... ... ... ... ... ... ... ... ... . ... ... ... ... ..N ... ... ... ... . ... .1835 Hindrov 1A Hindrov 2A 226 179 176 237 168 241 281 286 265 338 285 240 303 442 501 228 260 438 393_. Hindrov 18 Hindrov 28 170 131 ' 194 195 200 224 291 305 235 293 272 321 187 471 514' 219 315 436 462 Hindrov 18 Hindrov 28 147 206 200 181 214 333 276 239 409 310 214 313 320 511 253 351 375 477 Figure 5.40: Average wet bulk density (kg/m3) by location and turning for windrows l and 2. 186 approximately 68 % (s = 6.5 %) of the locations experienced increases in bulk density over time. Locations 6 and 9 (Figure 5.42) have the lowest incidence of bulk density increases over time with only 17 % (l of 6) turnings showing increases. When the effect of the 11 % coeffient of variation encountered in preliminary experiments is considered, the effect of time on windrow bulk density becomes even less conclusive. When this criterion is applied, decreases over time greater than 11 % were present as much as 56 % in some cases and were seen in all windrow turnings. 37 percent of the locations in each turning show changes in wet bulk density less than 1 ll % of the coefficient of variation. Increases in wet bulk density greater than 11 % occur at 36 percent of the locations while decreases in wet bulk densities greater than 11 % of initial values occur at 27 percent of the locations. Figure 5.43 summarizes these results. Until the core sampling method can be better calibrated, densities should be considered accurate in a relative as opposed to an absolute sense. Bulk density does appear to change with depth, but the effect of the process on bulk density changes with time was not clear with the method used in this study. §4§i Porosity Porosity at each sampling location is calculated using the results of the bulk density and volatile solids 187 .coMumooM an mEMu Lm>o mmmcmcu xuMmcmp xMSQ box .N¢.m wusmMm m=o_.~.o. .. . . . - o. .. . . . .. .. .. . w . ... .. .mEMu Lm>o mmmcmcu xuMmcmp stn .83 a.» u a. a.» a .. > > > > z z > z z > z » > > > > r > .. ....... .. ....... ... . .. . ... . .. z z z z z z > > x z > z z > > r z > > .. ....... .. ....... .o oo— nm co .M¢.m musmMm mo» a .o cw so.ch= a.» u om Z >- >.. (M ao.ucM= 188 Hindrov 14 Hindrov 18 Hindrov 18 Y N N N N Y N N I I Y Y -Y N Y -Y -Y Y Y -Y N Y Y -Y -Y -Y N -Y -Y - 10 50 4 40 - 30 40 4 30 - 56 22 1 22 Uindrov 2A Uindrov 28 81ndrcv 2C 8 N Y Y Y -Y Y N N N N Y N Y -Y Y N -Y Y» N Y -Y -Y ‘7 N Y Y Y - 22 56 4 22 - 22 33 4 4S - 22 22 + 56 Figure 5.43: Wet bulk density changes over time greater than 11 % coefficent of variation. calculations according to the method described in Chapter 4. The porosity measurement is therefore subject to all of the uncertainties described in the bulk density and volatile solids section. In addition, the method does not distinguish between porosity within the compost particles and that between the compost particles. Porosities for Windrows l and 2 are given in Figures 5.44 and 5.45, respectively. Average porosities for each turning are shown in Figure 5.46. Porosities vary from 0.98 to 0.97 at the surface of the windrow, to 0.95 to 0.92 in the windrow interior. As expected, porosity generally decreases with depth. The accuracy of the porosities cannot be evaluated until a bulk density model is developed. The effect of time on porosity 189 .H 3ochM3 LoM coMumqu up auMmoLOQ -..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... N.... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... . ... -..... ..... ..... ..... ..... ..... ..... ..... ..... .. ....... -..... ..... ..... ..... ..... ..... ..... ..... ..... ..... .. ....=.. -..... ..... ..... ..... ..... ..... ..... ..... ..... ..... .. ....... . ... .¢¢.m mLDmMm ..... ..... ..... ..... ..... ..... ..... ..... ..... -..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ....o ..... ..... ..... . ... 190 .N zoLocMz Lo. :oMumqu Mb qumoLom -..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... .. ....... -..... ..... ‘4 ..... - ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... .. ....... N.... ..... ..... ..... , ..... ..... ..... ..... ..... ..... ..... ..... ..... ..... . ..... ..... ..... .. ....... . an: a an: .m¢.m muamMm -..... ..... ..... ..... ....o ..... ..... ..... ..... -..... ..... ..... ..... ..... ..... ..... ..... ..... ...... ..... ..... ..... ..... ..... ..... ..... ..... . ... 191 81ndrov 1A Hindrov 24 0.972 0.976 0.980 0.970 0.977 0.969 0.964 0.961 0.969 0.959 0.960 0.969 0.961 0.941 0.933 , 0.973 0.967 0.942 0.944 windrow 18 Hindtov 28 0.976 0.974 0.974 0.973 0.971 0.973 0.959 0.956 0.968 0.959 0.966 0.956 0.959 0.933 0.928 _0.971 0.956 0.941 0.938 Hindrov 18 Hindrov 28 0.645 0.969 0.974 0.975 0.969 0.953 0.960 0.967 0.944 0.958 0.969 0.955 0.955 0.928 0.965 0.951 0.937 0.936 Figure 5.46: Average porosity by location and turning for Windrows l and 2. is similar to that for bulk density. Five of the six turnings had an average of 70 percent of their locations showing decreases over the 7 day compost turning period. Figure 5.47 presents this information. Porosity changes over time by location are shown in Figure 5.48. §;21 Free Air Space Free air space calculations are made using the results of the wet bulk density, moisture content and volatile solids content for each location as discussed in Chapter 4. Figures 5.49 and 5.50 give the free air space for each location in Windrows l and 2, respectively. Figure 5.51 presents the average free air space values by location for each windrow turning. Free air spaces of 0.87 to 0.71 are found near the 192 .coMumooM mo mEMu Lm>o mwmcmcu xuMmoLom .ma.m muomMm .8333 , o. . . . - o. o. .. ... . . . .. .. o. .. .. . . .. ... .mEMu Lm>o mmmcmcu muMmoLom ....m mLDmMm ... . .. ... . .. ... . .. . . . z z . z . z z x . . . . . . . z . . . . . . .. ....... .. ....... .. ....... ... . .N ... . .. ... . o. a z . - z z . . - . z . z a . . z . . . . . 2 3.3M: 3 3.2:: 5 3.2.5 193 -__~.° .H aouncfiz pew cofiumuoH an mumam umm mmum cnh.o ¢m~.° ”No.0 ¢o~.° ..~.° n-.° m-.° -.mu.o "no.0 ~.~.o ~am.¢ ”05.: aw~.° ~_o.o ¢a~.° m,¢.° mma.o ,cmn.° oug.o _.~.° m.~.o _n~.° c.~.° ”$5.0 ~.a.° n_a.o nn~.° N xaa 5 322:: m an: m...° ouh.o mac.o .oa.o ¢m~.° ..m.° «$5.0 ._o.o coa.° u. ,o.u=_= ~wn.o ¢°¢.° .nh.o m¢¢.° .w~.° .n~.° mwa.¢ ~o~.o “no.0 a.m.o a. ,o.u=_. , avm.o oan.° osm.° m-.° o_~.o ~n~.° __~.° ~.o.o mnm.o mah.o "m¢.m musmwm .an..o mn¢.° 5.5.o _m~.o ._~.° mao.° ..w.o n.o.° mo~.° - omm.o mn..° ~.~.° ~m~.° _¢c.o no~.° ¢m~.° www.° guc.c n.¢.o , ”Nn.o am..o mo~.° °_a.c ~n~.o _.~.° owo.o onm.° .«m.o _ ».a 194 -.N... .N sauna“; new cowumuoH up mumam ham mmpm ann.° ~n¢.¢ .a~.° _a~.° na..° ~a~.° .oa.o -o.o _o¢.° aom.° nou.o ~_~.. -~.o .o~.° .°¢.. .an.° . omm.° ¢~u.o ~m~.° nmh.o _.~.o _n~.° a_~.° na~.° .Na.° h ».a .mnm.o .nn.° a.w.o anh.o ._~.° ao¢.o moo.° o¢~.° ow .o.==_= .nn.° ”No.0 ~°~.° ~a~.° ”$5.0 ~.~.° .-.° nnm.o oaa.° .N ,o.u=_. ~°¢.o amn.o .a~.° ~n~.o ~am.o a¢~.° .m~.° ~.¢.° «N ,o..=_= n ».a "om.m muzofim -~¢m.o .om.o o-.° °a~.o $.m.° °°~.° w-.o nam.° m.c.° -n_n.o nhn.o _¢h.o mm~.o c.~.o -~.o ¢a~.o oNo.¢ a.a.° - aag.o nsn.° om~.° n.m.° a-.° nn~.° a-.° oa~.° snm.° _ >.¢ 195 Hindrov 18 Hindrov 2A . 0.790 0.834 0.845 0.776 0.739 0.735 0.777 0.718 0.589 0.535 0.754 0.688 0.740 0.527 0.760 0.598 0.643 0.935 0.792, Hindrov 18 Uindrou 28 ' 0.843 0.833 0.815 0.825 0.732 0.719 0.703 0.733 0.566 0.526 0.785 0.734 0.749 0.799 0.712 0.601 0.576 0.822 0.322, Hindrov 18 Hindrov ZC 0.531 0.811 0.818 0.835 0.803 0.694 0.746 0.783 0.627 0.717 0.804 0.712 0.705 0'529. 0.769 0.670 0.571 0.564 Figure 5.51: Average free air space by location and turning for Windrows 1 and 2. windrow surface. In the interior of the windrow, free air space ranges from 0.36 to 0.69. Free air space decreases with depth. No changes over time were observed (Figures 5.52 and 5.53). The coefficient of variation of the free air space measurements from the preliminary experiment is 7.5 %. Applying this i 7.5 % factor to free air space changes over time led to the result that for 69 percent of the locations, free air space does not show a change greater than 1 7.5 %, with 16 % showing an increase and 15 % showing a decrease in free air space for the 1 week compost turning period (Figure 5.54). .cofiumooH wo mew» um>o mwmcmcu mumam ppm ovum "mm.m mpsmmm m=o_.~uod -0. m a _ mm 2. cm co. a n N ¢_ 5. on mm on a . no ”a .mEMu um>o mwmcmnu mumam cam mmum "mm.m whomfim 196 mo» u 50 a.» n mm «o» u so > > > > z z > z > z z z > z > z > z p > z > > > > > ow :o.u=_: mm accec_: <~ so.ec_= no» u an no» u on no» m cm x a z z z a > > z z > z z > z z > > z > > > z > > > > a. .o.u=_= a. so.==_= <_ .o.u=_= 197 Nindrov IA Hindrov 18 Hindrov 1C 1 Y I N H Y N N N N Y N -Y N Y N N N N -1 N I N -Y -Y N N -Y -Y - 10 90 9 O - 30 60 9 10 - 22 56 4 22 Hindrov ZA Hindrov 28 Hindrov 28 N N H N N N Y I H N N N N Y -1 N N -Y Mr N Y -Y N Y N Y - 11 78 O 11 - 12.5 75 0 12.5 - 11 56 O 33 Figure 5.54: Free air space changes over time greater than 7 % coefficent of variation. 5.10: Additional Moisture and Physical Properties Volumetric moisture content ( 8 ), volumetric solids content ( fs ), void ratio ( e 1, and the degree of saturation ( s ) were calculated according to the methods described in Chapter 4. The results are presented in Figures 5.55 to 5.62. The implications of these results are discussed in Chapter 6. §4lli Windrow Color Changes When the windrows were pulled apart, three separate color zones could be seen. Profiles of the front half of Windrow l cross-sections are shown in Figure 5.63. Figure 5.64 shows cross-sections of the left and right side of the fron half of Windrow 2. The method of note-taking for these observations was not standardized until Windrow 2 was ..~.° a-.° -~.° m.~.° sa_.° s_~.° cm_.° om_.o 8 - mu oo~.° mom.o °.~.° -~.° a.~.° 3°~.° °¢_.o ~a..o nn_.° am_.° No..o _-.o a.~.° ~_~.c ._~.° o.~.o n~_.o _n_.° an..° m_~.° s ens .H 3ouocw3 no“ comumuoH an ucmucoo monumwos umbumEDHo> "mm.m whomwm -mm..o _m~.o hs~.o .s_.o ~o~.° .om.o s~_.o mm_.° _m_.o a. ,o.u=_. .m.n.° .mn.° n-.o ~n~.o oa_.° n-.° m...° .~_.o m~_.o -~.o ._ ,0..=_= -mam.o ~nm.o a-.¢ so_.o ~.~.o ._~.o o.~.o mn_.o -_.o -_.o ._ .o..=_= m ».a -mn..¢ ma~.° ¢m~.° mm_.o .w~.o awN.o .m_.° sn~.° _m~.° lasn.° mo..o m_~.° ¢-.o n¢~.o. an~.° ¢m_.o ~__.° a._.° ~m_.o so..o mm..o ~°~.o ms_.o -~.° .-.° wn_.o s._.° .w_.a _ ..a 1599 -m...o osm.° aon.° ma_.° s~_.o -..¢ ~o_.o a__.° a...° -~a~.° a-.° “a..° «.~.° _n~.° hg..° .5..° -mon.o m_n.° a°~.° m°~.° a_~.° an~.° n.~.o n¢_.o .m_.o a >.. .m Boupcwz u0m cowumuoH an ucmucou muzummos uwuumESHo> "mm.m whommm oom.o ~am.o mo~.° s-.o -owm.o o-.o mm~.o m~_.o mv~.o om~.o «on.o ww~.o ~m_.o mm..o oa_.o _v_.o ~m_.o UN soevc_a -ncm.o m_n.o mv~.o vh~.° -n_v.c ~mn.¢ vo~.o v~_.o nm_.o n_~.o ~o_.o e_~.o ~m~.o m~_.o ov~.o nn—.o mv_.o a~—.o an aoeac_= -nmm.° a.n.o °m_.o -.¢~.° osn.o nm~.° ~m_.o o_~.e nmn.o mo~.o vm~.o -~.o nw_.o no~.o mn~.o ~n_.o nv_.o cu socuc_a a sun _ x»: 200 .H 30hmqu3 hOw COMHMUOH an ucmucou mofiaom Hmcowuumpm -m.°.o ~.°.° ~.°.o a~°.o om°.° °.°.° mn°.° ~m°.° -nno.o ono.c ..o.o n.o.o a.°.° om°.° «No.0 «no.9 a~°.° ¢~°.° -_~°.° o.°u° _.°.° nm°.° __~.° _.°.° «No.0 n~°.° an_.° .mo.° ~ 1.: unm.m whammm -amo.o m.o.o ono.o ~mo.o moo.o _mo.o nwo.c vmo.o omo.o mno.o ~mo.o avo.o mvo.o w~o.o omo.c .no.o muo.o vmo.o u_ socuc_= . soo.o moo.o _vo.o svo.o -mwo.o m~c.o smo.c wmo.o wmc.o mvo.o -o.o vmo.o vvo.o avo.o n~o.o owo.o -o.o v~o.o mmo.o m~°.o a. scene—a -mmo.o mno.o neo.o ~mc.o woo.o omo.o mmo.° wwo.o «vo.o ~no.o _vo.o muo.o omo.o mno.o m~°.o ouo.c -o.o vao.o vuo.o c. soccc_= n 1.: _ 1.: 201 -:::.: :::.: 1::.: :::.: .::.: 111.: :::.: .::.: :1:.: -:::.: 1::.: :::.: :.:.: :1:.: 1:..: :::.: :::.: -:::.: :::.: :::.: :::.: :.:.: :::.: :::.: 1::.: :::.: 1 1:: .N 30uocmz uON cowumuofl mo ucmucou mpwaom Hmcomuomum 1mm.m mcommm -.::.: :::.: :::.: :.:.: - :.:.: .::.: 1.:.: 1::.: :1:.: :.:.: 1::.: :.:.: .::.: :::.: :::.: .::.: ~1:.: :: 1:.::_1 . 4:::.: :::.: :.:.: :::.: - :::.: :::.: .::.: :::.: :_:.: :.:.: 1::.: :::.: 1.:.: :::.: 1::.: 1::.: :::.: :::.: :: 1:.::_1 -:::.: :::.: :::.: - :.:.: m::.: 1::.: :::.: :::.: :.:.: :::.: :::.: :::.: :::.: :::.: :_:.: :::.: :::.: :: 1:_::_: m 1.: 1 1.: 202 .H 30poc13 10m cowumuoH >2 ofium: owo> .:.:1 ..1: :.1: .... .... .... :..~ ...: ...: -:.:1 1.:1 ..1: :.1: ...1 ...: ...: ...: .... :.:. -..:1 ...1 ...: ...: ...: ...: ...: :.:. ...: ...: 1 1.: ...: :.:1 ..:: ...: :.11 1... ...: 1.1: :1 .....11 -... ...1 ...: ..1: N..: ..:: ...: ..1. 1... n.:. :1 .....1: :.:1 1..1 ...: ..1: ...: ...: ...: ..1. .... ...: 11 .....1: : 1.: "mm.m musmmm :.11 1.:1 :..: ...: ..N: :.1: .... ...: :.:N -...1 1.11 ...: ...: ..1: :.:: ...: 1.1. :.1: :.:. 1... :.:1 :..~ 1... ...: ...: ...: :.1. :.:. 1 1.: 203 .m 3010:13 no. comumuofi >2 011mb U1o> ..11 1..1 ..11 ...: ..1: 1.11 1.:: ..1. 1... 1.:1 :.:1 ..1. 1.11 1.11 ... 1.:. 1... .1.:1 1.11 ..11 ..:. ...: ...: ...: ..1: .... 1 1.: ...1 ...1 1.11 :..1 :.:1 1.:. ..11 .... ...: 1: 1.1.11: ...1 1.11 n.11 .... ..:. 1.:. ...1 1.:: 1.:: 1. .....11 -...1 ...1 ..1: ...: ...: ...1 ...: 1.:: .1 1.1.:11 : 1.: “om.m whomwm ,1.:1 .... ...: :.:. :.:1 ..1. ...: :.:. 1.:. 1.:1 ...1 1.:: ..1: ...: ...: 1.:: :.:: .... 41.11 ..11 ...: .... ...: ...: .... ...: 1.:. 1 1.: 2(34 .H 3oupc13 pom cowumuoH xn co1umpsumm mo mwummo 1.1.: .11.: 1:1,: .11.. 1...: 1.1.: 1.1.: :11.: .:1.: .11.: 1.1.: :11.: 111.: 1.1.: :11.: 111.. 1.1.: :1 1.1.11: ,1::.: 111.: 1.1.: :11.: - .:1.: .:1.: .11.: :11.: 1.1.: 111.: ..1.: :11.: .:1.: .11.: 1.1.: :.1.: :11.: 1.1.: .11.: ::1.1 :1 1.1..1: ,1...: .:1.: ..1.1 .11.: - .1..: .11.: 1.1.: 1.1.: .11.: 1.1.: :11.: .11.: 1.1.: .11.: 1.1.: :11.: 1.1.: .11.: .11.. 1:1.: .1 1.1.11: s 1.9 m 11: "Hm.m musmwm 1::.: .:1.: :w1.: 1:1.: .m1.: 111.. :11.: .:1.: 111.: - 1:... 1:1.: .11.: .:1.: 111.: 1.1.: 1.1.: .11.: 1.1.: .:1.: 1...: :...: .:1.: .11.: .11.: 1:1,: .11.: :11.: ::1.: 1 1.: 205 101 :o1umu01 xn :OMumpoumm mo wmummo -11... .1 301oc13 1mm.m musmmm 1.... 111.: 1.1.. -..... 1.... 111.. 111.. -..... 111.. 1.1.. 111.. .11.. 1.... 111.. 111.. .11.. 111.. 1.1.: .11.: 111.. 111.. 111.. ..1.. ..1.. 111.. 11 1.1.111 .111.. .11.: 111.. -.1... 111.. .11.. 111.. -..1.. 111.. 111.: 111.. .11.. 1.1.. 111.. 111.. 111.: 111.. 1.1.. .11.. 111.. 111.. ..1.. 111.. 1.1.. 111.. 11 1.1.111 - 111.. 111.. 111.. .11.: , 111.. 111.. 111.. -111.. 111.. 1.1.. 111.. .11.. 1.1.. .11.: 111.. 111.. 111.. 1.1.. .11.. 111.. 111.. 111.. .11.. 111.. 111.. 1.1.: .1 1.11.11 1 1.. 1 1.: 1 1.: 206 E 63 X D 37 Z Windrow 1A A £11 0 542 D 46 z Windrow 1C 88 1 E D 12 X Windrow 10 Figure 5.63: Windrow l cross-section showing color zones. 207 Windrow 2A Windrow 2C Figure 5.64: Windrow 2 cross-section showing color zones. 208 started, so the profiles do not contain the same information. At the outside surface and in the toe area, the compost is dark brown color. Temperatures in this zone are moderately hot. A second zone of mixed brown and white colored material is underneath this first zone. This second zone is typically 15 cm thick near the center line of the windrow and 30 cm thick in the toe area. A well-defined color gradient is present in this zone with the white particles increasing towards the center of the windrow. There is a distinct white line at the interface between the second zone and the yellow third zone. High temperatures tend to correlate with the occurrence of the white particles; temperatures are very high in the white line interface. The white line is most noticeable in areas where there was a bloom of mushrooms on the windrow surface. The third zone is the color of the original uncomposted DMS, usually a yellow to yellowish-brown. The location of the three zones varies between the end of each successive windrow turning. This can be quantified by looking at the change in the percent of the total area represented by each color and by the depth of penetration of the black and white zone into the windrow at the centerline and bottom of the windrow. Table 5.4 presents the percent cross-sectional area for each windrow turning profile and color zone. With each successive turning, the average Windrow 2 Zone D area 209 Table 5.4: Windrow color changes. Windrow Color (% Area) Yellow Brown and Brown White 1A 37 63a 13 46 54 1C 12 88 2A Left 42 40 18 Right 45 28 27 Average 43 35 22 23 Left 33 38 29 Right 39 46 15 Average 36 42 23 2C Left 27 59 15 Right 34 56 10 Average 31 57 12 a Could not calculate brown and brown/white areas from notes. Area percentages represent non-yellow areas. 210 decreases. The percent area in Zone D in Windrow 1 decreases dramatically from Windrow 1A to Windrow 1C. The left and right halves of Windrow 2 are quite different. This is probably due to differences in windrow size and construction. Zone B in Windrow 2 expands inward with the interior edge moving faster than the exterior edge. Examining the penetration depth data presented in Table 5.5, the side and top penetration depths are very different. Top penetration depths stayed the same or increased slightly with time in both windrows. Average side penetration depths for Windrow 2 increased almost linearly with each turning. The top penetration depths are very consistent when compared to side penetration depths. Variances for top penetration depths are 0.33, 0.25 and 0.33 inches for turnings A, B, and C, respectively, while side penetration depth variances are 4.75, 98 and 108 inches. Side penetration is much more sensitive to the conditions encountered in each windrow: ambient temperature and relative humidity, substrate, and windrow size and shape. Table 5.5: 211 Windrow penetration depths. Windrow Penetration Depth (cm) Side Top 1A 89 33 13 76 36 1C 152 41 2A Left 79 36 Right 80 33 Average 80 34.5 28 Left 114 36 Right 69 36 Average 91.5 36 2C Left 107 40.5 Right 107 38 Average 107 39.25 CHAPTER 6 DISCUSSION 6.1: Experimental Methods 6.1.1: Probe Design Probe design was generally successful. Probe insertion did not cause the windrow to split apart during the one week turning periods. The probes could be inserted and were durable. Some difficulty was encountered with the repair of damaged gas collection lines; this was solved by taping new lines to the outside of the probe. New designs should allow easier probe assembly/disassembly. 6.1.2: Temperature Temperature monitoring equipment had a calculated measurement error of 11.5°C. Measured variations during calibration were much smaller than this on the order of 10.07°C. In retrospect, the ability of the experiment to test a key assumption could have been improved. Temperature and gas monitoring points were located in the windrow front half center section to allow replicate samples to be taken. This placement did not allow testing the assumption that longitudinal edge effects were negligible. While two of the three windrows with replicate monitoring locations did not 212 213 show evidence of longitudinal effects, these effects cannot be ruled out. The longitudinal effects could have been determined if probes had been placed in one quarter of the windrow. Additional measurements of temperatures at the windrow- air and windrow-concrete boundries would have improved the accuracy of temperature contour figures and aided future modeling efforts. 6.1.3: Gas Concentration Unexpectedly high oxygen concentrations in the windrow interior suggested that leakage of air had occurred during the experimental procedures. The gas sampling probe and withdrawal equipment functioned very well with no influx of air attributable to these components of the gas sampling and analysis system. The storage system that was initially adopted was able to function adequately over time periods of less than 2 weeks. Covering the septum top with silicon caulk stopped the sample from leaking out due to overpressures inside the storage tube and greatly extended the storage time. If this type of storage system is to be used in the future, longer term experiments should be conducted to determine the exact length of intact storage. The injection of a fluid into the tube while air samples are being withdrawn would also keep internal pressures closer to atmospheric pressure. The largest error in the gas system occurred when gas was withdrawn from either the sampling probe or from the 214 storage system. Correction factors based on experimental work (See Appendix D) accounted for most of the differences between the expected and measured values. Several correction factors were evaluated: low, high, and average influx. The low and high influx corrections were based on the observation of two separate linear segments in the influx vs. calculated sample pressure data. The low influx correction was based on the initial slope which occurred while calculated sample pressures were less than 1.0 atmospheres. The high influx correction was based on the steeper sloped influx vs. calculated internal pressure data. Calculated pressures in this case were less than 1.0 atmosphere. The physical explanation for these two sloped lines is most likely diffusion for the low influx segment and a pressure gradient for the high influx condition. The average influx approach represents a compromise due to the uncertainty of the actual initial pressure in the sample tube during sample removal. If this pressure were known, more accurate estimates of gas concentrations could be made. From a parameter estimation and modeling standpoint, the gas concentration data are probably accurate enough to illustrate basic trends and to do initial estimation and modeling work. Additonal data must be collected, however, to justify more rigorous parameter estimation and model validation efforts. Carbon dioxide data are subject to less 215 error than the oxygen concentrations because the influx mainly caused the addition of oxygen. The use of a gas valve on the sampling syringe would directly reduce influx and should be used in future experiments. A large quantity of gas samples were analyzed in this experiment. The available analytical method, a gas chromatograph, is quite time consuming. The use of equipment such as an infared C02 analyzer and a paramagnetic oxygen analyzer could speed up the analysis process considerably, particularly if it is onsite. 6.1.4: Gravimetric Moisture Content Gravimetric moisture content determinations had acceptable levels of measurement error with 12 % wb standard deviation. Measurement errors can be reduced with the use of scales accurate to 10.01 mg. Sample sizes with wet weights between 2.5 and 3.0 grams can further reduced the range of measurement error. Gravimetric moisture content determinations are also very time consuming. 6.1.5: Volatile Solids Volatile solids determinations made in this experiment were subject to unacceptable measurement errors with an average standard deviation of 17 percent. Dry sample weights of 0.7 grams coupled with the use of a scale accurate to 0.01 mg could produced standard deviations of approximately 1 percent. 216 6.1.6: Dairy Manure Solids Sampling and Bulk Density The rotary corer sampling method proved to be an acceptable way to sample compost materials for insitu bulk density, porosity and free air space determinations. Initial experimentation indicated that bulk density replicates taken at the same depth had coefficents of variation of 11 % for several depths. Windrow bulk densities obtained with the corer were in reasonable agreement with the overall windrow bulk density. This agreement could be more readily determined if samples were taken from deeper in the windrow's interior and if a windrow consolidation model was developed. Examination of the source of bulk density measurement variance indicated that the variance in effective radius measurement was the most significant factor. Future study of the core sampling method for composting materials should focus on the effects of corer speed, radius and compost material on bulk density determinations. 6.1.7: Porosity and Free Air Space Examination of the sources of porosity measurement variance indicated that errors in determining the particle density could affect porosity determinations. This was also an important error source in free air space determinations, but radius measurement errors were the largest source in this case. 217 6.1.8: Dairy Manure Solids Placement Dairy manure solids placement and operator skill were observed to have a large impact on the apparent bulk density of a windrow. Pushing DMS into position, dropping DMS in a large mass (as opposed to ”shaking" it out of the skid steer bucket), and leveling the windrow with the bucket all appeared to increase compaction. Although efforts were made to standardize the placement procedure, placement of DMS into the windrow probably had a large and unpredictable effect on bulk density and therefore on the transport properties in the experimental windrows. 6.1.9: Windrow Size Measurements The method of measuring windrow size changes had acceptable accuracy. Measurement errors were due to sag in the lines, problems with determining the windrow surface, and parallax. 6.2: Experimental Results 6.2.1: Temperature The temperatures encountered during both DMS composting runs were lower than those reported for uncontrolled sewage sludge windrows. A number of factors could have influenced this. The large quantity of water, as indicated by high moisture contents, could have absorbed the heat output of the reaction. This type of behavior was observed for decomposing oat straw in an adiabatic calorimeter by Bartholomew and Norman (1953) Conversely, the windrow size might have been small enough so that heat dissipated before 218 the temperature could increase. Finally, if the reaction rate was low, the heat output would be low and high temperature would not have occurred. 6.2.2: Gas Concentration There are two possible explanations that can account for the gas concentrations observed in this experiment, in particular the higher than expected oxygen concentrations in the windrow interior. The first hypothesis is that the reaction rate is limiting. In this case the flux of oxygen to the reacting sites is large enough so that oxygen is not rate limiting. The second hypothesis is that the reaction is transport limited. Oxygen flux, in this case, is great enough to supply all the oxygen demanded by the reaction. The high oxygen concentrations would be due to experimental error, in particular the influx mentioned earlier. The reaction limited hypothesis occurs under two different scenarios: (1) high transport rates/medium reaction rates (HTmR, small windrow), and, (2) medium transport rates/low reaction rates (MTLR, relatively non- reactive substrate). The system studied in this experiment is not a high rate system: oxygen transport is either by diffusion or natural convection, neither of which is a high-rate process. Forced aeration or even enhanced natural convection were not studied in this experiment. Therefore the possibility that the observed high oxygen concentrations are due to HTMR is not realistic. 219 The MTLR scenario is more plausible in light of the previous points. The lack of change in volatile solids levels indicates an unreactive substrate. There are several arguements against MTLR, however. The first argument is that volatile solids is not a good indicator of substrate consumption. The volatile solids measurement is not very specific: it cannot distinguish between the original substrate, added microbial mass and material that may be organic but cannot be degraded (Finstein, 1986). As is discussed later, the use of an ash solid basis is a better indication of substrate consumption. There is evidence that substrate is being consumed when volatile solids are expressed on an ash basis. If the possibility exists that the substrate is not limiting, then the low oxygen concentrations in the windrow interior become more troublesome for MTLR. In this case, these low concentrations could be limiting the reaction. In many microbial systems, including composting, the transport of oxygen across the water film is the limiting rate (Finger, 1975; Haug, 1980). While Haug's assertion is based on forced aeration systems that have high air flow velocities, Finger studied diffusion/natural convection systems that are similar to what is studied in this research. Furthermore, Shell's (1955) study of diffusion in windrows also pointed to the limited ability of the effective oxygen diffusion rate to supply oxygen to an actively composting material. 220 The strongest challenge to the short-term reaction limited hypothesis is that if the interior measured oxygen levels were not rate limiting, why were higher interior temperatures not observed? During temperature measurements for Windrow 1A, Locations 6 and 10 Figures 5.9 and 5.10) have an "S" shaped profile. Locations 5, 8 and 9, on the other hand, have steep increases followed by a leveling off of temperature. Substrate is initially the same at the all locations: it is not the cause of the different temperature histories. Possible high temperature limitations do not apply at Locations 6 and 10 as both these temperatures are solidly in the mesophilic range. Referring to the moisture content data presented in Figures 5.32 and 5.33, the wet basis gravimetric moisture contents at all locations are approximately the same, so no difference in the effect of moisture on the reaction rate or heat storage is expected. Examination of wet bulk densities (Figure 5.40) and free air space (Figure 5.51) indicate that the rate of transport of gases, whether by diffusion or natural convection, might be lower in the interior locations (6 & 10) than the locations closer towards the surface of the windrow (5, 8, 9). Given that other measurements do not provide a ready explanation for the shape of the temperature curves, oxygen transport rate limitations might provide an answer. In Figure 5.10, while Locations 8 and 9 are showing large temperature increases, Location 10 does not increase in 221 temperature. Only when the temperatures at Locations 8 and 9 level off does the Location 10 temperature increase. The increasing temperatures at the two higher locations increase the reaction rates and most of the available oxygen is consumed, leaving little to diffuse down to Location 10. As reactions at Locations 8 and 9 leveled off due to temperature inhibition, more oxygen diffused down towards Location 10. Oxygen ceased to become a limiting factor in the reaction at Location 10 and the increased reaction began to put out more heat, causing temperatures to increase. The course of the reaction at Locations 5 and 6 (Figure 5.9) is somewhat different. Temperatures begin to increase before the maximum temperature plateau is reached at Location 5. This could be due to a higher transport rate near the edge of the windrow (Location 6) as opposed to the windrow interior (Location 10). Evidence for this possibility will be discussed in a later section. Another feature of the temperature profiles that could be explained by oxygen limitations is the leveling off of Locations 6 and 10 before the higher temperatures of 5, 8 and 9 are reached. In this case, while the reaction rate at the higher temperatures is reduced, it is still occurring and consuming oxygen. Therefore, while more oxygen is being transported into the windrow interior, not enough is present in the free air space at Location 10 to permit a reaction that is not limited by low oxygen concentrations. If higher oxygen transport rates were present at Location 6, the 222 oxygen concentration might be higher than that at Location 10, although still rate limiting. The higher temperatures would be due to a higher rate of reaction caused by oxygen concentrations which are less limited in the free air space. A potential objection to the temperature level argument for oxygen transport limitations that heat conduction from the bottom of the windrow was enough to limit the temperature rise. This could indeed be an explanation, particularly if the ground was very cold, because of the large thermal mass of the earth. Measured slab temperatures during this period are very close to those at Location 10 and increase along with that temperature. In addition, one would expect this problem to occur in the winter when the ground was very cold. Both tests were conducted in the summer and fall, however, presumably after the ground had warmed up. Furthermore, DMS had been composted there for several months previously, so heat stored in the ground would have limited this effect. Based on the evidence presented above, I feel that a oxygen transport limitation is the most probable explanation for the gas concentration data. A common assumption in forced air composting is that the sum of measured oxygen and carbon dioxide concentrations equal 21 percent. This is roughly the concentration of oxygen in the ambient environment. The sum of oxygen and carbon dioxide at several locations in the compost windrows I studied was greater than 21 %, however. Initially, I 223 thought that I had encountered another indication of experimental error. Re-examination of the mushroom composting literature revealed that carbon dioxide concentrations could become as high as 30 % in windrow interiors (Lambert and Davis, 1934). The high carbon dioxide concentration in Lambert and Davis' study was correlated with zones of ”green" manure and they suggest that manure decomposition is retarded in this area. 6.2.3: Volatile Solids Volatile solids measurements had several problems associated with their use. The method chosen to measure volatile solids was subject to large measurement errors. Future work that uses volatile solids will need to use more accurate techniques, as described in Section 6.1.5, above. Expression of volatile solids on an ash solids basis is more sensitive to changes in volatile material than expression of volatile solids on a dry solid basis. Expressed as a percent decrease in the original ash fraction, the dry solids basis decreased only 8 %, while the ash solids basis decreased between 30 and 44 percent. The conversion to measuring substrate and substrate consumption on an ash solids basis correlates more closely with other indications that a composting reaction was occurring. The large changes in temperature, gas concentration and compost color can therefore be supported by evidence of substrate consumption. While the volatile solids determinations do provide a 224 rough measurement of biological activity, its usefulness is limited because it lacks both specificity and sensitivity. Finstein et a1. (1986) note: The test fails to discriminate among readily metabolized, putrescible material, less readily metabolized material, and organic material that is not metabolized during any reasonable composting period. Process performance is primarily concerned with the first of these, secondarily with the second, and not at all with the third. Also,...both fresh organic waste and stabilized organic residue are included in the VS test, which decreases sensitivity. Similarly, sensitivity is poor because a high percentage of the dry wei ht is volatile matter. This means that the decomposition of a large amount of VS may result in only a small change in percent VS. In the last 10 years most of the theoretically sound work with composting has been done with sewage sludge and garbage (Haug, 1980; Finstein and Morris, 1975). Volatile solids has been used successfully as an indicator of substrate consumption (Higgins et al, 1982) in those reactive substrates. I based my work closely on this experience. I neglected to consider the relatively inert nature of the DMS substrate, however, and volatile solids expressed on a dry basis showed little change. The mushroom composting industry deals with straw based substrates that are very similar to DMS. They express volatile solids consumption on an ash basis in their work (Burrows, 1951). It would appear that in dairy manure solids composting, volatile solids should also be expressed on an ash basis. Even though the use of the ash solids fraction allows substrate level changes to be observed, for detailed work 225 more specific measurements of the substrate are useful. The adoption of the procedures used in the mushroom composting industry as described by Mueller (1962) present an alternative. He recommended analyzing cellulose, lignin, pentosans (90 % of hemicellulose), ether and hot water extracts, and soluble and insoluble ash contents. Determination of the total carbon, total nitrate and ammonia nitrogen and nitrogen in lignin and alpha cellulose is also important for research work. 6.2.4: Moisture Content Moisture content changes expressed as a percent wet basis do not exhibit major changes in either of the two windrows: 81 to 77 % in Windrow l and 80 to 77 % in Windrow 2. This observation can be misleading, however. The wet basis expression can mask changes in the moisture content that are significant. Both windrows had final moisture contents of approximately 77 % wb. Initially Windrows l and 2 had 81 and 80 % wb, respectively. The corresponding percent dry basis values are 335, 426 and 400 % db. A 21 % decrease in the intial weight of water of Windrow 1 was observed; Windrow 2 had a 16 % decrease from the initial weight of water. Another advantage of the dry basis expression is that at the lower moisture contents it relates more directly to changes in the drying rate. Bohnhoff and Converse (1986) found that the point where the DMS drying rate changed from 226 a constant rate to a falling rate was at approximately 30 % db. This corresponds to 23 % wb. While the expression of gravimetric moisture contents on a dry basis is more sensitive to changes in water weight, the denominator is affected by the change in solids weight due to composting. Therefore it should also be corrected to an ash solids basis, although this will result in very large percentages. In addition to the problems mentioned with the wet basis moisture content, considering moisture contents on a gravimetric basis hides changes in water relationships that affect compost windrow heat and mass transfer. 6.2.5: Bulk Density Wet bulk density changed with depth but did not show a change over time. The change with depth is not suprising. In-situ measurments of bulk density at specific points in compost windrows have not been previously reported in the literature, however. Disturbed sample bulk density and gross bulk density have been measured but given the variation with depth found in this experiment, they do not adequately describe the local conditions inside the windrow. Given the importance that bulk density plays in a variety of yheat and mass transfer parameters, accurate bulk densities are crucial if distributed parameter heat and mass transfer modeling is to be conducted. Bulk density appeared to be relatively constant over time because of the relative lack of precision in the method 227 of measurement. The coefficent of variation of approximately 11 percent lead to the uncertainty of the bulk density determination of :17 kg/m3 for lower bulk densities and :51 kg/m3 for the largest bulk densities encountered. If longer term or high rate systems are to be studied, windrow consolidation and bulk density increases will need to be measured. Therefore, an improvement in the core sampling method to improve its accuracy and precision is imperative. 6.2.6: Additional Derived Physical Parameters A wide range of physical parameters are derived from the moisture content, volatile solids and wet bulk density measurements. The porosities encountered in this experiment are rather high, typically above 90 percent. This indicates that the pore spaces in this material occupy a high proportion of the total volume. Given that the compost substrate is a fibrous material with a high internal porosity and that the particle density of the DMS is fairly high, this is not too suprising. The volume of water also has important implications. This can be expressed in several fashions: free air space, volumetric moisture content and degree of saturation. When compared with the wet basis moisture contents that were fairly constant, all three of these parameters show that on a volume basis, water content increases with depth. The free air space measurement indicates that as depth 228 increases, there is less space occupied by voids where gases (:an be transported through. This can significantly reduce 'the transport of oxygen either through the void spaces or across the thicker liquid films that are present at greater depths. Experiments by Shell (1955) on diffusion of oxygen through a composting mass support this observation. The volumeteric moisture content (0) can affect thermal conductivity. Increasing the volume of water in the series- parallel models of thermal conductivity will increase k as water displaces air. Bohnhoff et al. (1983) observed just such a relationship between k and e. e is also more convenient to use when calculating fluxes of water through the composting mass. In comparison with other composting systems, the dairy manure solids system composts at very high moisture contents between 76 and 81 % wet basis. The 9 and high fa values point towards an explanation as to why composting could take place. Free air space values measured in the DMS windrow did not even approach the free air space limitations observed in other substrates at FAS values below 30 percent. This is probably due to the relative structural stability of the dairy manure solid particles that were able to form a porous windrow and the high moisture content of the particles themselves. The degree of saturation is useful in predicting the effective stress on a material when bouyancy forces are considered. Both the degree of saturation and the void 229 ratio are preferred expressions in soil mechanics where prediction of compaction and consolidation are important. Many of the properties discussed in this section are derived directly from the soil mechanics and soil physics literature. While they provide insight into the phySical processes in composting, a significant difference exists. Unlike most soil materials, compost particles are compressible. This means that under enough pressure, the internal porosity of a compost particle can be reduced. Therefore the nature and size distibution of pores may change significantly at greater depths. Intra-particle water may be expressed into inter-particle pore spaces. Particles that have been crushed during one windrow will respond differently to stress when redistributed and reloaded. Models of windrow bulk density must deal both with the effects of compressibility and changes in compressibility over time. 6.2.7: Thermal Properties An important simplifying assumption made by Finger (1975) was that thermal and physical properties in a compost windrow were neither spatially or temporally variable. From the results of this investigation, it appears that some of the physical properties do vary spatially. The question remains whether this physical property variance affects heat transfer properties. In order to evaluate the potential thermal property spatial variability, I calculated a number of thermal 230 properties using temperature, moisture content and bulk densities at 9 locations in Windrow 1B on days 1, 3, and 7. Empirical equations by Bohnhoff et a1. (1984) for thermal conductivity, specific heat and thermal diffusivity were used. The results of the calculations are shown in Table 6.1. Specific heat variability is very small in this windrow. The maximum difference between the largest and smallest values at any time is 2 percent. Thermal diffusivities exhibit greater spatial variability. Comparison of locations at any time reveals differences of 16 to 30 percent. Spatial variability in thermal conductivity is very large. The maximum percent difference ranges from 150 % to 320 percent. There is a pattern to this variability as well: the outer surface and windrow interior have low and high thermal conductivities, respectively. The low exterior thermal conductivities provide insight into why the interior the windrow tends to heat up. It would appear that the assumption of constant thermal and physical properties is not accurate. 6.2.8: Windrow Size Changes There was evidence of windrow size changes both during the windrow composting periods and between different periods. Typically, during a one week composting period, the center of the windrow would have the greatest decrease and the windrow toe would show a slight increase. The 231 Table 6.1: Effect of windrow physical property variability on the calculated thermal properties in Windrow 1B.- Day 1 Day 3 Day 7 50.8 50.8 49.7 52.4 42.8 51.2 57.5 53.9 53.4 62.7 64.9 58.5 55.7 68.1 61.4 59.4 53.3 45.7 44.9 65.1 61.5 51.0 48.7 46.0 57.7 60.1 55.6 Telperature, C (s = +l-1.5 C) 78.6 78.8 77.3 78.3 77.8 77.3 78.5 78.8 79.4 77.6 77.2 77.4 78.2 77.4 77.1 79.2 79.0 79.1 78.8 78.0 77.7 77.3 78.5 79.5 78.0 77.4 77.8 Hoisture Content, Ivb (s = 4!-2.2 106) 168 189 164 157 178 198 177 321 334 189 290 257 204 262 324 290 273 585 477 330 288 428 696 343 307 399 368 Bulk Density, kg/I‘3 (s = +l-11 I) 0.132 0.149 0.127 0.123 0.138 0.153 0.139 0.259 0.265 0.146 0.223 0.198 0.160 0.204 0.249 0.229 0.215 0.463 0.376 0.257 0.225 0.331 0.546 0.272 0.240 0.308 0.286 Volunetric Hoisture Content 0.074 0.082 0.071 0.072 0.070 0.084 0.083 0.140 0.142 0.092 0.137 0.114 0.092 0.130 0.146 0.132 0.116 0.239 0.186 0.157 0.133 0.174 0.301 0.134 0.136 0.178 0.157 Therlal Conductivity, HII‘Z C 3.51 3.52 3.47 3.50 - 3.48 3.47 3.5 3.54 3.48 3.46 3.47 3.50 3.47 3.46 3.53 3.52 3.53 3.52 3.49 3.48 3.47 3.51 3.54 ~ 3.49 3.47 3.48 p. 0 Specific Heat, Kj/kg C . 4 . . 1.40 . . 1.29 1.21 1.16 1.11 1.36 1.33 1.17 ‘ 1.23 1.10 Thernal Diffusivity, 10‘SXI‘2/s 232 largest changes in the center of the windrow occurred from day 1 to day 3. The experimental design did not include detailed sjze measurements immediately after windrow construction. If this had been included in the design, large changes might have been observed. The small changes in windrow size are not suprising due to the shortness of the experiment and the type of composting system studied. Other researchers that have noticed larger changes either monitored long term changes in windrow systems (Mears et al. 1975), or studied higher rate systems (i.e. forced aeration) (Steniford et al., 1984). The difference between the small changes that I observed in the experimental DMS windrows and those reported by Steniford et al. (1984) in co-composted forced aeration systems may indicate that high rate composting leads to changes in particle size that affect consolidation behavior. Chang and Rible's (1972) analysis of the particle size distribution and the nutrient content of fresh and composted dairy manure indicates that composting decreases the proportion of small particles in the overall particle size distribution. If longer term composting systems are to be studied, the effect of composting on particle size and windrow consolidation should be examined. 6.3: Reaction and Transport Characteristics 6.3.1: Type pf Reaction The oxygen concentration profiles presented in Figure 5.26 and 5.27 provide insight into the type of reaction that 233 is occurring. By analogy with the diffusion of a gas into a reactive spherical particle, when the rate of gas diffusion is less than the reaction rate, the gas concentration profile falls very quickly from the ambient level near the particle surface to very low levels that are maintained in most of the particle (Findlayson, 1980). This description is almost identical to the oxygen concentration profile mentioned earlier. In such cases, the Thiele modulus, or dimensionless ratio of the characteristic time for transport to the characteristic time of reaction, is much greater than 1. This occurs with a fast reaction and relatively slow transport of oxygen and heat (Findlayson, 1980). The exact value of the Thiele modulus is hard to determine because of the non-isothermal composting reaction. Difficulties that I encountered with the gas sampling and analysis system prevented accurate determination of this modulus. The active presence of fungal masses in the advancing color zone, the deeper penetration of the color zone with each windrow turning, and the relatively poor nutrient status of flushed and separated dairy manure solids suggests that substrate limitations were present and may have caused the classic microbial population progression to be short circuited to the fungal recolonization period. 6.3.2: Unreacted Core Model The overall course of the composting reaction behaved like an unreacted core model of a reactant being transported 234 into a reactive mass (Levenspiel, 1972). According to the unreacted-core model, the reaction proceeds at a narrow front which moves into the solid particle. The substrate is completely consumed as the front moves through the particle, leaving behind an unreactive mass. As applied to the composting windrow, the windrow itself would correspond to the particle. Evidence of a moving reaction front is provided by the observed movement of the temperature, oxygen, carbon dioxide and color profiles through the windrow over the course of the composting period. The rate at which each of the observed fronts moves through the windrow can also be accounted for with the unreacted core model. Initially, the reaction consumes the oxygen very rapidly. The controlling step quickly becomes the oxygen transport rate, and zones with low oxygen concentrations have lower or non-existant aerobic reactions. In those areas where aerobic reactions are occurring, the rate of heat transfer begins to affect the reaction. In areas with low heat transfer, the temperature builds up and reactions increase. This process continues until inhibitory temperatures are reached and the reaction rate decreases. With this rate decrease, less oxygen is consumed and more is available for transport to locations deeper in the windrow. Finally, as the reaction proceeds, substrate at the initial location is consumed to where it cannot support high rate reactions. The front between the "consumed" and fresh 235 dairy manure solids moves inward. The observed composting reaction was more complex than the classic unreacted core model. Additional transfer resistances due to transport inside the windrow and across the water film surrounding the DMS compost particles must be included. Another complication is that the composting reaction occurs across a diffuse front rather than a sharp interface between the spent and fresh solid zones. Finally, the presence of temperature gradients influences both reaction and transport. These complications can be dealt with, however. Levinspiel (1972) describes how the combination of resistances can be handled and what implications the different resistances have for what the controlling rate is. The diffuse front problem is considered by Wen (1968) and Ishida et al. (1971a). Wen and Wang (1970) treat the effect of non-isothermal reactions. 6.3.3: Pore Space Heat and Mass Transfer In the composting literature there is a consensus that heat and mass transfer in the pore spaces of windrow composting systems is primarily due to natural convection. Support for this comes from a variety of sources and includes anecdotal observations of steaming windrows, higher oxygen concentrations over buried tile inlets (Lambert and Davis, 1932), comparison of the maximum observed composting reaction rates to oxygen diffusion rates (Shell, 1955) and simple models of natural convection caused by temperature 236 and moisture content gradients (Haug, 1980). While there is no reason to doubt that these inferences are incorrect, there is no direct proof that natural convection as opposed to diffusion is the major heat and mass transfer mechanism. A major reason for this lack of proof is the difficulty in measuring natural convection in uncontrolled environments. In my observations of the depth of penetration of the color zones in the compost windrow, there is some evidence that two different types of transport were present. Comparison of the depth of penetration of white colored zone at the top and sides of each windrow turning at completion indicated that the side penetration was 2.4 to 3.1 times greater, depending on the length of composting. The depth of penetration of a reactive front over time is a rough indication of the ratio of the rate of transport to the rate of reaction or Thiele modulus. The observed difference between the top and side of the windrow could be due to different reaction rates or methods of heat and mass transfer. Reaction rates are affected by the substrate, oxygen availability, and temperature. After the DMS is mixed and placed into a windrow, the substrate is assumed to be the same in both locations. Oxygen concentrations and heat removal are affected by transport. Higher transport rates increase reaction rates by providing more oxygen. Higher heat transport lowers reactions initially but prolongs a reaction by delaying 237 inhibitory temperatures. The effect of higher transport rates on the Thiele modulus for oxygen is therefore somewhat ambiguous. If the rate of reaction of the substrate is considered, the effect is not ambiguous, however. Because of the higher oxygen concentrations and lower temperatures due to higher transport rates, the substrate is consumed more quickly. The reaction front moves inward at a more rapid pace where transport rates for oxygen and heat transfer are larger. This experimental observation provides strong evidence that two different rates of transport processes are present in windrow composting systems. These observations need to be better quantified before they are considered proven. The evidence presented above is also not sufficent to conclude what types of transport are present. Higher free air space in the side could lead to greater rates of diffusion in that area as compared with the top, for instance. Other potential mechanisms include diffusion or roll type convective cells in the top portion and in the windrow side, diffusion, natural convection, or even heat conduction through the slab. 6.4: Pathogen Survival The ultimate concern with studying DMS composting is to determine whether it can produce a bedding material that is sufficiently free of mastitis-causing organisms that it will not cause outbreaks of mastitis when used by dairy cattle. This concern can be addressed by considering the effect of 238 observed composting conditions on the survival of pathogenic organisms. Microbial survival and growth is affected by a number of factors including time-temperature conditions, gas concentrations, moisture content, water activity, pH and substrate availability. Time-temperature profiles are the most important factor in composting and will be the subject of the rest of this discussion. The discussion in the Literature Review indicated that there are a variety of standards for composting time- temperature curves. A common one is the maintenance of 55°C for three days. This criterion is based on refuse composting where there is a concern with micro-organisms that have higher sterilization temperatures than the mastitis-causing organisms. Temperatures from 50°C to 55°C are minimally to moderately inhibitory for E. coli. 35 to 45°C encompasses the optimum growth temperature range for E. coli. Accordingly, these three criteria were applied to the time-temperature data in Tables 5.2 and 5.3. The number of hours spent in each temperature range are presented in Table 6.2. An assessment of the impact of the time spent in each temperature range criteria as presented in Figure 6.1. Looking at Figure 6.1 it is apparent that while there are zones of temperatures lethal to coliforms there are also areas that have sublethal or even optimal temperature for growth. Generally, the toe area and the center of the windrow seem to have the largest potential for E. coli 239 Table 6.2: Length of time spent in selected temperature ranges by location and windrow. Time (hours) Windrow 1A 18 1C 2A 28 2C LOC T > 55 C 1 117 76 0 0 0 0 2 144 151 0 109 117 76 3 128 130 0 33 155 83 4 0 0 na 112 156 88 5 125 136 151 49 148 76 6 0 45 114 0 117 24 8 119 0 0 107 151 88 9 93 133 19 0 100 56 10 0 7 95 0 0 0 T > 50 C 1 159 131 0 40 12 0 2 155 158 24 127 161 99 3 146 156 56 62 158 89 4 0 54 na 126 160 96 5 142 150 158 80 162 84 6 0 94 132 0 148 47 8 143 157 0 123 159 96 9 126 157 135 8 128 72 10 0 73 116 0 38 13 35 < T < 45 1 0 0 71 21 123 73 2 0 0 95 2 2 0 3 0 0 63 41 0 4 4 137 36 0 2 2 0 5 14 0 0 38 0 8 6 74 0 0 73 2 33 8 12 0 164 2 0 0 9 21 0 17 64 4 2 10 136 41 5 31 57 31 214(3 63 L62 L L61 L61 L L 62 63 Nindrov 18 61 N L L L L L N N61 _ Hindrov 18 63 62 LN N61 62 62 LN L Nindrov 16 = T ) 55 C for 3 days. L L L 862 62 61 62 62 62 Uindrov 24 L L L L L 63 L L 627 Nindrov 28 L L L L N 62 L 62 62 Hindrov 20 61 = 0pt1:UI regrowth temperatures ) 10 hours. 62 = 30 ( Optinul regrowth teaperatures ( 100 hours. 63 = Dptinuo regrowth temperatures ) 100 hours. Figure 6.1: Summary of assessment of time- N = T ) 50 C for 3 days. temperature effects on microbial survival and growth. 241 survival and regrowth. Based on the number of locations that had temperatures greater than 50 or 55°C, both windrows seemed to be equally effective. When the length of time that a location had optimal E. coli growth temperatures is considered, Windrow 1 had only half as many survival zones as Windrow 2. Another way to evaluate the effect of a time- temperature curve on organism survival is to calculate the number of log reductions that it causes. Following the procedUre described by Haug (1980), kd as a function of temperature was calculated for "coliform bacteria" with time-temperature survival data originally developed by Ward and Brandon (1977). The equation developed was kd = 0.216 exp [0.145 * (TC—50)] (6.3) where TC is the material temperature in degrees C. Equation 6.3 was substituted into 1n (no / nt) = Rd (6.4) and converted into base 10 logs to achieve log (no / nt) = (0.216 / 2.303) * exp [0.145 * (TC - 50)] (6.5) ) Equation 6.5 can be used with time-temperature data to predict the number of decimal reductions (Dr) that are possible at a given location. The 9 locations in Windrow 18 have been analysed in this manner. Figures 6.2 through 6.4 show the Dr at each location over the one week composting period. Most locations have Dr occurring throughout the composting period. Locations 4, 6, 242 c m 2 u 3 no — 'o o m o 50 ‘ 1.4 0' o a-‘ 40- ‘U c a 1. so 1 U C O 20 - h 3 3 '0 LRZ o ' LR3 2' LRI g 0 LR3 LR]. R2 I U r U I r O 2 4 O 8 Time (Days) Figure 6.2: Windrow lB temperature and log reduction at Locations 1, 2, and 3. C 70 O "‘ T5 ‘J U '3 60 - m T6 m 50 — 3 v (7' 3 4O -« 7:. ‘U I: (U U o 1. LR6 q: 20 -4 h D u 2 1O -( 3 Q) E" O l I r 1-l I I I 0 2 4 O 8 Time (Days) Figure 6.3: Windrow lB temperature and log reduction at Locations 4, 5, and 6. 243 r: 70 O or. 3 79 '8 00 - Q . T10 I: 3 so '1 .ar-f' T8 0" .9. M '0 7° " r: to 8 “’5 O :3 1 - :1 U «I :3 IO .1 LR9 O. 5 LR8 L110 (— o — fi‘ 7 - t“ I I I I T I I O 2 4 0 I Time (Days) Figure 6.4: Windrow lB temperature and log reduction at Locations 8, 9, and 10. 244 8, and 10 show very low values. Figure 6.5 summarizes the Dr at all 9 locations for the entire composting period. Even Location 4, which has the lowest Dr, still has a total of 380 over the seven day period. 380 1180 3870 7480 3250 1890 3380 1310 510 Figure 6.5: Decimal reductions in Windrow 18 by location. Calculations of total Dr do not tell the entire story, however. The effect of micro-organism growth is not included. Problems with unheated, anaerobic zones exist. The effect of mixing on organism survival is similarly not covered. Both of these problems will now be discussed. Assuming that an nonreacting volume is represented by a sphere, the amount of time that it would take for the temperature at the center of the sphere TO to heat to 90 percent of the surrounding mass is calculated. The dimensionless time is given by the Fourier ratio (F0) F0 = (k t / 0 cp r2) (6.6) where Fo = Fourier ratio, dimensionless k = thermal conductivity, W / m-s t = time, s p = density, kg/m3 c = specific heat, Kj / kg °C radius, m From the Heisler charts (Holman, 1981), the value of the 245 Fouier number at a dimensionless temperature of 0.9 is 0.3. This is substituted into Equation 6.6 and the equation solved for the time t: t = 0.3 p c R2 / k (6.7) P Selected values of thermal and physical properties covering the range encountered in Windrow 18 are substituted into the above equation and the effect of radius is examined. Examination of Table 6.3 indicates that unreactive spheres with radii of 5 cm would require less than 2 hours to heat up to 90 % of the surrounding material. Spheres with a radius of 20 cm would take slightly more than a day for the same degree of heating. Large anaerobic compost balls are not expected in DMS solids composting, but it appears that even fairly large spheres could be heated adequately. Thermal inactivation during composting may also be limited because of non-uniform heat distribution followed by mixing of the compost. Cold pockets could allow pathogenic micro-organisms to survive or regrow. Mixing would then redistribute the organisms throughout windrow or bedding material. Haug (1980) described the effects of thermal inactivation of pathogens in a two zone windrow. One zone is assumed to have sublethal time-temperature profiles and the other is assumed to have lethal conditions. The windrows are turned at time intervals of At and thoroughly mixed so that a random redistribution of compost occurrs. 246 Table 6.3: Temperature penetration time as a function of clump diameter and thermal properties (6 = 0.9). Thermal Cond. (W/m—C): 0.301 0.174 Bulk Density (Kg/m“3): 696 428 Specific Heat (KJ/Kg-C): 3.51 3.47 Radius Penetration Time Radius Penetration Time (cm) (hours) (days) (cm) (hours) (days) 0.5 0.02 0.001 0.5 0.02 0.001 1.0 0.07 0.003 1.0 0.07 0.003 5.0 1.69 0.070 5.0 1.78 0.074 10.0 6.76 0.282 10.0 7.11 0.296 15.0 15.22 0.634 15.0 16.00 0.667 20.0 27.05 1.127 20.0 28.45 1.185 25.0 42.27 1.761 25.0 44.46 1.852 50.0 169.09 7.045 50.0 177.82 7.409 100.0 676.35 28.181 100.0 711.28 29.637 Thermal Cond. (W/m-C): 0.133 0.071 Bulk Density (Kg/m‘3): 288 164 Specific Heat (KJ/Kg-C): 3.48 3.47 Radius Penetration Time Radius Penetration Time (cm) (hours) (days) (cm) (hours) (days) 0.5 0.02 0.001 0.5 0.02 0.001 1.0 0.06 0.003 1.0 0.07 0.003 5.0 1.57 0.065 5.0 1.67 0.070 10.0 6.28 0.262 10.0 6.68 0.278 15. 14.13 0.589 15.0 15.03 0.626 20.0 25.12 1.047 20.0 26.72 1.113 25.0 39.25 1.635 25.0 41.75 1.739 50.0 156.99 6.541 50.0 166.98 6.958 100.0 627.97 26.165 100.0 667.93 27.831 247 , The resulting equation is: nt = no (£1 + fh expL"kd AtJN ) (6.8) where nt = number of organisms surviving no = number of organisms initially present fl = fraction of composting material in the low- temperature, sublethal'zone fh = fraction of composting in the high- temperature zone At = time interval between windrow turnings kd = thermal death coeffiecient N = number of windrow turnings BDd fl + fh = 1 Various assumptions about inital pathogen population levels, lethal zone fractions and time temperature thermal death factors are evaluated. Table 6.4 presents the results of such an analysis. Table 6.4: Effect of initial decimal reduction, fraction of lethal temperature and thermal death coefficent on the number of windrow turns required. Number of Turns Rd A t f1 / fh Dr -2 -3 -4 -5 —6 -7 a 0.11 2 3 4 5 6 7 a 0.25 3 5 6 8 9 10 m 0.50 7 10 14 17 20 24 m 1.00 8 12 16 20 24 27 A kd At value of infinity is assumed for this analysis because of the long one week residence times. The kd At value used in the previous time-temperature curve analysis was approximately 1.5 decimal reductions for a 15 minute 248 period. A range of fl/fh values was used to correspond to different levels of the lethal fraction. The lethal fraction was based on the assumption that each temperature measurement location represented an equal volume in the composting windrow. The 1.0 value results from the assumption that 3 days at 55 C are necessary for safe composting. The 0.25 and 0.11 values for fl/fh correspond to the assumption that temperatures over 50 C resulted in thermal inactivation in Windrows l and 2, respectively. From the results of the analysis presented in Table 6.4, it is apparent that a large number of turns are necessary to achieve thermal destruction by complete mixing of the windrow. If 5 Dr are desired, 5 turnings are required even under the most optimistic assumptions. This analysis would predict that the 3 turnings result in at best 3 Dr' A number of factors contribute to potential errors in the number of turns analysis. More accurate measurement of the lethal areas improves accuracy. Data developed in this study could provide a first approximation of lethal areas. However, it should be remembered that temperatures in the toe area and on the surface were not adequately monitored. A major confounding factor is the observation that the degree of substrate consumption could reduce the ability of a windrow to self heat on successive turnings. Finally, the effect of pathogen growth, both at optimal temperatures and on reduced substrates,is not directly included. 249 Analysis of time-temperature patterns in the compost windrows indicated that there were regions that would have significant reductions of coliform organisms due to long periods of high temperature. There were also windrow regions, however, that had episodes that were in the optimal temperature range for coliform growth. The locations of both regions changed with time, windrow size and windrow shape. The rationale for an improved heat and mass transfer model to predict windrow temperatures is justified. 6.5: Application of Com ostin to Dairy Manure Solids for Production gI—Bedding MageFTaI At this point I find it hard to advise dairy operators on how best to handle the manure solids, because I collected no microbial evidence to support the deductions based on the time-temperature curves. From a consideration of the time- temperature history alone, I would recommend that dairy operators using this type of windrow composting system should not mix the windrow immediately before using the composted dariy manure solids as bedding material. Based on the results of the time-temperature curve analysis, I would suggest that at the end of a three week composting period material from the toe area and the unreacted windrow interior should be set aside and the hot composted material should be mixed and used for bedding. The material set aside should be discarded or recycled back into the first week's composting material. In the long run, enhanced natural or forced convection 250 may prove to be the methods of choice because of the greater control offered by the forced convection, and the higher composting rate and smaller land requirements of both when compared with a windrow composting system. Microbial evidence that confirms conclusions based on time-temperature data should receive high research priority. If this evidence becomes available, research to optimize pathogen and substrate reduction in natural or forced convection systems should receive priority over further windrow studies. 6.6: Prelimipary Assessment pf Heat and Mass Transfer Models For Composting Dairy Manure Solids Without conducting a detailed analysis, the complex unreacted core model discussed earlier appears to describe the observed behavior of this type of composting system very well. The chemical engineering literature contains a large body of information on this subject. This should prove to be a fruitful avenue of exploration to understand the behavior of this type of composting system. As a first approximation of a heat and mass transfer model, a diffusion based model with spatially varied bulk density and constant moisture content should be considered. The effect of the latent heat of evaporation would be contained in a non-linear keff term. Respiration modeling should include the effects of substrate, oxygen concentration, and temperature on micro-organism growth and inactivation. An accurate unsaturated compaction model for the determination of the spatially varied bulk density, 251 volumetric moisture content and free air space variables that affect heat and mass transfer is essential. A second generation model that explicitly includes moisture generation and transport relationships would provide a theoretically more accurate picture of windrow processes. Longer term studies would require an unsaturated consolidation model to determine windrow bulk density. The effect of composting on particle size and structural stability should be included in such a model. Due to observed differences in the penetration depth of the treated dairy manure solids, a model that includes both natural convection and diffusion may be the only acceptable alternative. Due to the formidable requirements of applying this type of model to a windrow, it is suggested that such a model be attempted only after the previously mentioned diffusion based models prove to be inadequate. CHAPTER 7 CONCLUSIONS 1. The probe design and methods of determining temperature (5 a :1.5°C), gravimetric moisture content (s = t 2.0 % wb), and wet bulk density (CV = 11 %) had acceptable levels of accuracy. Gas concentration determinations were subject to errors caused by gas influx into storage containers. With the use of an empirical correction equation, gas concentrations were accurate to within :2 percent. Volatile solids, whether expressed on a dry or ash basis, had unacceptable measurement errors (sd = 7 %). 2. The assumption of constant physical and thermal properties used in previous models is not warranted in all cases. Bulk density, volumetric moisture content and the degree of water saturation increase with depth. Porosity, free air space and the void ratio decrease with depth. Spatial changes in the moisture content (% wb), volatile solids content (% db) and calculated specific heat were not observed. Since many heat and mass transfer properties are related to bulk density, volumetric moisture content and free air space, they should be considered spatially varied. With the methods of measurement used in this experiment, 252 253 there was no evidence indicating that bulk density varied with time for this type of compost substrate and windrow system- 3. The dairy manure solids behaved similarly to other composting substrates. Differences exhibited by dairy manure solids composting included the very high moisture contents (between 76 and 81 % wb) and low maximum temperature of 68°C. High moisture content composting occurred because of the high free air spaces caused by dairy manure solid particles. The low temperature could be caused by small windrow sizes, high moisture contents, or long-term substrate limitations. Experimental evidence points toward high moisture content or substrate limitations. Unlike sewage sludge and garbage composting, the expression of moisture content as percent dry basis and volatile solids on a percent ash basis is necessary to detect changes in these parameters. 4. The overall course of the composting reaction behaved like an unreacted core with a reactant being transported into a reactive mass. The observed composting reaction was more complex than a classic unreacted core model as evidenced by the lack of a sharp interface between the spent and fresh solid zones and because of a fast exothermic reaction that produced temperature gradients in the windrow. 5. An estimate of the magnitude of the Thiele modulus was made. The Thiele modulus was not constant throughout the 254 windrow. 6. Analysis of time-temperature patterns in the compost windrows indicated that there were regions that would have significant reductions of coliform organisms due to long periods of high temperature. There were also windrow regions, however, that had periods of temperature that were in the optimum range for coliform growth. The locations of both regions changed with time, windrow size, and windrow shape. 7. Dairy operators using a windrow compost system should not mix the entire windrow immediately before using the composted dairy manure solids as bedding material. A safer method of handling a three week old windrow of composted dairy manure solids is to set aside material from the windrow toe and central unreacted core for recycling or disposal. The remaining hot composted material can be used for bedding. 8. As a first approximation, a heat and mass transfer model of the dairy manure solids windrow composting process should have the following characteristics: Diffusion-based Spatially varied bulk density, predicted by an unsaturated compaction model Constant moisture content Respiration modeled with terms accounting for the effect of substrate consumption, temperature and free air space oxygen concentration. CHAPTER 8 DIRECTIONS FOR FUTURE RESEARCH 8.1: Experimental Methods 1. Bulk Density. The in situ bulk density of the composting material is a critical parameter that must be accurately measured in any futher work. Together with moisture content, it affects a number of heat and mass transport processes in the compost windrow. Two areas warrant particular study: (a) methods of in-situ measurement of bulk density, and (b), the spatial variance of bulk density as a function of the methods of mixing, placement and turning. a. In-situ Measurements. There is little discussion of insitu bulk density measurements in the composting literature. The soil and peat sampling literature provided the most fruitful sources of information that lead to the adoption of the bulk density sampling method used in this study. While the rotary corer method gave good repeatability and appeared to give good estimates of bulk density, this needs to be further substantiated. In particular, the effect of drill speed, the possible inclusion or exclusion of dms material in the sample, the effect of methods, rates 255 256 and vibration of insertion and the effect of dms particle size and shape on the effective height and radius should be examined. b. Spatial Variance pf Bulk Density. Once accurate methods for in-situ bulk density measurement have been developed, the spatial variance of bulk density as affected by compost substrate, methods of mixing, placement and turning should be investigated. Other investigators have observed that tremendous variations in bulk density beyond those expected due to compost compaction can be present in windrows. Variance estimates the variability are particularly necessary because of the difficulty in taking many bulk density samples during an experiment without affecting heat and mass transfer in the experimental windrow. 2. Moisture Content. Moisture contents expressed on a dry basis give more realistic indications of moisture content changes than wet basis moisture content. The moisture content measurement technique used in this experiment was very time consuming and did not provide frequent measurements. A quicker method could greatly reduce experimental efforts in this area. The development of real time moisture content determinations could complement temperature data measurements. This becomes more crucial in 1Windrow environments with enhanced natural convection or f0 rced aeration where more rapid moisture content changes 257 can be expected to occur. 3. Substrate. While volatile solids is a commonly used parameter in engineering studies of waste treatment systems, it is a rather crude measurement of substrate consumption. If volatile solids are to be measured in future studies, they should be expressed on an ash solid basis. Future studies should include additional substrate measurements in order to more completely reflect substrate degradation. Levels of amino acids, amino sugars, nitrate and ammonia nitrogen, hemicellulose, cellulose and lignin have been used to measure the degradation of other compost substrates. 8.2: Windrow Compost Processes 1. Respiration. The respiratory activity of micro organisms on dms needs to be better understood and quantified. It is doubtful that research on other substrates such a sewage sludge and ground garbage will be transferable beyond providing the general shape of respiration equations. If the dms respiration rate is not quantified, the difficulty in obtaining reasonably accurate estimates of respiration as well as dms heat and mass transfer parameters is greatly increased. The effect of temperature, oxygen and moisture content on respiratory activity needs to be examined. The consumption of various substrate components needs to be measured concurrently. 2. Moisture Content. If moisture content is to be included in a compostwindrow model, the role of water in the 258 composting process must be much better understood. Areas of investigation include the degree of free vs bound water on dms particles as a function of particle size and moisture content; the role of dms compressibility on the extrusion of water from particles to void spaces: and the quantification of water lost to seepage from the windrow base and evaporation from windrow surfaces. The use of tensiometers and relative humidity probes should be evaluated for this purpose. 8.3: Compost Heat and Mass Transfer Modeling 1. The development of an unsaturated model of dms windrow compaction is essential to the development of an in- situ bulk density sampler and to future modeling efforts. Since bulk density is a key parameter in predicting several heat and mass transfer properties, such a model must be available to predict bulk density. 2. Once the model described above is available, work should proceed with the development and validation of a heat and mass transfer model along the lines suggested in Chapter 7. A second generation model that explicitly includes moisture content would provide a more theoretically sound model. If longer term studies on windrow composting are to be conducted, an unsaturated consolidation model needs to be developed. The effect of the composting process on particle size distributions and dms particle compressibility needs to be investigated and included in the model. 259 3. Due to the observed differences in the penetration depth of the treated dms, a model that includes both diffusion and natural convection may be the only acceptable alternative. Because of the formidable requirements of applying this type of model to a windrow, it is suggested that such a model be attempted only after the previously mentioned diffusion based models prove to be inadequate. Model development should procede from simpler l and 2 dimensional cases before expanding to the complex 3 dimensional situation in dms windrows. APPEND I CES ~n-r‘,‘ 1%.- -._r1 ...—7' :1 APPENDIX A TEMPERATURE MEASUREMENT ERROR CALCULATIONS APPENDIX A TEMPERATURE MEASUREMENT ERROR CALCULATIONS The total thermocouple temperature measurement error is the sum of the errors in the reference junction temperature, the thermocouple output, the thermocouple voltage measurement, and the linearization error (difference between standard and polynomial approximations). A Fenwal Electronics UUTSlJl thermistor mounted in the center of the analog input terminal strip measures the 21X panel temperature. According to the Campbell literature, the "worst case" example adds a : 0.3°C error in the range of -35°C to +50°C. In CSI's experience, the overall accuracy is typically better than t 0.2°C. The thermistor error increases drastically if temperatures greater than 50°C or less than -35°C are imposed on the CR21X. Differences between the thermistor and actual reference junction temperature will become errors in measurement. The terminal cover provides a shield to reduce the gradient along the terminal strip. The CR21X manual (Campbell Scientific, 1985) gives an example in which the 21x was brought from -25°C to an ambient temperature of 20°C. After 100 minutes the temperature gradient between the air and the battery is still 15°C, but the deviation from the 260 261 measured temperature along the reference strip is only 0.3°C. This taken as the maximum error that could occur in a field situation with proper radiation shielding. The ANSI standards for thermocouple limits of error are given in the National Bureau of Standards Monograph 125 (1974). For copper constantan thermocouples, they are as follows: Table A.1: Limits of Error for Thermocouple Wire (Reference Junction at 0°C) Limits of Error Thermocouple Temperature (Whichever is greater) Type Range °C Standard Special T -200 to 0 1 l.0°C or 1.5 % 0 to 350 t l.0°C or 0.75 % t 0.5°C or 0.4 % In order to quantitatively evaluate thermocouple error when the reference junction is not fixed at 0°C, the limits of error for the Seebeck coefficient (slope of thermocouple voltage vs temperature curve) for the various thermocouples must be available. According the CSI, if this information is missing, a reasonable approach is to apply the percentage errors, with perhaps 0.25 % added on, to the difference in temperature being measured by the thermocouple. The temperatures in this experiment will range from 15°C to 70°C. The error due to the thermocouple would be l.0°C. The accuracy of a 21X voltage measurement is specified as 0.1 % of the full scale range being used to make the 262 measurement. The error in the temperature due to inaccuracy in the measurement of the thermocouple voltage is worst at temperature extremes, where a relatively large scale is necessary to read the thermocouple output. .In the environmental temperature range with voltage measured on an appropriate scale, error in temperature due to the voltage measurements is a few hundredths of a degree (Anonymous, 1985). Voltage to temperature conversions are accomplished using a proprietary 6th order polynomial. The limit of error on 21x thermocouple output linearization, relative to NBS Standards, is 10.001 over a range of -100°C to 100°C. When external reference junction boxes are used errors can arise if the reference junction temperature is outside of the linearization range. The reference temperature compensation range and linearization error relative to the NBS Standards for type T thermocouples are -100°C to 100°C and t0.001°C. The sources of error discussed above are summarized in Table A.2. 263 Table A.2: Summary of the sources of error in thermocouple measurements. Source Error (°C) Percent of Total Error Reference junction temperature 0.5 32.2 Thermocouple output 1.0 64.4 Voltage measurement 0.05 3.2 Reference linearization 0.001 0.1 Output linearization 0.001 9:1 Total Error 1.552 100.0 APPENDIX B SAMPLE THERMOCOUPLE WATER BATH CALIBRATION DATA APPENDIX B SAMPLE THERMOCOUPLE WATER BATH CALIBRATION DATA Table B.l Water bath thermocouple calibration-~Run 1. Thermocouple Hour:Mi 1 2 3 4 5 6 7 1418 -6999 27.86 27.84 27.86 27.86 27.84 27.86 1419 -6999 27.86 27.86 27.87 27.85 27.87 27.86 1420 -6999 27.84 27.82 27.85 27.82 27.84 27.83 1421 -6999 27.82 27.80 27.81 27.80 27.80 27.80 1422 -6999 27.75 27.74 27.75 27.73 27.76 27.75 1423 —6999 27.69 27.68 27.70 27.70 27.70 27.69 1424 -6999 27.64 27.64 27.67 27.65 27.68 27.66 1425 -6999 27.60 27.61 27.62 27.61 27.63 27.62 1426 -6999 27.56 27.57 27.58 27.57 27.59 27.59 1427 -6999 27.53 27.53 27.53 27.52 27.55 27.54 1428 -6999 27.50 27.49 27.50 27.49 27.51 27.51 1429 -6999 27.46 27.46 27.47 27.47 27.47 27.48 1430 -6999 27.41 27.42 27.42 27.42 27.43 27.42 1431 -6999 27.38 27.37 27.38 27.38 27.39 27.38 1432 27.35 27.34 27.33 27.35 27.33 27.36 27.34 1433 -6999 27.30 27.30 27.31 27.30 27.32 27.31 1434 27.26 27.26 27.24 27.27 27.26 27.25 27.26 1435 27.22 27.22 27.20 27.23 27.21 27.22 27.23 1436 27.19 27.17 27.17 27.19 27.18 27.19 27.19 1437 27.12 27.12 27.11 27.12 27.12 27.13 27.13 1438 27.08 27.08 27.08 27.10 27.08 27.09 27.09 1439 27.03 27.04 27.03 27.05 27.03 27.05 27.06 1440 27.03 27.04 27.03 27.06 27.03 27.04 27.06 1441 27.40 27.46 27.42 27.44 27.39 27.41 27.45 1442 27.69 27.71 27.70 27.67 27.71 27.73 27.68 1443 27.66 27.66 27.66 27.67 27.66 27.68 27.68 1444 27.62 27.62 27.61 27.63 27.62 27.64 27.63 1445 27.57 27.58 27.57 27.58 27.57 27.59 27.58 1446 27.52 27.53 27.51 27.53 27.53 27.54 27.54 1447 27.49 27.49 27.49 27.50 27.48 27.50 27.51 Ave. 27.35 27.48 27.48 27.49 27.48 27.49 27.49 Var. 0.05 0.06 0.06 0.06 0.06 0.06 0.06 78.1 265 Table 8.1 (Cont.). Thermocouple 8 9 10 ll 12 13 14 15 27.84 27.89 27.91 27.91 27.94 27.96 27.94 27.94 27.85 27.89 27.92. 27.92 27.94 27.95 27.95 27.94 27.84 27.87 27.88 27.89 27.92 27.92 27.92 27.91 27.80 27.84 27.85 27.87 27.88 27.89 27.89 27.88 27.75 27.78 27.80 27.81 27.84 27.84 27.85 27.84 27.68 27.73 27.76 27.77 27.79 27.80 27.80 27.80 27.66 27.70 27.72 27.73 27.77 27.77 27.78 27.78 27.62 27.66 27.69 27.70 27.73 27.73 27.74 27.75 27.58 27.62 27.65 27.66 27.69 27.69 27.72 27.72 27.53 27.58 27.61 27.62 27.64 27.66 27.66 27.67 27.50 27.55 27.57 27.58 27.61 27.62 27.63 27.64 27.47 27.52 27.54 27.55 27.57 27.59 27.61 27.60 27.41 27.46 27.48 27.50 27.52 27.54 27.54 27.55 27.38 27.42 27.45 27.47 27.48 27.49 27.51 27.51 27.33 27.38 27.37 27.38 27.44 27.45 27.45 27.45 27.30 27.35 27.37 27.35 27.40 27.41 27.43 27.40 27.26 27.30 27.32 27.33 27.36 27.38 27.38. 27.39 27.21 27.26 27.29 27.30 27.32 27.35 27.34 27.35 27.19 27.22 27.25- 27.26 27.28 27.29 27.31 27.31 27.12 27.17 27.18 27.19 27.23 '27.24 27.27 27.26 27.09 27.12 27.15 27.16 27.18 27.21 27.22 27.22 27.05 27.09 27.11 27.12 27.15 27.16 27.18 _27.19 27.05 27.09 27.08 27.08 27.13 27.13 27.15 27.17 27.45 27.45 27.40 27.37 27.41 27.41 27.45 27.45 27.69 27.74 27.71 27.73 27.76 27.77 27.81 27.80 27.66 27.71 27.73 27.74 27.78 27.78 27.81 27.80 27.63 27.67 27.68 27.69 27.73 27.74 27.76 27.76 27.57 27.61 27.64 27.65 27.68 27.69 27.71 27.71 27.54 27.58 27.60 27.61 27.64 27.65 27.68 27.68 27.49 27.53 27.55 27.57 27.60 27.60 27.63 27.64 27.48 27.53 27.54 27.55 27.58 27.59 27.60 27.60 0.06 0.06 0.06 0.06 0.06 0.06 0.05 0.05 266 Table 8.1 (Cont.). Thermocouple 16 17 18 19 20 21 22 23 27.91 27.91 27.99 27.94 27.94 27.94 27.91 27.91 27.95 27.92 28.00 27.95 27.93 27.92 27.94 27.92 27.91 27.91 27.97 27.93 27.91 27.90 27.91 27.91 27.91 27.88 27.96 27.91 27.89 27.87 27.89 27.89 27.85 27.84 27.91 27.86 27.84 27.83 27.84 27.83 27.82 27.80 27.87 27.82 27.80 27.80 27.80 27.79 27.79 27.78 27.85 27.80 27.77 27.76 27.77 27.76 27.75“ 27.75 27.81 27.75 27.75 27.72 27.74 27.72 27.72 27.72 27.79 27.72 27.70 27.69 27.70 27.69 27.69 27.67 27.74 27.69 27.66 27.65 27.65 27.63 27.65 27.63 27.71 27.64 27.64 27.62 27.62 27.60 27.62 27.60 27.66 27.63 27.59 27.58 27.59 27.57 27.56 27.56 27.63 27.56 27.56 27.53 27.53 27.52 27.53 27.51 27.59 27.53 27.51 27.49 27.50 27.47 27.47 27.47 27.53 27.50 27.47 27.46 27.46 27.44 27.45 27.42 27.49 27.45 27.45 27.41 27.43 27.41 27.40 27.39 27.46 27.41 27.39 27.38 27.38 27.36 27.37 27.35 27.43 27.38 27.36 27.34 27.34 27.32 27.32 27.31 27.38 27.35 27.33 27.31 27.32 27.28 27.28 27.27 27.35 27.31 27.29 27.25 27.26 27.24 27.26 27.23 27.31 27.27 27.24 27.22 27.22 27.21 27.21 27.20 27.27 27.23 27.21 27.18 27.18 27.17 27.21 27.17 27.27 27.22 27.19 27.18 27.17 27.15 27.56 27.46 27.57 27.64 27.59 27.56 27.57 27.54 27.84 27.81 27.90 27.90 27.87 27.86 27.85 27.84 27.84 27.82 27.90 27.85 27.83 27.81 27.81 27.79 27.79 27.78 27.86 27.81 27.78 27.76 27.76 27.74 27.75 27.74 27.81 27.76 27.73 27.71 27.72 27.70 27.70 27.70 27.78 27.70 27.68 27.68 27.66 27.65 27.67 27.65 27.73 27.67 27.64 27.63 '27.63 27.60 27.63 27.61 27.68 27.64 27.62 27.60 27.61 27.59 0.05 0.05 0.05 0.05 0.05 0.05 0.05 0.05 267 Table 8.1 (Cont.). Thermocouple 24 25 26 27 28 29 30 31 27.94 27.91 27.94 27.91 27.91 27.84 31.20 31.37 27.95 27.93 27.92 27.89 27.88 27.84 31.17 31.39 27.91 27.90 27.89 27.88 27.85 27.82 31.17 31.34 27.89 27.87 27.88 27.84 27.82 27.79 31.25 31.44 27.84 27.82 27.82 27.78 27.77 27.73 31.28 31.53 27.78 27.78 27.77 27.72 27.71 27.67 31.26 31.47 27.75 27.74 27.75 27.69 27.68 27.63 31.24 31.47 27.71 27.70 27.71 27.65 27.64 27.60 31.31 31.50 27.68 27.66 27.66 27.60 27.58 27.54 31.25 31.45 27.64 27.62 27.62 27.56 27.55 27.50 31.24 31.40 27.60 27.58 27.58 27.54 27.51 27.47 31.21 31.41 27.57 27.55 27.56 27.50 27.48 27.45 31.25 31.45 27.52 27.51 27.51 27.45 27.44 27.39 31.32 31.50 27.47 27.45 27.47 27.42 27.39 27.35 31.25 31.51 27.45 27.43 27.44 27.37 27.35 27.31 31.35 31.55 27.41 27.39 27.39 27.34 27.31 27.28 31.30 31.58 27.35 27.34 27.34 27.28 27.28 27.22 31.46 31.66 27.33 27.30 27.30 27.25 27.24 27.19 31.44 31.69 27.28 27.28 27.27 27.20 27.19 27.14 31.51 31.64 27.24 27.21 27.21 27.16 27.13 27.09 31.46 31.55 27.20 27.17 27.17 27.12 27.10 27.06 31.49 31.55 27.15 27.14 27.14 27.07 27.06 27.02 31.46 31.49 27.15 27.13 27.12 27.10 27.07 27.01 31.48 31.51 27.54 27.53 27.53 27.50 27.47 27.44 31.48 31.54 27.83 27.82 27.80 27.76 27.74 27.70 31.51 31.58 27.78 27.77 27.76 27.70 27.69 27.64 31.51 31.58 27.74 27.71 27.71 27.65 27.63 27.60 31.51 31.56 27.68 27.67 27.67 27.61 27.59 27.56 31.56 31.61 27.64 27.61 27.62 27.57 27.54 27.50 31.59 31.64 27.60 27.58 27.57 27.52 27.50 27.47 31.51 31.57 27.59 27.57 27.57 27.52 27.50 27.46 31.37 31.52 0.06 0.06 0.06 0.06 0.06 0.06 0.02 0.01 268 Table 8.1 (Cont.). 32 Ave. Var. 8.0. 31.59 27.91 0.002 0.040 31.61 27.91‘ 0.002 0.040 31.61 27.88 0.002 0.039 31.63 27.86 0.002 0.043 31.63 27.81 0.002 0.046 31.63 27.76 0.003 0.053 31.64 27.73 0.003 0.058 31.66 27.69 0.003 0.058 31.66 27.65 0.004 0.064 31.66 27.61 0.004 0.063 31.68 27.57 0.004 0.062 31.70 27.54 0.004 0.061 31.70 27.49 0.004 0.061 31.71 27.45 0.004 0.061 31.72 27.41 0.004 0.060 31.73 27.37 0.003 0.058 31.74 27.33 0.004 0.062 31.75 27.29 0.004 0.064 31.76 27.26 0.004 0.064 31.76 27.20 0.005 0.070 31.77 27.17 0.005 0.068 31.78 27.13 0.005 0.070 31.78 27.12 0.004 0.065 31.79 27.48 0.005 0.069 31.80 27.78 0.005 0.068 31.80 27.75 0.005 0.070 31.81 27.70 0.005 0.070 31.82 27.66 0.005 0.069 31.83 27.61 0.005 0.072 31.83 27.57 0.005 0.070 31.72 27.56 .00 0.06 0.01 0.06 .00 .00 APPENDIX C EFFECT OF SEQUENTIAL GAS SAMPLES AND LOCATION ON CARBON DIOXIDE AND OXYGEN CONCENTRATION APPENDIX C EFFECT OF SEQUENTIAL GAS SAMPLES AND LOCATION ON CARBON DIOXIDE AND OXYGEN CONCENTRATION Table C.1: Effect of sequential gas samples and location on C02 and 02 concentrations. Sequential Volume Location 1 Location 5 Location 8 Location 10 Withdrawn C02 02 C02 02 C02 02 CO2 02 5 0.63 19.54 0.80 18.54 0.27 16.64 15 7.85 12.64 9.02 10.17 3.40 16.06 16.38 5.82 25 7.80 13.01 15.95 4.65 4.40 16.40 19.08 3.96 35 7.82 13.11 16.87 3.81 4.00 15.73 45 7.61 12.75 15.73 5.63 4.22 16.33 18.63 4.99 55 7.37 12.95 15.44 2.96 4.03 16.30 19.32 4.06 65 7.78 13.49 15.66 4.36 20.05 3.49 75 7.94 13.23 16.04 4.61 2.62 16.87 19.12 3.87 AVG * 7.74 13.03 15.95 4.34 4.16 16.19 19.24 4.07 SD 0.18 0.27 0.46 0.82 0.16 0.27 0.46 0.50 CV 2.29 2.05 2.86 18.87 3.86 1.65 2.41 12.18 *Statistics are calculated from the linear portion of the data. APPENDIX D GAS CONCENTRATION CORRECTION FACTOR CALCULATIONS APPENDIX D GAS CONCENTRATION CORRECTION FACTOR CALCULATIONS The procedures used to determine the gas concentration correction factor due to withdrawal of samples from the "venoject" sample containers with a syringe prior to injection into the gas chromatograph were described in general in Chapter 4. This appendix describes in more detail the procedures and calculations used to arrive at these results. The experiment consisted of evacuating six 3 cc venojects and filling them to an initial pressure of 1.5 atmospheres. This duplicated the inital conditions of sample storage in the main experiment. 1.05 cc withdrawals were made from each venoject and injected into the gas chromatograph to check for the oxygen concentration. The lengths were recorded and oxygen concentrations were calculated based on a four point calibration curve. The presence of any oxygen was assumed to be due to oxygen present in the syringe tip before sample withdrawal, diffusion of oxygen into the syringe needle after withdrawal or the flow of ambient air into the syringe due to a pressure difference. The initial lengths of the gas chromatograph traces and the calculated oxygen 270 271 concentrations are shown in Table D.l and Table D.2. The calibration curve is given at the bottom of Table D.2. The internal pressure of the sample container and syringe were calculated for several pressures, sample container volumes and syringe volumes. The equation used in this calculation was Pf = (Pi * vC)/(vC + vs) (0.1) where Px = internal pressure of x, atmospheres V volume of y, cubic centimeters final initial sample container sample syringe MOI-"H: *4 For example, if the initial conditions were 1.5 atmospheres, 3 cc and 1.05 cc for the initial pressure, container volume and sample volume, respectively, the final internal pressure is Pf = 1.50 * 3.00 /(3.00 + 1.05) = 1.11 atmospheres The results of these calculations are shown in Table 0.3. Plotting the calculated internal pressure for each sample withdrawn against the measured oxygen concentration indicated two linear segments. These were shown in Figure 4.9 of the text. Influx was calculated by assuming that the initial and atmospheric oxygen concentrations were 0.0 and 20.9 percent, respectively. Equation D.2 was used: om = In * oe + F * 0a (0.2) where 0x = oxygen concentration at x, percent Table 0.1: 272 Gas chromatograph trace withdrawal number. length as a function of With. Replication Number 1 2 3 4 5 6 AVG STD CV 1 1.90 1.30 1.55 1.40 2.30 1.75 1.70 0.34 19.73 2 2.25 2.30 2.95 2.70 3.70 2.85 2.79 0.48 17.29 3 5.85 6.50 6.50 6.45 7.15 6.85 6.55 0.40 6.09 4 8.40 9.00 9.10 9.10 8.70 9.90 9.03 0.46 5.10 5 11.10 11.80 11.45 11.45 0.29 2.50 6 12.40 13.00 12.70 12.70 0.24 1.93 Table .2: Oxygen concentrations as a function of withdrawal number. * With. Replication Number 1 2 3 4 5 6 AVG STD CV 0 0 - 0 0 0 0 0 0.00 1 1.56 0.99 1.22 1.08 1.97 1.42 1.37 0.33 23.89 2 1.92 1.97 2.66 2.39 3.49 2.55 2.50 0.52 20.97 3 6.06 6.88 6.88 6.82 7.72 7.33 6.95 0.51 7.32 4 9.37 10.19 10.32 10.32 9.78 11.43 10.24 0.63 6.16 5 13.11 14.12 13.61 13.62 0.41 3.01 6 14.99 15.86 15.42 15.43 0.36 2.32 * Calibration equation: 02 = 0.7217 * L ‘ 1.2048 273 Table 0.3: Effect of initial pressure, sample container volume, and sample size on sample container and syringe internal pressure after sampling. Init. Press.: 1.500 1.375 1.250 1.125 1.000 Stor. Vol.: 4.00 4.00 4.00 4.00 4.00 Sample Vol.: 0.55 0.55 0.55 0.55 0.55 SAMPLE PRESSURE AFTER SAMPLE NUMBER (Atmospheres) 0 1.500 1.375 1.250 1.125 1.000 1 1.319 1.209 1.099 0.989 0.879 2 1.159 1.063 0.966 0.869 0.773 3 1.019 0.934 0.849 0.764 0.679 4 0.896 0.821 0.747 0.672 0.597 5 0.788 0.722 0.656 0.591 0.525 6 0.692 0.635 0.577 0.519 0.462 7 0.609 0.558 0.507 0.457 0.406 8 0.535 0.491 0.446 0.401 0.357 Init. Press.: 1.500 1.375 1.250 1.125 1.000 Stor. Vol.: 4.00 4.00 4.00 4.00 4.00 Sample Vol.: 1.05 1.05 1.05 1.05 1.05 SAMPLE PRESSURE AFTER SAMPLE NUMBER (Atmospheres) 0 1.500 1.375 1.250 1.125 1.000 1 1.188 1.089 0.990 0.891 0.792 2 0.941 0.863 0.784 0.706 0.627 3 0.745 0.683 0.621 0.559 0.497 4 0.590 0.541 0.492 0.443 0.394 5 0.468 0.429 0.390 0.351 0.312 6 0.370 0.340 0.309 0.278 0.247 7 0.293 0.269 0.245 0.220 0.196 8 0.232 0.213 0.194 0.174 0.155 274 Table D.3: (cont'd.). Init. Press.: 1.500 1.375 1.250 1.125 1.000 Stor. Vol.: 3.00 3.00 3.00 3.00 3.00 Sample Vol.: 0.55 0.55 0.55 0.55 0.55 SAMPLE PRESSURE AFTER SAMPLE NUMBER (Atmospheres) 0 1.500 1.375 1.250. 1.125 1.000 1 1.268 1.162 1.056 0.951 0.845 2 1.071 0.982 0.893 0.803 0.714 3 0.905 0.830 0.754 0.679 0.604 4 0.765 0.701 0.638 0.574 0.510 5 0.646 0.593 0.539 0.485 0.431 6 0.546 0.501 0.455 0.410 0.364 7 0.462 0.423 0.385 0.346 0.308 8 0.390 0.358 0.325 0.293 0.260 Init. Press.: 1.500 1.375 1.250 1.125 1.000 Stor. Vol.: 3.00 3.00 3.00 3.00 3.00 Sample Vol.: 1.05 1.05 1.05 1.05 1.05 SAMPLE PRESSURE AFTER SAMPLE NUMBER (Atmospheres) 0 1.500 1.375 1.250 1.125 1.000 1 1.111 1.019 0.926 0.833 0.741 2 0.823 0.754 0.686 0.617 0.549 3 0.610 0.559 0.508 0.457 0.406 4 0.452 0.414 0.376 0.339 0.301 5 0.335 0.307 0.279 0.251 0.223 6 0.248 0.227 0.206 0.186 0.165 7 0.184 0.168 0.153 0.138 0.122 8 0.136 0.125 0.113 0.102 0.091 275 m = measured e = environmental a = actual In = influx fraction, decimal F = original fraction, decimal = l - In Substituting known values and rearranging, we get In = O / 20.948 (D.3) m Figure 4.9 of the text shows influx for the 3cc sample container as a function of internal pressure. Table D.4 includes the calculated average 3cc influx values (In3) for this experiment. Low error intermediate calculations. Table 0.4: Sample Vol P4 02M In3 In4 Number Removed 0 0.00 1.500 0.0 0.000 0.000 1 1.05 1.111 1.3 0.066 0.049 2 2.10 0.823 2.5 0.119 0.089 3 3.15 0.610 7.0 0.332 0.249 4 4.20 0.452 10.2 0.489 0.366 5 5.25 0.335 13.6 0.650 0.487 6 6.30 ‘0.248 15.4 0.736 0.552 The 3 cc results were converted to those for a 4 cc container by (a) converting the 3 cc fractional influx to a volume influx and (b) calculating the fraction of 4 cc that was represented by (a). For samples of 1.05 cc, the 3 cc fractional value was 0.066, the volume was 3 * 0.066 = 0.198, and the 4 cc fractional volume was 0.198 / 4 = 276 0.0495. These results are also shown in Table D.4. The calculated internal pressures that corresponded to the volume withdrawn from the 3 cc sample container were used with the 4 cc influx values in a linear regression. In Table D.4, the pairs of numbers corresponding to the sample numbers 0, 1 and 2 in columns 3 and 6 were used for Y and X, respectively. The calculated pressures from the 3 cc container were used for the same sample number from a 4 cc container because it was the pressure conditions in the 3 cc container that produced the calculated pressure. The equation developed (r = 0.9996) was In4P = -7.619 * Pf + 1.495 (0.4) Next, a linear regression was run with the withdrawal number and the assumed internal pressure of a 4 cc container (P4) as x and Y respectively. The resulting equation (r = 0.9971) was P4 = -0.1508 * S + 1.4802 (D.5) Using Equation 0.5, P4 values for the main experiment sample numbers were caclutaed. The results of this calculation are shown in Table D.5. The influx into a 4 cc container was calculated using the output of the previous step and Equation D.4. A linear regression was performed on the In4 as X and S as Y. The equation developed (r = 1.000) was In4 = 0.01977 * S + 0.002 (D.6) This equation was used to calculate influx values as a function of the sample withdrawal number. 277 Table 0.5: Final gas correction calculations. Sample Low Error High Error Number P4 In4 P4 In4 0 1.480 0.0020 0.9868 0.0000 1 1.329 0.0218 0.8862 0.0914 2 1.179 0.0415 0.7856 0.1726 3 1.028 0.0613 0.6850 0.2588 4 0.877 0.0811 0.5844 0.3417 In order to develop a worst case analysis, it was assumed that the worst situation would occur if the sample gas leaked out of the container to the point where the internal pressure was at one atmosphere. In this situation, the removal of gas with the syringe would immediately produce a relative vacuum in the syringe and influx of ambient air would occur. It was further assumed that the worst case was represented by the second linear section of the pressure vs. measured oxygen curve (Figure 4.9). The calculation procedure was similar to that of the previous error calculation. The fractional influx was determined and the 3 cc influxes were converted to 4 cc influxes. As in the small error case, the calculated P3 was assumed to be the cause of the gas transport and a linear regression was run on In4 and P4 (= P3). The data used corresponded to the rows with sample numbers 3 to 6 in Table 0.9976) was D.4. The resulting equation (r = 94 = -1.160 * In4 + 0.8934 (0.7) 278 Since all gas under pressure was assumed to have escaped the sample container in this worst case scenario, the initial pressure was assumed to be 1.0 atmospheres. The linear regression of sample numbers and internal pressure used the values corresponding to the calculated pressures in a 4 cc sample container initially at 1.0 atmosphere with sequential 0.55 cc syringe samples being withdrawn. These values are shown in Table D.3. The equation that resulted from this regression was In4 = -l.160 * S + 0.9868 (r = -0.9971) (D.8) New values of the 4 cc internal pressure were calculated using this equation. The results are shown in Table D.5. The slope of Equation D.8 was assumed to be accurate and the "a" term was adjusted so that at S = 0, the influx would be zero. The resulting equation was In4 = (P4 - 0.9868) / -l.166 (D.9) New values of the influx were calculated using the above equation and the results are shown in Table 0.5. A linear regression was run on the sample number and the In4 influxes and the following equation (r = 1.000) was developed: In4 = 0.0851 * S + 0.0027 (D.10) This equation was used to assess the high error case. For a given sample volume the effect is linear and can be described as follows for each gas: 0 = O a - In * Oe / F (D.ll) m Ca = c:m / F (v.12) 279 where all terms are as defined before. Data on the sample number of each gas measurement were available. With this, the measured concentrations, Equations 0.6 and D.10 for the low and high influx calculations, and Equations 0.11 and D.12 for the actual oxygen and carbon dioxide, the actual gas concentrations could be calculated. In order to provide sample calculations a number of combinations of gas concentration and sample withdrawal numbers are shown in Table D.6. Three combinations of carbon dioxide and oxygen concentrations with high, medium and low levels of each gas are shown in this table. Withdrawal numbers ranging from 1 to 3 are included. Most sample withdrawal numbers in the study were either 1 or 2; only a few were 3. As can be seen from the calculated concentration columns in Table D.6, the correction equations do adjust the concentrations in the correct direction and larger withdrawal numbers do lead to greater influxes and larger adjustments to the concentrations. One drawback is that the sum of the calculated concentrations is affected by the starting concentration and the number of withdrawals, with lower C02 and higher withdrawal numbers lowering the sum of the two calculated concentrations. Two additional estimates of gas concentration corrections were examined. The third estimate was to take the "average" of the low and high influx error estimates. This was intended to approximate the uncertainty that the 280 Table D.6: Gas concentration calculations. Withdrawal Influx Influx Caluclated Sum C02 02 Number Eqn C02 02 17 4 1 D.6 0.022 17.4 3.5 20.9 12 9 1 D.6 0.022 12.3 8.5 20.8 2 19 1 D.6 0.022 2.0 18.5 20.6 17 4 2 D.6 0.042 17.7 3.1 20.8 12 9 2 D.6 0.042 12.5 8.1 20.6 2 19 2 D.6 0.042 2.1 18.1 20.2 17 4 3 D.6 0.061 18.1 2.6 20.7 12 9 3 D.6 0.061 12.8 7.6 20.4 2 19 3 D.6 0.061 2.1 17.6 19.8 17 4 1 0.10 0.088 18.6 2.0 20.6 12 9 1 0.10 0.088 13.2 7.0 20.1 2 19 1 D.10 0.088 2.2 17.0 19.2 17 4 2 0.10 0.091 18.7 1.9 20.6 12 9 2 D.10 0.091 13.2 6.9 20.1 2 19 2 0.10 0.091 2.2 16.9 19.1 17 4 3 D.10 0.093 18.7 1.8 20.6 12 9 3 D.10 0.093 13.2 6.8 20.1 2 19 3 0.10 0.093 2.2 16.8 19.1 17 4 1 Average 18.0 2.8 20.8 12 9 1 Average 12.7 7.8 20.5 2 19 1 Average 2.1 17.8 19.9 17 4 2 Average 18.2 2.5 20.7 12 9 2 Average 12.9 7.5 20.4 2 19 2 Average 2.1 17.5 19.6 17 4 3 Average 18.4 2.2 20.7 12 9 3 Average 13.0 7.2 20.2 2 19 3 Average 2.2 17.2 19.4 17 4 1 C02 Correct 17.0 3.9 20.9 12 9 1 C02 Correct 12.0 8.9 20.9 2 19 1 C02 Correct 2.0 18.9 20.9 17 4 2 C02 Correct 17.0 3.9 20.9 12 9 2 C02 Correct 12.0 8.9 20.9 2 19 2 C02 Correct 2.0 18.9 20.9 17 4 3 C02 Correct 17.0 3.9 20.9 12 9 3 C02 Correct 12.0 8.9 20.9 2 l9 3 C02 Correct 2.0 18.9 20.9 281 exact initial pressure in the sample container was unknown. The fourth estimate was based on the assumption that the measured carbon dioxide value was correct and that the oxygen concentration was equal to 20.95 % minus the measured carbon dioxide concentration. Increases in carbon dioxide concentration at high measured concentrations were 1.6 % for the high influx correction, 1.0 to 1.4 % for average influx, and 0.4 to 1.0 % for the low influx correction. Increases at low measured carbon dioxide concentrations were less than 0.2 percent. Oxygen concentration corrections achieved by the high, average, and low influx calculations were all less than the measured values. The largest corrections occurred at the lower measured concentrations. The high influx correction resulted in an approximately 2 % calculated oxygen decrease from measured concentrations of 4 percent. The corresponding decreases for the average and low influx calculations were 1.2 to 1.8 % and 0.5 to 1.4 % respectively. Corrected oxygen concentrations obtained from the fourth method tended to be higher than the measured concentrations by as much as 1 % oxygen. The low influx correction is based on the assumption that all gas samples were placed in the cylinders at the same pressure and no gas leaked out. The high influx assumes that all gas at pressures over 1 atmosphere leaked out. As indicated in Chapter 4 of the text, the assumptions underlying both these methods are open to question. The 282 initial pressure under which gas was inserted into the sample containers was unknown. Preliminary experiments also indicated that there was some leakage out of the storage system but that uncovered samples did not decrease to 1 atmosphere pressure in the two week period before they were analyzed. All that can be said, therefore, about the internal sample container pressure at the start of analysis is that it was between 1.0 and 1.5 atmospheres. The fourth correction method was abandoned because it did not result in the desired corrections. The average influx calculation was chosen for use because of the uncertainty over the internal sample pressure at the start of analysis. This method is not without its problems, however. In realistic conditions, the actual influx would be governed by the low influx correction until the internal pressure fell below 1.0 atmospheres, at which time the high influx equation would apply. Since the initial pressure is unknown and the number of samples that could be withdrawn before the high influx conditions would prevail is therefore uncertain, the average influx calculation is the best compromise that can be achieved. APPENDIX E VOLUMETRIC GAS STANDARDS ERROR CALCULATIONS APPENDIX E VOLUMETRIC GAS STANDARDS ERROR CALCULATIONS Two of the four standards used in calibrating the gas chromatograph were made up volumetrically. The following discussion describes the procedure that was used and the error calculations that were made. The standards were made using 160 cc sample bottles with rubber caps. The capped bottles were evacuated using a vacuum pump and then filled with pure nitrogen gas. This procedure was repeated 3 times with the sample bottles being left in the vacuum state. Oxygen, carbon dioxide and nitrogen were then added by gas syringes of different sizes to make up the desired concentrations at 2 atmospheres of pressure. Two syringe sizes were used: 30 cc and 60 cc. The two syringes had variances of 0.5 and 1 cc respectively. Table E.1 summarizes the concentrations and numbers of syringe fillings that were needed for each concentration. The calculation for gas concentration in percent has the form CC = 100 * V1 / Vt (3.1) where GC gas concentration, percent Volume, cc ith gas 283 284 Table E.1: Concentration and syringe use data. Standard Gas Concentration Number of Syringe Uses Number C02 02 C02 02 N2 (percent) 30 60 30 60 30 60 3 10 10 2 0 2 0 l 4 4 20 l l 1 l 0 l 4 t = Total Applying the product rule for the propagation of errors (See Appendix G), we have 59 = [ (aGC2/ avi’) * Si: + (aGCZ/ avt2 ) * st: 1°-5 (3.2) Taking the paritial derivatives of Equation E.l with respect to Vi and Vt we get values of (lOO/Vt) and (lOOVi/TZ) respectively. Values of the variance for Vi and Vt are arrived at by multiplying the individual variance of each syringe by the number of times that syringe is used. The variance of the 10 % C02 standard can be calculated as follows: Vi = 2 * 0.5 cc = 1.0 cc Vt = [2 * 0.5 cc] + [2 * 0.5 cc] + [4 * 1.0 cc + 1 * 0.5 cc] = 6.5 cc SCOZZ = [(100/320)2 * (1.0) + (100*32/(320z ))2 * (6.5) = 0.104 % The variance for the oxygen concentration of standard 3 is 285 the same as that for carbon dioxide. Variances of standard 4 are 0.1719 % and 0.049 % for carbon dioxide and oxygen, respectively. APPENDIX F PRELIMINARY BULK SAMPLER EXPERIMENTS APPENDIX F PRELIMINARY BULK SAMPLER EXPERIMENTS Table E.1: Effect of sampler type and sample depth on calculated bulk density, porosity and free air space. TYPE DEPTH REPLICATION AVG STD CV 4 5 7 DEV Moisture Content RING 1 80.6 81.4 81.8 81.3 0.49 0.006 2 82.1 83.0 81.2 82.1 0.72 0.009 3 81.6 81.1 81.7 81.5 0.26 0.003 AVG 81.4 81.8 81.6 STD 0.61 0.83 0.24 CV 0.007 0.010 0.003 SHORT 1 81.0 81.4 81.9 81.4 0.36 0.004 2 80.8 81.0 81.8 81.2 0.43 0.005 3 82.6 80.7 80.9 81.4 0.85 0.010 AVG 81.5 81.1 81.5 STD 0.83 0.26 0.42 CV 0.010 0.003 0.005 LONG 1 79.6 80.9 81.9 80.8 0.91 0.011 2 80.3 81.2 81.6 81.0 0.54 0.007 3 80.1 80.7 81.4 80.7 0.53 0.007 AVG 80.0 80.9 81.6 STD 0.28 0.19 0.18 CV 0.004 0.002 0.002 Table F.1 (cont'd.). TYPE DEPTH REPLICATION AVG STD CV 4 5 7 DEV Volatile Solids RING 1 93.5 94.3 94.4 94.0 0.42 0.004 2 94.8 94.6 94.8 94.7 0.09 0.001 3 93.8 94.9 94.7 94.5 0.50 0.005 AVG 94.0 94.6 94.6 STD 0.56 0.25 0.17 CV 0.006 0.003 0.002 SHORT 1 94.1 95.0 94.3 94.5 0.36 0.004 2 94.9 94.9 94.3 94.7 0.26 0.003 3 94.8 95.0 94.2 94.7 0.37 0.004 AVG 94.6 95.0 94.3 STD 0.33 0.06 0.07 CV 0.004 0.001 0.001 LONG 1 94.3 94.6 94.4 94.4 0.15 0.002 2 94.5 94.0 94.5 94.3 0.21 0.002 3 94.4 94.6 93.7 94.3 0.38 0.004 AVG 94.4 94.4 94.2 STD 0.09 0.28 0.34 CV 0.001 0.003 0.004 Table F.1 (cont'd.). 288 TYPE DEPTH REPLICATION- AVG STD CV 4 5 7 DEV Full Bag Weight (gm) RING 1 467.50 495.80 413.75 2 460.70 556.20 492.50 3 500.20 446.50 508.80 SHORT 1 _296.33 317.43 311.67 2 425.91 409.90 371.03 3 447.10 408.50 454.30 LONG 1 221.08 192.38 189.91 2 237.26 224.72 216.72 3 291.44 312.68 276.08 Net Solids Weight (gm) RING 1 458.00 486.30 404.25 2 451.20 546.70 483.00 3 490.70 437.00 499.30 SHORT 1 286.83 307.93 302.17 2 416.41 400.40 361.53 3 437.60 399.00 444.80 LONG 1 211.58 182.88 180.41 2 227.76 215.22 207.22 3 281.94 303.18 266.58 Length (in) RING 1 2.0 2.0 2.0 2.0 0.00 0.000 2 2.0 2.0 2.0 2.0 0.00 0.000 3 2.0 2.0 2.0 2.0 0.00 0.000 SHORT 1 12.0 12.0 12.0 12.0 0.00 0.000 2 11.5 12.5 12.0 12.0 0.41 0.034 3 9.5 7.0 6.5 7.7 1.31 0.171 LONG 1 13.0 12.5 14.0 13.2 0.62 0.047 2 10.5 12.5 10.0 11.0 1.08 0.098 3 10.0 10.0 10.0 10.0 0.00 0.000 289 Table F.1 (cont'd.). TYPE DEPTH REPLICATION AVG STD CV 4 5 7 DEV Bulk Density (kg/m‘3) RING 1 278 295 245 273 20.7 0.076 2 274 332 293 300 24.1 0.080 3 298 265 303 289 16.7 0.058 AVG 283 297 281 STD 10.5 27.2 25.2 CV 0.037 0.092 0.090 SHORT 1 198 213 209 206 6.1 0.030 2 300 265 250 272 21.0 0.077 3 382 472 567 474 75.7 0.160 AVG 293 317 342 STD 75.1 112.0 160.1 CV 0.256 0.354 0.469 LONG 1 202 182 160 181 17.2 0.095 2 270 214 258 247 23.9 0.097 3 350 377 331 353 18.7 0.053 AVG 274 258 250 STD 60.6 85.4 70.1 CV 0.221 0.331 0.281 Table F.1 (cont'd.). 290 TYPE DEPTH REPLICATION AVG STD CV 4 5 7 DEV Porosity RING 1 0.967 0.964 0.971 0.967 0.003 0.003 2 0.884 0.859 0.876 0.873 0.010 0.012 3 0.964 0.969 0.964 0.965 0.002 0.002 AVG 0.938 0.930 0.937 STD 0.039 0.050 0.043 CV 0.041 0.054 0.046 SHORT 1 0.976 0.974 0.975 0.975 0.001 0.001 2 0.962 0.967 0.970 0.967 0.003 0.003 3 0.957 0.941 0.930 0.942 0.011 0.012 AVG 0.965 0.961 0.959 STD 0.008 0.014 0.020 CV 0.008 0.015 0.021 LONG 1 0.973 0.977 0.981 0.977 0.003 0.003 2 0.965 0.974 0.969 0.969 0.003 0.004 3 0.955 0.953 0.960 0.956 0.003 0.003 AVG 0.964 0.968 0.970 STD 0.008 0.011 0.009 CV 0.008 0.011 0.009 Table F.1 (cont'd.). 291 TYPE DEPTH REPLICATION AVG STD CV 4 5 . 7 DEV Free Air Space RING 1 0.741 0.724 0.770 0.745 0.019 0.026 2 0.743 0.688 0.726 0.719 0.023 0.032 3 0.721 0.752 0.716 0.730 0.016 0.022 AVG 0.735 0.721 0.737 STD 0.010 0.026 0.023 CV 0.013 0.036 0.032 SHORT 1 0.815 0.801 0.805 0.807 0.006 0.007 2 0.720 0.752 0.766 0.746 0.019 0.026 3 0.641 0.559 0.471 0.557 0.070 0.125 AVG 0.726 0.704 0.681 STD 0.071 0.104 0.149 CV 0.098 0.148 0.219 LONG 1 0.812 0.830 0.850 0.831 0.015 0.019 2 0.749 0.800 0.759 0.769 .0.022 0.029 3 0.674 0.648 0.690 0.671 0.017 0.026 AVG 0.745 0.760 0.766 STD 0.057 0.080 0.065 CV 0.076 0.105 0.085 292 Table F.2: Statistical analysis of bulk density and free air space data. COMPARISON STD DEV F RESULT POOLED STUDENT SIGNIF. RATIO * (95%) STD DEV T ‘ LEVEL Bulk Density Ring/ 25.3/7.529 11.3 nsd 40.00 2.03 nsd 95 Short 29.519/25.772 1.3 nsd 29.13 1.18 nsd 95 20.414/92.956 16.0 nsd 114.60 1.98 nsd 95 Ring/ 25.3/21.07 1.4 nsd 54.20 2.07 nsd 95 Long 29.519/29.258. 1.0 nsd 38.98 1.65 nsd 95 20.414/22.845 1.2 nsd 40.16 1.96 nsd 95 Short/ 7.529/21.070 7.8 nsd 19.83 1.54 nsd 95 Long 25.772/29.258 1.3 nsd 28.08 1.07 nsd 95 92.956/22.845 16.6 nsd 89.61 1.65 nsd 95 Free Air Space Ring/ 0.0233/0.007 11.1 nsd 0.037 2.04 nsd 95 Short 0.0282/0.0236 1.4 nsd 0.028 1.20 nsd 95 0.0195/0.085 19.0 sd 0.110 1.94 nsd 95 Ring/ 0.0233/0.019 1.2 nsd 0.058 2.08 nsd 95 Long 0.0282/0.027 1.1 nsd 0.037 1.66 nsd 95 0.0195/0.212 1.2 nsd 0.037 1.95 nsd 95 Short/ 0.007/0.019 7.4 nsd 0.018 1.61 nsd 95 Long 0.0236/0.027 1.3 nsd 0.026 1.08 nsd 95 0.085/0.0212 16.1 nsd 0.083 1.68 nsd 95 * Larger standard deviations placed on top for calculations. n = 3 for all treatments. APPENDIX G DERIVATION OF ERROR PROPAGATION EQUATIONS APPENDIX G DERIVATION OF ERROR PROPAGATION EQUATIONS Assume that u = f(x,y) and that all errors are independent and may be treated as random. Further assume that all deviations Gxi = xi - x and Gyi = Yi - y are relatively small. Applying the Taylor series expansion, neglecting higher order terms, we have “i = f([x + 6xi], [y + Gyil) = f(x,y) + an * 6x- + au * Gy- (G.l) 3; 1 5y 1 and, 6u- = u- - u = au 6x. + an Gy- (6.2) 1 1 5; 1 5? 1 By definition, the square of the standard deviation in u, SD, is su2 = 2 (Gui)z / n (6.3) Squaring 6.2 gives (Gu-)z = (8u)z(6x-)z + 2 an au 6x- 6y- + (au)2(6y-)2 1 8i 1 5i 5? 1 1 5? 1 (6.4) Placing this expression into 6.3, we have s 2 = (Bu)2 2(bx-)z + 2 au 8u £(bx- by-) + (an)2 £(by-)2 u 3? 1 5i 3? 1 1 5? 1 n (G.5) As n increases, the sum (dxiéyi) goes to 0 if xi and Yi are independent because any bxi byi is likely to be positive as negative. Since 293 294 5,,2 = £(6xi)2/ n and 5Y2 = £(6yi)2/ n (6.6) can be substituted into G.5 to get 2 z 2 z z s = (an) s + (an) s (6.7) u 5; x §§ y For more than 2 terms, 6.7 can be generalized to s 2 = 2(3u )2 s -2 (6.8) u 5;. X] 1 The fractional variance in U is written by dividing su by u: (su / a )2 = (2(au)z sijJ / a: (0.9) X The fractional variance should not be used unless the 0 of each xo 1 and u scales are physically significant (Parrot, 1961). Sum or Difference Let u = x i y then ' 8u = l and Bu = :1 ’52? ”517 substituting into G.8, we have su’ = 5,,2 + syz (6.10) The fractional standard deviation is - 2 - 2 2 "'2 (sn / U) - (5x + sy ) / U (6.11) Product 95 Quotient Let ' u = x8 + Yb (G.12) with a and b assumed to be exact constants. Then au = axa_lyb and Bu = bxayb'l (G.13) F? F? Substituting the above equation into equation G.8, we 295 have 2 = U a s 2 x2(a-1)Y2bs 2 + b2x2ay2(b—l)sy2 (0.14) X The fractional variance of a product is given by (s / U)2 = (a2x2a-2Y2bsx2/ 02) + (bZXZaYZb-25y2/ U2) U 2X2a-2Y2bsx2/ §2a§2b) + (bZXZaYZb—Zsy2/ §2a§2b) (a = (aZsz/ 22) + (bzsyz/ Y2) (0.15) APPENDIX H WINDROW SIZE, TEMPERATURE, GAS CONCENTRATION AND PHYSICAL PROPERTY SAMPLING LOCATIONS AND DATA AVAILABILITY APPENDIX H WINDROW SIZE, TEMPERATURE, GAS CONCENTRATION AND PHYSICAL PROPERTY SAMPLING LOCATIONS AND DATA AVAILABILITY Complete data on windrow size, temperature and ambient conditions, gas concentration and physical properties are available from the author at the following address: Stephen E. Ferns c/o Dr. John Gerrish Department of Biological and Agricultural Engineering A.W. Farrall Hall East Lansing, Michigan 48823 USA The data has the following formats and requirements: Temperature: Symphony (1) 7 360 KB floppy disks; Windrow Elevations, Gas Concentrations, and Physical Properties: Lotus 123 (1A) 2 360 KB floppy disks. Please send the necessary disks and disk mailer with your request. 296 L I ST OF REFERENCES J--" '- LIST 93 REFERENCES Acott, K. M. and T. P. Labuza. 1975. Microbial Growth Response to Water Sorption Preparation. J. Food Technol. 10:603-611. Cited in Christian, 1980, op cit. Adams, D. M. 1978. Heat Injury of Bacterial Spores. Advances in Applied Microbiology 23:245-261. Allen, S., M. Wallentine, S. Austin, P. Burch and K. Hoopes. 1979. Recycled Manure Solids as a Free-stall Bedding for Lactating Dairy Cows, and its Association with Mastitis. Report from the National Mastitis Council. pp. 121—127. Allwood, M. C. and A. D. Russell. 1970. Mechanisms of Thermal Injury in Nonsporulating Bacteria. Advances in Applied Microbiology 15:245-261. Andrews, J. F. and K. Kambhu. 1973. Thermophilic Aerobic Digestion of Organic Solid Wastes. Office of Research and Development, 0.8. EPA NTIS PB-222-396. Cited in Haug, 1980. op. cit. ASTM. 1986. Annual Book of ASTM Standards: Soil & Rock, Building Stones. SEction 4, Vol. 04.08. - American Society of Testing and Materials. Washington, D.C. AWWA. 1976. Standard Methods for the Examination of Water and Wastewater. “IIfE'EaT‘1976. AWWA, APHA, afia“‘ WPCF TJointly published), Washington, DC. Atlas, R. M. and R. Bartha. 1981. Microbial Ecology: Fundamentals and Applications. Addison-Wesley Publishing Company, Reading, Mass. Baranowski, R. 1983. Evaluation of the Useability of Procedures for the Determination of Soil Physical Properties in Tillage Experiments. Institute of Soil Science and Plant Cultivation. 55-230. Luskowice Olawskia, Poland. Cited in Raper, R. L. and D. C. Erbach. 1985. Accurate Bulk Density Measurements Using a Core Sampler. ASAE Paper 85-1542. 297 298 Bartholomew, W. V. and A. G. Norman. 1953. Microbial Thermogenesis in the Decompositon of Plant Materials IV. Influence of Moisture Content and of Initial Temperature. J. Bacteriol. 65:228-232. Bauer, L. D. 1959. Soil Physics. Third Edition. John Wiley and Sons. Cited in Raper, R. L. and D. C. Erbach.. 1985. Accurate Bulk Density Measurements Using a Core Sampler. ASAE Paper 85-1542. Bailey, J. E. and D. F. Ollis. Biochemical Engineering Fundamentals. McGraw-Hill, San FranciSco, CA. Beckwith, C. P. and J. W. Parsons. 1980. The Influence of Mineral Amendments on the Changes in the Organic Nitrogen Components of Composts. Plant and Soil 54:259-270. Bejan, A. 1984. Convection Heat Transfer. John Wiley and Sons. New York, NY. Bell, R. G. and J. Pos. 1971. Design and Operation of a Pilot Plant for Composting Poultry Manure. Trans. of the ASAE 14(6):1020-1023. Bishop, J. R., J. J. Jansen, and A. B. Bodine. 1980. Effect of Ambient Environments on Survival of Selected Bacterial P0pulations in Dairy Waste Solids. J. Dairy Sci. 63(3):523-525. Bishop, J. R., J. J. Janzen, A. B. Bodine, C. A. Caldwell and D. W. Johnson. 1981. Dairy Waste Solids as a Possible Source of Bedding. J. Dairy Sci. 64(4): 706-711. Blosser, T. H. 1979. Economic Losses from and the National Research Program on Mastitis in the United States. J. Dairy Sci. 62(1):ll9-127. Bohnhoff, D. R., J. C. Converse and J. B. Petersen. 1984. Thermal and Physical Properties of Separated Manure Solids. ASAE Paper No. 84—4077. Am. Soc. of Agr. Eng., St. Joseph, MI 49085. Bohnhoff, D. R. and J. C. Converse. 1986a. Engineering Properties of Se arated Manure Solids. Agricultural Wastes (in press . Bohnhoff, D. R. and J. C. Converse. 1986b. Water Desorption Properties of Separated Manure Solids. Agricultural Wastes (in press). Braithwaite, R. L. 1956. A Study of Garbage Composting in Controlled, Insulated Barrels. Unpublished M.S. 299 Thesis. Michigan State University, E. Lansing, MI 48824. Bramely, A. J. 1974. The Aetiology and Control of Coliform Mastitis in Dairy Cattle. PhD Dissertation, University of Reading, Reading, England. Cited in Carroll, L. J. 1977. Environmental factors in bovine mastitis. JAVMA l70(10):1l43-ll49. Bramley, A. J. and F. R. Neave. 1975. Studies on the Control of Coliform Mastitis in Dairy Cows. Brit. Vet. J. 131:160-169. Brander, G. C. 1972. Dairy Herd Environment and the Control of Mastitis. Vet. Rec. 92:501-506. Brandon, J. R. 1976. Sandia's Sludge Irradiation Program. In Slud e Mggagement, Disposal and Utilization. Information Tranéfer. Rockville, Md. Cited in Haug, 1980, op cit. Bretzloff, C. W. and M. S. Fluegel. 1962. Chemical Compositon of Mushroom Compost During Composting and Cropping. Mushroom Sci. 5:46-60. Brooker, D. B., F. W. Bakker-Arkema and C. W. Hall. 1981. Dr in Cereal Grains. AVI Publishing Company. Wesfporf, CN. Brown, M. H. and O. Emberger. 1980. Oxidation-reduction Potential. In Microbial Ecolo of Foods I. Factors Affecting Life and Death of MicroEFganismST— J.H. SilliRer et al., eds. *Acidemic Press. New York, NY. Browne, C. A. 1929. The Spontaneous Combustion of Hay. U.S. Dept. of Agric. Tech. Bull. 141:1-38. Cited in Cooney and Emerson, 1964, op. cit. Burge, W., D. Colacicco, W. Cramer and E. Epstein. 1978. Criteria for Control of Pathogens During Sewage Sludge Composting. In Proceedings of the National Conference on Desi n of*Munici a1 SIfidge Compost Facilities. InformatTEn TransIer. RockvilIe, Md. Busta, F. F. 1978. Introduction to Injury and Repair of Microbial Cells. Adv. Appl. Microbiology 23:195-201. Burrows, S. 1951a. The Chemistry of Mushroom Composts. I. General Introduction and Methods of Investigation. J. Sci. Food Agric. 2(9):395-403. 300 Burrows, S. 1951b. The Chemistry of Mushroom Composts. II. Nitrogen Changes During the Composting and Cropping Processes. J. Sc1. Food Agric. 2(9): Campbell Scientific. 1985. CR21X Microlo er Instruction Manual. CampEeII Scientific, Inc. Provo, Utah. Carroll, E. J. 1977. Environmental Factors in Bovine- Mastitis. JAVMA 170(10):1143—1149. Carroll, E. J., N. C. Jain, O. W. Schalm, and J. Lasmanis. 1973. Experimentally Induced Coliform Mastitis: Innoculation of Udders with Serum-sensitve and Serum Resistant Organisms. Am. J. Vet. Res. 34(9):1143-1146. Carroll, E. J. and D. E. Jasper. 1978. Distribution of Enterobacteriaceae in Recycled Manure Bedding on California Dairies. J. Dairy Sci. 61:1498-1508. Carroll, E. J. and D. E. Jasper. 1979. Coliform Populations in Bedding Materials and Coliform Mastitis Incidence. University of California, Davis, CA. Chang, A. C. and J. M. Rible. 1975 Particle-size Distribution of Livestock Wastes. In Managing Livestock Wastes--Proceedings of the 3rd International Symposium on LivEEtEEK WEEtes. American Society ongriculturaI Engineers, St. Joseph, MI. pp. 339-342. Chang, C. S., F. S. Lai and B. S. Miller. 1980. Composting of Grain Dust. Trans. of the ASAE 23(3):709-711. Chang, Y. and H. J. Hudson. 1967a. The Fungi of Wheat Straw Compost I. Ecological Studies. Trans. Br. mycol. Soc. 50(4):649-666. Chang, Y. 1967. The Fungi of Wheat Straw Compost II. Biochemical and Physiological Studies. Trans. Br. mycol. Soc. 50(4):667-677. Chatt, E. M. 1953. Cocoa: Cultivation, Processin , Analysis. Interscience. New York, NY. C1ged in Cooney and Emerson, 1964, op.cit. Chau, K. V., J. J. Gaffney and S. Bellagha. 1984. Simulation of Heat and Mass Transfer in Products with Internal Heat Generation and Transpiration. ASAE Paper No. 84-6513. Am. Soc. Agr. Eng. St. Joseph, MI 49085. 301 Chen, Y. R. 1982. Engineering Properties of Beef Cattle Manure. ASAE Paper No. 82-4085. Am. Soc. Agr. Eng. St. Joseph, MI 49085. Chen, Y. R. 1983. Thermal Properties of Beef Cattle Manure. Agricultural Wastes 6:13-29. Chen, W. W. H, R. E. Zimmerman and A. G. Franklin. 1977. Time Settlement Characteristics of Milled Urban Refuse. In Geotechnical Practice for Disposal of Solid Waste Materials. Am. Soc. Civil Eng., New York, NY. pp. 136-152. Cohn, F. 1889. Uber thermogene Wirkung von Pilzen. Jahresber. Schles. Geselsch. Vaterl. Kult. 66: 150-156. Cited in Cooney and Emerson, 1964, op. cit. Cooney, D. G. and R. Emerson. 1964. Thermophilic Fungi: An Account of Their Biolo , ActiVities, and CIass1f1cati5n. W. H. reeman andiCompany, San Fiansisco, CA. Corlett, Jr., D. A. and M. H. Brown. 1980. pH and Acidity. In Microbial Ecology of Foods 1; Factors Affecting Life and Deat g_ 1croorganisms. J.H. SiIliker et aI., eds. Academic Press, New York, NY. Christian, J. H. B. 1980. Reduced Water Activity. In Microbial Ecolo of Foods 1. Factors Affecting Life and Death 9: M1cr05?gan1sm§:— J.H. S111iker éf al., 53—. Academic Press, New York, NY. Cussler, E. L. 1984. Diffusion Mass Transfer in Fluid S stems. Cambridge University Press, Cambr1dge, EngIand. Davis, M. L. 1981. Water and Wastewater Analysis. Department of C1v11 Engineering, MiChiganiState University, East Lansing, MI. Dilley, D. 1986. Professor, Department of Horticulture, Michigan State University, East Lansing, Michigan 48824. Personal Communication. Douglas, F. R. Hambleton, and G. J. Rigby. 1973. An Investigation of the Oxidation-Reduction Potential and of the Effect of Oxygen on the Germination and Outgrowth of Clostridium Butyricum Spores, Using Platinum Electrodes. J. Appl. Bacteriol. 36:623- 633. Cited in Brown and Emberger, 1980, op. cit. Eberhart, R. J. 1977. Coliform Mastitis. JAVMA 170(10): 302 1160-1163. Eberhart, R. J. and J. M. Buckalew. 1972. Evaluation of a Hygine and Dry Period Therapy Program for Mastitis Control. J. Dairy Sci. 55:1683-1690. Farnsworth, R. J. 1977. Significance of Fungal Mastitis. JAVMA l70(10):1173-1174. Finger, 5. M. 1975. Aerobic Microbial Growth in Semi -Solid Matrices. Ph.D Dissertation. University of Maryland, College Park, Md. Finger, S. M., R. T. Hatch, and T. M. Regan. 1976. Aerobic Microbial Growth in Semi-Solid Matrices: Heat and Mass Transfer Limitation. Biotechnology and Bioengineering 18:1193-1218. Finnish Peatland Society. 1982. Peatlands and Their Utilization in Finland. Finnish Peatland Society, Hel§inEi, FinIand. Finstein, M. S. 1980a. Composting Microbial Ecosystem: Implications for Process Design and Control. Compost Science/Land Utilization 21(4):25-27. Finstein, M. S., J. Cirrello, S. T. MacGregor, F. C. Miller, and K. M. Psarianos. 1980b. Slud e Composting and Utilization: Rational Approacfi to Process Contr6i} Department of Env1ronmenta c1efiEe,iRutgers--The State University of New Jersey, Cook College. New Brunswick, New Jersey 08903. USEPA P882-136243. Finstein, M. 8., F. C. Miller, P. F. Strom, S. T. MacGregor and K. M. Psarianos. 1983. Composting Ecosystem Management for Waste Treatment. Bio/Technology 1(4):347-353. Francis, P. G., J. Sumner and D. A. Joyce. 1981. The Influence of the Winter Environment of the Dairy Cow on Mastitis. The Bovine Practitioner 16(11):24-27. Freese, E., C. W. Sheu and E. Galliers. 1973. Function of Lipophilic Acids as Anitmicrobial Food Additives. Nature (London) 241:321-325. Cited in Corlett and Brown, 1980, op. cit. Frost, A. J., D. D. Wanasinghe and J. B. Woolcock. 1977. Some Factors Affecting Selective Adherence of Micro-organisms in the Bovine Mammary Gland. Infection and Immunity 15(1):245-253. Garner, W. W. 1946. The Production of Tobacco. Blakiston, Phiadelphia. Cited in Cooney and Emerson, 1964, 303 op.cit. Geankoplis, C. J. 1983. Transport Processes and Unit Operations. Allyn and Bacon, Inc. Boston, MA. Gerrits, J. P. G., H. C. Bels-Koning and M. Muller. 1965. Changes in Compost Constituents During Composting, Pasteurization and Cropping. Mushroom Sci. 5: 225-243. Gerrits, J. P. G. 1972. The Influence of Water in Mushroom Compost. Mushroom Sci. 8:43-57. Glathe, H. 1960. Selbsterhitzung und selbstenzundung von erntestoffen und ihre verhutung. Ergeb. landwirtsch. Forsch. 3:83-98. Cited in Cooney and Emerson, 1967, op. cit. Gregory, P. H., M. E. Lacey, G. N. Festenstein and F. A. Skinner. 1963. Microbial and Biochemical Changes During the Moulding of Hay. J. gen. Microbiol. 33: 147-174. Cited in Finger and Morris, 1975, op. cit. Griffis, C. L. and C. R. Mote. 1978. Cotton Gin Trash Composting Studies. Ark. Farm Res. Ark. Ag. Exp. Stn. 27(4):3. Groffman, P. 1986. Personal Communication. Research Associate. Department of Crops and Soils Sciences. Michigan State University, East Lansing, MI 48824. 6/86. Hansen, N. H. and H. Riemann. 1963. Factors Affecting the Heat Resistance of Nonsporing Organisms. J. Appl. Bact. 26(3):314-333. Hajny, G. J. 1966. Wood Chip Storage: A Review and Bibliography. TAPPI 49(10):97A-105A. Haug, R. T. 1979. Engineering Principles of Sludge Composting. JWPCF 51(8):2189-2206. Haug, R. T. 1980. Compost Engineering: Principles and Practice. Ann Arbor Science Publishers. Ann Arbor, MI. Hauser, E. and J. W. Sinden. 1953. The Nature of the Composting Process and its Relation to Short Composting. Mushroom Sci. 2:123-130. Hermanson, R. E. 1985. Flush Cleaning Dairy Barns: Case Studies. In Agricultural Waste Utilization and Management--Proceedings of the Fifth International Sym osium on AgriculturaI_Wastes. December 16-17, 1985. Afi.-Soc. Agr. Eng., St.iJOSeph, Michigan. 304 pp. 590-597. Hewitt, L.F. 1950. Oxidation-reduction Potentials in Bacteriology and Biochemistry. Sixth EditiofiT Livingston, Edinburgh. Cited in Brown and Emberger 1980, op. cit. Higgins, A. J., A. J. Kaplovsky and J. V. Hunter. 1982. Organic Composition of Aerobic, Anaerobic, and Compost-stabilized Sludges. JWPCF 54(5):466-473. Hillel, D. 1982. Introduction to Soil Physics. Academic Press, New York: NY. Hoitink, H. A. J. and H. A. Poole. 1980. Bark Compost Use in Container Media. Compost Science/Land Utilization 21(3):38-4l. Houkom, R. L., A. F. Butchbaker and G. H. Brusewitz. 1972. Thermal Properties of Beef Manure. ASAE Paper No. 72-316. Am. Soc. Agr. Eng. St. Joseph, MI 49085. Hoyle, D. A. and G. E. G. Mattingly. 1959. Studies on Composts Prepared from Waste Materials I: Prepartion, Nitrogen Losses and Changes in 'Soluble Nitrogen.‘ J. Sci. Food Agric. 5(1):54-65. Hummel, J. W. and G. B. Willson. 1975. High Rate Composting of Dairy Manure. In Proceedings of the International Symposium on Livestock Wastes,_AmT_— Soc. Agr. Efi§., St. Josepfi, MI pp. 4ST3434. Hurst, A. 1980. Injury and Its Effect on Survival and Recovery. In Microbial Ecolo of Foods I. Factors Affecting Life and Death of M1cr6:organ1sfi§. J.H. Silliker et al., eds. Academic Press. New York, NY. Ishida, M. and C. Y. Wen. 1971a. Comparison of Zone- Reaction Model and Unreacted Shrinking-core model in Solid-gas Reactions--I. Isothermal Analysis. Chem. Engng Sci. 26(7):1031-104l. Ishida, M., C. Y. Wen and T. Shirai. 1971b. Comparison of Zone-reaction Model and Unreacted Shrinking-core Model in Solid-gas Reactions--II. Non-isothermal Analysis. Chem. Engng Sci. 26(7):1043-1048. Jackson, H. and M. Woodbine. 1962. The Effect of Sublethal Treatment on the Growth of Staphylococcus Aureus. J. Appl. Bact. 25, viii. Cited in Hansen and Riemann, 1963, op. cit. Jackson, R. D. 1964. Water Vapor Diffusion in Relatively 305 Dry Soil: 1. Theoretical Considertions and Sorption Experiments. Soil Sci. Soc. Am. Proc. 28:172-176. Cited in Hillel, 1982, op.cit. Jain, N. C. 1979. Common Mammary Pathogens and Factors in Infection and Mastitis. J. Dairy. Sci. 62(1): James, L. H., L. F. Rettger and C. Thom. 1928. Microbial Thermogenesis II. Heat Production in Moist Organic Materials with Special Reference to the Part Played by Micro-organisms. J. Bacteriol. 15:117-141. Jarrett, J. A. 1981. Mastitis in Dairy Cows. Veterninary Clinics of North America: Large Animal Practice 3(2):447-455. Jarret, P. 1986. Personal Communication. Professor of Civil Engineering. Royal Military College of Canada. Kingston, Ontario K7L2W3. 5/86. Jasper, D. E. 1977. Mycoplasma and Mycoplasma Mastitis. JAVMA l70(10):1167-ll72. Jasper, D. E., J. D. Dellinger, and R. B. Bushnell. 1975. Herd Studies on Coliform Mastitis. JAVMA 166(14): Jeris, J. S., and R. W. Regan. l973a. Controlling Environmental Parameters for Optimum Composting, Part I: Experimental Procedures and Temperature. Compost Science. Jan/Feb: 22-28. Jeris, J. S., and R. W. Regan. 1973b. Controlling Environmental Parameters for Optimum Composting, Part II: Moisture, Free Air Space, and Recycle. Compost Science. March/April: 8-22. Jeris, J.S., and R.W. Regan. 1973c. Controlling Environmental Parameters for Optimum Composting, Part III: pH, Nutrients, Storage and Paper Content Relative to Composting. Compost Science. July/August: 16-22. Kane, B. E. and J. T. Mullins. 1973. Thermophilic Fungi in a Municipal Waste Compost System. Mycologia 65: 1087-1100. Karak, S. and Y. Yener. 1979. Heat Conduction. Middle East Technical University Press. Ankara, Turkey. Kaufmann, C. W., L. G. Harmon, O. C. Pailthorp and I. J. Pflug. 1959. Effect of heat treatment on the growth of surviving cells. J. Bact. 78:834. Cited 306 in Hansen and Riemann, 1963, op. cit. Keys, Jr., J. E., L. W. Smith, and B. T. Weinland. 1977. Response of Dairy Cattle Given a Free Choice of Free -sta11 Location and Three Bedding Materials. J. Dairy Science 56(6):1157-1162. Lacey, J. 1973. In Actinomycetales: Characteristics and Practical Importance. G. Sykes and F.A. Skinner, eds. pp. 231-251. Academic Press. New York. Cited in Finstein and Morris, 1975, op. cit. Lambert, E. B. 1941. Studies on the Preparation of Mushroom Compost. J. Agric. Res. 62(7):415-422. Lambert, E. B. and A. C. Davis. 1934. Distribution of Oxygen and Carbon Dioxide in Mushroom Compost Heaps as Affecting Microbial Thermogenesis, Acidity, and Moisture Therein. J. Agric. Res. 48(7):587-601. Latif, N. and E. Lissik. 1986. Respiration Model for Heat and Moisture Release During Grain Storage. ASAE Paper No. 86-6508. Am. Soc. Agr. Eng., St. Joseph, MI 49085. Lembke, A. 1937. Die Hitzewiderstandsfaghigkeit der Colibakterien und die verwendbarkeit dieser eigenschaft als vergleichmasstab fur die Beurteilung von Milcherhitzern. Zbl. Bakt. (Abt. 2) 96:92. Cited in Hansen and Riemann, 1963, op. cit. Lindemann, E. R., J. M. Sweeten, and J. P. Burt. 1985. Sludge Removal from Dairy Lagoons. In Agricultural Waste Utilization and Management--Proceedings of*the Fifth InternationaI—S osium on Agriculturai WESEEE, Am. Soc. Agr. Eng., St. UosephT—MI. pp. Lombardo, P. 1977. Septage Composting. Compost Science 18(6):12-14. Marshall, B. J., D. F. Ohye and J. H. B. Christian. 1971. Tolerance of Bacteria to High Concentrations of NaCl and Glycerol in the Growth Medium. Appl. Microbiol. 21:363-364. Cited in Christian, 1980, op. cit. Martin Jr., J. H., M. Decker Jr., and K. C. Das. 1972. Windrow Composting of Swine Wastes. In Waste Management Research: Proceedings of the I972 Cornell Agricultual Waste Managemefit Conference. Giaphics Management Corp., Washington, D.C. pp. 159-172. McKinley, V. L. and J. R. Vestal. 1984. Biokinetic 307 Analysis of Adaptation and Succession: Microbial Activity in Composting Sewage Sludge. Appl. Environ. Microbiol. 47(5):933-941. Mears, D. R., M. E. Singley, G. Ali and F. Rupp III. 1975. Thermal and Physical Properties of Compost. In Ener , Agriculture and Waste Management: Proceedings ofthe 1975 COrnelI AgricuItural Waste Management Conference. W.J. JewelI, ed. Ann Arbor Science Publishers. Ann Arbor, MI. Meijering, A. G., W. H. Clifford and F. W. Bakker-Arkema. 1972. Drying a Bed of Composted Waste. Trans. of the ASAE. 15(1):116-120. Miehe, H. 1907. Die Selbsterhitzung des Heus. Eine Biologische Stuie. Gustav Fischer. Jena. ppl-127 Cited in Cooney and Emerson, 1964, op. cit. Miehe, H. 1930. Die Warmebildung von Reinkulturen im Hinblick auf die Atiologie der Selbsterhitzung pflanzlicher Stoffe. Arch. Mkrobiol., 1:78-118. Cited in Cooney and Emerson, 1964, op. cit. Moats, W. A. 1971. Kinetics of Thermal Death of Bacteria. J. Bacteriol. 105(1):165-l71. Moore, A.F. 1958. Oxygen Uptake Rates and Respiratory Quotients of an Aerobically Decomposing Synthetic Garbage. M.S. Thesis. Sanitary Engineering Department, Michigan State University, E. Lansing, MI 48834. Mote, C. R. and C. L. Griffis. 1980. Variations in the Composting Process for Different Organic Carbon Sources. Agricultural Wastes 2(3):215-223. Mote, C. R. and C. L. Griffis. 1980. Water Management for Sustaining the Composting Process in Windrows of cotton gin trash. Report Series 256, Agricultural Experiment Station, Division of Agriculture, University of Arkansas, Fayetteville, Arkansas. Mote, C. R. 1985. Personal Communication. Professor. Department of Agricultural Engineering, University of Tennesee--Knoxville. Knoxville, Tennesee. 12/85. Muller, M. 1965. Some Thoughts About Composting. Mushroom Sci. 5:213-223. Nat. Mastitis Counc. 1978. Current Concepts of Bovine Mastitis. National Mastiti§ Council, THC. Washington, DC. 308 Nat. Res. Counc. Can. 1979. Peat Testing Manual. Nat. Res. Counc. Can., Assoc. Comm. Geotech. Res. [Muskeg Subcomm.]. Technical Memo 125. pp 1-8. Neave, F. K. and J. Oliver. 1962. The Relationship Between the Number of Mastitis Pathogens Placed on the Teats of Dry Cows, Their Survival, and the Amount of Intramammary Infections Caused. J. Dairy Res 29:79-93. Olson Jr., J. C. and P. M. Nottingham. 1980. Temperature. In Microbial Ecolo of Foods I. Factors Affectin Life and Deatfi g: 31CFB-6?§35i§fisT_—UTHT SiiiiREFS et al., edS. Academic Press, New York, NY. Packer, R. A. 1977a. Bovine Mastitis Produced by Cornebacteria. JAVMA l70(10):1164-1165. Packer, R. A. 1977b. Bovine Mastitis Caused by Pseudomonas Aeruginosa. JAVMA l70(10):1175. Parratt, L. G. 1961. Probability and Experimental Errors In Science. John Wiley and Sons, New York, NY. Pizer, N. H. 1950. Horse Manure Composts. Mushroom Sci. 1:46‘510 Poincelot, R. P. 1977. The Biochemistry of Composting. In Composting of Munici a1 Residues and Sludges, Proceedings g£_fhe 1977 National Conference. Information Transfer. RockvilIe, MD. pp. 33-39. Pratt, A. W. 1969. Heat Conduction in Low Conductivity Materials. In Thermal Conductivity, Vol 1. R.P. Tye, ed. Academic Press. New York, NY. Raper, R. L. and D. C. Erbach. 1985. Accurate Bulk Density Measurements Using a Core Sampler. ASAE Paper No. 85-1542, Am. Soc. Agr. Eng., St. Joseph, MI 49085. Rendos, J. J., R. J. Eberhart and E. M. Kesler. 1975. Microbial Populations of Teat Ends of Dairy Cows and Bedding Materials. J. Dairy Sci. 58(10):l492-1500. Rodriguez, A. C., G. H. Smerage, A. A. Teixeira, and F. F. Busta. 1986. Kinetic Effects of Lethal Temperatures on Spore Population. Paper No. 86- 6533. Am. Soc. Agr. Eng., St. Joseph, MI 49085. Rose, W. W., J. E. Chapman, S. Roseid, A. Katsuyama, V. Porter and W. A. Mercer. 1965. Composting Fruit and Vegetable Refuse. Compost Science 6(3):l3-25. 309 Rosenberg, S. L. 1975. Temperature and pH Optima for 21 Species of Thermophilic and Thermotolerant Fungi. Can. J. Microbiol. 21:1535-1540. Ryan, D., R. G. Carbonell and S. Whitaker. 1981. A Theory of Diffusion and Reaction in Porous Media. AIChE Symposium Series 77(202):46-62. Schalm, O. W., E. J. Carroll and N. C. Jain. 1971. Bovine Mastitis. Lea and Febiger, Philadelphia, PA. pp. Schulze, K. L. 1962. Continuous Thermophilic Composting. Applied Microbiology 10:108-122. Scott, W. J. 1957. Water Relations of Food Spoilage Microorganisms. Adv. Food Res. 7:83-127. Cited in Christian, 1980, op. cit. Selna, M. and D. Smith. 1976. Pathogen Inactivation During Composting. Technical Services Department, Los Angeles Count Sanitation Districts, Los Angeles, CA. Cited in Haug, 1980, op cit. Shell, B. J. 1955. The Mechanism of Oxygen Transfer Through a Compost Material. Ph.D Dissertation. Sanitary Engineering Dept., Michigan State University, E. Lansing, MI 48824. Sinclair, C. G. and D. N. Ryder. 1975. Models for the Continuous Culture of Micro-organisms under Both Oxygen and Carbon Limiting Conditions. Biotechnology and Bioengineering 17: 375-398. Singley, M. E., A. J. Higgins and M. Frumkin-Roesengaus. 1982. Sludge Composting and Utilization: A Desi n and Operat1n Manual. New Jersey AgricuItuFaI Experiment tation, Cook College, Rut ers--The State University of New Jersey. New Brunsw1ck, NJ. Snell, J. R. 1957. Some Engineering Aspects of High-rate Composting. J. San. Eng. Div., ASCE, Paper 1178. Cited in Haug, 1980, op. cit. Stanek, M. 1972. Micro-organisms Inhabiting Mushroom Compost During Fermentation; Mushroom Sci. 8:797- 811. Stanek, W. 1980. Comparison of Three Methods for Determining the Bulk Density of Peat. Bimonthly Research Notes, Forestry Service, Environment Canada. 36(6):29; 310 Steniford, E. I., D. D. Mara and P. L. Taylor. 1984. Forced Areation Co-composting of Domestic Refuse and Sewage Sludge in Static Piles. In Composting of Agricultural and Other Wastes. Elsevier AppliEd ScienceiPUinshers. London, England. pp. 42-55. Strom, P. F., M. L. Morris and M. S. Finstein. 1980. Leaf Composting Through Appropriate, Low-level, Technology. Compost Science/Land Utilization 21(6):44-48. Suzuki, M. and K. Kumada. 1977. Nitrogen Transformation During the Rotting Process of Rice Straw Compost. Soil Sci. Plant Nutr. 23(2):163-174. Topping, J. 1951. Errors of Observation and Their Treatment. Tfie Insiitute of Physics, London, England. Troller, J. A. and Christian, J. H. B. 1978. Water Activity and Food. Academic Press. New York, NY. Cited in Cfiristian, 1980, op. cit. Waksman, S. A., T. L. Gordon and N. Hulpoi. 1939. Influence of Temperature Upon the Microbiological Population and Decomposition Process in Composts of Stable Manure. Soil Sci. 47:83-113. Ward, R. L. and J. R. Brandon. 1977. Effect of Heat on Pathogenic Organisms Found in Wastewater Sludge. In Composting of Munic a1 Residues and Sludges, ' Pioceedings_3f tfie I977 National Conference. Information T?ansfer,Rockvi11e, Md. pp. 122-127. Wei, J. H. and S. L. Chang. 1975. A Multi-poisson Distribution Model for Treating Disinfection Data. In Disinfection: Water and Wastewater. J.D. Johnson, ed. Ann Arbor SEience PUinShers, Ann Arbor, MI. Cited in Haug, 1980, op. cit. Wells, C. B. 1983. Core Samplers for Soil Profiles. J. Wen, C. Y. 1968. Noncatalytic Heterogeneous Solid Fluid Raction Models. Ind. Eng. Chem. 60(9):34-54. Wen, C. Y. and S. C. Wang. 1970. Thermal and Diffusional Effects on Noncatylitic Solid Gas Reactions. Ind. Eng. Chem. 63(8):30-51. Whang, D. S. and G. F. Meenaghan. 1980. Kinetic Model of Composting Process. Compost Science/Land Utilization 21(3):44-46. 311 Wiley, J. S. 1957. Progress Report on High-rate Composting Studies. In Proceedings gf the 12th Industrial Waste Conference. PurdueUniversity, West Lafayette, IN. Wiley, J. S. and G. W. Pierce. 1955. A Preliminary Study of High Rate Composting. Proceedings ASCE 81:846, American Society of Civil Engineering. New York, NY. Willson, J. F., E. Parr, E. Epstein, P. B. Marsh, R. L. Chaney, D. Colacicco, W. D. Burge, L. J. Sikora, C. F. Tester and S. Hornick. 1980. Manual for Composting Sewage Sludge by the Beltsville Aerated- Pile Method. Municipal Environmental Research Laboratory, Office of Research and Development, US-EPA, Cincinnati, Ohio, 45268. EPA-600-8-80-022. May, 1980. p. 31. Zimmerman, R. E., W. W. H. Chen and A. G. Franklin. 1977. Mathematical Model for Solid Waste Settlement. In Geotechnical Practice for Dis osal of Solid Waste Materials. American Society of CiviI Engineers. New York. pp. 210-226. Ipll I !.|.‘ l. I: ll... nICHIan STnTE UNIV. LIBRARIES (Iimm]mW(WWW,(W("IIIHIIWWI 31293016279816