bug A 1 . . w . 4 17,: Q I}. .. xii? 0‘1 ant»... guard 1 7|:vns‘n‘rl 3’] 1.. a .1. A 1.4.1.5.‘35 . . It. :3. v I. 7131‘ .5... dhrunif‘. : .2141. .3, 11‘! . LIBRARY Michigan State University This is to certify that the dissertation entitled START UP OF ANAEROBIC BIOREACTOR LANDFILLS IN COLD CLIMATES BY INTERMITTENT AIR INJECTION presented by Reem Raji Musleh has been accepted towards fulfillment of the requirements for the Ph.D degree in Environmental Engineering Major Professor’s Signature “[3012009? Date MSU is an Affinnative Action/Equal Opportunity Institution ~—-o---o-o-n--n--o-o-a----«-n- - _«-.-._._.—.--.--.-u--n--o-n-.-a-.-n---.au_.g.—u-n-a-n-n-o--—.—.:— PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:/ClRC/DateDue.indd-p.1 START UP OF ANAEROBIC BIOREACTOR LANDFILLS IN COLD CLIMATES BY INTERMITTENT AIR INJECTION By Reem Raji Musleh A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Civil and Environmental Engineering 2005 ABSTRACT START UP OF ANAEROBIC BIOREACTOR LANDFILLS IN COLD CLIMATES BY INTERMITTENT AIR INJECTION By Reem Raji Musleh Intermittent air injection may be one of the most economical solution to enhance waste temperature without toxicity to methanogenesis. Waste placed in landfill at lower temperatures may remain at lower temperature, leading to slow startup of anaerobic digestion process in biodegradation in landfills. In this study, waste placed at ambient temperature of 3.2i0.3 °C (range between -30°C to +30°C), was subjected to leachate injection to increase waste moisture content. This has resulted in waste temperature of 10.4i1 .3°C. After leachate injection, the was subjected to intermittent horizontal air injection increasing the temperature by 10.2:2.6°C over the entire cell using 2.5hp compressor. The objectives of this study were to quantify the relationship between amount of heat consumed and amount of heat production (known as heat generation factor), to determine Point oxygen consumption rates (i.e., that is oxygen consumption rate at particular location in the landfill not normalized to the waste mass), and to evaluate the impact of air injection on methanogenesis process. The average heat generation factor under field conditions was 341:84 kJ/mole, using dynamic calorimetric method. First order POCR averaged 00867100422 h-l after 15 days of air injection and increased to O.2450i0.0843 h-l after 21 days of air injection. This study demonstrated that air injection is key to a successful startup of anaerobic digestion of bioreactor landfills in colder climates. The average rate of increase in percent methane at all locations exposed to air in the landfill increased significantly from 0.15:0.042 °C day-1 to O.68:t0.15°C day-1. This was accompanied by a decrease in the median of hydrogen concentration from 4,085 ppm to 953 ppm. The lag time also decreased from 137 days to 15 days, in response to air injection. Cepyright by REEM RAJI MUSLEH 2005 DEDICATION To my family ACKNOWLEDGEMENTS I want to convey my thanks to my advisor Dr. Syed Hashsham and to Dr. Xianda Zhao for the guidance and support they provided through this research. I also convey my gratitude to my committee members Dr. Thomas C. Voice, Dr. Susan Masten, and Dr. Stephen Boyd. This project was supported by grants from the Environmental Research and Education Foundation with funding from the Department of Energy (under special project number DE_FG36-016011096) and from Waste Management, Inc. I am especially thanka to Dr. Ed Repa, whose initiative and continuous support was critical for this project. I would like to extend my gratitude for the staff at Northern Oaks Waste Recycling Facility at Harrison, M1 for their cooperation. I would also like to acknowledge the support provided by the Fullbright Fellowship in support of my graduate studies. It is also fair to mention the name of my colleague Irfan Aslam, with whom I could discuss my research problems and get a meaningful critique and ideas to work on. I feel obliged to mention here my uncle Tank and his family as I was privileged to enjoy their love, company and continuous support since my first day at MSU. I want to thank my parents, and my sister for their continuous motivation and support that helped me throughout my studies. Special thanks to all my friends at MSU who were always there for me: Rana, J ameel, Irfan and Dominique. vi TABLE OF CONTENTS List of Tables - ix List of Figures- - xii 1. INTRODUCTION 1 1.1. Temperature .................................................................................................. 2 1.2. Heat Generation Factor ................................................................................. 3 1.3. Oxygen Consumption Rate ........................................................................... 3 1.4. Impact of Air Injection on Methanogenesis .................................................. 4 1.5. Heterogeneity ................................................................................................ 5 1.6. Objectives ..................................................................................................... 8 2. LITERATURE REVIEW - 10 2.1 . Temperature .................................................................................................. 10 2.2. Enhanced air diffusion .................................................................................. 10 2.2.1 Air injection .......................................................................................... 12 2.2.2. Radius of influence .............................................................................. 14 2.3. Heat Generation Factor ................................................................................. 14 2.4. Oxygen Consumption Rate ........................................................................... 17 2.4.1 Effect of process variables on aerobic degradation .............................. 18 2.5. Anaerobic Digestion of Solid Waste ............................................................. 20 2.5.1 Effect of temperature on anaerobic digestion ....................................... 22 2.5.2. Inhibition of methanogenesis due to low pH and high VFA ............... 23 2.5.3. Inhibition of methanogenesis due to oxygen ....................................... 24 2.5.4. Hydrogen and anaerobic digestion process .......................................... 24 3. SITE DESCRIPTION AND METHODS 27 3.1 Site Description and Bioreactor Landfill Cell Construction .......................... 27 3.2. Monitoring Devices ...................................................................................... 28 3.2.1. Automated and onsite measurements ................................................... 33 3.2.2. Gas and leachate sampling protocol for off-line analysis .................... 34 3.3. Analytical Methods ....................................................................................... 34 3.3.1. Hydrogen .............................................................................................. 35 3.3.2. Oxygen, nitrogen, carbon dioxide and methane .................................. 35 3.3.3. Electrical conductivity, pH, COD, VOC ............................................. 35 3.4. Experimental Procedures .............................................................................. 36 3.4.1.. Performance of intermittent air injection system (Objective 1) ........... 38 3.4.2. Oxygen consumption rate (Objective 2) .............................................. 41 3.4.3. Evaluation of anaerobic digestion process after startup or air perturbation (Objective 3) .................................................................................................. 41 vii 3.5. Statistical Methods ........................................................................................ 42 3.5.1. Normality assumption .......................................................................... 43 3.5.2. Test difference between two independent samples .............................. 45 3.5.3. Test difference between two dependent samples ................................. 47 3.5.4. Tests the linearity of a relationship between two variables ................. 49 3.6. Method to determine heat generation factor ................................................. 51 3.7. Method to determine oxygen consumption rate ............................................ 57 3.7.1. Oxygen concentration vs. time curve ................................................... 58 3.7.2. Calculation of liquid phase oxygen concentration ............................... 59 3.7.3. Calculation of kinetic coefficients for POCR using Monod equation .62 4. RESULTS AND DISCUSSION - - 70 4.1. Evaluation of Intermittent Air Injection in Northern Oaks Bioreactor Landfill .............................................................................................................................. 7 1 4.1.1. Establishment of base line temperature before air injection ................ 71 4.1.2. Effect of air injection on waste temperature ........................................ 79 4.1.3. Effect of moisture on rate of temperature increase .............................. 90 4.1.4. Temperature response after air injection tests ..................................... 93 4.1.5. Heat Generation Factor ........................................................................ 97 4.1.6. Change in oxygen, nitrogen, carbon dioxide and methane concentration due to air injection ......................................................................................... 109 4.1.7. Impact of air injection on leachate characteristics .............................. 111 4.2. Oxygen Consumption Rate ........................................................................... 119 4.2.1. Operational conditions ......................................................................... 120 4.2.2. Fitting POCR data to Monod equation ................................................ 122 4.2.3. Fitting first order kinetics to POCR data ............................................. 126 4.2.4. Factors affecting POCR ....................................................................... 129 4.2.5. Comparison between POCR and OCR from the landfill ..................... 133 4.3. Evaluation of methanogensis ........................................................................ 136 4.3.1. Methane generation and hydrogen accumulation during startup ......... 136 4.3.2. Methane generation and hydrogen accumulation after stopping air injection ............................................................................................................. 149 4.3.3. Factors affecting methanogensis establishment during bioreactor landfill startup ............................................................................................................. 159 4.3.4. Factors affecting methanogenesis establishment afier air injection was stopped ........................................................................................................... 164 4.3.5. Comparison of methanogensis establishment between startup and after air injection .......................................................................................................... 167 5. CONCLUSIONS - - - -- -- 169 6. ENGINEERING SIGNIFICANCE - 172 7. REFERENCES - -- 175 viii LIST OF TABLES Table 1-1: Temperature measured in the landfill on 5/9/03 ..................................... 7 Table 2-1: Air flow rate and temperature change as reported in aerobic bioreactor landfills ...................................................................................................................... 13 Table 2-2: OCR reported for solid waste ................................................................... 20 Table 3-1: Chronological order of filling the Northern Oaks Bioreactor Landfill ....29 Table 3-2: Air injection test 1 (lower lift) — Percent flow in each injection line ....... 40 Table 3-3: Air injection test 2 (upper 1ifis)—injected air volume distribution among the injection lines ............................................................................................................. 40 Table 3-4: Comparison between parametric and nonparametric tests ....................... 43 Table 3-5: Calculation of the standard error of the difference and the degree of freedom for the t-test under homogeneous and non-homogeneous variances ......................... 46 Table 4-1: Kruskal-Wallis Test for testing differences between lifts waste lifts with respect to a) ambient temperatures to which waste was exposed and b)waste temperature .................................................................................................................................... 73 Table 4-2: AN OVA test for waste temperature differences for lifts 1 through 4 after leachate recirculation (9/5/03) ................................................................................... 77 Table 4-3: Operational conditions for each air injection event .................................. 80 Table 4-4: Rate of temperature change for air injection events in each lift ............... 81 Table 4-5 Air injection events for which the heat balance was conducted and their operational parameters ............................................................................................... 98 Table 4-6: Calculation results of input parameters that varied for each air injection .................................................................................................................................... 99 Table 4-7: Input parameters for the calculation of heat balance equation ................. 100 Table 4-8: Heat loss, production and accumulation in the bioreactor landfill ........... 105 Table 4-9: Average heat generation factor ............................................................... 106 ix Table 4-10: Comparison between laboratory values for heat generation factor in laboratory (single substrates) and field scale (solid waste) ....................................... 106 Table 4-11: Total heat loss and gain and the percent contribution of each term of the heat balance equation ......................................................................................................... 108 Table 4—12: Average concentrations of VOC in lift 1 ................................................ 113 Table 4-13: Henry’s constant for VOCs found in leachate samples from Northern Oaks bioreactor landfill ...................................................................................................... 1 18 Table 4-14 Ratio of RL/RT VOCs found in leachate samples from Northern Oaks bioreactor landfill for kg/kl=100 .............................................................................. 119 Table 4-15: Sampling locations, temperature, moisture content and oxygen concentration before stopping air injection at which POCR was conducted (Test 1) ...................... 121 Table 4-16: Sampling locations, temperature, moisture content and oxygen concentration before stopping air injection at which POCR was conducted (Test 2) ...................... 122 Table 4-17: Comparison between the range of data used and the required to estimate Monod kinetics parameters ........................................................................................ 124 Table 4-18: First Order POCR ................................................................................... 127 Table 4-19: Tests of Normality for POCR, temperature, moisture content and background oxygen concentration ................................................................................................. 128 Table 4-20: Wilcoxon Signed Ranks Test for comparison between POCR tests 1 and 2 .................................................................................................................................... 129 Table 4-21: Paired samples t-test for comparison between tests 1 and 2 .................................................................................................................................... 131 Table 4-22: pH, COD, and temperature at sampling locations during landfill startup .................................................................................................................................... 139 Table 4-23: Hydrogen during and after lag time, initial and final methane concentration, rate of increase in percent methane and lag time at sampling locations during landfill startup ......................................................................................................................... 141 Table 4-24: ANN OVA test to determine differences between the different lifts in rate of increase in percent methane .................................................................................................................................... 143 Table 4-25 Normality tests for methane concentration, hydrogen concentration and rate of increase in percent methane .................................................................................................................................... 144 Table 4-26: t-test to test the difference between upper and lower lifts with respect to the rate of increase in methane percent .................................................................................................................................... 144 Table 4—27: Wilcoxon Signed Ranks Test to study the difference in hydrogen concentration between the two periods: before and lag time .................................................................................................................................... 146 Table 4-28: Kruskal Wallis Test to study the difference in hydrogen concentrations between the different lifts .................................................................................................................................... 147 Table 4-29: Wilcoxon Signed Ranks Test to test differences between hydrogen concentration during and after lag time for each lift .................................................................................................................................... 148 Table 4-30: Waste temperature, hydrogen concentration, lag time, final methane concentration and rate of increase in percent methane for locations exposed to air .................................................................................................................................... 151 Table 4-31: Normality tests for concentration of methane, final methane concentration, hydrogen concentration and rate of increase in percent methane in the landfill as a whole .................................................................................................................................... 152 Table 4-32: Tests of normality per lift for rate of increase in percent methane, methane concentration and hydrogen concentration ................................................................ 153 Table 4-33: ANN OVA table to test the difference in final methane concentration in between lifts 1, 2 and 3 and in rate of increase in percent methane for lifts 1, 2, 3 and 4 .................................................................................................................................... 154 Table 4-34: Test statistics for Wilcoxon Signed Ranks Test .................................................................................................................................... 156 Table 4-35: Kruskal Wallis Test to test difference between hydrogen concentrations among the different lifts ............................................................................................ 157 Table 4-36: Wilcoxon Signed Ranks Test to study the difference in hydrogen concentration between before and after lag time ....................................................... 157 xi LIST OF FIGURES Figure 1-1: Ambient Temperature during waste placement ......................................... 7 Figure 3-1: Bioreactor landfill cell at Northern Oaks Recycle and Disposal Facility. A. The layout of the bioreactor landfill cell, B. View of the cell from the north during filling, C. View of the cell from north after filling was completed. ....................................... 28 Figure 3-2 Layout of bioreactor landfill at NORDF .................................................... 30 Figure 3-3: Gas sampling port ..................................................................................... 31 Figure 3-4: Leachate sampling port ............................................................................. 32 Figure 3-5: Automata TDR moisture probe and hand held reader .............................. 32 Figure 3-6: Gro.point TDT moisture probe ............................................................... 33 Figure 3-7: Relative location of the weather station on site (This picture was taken from east side of the bioreactor landfill) ............................................................................... 33 Figure 3-8 Overview of experiments timeline and sampling events ........................... 37 Figure 3-9: Gas concentration change after stopping air injection location G143 ...... 58 Figure 3-10: A hypothetical example for determination of the initial rate for Lineweaver- Burk plot ...................................................................................................................... 68 Figure 4-1: Average waste temperature vs. Average ambient temperature waste is exposed to in each lift .............................................................................................. . ....72 Figure 4-2: Initial waste temperature in the landfill .................................................... 74 Figure 4-3: Gas composition alter waste coverage ...................................................... 76 Figure 44 Average waste temperature in each lift and air temperature during leachate recirculation ................................................................................................................. 78 Figure 4-5: Average waste temperature in five lifts in Northern Oaks bioreactor landfill ...................................................................................................................................... 82 xii Figure 4-6: Temperature increase in waste during air injection in L14 for lift2 ........ 83 Figure 4-7: Temperature change during air injection test 1 ......................................... 84 Figure 4-8: Rate of waste temperature increase during test 1 ...................................... 85 Figure 4-9: Temperature change during air injection test 2 ......................................... 87 Figure 4-10: Average rate of waste temperature increase during test 2 while injecting air through liftl ................................................................................................................ 88 Figure 4-11: Average rate of waste temperature increase during test 2 while injecting through lifts 3 and 4 .................................................................................................... 89 Figure 4-12: Comparison of the air injection volume between events 1 and 2 .......... 91 Figure 4—13: Comparison of rates of waste temperature increase between air injection testl and test 2 when injecting air through lift 4 .......................................................... 91 Figure 4-14: Comparison of the rate of waste temperature increase between test 1 and test 2 when injecting air through lift 1 ................................................................................ 92 Figure 4-15: Waste and air temperature after stopping second air injection event ..... 93 Figure 4-16: Average temperature in the landfill after stopping air injection tests 1 and 2 ...................................................................................................................................... 94 Figure 4-17: Average temperature in lifts 4 and 5 after stopping air injection test 1 with leachate recirculation periods marked ......................................................................... 95 Figure 4-18: Average temperature in lift 3 after stopping air injection test 1 with leachate recirculation period through lift 4 marked ................................................................... 96 Figure 4-19: Determination of the rate of temperature increase in the bioreactor landfill for air injection event 9 .............................................................................................. 101 Figure 4-20: Change in gas phase concentration during air injection at location G211 .. .................................................................................................................................... 110 Figure 4-21 Change in gas phase concentration during the initial hours of air injection at location G21 1 ............................................................................................................. 1 1 1 xiii Figure 4-22: COD concentration before and after air injection in location lift 1 ...... 113 Figure 4-23: COD concentrations decrease during air injection in lift 3 ................... 114 Figure 4-24: COD concentration vs. pH before and after air injection in sampling location W111 — Lift 1 ............................................................................................................. 114 Figure 4-25: Acetone concentrations before air injection in lift 1 (9/2/2003) ........... 117 Figure 4- 26: Acetone concentrations after air injection in lift 1 (10/15/2004) ......... l 17 Figure 4-27 Fitting Monod equation to the experimental data for location G221 ..... 123 Figure 4-28 Monod equation based on the estimated parameters for location G221, T = 50°C,y=1.15 .............................................................................................................. 125 Figure 4-29: First order POCR for location G221 ..................................................... 128 Figure 4-30: First order POCR for two tests separated by 7 days of air injection ..... 129 Figure 4-31: Temperature during the two tests separated by 7 days of air inj ection..130 Figure 4-32: Volumetric moisture content during the two tests separated by 7 days of air injection ...................................................................................................................... 131 Figure 4-33: Oxygen concentration before turning off the air injection .................... 132 Figure 4-34: POCR Difference between of tests 1 and 2 vs. their temperature difference .................................................................................................................................... 133 Figure 4-35: Oxygen consumption rate at vent L6 (Date: 10/24/03) ......................... 134 Figure 4-36: Comparison between POCR and oxygen consumption rate in vents during test 1 ........................................................................................................................... 135 Figure 4-37: Comparison between POCR and oxygen consumption rate in vents during test 2 ........................................................................................................................... 135 Figure 4-38: Methane concentration as a function of time during startup in a sampling location in lift 2 .......................................................................................................... 137 xiv Figure 4-39: Methane concentration as a fimction of time during startup in a sampling location (G134) in lift 1 ............................................................................................. 138 Figure 4-40: Distribution of the rate of methane generation in Northern Oaks bioreactor landfill during startup ................................................................................................. 139 Figure 4-41: Rate of increase in methane concentration (%lday) for each lift ......... 143 Figure 4-42: Methane concentration during start up of bioreactor landfill lift 4 ....... 145 Figure 4-43: Hydrogen concentration during and after lag time .............................. 147 Figure 4-44: Hydrogen concentration during start up of bioreactor landfill lift 4 ..... 148 Figure 4-45: Rate of increase in methane concentration (%lday) for each lift ......... 150 Figure 4—46: Methane regeneration after air injection was stopped in lift 2 .............. 155 Figure 4-47: Hydrogen concentration during and after lag time after air injection .13; Figure 4-48: Hydrogen concentration change after air injection was stopped in lift 2.... .................................................................................................................................... 158 Figure 4-49: Methane concentration and waste temperature during the startup period for lifts 1 and 2 ................................................................................................................. 161 Figure 4-50: Methane concentration and waste temperature during the startup period for lifts 3 and 4 ................................................................................................................. 161 Figure 451: Rate of increase in methane concentration vs. waste temperature during startup period ............................................................................................................. 162 Figure 4-52: COD vs. initial waste temperature in the landfill .................................. 162 Figure 4-53: Methane concentration after 400 days vs. COD ................................... 163 Figure 4-54: Rate of increase in methane concentration vs. pH ................................ 164 Figure 4-55: Rate of increase in methane concentration vs. waste temperature after air injection ...................................................................................................................... 165 XV Figure 4-56: Methane concentration after 62 days from stopping air injection for lifts 1, 2 and 3 and after 40 days from stopping air injection for lift 4 vs. waste temperature ...... .................................................................................................................................... 166 Figure 4-57: Rate of increase in percent methane after air injection vs. before air injection .................................................................................................................................... 166 Figure 4-58: Comparison between the rate of methane concentration increase during startup up and after stopping air injection .................................................................. 167 Figure 4-59: Comparison of the hydrogen concentration during lag time between during startup up and after stopping air injection .................................................................. 168 Figure 4-60: Comparison of the hydrogen concentration after lag time between during startup up and after stopping air injection .................................................................. 168 xvi 1. INTRODUCTION Bioreactor landfills are receiving considerable attention as an alternative to traditional “dry tomb” landfills. This is because bioreactor landfills have several advantages over traditional landfills, including faster methane recovery (in case of anaerobic bioreactor landfills), additional space due to degradation of organic waste and settlement, and lower post-closure maintenance [1]. Based on the electron acceptor conditions, bioreactor landfills can be operated in four different modes: aerobic, anaerobic, nitrifying, and a combination of aerobic/anaerobic processes[2]. However, anaerobic mode of operation is the most common [2, 3]. The first bioreactor landfill was probably tested around 19805 [4]. An accurate count of the number of bioreactor landfills is seldom available due to the disagreement among the practitioners about its definition. According to the United States Environmental Protection Agency (US EPA) definition, “A bioreactor landfill operates to rapidly transform and degrade organic waste. The increase in waste degradation and stabilization is accomplished through the addition of liquid and air to enhance microbial processes [4].” The Solid Waste Association of North America (SWANA) defines a bioreactor landfill as “Any permitted subtitle D landfill or landfill cell where liquid and/or air, in addition to leachate and landfill gas is injected in a controlled fashion into the waste mass in order to accelerate or enhance biostabilization of the waste.” However, according to a survey reported by SWANA in 2002, there were 20 full-scale demonstration bioreactor landfills in the US. [2]. Most field scale anaerobic bioreactor landfills focus on leachate recirculation as a major variable [5-13]. Many other variables that may influence the extent of biodegradation of waste in bioreactor landfills are studied to a lesser extent. These variables include alternative daily and intermediate covers and final caps, gas extraction, compactness, leachate seeps, addition of nutrients, preprocessing of waste, and heterogeneity of waste [2]. Field scale studies focusing on temperature enhancement in bioreactor landfills in cold climates, oxygen consumption and heat generation rates, impact of aerobic start up on methanogenesis, and measures of biostabilization end points have not yet been reported. Heterogeneity which is always a concern with most measurements associated with solid waste studies is also described in the context of the above parameters. The following paragraphs describe some of these factors in more detail. 1.1. Temperature Reported temperature in traditional landfills is either in the mesophilic [14] or therm0philic range [7, 15-17]. Mesophilic refers to a temperature range of 20 to 40 0C [18, 19], and thermophilic refers to temperatures above 45 0C [20, 21]. In cold climate, waste filled during winter may be at temperatures below freezing. At these temperatures, only psychrophilic microorganisms are expected to be active. Psychrophilic microorganisms have optimum temperature below 20 °C [20, 22]. If the temperature is close to zero, even the psychrophilic microorganisms may work at much slower rates. Therefore, low temperature can become a major issue for bioreactor landfills started and/or operated in cold climates. To our knowledge, no field scale bioreactor landfill study has been reported at temperatures below 10 °C. It is obvious that besides leachate recirculation, enhancement in temperature may be required for bioreactor landfills operated in cold climates to obtain higher biodegradation rates. 1.2. Heat Generation Factor As mentioned above, the increase in temperature due to air injection is a function of the amount of heat produced per kg of oxygen consumed within the system and the amount of heat lost fiom the system. The amount of heat produced per kg of oxygen consumed is known as “heat generation factor”. Heat generation factor for bioreactor landfills has not been reported yet. Estimation of this factor combined with the heat loss is critical to quantitatively link the increase in temperature increase to the rate of air supply. 1.3. Oxygen Consumption Rate A cost effective approach to increase the temperature in landfills is through aerobic oxidation of waste. Enhanced air diffusion during waste placement [23] and forced aeration after waste placement are the two main approaches that may enhance waste temperature. Air injection is less intrusive to landfill operations when compared to enhanced air diffusion. Air injection leads to creation of aerobic zones within the landfill ensuing exothermic biodegradation process and increased temperatures. The amount of heat produced per unit mass of organic matter degraded is at least 20-fold higher during aerobic process than the heat produced during anaerobic process [24]. The design of air injection system is still in its infancy. Often a separate pipe network is not necessary. The leachate and gas collection system may serve as part of an air injection system, provided care is taken to insure fire prevention and allow onset of methanogenic activity. Currently, design of air injection system mostly amounts to evaluating the amount of total air that should be provided at a given flow rate. It is evident that the rate of air supply will be a fimction of the oxygen consumption rate (OCR). Oxygen consumption rate may be defined as the decrease in oxygen concentration with time at a given location in a bioreactor landfill. The oxygen consumption rate is a function of temperature and moisture. Oxygen consumption rate expressed per unit mass of solid waste is also known as oxygen uptake rate. Ideally, only the organic fraction of the solid waste should be considered in defining oxygen uptake rate. Data about the rate of oxygen consumption in landfills is currently not available. In composting, however, the effect of temperature on oxygen consumption has been quantified using Arrhenius equation [25, 26] or by empirical relationships [24]. Arrhenius equation only considers temperature but some of the empirical equations also include the effect of moisture [27]. To design an air injection system to enhance temperature in bioreactor landfills, it is essential to quantitatively relate the oxygen consumption rate to temperature and moisture content. 1.4. Impact of Air Injection on Methanogenesis Air injection will obviously impact the anaerobic processes leading to methane production [28]. Methanogens are strict anaerobes and exhibit toxicity to small concentrations of oxygen [29-31]. If intermittent air injection is to be used successfully in enhancing temperature in anaerobic bioreactor landfills, it is important to evaluate the extent of impact on methanogenesis and the time of recovery for methanogenesis after air injection is stopped. Recovery of methanogenesis after air injection is similar to the start up phase of an anaerobic digestion process and air injection is akin to a perturbation to the anaerobic process. Thus some measure of the anaerobic process may be helpful to quantify the rate of recovery of methanogenesis. Besides methane production rate, accumulation of hydrogen is a common indicator of upset anaerobic process. Elevated levels of hydrogen may serve as an indicator of inhibited methanogenesis similar to those reported for anaerobic digesters. Indeed, elevated H2 levels correlated with low pH and decreased rate of methanogenesis in the only study (to our knowledge) reporting hydrogen concentrations in solid waste samples from a landfill [32]. 1.5. Heterogeneity Heterogeneity in waste characteristics is one of the most difficult problems encountered when studying bioreactor landfills. Often, aggregate parameters for the whole landfill are the only measures of its performance. For example, the extent of biodegradation is measured mainly by total gas produced or total leachate generation and its characteristics. Sometimes, variability in the extent of biodegradation in landfills is studied by excavation of the waste followed by measurements in the laboratory for various parameters, such as pH, lignin and cellulose content, biological methane potential, volatile solids, and moisture content. Temporal measurements can not be made at the same location using excavated samples. Alternative strategies, e.g., use of sampling ports may be necessary to enable multiple measurements at the same location at least for parameters that can me measured without destructive sampling. This study will also present unique sampling strategies for leachate and gas sampling from within the landfills to characterize heterogeneity. We have been working to establish a bioreactor landfill at the Northern Oaks Waste Recycling Facility at Harrison, MI since August 2002. The cell construction took one year during which the waste was exposed to temperatures below freezing due to the winter season. During the filling period, ambient temperature ranged between -33°C to +33°C (Figure 1-1) and the bioreactor landfill cell height increased to about 65 feet with approximately 10 feet lifts. At the close of the waste placement, the temperature within the waste was in the range between -5 °Cand 18 °C (Table 1-1) in the landfill. In lift 5 freezing conditions existed. Although ambient temperatures increased to 33°C during summer months, the waste temperature remained low. Table 1-1 summarizes the average waste temperature for each lift in the landfill on 9/5/03 (Day 420 from the start of filling lift 1). Anaerobic digestion at such low temperatures is very slow if any, hence increasing the landfill temperature is critical to enhance biodegradation. Landfill height (ft) Figure 1-1: Ambient Temperature during waste placement Table 1-1: Temperature measured in the landfill on 5/9/03 Duration 0 after Temperature( C) Lift Filling completion of duration filling of the . corresponding Average SD Max Mm lift (months) Jul 02- 1 Aug 02 13.7 12.6 1.1 14.7 10.5 Sep 02- 2 Oct 02 10.7 13.4 2.5 18.0 8.6 Nov 02- 3 Dec 02 9.4 9.5 4.1 16.2 3.7 Dec 02- 4 Jan03 8.4 11.1 3.7 14.7 4.4 Jan 03- 5 Feb 03 6.5 1.6 9.0 14.9 -4.8 1.6. Objectives The goal of this research is to develop an engineered approach to enhance anaerobic biodegradation of solid waste in bioreactor landfills located in cold climates. Intermittent air injection system is the method selected for increasing the temperature. To obtain design parameters needed to operate bioreactor landfills in cold climates, the following specific objectives are set. Objective 1: To establish and quantify relationship between amount of heat produced (kJ) per unit mass of oxygen supplied (kg), and to characterize the performance of air injection with respect to temperature and leachate characteristics. Hypothesis: Heat generation factor determined in laboratory conditions for single substrates may differ from that obtained for solid waste under field conditions. Objective 2: To establish the spatial variability in point oxygen consumption rate (decrease in oxygen concentration with time at a given location in a bioreactor landfill), and study the effect of temperature, oxygen concentration and moisture content on point oxygen consumption rate. Hypothesis: Point oxygen consumption rate is a function of moisture content and temperature. The increase in moisture content will inhibit POCR however the increase in temperature will enhance POCR. Objective 3: To quantify the performance of anaerobic digestion at startup and following air injection with respect to methanogenesis establishment, using methane and hydrogen concentrations as indicators for anaerobic digestion process. Hypothesis: Air injection will affect methanogenic activity negatively however this affect can be altered by improved conditions of enhanced temperatures. Thus air injection will enhance the methanogenic activity after the aeration process is ceased compared to before air injection. 2. LITERATURE REVIEW Although a large number of studies exist on traditional landfills, studies related to bioreactor landfills are relatively few. Hence the following paragraphs extract information related to i) temperature, ii) oxygen uptake rate, iii) heat generation factor, and iv) potential impact of intermittent aeration on methanogenesis in bioreactor landfills from published literature on bioreactor landfills a well as traditional landfills. 2.1. Temperature Temperatures reported in landfills at the time of placement generally range between 10 to 20°C [15 , 16]. This temperature may be lower in colder climates but no published data is available for landfill temperatures in cold climates, especially during initial stages of filling. Because the focus of this project was to enhance the temperature of waste placed in bioreactor landfills in cold climates, studies focusing on the behavior of temperature with time and various operations are reviewed. Two main strategies, namely air diffusion during waste placement and active air injection, are reviewed below for their effectiveness in enhancing landfill temperature through aerobic degradation. 2.2. Enhanced Air diffusion Enhanced air diffusion can be achieved by either increasing the porosity of waste or lengthening the waste placement. Under normal filling operations, a 10-15 °C increase in temperature can be achieved by enhanced air diffusion [33, 34]. Lowering the refuse density and adding wood chips or aerobically stabilized waste are the two main 10 approaches to increasing porosity [35]. Lengthening the exposure time to air was used in one case with approximately 25°C change in temperature [15]. Ambient temperatures do not significantly impact the waste temperature after the landfill is covered. This has been observed in many cases [36, 37] and is due to the elimination of air diffusion and thermal insulation by the waste. Lifts closer to the surface of the landfill generally have lower temperatures compared to lifts that are closer to the base [15, 17, 38, 39]. However, exceptions to this may exist due to seasonal variations during waste filling. Waste filled at higher ambient temperatures will have steady state temperatures that are higher than those filled at lower temperatures. In one case, it was observed that temperature in various lifts may slowly approach a constant value for each lift in the absence of leachate recirculation [37]. This steady state temperature may depend upon the initial temperature at which the waste was filled. For example, at Yolo County bioreactor landfill, CA, lift one or the deeper layer was filled at ambient temperature of around 15°C, which increased to about 25°C during the aerobic phase. Ambient temperature during filling the second and third lifts was about 25°C, which rose to about 50°C immediately after filling due to aerobic oxidation but subsequently decreased to about 31°C [37]. Moreover, it was observed from figures published in [37] that temperatures in waste between consecutive lifts, if different from each other, might stay different for long periods of times. This can be explained by the minimal heat conduction occurring in the waste matrix due to its low thermal conductivity 0.1i0.04 Wm'lK.1 [15]. Such an 11 insulation property is very close to that of asbestos which is known for its ability to insulate heat and has a thermal conductivity is 0.05-0.15 Wm'lK'l [40, 41]. Although temperature increase can be achieved by diffusion, it is not preferred in practice because filling must be continuous; leaving the air space unfilled for air diffusion is not an option. Air injection is a better option to achieve temperature increase because it does not interfere with the landfill operations and does not delay the utilization of air space. 2.2.1. Air injection Air injection is used to change the anaerobic conditions in a landfill into aerobic state. If aeration is adopted at the whole bioreactor landfill for very long periods, the cell will obviously turn into an aerobic bioreactor landfill. Partial aeration of the landfill results in hybrid landfills; in hybrid landfills the upper layers of waste are aerated, while bottom layers are operated under anaerobic conditions [4]. There are few reported aerobic bioreactor landfill demonstration projects [42-47] and two hybrid bioreactor landfills [16, 48]. Air flow rate reported in these bioreactor landfills varies considerably, ranging from 0.2 to 2.9 m3air/m3 waste per day (Table 2-1). Such flow rates are sufficient to cause an increase in temperature (Figure 2-6). The rate of temperature increase ranged from 0.04 °C/day to 1.3 °C/day (Table 2-1). Occasionally, much faster rates of increase have also been observed, e.g., an increase of 20 °C in 5 days was reported due to air injection in a landfill in Korea [46]. This is the fastest rate of increase in temperature reported for landfills. The total increase in temperature can also vary significantly. For example at 12 Live Oak landfill (Georgia), the average temperature reported was 37.8 °C but ‘hot pockets’ of up to 71 °C were also reported [42]. The increase in temperature is generally limited to the duration of aeration. In the hybrid Metro Bioreactor Landfill (Wisconsin), as soon as the air injection ceased, the temperature started to decrease [16]. However, at steady state, temperatures stabilize at higher temperatures compared to those before aeration. This pattern of changes and stabilization at higher temperature is similar to that observed during start up of landfills as a result of air diffusion. Table 2-1: Air flow rate and temperature change as reported in aerobic bioreactor landfills Air 12°"; t Rate if Waste Flow rate m emper a ure Volume Rate Landfill Name air/m3 increase Location ’ Reference 0 (m3 ) (m3/ min) waste day) ( C/day) Live oak 2.9 Not reported 49,000 100 landfill, [45] Georgia Live Oak 0.2 0.3 53,500 6.1 Landfill, [42, 49] Georgia 0 3 Columbia 1.8 ° 45,200 56 county landfill, [43-45] Georgia Williamson 28.3 x 3 = County 0.7 0.04 185,000 84.9 bioreac tor, [47] Texas 4 Not 590,000 reported Seoul, Korea [46] 1.3 Not Not Metro Landfill, [16] reported reported Wisconsin l3 2.2.2. Radius of influence Air injection into landfills is a relatively new practice. Hence, parameters such as the extent of dissemination of oxygen and nitrogen (described by the radius of influence) during air injection through vertical wells or horizontal pipes is rarely estimated. The radius of influence with respect to pressureiof an injection well may be defined as the distance from the injection point at which a pressure change of 0.1 inch of water, or about 10% of the injection pressure is observed [46]. Injection rate and pressure distribution are important for the design of the air injection and distribution systems. As the radius of influence increases a less dense air injection system network is required. The radius of influence with respect to oxygen is reported to be less than that given by pressure changes because oxygen is consumed during dissemination [46]. This radius of oxygen is the distance from the injection line to the point where the oxygen concentration becomes negligible [46], and is an important parameter for designing air injection system because it determines the maximum allowable distance for the placement of pipes used for air injection. The difference between the radius of oxygen and nitrogen is determined by the kinetics of oxygen consumption. A faster rate of oxygen consumption results in bigger difference between pressure and oxygen radii of influence. Both radii are a function of the rate at which air is being injected and characteristics of the waste. Determining this rate under field conditions is critical for designing air injection systems. 2.3. Heat Generation Factor Heat is produced by aerobic biological activity [50]. The amount of heat generated per unit time is known as the heat generation rate. This rate is proportional to the oxygen 14 uptake rate as well as the carbon dioxide production rate and the biomass concentration [50—52]. The proportionality constant linking the heat generation rate to the oxygen uptake rate is called the heat generation factor. The heat generation factor is sometimes also referred to as heat yield on oxygen. To measure the heat generation factor, a heat balance around the bioreactor is conducted using a heat balance, i.e., by stating that accumulation of heat equals the heat generation in the system minus the heat loss from the system. Both heat accumulation and heat loss can be quantified [51, 53-57]. Substituting these values in the heat balance equation and solving the resulting equation yields the heat generation factor. The heat generation factor can be measured in the laboratory either by dynamic or continuous calorimetry [52]. The dynamic method allows the temperature to rise in the bioreactor, hence accumulation of heat takes place with time [51, 53-56]. The slope of curve in which temperature is plotted against time is used to calculate the accumulation of heat. The loss of heat in the system is measured and substituted into the heat balance equation. This equation is then solved to obtain the amount of heat produced at several time points. The slope of the calculated amount of heat produced vs. the oxygen uptake is the heat generation factor. Continuous method maintains a constant temperature by heating or cooling the reactor [54, 55, 57]. In the continuous method, the accumulation of heat is zero because the temperature is constant. Therefore, heat generation equals the heat loss in the system. In laboratory scale reactors, it is feasible to cool and/or heat the system. However it is not feasible to maintain constant temperatures in large-scale or field-scale reactors. Therefore, the dynamic method is the method of choice under field conditions. It is also acceptable because heat generation factors measured by both 15 methods are comparable. Turker [55] obtained a heat generation factor for molasses of 444 kJ/mole Oz and 431 kJ/mole 02 using continuous and dynamic methods, respectively [55]. There is lack of data for heat generation factor for solid waste. However, studies related to heat generation factor for single substrates including glucose, molasses, soybean meal, ethanol, lactose, and acetate are plentiful [24, 51, 52, 55, 56]. Turker reported the range of heat generation factors for the above single substrates in the range of 397 to 543 kJ/mole 02 using laboratory bioreactors [55]. At the average heat generation factor for the above substrates was 456:38 kJ/mole Oz indicating a relatively small variability [55]. In another study, the slope of heat generation rate vs. oxygen consumption rate for different single microbial species grown on different single substrates gave a value of 444 kJ/mole Oz [50]. Values for heat generation for synthetic solid waste are lower (304 kJ/mole Oz [24]) than for single substrates. Only one study considered the use of large scale bioreactors as calorimeters to collect data for heat generation factor [55]. However, in most instances, heat generation factor obtained for single substrates have been routinely used for complex substrates including solid waste. For example, a heat generation factor of 443 kJ/mole 02 was used in composting operations [58] and a value of 460 kJ/mole 02 was used for a landfill [23]. It is evident that determination of heat generation factor for solid waste preferably in situ is critical to obtain a better understanding of temperature distribution in 16 bioreactor landfills and predict the amount of oxygen required to increase landfill temperature. 2.4. Oxygen Consumption Rate Oxygen consumption and /or carbon dioxide production have been used extensively to indicate the status of aerobic degradation [25, 53, 59-72]. The most common way of expressing the oxygen consumption rate is by using the amount of oxygen consumed (weight of oxygen consumed per hour (g Oz/hl') to the mass of solid waste in a control volume (kg)) [72]. The mass of solid waste obviously requires further description about the type of waste. Total organic matter [62] , volatile solids [61], and volume [55] of solid waste have also been used instead of total mass (Table 2-2). The oxygen consumption rate can be determined in the laboratory by respirometry either under dynamic [53, 59-65] or static conditions. However, the determination of the oxygen consumption rate by respirometry requires high levels of oxygen in the sample and outlet gas (above 14%). In bioreactor landfills, the concentration of oxygen in the effluent air in response to air injection might not reach such high levels. Therefore, the respirometric method can not be applied under filed conditions. It is possible, however, to determine the rate of oxygen consumption at a given point in the landfill cell. Determining this rate under field conditions is necessary because rates determined under laboratory conditions are not reflective of the field rates. For example, the OCR reported for full-scale plants were almost twice as much as those determined in a laboratory scale system [73]. Oxygen uptake rates reported in literature vary in the range of 1.0-1.5 mg g'1 hr.1 for an active composting process; above 1.5 mg g'1 hr'l for very active (considered unstable); 17 and up to 3.5 mg g'1 hr.l for extremely high activity [72]. Oxygen consumption rate for landfills have not been determined before, therefore only values for composting of solid waste is reported. 2.4.1. Effect of process variables on aerobic degradation Several studies report correlations between oxygen consumption, carbon dioxide production, and aerobic biodegradation rates with other process variables such as temperature [24, 25, 72, 74-77], dissolved oxygen concentration /gas phase oxygen concentration [25, 78], maturation of waste [62, 73], air filled porosity [79, 80], and moisture content [72, 81]. The influence of these parameters on oxygen uptake rate varies. Kulcu [27] summarized that a minimum concentration of 5% in the pore space is required for composting piles. Temperature is an important factor for oxygen uptake, however it less influential than moisture [72]. A lag phase occurs at the onset of aerobic degradation regardless of moisture content. This lag phase decreases as the temperature increases [72]. OCR depends on the waste moisture content [72]. Moisture contents of 50, 60, and 70 % did not significantly impact the OCR. However, moisture contents of 30, 40, and ~50% significantly impacted the OCR. A minimum moisture content of 50 % was required for active composting, with no upper limit in the moisture content at the ranges studied (30-70 %) [72]. Similarly, air filled porosity is also known to affect microbial kinetics. Maximum OCR were observed at air filled porosity of 25-30 % [80]. As described above, several factors affect the oxygen consumption rate. However the effect of temperature is most commonly studied. In the composting literature, it is 18 observed that temperatures below 20 °C severely slowed down the composting process [72]. The decomposition rate in compositing operations are also reported to slow down under high temperature conditions. Temperatures above 60 °C were reported to reduce the microbial activity and these above 82 oC severely hindered the process [72]. The highest OCR in composting is at temperature range of 52 to 60 °C [72]. Within this optimum composting temperature range, the oxygen uptake rate increased with an increase in temperature. Several empirical and mechanistic equations are available to quantify the effect of temperature on oxygen uptake rate. Among the mechanistic equations, the Arrhenius equation is the most common [25, 26]. It can be applied to the individual temperature ranges in which microbial activity takes place (i.e., psychrophilic, mesophilic, and thermophilic) or to the entire range [26]. Often empirical equations were found more appropriate as they can better describe the effect of temperature [24]. 19 Table 2-2: OCR reported for solid waste. OCR* Process Reference 0.11-4.73( g 02/ (kg TOM Organic fraction of municipal solid waste [62] hr)) (OFMSW) composting operation 0.130-0.930 (g Oz/(kg vs [61] hr» 10'5-10'3 (mol oz /( m3 sec)) Solid waste laboratory column fed by [15] 0806-8064 (11 (m3 hr)) diffusion only 0.0036036 (mol/ (m3 hr)) Mechanical-biological end composting 0500 (g 02 / (kg VS hr)) process (pilot scale) -Active phase [73] Mechanical-biological end composting 0280 (g 02 /( kg VS hr)) process (full- scale) — Active phase [73] 1-0 (mg/(8110) [72] ‘Note that OCR is reported in the original units which may be different for each reference. 2.5. Anaerobic Digestion of Solid Waste Anaerobic digestion is characterized by greater degree of metabolic specialization in comparison to that of aerobic degradation [28]. The first step in the anaerobic digestion of complex waste is hydrolysis in which particulates substrates such as lipids, polysaccharides, proteins and fats are hydrolyzed and fermented into soluble organic materials [82-86]. The second step is the conversion of these soluble compounds by acidogens into volatile fatty acids (propionate, butyrate etc.), carbon dioxide, and hydrogen [85]. The third step is the conversion of these fatty acids by acetogens into acetic acid and hydrogen. The last step of anaerobic digestion is methanogenesis, in 20 which hydrogen and carbon dioxide are converted to methane by COz-reducing methanogens or acetate is cleaved to methane and C02 by aceticlastic methanogens. Approximately 75 % of the methane produced is a result of acetate conversion into methane and carbon dioxide [28, 86, 87]. Hydrogen utilizers are responsible for producing the rest of the methane [86, 88, 89] In line with the anaerobic digestion process, anaerobic decomposition of solid waste is marked by an increase in chemical oxygen demand (COD) and volatile fatty acids (VFA), and a decrease in pH [90-93]. This stage is usually followed by an exponential increase in methane generation, pH increase, and a decrease in VFA and COD. At the end, methane production starts to decrease and COD stabilizes at low concentrations. The only difference under field conditions may be in the rates at which the above processes occur [94]. The start-up of anaerobic digestion is considered the most critical step in the operation of anaerobic digester [95]. It is characterized by low abundance of methanogens and acetogens and high abundance of acidogens growing at a fast rate [96]. This imbalance leads to the accumulation of hydrogen and VFA (and therefore low pH). These intermediates further inhibit acetogens and methanogens. Methanogenesis may further be affected by lower temperatures, leading to a very slow start up of the process under cold climates [97]. 21 2.5.1. Effect of temperature on anaerobic digestion Mesophilic operation of anaerobic digesters is most common followed by the thermophilic process. Landfill temperatures above 30 °C have been reported to be favorable for the anaerobic digestion of solid waste [35]. The optimum temperature range for refuse degradation in the laboratory was found to be 30-35 °C [35]. Engineered anaerobic digesters are probably nonexistent under psychrophilic conditions (10 to 20 °C). Microbial degradation rates are expected to be very slow under psychrophilic conditions. When the temperature decreases from mesophilic to psychrophilic range, methane production decreases [98]. This is a known to be a problem during winter [83]. The start up of anaerobic reactors at low temperatures (15 °C and 25 °C) is problematic even under laboratory conditions [97]. Besides affecting the startup of a bioreactor, psychrophilic temperatures also impact the structure of the microbial community. For example under psychrophilic conditions, acetoclastic methanogens are reported to be the main microorganisms for methane production [99, 100]. Hydrogen consumers are hindered more at such temperatures leading to accumulation of hydrogen. Homoacetate bacteria strongly compete with methanogens and might be considered “the most typical participants of the anaerobic decomposition community in psychrophilic conditions” [97]. Thermophilic anaerobic digestion process has shorter startup time and higher gas production rates compared to the mesophilic process [95]. Growth rates of methanogens under thermophilic conditions are higher than those observed under mesophilic 22 conditions. As the temperature increases to the hyper-thermophilic range (i.e., above 80 °C), hydrogen consuming methanogens become more active, often exceeding the rate of hydrogen producing syntrophs [101]. The optimal temperatures for methane generation from acetate are 55 °C, for hydrolytic and fermentative bacteria are 55 to 70 °C, for hydrogen utilizing methanogens 65 oC, and for acetate, forrnate and butyrate consuming populations 60 oC. Thus, in thermophilic anaerobic digestion process, different groups of microorganisms require different optimal temperature, complicating the optimal performance [82]. 2.5.2. Inhibition of methanogenesis due to low pH and high VFA The pH in an anaerobic digestion process is an indicator of imbalance and stability in anaerobic digestion system. The vulnerability of reactors to pH changes is due to variability in buffering capacity (V FA and bicarbonate) [98]. High VFA concentrations are associated with low pH values. The methanogens are inhibited by the unionized acid concentration, leading to a decrease in methane production [88, 102, 103]. Inhibition during the anaerobic digestion of solid waste due to VFA is well documented [91]. However, the concentration of VFA at which inhibition occurs in solid waste is much higher compared to anaerobic digesters [91]. Optimum pH for hydrolysis is slightly acidic (~pH 5) and inhibition occurs at pH of 4.0 [85]. Optimal pH for methanogens is between 7 and 7.2. Inhibition occurs at pH values below 6.2 or above 7.8 [20, 87, 104]. Butyrate degradation is optimum in the pH range of 6.8 to 7.3, with inhibition at pH 6.0 to 6.4 [105]. 23 2.5.3. Inhibition of methanogenesis due to oxygen Methanogens are strict anaerobes, hence exposure to air is detrimental to the anaerobic process [29]. Inhibition constants (the concentration at which 50% inhibition occur) for oxygen are at very low concentrations 0.0001-0.0005 gfir L'1 [31]. However, it is also reported that granular sludge survived up to 18 days when exposed to air and substrate was available [29]. High dissolved oxygen concentration (8.1 ppm) in the influent wastewater was inhibitory to acetogens and methanogens [29]. A slight decrease on acetate and propionate utilization rate was found at 8.1 mg L'1 dissolved oxygen when operated for 2 months. This rate declined further over prolonged operation. The structure of the biofilm was able to protect the strict anaerobes from oxygen toxicity when exposed for one month to oxygen. The tolerance of methanogens to oxygen was attributed to their presence in structured granules, protected by facultative bacteria and aerobes at the outer parts limiting the diffusion of oxygen. Longer aeration periods leads to the destruction of these aggregates, leading to easier diffusion of oxygen and penetration of oxygen to deeper layers [29]. 2.5.4. Hydrogen and anaerobic digestion process Hydrogen is produced (together with acetate) by fermenters and syntrophs as a by product during anaerobic digestion process. Acetate is produced by three paths: Approximately 30% of the acetate is produced by the fermenters. Sixty percent of the total acetate is produced by the conversion of butyrate, propionate, valerate, isovalerate, and isobutyrate to acetate by acetogens (syntrophs). The rest is formed from forrnate, hydrogen and carbon dioxide by the homacetogenic bacteria [106]. Acetate production 24 by syntrophs is inhibited by elevated levels of hydrogen concentrations. A theoretical value of 100 ppm of hydrogen in the gas phase is shown to thermodynamically inhibit the syntrophs [92]. In addition methanogens that utilize acetate at thermophilic conditions such as Methanosarcina therrnophila TM-l are also inhibited by hydrogen concentrations above 250 Pa (~2500 ppm) [107]. The percent inhibition at 76, 101, 253, 507, 1010 and 2030 Pa is 0%, 5% 13%, 31%, 82%, and 100%, respectively. This pressure inhibited syntrophic butyrate degrading cultures, but at lower level than that of acetate utilizing methanogens [107]. Syntrophs particularly butyrate degraders at hydrogen pressure of 76, 101, 253, 507, 1010, 2030 and 3040 Pa were inhibited by 0 %, 3 %, l7 %, 32 %, 52 %, 75 %, and 100 %, respectively. It seems at lower hydrogen concentrations (up to 507 Pa) the inhibition of syntrophs is slightly worse than those of acetate utilizing methanogens, however as the hydrogen pressure further increases syntrophs are less inhibited. These values for inhibition of syntrophs are at much higher levels than 100 ppm. This implies that the methanogens that utilize acetate, known to produce more than 70 % of methane in anaerobic digestion, are more affected than syntrophs, indicating that the anaerobic digestion under such conditions will further be inhibited, due to acetate accumulation and reduction in pH. The level of inhibition by hydrogen has been reported to vary significantly between suspended and attached grth anaerobic digestion. A low dissolved hydrogen concentration in anaerobic digestion is an indication of a well functioning anaerobic process. Digesters with upsets experience elevated levels of hydrogen [86]. It is reported to be a useful indicator of process upsets because the effect on hydrogen is more rapid than on methane [93, 104]. Others state that hydrogen is not a good indicator because of 25 mass transfer limitations [101, 107] and fast dynamic behavior [97, 108]. Its relevance as an indicator is less documented in attached growth systems. Some researchers believe that aggregates and biofilrns provide an efficient mechanism for hydrogen delivery and its removal [101]. In granular sludge microcolonies have been observed between syntrophs such as propionate or butyrate degraders and hydrogen utilizing methanogens. Destruction of these aggregates lead to a decreased rate of activity for the syntrophic microorganisms [102]. The proposed structure of methanogens and syntrophs in the aggregate includes acidogenic bacteria in the outer layer and hydrogen utilizing bacteria in the inner layer. Within such structures tolerance to inhibitory concentrations of substrates is quite likely [102]. 26 3. SITE DESCRIPTION AND METHODS 3.1. Site Description and Bioreactor Landfill Cell Construction The bioreactor landfill is located at Northern Oaks landfill and Recycling Facility- Harrison, Michigan which opened in December 1992. The facility, occupies 320 acres, of which 76 acres is designated for active use. Nearly 75 % of the leachate collected on site is recirculated either to the bioreactor landfill cell or to the working face of other cells [109]. The bioreactor landfill cell (the subject matter of this study) is 60 feet deep, has 1.2 acres footprint and contains 71,000 cubic yards of municipal solid waste (Figure 3-1). This cell was filled between 7/12/2002 and 3/4/2003 and final cap in place by 6/25/2003. The total volume of cover soil used was 17,000 yd3 which is approximately 19 % of the landfill volume. The cell includes horizontal leachate recirculation trenches, horizontal leachate recirculation lines placed on geocomposite, horizontal gas extraction lines, and settlement profilers. The landfill cell was constructed in six 10-ft lifts referred to as lifts l to 6, starting from the bottom. At the top of lift 1 leachate recirculation lines, gas extraction lines, sampling ports, and probes were placed. The leachate distribution lines and ports were placed at the top of lifts 1 to 5. Gas extraction lines were placed in the middle of lifts 3, 4, 5, and 6, i.e., at 25, 35, 45, and 55 ft from the bottom, and at the top of lifts 1 and 2 at 10 and 20 ft from the bottom. The waste was covered daily with soil according to the regular operation of the landfill at Northern Oaks. The gas extraction lines were connected to a flare through a gas header. Leachate from the bioreactor cell was collected through drainage system constructed above the liner and connected to a sump. 27 The sump was equipped with a level controlled pump to meet the regulatory requirements of 1 ft of leachate head above the liner. Table 3-1 lists the events of construction and placement of monitoring devices at the bioreactor cell in a chronological order. Figure 3-1: Bioreactor landfill cell at Northern Oaks Recycle and Disposal Facility. A. The layout of the bioreactor landfill cell, B. View of the cell from the north during filling, C. View of the cell from north after filling was completed. 3.2. Monitoring Devices Monitoring devices at the bioreactor landfill site included leachate and gas sampling ports, moisture and temperature probes, weather monitoring station, and settlement profilers. A total of 48 ports for gas and leachate sampling were placed within the landfill in a three-dimensional grid. Moisture and temperature probes were installed next to the sampling ports. The distribution of these monitoring locations was as follows: 19 in lift 1, 13 in lift 2, 6 each in lift 3 and 4, and 4 in lift 5 (Figure 3-2). A perforated 1 inch diameter and 3 inch long PVC pipe connected to a long tubing extended outside the cell to be used as a gas sampling port (Figure 3-3). The leachate sampling port consisted of a 14 L plastic bucket, with a perforated manifold that was placed inside the 28 Table 3-1: Chronological order of filling the Northern Oaks Bioreactor Landfill Start End Description Jate Day“ Date Day flZ/ZOOZ 0 8/1/2002 20 Fill Lift 1 fi4/2002 12 7/24/2002 12 Install weather station \8/24/2002 40 9/13/2002 63 Install monitoring devices on Lift 1 M/ZOOZ 82 10/8/2002 88 Install leachate recirculation and gas extraction lines on Lift 1 M002 89 10/28/2002 108 Fill Lift 2 M2002 111 11/8/2002 119 Install monitoring devices on Lift 2 M002 122 11/11/2002 122 Install leachate recirculation lines on Lift 2 M2002 123 11/22/2002 133 Fill Lift 3 (First 5 ft) M002 136 11/25/2002 136 Install gas extraction lines on Lift 2.5 11/2 6/2002 137 12/6/2002 147 Fill Lift 3 (Second 5 ft) 12/ l 4/ 2002 155 12/16/2002 157 Install monitoring devices on Lift 3 12/ l 7/ 2002 158 12/17/2002 158 Install leachate recirculation lines on Lift 3 12/1 7/2002 158 12/26/2002 167 Fill Lift 4 (First 5 ft) 12/27/ 2002 168 12/27/2002 168 Install gas extraction lines on Lift 3.5 12/27/2002 168 1/7/2003 179 Fill Lift 4 (Second 5 ft) 1/24/ 2003 196 1/26/2003 198 Install monitoring devices on Lift 4 _1/2\8/‘2003 200 1/28/2003 200 Install leachate recirculation lines on Lift 4 $842003 200 2/4/2003 207 Fill Lift 5 (First 5 a) £003 208 2/5/2003 208 Install gas extraction lines on Lift 4.5 M03 208 2/13/2003 216 Fill Lift 5 (Second 5 ft) _2/1\3/2003 216 2/14/2003 217 Install monitoring devices on Lift 5 -2/1\4/2003 217 2/14/2003 217 Install leachate recirculation lines on Lift 5 M003 220 2/19/2003 222 Fill Lift 6 (First 5 ft) £2003 223 2/20/2003 223 Install gas extraction lines on Lift 5.5 #244003 227 2/26/2003 229 Fill Lift 6 (Second 5 ft) £10003 230 2/27/2003 230 Install leachate recirculation lines on Lift 6 _3/{/2003 234 3/4/2003 235 Fill Final Lift Mom 306 5/14/2003 306 Install gas ports and TDT in cap below geomembrane Mom 339 6/20/2003 343 Install geomembrane cap __6/1&2003 341 6/25/2003 348 Covering geomembrane with soil £2003 396 8/12/2003 396 Install gas ports and TDT in cap above geomembrane MIL/2003 431 9/16/2003 431 Install three flux chambers 29 m nu B n .m m m m. S elvO mp, u T, ///flw ;’ [l // // s l 2 3 4 5 d r m fi fi fi fi h n r 0 on .I :I .1 :l .1 e 0 m a L L L L L m... e m .m .m m m .m S o A x n_ x R 2 t sass; L new...” m I $53.. . Essex... L I n $Nw$ $3 .3. . 4. »$ $3 «3 , £33,: «.25 tor landfill at NORDF Figure 3-2 Layout of bio 30 l': ‘2' w 1:1, it'. ....... :x fOtZ' (in-I .4 ltuj-Z'. ' 3,; t a . x.. I m; .. .»:t.c:;l:ztzzl:tISiulurtaW Imam- Figure 3-3: Gas sampling port bucket, and with a long tube that extended to outside the cell for sampling. The bucket was covered with a perforated plate to prevent large slid objects from entering the bucket (Figure 3-4). Two types of moisture probes were used: Time Domain Reflectometry (TDR; Automata, Inc.) and Time Domain Transmissometry (TDT; GroPoint soil moisture sensor). Both had the capability to measure temperature. Figures 3-5 and 3-6 show these probes. TDR probes were placed at all the 48 locations mentioned above and TDT probes were placed below the geomembrane cap and within the topsoil in intermediate cover. A weather station (Campbell Scientific, Inc.) located at the east side of the cell (Figure 3-7), provided continuous measurement of air and soil temperature, and precipitation. Temperature and moisture probes and weather station were connected to a data logger with multiplexer and remote data collection capabilities. Leachate flow to the sump was monitored by a flow meter (model 52 EP125, EPG companies Inc.), which operates in the 31 range of 5.7-103.4 gpm. The total gas produced from the landfill was measured at the header using a gas flow meter (model TA2, Magnetrol International, IL). Figure 3-5: Automata TDR moisture probe and hand held reader 32 Figure 3-6: Gro.point TDT moisture probe Figure 3-7: Relative location of the weather station on site (This picture was taken from east side of the bioreactor landfill) 3.2.1. Automated and onsite measurements Online measurements included moisture, temperature, weather related data (wind speed and direction, rainfall, snow accumulation, etc.), and leachate and gas flow. In addition, gas composition and temperature were measured onsite at the venting and extraction lines using a gas analyzer (Gem-meter 2000). Pressure was also measured in the dry gas sampling tubes using a manometer (Extech Instruments, NY), except in those tubes that were full with leachate. 33 3.2.2. Gas and leachate sampling protocol for off-line analysis Gas and leachate samples were collected using a peristaltic pump with a capacity to pump gas and leachate at 150 ml min'1 and 300 ml min”, respectively. The pump was connected to the sampling ports described above at the time of each sampling event. Sampling tubes were purged for 15 minutes (3 times the volume of the longest tube) before taking samples. Leachate sampling-buckets were emptied 7 to 14 days in advance before each leachate sampling event. Serum bottles (70 ml total volume) were prepared as follows for gas sampling: bottles were completely filled with acidified water (pH < 2 using concentrated hydrochloric acid) and sealed with butyl rubber stoppers and aluminum crimp caps. Gas samples were collected by water displacement method. During winter months the acidified solution also contained 20 % sodium chloride to prevent freezing. Bottles were stored at room temperature until the time of analysis. 3.3. Analytical Methods Gas samples were analyzed for oxygen, nitrogen, carbon dioxide, methane and hydrogen. A l-ml gastight syringe (Hamilton) was used to take a sample from the headspace of the gas sampling bottles. Leachate samples collected from the sampling ports were analyzed for pH, electrical conductivity (EC), chemical oxygen demand (COD), and volatile organic compounds (V DC). Protocols for measuring the above parameters are briefly summarized below. 34 3.3.1. Hydrogen Hydrogen was analyzed by injecting one mL of headspace sample in a reduction gas analyzer (RGA3, Trace Analytical) equipped with Spherocarb MoleSieve column 60/80; 31 1/4" x1/8" column at 53 0C, with the detector temperature set at 265 °C. The retention time of hydrogen was 0.2 minutes using Nitrogen as a carrier gas with flow rate 18-20 mL / min. The peaks were integrated using Turbochrom software (Perkin Elmer, MA). 3.3.2. Oxygen, nitrogen, carbon dioxide and methane Oxygen, nitrogen, carbon dioxide, and methane were analyzed using Gas Chromatography HP5890 equipped with a thermal conductivity detector (TCD). Column used was Supelco 60/ 80 Carboxen — 1000 and 15 ft 1/8 in outer diameter. Carrier gas was set to a flow rate of 40 ml/min. The temperature for the injector and detector were set at 110 °C and 200 °C, respectively. Oven temperature was set at 45 °C for 5 minutes as isothermal then increased at a rate of 40 °C/min to final temperature of 200 °C. This final temperature was kept for 2 min before it was allowed to cool down to the initial temperature. All peaks were quantified with Turbochrom software. 3.3.3. Electrical conductivity, pH, COD, VOC Leachate COD was measured using either 0-1500 ppm or 0-15,000 ppm range commercially available kits (HACH Company, CO). Leachate samples after appropriate dilution were digested in a COD reactor for 2 hours at 150 °C. Absorbance was read 35 with a spectrophotometer (Milton Roy Spectronic 20D) at 620nm wavelength and concentration was determined using standrads. The pH of leachate samples was measured in the field using a pH probe (Cole Parmer, IL) and pH meter (Hanna instruments, RI) after calibrating the pH meter using buffer solutions of pH 4, 7, and 10. Eletrical conductivity (BC) was measured using YSI conductivity meter (30/25 FT). “Check standard” (Hanna Instruments conductivity calibration solution 12.880 mS) was used to standardize the meter. VOC was measured using gas chromatography mass spectrometer (GC-MS). The GC used was (Agilent 6890) with Agilent 5973 network mass selective detector and J&W Scientific 30mX0.250mm column (DB-624 Liquid phase coating). Column temperature was held at 40 °C for 8 minutes, then increased at 5°C/min to a final temperature of 200 °C. The method was same as EPA prescribed method for measuring VOCs. 3.4. Experimental Procedures To accomplish the four objectives of this research, a series of experiments were conducted on Northern Oaks bioreactor landfill. Figure 3-8 shows the title and duration of each experiment followed by the associated sampling events. 36 35>“... wow—mama 98 3225“ 358598 mo 32225 “arm «.5»:— Gm» 88 a: 963288 302550 93695 95% Sea mew—anew 8383 a: $55 mam—9:8 03:83 6% 35.6 mam—95m mew nouoo E be Bow mmmcowoqafiowz v.68 cowofismaoo commxo N “moo cocoa E has _ $2 compo E .__< 95.8% waist mmmoeowocmfioE oomosemeoo E23 oo: ooo _ ooo oow ooh ooo oom oov oom ooN oo_ 37 3.4.1. Performance of intermittent air injection system (Objective 1) Objective 1 was to evaluate the intermittent air injection in terms of its impact on temperature, oxygen and nitrogen dissemination into the landfill, and leachate characteristics. Air injection was accomplished using a Fugi Ring compressor with capacity 2.5 hp (500p-2T) with a maximum pressure of 80 inch and maximum flow of 154 scfrn at two different time points. Air injection test 1 was conducted from Sep 03 to Oct 03, while air injection test 2 was conducted from May 04 to Dec 04. Air injection test 1 occurred before the operation of the gas extraction system. Hence, during air injection test 1, the horizontal gas extraction lines were used for air injection and the horizontal leachate collection lines were used to vent the injected air. Air injection test 1 had two phases. Air injection during phase 1 was conducted through pipes in lift 1, while during phase 2 air was injected through lift 4. Air injection test 2 was conducted after the installation of the gas extraction lines. Hence leachate recirculation lines were used to inject air. Air injection test 2 had two phases as well. During phase 1 air was injected through lift 1 and during phase 2 through lifts 3 and 4. Detailed description of each test is presented below. During the two air injection tests, gas samples collected from most of the ports were analyzed for oxygen, nitrogen, carbon dioxide and methane. In addition leachate samples were collected before and after air injection experiments to evaluate the effect of air injection on leachate quality. Leachate samples were analyzed for COD, pH, and VOC. Composition of the vented gas and the flow rate of injected air were measured. Ambient and landfill temperature and humidity were monitored continuously. 38 Air injection test 1 was conducted between 9/5/03 and 10/24/03. During Phase 1, air was injected into gas extraction line G12 and vented from leachate recirculation lines L32 and L42. The air was injected at a rate of 91.5 scfrn for 21 days, from 9/5/03 to 10/2/03. The volume of air injected during Phase I was ~2.59 million cubic feet. During Phase 11, air was injected into gas extraction line G4 and vented from leachate recirculation lines L32, L5, and L6. The air was injected at a rate of 82.7 scfm for 22 days, from 10/2/03 to 10/24/03. The volume of air injected during Phase II was also ~2.60 million cubic feet. Air Injection test 2 was conducted between (5/25/04 to 12/8/2004). During Phase 1 of air injection test 1 (5/25/04 and 10/14/04), lifts L11, L12, L13, and L14 were used for air injection. The active aeration period lasted for 130 days and air injection flow rates varied for each line ranging from 41 to 133 scfrn. The volume of air injected during this period was 12 million cubic feet. The percentages of air-injected volume to each line in comparison to the cumulative volume was 36, 26, 16, and 22 % for lines L11, L12, L13, and L14, respectively (Table 3-2). This was calculated based on flow rate in each line. Phase 2 of air injection test 2 spanned from 10/15/04 and 12/08/04 and targeted lines L31, L32, L33, L41 and L42 in lifts 3 and 4. The active aeration period lasted for 45 days. Air injection flow rates varied for each line ranging from 46 to 122 scfrn. The volume of air injected during this period was 5.77 million cubic feet. The percentages of air volume injected in each line as a percent of total air injected was 21, 17, 8, 9 and 45% for lines L31, L32, L33, L41, and L42, respectively (Table 3-3). During air injection test 2 (both phases), the gas extraction line was operational. 39 Table 3-2: Air injection test 2 (lower lift) - Percent flow in each injection line Distribution of air . . Volume Duration of aeration Injection pipe injection volume (scfm) (Day) (%) L11 4,331,440 36 39 L12 3,072,482 26 39 L13 1,962,510 16 20 L14 2,621,045 22 32 Total 1 1,987,476 100 130 Table 3-3: Air injection test 2 (upper lifts) —injected air volume distribution among the injection lines Distribution of air Duration of hip???“ Vr()lume injection volume aeration “0 (%) (Day) L31 1,224,891 21 6.9 L32 981,675 17 L33 459,284 8 4 L41 519,737 9 7.8 L42 2,588,227 45 19.4 Total 5,773,815 100 44.7 Computation of heat generation factor amounts to performing a heat balance around the bioreactor landfill. The data for this objective comes from the two air injection tests. Parameters needed to calculate the heat generation factor include, temperature and moisture of the waste, temperature beneath the landfill, the ambient air temperature and humidity, the air injection flow rate, the gas venting composition, and the volume of the 40 waste. The approach to compute the heat generation factor is described in more detail in section 3-6. 3.4.2. Oxygen consumption rate (Objective 2) The experiment to calculate oxygen consumption rate was conducted during the air injection test 1 in 2003. Briefly, after maintaining air injection for several days, it was stopped so that decrease in oxygen concentration at each sampling location could be monitored. A background sample was taken just before stopping the air injection, and then gas sampling continued until the oxygen levels at the sampling locations dropped to very low concentrations. This experiment was repeated twice at different initial temperatures, moisture contents and starting oxygen concentrations using sampling ports located in lifts 1, 2, 3 and 5. Lifts 1, 2 and 3 were at similar moisture content as given by moisture probes however lift 5 had lower moisture content compared to all others. 3.4.3. Evaluation of anaerobic digestion process after startup or air perturbation (Objective 3) Anaerobic digestion process was evaluated at the start of the bioreactor landfill cell operation (i.e., prior to air injection test 1) as well as following air injection test 1. It was accomplished by measuring methane and hydrogen concentration in the sampling ports inside the bioreactor landfill. During the startup phase of the bioreactor landfill (1/19/03- 9/5/03), 19 gas sampling tests occurred. One sampling event amounted to collecting samples from most of the 48 locations for gas and leachate analysis. Gas and leachate monitoring after air injection lasted up to 2 months. Temperature within the landfill was continuously measured during this period. Samples were analyzed for 41 methane, carbon dioxide, oxygen, nitrogen and hydrogen by methods described above. Leachate samples were analyzed for VOC, COD, and pH. 3.5. Statistical Methods Statistical analyses were conducted using SPSS version 11. The tests used to analyze the data include: t-test, Mann-Whitney U Test, paired t-test, Wilcoxon T Test, Pearson and Spearrnan's correlation, and partial correlation. T-test and Mann-Whitney U Test were used to test the difference in the mean and median of two independent variables, respectively. Paired t-test and Wilcoxon T Test were used to compare two dependent variables in terms of their mean and median, respectively. Pearson and Spearrnan’s correlations study the significance and strength of a linear relationship between two variables. The above tests are grouped into two types; parametric and nonparametric (Table 3-4). Parametric tests are conducted when the normality assumption is verified. The normality assumption means that the data is normally distributed. For non-normal distribution, parameters of estimation used to describe the normal distribution such as mean and standard deviation cannot describe the data any more. Therefore, nonparametric tests do not use these parameters. All nonparametric tests require ranking of the values. In order to do so the values from both samples are combined and ordered from lowest to highest. A rank is given for each value. In case of a tie in the values of the observation, the range of the ranks for the 42 group will be calculated, and an average rank for each observation in the group will be given. Table 3-4: Comparison between parametric and nonparametric tests Reason to test Parametric test Non-parametric analogue [1; est difference Independent Mann-Whitney U etween two . samples t-test Test independent samples TCSt difference Paired sam les t- between two dependent test p Wilcoxon T Test samples Tests the linearity of a , , . . Pearson s Spearrnan 5 relationship between . . . correlatlon correlation two varlables 3.5.1. Normality assumption Several methods are available to test the normality assumption. The first method involves calculation of skewness (3-1) and kurtosis (3-2) of the variable and visual observation of its distribution. Kurtosis is the measure of whether the data are peaked or flat in comparison to a normal distribution. Skewness is a measure of lack of symmetry. If the range of kurtosis and skewness is between -1 and +1, then the assumption is valid. If these parameters are not within the normality range, the central limit theorem can be applied. This theorem indicates that if the number data points is above 50, then parametric tests can be used. _ 4 2:100 " y) (n —1)SD4 kurtosis = — 3 ................................................................................... (3-1) 43 where, E is the mean, SD is the standard deviation, and n is the number of data points. n -3 Zi=](yi -y) 3 ....................................................................................... (3-2) (n —1)SD skewness = The other two methods to test the normality of distribution of the data are Kolmogorov- Smimov (D) given by equation (3-3) and Shapiro-Wilk tests (W) given by equation (3- 4). D = max (F(Y,-))— i ................................................................................................ (3-3) lSiSn n where, F is the theoretical cumulative normal distribution ’1 Zaixo) '=l W = —n’————— ...................................................................................................... (3-4) (x,- " 3‘22 E where, x(i) are the ordered sample values (x(l) is the smallest), and a,- are constants generated from the means, variances and covariances of the order statistics of a sample of size n from a normal distribution. 3.5.2. Test difference between two independent samples The two tests used to evaluate the difference between two independent samples are t-test and Mann-Whitney U Test. The t-test is used when the distribution of data in the two samples is normal; otherwise Mann-Whitney U Test is used. The t-test is conducted on the sample values directly, while Mann-Whitney U Test is conducted on ranked data. In order to conduct t-test, t given by equation (3—5) and the degrees of freedom (df) are calculated. Then using values of (t) and (df), the significance of the test is read from statistical tables. If the significance of the test is less than 0.05, then there is a significant difference between the two samples being tested. The standard error of the difference in mean (SE difi" ) used in equation (3-5) and the degrees of freedom are calculated by either of the following methods. The choice of method depends on the homogeneity of the variances of the two samples. Homogeneous variances mean equal variance in both samples that are tested for difference in their mean. These equations for the two cases are shown in Table 3-5. M . t= “ff .................................................................................................................. (3-5) SEdl-fl‘ where, Mdifl =|M1—M2 ................................................................................................... (3-6) 45 M1 and M 2 are the means of first and the second samples, respectively. SE at)? is the standard error of the difference in mean (Table 3-5) Table 3-5: Calculation of the standard error of the difference and the degree of freedom for the t-test under homogeneous and non-homogeneous variances . Non-homogeneous Homogenous variances variances Standard error of the 2 2 SD SD _ l 2 difference SE diff = (SDI + SD2 ) SEd'fl —\/ n + n J n1 + n2 1 2 2 2 2 [SD/ 313/] + ”1 "2 Degree of freedom _ _ _ df = 2 2 df - (n1 1) + (m 1) 5012 3022 "1 "2 + n1 — n n2 —1 Were, n1 and n2 are the sample size of first and the second samples, respectively, and SDI and SDZ are standard deviations of the first and second samples In order to conduct Mann-Whitney U Test, U value is calculated (3-7). If U value is bigger than that in the statistical tables that corresponds to the number of the sample points at the 0.05 significance, the test is significant. M-R, ......................................................................................... ,(3-7) U=nn + 12 2 46 where, R] is the sum of the ranks in sample 1, and n1, n2 are the sample size of 1St and 2nd samples. 3.5.3. Test difference between two dependent samples The difference between two dependent samples can be evaluated by using either paired t- test or Wilcoxon t-test. Paired t-test is used when both samples are normally distributed; otherwise its nonparametric analogue (Wilcoxon t-test) is used. To conduct a paired t-test, the degrees of freedom (df) given by equation (3-8) and t value given by equation (3-9) are calculated. Using statistical tables, the significance that computed t value and the degrees of fi‘eedom is read. If the significance is less than 0.05 then there is a significant difference between the two groups of samples. df = n —1 ................................................................................................................... (3-8) Where, n is the number of the pairs. t = d .................................................................................................................. (3-9) SE difl Where, 47 E is the mean of the difference in value of the paired data points, given by equation (3- 10), and SEdifi‘ is the error of the difference, given by equation (3-11) 3:21—49 ................................................................................................................... (340) n Where, d,- is the difference in value for the paired data points SEW = ........................................................................................................ (3-11) SDdiff is the standard deviation of the difference in value of the paired data points, given by equation (3-12) ._— 2 2(‘1' d) ............................................................................................. (3-12) n—l SDdIflr = To conduct Wilcoxon t Test the data is ranked as discussed above for all nonparametric tests. If there are no ties in the ranking, then equations (3-13) through (3-15) are used. 2 J31?) .............................................................................................................. (3-13) J VAR 48 Where, VAR is the variance of the expected rank, given by equation (3-14) ER is the expected rank sum, given by equation (3-15) R is the higher rank sum of the two samples, given by equation (3-16), VAR = "(n + ”(2" +1) ............................................................................................... (3-14) 24 ER = w ............................................................................................................ (3-15) There may be two kinds of ties in this test. In the first case the values of the pair are exactly the same. In the second case, the difference in the values of the pair is the same. In the first case, the measured values are ignored and n is adjusted accordingly. In the second case, the variance calculated by equation (3-14) will be reduced by a value of x, which is calculated by equation (3-16). 3 t -t = ................................................................................................................... 3-16 x 48 ( ) Where, t is number of the tied ranks. 3.5.4. Tests the linearity of a relationship between two variables Quantifying the linearity of relationship between two variables is established by Peasron or Spearman’s correlations. When the dependent variable is normally distributed, 49 Pe' Cl Pearson correlation (ryx) is used given by equation (3-17). Otherwise the nonparametric analogue, Spearman’s correlation (r,) is used, given by equation (3-19). To conduct Spearman’s correlation values of samples are ranked, and the Spearman’s correlation is calculated based on the ranks. EPG-50’1"?) S r ="=‘ = "y ...................................................................... (3-17) y" Jams}, 5,0,3” Where Sxx, Syy are the variances of x and y variables, respectively. Variance is given by equation (3-18), 3 and E are the average of the values of the variables y and x, respectively, and n is the number of samples. n — SW = Z(y,~ -y)2 .................................................................................................... (3-18) i=1 6 d2 ,._..-z_ ....................................................................................................... .3-... n(n -1) Where, d is the difference in ranks between the two variables 50 Partial correlations quantify the linear correlation between two variables while controlling for a third one. The formula to determine the partial correlation coefficient ryxc, where c is the control variable, and y is dependent on x is given by equation (3-20). If partial correlation equals the correlation without control of variable c, then the control variable does not affect the correlation of variables x and y. A partial correlation that equals zero indicates a false correlation between variables x and y, and the relationship is not direct between these variables. ryx—rycrxc = ............................................................................................ 3-20 ryxc ‘ll _ ryc J1 _ rxc ( ) 3.6. Method to Determine Heat Generation Factor Heat generation factor is defined as the amount of heat produced (kJ) per unit mass of oxygen (kg) consumed. To compute heat generation factor, a heat balance around the bioreactor landfill needs to be performed. Heat balance equation states that accumulation of heat equals to the total heat gained or produced minus the total heat loss. Sources of heat generation in a bioreactor landfill are microbial exothermic reactions and sensible heat gain. Losses in heat can be attributed to sensible heat, latent heat and heat loss to boundaries. When including these heat losses and gains the heat balance equation would be accumulation of heat = i sensible heat — latent heat + metabolic heat generation - heat loss to boundaries Mathematically the above equation can be written as: q accumulation = 313‘] sensible i qlatent + q metabolic i q boundaries ---------------------------------- (3'21) 51 The following assumptions are made to solve equation (3-21): The solid and gas phases are in equilibrium. This means that the temperature of air and waste are equal in the bioreactor. This is usually named as homogenous assumption, and is frequently used when conducting heat modeling in composting operations[24]. The gas phase within the bioreactor landfill and the effluent gas are saturated with water vapor. This is true because at low aeration rates and high waste moisture content, moisture transfer from solid waste matrix to air encounters no resistance [24]. Changes in the pore space and reactor working volume were assumed to be negligible during the period of observation used in the heat balance study. This assumption is justified because the duration of each air injection is less than 23 day. Oxygen enters the reactor only through forced aeration (advection flow) and the diffusion of oxygen into the landfill is negligible. The landfill cell was capped with HDPE geomembrane minimizing oxygen diffusion. Effluent gases are at a constant pressure. This assumption also been used in previous studies [24]. Pressure change in the landfill is measured and was in the range of 1- 2 mbar approximately about 0.2% of the atmospheric pressure. The moisture content of the solid waste matrix is assumed to be constant. The heat generation factor is constant and is a result of aerobic reaction only, and heat generation due to anaerobic process is negligible [74]. Although some 52 anaerobic reaction might be taking place, the heat produced of aerobic reaction is much higher than of the anaerobic one. Using the above assumptions and substituting with the appropriate equations for each heat term in equation (4-2), the heat balance equation is given by equation (3-22). d pCpVEU") =FLCprL(TL "Twaste) +FaiGCairpair(Tair ”Twaste)+ ............ (3-22) + AHw(xw,out — xw,in ) + QR(02) + UA(Tout '— Twaste) Each of the five terms is described in detail below. 1. Accumulation of heat is given by equation (3-23) d qaccumulan-on = pCszi-t-(T) ..................................................................................... (3-23) where, p is the apparent density of the waste (kg/m3), V is volume of the waste in the landfill (m3), T is temperature (°C), t is time (sec), dT/dt is the rate of temperature increase during air injection. This rate equals the slope of temperature vs. time curve, and Cp is the specific heat of solid waste (J/kg °C), given by equation (3-24) Cp = Cp-solids °msolids + Cp—water 'mwater .............................................................. (3-24) where, 53 Cp-so]ids is the heat capacity of solids and ranges between 0.5-2 J g'1 0C'1 for solid waste [58]. A typical value of 1.2 J g.1 0C1 is generally used, including this study, Cp-wate, is the heat capacity of water 4.1818 J g.1 °C.1 [110], mwatc, is the mass fraction of water, and msonds is the mass fraction of solids. 2. Sensible heat change is the heat change in the landfill due to transport of fluids and air [55]. It is given the following equation: qsensible = FL CpL pL (TL - Twaste ) + Fair Cpair pair (Tair '" Twaste) ----------------------- (3'25) where, Twaste is the temperature in the waste (°C), Tai, is the temperature of the incoming air (°C), TL is the temperature of the incoming leachate (°C), F is flow rate (m3/s), Cp is the specific heat (J/kg°C), p is density (kg/m3), and subscripts air and L refer to the air or leachate moving injected in the landfill. 54 3. Latent heat loss is due to evaporation of water, and can be significant if humidity of the injected air is low. It is given by equation (3-26) qlatent = AHW(xw,om - xw,in) .................................................................................. (3’26) Where, AH“, (J/kg) is the heat evaporation of water, x w, in is water mass in the coming air per unit time (kg/s), given by equation (3-27), and x w, out is water mass in the outgoing air per unit time (kg/s), given by equation (3-28). M P - xw in = W W" Fair ..................................................................................... (3-27) , Vstp P0 " Pw,in MW PW,0W Tair x , , = - ...................................................................... (3-28) w ou Vstp PO _ Pw,0ut arr T waste where, MW is the molecular weight of water (0.01 8 kg/mole), Vstp is the volume of one mole of water vapor at STD (22.4 L/mole), P0 is the atmospheric pressure (atm), Fair is the air flow rate at STD (L/s) Pw, out is the partial pressure of water inside the landfill and equals to saturated vapor pressure at the waste temperature (atm), 55 T air is temperature of the incoming air (°C) T waste is the temperature of the waste (°C) Pw,in is the partial pressure of water in the incoming air (atm) and given by equation (3-29), RH 0 PW, in = 1—06 . w,in ...................................................................................................... (3'29) where, RH is the relative humidity in the incoming air (%), and O Pan is the saturated vapor pressure at the air temperature 4. Metabolic heat generation is given by the following equation: qmetabolic = Q.R(02) ............................................................................................... (330) Where, Q is the heat generation factor (J /mol 0;), and R(02) is the amount of oxygen consumed per unit time (mol Oz/s), given by equation (3-31). x - - X Fair (%) 12(02) = V ............................................................................................. (3-31) stp Where, x is the concentration of oxygen in the venting or extraction lines (%), xai, is the concentration of oxygen in air (%), 56 Vstp is the volume of one mole of water vapor at STD (22.4 L/mole), and Fair is air flow rate at standard temperature and pressure (Us) 5. Heat loss to the boundaries is given by equation (3-32) (1,0,, = UA(Tou, — Twm) ......................................................................................... (332) where U is the heat transfer coefficient (J s.I m.2 -°C), A is the area of the bottom of the landfill, footprint area (m2) T waste is the waste temperature (°C), and Tout is the temperature of soil undemeath the landfill (°C). 3.7. Method to Determine Oxygen Consumption Rate Ideally, oxygen consumption rate should be determined in a manner that allows normalization of the rate to the amount of organic matter in the waste. Under field conditions, assigning an active volume of waste to a given rate of consumption is not feasible. Therefore, a new term, called “Point” oxygen consumption rate (POCR) was coined in this study. Point oxygen consumption rate is defined as the rate of decrease in oxygen concentration with time at a given location in a bioreactor landfill. Although POCR is not ideal, it has the advantage of measuring the consumption rate under field conditions. Neither oxygen consumption rate, nor POCR have been measured under field conditions. 57 The approach to determine POCR includes (1) oxygen concentration vs. time curve (data collected), (2) calculation of liquid phase oxygen concentration, and (3) calculation of kinetic coefficients for POCR using Monod equation. 3.7.1. Oxygen concentration vs. time curve The method to obtain POCR involved measuring oxygen concentration at a sampling location for a period of approximately 8 hours. This period included a background reading before stopping the air injection. Figure 3-9 shows an example of change in the concentration of oxygen and other gas phase constituents concentration after stopping air injection. The concentration of oxygen decreased from 13% to 2% in 7 hours. During this time, carbon dioxide concentration increased from 4% to 9%. The nitrogen and methane concentrations stayed constant during the entire period of the experiment. ,HW . “r h— 2:: s .o u: :1 lg g 8 8 ii: 8 :0 e0 0 | 2‘ l l o 1 2 3 4 5 6 7 8 E Time(hr) I | :o— CH4 -—c02 +02 -x-N2 Figure 3-9: Gas concentration change after stopping air injection location 6143 58 3.7.2. Calculation of liquid phase oxygen concentration POCR was computed using liquid phase oxygen concentrations. Gas phase oxygen concentration was converted to liquid phase concentration and corrected for temperature and salting out effect. For each sampling location, Henry’s constant was calculated based on the corresponding temperature at that location, and corrected for the salting out effect. Then Henry’s constant combined with the partial pressure of the oxygen at each location were used to calculate the liquid phase oxygen concentration. The underlying assumption when using this equation is that the gas and liquid phase oxygen concentrations are at equilibrium. The liquid phase concentration of oxygen was calculated by equation (3- 33). 0 = y'P ............................................................................................................ (3-33) 100H2 Where, y is the percent of oxygen in the gas phase at a given location (%), H2 is corrected Henry’s constant for temperature and salting out effect (L atm/mole), and P is the pressure within the landfill (atm). Pressure measured in the landfill at sampling locations during the POCR study did not vary significantly compared to atmospheric. Hence, atmospheric pressure was used in equation (3-33). Temperature correction for Henry’s constant The Van’t Hoff -Type equation was used to correct Henry’s constant for temperature, and is given by equation (3-34). 59 H2 =H1 exp-£[i—i] .............................................................................. (3-34) R T2 T1 Where, H1 is Henry’s constant at known temperature T1 (K), H2 is Henry’s constant at unknown temperature T2 (K), AH° is the standard enthalpy change in water due to the dissolution of oxygen in water (kcal/kmol), R is the universal gas constant (1.987 kcal/kmol.°K), Another form of equation (3-34) is given by equation (3-35). O 10307) = - + C ................................................................................................ 3-35 RT ( ) The constants C and AH° for oxygen in equation (3-35) equal to 7.11(tmit less) and 1.45 103 (kcal/kmol), respectively [111]. Using these values Henry’s constant (H) for oxygen was calculated in units of atmosphere. Henry’s constant units of (atm) can be converted to units of (L atm mole-1) by equation (3-36). H[ My") = ”(atm) 1 ................................................................................... (3-36) m0 8 m0 3 MVwater( L j 60 Where, M Vwam is the molar volume of water and equals to 55.6 mole L'l. Salting out effect correction for Henry’s constant As described in [112], correcting for the salting out effect on Henry’s constant is given by equation (3-37). H 2 = H17 .................................................................................................................. (3-37) Where, H2 is Henry’s constant corrected for both salting out effect and temperature (L atm/mole), H1 is the Henry’s constant corrected for temperature in (L atm/mole), and y is the activity coefficient for oxygen solubility in salt water, given by equation (3-38). log(7) = ks! .............................................................................................................. (3-38) Where, k8 is the salting out coefficient and equals 0.132 for oxygen, and I is the ionic strength, given by equation (3-39). 1 = 0.016EC .............................................................................................................. (3-39) Where, EC is the electronic conductivity [fig]. cm 61 3.7.3. Calculation of kinetic coefficients for POCR using Monod equation Monod kinetics, given by equation (3-40) is the most common model used for aerobic substrate degradation [1 13-1 15]. V _ dS = _kmax.S.X — — ............................................................................................... (3-40) dt K S + S Where, V is the substrate utilization rate (mg L.1 h.1 ), S is the concentration of substrate (mg L'l), . . —l . . . Ks rs the half saturation constant (mg L ); re the substrate concentration at which the substrate utilization rate is half the maximum substrate utilization, X is concentration of biomass (mg /L), kmax is the maximum specific substrate utilization rate (L mg '1 h'l), and t is time (h). The kinetic coefficients (kmax, K5) are classified into two groups based on the ratio of the substrate to the biomass concentration. When the ratio of substrate to biomass is high, unrestricted growth occurs and the kinetic coefficients are called intrinsic. When the ratio is small the kinetic coefficients are called extant [116]. While determining extant kinetics the growth is minimized thus the concentration of biomass X is relatively 62 constant [116]. When change in the biomass concentration is negligible, Monod equation (3-40) can be reduced to equation (3-41) [117]. dS k.S V = — = .................................................................................................... 3-41 dt Ks +S ( ) k is the maximum substrate consumption rate (h'l) , given by equation (3-42). k = kmaxX ................................................................................................................ (3-42) In order to obtain a solution for Monod equation (3-41), kinetic parameters (k, K.) have to be unique and independent. Knights et. al. [118] found that a unique solution for Monod equation can be obtained when the ratio of the initial substrate concentration (So) / to the half saturation constant (Ks) to be larger than 0.1 iii—0 > 0.1]. Moreover, the S uniqueness of estimates for k and Ks depends on the number of observations and the precision of the measurements [118]. Estimation of Monod coefficients k and KS is rather difficult because these two kinetic coefficients are not completely independent [114]. The experimental conditions at which these parameters are estimated can be controlled to lead to minimal dependence among the parameters [114]. When the ratio of SOIKs is bigger than 1, a good separation between the coefficients k and Ks can be obtained [118]. 63 To estimate the kinetic coefficients, substrate concentration (S) must be distributed over a large concentration range of the rate curve. This is ensured if the substrates concentrations used cover the following three regions [114]: 1) Region in which the slope is very steep and constant (0 < S/Ks S2, 0 > K S) then the degradation will follow zero order kinetics with respect to substrate. This is given by equation (3-45). dS kS V=—=-—=—k=-k ........................................................................................ 3-45 dt S o ( ) Integrated form of equation (3-45) is: S = S0 — kot ............................................................................................................. (3-46) Where, k0 — Zero order rate constant of substrate degradation (mg L'1 h'l) Three methods to solve Monod equation (5-9) are: the integrated solution of Monod equation, linearized form of Monod equation, and nonlinear regression methods. The following two sections describe 1) the methods to solve Monod equation, and 2) fitting POCR data to the Monod equation. Methods to solve Monod Equation The first method to solve Monod equation (3-41) is by its integration and then solving the analytical solution by least square method [119]. Integration of equation (3-41) is given in equations (3-42) through (3-45). 65 KS+S — d5 = dt ...................................................................................................... (3-42) 'IESS—dS +%dS = —dt ................................................................................................ (3-43) Kks 1n(SO)+,1C—SO—Kksln(S)-—}c—S=t .................................................... (3-44) %1n(%9) +£(sO — S) =z ...................................................................................... (3—45) The integrated solution given by equation (3-45) is in the form of time (t) as a fimction of substrate concentration. The least square method minimizes the difference between the measured time and the calculated time by equation (3-45) for each measured substrate concentration, using variable kinetic coefficients k and Ks. In order to estimate the coefficients k and Ks by varying time rather than substrate concentration, measurements of substrate concentration must be accurate. The second method to solve Monod equation uses the linearized form of equation (3-41), also known as Lineweaver-Burk plot [113, 120]. The linarized form of Monod equation is given by equation (3-46). -53. k + ........................................................................................................ (3-46) Cal—- Pelv— _1_ V In this solution the rate dS/dt is substituted with a measured value of the substrate utilization rate (V). To accomplish this, laboratory experiments involving concentration vs. time curves with different initial substrate concentrations are conducted. Using these 66 curves, initial substrate utilization rates are calculated. These rates equal the slope of the initial linear portion of the concentration vs. time curves (Figure 3-10). For each initial substrate concentration, an initial rate is obtained. For example the rate in Figure 3-10 equals 3.23 mg L.1 h'1 at an initial concentration of S=8mg/L. (The data in this figure are hypothetical and for the purpose of explaining the method). Substituting the rate (V) and concentration (S) in equation (5-15) will result in a linear plot in which Ks/k is the slope and l/k is intercept. The advantage of this method is its mathematical simplicity. However, this method suffers from two disadvantages. First it assumes that the errors are normally distributed, which might not be valid. Second, this method cannot be applied under field conditions. This method has been used by many researchers in the past because of the complexity of nonlinear regression methods that involve iterative techniques to obtain optimal values of estimates of k and Ks [118]. With the improvement in the computerization techniques, nonlinear regression techniques are now preferred. 67 9 . I A 8 ' ——— —. _ .- ___, 2w, 7% E}, 7 ,__l=-3.23x+7.8027_ , g V 2— g 6 ”.2 R;o.asa5__-_ -flfi __ .g ____ __ __ g 5 G 4 ‘ .2_ _ _ _ _____ ___ ‘2- § § 1 ° l l 8 3 p --+.—---— ——~ fl ————~ ~——- * 2 w + h m m a i ? 1 +__ '_'_—4"§".—; ' _——___ 7 "fl — I i O 0 l .——~ ————————?~ ’ -»—-+—--—0— o ~~ '1 i ‘ o 2 4 6 8 10 12 , i Trme (hr) E Figure 3-10: A hypothetical example for determination of the initial rate for Lineweaver- Burk plot The third method to solve Monod equation (3-41) is by nonlinear regression analysis, which is now routine due to improvements in computational methods [114, 118, 121- 123]. In addition, the use of this method does not require constructing several curves of substrate vs. time. The rate can be determined using one curve. This is particularly important when data collection is expensive or it is not feasible to obtain several curves. In this method, given initial values for k and Ks the residual sum squares between the observed and predicted concentration is minimized. The values of k and Ks when the residual sum squares reaches minimum is the optimum solution. In this study, the method chosen to solve Monod equation (3-41) was nonlinear regression. The differential equation was solved numerically using Runge-Kutta approximation technique. Given an initial concentration of oxygen and trial set of k and 68 Ks, the residual sum squares between the observed and predicted concentration was evaluated using frnincon function in Matlab, given by equation (3-42). ’.’ s,—s 2 k . f(6)= \[ZIJ ) ,9: K, ........................................................................ (3-42) n so Equation (3-42) attempts to find a constrained minimum of a scalar function of several variables starting at an initial estimate. The least square minimization routine based on Levenberg Marquardt algorithm was used to minimize frnincon function. Coefficients refining continued until frnincon reached global minimum. The coefficients corresponding to the global minimal value were taken as the best fit. 69 4. RESULTS AND DISCUSSION Low ambient temperatures in cold climates during waste filling period can lead to persistent low temperatures inside a bioreactor cell. Air injection is a practical option to increase waste temperature. In Northern Oaks bioreactor landfill, low temperature in the waste was an issue leading to a slow start of the anaerobic process. The three objectives in this study were: Objective 1: To establish and quantify relationship between amount of heat produced (1d) per unit mass of oxygen supplied (kg), and to characterize the performance of air injection with respect to temperature and leachate characteristics. Objective 2: To establish the spatial variability in point oxygen consumption rate (decrease in oxygen concentration with time at a given location in a bioreactor landfill), and study the effect of temperature, oxygen concentration and moisture content on point oxygen consumption rate. Objective 3: To quantify the performance of anaerobic digestion at startup and following air injection with respect to methanogenesis establishment, using methane and hydrogen concentrations as indicators for anaerobic digestion process. Two air injection tests for evaluation the correlation between air injection and temperature increase were conducted. POCR was determined during the first test by stopping the air injection for about 8 hours. Methanogenic activity was studied before air 70 injection test 1 (i.e., during startup of the bioreactor landfill) and after the air injection test 1 was stopped. 4.1. Evaluation of Intermittent Air Injection in Northern Oaks Bioreactor Landfill Two separate air injection experiments were performed to quantify relationship between increase in waste temperature inside and amount of air injected. This section starts by establishing the base line for waste temperature before air injection. Then evaluation the response of temperature to air injection is described in the second subsection. The third subsection addresses the response of waste temperature response after stopping air injection. In the forth heat generation analysis are presented. The fifth section describes the gas phase concentration—time profile during the air injection. In the last section, the impact of air injection on leachate characteristics is characterized. 4.1.1. Establishment of base line temperature before air injection Base line of temperature was established by studying the ambient temperature and its relation to initial waste temperature. In addition, the effect of leachate recirculation before air injection on waste temperature is discussed in this section. Correlation between waste and ambient temperature Low temperatures such as those observed in Northern Oaks bioreactor landfill have not been reported in literature before. Temperatures reported in landfills in the range between 10 and 70 °C [15, 17, 33, 34, 36-39]. Lower ambient temperatures during filling and before waste coverage resulted in lower initial waste temperatures (Figure 4- 1). In this study, the average of waste temperature until five days after covering it with 71 subsequent waste is termed initial waste temperature. Figure 4-1 illustrates the average of initial waste and ambient temperature for the five lifts. 1'- 95% confidence interval of mean 0 Mean :3 I: .. ”1337-7: e: 20 he ———~*~ my. 5 15: - / -— § 10 I "T ‘Tm‘Lifl3 g 5 i 'Lifl4wi7(W—_ny=l.6n301x+10.858 g 5 i- / _ _ RITOL9855 -15 -10 -5 0 5 10 15 20 Airtemperature(C) Figure 4-1: Average waste temperature vs. Average ambient temperature waste is exposed to in each lift The range and average of ambient temperature for which waste was exposed to were (0.5 to 32 °C, 18:0.3 °C), (-8 °C to 20 °C, 38:03 °C), (-24 °c to 10 °C, 2810.4 °C), (-28 °C to 10 °C, -6.5i0.4 °C), and (-30 °C to -1 °C, -103-_0.4 °C), for lifts 1, 2, 3, 4 and 5, respectively. The difference in ambient temperatures that each lift was exposed to was significant as determined by Kruskal-Wallis Test (Table 4-1). The resulting initial waste temperatures reflected the above variation in ambient temperatures. Waste temperatures were significantly different among the lifts as determined by Kruskal-Wallis Test (Table 72 4-1). Average initial waste temperature in lifts 1, 2 3, 4, and 5 were: 27.73205 oC, 17.0:10 °C, 54:27 °C, 20:13 °C, and -6.5i1.3 °C, respectively (Figure 4-2). Table 4-1: Kruskal-Wallis Test for testing differences between lifts waste lifts with respect to a) ambient temperatures to which waste was exposed and b) waste temperature Lift N Mean Cbi- d f Asymp. number Rank Square Sig. Lift 2 816 2625 - Lift 3 911 1750 A" 1543 3 <0.001 temperame Lift 4 1055 1211 Lift 5 502 758 Lift 2 69 103 Lift 3 25 56 waste 103 3 <0.001 temperature Lift 4 30 40 Lift 5 16 9 73 50 4o. 8 30‘ E 1— g 20: I at 8" —— “l 6 O E 101 —-[— .93. I 53 0. fl ‘ i 3 _r_ ._.__-______ -101 -20 . - N = 375 69 25 30 16 Liftl Lift2 Lift3 Lift4 Lift5 Figure 4-2: Initial waste temperature in the landfill Initial waste temperatures were higher than ambient temperatures to which waste was exposed. The differences between ambient and waste temperatures were: 9.6, 13.2, 8.2, 8.4, and 3.6 °C for waste lifts 1 through 5, respectively. Waste temperature correlated strongly with ambient temperatures (R2=0.99). The relation between initial waste temperature and ambient temperature to which waste was exposed for lifts 2, 3, 4 and 5 was established in this study as: Twa,,,=1.6 Tambiem + 10.9 ...................................................................................... (4.1) Where Twaste is the waste temperature in a lift (°C), and 74 Tambiem is the ambient temperature the waste was exposed to in that lifi (°C). Lift 1 was excluded from the analysis, because it was uncovered for a period of 90 days, however lifts 2, 3, 4 and 5 were exposed to air for about 2 weeks each, making them comparable. The increase in waste temperature above ambient temperatures was attributed to heat generating aerobic biodegradation reactions. Measured oxygen and carbon dioxide concentrations in the landfill indicated that oxygen consumption and carbon dioxide production were taking place in the waste mass. The oxygen and carbon dioxide concentrations profiles in the gas phase for lift 4 are shown in Figure 4-3. Oxygen concentration decreased from 9.0i2.9 % to 4.4i1 .7 % within two days after coverage and further decreased to 1.7i0.3 % by day 18 after waste coverage. This decrease in oxygen concentration was accompanied by an increase in carbon dioxide concentration from 20.7:70 % to 30.6i7.7 % during the first two days, and to 33.71:] .67 % by day 18. 75 Carbon—dixoide concentration (%)i ’ j Days after covering lift 4 .. _.._ _1 E mm -rmx omedian —-—mean; Figure 4-3: Gas composition after waste coverage Effect of leachate recirculation on waste temperature Leachate recirculation is used in bioreactor landfills to enhance biodegradation. However the effect of leachate recirculation on waste temperature is not yet documented. Before the impact of air injection on waste temperature can be studied, it is important to understand the effect of leachate injection on waste temperature. Leachate injection started as soon as the waste was covered. The temperature of the injected leachate was constant at about 125°C throughout the leachate recirculation period. After the start of leachate injection, the difference between waste and leachate temperatures began to decrease. When leachate was not practiced in a lift waste 76 temperature remained constant (Figure 4-4). Waste temperatures decreased in lifts 1 and 2 and increased in waste temperature in lifts 3, 4, and 5 as a result of leachate injection. Average waste temperatures on the last day of leachate recirculation (9/5/03) were 10.9:06 °C, 12.5i2.4 °C, 10.1:42 °C, 10.4i4.1 °C, and 1.5:r4.0 °C, for lifts 1, 2, 3, 4, and 5, respectively. The waste temperature in lifts 1, 2, 3, and 4 was not statistically different after this period of leachate recirculation as determined by ANNOVA test (Table 4-2). The volume of leachate injected expressed as a percent waste volume for each lift were 8.1, 7.8, 7.4, 8.5, and 4.9 %, in lifts l, 2, 3, 4, and 5, respectively. Table 4-2: AN OVA test for waste temperature differences for lifts 1 through 4 after leachate recirculation (9/5/03) Sum of Mean . Squares df Square F 813' Between 32.48 3 10.8 Groups Wlthln 360.20 38 9.5 1.142 0.344 Groups Total 392.68 41 77 noun—3:8» 8382 meta—u 8356.58“ he EH a: some 5 053.6988 8mm? owflo>< ”Viv ensur— 658688 2383 53862 2383 Em - 1.. Base 6883 2283 ll _ 5' N a: .. m as . :3 x :E . 98H? (3) armatodura 1, 78 4.1.2. Effect of air injection on waste temperature During test 1, the waste temperature in lift 5 increased from ~O °C to more than 30 °C. In the remaining lifts 1, 2, 3 and 4 the temperature increased from 11.23209 °C to 17.8i2.3 °C (Figure 4-5). During air injection test 2, waste temperature increased from 14.5i1.2°C to 29.0:39 °C. The rate of temperature increase was calculated for a total of 26 air injection events that took place as part of air injection tests 1 and 2. Detailed description of these air injection events including their duration, start and end date, air flow rate and the location of the air injection line are summarized in Table 4-3. The rate of waste temperature increase was calculated as the slope of the linear trend line fitted to temperature data over the duration of each air injection event. An example of this calculation is shown in Figure 4-6 for two air injection events. The first air injection event shown in the figure was initiated on 7/2/2004 and continued for 16 days (air injection event 9 in Table 4-3). Three days later, the air injection resumed again (air injection event 10 in Table 4-3). The rate of temperature increase was calculated for the two events separately. The rate of waste temperature increase for event 9 was 0.28 °C/day and for event 10 was 0.11 °C/day. The average and median of all rates of temperature change in the bioreactor landfill during both air injection tests was 0.10:0.04 °C, and 0.05 °C/day. The average and median rates of temperature change were: (0.14:0.08, 0.09), (0.10:1:0.08, 0.05), (0.14:0.08, 0.07), (0.05:t0.05, 0.03), and (0.08:0.05, 0.04) °C/day, for lifts 1, 2, 3, 4 and 5, respectively. Table 4-4 displays the 26 air injection tests and the rate of temperature change calculated for each lift. 79 Table 4-3: Operational conditions for each air injection event Chronological Start End date Duration Injection Air flow rate numbering date (day) line (scfm) Injection through lift 1 1 9/5/2003 9/7/2003 2 G12 87 2 9/8/2003 9/16/2003 8 612 89 3 9/22/2003 9/29/2003 7 612 94 4 9/30/2003 10/2/2003 2 G12 97 6 5/25/2004 6/17/2004 23 L11 41 7 6/17/2004 6/25/2004 8 L12 65 8 6/25/2004 7/2/2004 7 L13 70 9 7/2/2004 7/18/2004 16 L14 56 10 7/20/2004 8/5/2004 16 L14 58 11 8/5/2004 8/18/2004 13 L13 67 12 8/19/2004 8/24/2004 5 L12 61 13 8/24/2004 9/2/2004 9 L12 55 14 9/2/2004 9/3/2004 1 L1 1 127 15 9/7/2004 9/8/2004 1 L1 1 1 16 16 9/9/2004 9/17/2004 8 L11 124 17 9/20/2004 9/23/2004 3 L1 1 134 1 8 9/24/2004 9/27/2004 3 L1 1 130 19 9/27/2004 10/14/2004 17 L12 49 Injection through lift 3 20 10/15/2004 10/22/2004 7 L31 122 21 10/22/2004 10/24/2004 2 L32 91 22 10/26/2004 10/31/2004 5 1.32 98 23 10/31/2004 1 1/1/2004 1 L33 82 24 11/8/2004 1 1/11/2004 3 L33 88 Injection through lift 4 5 10/2/2003 10/24/2003 22 G41 83 25 11/11/2004 12/2/2004 19 L42 94 26 12/2/2004 12/8/2004 8 L41 46 80 Table 4-4: Rate of temperature change for air injection events in each lift Rate of temperature change (°C/day) Chronological . Lift 1 Lift 2 Lift 3 Lift 4 Lift 5 numbering Injection through lift 1 1 0.13 0.13 0.00 0.01 0.00 2 0.24 0.21 0.15 0.00 0.00 3 0.44 0.58 0.10 0.04 0.01 4 0.37 0.60 0.21 0.02 0.02 6 0.18 0.22 0.15 0.00 0.01 7 0.07 0.12 0.03 0.00 0.03 8 0.04 0.40 0.00 -0.04 0.02 9 0.05 0.28 0.09 0.01 0.03 10 0.05 0.1 l 0.07 0.00 0.04 1 l 0.1 1 0.24 0.07 0.00 0.05 12 0.15 -0.11 0.01 0.05 0.04 13 0.14 -0.08 0.06 0.03 0.04 14 0.44 -0.14 0.00 0.00 0.03 15 0.47 -0.08 0.09 0.05 0.06 16 0.44 -0.03 0.08 0.02 0.05 17 0.47 -0.05 0.05 0.00 0.04 18 0.33 0.00 0.00 0.03 0.05 19 0.07 -0.02 0.13 0.03 0.05 Injection through lift 3 20 -0.10 0.04 0.41 0.03 0.00 21 -0.04 0.01 0.58 0.03 -0.03 22 -0.10 0.06 0.35 0.06 -0.23 23 -0.04 0.08 0.77 0.06 0.06 24 -0.10 0.11 0.07 0.05 0.06 Injection through lift 4 5 -0.06 -0. 10 0.07 0.60 1.56 25 -0.05 0.03 0.07 0.13 0.06 26 -0.05 0.04 0.00 0.15 0.07 81 Eggs 88385 830 805.52 5 ma: 2:.“ E ogfiomfiou 8m“? owmco>< “m4 95E."— (Q) armamdural m5.-le§t-lm§r 1.35-1-5: e88E8a2m1-|=88.§e< magi- M $63 ea: 82 82 8» Se 84 com o _ _ _ — 2 owl i ON! o 8 h 9. _ i 25 u 8 m e: N a: _ All a HWDH. Nash :fi cw 82 y=0.11x+ 26.27 *—_b— I i 6 l 2 V R = 0.98 i —*—+~we~~- e I | E 23 ,_______..___ —_——-— —— 1 1 e — ”_.__— _.__ 3 I y = 0.28X + 24.42 I I g - R2 = 0.99 J l 21 _g _L _.__ _Lw .___,__~ .2“ _ _ . _.___ _W M 3 ' | I 3’9 | g 19i——— —- ———~ --' I — e- 3: i | I | I 17 FT- .Wwv __ _._ .. . _ ‘ l l l 15 i——— r r 4.- r t i i 0 5 10 15 20 25 3O 35 40 Days from starting air injection F. [— — Airiry'ectionpaused o Event9 I Event 107‘ Figure 4-6: Temperature increase in waste during air injection in L14 for 1ift2 When air injection was conducted through lifts 3 and 4, waste temperature in the lower lifts decreased or remained constant. The decrease in waste temperature was due to net heat loss to the bottom of the landfill. When air injection was conducted through lift 1, the temperature did not change significantly in the upper lifts. During air injection test 1 air was initially injected through lift 1 and shifted to lift 4 after 27 days. The temperature change that corresponded in all lifts during air injection test 1 is shown in Figure 4-7. The start of air injection in lift 1 is shown by the vertical dotted line at day 420. Air 83 Temperature (C) 40 1 é-if: o - a...» - 4 F i'. au 1 i 5 I ‘1 y i: I I i' -5 i j . , l '10 7 l I 1 l 400 420 440 460 480 500 Time (days) ------ —AirTemp 1' Lift5 x Lifi4 A Lifi3 I Lift2 —-Lift1 i —-—Stop air iry'ection — - - Start air injection 17 - - -Switchtoupperliftsinjection Figure 4-7: Temperature change during air injection test 1 84 injection resulted in temperature increase in lifts 1, 2 and 3 at rates of 0.29i0.22 oC/day, 0.38:0.39 oC/day, and 0.15:0.14 oC/day, respectively (Figure 4-8). Lifts 4 and 5 had negligible increase in temperature during the same period. On day 447, indicated by the second dotted line in Figure 4-7, the air injection was shifted to lift 4. This shift in location of air injection resulted in an increase in waste temperature in lifts 3, 4 and 5 at an average rate of 0.07 °C/day, 0.6 °C/day, and 1.56 oC/day, respectively (Figure 4-8). This shift in location also resulted in a decrease of temperature in lifts 1 and at an average rate of -0.06 °C/day and -0.10 °C/day. This is due most likely to heat loss to the bottom of the bioreactor landfill cell. 2.0 1.51 Rate of temperature change (C/day) ES [3612 0.0 . [Tr—1 [3041 Lift] Lift2 Lift3 Lirt4 Lifts Figure 4-8: Rate of waste temperature increase during test 1 85 The temperature response during air injection test 2 is plotted in Figure 4-9 is a representation of the same picture occurred in test 1, however for test 2. The temperature increased in response to air injection through lift 1 at a rate of 0.21:0.10 °C/day, 0.23:0.11 °C/day, 0.08i0.03 °C/day, for lifts 1, 2, and 3 respectively (Figure 4-10). Temperature increase in lifts 4 and 5 was negligible, with a calculated rate of 0.03:0.01 °C/day and 004230.01 °C/day, respectively. During the air injection through lift 4, the waste temperature increased at a rate of 0.03i0.06 °C/day, 0.04:0.46 °C/day, 0.14:0.18 °C/day and 0.07i0.18 °C/day for lifts 2, 3, 4 and 5, respectively (Figure 4-11). The temperature in lifts 1 decreased at rate of -0.05i0.01 oC/day. When air injection was conducted through lift 3, the temperature increase in lifts 2, 3 and 4 increased at a rate of 0.06i0.05 °C/day, 0.43i0.32 °C/day, and 0.0532002 °C/day, respectively (Figure 4-1 1). 86 40 35 30 25 N 0 Temperature (C) G 10 5 0 -5 -10 650 700 750 800 850 Time (days) [ Kites... .. Lift 5 L :1: Lift 4 1 Lift 3 I Lift 2 —Lifl 1 — Stop air injection — - - Start air injection - :7-78witchto upper lifts injection 900 Figure 4-9: Temperature change during air injection test 2 87 45D 3' NEW :Ai E: swag bu misc? 3E3 N 58 macaw 080.83 EEEomEB 833 no 88 omSo>< 57v o.— 5:— (map/3) afiueqo ammrodwof, 30 91123 88 23B NED m8. «3” 5. V 9a n ma: Amsohu wage? 233 N “we macaw 83.8.: 0538983 883 we 88 owfio>< A _-v 0.59,.— SE- m .5..— v #5 m «E I .. do (flap/3) afiueqo armerodtuai 50 91123 89 4.1.3. Effect of moisture on rate of temperature increase Variation in moisture content was expected to be a major factor impacting the rate of temperature increase in response to air injection. Lifts 4 and 5 had lower moisture content during air injection test 1 than air injection test 2. In order to increase the moisture content in lifts 4 and 5 leachate was injected in the period between tests 1 and 2. The cumulative amount of leachate recirculated before air injection test 1 and air injection test 2 is shown in Figure 4-12. Waste with lower moisture content has lower heat capacity leading to higher increase in temperature. The resulted in enhanced rate of temperature increase in lifts 4 and 5 during air injection test 1 compared to that observed during test 2. The rates of waste temperature increase during test 1 for lifts 4 and 5 were 1.56 °C/day and 0.07:0.18 °C/day, respectively. The rate of temperature increase during test 2 was 0.60 °C/day and 0.14 °C/day for lifts 4 and 5, respectively (Figure 4-13). Lifts 1, 2 and 3 had the same moisture content during tests 1 and 2. As a result, the rate of waste temperature increase did not differ significantly between tests 1 and 2 (Figure 4- 14). Lifts 4 and 5 had lower moisture content in test 1 than test 2, thus leading to enhanced rate of waste temperature increase during air injection test 1. Moreover, lifts l, 2 and 3 had the same moisture content during tests 1 and 2; therefore the rate of waste temperature increase did not differ for these lifts. 9O 600,000 I 1 500,000 T —4 400,000 j 300,000 -._ 200,000 +1 100,000 “fl Leachate Injection (gal) 1 0 . . l First Second Third Forth Fifth Sixth . Lifts ! l D Befbre air im'ection 1 Before air injection 2 i i Figure 4-12: Comparison of the leachate injection volume between tests 1 and 2 2.0 3‘ '3 $3. 1.5l 0 DD 5 4: O 2 1.0. *3 a Q. E 8 «5 .5' 8 _ m .A M .Testl 0.0 .—, f—j—1 ‘ I “—1 [:]Test2 Lift2 Lifi3 -Lifl4 Lifts Figure 4-13: Comparison of rates of waste temperature increase between air injection testl and test 2 when injecting air through lift 4 91 9. g .61 B “33 .54 2 U .E .4 0 E g .3 d.) o. E. .2 “-4 o 2 .l .Testl 0.0 _ _ _ [:lTestZ = 4 14 4 6 4 14 Lifil Lift2 Lifi3 Figure 4-14: Comparison of the rate of waste temperature increase between test 1 and test 2 when injecting air through lift 1 4.1.4. Temperature response after air injection tests Waste temperature decreased after stopping air injection for both air injection tests. A decrease in the waste mass is attributed to injection of colder leachate than the waste mass (temperature of injected leachate was 9.8 °C), heat loss to the bottom of the bioreactor landfill, and heat conduction within the waste mass. Heat losses to the surface of the landfill were eliminated because waste temperatures in lifts 4 and 5 did not decrease when leachate recirculation was not taking place regardless of ambient temperature (Figure 4-15). Temperature (C) ‘ 0 50 100 150 200 250 "4 Air tenperature .2004 lift 5 . 2004 m" 4’ ‘ _l Figure 4-15: Waste and air temperature after stopping second air injection test 93 Average temperature on the day of stopping air injection test 1 was 19.5:t3.0 °C, within 213 days the temperature decreased to 14.5:3.7 °C. After air injection test 2 the temperature decreased from 28.2:3.6 °C to 22.1:22 °C after 232 days (Figure 4-16). The rate of the temperature change was studied on the temperature of the landfill as a whole, the temperature after both tests decreased, and the rate of decrease after tests 1 and 2 was calculated to be -0.017 °C/day and -0.029 °C/day, respectively (Figure 4-16). ‘ 35 yet -0.0171x #2152745 y =’ -0-0239X + 53-167 i L R2 = 0.8865 R2 = 0.9308 1 30 N M Mean temperature (C) N o ; wt : : ~ ~ ~~~——-——~ ' i I I I: a‘ 10 .W . W Y" 1 W ___T_________,__ .. . 400 600 800 1000 1200 1; Days i i 0 After air injection 2 I After air injection event 1]! i i- - - - Leachate injection — - - Smnp removed Figure 4-16: Average temperature in the landfill after stopping air injection tests 1 and 2 Moreover, the variability of the rate of waste temperature change was calculated for each individual location in the landfill to study if there is conduction in heat within the waste mass. The variability in the direction of temperature change in the waste indicated that 94 some of the heat conduction was taking place within the waste mass. Some of the measured locations had temperature increase while others showed a decrease in temperature. The largest decrease in temperature after air injection test 1 was observed for lifts 4 and 5 in which leachate recirculation was practiced (Figure 4-17). Waste temperature also decreased in the lift underneath the lift receiving leachate (Figure 4-18). 1 Inject 12,100 gal in lift 5 133011313 93°F“, , 1 45 injection inlifls __ injection In hit 4 fl _ 1 u 5 and 6 i ; 40 ,_ - -- ---— - -_.. I | I I I | . 35 | l I l Q I I I e 5 I J 5 I I I i i . 2% I E— - i 10 i l l l l ‘ hi I I I I ij—M_ _ —— I I I O LWW__. _W I I____ l 1 WI I -2 48 98 148 198 Days 1 IWWW WWW, _ ! L 1 o Lrit 4 A Lift 5 - - - Leachate 'njection period} Figure 4-17: Average temperature in lifts 4 and 5 after stopping air injection test 1 with leachate recirculation periods marked 95 “‘ 1Q 401“ — —_ #— —— #* I I ”__“w— E 30 _.__ — e I I ~ A“ e l I I E“ 20 k'h‘w —J B . - _ 8 10" I I a I l 3 O ‘T_‘—#" '_“ —T C“ ‘ l l +— 'T 1 0 50 100 150 200 250 Days from stopping air injection ‘ - - - - Leachate injection 136de in lift 4 l ! r—_._f22 __.. _._ Figure 4-18: Average temperature in lift 3 after stopping air injection test 1 with leachate recirculation period through lift 4 marked 96 4.1.5. Heat Generation Factor Input parameters and operational conditions The heat balance was conducted on the bioreactor landfill using data from 13 air injection events. Table 4-5 shows the operational conditions for each air injection event. These include the duration of air injection, air injection flow rate, average air temperature, average relative humidity, average waste temperature, and average oxygen concentration leaving the landfill during each air injection. Table 4-6 shows calculation results of input parameters that varied for each air injection. These include vapor pressure, oxygen consumption rate, and rate of temperature increase in the landfill. Vapor pressure at waste and air temperatures was obtained from the CRC handbook of chemistry and physics [110]. The amount of oxygen consumed was calculated by applying equation (3- 31). The rate of temperature increase was equal to the slope of temperature vs. time curve shown in Figure 4-19. Table 4-7 shows the constant input parameters for the heat balance equation. All input parameters shown in Table 4-7 were either measured or obtained from chemical and physical tables [110], except for the temperature underneath the landfill and the heat transfer coefficient. Discussion on estimation of these parameters is described below. 97 N.N fiNw Nde wNN md v.2 §N\N\N~ vooQ _ C _ _ N3 2 02 ".3. mwmdd v.mN dd 0.5 .3de (a 383R :4 Nd 9M: m.Nm ocNdd EMN m d Nd vooQNE XENENE N5 ~ ~ 03 93 chdd wNN 5m m voonQw 3dN\d:w N5 3 mad 04K w _ mod YNN Nd m— RENE :w RENEE m S a Nd_ mdc ENdd N. _N Nd m: 3dN\m\w vodeNR 34 m 0.3 ddh moNdd wd~ _d E 33% :5 vodNNR 34 N. E; wdc dedd 03 dd 5 vodNNR §N\mN\o m 3 c 92 QNN nomod 0.2 d d w 3oN\mN\c .3de to N: m 5.2 mNu 38d QE fiN MN vodNRCc 33ka :4 v Md ddh domed QB wd NN mdthsNBfi modN\N\o~ :5 m md v. S Nde 02 dd _.h mddNRNR mddNNNR N~O N 0.5 ".2. w :dd d: _d fiw mddQER mddNER N #0 fl Ems Gov Afib Shmv Gov “scum“ «a 333 8.5 3.5 35 5393.5 2522—88 3:55... 8a.. 23.22.88 net-559:8 562:9 dam 25m cowoo E a? :4 9523— ”a.“ 33>? 5»sz “2 Eogogm 385830 :05 can 36358 mm? ooze—an «no: 05 :02? 8m 3:26 8302.5 5w ”m... 035. 98 Table 4-6: Calculation results of input parameters that varied for each air injection Vapor pressure Vapor pressure 0 Rate of Air at arr at waste xygen temperature injection temperature temperature consumed increase (10'2atm) (10"atm) (“‘"WS’ (°C/day) 1 1.99 1.30 0.29 0.15 2 1.17 1.49 0.43 0.33 3 1.08 1.90 0.28 0.18 4 1.76 1.78 0.17 0.14 5 1.68 1.99 0.30 0.07 6 1.92 2.08 0.31 0.13 7 2.12 2.27 0.26 0.11 8 2.20 2.48 0.26 0.07 9 1.77 2.68 0.31 0.12 10 1.72 3.69 0.23 0.04 11 2.11 2.90 0.25 0.05 12 2.10 3.21 0.55 0.18 13 0.71 3.70 0.42 0.02 99 Table 4-7: Input parameters for the calculation of heat balance equation Parameter Value Unit The landfill footprint area (A) 4856 m2 Waste Density (P) 599.21 kg/m3 Waste volume (V) 41,330 m3 Specific heat of water 4.1818 J/ kg °C Specific heat of solid fraction of waste 1.2 J/ kg °C Specific heat of wet waste (Cp) 2,491 J/ kg °C Specific heat of dry air (Cpair) 1,030 J/ kg °C density of dry air 1.2929 kym3 R (gas constant) 0.082057 atm L/mol K Heat of vaporization for water 2.444667 kJ/g Pressure in the landfill 1 atm Heat transfer coefficient (U) 3.18E-04 kJ/(s m2 °C) Temperature beneath the landfill 11 °C Moisture content 43.3 % 100 25 _._... --——~ ---——— «- ‘ y=0.1221x- 70.604 R2 = 0.9965 i i N W Waste temperature (C) N N T— _— 754 756 758 760 762 764 766 768 770 Days Figure 4-19: Determination of the rate of temperature increase in the bioreactor landfill for air injection event 9 Temperature of groundwater in Michigan was used to represent the temperature of soil underneath the landfill, because the temperature of soil underneath the landfill could not be measured directly. Groundwater nearby the surface is at constant temperature during the year [124]. The average shallow groundwater temperature in Michigan is 11 °C [125]. A study conducted in Oakland County in Michigan concluded that groundwater temperature samples ranged from 10.4 °C to 15.5 °C, with a mean of 12 °C [126]. The temperature of ground water is usually 1 to 2 oC higher than the mean annual air temperature [127]. The mean air temperature in Harrison in 1993 and 1994 was 6.7 °C 101 as measured at the weather station in Northern Oaks bioreactor landfill. This results in a groundwater temperature of about 8 °C as described by Todd [127]. As a choice of input parameter, temperature of 11 °C was used to represent the soil temperature underneath the landfill. Heat transfer coefficient was calculated for the bioreactor landfill by conducting heat balance after the second air injection test was stopped. During this period neither air nor leachate were recirculated in the landfill. Therefore the value of latent, sensible, and metabolic heat can be taken as zero. Substituting in equation (3-22), the heat balance equation after the second air injection test is given by equation (4-2). pCpV % (T) = UA(T0u, — Twang) ............................................................................. (4-2) Rearranging, equation (4-2), we get: dT = pVCp /dt A(T out ’ T waste) ................................................................................................ (4-3) Where, dT/dt is the rate of temperature decrease in the landfill after the second air injection test and equals 0.0289 °C/day given by Figure 4-16, and Twaste average waste temperature during this period equals to 24.36 °C. 102 The computed value of heat transfer coefficient was 3.18 10'4 kJ/(s m2 °C). Heat balance analysis Dynamic calorimetric method was used to calculate the heat generation factor of waste in this study. This study represents the first reported effort to use landfill as a calorimetry to determine heat generation factor. The only other study reporting the use of larger size reactors than laboratory scale reactors, used a size of 21.4 m3 in 1856, 150 years ago [128] cited in [55]. Another study indicated the use of commercial size to determine heat generation, but they did not specify reactors making the comparison with difficult [55]. The size 21.4 m3 is 1,931 times less than the bioreactor cell. Field scale evaluation of heat generation factor are more valuable than those obtained using lab scale reactors, because it reflects the amount of heat generated under field conditions. The results of the heat balance analysis for each air injection test are presented in Table 4-8. This table includes all five terms in the heat balance equation and the value obtained for the heat generation factor. Heat generation factor ranged from 200 kJ/mole 02 to 626 kJ/mole 02 with an average of 354188 kJ/mole Oz. The range observed in this study is larger than the range observed for single substrates under laboratory conditions (397 to 543 kJ/mole 02) [55]. The range of heat generation factor calculated for Northern Oaks bioreactor landfill was within the same range of that observed for single substrates under laboratory conditions (Table 4-9). To perform this analysis, 29 reported heat generation factors were obtained from a study that compiled the heat generation factor in literature [55] and a t-test was conducted to compare these values with those calculated in this 103 study (4-10). A t-test indicated that the average obtained under field conditions for solid waste is significantly less than that obtained under laboratory conditions for single substrates. Although the heat generation factor for single substrates (29 measurements) was found to be high compared to the heat generation factor values determined in the bioreactor landfill, many more field studies measurements spread over different bioreactor landfill must be performed to conclude that single substrate values are always higher than complex substrate measurements. Indeed, in one study the heat generation factor for a synthetic solid waste was determined under laboratory conditions to be 304 kJ/mole [24]. If the value determined for solid waste under field conditions is truly lower than that observed for single substrates then the heat generation factor of single substrates under laboratory conditions, then using the single substrate valued would be an overestimation of the amount of heat produced for each mole of oxygen consumed. Studies that have focused on heat transfer and modeling in landfill environments have used values for single substrates [23, 58]. 104 Table 4-8: Heat loss, production and accumulation in the bioreactor landfill Ai -HGF Heat Metabolic Sensible Heattoloss L133“ injectrion (ET/mole) accumulation beat heat due boundary loss (kJ/s) (kJ/s) (kJ/s) (kJ/s) (kJ/s) 1 370 109 109 0.36 -0.12 0.16 2 559 237 241 -0.23 -3.60 —0.46 3 490 127 137 -O.44 -9.07 -0.80 4 627 99 107 -0.01 -7.58 -O.20 5 200 48 59 -O.1 1 -10.27 -O.46 6 328 89 101 -0.05 -11.28 -O.52 7 349 76 89 -0.04 -13.54 -0.31 8 237 47 63 -0.07 -15.71 -O.52 9 345 87 106 -0.28 -17 .65 -0.86 10 245 29 57 -0.48 -25 .96 -1.40 11 220 34 55 -O.18 -19.71 -0.59 12 275 126 150 -O.54 -22.30 -1.88 13 191 13 81 -1.51 -26.02 -2.54 (-3736‘) * Sensible heat loss due to leachate recirculation 105 Table 4—9: Average heat generation factor Std. Heat generation factor Std. (kJ/mole) N Mean Deviation Error Mean Literature data for single ‘ substrates [55] 29 445.6 37.2 6.9 Synthetic solid waste [24] 1 304 This study 13 341.2 139.6 38.7 Table 4-10: Comparison between laboratory values for heat generation factor in laboratory (single substrates) and field scale (solid waste). Si (2_ Mean Std. 95% Confidence t (if “Sic d) Difference Error Interval of the Difference Difference Lower Upper 2.65 12.8 0.02 104.3 39.34 19.2 189.5 Table 4-11 shows total heat loss and heat gain and the percent contribution of each term in the heat balance equation. Heat accumulation ranged from 51 % to 100 % of the produced heat when no leachate recirculation was taking place. With leachate recirculation, the heat accumulation was 17 % of the total heat produced. This indicates the strong effect of leachate recirculation on the temperature of the waste, particularly when the leachate temperature is low. The temperature of the recirculated leachate was 10 oC. Fifty percent of the heat loss during leachate recirculation was sensible heat loss. 106 The above data indicate that leachate recirculation can be used as a temperature control mean to minimize fire hazards during air injection. Sensible heat due to air injection was minimal and ranged between zero percent and 5.3 %. This term is much lower than found in commercial size reactors, where 9-12 % of the total heat loss was sensible heat [58]. The latent heat is usually assumed to be the biggest mechanism of heat loss in larger reactors, reaching 80 % of the total heat loss [58]. In this study it ranged between 0 % and 10.7 %. This low percentage of heat loss can be attributed to the high relative humidity in the injected air. Heat balance studies on relatively big reactors with high volume to surface ratio, usually ignore the heat loss to the boundaries [24, 55, 58]. In this study, it was found that the heat loss to the bottom of the landfill was significant. Under no leachate recirculation conditions the heat loss to the boundaries ranged from 84 % to 100 % of the total heat loss (Table 4-11). 107 8.3388. 228.... e 8.. a... .8. ”3.80m .. Came N.N w.m 9mm 2 E #60. 2 N.N 0.5 Nda V» 02 WVN- N— md QN Ném no mm WON- u w NA 06 m.m¢ _m hm wfim- A: WA 9? ada mm 9: w.w~- m v.0 N.m #60 E. mw md—i w Md m.N 155 Va aw Q2- 5 m6 vi Nda mm 2: m.:- @ o4 m6 Eva mm mm wd—t m fio WN vfia ma 2: QB. v m4... mg. mfim Na wm— WOT m Md 5A: 64% ma ~VN m4». N Cd od odo— co“ 9: fic- m Ae\ev AX; ax; 68.69... 3...: 3...: .é a... .88 a...» .8 5.8.3. m8. .3.. 2.3.5 b.652— ue 2.3.2— 2. . _ .._< .3 o .28 e mm: a .— 3. m a ‘— 23.. «no: «no: 132......83. an... .38 he 2.8.8: a as mne— «no: nouns—co 02.2.3 .8.— ofi mo 8.8. some .«0 8:35.80 2.8.3 on. v...» 5mm EB 32 .3: :33. u. 7.. 033—. 108 4.1.6. Change in oxygen, nitrogen, carbon dioxide and methane concentration due to air injection The gas phase in anaerobic bioreactor landfill is dominated by carbon dioxide and methane. Air injection into the bioreactor landfill leads to the change in this gas phase composition at the affected locations. Figure 4-20 is an example of gas phase concentration change at a single sampling port in the landfill during the air injection experiments. Air injection leads to an increased oxygen and nitrogen concentrations and decreased methane concentrations in the landfill. The injected air penetrates into the landfill pores, replacing both carbon dioxide and methane. Over time, it reaches to a relatively stable condition, at which change in concentrations is minimal as long as air injection is still in operation. The process of replacement is very quick; within a few hours the concentration of methane and carbon dioxide decreases and the oxygen and nitrogen concentration increases. Initially the rate of change is faster which slows down gradually. The time required to reach the steady state condition was not monitored closely, however it was seen that after 4 days of aeration, the concentration of nitrogen reaches a stable value. The concentration of the gas components after 4 days of aeration was similar to that after 11 days. Figures 4-20 and 4-21 show specific examples of this process at a location that is 20 ft away from the air injection line. Figure 4-21 shows that 40 minutes prior to the start of air injection, the concentration of carbon dioxide, methane, nitrogen, and oxygen were about 65 %, 30 %, 10 % and zero, respectively. Within 5 hours of starting air injection the oxygen and nitrogen increased to about 10 % and 40 %, while methane and carbon dioxide concentrations were reduced to 10 % and 109 40 % (Figure 4-21). As time progressed (Figure 4-24) these concentration changed further, towards higher oxygen and nitrogen concentrations, reaching a state of ‘steady state’, where change became minimal. The close to zero methane concentration and 5% carbon dioxide concentration indicates that flushing of the gas phase was taking place. Most of the methane produced and accumulated in the locations exposed to air was flushed out by the injected air. Other ports (15) exposed to air showed similar. Time (days) {—9-CH4 +c02 +02 +N2 i 80 1 70 1 A60 ~ @150 . 8 2:40 -_ E 830 c: 152° 4><' 1 1. N 1 \- ' O l I a __.-———— i 0 l 2 3 4 11 1 1 Figure 4-20: Change in gas phase concentration during air injection at location G211 110 OIL 0U 1,. “I: :79 ‘rStartairirjection _ N2 _-_. 13 1. 4414—4 — :§ “_50 (:02 fill—_l 4 .. 4 >— 15 - 4 + . o : G O I) 0.00 0. 04 0.09 0.14 0.19 0.24 Time (days) 1 15.79114 +C02 +02 +102] Figure 4-21 Change in gas phase concentration during the initial hours of air injection at location G211 4.1.7. Impact of air injection on leachate characteristics Laboratory and field scale bioreactor landfills have demonstrated that leachate can be treated by recirculation. The process of cleaning leachate aerobically is faster than that of anaerobic process at optimum conditions, because of the faster degradation rates. At Northern Oaks bioreactor landfill, the pH, COD and VOC of leachate was studied within the waste using sampling ports and at the sump. Leachate samples from the sump were collected monthly, but from the sampling ports it was collected less frequently for analysis. Leachate samples from sampling ports were collected every three months and before and after air injection. However, during the air injection test 2, leachate samples were collected on a weekly basis. Leachate sampling from some of the ports was not 111 possible due to lack of sufficient leachate in the sampling basin. On an average only 20 to 35 ports were usable out of the 48 ports. Before the air injection experiments, very high concentrations of COD and VOC were observed that was accompanied by low pH in all ports. After air injection, COD and VOC dropped rapidly and pH increased. Figure 4-22 shows the average concentration of COD in lift 1 for 13 sampling ports before and after the two air injection tests (i.e., before air injection test 1 and after air injection test 2). The average COD concentration for lift 1 before air injection test 1 was seven times higher than that after air injection test 2, a decrease from 21 g/L to 2.9 g/L. Similar decrease in COD occurred in lift 3 (Figure 4-23; three separate ports). The third port did not have enough leachate in the beginning. The first sample obtained was 20 days after air injection, which was ~ 23 g/L. It decreased to 3 g/L after air injection. More than 90% reduction in leachate COD was observed at most ports. The impact of air injection and decrease in COD on pH was expected. Figure 4-24 depicts the correlation between pH and COD before and after air injection. Before air injection, the pH was in the range of 5 .3 to 6.3, and the average COD was ~34 g/L. After air injection, the pH was in the range of 7.2 to 8.3, while average COD was about 3.2 g/L. VOC concentrations were also measured in leachate, although less frequently than other parameters. The following VOCs were detected in the leachate: acetone, methyl chloride, methyl ethyl ketone (MEK), methyl isobutyl ketone (MIK), toluene, and xylene. MEK and acetone had the highest concentrations and methyl chloride and xylene had the lowest concentrations among all the measured VOCs. Table 4-12 lists the VOCs and 112 their concentrations for two sampling events, one before the air injection and one after. As with COD, the concentration of VOCs reduced significantly after air injection. However, the sampling events for VOCs also included the impact of losses due to volatilization and aeration over a period of one year. During this period, air injection was in place and lift 1 was exposed to air for more than five months. Table 4-12: Average concentrations of VOC in lift 1 Methyl chloride MEK MIK Toluene Xylene (Hg/L) (pg/L) (113/L) (us/L) (pg/L) (Hg/L) Acetone 9/2/2003 24,681i13,685 79i87 34,948i26,555 522:1:498 108:1:60 41:4 10/15/2004 1,252i1,670 0 301-59 5:17 31i26 4i5 1 40 : 4444 ----- « 1 ’ 4 444 4 4 -— COD before air injection —444. .. 1 I 241- 31 - 444— coo afterairinejction / - Miriam a 2.897 b) (It Thousands N b.) u. o p—n M Average COD (mg/L) N o —v ...-A OUIO I K V 650 700 750 800 850 900 950 1000 Figure 4-22: COD concentration before and alter air injection in location lift 1 113 A -8 60 I A "*fl i E g 50 " ‘“’ Y 1 V .1: N A g 30 1— -— 4 - __.. 0 g 20 _l__ “_.__fi __ __ a | O 10 7*“; -—-—--- - .3 U | K; 0 “i“ ‘ ' 1 7 1 1 1 1 Figure 4-23: COD concentrations decrease during air injection in lift 3. 100,000 444 20,000 .1: ‘ i: L o e 16000.5 ,2 A ‘ —¢-—¥—- 34,387 ; 4 g . ..D 0 fi ' O a 3 g g 1 9 412,000,, g g g 10,000 1 E g . a . § 1,; 1 4 8,000 ,3; g .. 1 Q 1 I "f 4,000 O 8 ‘ F. 3,180 8 1,000 4.4—4444 4 - 4 4 - 0 5.00 6.00 7.00 8.00 9.00 pH *7 Beforegfiijfiectiofi ‘ __ Average 9 I _ I Aflerairinjecfion _ l, Figure 4-24: COD concentration vs. pH before and after air injection in sampling location W111 — Lift 1 114 Figures 4-25 and 4-26 show the distribution of acetone concentrations within lift 1. Figures 4-25 show that the concentration before air injection (2003) was in the range of 7-42 mg/L, with an average of 25 mg/L. In 2004, acetone concentration decreased to below 2 mg/L (Figure 4-26). It was interesting to note that the port at which acetone was still present after one year, were the locations where it was low in 2003 indicating the role of air injection as a major mechanism of removal. The role of air injection as a major mechanism of removal was studied by calculating the resistance to mass transfer from the liquid and gas phase for the VOCS found in leachate in Northern Oaks bioreactor landfill To determine which phase controls the mass transfer, the following equation was used [129]: gim) = 1001 ...................................................... , .................................... (44) T 1 + (kg / k1 j; Where, RL is the liquid phase resistance to mass transfer, RT is the overall resistance to mass transfer, k3 is the gas-phase mass transfer coefficient that describes the rate at which contaminant A is transferred from the air—water interface to the bulk gas phase, 115 k1 is the liquid-phase mass transfer coefficient that describes the rate at which contaminant A is transferred from the bulk aqueous phase to the air-water interface, and H is Henry’s constant in (Lliquid/LAjr ). The ratio (kg/k1) ranges between 40 to 200 depending upon the type of aeration or stripping system used [129]. For a ratio (kg/kl) of 100, RL/RT was calculated by equation (4-4), using Henry’s constants given in Table 4-13. The results of the calculation are shown in Table 4-14. For MEK and MIK the mass transfer is controlled by the gas phase only. For methylene chloride, toluene, and xylene the mass transfer is controlled by both phases, but predominantly by the gas phase. With temperature increase the contribution of resistance from liquid phase becomes closer to that of the gas phase. For acetone mass transfer is controlled mainly by the gas phase, as temperature increase the contribution of the liquid phase increases slightly. The volume of injected air was 17 times the volume of the waste. All of theses chemicals are easily volatile and some are also easily biodegradable such as acetone. Therefore, it is possible that all chemicals were volatilized before experiencing any significant degradation however for acetone a combined mechanism for loss is more likely. 116 2604.00 42000 36760 31500 2556.00 26250 21000 15750 10500 5250 2828.001111] 0 200.0000] 2506.00 2534.00 2502.00 2590.00 Figure 4-25: Acetone concentrations before air injection in lift 1 (9/2/2003) Between 2,600 and 5600 jig/L 2604.00 + + + 2656.00 4‘ 4» Less than 2,600 ng/L + 252000011] 2478.00[m] 2508.00 2534.00 2562.00 Figure 4- 26: Acetone concentrations after air injection in lift 1 (10/15/2004) 117 Table 4-13: Henry’s constant for VOCs found in leachate samples from Northern Oaks bioreactor landfill [129] Henry’s constant Name (Lamar/Lair) Temperature 10°C 20 °C 25 °C 30°C Acetone‘ 8.4310“4 1.3010“3 1.5910‘3 1.9110‘3 MEK 2.80104 1.9010‘l 1.30104 1.1010’4 MIK 6.60104 2.90104 3.9010‘4 6.8010'4 Methylene chloride 1.4010"3 2.4410”3 2.9610'3 3.6110“3 Toluene 3.8110‘3 5.5510‘3 6.4210‘3 8.0810’3 -3 -3 -3 -3 m-Xylene 4.11 10 5.98 10 7.4410 8.87 10 -3 -3 -3 -3 p-Xylene 4.2010 6.45 10 7.4410 9.45 10 -3 -3 -3 -3 o-Xylene 2.8510 4.74 10 4.8710 6.2610 *Source for acetone data: http://www.epa.gov/athens/learn2model/part- two/onsite/esthenry.htm 118 Table 4-14 Ratio of RL/RT VOCs found in leachate samples from Northern Oaks bioreactor landfill for kg/kl=100: Name RL/RT (%) Temperature 10 °C 20 °C 25 °C 30 °C Acetone 8 12 14 16 MEK 3 2 1 1 MIK 6 3 4 6 Methylene chloride 12 20 23 27 Toluene 28 36 39 45 m-xylene 29 37 43 47 p-xylene 30 39 43 49 o-xylene 22 32 33 38 4.2. Oxygen Consumption Rate Attaining optimal temperature of waste is an essential requirement for enhanced biodegradation. This is especially true for colder climates when the waste is placed at lower ambient temperatures. The fastest and most economical method to increase temperature in a landfill is by air injection. Air injection creates aerobic conditions in the waste resulting in exothermic reactions. The rate of aerobic microbial degradation affects the size of the aerobic zone for a given air injection flow rate. Higher oxygen consumption rates requires higher air flow rate. To properly design an air injection system, the rate of oxygen consumption must be known. This is accomplished in 119 laboratory scale reactors under highly controlled conditions. Such conditions cannot be achieved in a bioreactor cell. The difference in conditions and scale may lead to difference in the oxygen consumption rate, because of heterogeneity of waste. In this section oxygen consumption rate inside the landfill and at vents are described. The effect of background oxygen concentration, moisture content, and temperature on this rate is discussed in this section. 4.2.1. Operational conditions Oxygen vs. time curves were obtained at the sampling locations twice, separated by 7 days of cumulative air injection period during the year of 2003. The two tests were apart by 7 days of active air injection. The sampling locations at which POCR was determined in the first and second tests are shown in Tables 4-15 and 4-16, respectively. These tables describe the oxygen concentration before stopping air injection, moisture content, and waste temperature during the period of POCR determination for each location. The temperature at locations in Test 1 ranged between 3.7 and 24.7 °C, volumetric moisture content ranged between 0.31 and 0.65, and starting oxygen concentration ranged between 4.4 and 19.8%. The ports sampled in the second test had a range of temperature, moisture content and starting oxygen concentrations of 10.5-51. 0°C, 0.35-0.66, and 2.2-20.3%, respectively (Table 4-16). Together, these experiments provided the data for studying the effect of temperature, moisture and starting oxygen concentration on POCR. Electronic Conductivity in Leachate samples was measured before and after the air injection, an average of the two samples was used. At locations were leachate samples 120 were not obtained, the average of the leachate value for the lift from which the location was measured was used. Table 4-15: Sampling locations, temperature, moisture content and oxygen concentration before stopping air injection at which POCR was conducted (Test 1) Average Location temperature Moisture Starting oxygen content concentration (%) (° C) 9/16/03 G132 14.8 0.46 4.4 G143 16.7 0.53 12.5 G211 16.7 0.64 17.8 G213 14.0 0.60 13.5 G214 13.3 0.65 9.1 G221 23.7 0.45 19.7 G222 13.4 0.60 17.9 6223 18.1 0.60 16.7 G224 20.1 0.60 14.9 G231 11.9 0.64 19.8 G321 11.6 0.46 12.2 G322 10.5 0.57 12.4 10/17/03 G512 6.2 0.31 19.6 G521 3.7 0.48 19.2 G522 7.4 0.51 18.7 121 Table 4-16: Sampling locations, temperature, moisture content and oxygen concentration before stopping air injection at which POCR was conducted (Test 2) Location Average temperature Moisture content Starting oxygen (°C) concentration (%) 9/29/03 G123 10.5 0.52 2.2 G213 15.8 0.54 11.5 0214 16.7 0.66 8.0 G221 51.0 0.54 20.3 G222 20.3 0.54 10.0 G223 21.2 0.54 15.4 G224 24.8 0.54 13.1 G225 19.7 0.54 2.7 G231 15.6 0.65 20 G233 15.2 0.53 8.9 G321 15.8 0.45 1.6 G322 12.4 0.57 7.2 10/24/03 G512 39.2 0.35 13.9 G522 18.4 0.49 10.8 4.2.2. Fitting POCR data to Monod equation Liquid phase oxygen concentrations over a range of approximately 8 hours were fitted to Monod equation. An example of the fitted values for sampling location G221 is given below. For this sampling location, the parameters estimated for k and Ks with 95% confidence interval were 4.76:0.02 per h, and 4.43i0.0003 mg/L, respectively. Initial substrate concentration was estimated to be 3.22i0.01 mg/L. Concentration in the gas phase for this sampling location ranged between 0.1 and 21%. This range is the 122 maximum range that can be obtained when injecting air. A fit was obtained (Figure 4-27) with the minimization function of a value f=0.0047 (equation 3—42). A wide range of initial values were used to avoid local minima, forty iterations performed using different initial values of k, Ks and So. 3.5: b.) T- 1" Ln Oxygen concentration (mg/L) Figure 4-27 Fitting Monod equation to the experimental data for location G221 123 The validity of these parameters in comparison to the requirements for their determination is presented below. In order to obtain a unique solution the ratio So/Ks should be bigger than 0.1, this requirement was met as the ratio obtained was 0.68. This is supported by the results of the 40 iterations performed which gave standard error for k and Ks of 0.5% and 0.007%, respectively. In order to obtain independent coefficients, the ratio So/Ks should be larger than 1. This requirement was not met as the ratio calculated was 0.678. In order to obtain valid estimates of parameters, the range of the substrate concentrations should in the three regions defined by Monod equation as described (Table 4-17). Table 4-17: Comparison between the range of data used and the required to estimate Monod kinetics parameters Range of data measured Range of data required Minimum Maximum RegionI RegionII Region III s/ks 0.0370 1 .00 0-2 2-9 >9 V/k 0.0357 0.50 0—0.67 0.67-0.9 >0.9 For sampling location 6221, the rate of oxygen consumption was calculated based on the estimated kinetic coefficients using Monod equation (3-41). This rate is plotted as a function of substrate concentration in Figure 4-28. In this figure the theoretical rate is extended to cover the three regions as described above. The first region is located at concentration range from 0 -9mg/L (shown by the first dotted line and labeled as region I in which 0 < S/KS $2). The second region is in the range between 9mg/L and 40mg/L (2 < S/Ks S9). The third region is at concentrations higher than 40mg/L corresponding to 124 (S/Ks 29). The oxygen concentrations measured in the field were all located within the first region located between 0-4.4 mg/L. From this analysis we conclude that the requirement to cover the three regions in determining Monod equation coefficients was not met (Table 4-17). 1 Region] _*fi _ fi_ __— __— 4 4 _ ._.._--_._.__ l ————————— I I I I I I I I T— . FH‘” ' ' n FT 9 " --..-- _ I -f _~ 1 n. . _ ' I l 1!____. _ __ r ____ , ‘—— -ul—-”” 1 . i . : Reglon II I Reglon III . l i _ -____.|__ __. 1 4 Oxygen =21% 4W4" 44—414444 >44 4 i , -_ l __ i 20 30 40 50 60 ' 3 Oxygen Concentration (mg/L) Figure 4-28 Monod equation based on the estimated parameters for location 6221, T = 50 °C, y=l.15 In conclusion Monod equation cannot be used to model POCR data. Although the measured data appears to fit Monod equation (Figure 4-27), the requirements to use this model were not met for to be applicable. Moreover, the range of data required to obtain a valid fit for the data is impossible because this range extends beyond the saturated oxygen concentration. The parameters estimated are dependent, and located within the region Monod equation reduces to first order kinetics. 125 4.2.3. Fitting first order kinetics to POCR data POCR at all the locations were modeled successfully using first order kinetics. An example of the fit is shown in Figure 4-29 for location 6221. In each test POCR was determined for 15 locations within the landfill, of which 12 locations were studied in both tests 1 and 2. The first order kinetic for this location is 0.51 per h with R2 of 0.99. First order POCR was calculated for tests 1 and 2 (Table 4-18). The mean POCR of the 15 sampling locations for tests 1 and 2 were 0.0867i0.0422 per h, and 0.2450i0.0843 per h, respectively. To evaluate the POCR difference between the two tests Wilcoxon signed ranks test was used. This test was used because POCR in test 1 was not normally distributed (Table 4-19). The result of this test indicates that there is a significant difference between the median of the two tests (Table 4-20). In Wilcoxon signed ranks test, median is a better representation of the data than mean. The medians of POCR in tests 1 and 2 were 0.041 (per hr) and 0.23 (per h), respectively (Figure 4-30). 126 Table 4418: First Order POCR First order POCR First order POCR Location Lift number for Test 1 for Test 2 (per h) (per h) 1 6123 Lift 1 0.106 2 6132 Lift 1 0.201 3 6143 Lift 1 0.250 4 6211 Lift 2 0.197 5 6213 Lift 2 0.017 0.038 6 6214 Lift 2 0.058 0.201 7 6221 Lift 2 0.046 0.510 8 6222 Lift 2 0.039 0.399 9 6223 Lift 2 0.022 0.219 10 6224 Lift 2 0.044 0.123 11 6225 Lift 2 0.146 12 6231 Lift 2 0.044 0.125 13 6233 Lift 2 0.206 14 6321 Lift 3 0.049 0.376 15 6322 Lift 3 0.041 0.122 16 6512 Lift 5 0.167 0.331 17 6521 Lift 5 0.042 0.238 18 6522 Lift 5 0.086 0.537 Number 15 15 127 Table 4-19: Tests of Normality for POCR, temperature, moisture content and background oxygen concentration Variable Kolmogorov- Smirnov Shapiro-Wilk D df Sig. w df Sig. First order POCR(testl) 0.303 12 0.003‘ 0.706 12 0001" First order POCR(test 2) 0.158 12 0.200 0.939 12 0.489 Temperature (test 1) 0.166 12 0.200 0.972 12 0.935 Temperature(test2) 0.262 12 0.022‘ 0.813 12 0.013‘ BaCkgm‘md °"t§f‘l‘)°°“°ent’ati°n( 0.174 12 0.200 0.882 12 0.075 Background oxygen concentration( 0 101 12 0 200 0 972 12 0 913 test Q Moisture content (test 1) 0.227 12 0.088 0.876 12 0.079 Moisture content (test 2) -.233 12 0.072 0.925 12 0.334 * Significant (i.e. does not satisfy normality assumption) 1 1 1 A ‘41 1 '3}, 1 1 E 1 1 g _. 1 1 -a - x 1 E y=4.8745e0'51 _ 1 1 4g 1" R2=0.9916 _1 1 8 1“ “"1 1 1 =1 1 1 1 1 86 1 -1 . .-.--..-.. __-__.__.1 1 O 0 . , 1 1 1 0 2 4 6 8 1 1 1 Time (hr) 1 1 1 Figure 4-29 First Order POCR for location 6221 128 Table 4-20: Wilcoxon Signed Ranks Test for comparison between POCR tests 1 and 2 Tests value First order POCR Temperature Z -3.059 -3.059 Asymptotic Significance (2-tailed) 0.002 ' 0.002 . * Significant at 95% confidence interval .6 A .51 {—4 3 O r: .41 13 81 at .31 O 0 fi .21 Q) * “U lI-l 3 1‘ 0 LL 0.01 -.1 _ _ N= 12 12 Testl Test2 Figure 4-30: First order POCR for two tests separated by 7 days of air injection 4.2.4. Factors affecting POCR In order to analyze the reason for difference in POCR between the two tests, differences and similarities for initial oxygen concentration, moisture content and temperature are described below. Wilcoxon signed ranks test was used to test the difference in 129 temperature. This test was used because temperature data during the second POCR test was not normally distributed during the second test (Table 4-19). The median temperature during test 1 was 12.6 °C and during test 2 was 17.5 °C (Figure 4-31). The difference in median was significant between the two tests (Table 4-20). Paired t—test was used to study the difference of background oxygen concentration and volumetric moisture content because they were normally distributed (Table 4-19). The volumetric moisture content did not differ between the two tests (Table 4-21). The average volumetric moisture content was 0.54:0.06 during test 1 and 0.53:0.05 during test 2 (Figure 4-32). Background oxygen concentration differed significantly between the two tests (Table 4421). Average background oxygen concentration of test 1 was 15.282t2.82% and 11.96i3.12% for test 2 (Figure 4-33). 60 504 * 6 40‘ o d) 5 e 301 O) E‘ —— _.— g 201 101 0 w I N: 12 12 Testl Test2 Figure 4-31: Temperature during the two tests separated by 7 days of air injection 130 Table 4-21: Paired samples t-test for comparison between tests 1 and 2 Tests value Moisture content Background 0.“ gen concentration 1 0.747 2930 df 11 12 Significance .. . (2-tailed) 94“ 0.013 * Significant at 95% confidence interval, ** Not significant .7 § .61 c: o O O 5 .51 .2 O E O E .4 ’8’ .2 9 31 _— .2 - ' N= 12 12 Testl Test2 Figure 4-32: Volumetric moisture content during the two tests separated by 7 days of air injection 131 g 30 0 8 J: 0.. 26 00 o 201 __.— —— E. .E C .2 *6 E 8 10‘ _L__ 1:: O Q t: & >~ __L_ 5 0 N= 12 12 Test 1 Test 2 Figure 4-33: Oxygen concentration before turning off the air injection Although POCR did not correlate with temperature, oxygen concentration or moisture content, the difference in POCR between test 1 and test 2 did correlate with difference in temperature (Figure 4-34). Except one point (circled), the increase in temperature correlated well with POCR increase. Most of the increase in POCR was for temperature increase up to 12 °C. Additional temperature increase from 12 °C to 27 °C did not give significant additional increase in POCR. The outlier (circled) has an average moisture content of 0.35, while all other locations had a moisture content of 0.55:0.04. We believe that this lower moisture content is the reason for the low increase in POCR. Because there was a large temperature increase (33 °C) at this location even with 0.35 mositure content, activity of microorganisms was not limited. Hence the reason for lower 132 increase in POCR maybe related to either smaller heat capacity of waste at lower moisture content, or to larger void volume of lower moisture content. N .5 § . x 'U g .4. § X a D 8 .3- E O ..D a: 8 .2' x1> . m Lffi nUIanr .S x 8 1 D Lift 5 § - 0 xx APOCR = 0.127371n(AT)—0.0491 0 Lift 3 g R =o.731 D 0.0 X _ - _ >< Lift2 o 10 20 30 40 Difference in temperature between test 1 and test 2 Figure 4-34: POCR Difference between of tests 1 and 2 vs. their temperature difference 4.2.5. Comparison between POCR and OCR from the landfill Oxygen consumption rate was also determined at the vents at the time of POCR were measured in the landfill. The oxygen concentration in the vents decreased following first order kinetics (Figure 4-35). OCR values are shown as vertical lines for different vents for tests 1 and 2 in Figures 4-36 and 4-37. The horizontal lines represent POCR measured inside the landfill in terms of mean with 95% confidence interval and range. 133 For test 1 vent L32 had a value above the range measured in the landfill, while for test 2, OCR values at the vents were within same range of POCR determined inside the landfill. Moreover, OCR was not measurable at one vent because oxygen concentrations at the vent were undetectable. These measurements demonstrate that measuring POCR is more accurate representation for aerobic activity compared to OCR determined at the vents. From this limited data set, it is not possible to determine the relationship between OCR at the vents and POCR inside the landfill. Si . : ~ _.__ w_~ i .8 ; E ______ C.‘ l s . t 8 . - - __..— ______ 5 ° . OD - _ _ _ >3 X l o +_ o 1 2 3 4 5 6 J w Time (hr) Figure 4-35: Oxygen consumption rate at vent L6 (Date: 10/24/03) 134 I LIT i A “-12:35020 Confidence interval for Mean 7 I F O W a 0.2909 #7 I I «r 0.201 , I 0.3593 I I min I I rmx I I x i / 4 L32 ( Lifl 5) I i / I I I L6 ’7 - I I I L32 (hfts 1,2 &3) I 0 0.1 0.2 0.3 0.4 0.5 0.6 I Rate of oxygen consumption during test 2 (h' 1) Figure 4-36: Comparison between POCR and oxygen consumption rate in vents during test 1 I ’ '17.:T—‘17'il ’— "" ..- .H .._~. _- _-_ _..._ _i:__.4__.__.fi,'— _. #——TI I I --—~><-—~—- 95% Confidence Interval for Mean I I I I I . . F 'I I I 5/ L32 (hits 1, 2 & 3 ) I I I 3 T 0.1154 I I . . I» 0.3925 I I Imin s i m I I I ' i i ' I ' I x..-” I I I I ; i L32 ( Lift 5) /' I ‘ | l L6 . I I_ ’r . . _ ._ . __ 0 I I I O 0.1 0.2 0.3 0.4 0.5 3 I Rate of oxygen consumption during test 1 (h' 1) I Figure 4-37: Comparison between POCR and oxygen consumption rate in vents during test 2 135 4.3. Evaluation of Methanogensis The goal of increasing temperature in bioreactor landfill is to improve the operating conditions for methanogenesis. However, intermittent air injection leads to aerobic conditions within the bioreactor landfill. Exposure to air is detrimental to methanogens, because methanogens are strict anaerobes. Methanogenic activity in bioreactor landfill cell was evaluated in terms of lag time, the rate of increase in percent methane, and hydrogen accumulation. Lag time was defined as the time during which methane concentration change was negligible. Methane and hydrogen concentrations were measured at sampling ports within the landfill for almost a year during startup and for several weeks after stopping air injection. To evaluate the effect of air injection on methanogenic activity, a comparison of hydrogen concentration, methane concentration and lag time was made during startup (i.e., before air injection) and alter air injection. This section is divided into the following: 1) methane generation and hydrogen accumulation during startup, 2) methane generation and hydrogen accumulation afier air injection was stopped, 3) factors affecting methanogenesis establishment during the startup period, 4) factors affecting methanogenesis establishment after air injection was stopped, and 4) comparison of methanogenic activity between the two periods. 4.3.1. Methane generation and hydrogen accumulation during startup The startup period was divided into two: lag time period with negligible increase in methane concentration and a period during which methane concentration was increasing. Methane concentration remained constant for 137i27 days in lifts 2, 3 and 4, after which 136 it started to increase. For lift 1 methane concentration at day 88 (first day of sampling) was 24.6:5.0 % and continued to increase afterwards. Thus lag time for lift 1 was not measured, but it is known to be significantly lower than other lifts. An example of the shape of methane concentration-time profile is shown for Figures 4-38 for location G213 in lift 2, and in Figure 4-39 for location G134 in lift 3. The rate of methane percentage increase was calculated for each individual location in the landfill for the active period regarding methane percent increase. The rate was calculated as the difference between the initial and final methane concentration during the active period divide by the time elapsed during this change. For location 6213, this rate was calculated as the difference between the final concentration (36.6%) and the background concentration (6.1%), divided by the time difference (331-222). For location 6134, the active period was the initial one, and the rate was calculated as the difference between final concentration (60%) and background concentration (24.6%) divided by the time (290-88) days. 40 __ Q 35 *I' — ——————— V’- I 8.; 30 L.-- __.__.._ _fl _- __ _- _ t: .8 25 ‘. 1 a 20 m2 _- - § 15 I — —— — I 8 10 I 3___ W a o 3 o g ;. __ ___ _ 9 _ ____;__ ”bmrfr 1%; 3;: _ ' E . - ‘ l T l j I 0 50 100 150 200 250 300 350 I I Days fiom start of filling lift 2 (Location 6213) r o G213 —LagtimemarkI Figure 4-38: Methane concentration as a function of time during startup in a sampling location in lift 2 137 100 200 300 400 500 I Days from start of filling lift 1 Methane concentration (%) N o o o o o v I I I I I I I Figure 4-39: Methane concentration as a function of time during startup in a sampling location (G134) in lift 1 The average rate of methane percentage increase for the locations in the landfill during the active period was 0.15:0.03% day'l (Figure 4-40). Rate of methane percentage increase was normally distributed; hence averages would be used to describe this rate (Figure 4-40). This rate was calculated for 37 locations within the landfill (Table 4-22). Conditions under which the rate was determined such as, initial waste temperature, temperature of the waste during lag time, temperature of the waste, COD and pH are shown in Table 4-22. For these locations, the calculated rate of methane percent increase, hydrogen concentration during and after lag time, lag time, background concentration and final methane concentration are shown in Table 4-23. 138 Rate of methane generation Std. Dev = .09 Mean = .15 N = 37.00 Figure 4-40: Distribution of the rate of increase in methane concentration in Northern Oaks bioreactor landfill during startup Table 4-22: pH, COD, and temperature at sampling locations during landfill startup Initial waste Temperature Location Lift number tem erature 0f the waste Temperature COD H p during the lag of the waste p (C) time 0112 Lifil 35.4 16.5 12.3 38,600 603 G122 Lifil 28.7 14.3 12.6 36,000 6.22 6123 Lifil 26.8 11.0 10.0 N/A 6.13 0124 Liftl 28.9 N/A N/A 37,800 6.13 0131 Lifi 1 24.1 13.5 10.5 N/A G132 Lift 1 25.7 13.1 12.4 N/A 6.1 0133 Lifil 29.0 13.7 12.2 N/A 6.19 0134 Liftl 29.1 14.4 11.2 43,700 6.34 GI35 Lifil 28.2 12.3 11.1 39,300 5.96 6142 Lifil 22.5 13.8 11.8 N/A N/A 139 Table 4-22: pH, COD, and temperature at sampling locations during landfill startup 140 (continued) L'ft Initial waste 1:11;:333 Timlferat“? e . I o t e waste Location mber tempzrature during the COD pH ( C) . 0 (°C) lag time ( C) 0143 Lifil 22.1 13.4 12.1 N/A N/A 0144 Lifi 1 24.1 N/A N/A N/A N/A (3145 Lifi 1 22.8 13.6 11.6 N/A N/A G146 L111 1 24.0 12.7 12.2 36,223 5.79 G211 Lift 2 21.0 13.4 13.5 N/A N/A G213 L111 2 15.5 11.7 12.0 4,822 6.02 G214 Lifi 2 13.9 9.4 10.5 N/A N/A G215 Lift 2 N/A 11.5 11.6 N/A N/A G221 Lifi 2 20.0 14.7 14.0 N/A N/A G222 L111 2 17.2 11.7 11.9 32,261 6.09 G223 L111 2 23.5 20.9 17.0 34,375 5.91 G225 L111 2 22.5 16.9 17.1 N/A N/A. 0231 Lift 2 14.8 19.8 19.5 26,089 6.23 G232 Lifi 2 10.7 10.9 10.7 N/A N/A G233 Lifi 2 12.0 11.0 11.6 N/A N/A G311 Lifi 3 3.9 7.4 8.6 N/A N/A G312 L111 3 1.7 7.8 8.6 28,076 6.22 G313 L111 3 17.4 14.1 15.5 28,034 5.74 G321 L111 3 0.2 8.1 8.6 N/A N/A G322 Lift 3 3.9 8.4 8.6 28,372 5.8 G323 L111 3 N/M N/A N/A 15,181 5.61 G411 L111 4 -0.7 3.3 4.8 12,686 6.1 G412 L111 4 2.9 4.8 5.6 21,607 6.08 G413 L111 4 -1.6 0.5 0.7 22,241 6.16 G421 L111 4 -1.0 2.5 3.2 11,502 6.2 G422 L111 4 4.5 4.0 4.5 25,920 6.22 G423 L111 4 7.7 6.9 7.0 27,568 6.06 Total 37 35 34 34 20 23 N/A Not applicable (sample was not obtained) Table 4-23: Hydrogen during and after lag time, initial and final methane concentration, rate of increase in percent methane and lag time at sampling locations during landfill startup Hydrogen Hydrogen Background Final 11:33:: Name co:centrslitlon concentration methane La methane in uring ag a ter lag time 0 time time (ppm) (ppm) (AI) (%) percent methane G112 N/A 398 27.02 N/A 52.6 0.130 6122 N/A 123 14.11 N/A 61.8 0.150 G123 N/A 4,829 22.85 N/A 62.9 0.130 G124 N/A 162 32.76 N/A 63.8 0.120 G131 N/A 66 43.06 N/A 65.4 0.320 G132 N/A 123 32.07 N/A 66.9 0.130 G133 N/A 325 18.49 N/A 61.1 0.220 G134 N/A 867 18.25 N/A 59.3 0.210 G135 N/A 1,497 32.03 N/A 52.6 0.090 G142 N/A 5,287 18.69 N/A 56.4 0.240 G143 N/A 200 16.35 N/A 55.7 0.200 G144 N/A 587 16.62 N/A 56.7 0.210 G145 N/A 122 19.36 N/A 60.4 0.210 G146 N/A 640 33.3 N/A 65.1 0.170 G21 1 5,451 473 5.02 222 30.9 0.240 G213 8,254 2,618 6.13 222 36.6 0.280 G214 4,846 385 16.67 130 46.7 0.170 G215 9,479 822 15.46 222 47.9 0.381 G221 11,560 7,001 8.15 156 38.3 0.172 6222 383 217 11.99 222 33.7 0.199 G223 N/A 2,883 11.83 40 39.8 0.096 G225 10,948 7,752 3.18 191 42.5 0.339 G231 525 123 3.43 222 23.5 0.184 G232 9,866 12,561 3.11 222 11.2 0.095 141 Table 4-23: Hydrogen during and after lag time, initial and final methane concentration, rate of increase in percent methane and lag time at sampling locations during landfill startup (continued) H Rate of Nam 801.33.131.12... 8033353530.. “13:53:31 3.23. 1.53221... “1.1"“ during lag after lag time percent time (ppm) (ppm) (ppm) (days) (%) methane (%lday) G233 5,751 2,276 3.3 102 32.2 0.141 G311 N/A 27,350 0 55 10.5 0.040 G312 N/A 9,267 1.4 68 3.6 0.010 G313 106,414 60,178 0 122 8.1 0.050 G321 29,571 1,097 3.3 157 18.7 0.110 G322 50,055 20,145 0.3 96 16.3 0.080 G323 N/A 816 0.2 41 8.9 0.040 G411 17,167 20,247 0.3 122 7.8 0.050 G412 14,229 18,640 0.7 122 27.7 0.140 G413 17,492 38,161 0.9 122 15.3 0.070 G421 6,622 19,496 0.6 122 21.8 0.1 10 G422 6,486 15,718 0.4 61 13.9 0.090 G423 27,359 52,996 1.3 122 20.5 0.120 37 19 37 37 23 37 37 N/A Not applicable (sample was not obtained) The average rate of percent methane increase for lifts 1, 2, 3 and 4 were 0.18:0.035, 0.2110062, 0.055i0.037, and 0.097i0.035 % day", respectively (Figure 4-41). This rate tested to be significantly different between the lifts based on ANNOVA test (Table 4-24). ANNOVA test was used because the rates were normally distributed within each lift (4- 25). The difference between the groups is observed to be between the lower lifts (1-2) and the upper lifts (3-4). Differences in average rate of increase in percent between lifts 142 1 and 2, and between lifts 3 and 4 were not significant. Difference in this rate between lower (1-2) and upper lifts (3-4) was significant based on t-test (Table 4-26). The rate of increase in percent methane in the lower lifts was 0.193i0.031 % day'1 and for the upper 1111s 007610.025 % day". .2: .1. _I_' __ 0.0‘ Rate of increase in methane concentration N= 14 11 6 6 Lifil Lift2 Lift3 Lifi4 Figure 4—41: Rate of increase in methane concentration (%/day) for each lift Table 4-24: ANN OVA test to determine differences between the different lifts in rate of increase in percent methane Sum of Mean , Squares Df Square F 3'8- Between Rate of Greg’s 0.1212 3 0.0404 methane , , concentration grthm 0.1461 33 0.0044 9.13 0.000152 increase roups Total 0.2673 36 143 Table 4-25 Normality tests for methane concentration, hydrogen concentration and rate of increase in percent methane Kolmogorov-Smirnov Shapiro-Wilk L‘fi Statistic df Sig. Statistic df Sig. number Liftl 0.121 14 0.200 0.949 14 0.545 Final L1112 0.175 11 0.200 0.929 11 0.399 methane . concentration L1113 0.205 6 0.200 0.951 6 0.750 L1114 0.213 6 0.200 0.934 6 0.615 Liftl 0.337 14 0.000 0.609 14 0.000 Hydrogen L1112 0.276 11 0.019 0.801 11 0.010 concentration annlngnme Lift3 0.202 6 0.200 0.862 6 0.196 Lift4 0.356 6 0.017 0.794 6 0.051 Liftl 0.155 14 0.200 0.936 14 0.368 Rate of methane Lift2 0.178 11 0.200 0.932 11 0.433 concentration L1113 0.223 6 0.200 0.949 6 0.735 increase Lift4 0.156 6 0.200 0.981 6 0.955 Hydrogen Lift2 0.151 10 0.200 0.909 10 0.276 °°“°e.““a“°“ L1113 0.319 0.108 0.769 5 0.044 during lag time L1114 0.203 6 0.200 0.912 6 0.452 Table 4-26: t-test to test the difference between upper and lower lifts with respect to the rate of increase in methane percent 95% Confidence Std. Interval of the t (if Sig. (2-tailed) .Mean Error Difference Difference . Difference Lower Upper 4.99 35 1.6310'5 0.117 0.023 0.070 0.16 144 To Show the variations within each lift in the methane concentration-time profiles, an example of this profile for all six sampling locations in lift 4 is shown in Figure 4-42. The lag time for lift 4 was 80 days. The concentration increased after from below 1% to a concentration range between 12.3 % and 27.7% by day 238 from start of lift 4. A 30 I— 2- -~———--- —— —— — , _-_w. _ . /3 g 20 I_._. P— _7— "——_—”_—"‘ ”fl” 7 A‘ “ / I 5 15 _I *5 __.__,_____ __.- . - . _w/[q —I g i / . 8 10 I""_ “u“ — —_ __’ “#1" “’—‘——G o . 8 s -I--———- —~ » ,/ g 0 1..-».-- - -----» . C . 0 50 100 150 200 250 Days fromfillinglift4 Figure 4-42: Methane concentration during start up of bioreactor landfill lift 4 Hydrogen concentration during the startup phase of the anaerobic bioreactor landfill was also measured as a second indicator of methanogenic development. Hydrogen concentrations were much higher than the 100 ppm level known to be thermodynamically unfavorable to syntrophs (butyrate and propionate utilizers). Hydrogen concentrations in the waste during and after lag time were not normally distributed (Table 4-25), hence the median is used to represent the values of hydrogen in the waste. The range of hydrogen concentrations during lag time for all the locations in the landfill was between 383 and 106,414 ppm, with a median of 9,866 ppm. The range of hydrogen concentration after the lag time was between 66 ppm and 60,178 ppm, with a median of 145 1497 ppm. There was no statistical difference between hydrogen concentration before and after the lag time, as concluded from Wilcoxon Signed Ranks Test (Table 4—27). Table 4-27: Wilcoxon Signed Ranks Test to study the difference in hydrogen concentration between the two periods: before and lag time H dro en concentration Z Asymp. Sig. (2-tailed) I l Before and after lag time -0.926‘ 0.355 I 'Based on positive ranks To study the variation in the lifts, the median for each lift was calculated. Median concentration of hydrogen before lag time was 1,549 ppm, 20,145 ppm and 19,872 ppm for lifts 2, 3 and 4, respectively (Figure 4-43). Lift 1 had lag time equals to zero hence hydrogen concentration during that time was not applicable. The median of hydrogen concentration after the lag time was 362 ppm, 2,276 ppm, 14,706 ppm and 19,872 ppm for lifts 1, 2, 3 and 4, respectively. Concentrations of hydrogen differed significantly between the lifts (Table 4-28). The variability in hydrogen concentrations was big as indicated by the following ranges observed in each lift. The range of methane concentration after lag time was from 66 to 5,287 ppm for lift 1, from 123 to 12,561 ppm for lift 2, and from 816 to 60,178 ppm for lift 3 and from 15,718 to 52,996 ppm. In addition, hydrogen concentrations during the lag time were different than those after the lag time, when compared lift by lift (Figure 4-43). Lifts 2 and 3 had significantly lower hydrogen concentrations after the lag time, while lift 4 had significantly higher hydrogen concentrations. 146 Table 4-28: Kruskal Wallis Test to study the difference in hydrogen concentrations between the different lifts Hydrogen Lift Mean Chi- Df Asymp. concentration number Rank Square Sig. . Lifi2 10 6.50 , , Lift3 18.00 During lag time 10.57 2 0.00507 Lift4 6 11.83 Total 19 Lift2 11 17.45 Lift3 6 27.17 After lag time Lift4 6 32.33 20.40 3 0.00014 Total 37 Liftl 14 11.00 70,000 A 60,0001 E o. 8: 50,0001 1:: .2 5 40,0001 8 O 0 30,0001 1: 8 5 20,000' -x- ‘10 La time g 10,0001 3: _ g >~. . I 01 _L 9% __ EB; [JDurmg -10,000 - _ _ _ EjAfter N= 5 12 12 5 5 5 5 Liftl Lift2 Lift 3 Lift4 Figure 4-43: Hydrogen concentration during and after lag time 147 Hydrogen concentration -time profiles for all the locations in lift 4 is Shown in Figure 4- 44. Hydrogen concentrations in lift 4, measured 40 days after the start of lift 4 was at a median of 7,835 ppm.‘ With time, hydrogen concentrations increased further. The median H2 concentration after 238 days was 30,308 ppm. The results indicate that even after 238 days, the landfill was severely inhibited. Moreover, the hydrogen concentration increased over the 238 days, thus indicating the inability of hydrogen utilizing methanogens to uptake the produced hydrogen. Indeed, these elevated hydrogen concentrations were accompanied by low methane concentrations (18.59%) (Figure 4- 41). i E 120,000 ”_-_ , __ 3 100,000 8 80,000 '33 § 60,000 § 40,000 I“ O . g 20,000 .8” ° . E‘ 0 50 100 150 200 250 I Days from filling lift 4 I Figure 4-44: Hydrogen concentration during start up of bioreactor landfill lift 4 Table 4-29: Wilcoxon Signed Ranks Test to test differences between hydrogen concentration during and after lag time for each lift Z Asymp. Siia-tailed) lift 2 -2.4973 0.012515 lift 3 -2.0226 0.043114 lift 4 -2.2014 0.027708 148 4.3.2. Methane generation and hydrogen accumulation after stopping air injection After air injection, percent increase in methane concentration and hydrogen concentrations were studied only in the locations that were exposed to air. The locations studied after air injection are shown in Table 4-30, with average waste temperature during this test, rate of increase in percent methane, and hydrogen concentration. For all the locations studied, the methane concentration was reduced to a median of 2.5% at the end of the air injection. Median was used for data that are not normally distributed; otherwise average was used to represent the center of the data (Table 4—31). Median of the lag time observed at all studied locations before was 11 days. After lag time, median of rate of percent increase in methane concentration was 0.255 % day'l. Methane concentration in the waste reached 40.5:7.6% for lifts l, 2 and 3 after 62 days fi'om stopping the air injection, and 21.1i8.0% for lift 4 after 40 days. The concentration of methane was not monitored after the 40 days, due to freezing conditions in the sampling ports. To consider the variations in methanogenic development within the waste, variations in each lift is studied. Representation of the center of the data was made using average values for normally distributed data, otherwise the median was used (Table 4-32). Rates of methane generation for lifts l, 2, 3 and 4 were 1.148i0.600, 0.7181016], 0.452i0.215, and 0.494:t0.170 % CH4 per day, respectively (Figure 4-45). Rate of methane concentration increase was significantly different between the lifts based on ANNOVA test (Table 4-33). Final methane concentrations at day 62 from turning off the air injection were 60.7i7.3 %, 41.3i7.2 %, and 18.0-336.5 %. Methane concentration for 149 lift 4 was 21.1i8.0 % measured on day 40 from turning off the air injection. The difference in concentration of lifts 1, 2, and 3 was significant based on ANNOVA test (Table 4-33). t: 1.8 .2 g 1.61 8 I g 1.41 8 a) 1.2‘ _— 5 1—1 E. 1.01 d.) E a .8‘ o __ 58 6I I at; :— .2 4‘ —“‘ _— ‘4—1 0 _1_ $3 .21 g 0.0 _ . . _ = 4 11 5 5 Lift 1 Lift2 Lift3 Lift4 Figure 4-45: Rate of increase in methane concentration (%/day) for each lift 150 Table 4-30: Waste temperature, hydrogen concentration, lag time, final methane concentration and rate of increase in percent methane for locations exposed to air H d H d Rate of ro en ro en . increase Location temYSSure :uringg 5t" 151g tlir‘i'rge mftgzire in lag time time percent (°C) (ppm) (ppm) (days) (%) methane (%lday) G141 19.9 N/A 2,372 0 53.5 0.829 G142 19.6 N/A 274 0 56.6 0.825 G143 22.9 N/A 527 0 60.8 1.384 G144 N/A N/A 550 0 64.9 1.554 G21 1 25.3 1,746 963 6 26.5 0.400 G213 19.2 8,103 4,432 11 50.3 0.640 G214 17.5 972 193 19 57.0 0.680 G215 16.4 764 931 28 56.7 1.140 G221 39.8 886 1,949 11 33.1 0.560 G222 22.4 803 320 1 1 36.7 0.420 G223 20.8 1,729 306 28 44.3 0.800 G224 25.5 2,102 1,114 11 51.8 0.830 G225 19.8 6,466 3,325 28 41.7 0.940 G231 16.8 372 233 28 22.1 0.520 G233 14.7 1,658 1,282 28 42.0 0.970 G311 15.7 1,628 988 28 23.9 0.560 G312 14.0 11,008 8,781 28 13.2 0.300 G313 18.9 23,471 34,486 40 21.1 0.670 G321 20.5 1,462 650 28 20.4 0.470 G322 13.7 11,519 13,198 28 12.0 0.260 G41 1 22.6 279 206 6 31.3 0.720 G412 26.4 384 192 6 19.2 0.440 G413 10.0 19,210 61,459 6 15.7 0.370 G421 33.2 1,978 942 6 23.2 0.550 G423 16.8 2,805 2,026 6 16.1 0.390 151 Table 4—31: Normality tests for concentration of methane, final methane concentration, hydrogen concentration and rate of increase in percent methane in the landfill as a whole Kolmogorov-Smirnov Shapiro-Wilk Statistic df Sig. Statistic df Sig. BaCkgr°und netham 0.261 27 5.2410'5 0.732 27 1.1210'5 concentration Fmalmethfme 0.142 27 1.7110'1 0.933 27 0.081 concentration Hydmg‘?“ “me.“mtw“ 0.350 27 2.8210'9 0.631 27 4.8410'7 during lag trme Hydmge“ °°“°.ent’a"°” 0.378 27 5.8110’” 0.429 27 3.4710'9 after lag trme Lagtime 0.250 27 1.4110“1 0.849 27 110103 Rate of increase in methane concentration 0.134438 25 0.2 0.913981 25 0.037432 152 Table 4-32: Tests of normality per lift for rate of increase in percent methane, methane concentration and hydrogen concentration Kolmogorov-Smirnov Shapiro-Wilk ”‘1 Statistic df Sig. Statistic df Sig. number - Lift] 0.301 4 N/A 0.825 4 0.156 Rate of increasein Lift2 0.109 11 0.20 0.962 11 0.792 percent Lift3 0.210 5 0.20 0.941 5 0.671 methane . L1114 0.246 5 0.20 0.883 5 0.323 L1111 0.162 5 0.20 0.969 5 0.866 Final Lift2 0.119 12 0.20 0.960 12 0.779 methane . concentration L1fi3 0.269 5 0.20 0.883 5 0.321 Lift4 0.217 5 0.20 0.877 5 0.296 Hydrogen L1112 0.344 12 3.27104 0.685 12 6.08104 “minimum Lifi3 0.225 5 0.20 0.884 5 0.327 during lag time Lift4 0.404 5 7.5710'3 0.668 5 4.3210’3 Liftl 0.299 5 0.164 0.849 5 0.191 Hydrogen Lift2 0.249 12 0.039 0.789 12 7.1310'3 concentration . annlagnme L1113 0.255 5 0.20 0.844 5 0.176 Lift4 0.457 5 9.9610“4 0.576 5 2.9110“1 153 Table 4-33: ANNOVA table to test the difference in final methane concentration in between lifts 1, 2 and 3 and in rate of increase in percent methane for lifts 1, 2, 3 and 4 Sum of Mean . Squares df Square 813' Between Final methane Grows 4567 2 2284 concentration Within 26 3. 54 10-6 (lifis 1,2,and 3) Groups 1662 19 87 Total 6229 21 Bé’twee“ 1.321709 3 0.44057 Rate of methane roups concentration Within 1.201084 21 0.057194 7.7 0.001174 increase Groups Total 2.522793 24 Figure 4—46 shows the increase in percent methane concentration in lift 2 after air injection was stopped. The average concentration of methane after stopping the air injection was 2.1.-1:16 %. The concentration of methane increased at an average rate of 0.72:0.16 % day", reaching an average value of 42.0i6.9 % methane after 62 days. Change in concentration beyond the 62 day period could not be followed because the sampling ports were frozen (12/3/03). Methane concentration 4.5 months later (4/20/04) was 53.8:t7.0, indicating further increase in % methane. During this sampling event, fewer variations in the methane concentrations were observed. After approximately 6 months from end of air injection test 1, the concentration of methane had less variations within lift 2 with significant methane production. Note that the average concentration of methane before air injection in lift 2 was 30.8:7.4 %, with a range from 11-50 %. 154 A70- —— ~——* -—— _-_ c,\" . :“r‘“‘“ J,,. .3 50 ”I‘ __f _- L . § 40I w«_ —— __.. L3 __.", ...» / / g 30 " 725 ~ 7 ‘1 g» 0 W ,_./ 0 10 20 30 40 50 60 70 Days fiom stopping air injection Figure 4-46: Methane regeneration after air injection was stopped in lift 2 The concentration of hydrogen was also measured at the end of air injection tests. Hydrogen concentration during the lag time for all the sampling locations exposed to air was at a median of 1694 ppm. This hydrogen concentration was variable in the waste with a range from 279 to 23,471 ppm. After the lag time the concentration of hydrogen became 964 ppm. The variability in hydrogen concentrations among the sampling locations was even higher than what was observed during the lag time. The range of hydrogen concentration after lag time was from 192 ppm to 61,459 ppm. There was no statistically significant difference between the concentration of hydrogen before and after the lag time, when studied for all the ports at in all the lifts (Table 4-34). Table 4-30 shows for each location the hydrogen concentrations during lag time and after lag time (Figure 4-47). 155 70,000 A 60,000 . * E Do 9; 50,000 i c: .2 *5 40,000 I E o 8 30,000 I c: 8 5 20,000 1 on o g 10,000 . 3t Lag trme >1 I 0 I AL 2% D DWI; -10,000 [:lAfter N= '5 12'12 5'5 5'5 Liftl Lift2 Lift3 Lift4 Figure 4-47: Hydrogen concentration during and after lag time after air injection Table 4-34: Test statistics for Wilcoxon Signed Ranks Test H dro en concentration Z Asymp. Sig. (2-tailed) I I Before and after lag time 4.5422 0.12302 I 'Based on positive ranks Median hydrogen concentrations during lag time were 1,315 ppm, 11,008 ppm and 1,978 ppm for lifts 2, 3 and 4, respectively (Figure 4-47). Ranges of hydrogen concentrations were from 372 ppm to 8,103 ppm for lift 2, from 1462 ppm to 23,471 ppm for lift 2 and from 279 ppm to 19,210 ppm for lift 4. Lift 1 had zero lag time. Median hydrogen concentrations after lag time were 550 ppm, 947 ppm, 8781 ppm, and 942 ppm for lifts l, 2, 3, and 4 (Figure 4-47). The ranges for hydrogen concentration after the lag time were 156 fi'om 274 to 2372 ppm for lift 1, from 193 ppm to 4,432 ppm for lift 2, fiom 650 ppm to 34,486 ppm for lift 3, and from 192 ppm to 61,459 ppm for lift 4. The hydrogen concentration data indicated that there is a big variation in hydrogen concentration among the lifts. The big variation within each lift, and the existence of very low concentrations in each lift, resulted in a non-significant test results between the lifts (Table 4-35). Hydrogen concentrations during and after lag time were not statistically different between the lifts, as indicated by Wilcoxon Signed Ranks Test for lifts 3 and 4, however the concentration after lag time was significantly lower than that during lag time (Table 4-36). Table 4-35: Kruskal Wallis Test to test difference between hydrogen concentrations among the different lifts Lift N Mean Chi- df' Asymp. number Rank Square SiL Lift 2 12 9.83333 Hydrogfim °°ncéntrati°n Lia 3 5 16 3.22134 2 0.19975 during lag t1me Lift 4 5 11 Total 22 Lift 1 5 12.4 , Lift 2 12 12.5833 Hydmge“ °°“°.e““a“°“ Lift 3 5 20.2 3.75053 3 0.28969 after lag t1me _ Lift 4 5 12.8 Total 27 Table 4-36: Wilcoxon Signed Ranks Test to study the difference in hydrogen concentration between before and after lag time Z Asymp. Sig. (Z-tailed) lift 2 ~2.118 0.034 lift 3 -0.135 0.893 lift 4 -O.674 0.500 157 To further show the type of variations within each lift, an example of hydrogen concentration-time profile for all the locations in lift 2 is shown in Figure 4-48. The median of hydrogen concentration after stopping the air injection for lift 2 was 809 ppm, which increased after 6 days to 2,026 ppm and then decreased to 650 ppm. The concentration of 650 ppm was observed after 62 days from stopping the air injection. The marginal increase in hydrogen indicates that hydrogen utilization has improved over the 62 days operating in anaerobic mode, and short periods of air injection for the purpose of increasing the temperature can be handled adequately by the methanogens. The concentration of hydrogen further decrease 4.5 months later (200 days from stopping the air injection) in lift 2, the median of hydrogen concentration was 278 ppm. 16,000 _.__- _ 14,000 12,000 . — 10,000 -. 6,000 . 4,000 2,000 . Hydrogen concentration (ppm) 00 ”o o O t | Days fi‘om stopping air injection Figure 4-48: Hydrogen concentration change after air injection was stopped in lift 2 Higher hydrogen accumulation and lower methane concentration at the start up and after air injection indicated that initial startup of the bioreactor landfill was more problematic compared to the restart period after air injection. The hydrogen concentrations continued 158 to increase during the startup, perhaps because methanogens were yet to establish in a newly started bioreactor landfill. However, during the period following air injection, hydrogen accumulated to some extent but decreased quickly. The normal methane concentrations for stable anaerobic digestion (40-60 %) even in the presence of more than 100 ppm hydrogen indicated that the syntrophs may be able to remain active perhaps in a protected environment that has lower hydrogen concentrations. This indicates that after 62 days from stopping the air injection test 1, methanogenesis was well established with lower hydrogen concentrations and higher methane concentrations in comparison to those observed during landfill startup. 4.3.3. Factors affecting methanogensis establishment during bioreactor landfill startup Waste temperature, pH and COD are some of the most important parameters that affect anaerobic biodegradation. Methane concentration in the landfill did not correlate with waste temperature of the waste (Figure 4-49, 4-50). Waste temperatures decreased from 26.5 °C to 11.7 °C in lift 1 and from 17.1 to 13.6 °C in lift 2 and remained constant at the final temperature throughout the startup period (Figure 4-49). Methane concentration increased from 24 % to 50 % in lift 1 and from 8 % to 30 % in lift 2 during the same period. However, the increase in percent methane in lift 2 had a considerable lag period, perhaps due to delayed supply of leachate. Waste temperature increased from 3.9 °C to 8.6 °C in lift 3 and from 1.1 °C to 4.6 °C in lift 4, and remained constant at the final temperature (Figure 4—50). Methane concentration in these lifts increased from values close to 0 % to 19 % in both lifts during the same period. This data supports that 159 methanogenic activity is substantial in bioreactor landfills even at lower temperatures (e. g., in this case at an average temperature of 10.7i0.7 °C). The rate of increase in percent methane (computed as difference in % increase divided by the time between the initial and final % methane) did not correlate with waste temperature (Figure 4-51). The rate of increase in percent methane for lifts 3 and 4 was less than for lifts l and 2. The time for which lift 2 had constant methane concentration at approximately 8% was 177 days from the start of lift 2 (Figure 4-49). Lift I remained open for about 60 days after filling, while the remaining lifts remained open for only 2 weeks after filling. The initial temperature of waste (which was kept open until 5 days after coverage) was different in each lift with temperatures of 26.5, 17.1, 3.9 and 1.1 °C for lifts 1, 2, 3 and 4, respectively. The exposure to aerobic conditions for varying amounts of time may be the main reason for the difference in rate of increase in percent methane. Aerobic biodegradation and hydrolysis at higher temperatures and longer periods may have lead to the increase in the readily available substrates in the waste for methanogens. This is supported by a correlation between COD and initial waste temperature (Figure 4-52). Moreover, methane concentration measured after approximately 200 days from the start of each lift also correlated with COD (Figure 4- 53). This indicated that hydrolysis of particulate organic matter into soluble COD was enhanced at a higher rate than hydrolysis observed at locations with lower temperatures. This process resulted in improved methanogenic activity, measured as increase in % methane and increase in rate of increase in percent methane. 160 05 O A ~ 50 28, e s 2 40 '5 g E a 3° ‘2’ E o a 20 3 g 5 . 10 ‘15. 3 2 . . 0 0 100 200 300 400 500 Days I '7'1—m—‘Tm'im1 . Ten'peraturein 1112‘ ‘ ’ . + Metlnne concentration in lift 1 ~-~~><« Methane concentration 'm lift 2 L Figure 4-49: Methane concentration and waste temperature during the startup period for lifts 1 and 2 D) O 1 O\ O A 25 - 50 :5 8 J 8 2 20 -» — - 40 IE '3 e a 15 a—— ~ - 30 g E 1 8 g 10 I g 1 § 5 ‘3 ( E 7 0 100 I I 2.-.... Terrperature in; lift 3 -+-— Metlune concentration 'n lift 3 i x Terrperatureinlift4 +Methaneconcennation'nlift4i Figure 4-50: Methane concentration and waste temperature during the startup period for lifts 3 and 4 161 Rate of increase in methane concentration .4 D :1 *- .3I 1:1 * Cl 9(- 99+ '2‘ a- 0 Liftnumber 1:1 * 1:1 51* OLift4 O * are 11 0 q, a 'Lift3 O D C’Lift2 0.0 *Liftl 0 10 20 30 Temperature of the waste Figure 4-51: Rate of increase in methane concentration vs. waste temperature during startup period 50,000 * 40,0001 * 'X- * -x- a A :1 a 30,0001 . > 3 O O 0 Lift umbe 8 20000 O O n r U r ' O Lifi4 ' Lift3 10,0001 8 D Lift2 D 0 _ _ - ' * Lift 1 -10 0 10 20 30 40 Initial waste temperature (C) Figure 4-52: COD vs. initial waste temperature in the landfill 162 7O *- 60' * 8 g 50' * * b 5 o 40‘ r: 8 Lift number 0 301 5 * 0Lm4 5 a) 201 O . 2 1:1 ' ert 3 O 101 D 00 . O ' c1 ‘3 Lift2 o _ _ °’_ _ *Lmr 0 10,000 20,000 30,000 40,000 50,000 COD Figure 4-53: Methane concentration after 400 days vs. COD Rate of increase in % methane did not correlate with pH (Figure 4-54) The pH range in the bioreactor landfill cell was between 5.6 and 6.3 which is on the lower side of the acceptable pH for methanogens (which is pH 6 to 8.0 [130]). Thus variations in pH were not expected to be the primary reason for variations in the rate of increase in percent methane. Nevertheless, all the locations that had hydrogen concentrations above 40,000 ppm had a pH below 6. Except for lift 4, hydrogen concentration decreased with an increase in methane concentration. The increase in hydrogen in lift 4 may be due to lower methanogenic activity in comparison to acidogenic activity, i.e., methanogens were not active enough to remove the hydrogen produced by acidogenes leading to hydrogen accumulation. 163 .5 .3 E Cl C.‘ o o C‘. 8 * .. o .2' 1:1 5 o .5 -x- ” «x- E 0 Lift number :1 * * x- .u-a O * ‘1’ O 0 Lift 4 Ki 1 ‘ D 2 . +1- 0 . a O ’ Lift 3 “5 > ’ O 0 Lift 2 8 g 0.0 ' ' ' * Lift 1 5.6 5.8 6.0 6.2 6.4 pH Figure 4-54: Rate of increase in methane concentration vs. pH 4.3.4. Factors affecting methanogenesis establishment after air injection was stopped Rate of increase in methane concentration and final concentration of methane did not correlate with the waste temperature during the test period (Figures 4-55 and 4-56). However the rate of increase in percent methane after air injection correlated with the rate of increase in percent methane during startup at a 2-tailed significant level of 0.020 (Pearson correlation coefficient = 0.482). Thus because the spatial trend in rates of increase did not change it is evident that the inhibitory effects of air injection on methanogenic activity were temporary. It is possible that methanogens were protected to an extent from oxygen concentrations in the liquid phase by aggregates. Therefore as soon as the air injection was stopped the concentration of methane started to increase at a 164 rate that is higher than that obtained before, and was higher for those locations where it was higher before. 1A Lfi Rate of increase in methane concentration 9° * Cl C] U #0 D O D E D Cl C D O o O D D D D 0 1'0 20 30 40 50 Temperature of the waste (C) Lfimmmr OLift4 ’Lift3 DLift2 *Liftl Figure 4-55: Rate of increase in methane concentration vs. waste temperature after air injection 165 70.0 *- g 60.0« * '5 DU 9(- g .. D 5 50.01 a O 1:: 8 a a” 1, 40.01 g a a D a E 30.01 0 _.. D ‘3 . o i: 20.0. ‘3 n O o O 10.0 r’ 0 10 20 30 40 50 Figure 4-56: Methane concentration after 62 days from stopping air injection for lifts l, 2 and 3 and after 40 days from stopping air inj ection for lift 4 vs. waste temperature Temperature of the waste 1.6 1.4- 1.2- 1.01 .8I After air injection .61 .4- *- Figure 4-57: Rate of increase in percent methane after air injection vs. before air Before air injection injection 166 Liftnumber C)Lift4 ’Lifl3 E3Lift2 *Liftl OLift4 ’Lifi3 DLiftZ *Liftl 4.3.5. Comparison of methanogensis establishment between startup and after air injection We hypothesized that temperature increase will lead to a higher rate of increase in percent methane, higher methane concentrations, lower hydrogen concentrations, and lower lag time. All the parameters used in this study to evaluate anaerobic digestion in the bioreactor landfill indicated a better performance of the bioreactor landfill after air injection was stopped in comparison to the startup period. The average rate of increase in percent methane at all locations exposed to air in the landfill increased significantly from 0.15i0.042 °C day'l to 0.68:0.15°C day“1 (Figure 4-58). This was accompanied by a decrease in the median of hydrogen concentration from 4,085 ppm to 953 ppm (Figure 4-59 and 4-60). The lag time also decreased from 137 days to 15 days, in response to air injection. This study demonstrated that air injection is a key to a successful startup of anaerobic digestion of bioreactor landfills in colder climates. 55 2.0 3; e) g 1.5. 0* <6 _.__— 5 0 E E 1.01 0 2 a .E -5' :1: § . I 3 —I— 00‘ ._J__. ‘4— o 8 £3 -.5 _ _ N= 23 23 Startup After air injection Figure 4-58: Comparison between the rate of methane concentration increase during startup up and after stopping air injection 167 a 120,000 C. 3 «x- ; 100,000' 00 -‘—° 80,000: E” "O 60,000I S .5 o g 40,0001 2‘3 8 20,000 i c: a ,, g 5 01 —l_ m E -20,000 . ' = 21 21 Startup Afterairirrjectbn Figure 4-59: Comparison of the hydrogen concentration during lag time between during startup up and after stopping air injection 8 70,000 ca. 8' a» 60,0001 0 * .§ ... 0 on 50,0001 ..‘9. l— a 40,000 CU *- 5 30,000. *5 b 5 20,0001 8 -x- 8 10,0001 0 a a) m 5 °‘ E ~10,000 _ = 24 24 Startup After air injection Figure 4-60: Comparison of the hydrogen concentration after lag time between during startup up and after stopping air injection 168 5. CONCLUSIONS From the experiments conducted in Northern Oaks bioreactor landfill, the following conclusions can be drawn. The conclusions are listed corresponding to each objective for clarity. A major part of Objective 1 was to establish and quantify the relationship between the amount of heat produced (Id) and the amount of oxygen supplied (mole), termed as heat generation factor. Characterization of performance of the air injection with respect to temperature and leachate characteristics was also part of Objective 1. The conclusions drawn from the intermittent air injection experiments are: 0 Heat generation factor was calculated to be 341i84 kJ/mole of oxygen under field conditions. 0 Leachate recirculation can be used as a control mechanism for extreme temperature increases in bioreactor landfills. This conclusion is drawn because leachate recirculation, if practiced with leachate temperatures lower than the waste temperature, leads to immediate heat loss (50% of the total heat loss), decreasing the waste temperature. 0 Waste temperature increased above the ambient temperatures by 9.6, 13.2, 8.2, 8.4, and 3.6 °C for waste in lifts 1, 2, 3, 4 and 5, respectively. The relationship 169 describing the increase of waste temperature above ambient temperatures was T -1.1803 T was-re- ambient established from the data in this study to be: + 85572 0 Leachate Injection at a rate of approximately 8% (leachate volume to waste volume) is sufficient to provide a uniform temperature within the waste. This conclusion was valid at an injected leachate temperature of 125°C and waste temperature between 4°C and 27 °C. 0 Air injection (21 days at full capacity using 2.5 HP compressor lead to and increase in waste temperature from 0°C to 30°C for waste with low volumetric moisture content (0.47:0.03), and from 11°C to 18°C for waste with higher volumetric moisture content (0.56i0.19) during. 0 In response to air injection (139 days), COD of leachate in lift 1 decreased fi'om 21 g L-l to 2.9 g L-l. Objective 2 was to determine the POCR and study the effect of moisture and temperature on POCR. The following conclusions were drawn from the data related to POCR. 0 First order POCR averaged 0.0867i0.0422 h-l after 15 days of air injection. This increased to 0.2450i0.0843 h-l after 21 days of air injection. The increase in POCR correlated with increase in temperature for temperature change less than 10°C. 170 Monod equation was not a good fit for POCR, because the range of oxygen concentrations of 0 to 21% is smaller than the range required to obtain independent values of Monod coefficients k and Ks. Because it is impossible to have an oxygen concentration higher than 21% in air, it is not expected that Monod equation will provide a good fit in such experiments. Objective 3 was to determine the performance of anaerobic digestion at startup and following air injection. Performance was evaluated with respect to methanogenesis establishment, using methane and hydrogen concentrations as indicators for anaerobic digestion process. The following conclusions were drawn with respect to performance of anaerobic digestion: Air injection is key to a successful startup of anaerobic digestion of bioreactor landfills in colder climates because of its potential to increase the waste temperature without impacting the methanogenesis. The average rate of increase in percent methane at all locations exposed to air as a result of air injection increased significantly from 0.15i0.042 °C day"1 to 0.68:1:0.15°C day]. This was accompanied by a decrease in the median of hydrogen concentration from 4,085 ppm (before air injection) to 953 ppm (after air injection). There was also a decrease in lag time (i.e., time during which change in % methane is negligible) also decreased from 137 days to 15 days, in response to air injection. 171 6. ENGINEERING SIGNIFICANCE Air injection can be used for increasing waste temperature in landfills. Air injection at a flow rate of 40 to 120 scfrn was practiced, that resulted in a rate of temperature increase of 0.3 °C/day i 0.3 with a range of 0.1 to 1.6 °C /day. In this study air injection did not result in fires, even at temperatures above 55 °C, because leachate recirculation was practiced to reduce waste temperature. In addition, turning off the air injection resulted in a slow decrease in temperature due to heat conduction within the waste and to heat loss to the boundaries of the landfill. In this study, both leachate recirculation and gas extraction lines were used to inject air, hence no additional construction was necessary. . To control temperature, it is recommended that leachate recirculation lines are constructed in proximity of air injection lines, or place the air injection lines in a manner that allows use for leachate recirculation. Heat generation factor can be used to predict temperature increase in the landfill cell due to air injection. In this study, the heat generation factor was computed to be in the range of 200 to 626 kJ per mole of oxygen, with an average of 341:1:84 lemole of oxygen. The heat generation factor determined in this study indicates that it is within the same range reported in literature for waste and other substrates, although variations observed in the field were higher than those obtained in laboratory scale reactors However, heat generation factor must be determined in other landfills to verify if the values obtained in this study are significantly different from that obtained for pure substrates. The results from this study support that landfills can be used as a calorimeter to determine tin situ heat generation factor for solid waste. 172 Point oxygen consumption rate in the landfill can be determined in the field using first order kinetics. This term was obviously a compromise because of the problems associated with apportioning a certain amount of solid waste to a measured oxygen consumption rate. However, it can still be used to assess the aerobic activity of waste in situ. For waste with higher POCR higher air injection flow rate or higher intensity of air injection network would be required. POCR found in this study was 0.09 to 0.25 per hour after 14 and 21 days of air injection. The POCR was affected by moisture content as well by temperature. This parameter can be used in two dimensional and three dimensional modeling of oxygen flow within the waste. This may help determine the distance of travel for the oxygen in the injected air. Intermittent air injection can be used as means to improve the conditions for anaerobic digestion process. In this study an air injection up to 25 days at a maximum flow rate of 120 scfrn was use. Subsequent to this air injection event, methanogensis was established within 60 days as supported by increase in methane concentration to ~ 60% at locations exposed to air after stopping the air injections. Continuous air injection for more than 25 days was not practiced in this study hence the effect of longer term aeration on the recovery of methanogens was not studied. These results indicate that intermittent air injection does not inhibit methanogens for significant periods. Air injection can also be used to improve leachate quality for COD and VOC. In this study, both COD and VOC concentration decreased after air injection. The mechanism 173 of VOC removal from leachate was not determined.VOC could have either volatilized or degraded in the process. Further research on this issue need to be conducted. 174 10. 11. REFERENCES Pacey, J., D.R. Reinhart, D.L. Hansen, G.T. Townsend, and W.H. Johnson. Landfill Bioreactor-An Innovation in Solid Waste Management. in SWANA 22nd Annual LFG Symposium. 1999. I SWANA, The Solid Waste Manager's Guide to the Bioreactor Landfill. 2002: Solid Waste Association of North America. Wolfe, K., Aerobic bioreactor landfill technology: An innovative approach to municipal solid waste management CE news, 2000: p. 62-67. USEPA, http://www.epa.gov/epaoswer/non-hw/muncpl/landfill/bioreactors.htm. Warzinski, J ., B.T. Waterrnolen, and DR. Genthe, A superior approach to recirculation Waste Age, 2000. Pagano, J .J ., R.J. Scrudato, and GM. Summer, Leachate recirculation at the Nanticoke Sanitary Landfill using a bioreactor trench NYSERDA report 98—6, 1998. Augenstein, D., R. Yazdani, R. Moore, and K. Dahl, Yolo county controlled landfill demonstration project 1997: p. 43-83. Moore, R., D. Karina, and R. Yazdani. Hydraulic characteristics of Municipal Solid Waste: Finding of the Yolo county Bioreactor landfill Project. in The thirteen International Conference on Solid Waste Technology and Management. 1997. Philadelphia, PA. Dahl, K. Reuse of shredded waste tires for landfill gas collection and leachate injection systems in Yolo county's landfill bioreactor demonstration project. in Proceedings from SWANA's 21st Annual Landfill gas symposium. March 1998. 1998. Yolo county, California. Townsend, T.G., W.L. Miller, R.A. BISHOP, and 1H. Carter, Combining systems for Leachate recirculation and landfill gas collection Solid waste technologies, 1994: p. 18-24. Barry, RM. and DC. Demme, Design innovations at the DSWA's latest landfill expansion project WASTECON Proceedings, 1997. 175 12. 13. 14. 15. l6. 17. 18. 19. 20. 21. Maier, T.B.e.a., E.S. Steinhauser, N.C. Vasuki, and F.G. Pohland. Integrated Leachate and Landfill Gas Management. in Sardinia 95: Fifth international Landfill Symposium. 1995. Sardinia, Italy: CISA, Environmental Sanitary Engineering Center. Wilson, V.L., W.D. Evans, T.D. Stark, and E. (c/o Repa. An Interim Slope Failure Involving Leachate Recirculation. in Municipal and Industrial Solid Waste Disposal Technology Waste Tech 2000. 2000. Orlando, Florida. Akesson, M. and P. Nilsson, Material dependence of methane production rates in landfills. Waste Management & Research, 1998. 16(2): p. 108-118. Lefebvre, X., S. Lanini, and D. Houi, The role of aerobic activity on refuse temperature rise, 1. Landfill experimental study. Waste Management & Research, 2000. 18(5): p. 444-452. Green, R.B., W.G. Vogt, and RS. Sullivan. Comparison of emission from bioreactor and conventional subtitle D landfills. in U. S. EPA workshop on landfill bioreactors. 2000. Arlington, VA. Townsend, T.G., W.L. Miller, H]. Lee, and J .F.K. Earle, Acceleration of landfill stabilization using leachate recycle. Journal of Environmental Engineering- Asce, 1996. 122(4): p. 263-268. Van Lier, J .B., F. Boersma, M. Debets, and G. Lettinga, High-Rate Thermophilic Anaerobic Waste-Water Treatment in Compartrnentalized Upflow Reactors. Water Science and Technology, 1994. 30(12): p. 251-261. Chou, H.-H., J.-S. Huang, and W.-F. Hong, Temperature dependency of granule characteristics and kinetic behavior in UASB reactors. Journal of Chemical Technology and Biotechnology, 2004. 79: p. 797-808. El-Mashad, H.M., G. Zeeman, W.K.P.v. Loon, G.P.A. Bot, and G. Lettinga, Effect of temperature and temperature fluctuation on thermophilic anaerobic digestion of cattle manure. Bioresource Technology, 2004. 95: p. 191-201. Ugwuanyi, J.O., L.M. Harvey, and B. McNeil, Effect of digestion temperature and pH on treatment efficiency and evolution of volatile fatty acids during thermophilic aerobic digestion of model high strength agricultural waste. Bioresource Technology, 2005. 96: p. 707-719. 176 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. Kashyap, D.R., K.S. Dadhich, and SK. Shanna, Biomethanation under psychrophilic conditions: a review. Bioresource Technology, 2003. 87: p. 147153. Lanini, S., D. Houi, O. Aguilar, and X. Lefebvre, The role of aerobic activity on refuse temperature rise: 11. experimental and numerical modeling Waste Management & Research, 2001. 19: p. 58-69. VanderGheynst, J.S., L.P. Walker, and J.-Y. Parlange, Energy transport in high- solids aerobic degradation process: mathematical modeling and analysis. Biotechnology progress, 1997. 13: p. 238-248. Vollertsen, J., M.D.C. Ahneida, and T. Hvitved-Jacobsen, Effects of temperature and dissolved oxygen on hydrolysis of sewer solids. Water Research, 1999. 33(14): p. 3119-3126. Hu, H.-Y., K. Fujie, and K. Urano, Effect of Temperature on the Reaction Rate of Bacteria Inhabiting the Aerobic Microbial Film for Wastewater Treatment journal of Fermentation and Bioengineering, 1994. 78(1): p. 100-104. Kulcu, R. and O. Yaldiz, Determination of aeration rate and kinetics of composting some agricultural wastes. Bioresource Technology, 2004. 93: p. 49- 57 . Zinder, S.H., Microbiology of Anaerobic Conversion of Organic Waste to Methane: Recent Developments. ASM News, 1984. 50(7): p. 294-298. Shen, CF. and SR. Guiot, Long-Term Impact of Dissolved 02 on the Activity of Anaerobic Granules. Biotechnology and Bioengineering, 1996. 49: p. 611-620. Liu, J.-S., I.W. Marison, and U. von Stockar, Microbial Growth by a Net Heat Up-take: A Calorimetric and Thermodynamic Study on Acetotrophic Methanogenesis by Methanosarcina bakeri. Biotechnology and Bioengineering, 2001. 75(2): p. 170-180. Haarstrick, A., DC. Hempel, L. Ostermann, H. Ahrens, and D. Dinkler, Modeling of the biodegradation of organic matter in municipal landfills. Waste Management & Research, 2001. 19(320-331). Morrnile, M.R., K.R. Gurijala, J.A. Robinson, M.J. McInemey, and J .M. Suflita, The importance of Hydrogen in Landfill F ermentations. Applied and Environmental Microbiology, 1996. 62(5): p. 1583-1588. 177 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. Molfetta, AD. and A. Sethi. Heat generation and transport from MSW landfill to groundwater. in Sardinia 2001, Eighth International Waste Management and Landfill Symposium. 2001. Sardinia, Italy: CISA, Environmental Sanitary Engineering Center. Haug, R.T., The practical handbook of compost engineering. 1993: Boca Raton: Lewis Publishers. James, A.G., J.A. Watson-Craik, and E. Senior, Use of Three-Stage Laboratory Model to Determine the Effects of Temperature on Individual Trophic Groups of Methanogenic Associations Isolated from Anoxic Landfill. Journal of Chemical Technology and Biotechnology, 1996. 67: p. 333-338. Norstrom, J.M., H.J. Bourque, and RA. Smith. Operation and monitoring of a bioreactor landfill in Sainte Sophie, Quebec. in SWANA 7th Annual Landfill Symposium. 2002. Louisville, Kentucky: SWANA. Yazdani, R., J. Kieffer, H. Akau, and D. Augenstein, Monitoring the performance of anaerobic landfill cells with fluids recirculation. 2003, Yolo County, Planning and Public Works Department- Division of integrated Waste Management: Woodland. p. 108. Yoshida, H. and H. Hozumi. Theoretical Study on temperature distribution in sanitary landfill. in proceedings of the second international congress on Environmental Geotechnics. 1996. Osaka, japan. Reinhart, D.R., T.G. Townsend, and J. Timmons. Florida bioreactor demonstration project - Instrumentation. in Waste Tech 2002- Landfill Conference. 2002. http://srdata.nist.gov/insulation/Insul query result10.a_sp. http://hypertextbool_<.com/nhysics/thermal/conduction. Johnson, W.H. and J. Baker. Operational characteristics and enhanced bioreduction of municipal waste landfill mass by a controlled aerobic process. in Proceedings from SWANA's 4th Annual Landfill Symposium. 1999. Denever, Colorado: SWANA. Read, A.D., M. Hudgins, S. Harper, P. Phillips, and J. Morris, The successful demonstration of aerobic landfilling - The potential for a more sustainable solid waste management approach? Resources Conservation and Recycling, 2001. 32(2): p. 115-146. 178 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. Read, A.D., M. Hudgins, and P. Phillips, Aerobic landfill test cells and their implications for sustainable waste disposal. Geographical Journal, 2001. 167: p. 235-247. Read, A.D., M. Hudgins, and P. Phillips, Perpetual landfilling through aeration of the waste mass; lessons from test cells in georgia (USA). Waste Management, 2001. 21(7): p. 617-629. Lee, J.Y., C.H. Lee, and K.K. Lee, Evaluation of air injection and extraction tests in a landfill site in Korea: implications for landfill management Environmental Giology, 2002. 42: p. 945-954. Wolfe, K. Applied Research: Williamson County Aerated Bioreactor Landfill. in 2nd Intercontinental Landfill Research Symposium. 2002. Asheville NC, USA. Hater, G.R., G. Hamblin, and D.T. Lane. F ull-Scale Demonstration of a Sequential Aerobic-anaerobic Landfill Bioreactor. in Waste Tech 2000. 2000. Orlando- Florida. Smith, M.C., D.K. Gattie, D.D.H. Boothe, and KC. Das, Enhancing aerobic bioreduction under controlled conditions in a municipal solid waste landfill through the use of air injection and water recirculation. Advances in Environmental Research, 2000. 3(4): p. 459-471. Marison, I., A calorimetric investigation of the aerobic cultivation of Kluyveromyces fragilis on various substrates. Enzyme Microb. Technol., 1987. 9(Jan): p. 32-43. Cooney, C.L., D.I. Wang, and RI. Mateles, Measurement of heat evolution and correlation with oxygen consumption during microbial growth Biotechnology and Bioengineering, 1968. 11(3): p. 269-281. Biron, B., I.W. Marison, and UV. Stockar, Calorimetric investigation of aerobic ferrnentations. Biotechnology and Bioengineering, 1987. 30: p. 650-660. Mou, D.-G. and CL. Cooney, Application of dynamic calorimetry for monitoring fermentation processes. Biotechnology and Bioengineering, 1976. 18: p. 1371- 1392. Marison, I.W. and UV. Stockar, The application of novel heat flux calorimeter for studying growth of Escherichia coli W in aerobic batch culture. Biotechnology and Bioengineering, 1986. 28: p. 1780-1793. 179 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. Turker, M., Measurement of metabolic heat in a production-scale bioreactor by continuous and dynamic calorimetry. Chemical Engineering Commission, 2003. 190: p. 573-598. Luong, J .H.T. and B. Volesky, Determination of the heat of some aerobic fermentations. The Canadian Journal of Chemical Engineering, 1980. 58: p. 497-504. Ishikawa, Y. and M. Shoda, Calorimetri analysis of Escherichia coli in continuous culture. Biotechnology and Bioengineering, 1983. 25: p. 1817-1827. Bach, P.D., K. Nakasaki, M. Shoda, and H. Kubota, Thermal balance in composting operations. Ferment. Technology, 1987 . 65(2): p. 199-209. Adani, F., G. Gigliotti, F. Valentini, and R. Laraia, Respiration index Determination: A Comparative Study of Different Methods. Compost Science & Utilization, 2003. 11(2): p. 144-151. Tremier, A., A.d. Guardia, C. Massiani, E. Paul, and IL Martel, A respirometric method for characterising the organic composition and biodegradation kinetics and the temperature influence on the biodegradation kinetics, for a mixture of sludge and bulking agent to be co-composted Bioresource Technology, 2005. 96: p. 169-180. Veeken, A., V.d. Wilde, H. Woelders, and B. Harnelers, Advanced biocoversion of biowaste for production of a peat substitute and renewable energy. Bioresource Technology, 2004. 92: p. 121-131. Gea, T., R. Barrena, A. Artola, and A. Sanchez, Monitoring the biological activity of the composting process: oxygen uptake rate (OUR), rspirometric index (RI), and respiratory quotient (RQ). Biotechnology and Bioengineering, 2004. 88(4): p. 520-527. Scaglia, B., F. Tarnbone, P.L. Genevini, and F. Adani, Respiration Index Determination: Dynamic and Static Approaches. Compost Science & Utilization, 2000. 8(2): p. 90-98. Adani, F ., P. Lozzi, and P. Genevini, Determination of Biological Stability by Oxygen Uptake Rate on Municipal Solid Waste and Derived Products. Compost Science & Utilization, 2001. 9(2): p. 163-178. Eggen, T. and O. Vethe, Stability Indices for Different Composts. Compost Science & Utilization, 2001. 9(1): p. 19-26. 180 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. Richard, D. and R. Zimmerman, Respiration Rate - Reheating Potential: A Comparison of Measures of Compost Stability. Compost Science & Utilization, 1995. 3(2): p. 74-79. Koening, A. and Q.H. Bari, Application of Self-Heating Test for Indirect Estimation of Respirometric Activity of Compost: Theory and Practice. Compost Science & Utilization, 2000. 8(2): p. 99-107. Adani, F., P. Genevini, and F . Tambone, A New Index of Organic Matter Stability. Compost Science & Utilization, 1995. 3(2): p. 25-37. Iannotti, D.A., M.E. Grebus, B.L. Toth, L.V. Madden, and HA]. Hoitink, Oxygen Respirometry to Assess Stability and Maturity of Composted Municipal Solid Waste. Journal of Environmental Quality, 1994. 23: p. 1177-1183. Lasaridi, KB. and BI. Stentiford, A simple respirometeric technique for assessing compost stability. Water Research, 1998. 32(12): p. 3717-3723. Randolph, T.W., I.W. Marison, D.E. Martens, and UV. Stockar, Calorimetric Control of Fed-Batch Fermentations. Biotechnology and Bioengineering, 1990. 36: p. 678-684. Liang, C., K.C. Das, and R.W. McClendon, The influence of temperature and moisture contents regimes on the aerobic microbial activity of a biosolids composting blend Bioresource Technology, 2003. 86: p. 131-137. Adani, F., L. Scatigna, and P. Genevini, Biostabilization of mechanically separated municipal solid waste fraction Waste Management & Research, 2000. 18(5): p. 471-477. Finger, S.M., R.T. Hatch, and TM. Regan, Aerobic Microbial grth in semisolid matrices: Heat and mass transfer limitations. Biotechnology and Bioengineering, 1976. 18: p. 1193-1218. Liwarska-Bizukojc, E. and S. Ledakowicz, Stoichiometry of the aerobic biodegradation of the organic fraction of municipal solid waste (MSW). Biodegradation, 2003. 14(1): p. 51-56. Nielsen, H. and L. Berthelsen, A Model for temperature Dependency of Therrnophilic Composting Process Rate. Compost Science & Utilization, 2002. 10(3): p. 249-257. 181 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. Weppen, P., Determining Compost Maturity: Evaluation of Analytical Properties. Compost Science & Utilization, 2002. 10(1): p. 6-15. Beck-Friis, B., S. Smars, H. Jonsson, Y. Eklind, and H. Kirchmann, Composting of source-separated household organics at different oxygen levels: Gaining an understanding of emission dynamics. Compost Science & Utilization, 2003. 11(1): p. 41-50. Veeken, A., V.d. Wilde, and B. Hamelers, Passively Aerated Coomposting of Straw-Rich Pig Manure: Effect of compost bed porosity. Compost Science & Utilization, 2002. 10(2): p. 114-128. Oppenheimer, J .R., A.G. Martin, and LP. Walker, Measurements of air-filled porosity in unsaturated organic matrices using a pycnometer. Bioresource Technology, 1997. 59: p. 241-247. Richard, T.L., B. Hamelers, A. Veeken, and T. Silva, Moisture Relationships in Composting Processes. Compost Science & Utilization, 2002. 10(4): p. 286-302. Ahring, B.K., A.A. Ibrahim, and Z. Mladenovska, Effect of temperature increase from 55 to 65 degrees C on performance and microbial population dynamics of an anaerobic reactor treating cattle manure. Water Research, 2001. 35(10): p. 2446- 2452. Yadvika, Santosh, T.R. Sreekrishnan, S. Kohli, and V. Rana, Enhancement of biogas production from solid substrates using different techniques - a review. Bioresource Technology, 2004. 95: p. 1-10. de Cortazar, A.L.G., J.H. Lantaron, O.M. Fernandez, I.T. Monzon, and M.F. Lamia, Modelling for environmental assessment of municipal solid waste landfills (Part 2: Biodegradation). Waste Management & Research, 2002. 20: p. 514- 528. Veeken, A., S. Kalyuzhnyi, H. Scharff, and B. Hamelers, Effect of pH and VFA on hydrolysis of organic solid waste. Journal of Environmental Engineering- Asce, 2000. 126(12): p. 1076-1081. Masse, DJ. and R.L. Droste, Comprehensive model of anaerobic digestion of swine manure slurry in a sequencing batch reactor. Water Research, 2000. 34(12): p. 3087-3106. 182 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. F ukuzaki, S., N. Nishio, and S. Nagai, Kinetics of the Methanogenic Fermentation of Acetate. Applied and Environmental Microbiology, 1990. 56(10): p. 3158- 3163. Moletta, R., D. Verrier, and G. Albagnac, Dynamic Modeling of Anaerobic Digestion. Water Research, 1986. 20(4): p. 427-434. Vavilin, V.A. and L.Y. Lokshina, Modeling of Volatile Fatty Acids Degradation Kinetics and Evaluation of Microorganism Activity. Bioresource Technology, 1996. 57: p. 69-80. Barlaz, M.A., D.M. Schaefer, and R.K. Ham, Bacterial population development and chemical characteristics of refuse decomposition in a simulted sanitary landfill. Applied and Environmental Microbiology, 1989. 55: p. 55-65. Wang, Y.S., W.S. Odle, W.E. Eleazer, and MA. Barlaz, Methane potential of food waste and anaerobic toxicity of leachate produced during food waste decomposition. Waste Management & Research, 1997. 15(2): p. 149-167. Pohland, F.G. and J .C. Kim, Microbially mediated attenuation potential of landfill bioreactor systems. Water Science and Technology, 2000. 41(3): p. 247-254. Cecchi, F., P. Pavan, A. Musacco, J. Mataalvarez, C. Sans, and D. Defaveri, Monitoring a Fast Thermophilic Re-Start-up of a Digester Treating the Organic Fraction of Municipal Solid-Waste. Environmental Technology, 1993. 14(6): p. 517-530. Vieitez, E.R., J. Mosquera, and S. Ghosh, Kinetics of accelerated solid-state fermentation of organic- rich municipal solid waste. Water Science and Technology, 2000. 41(3): p. 231-238. Griffin, M.E., K.D. McMahon, R.I. Mackie, and L. Raskin, Methanogenic Population Dynamics during Start-Up of Anarobic Digesters Treating Municipal Solid Waste and Biosolids. Biotechnology and Bioengineering, 1998. 57(3): p. 342-355. Masse, D.I., L. Masse, A. Verville, and S. Bilodeau, The Start-up of anaerobic sequencing batch reactors at 20C and 25C for the treatment of slaughterhouse wastewater. Journal of Chemical Technology and Biotechnology, 2001. 76: p. 393-400. 183 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. Nachaiyasit, S. and DC. Stuckey, Effect of Low Temperatures on the Performance of an Anaerobic Baffled Reactor (ABR). Journal of Chemical Technology and Biotechnology, 1997. 69: p. 276-284. Liu, J ., G. Olsson, and B. Matiasson, Monitoring and Control of an Anaerobic Upflow Fixed-Bed Reactor for Hi gh-Loading-Rate Operation and Rejection of Disturbances. Biotechnology and Bioengineering, 2004. 87(1): p. 43-53. Vavilin, V.A., L.Y. Lokshina, S.V. Rytov, and OR. Kotsyurbenko, Description of two-step kinetics in methane formation during psychrophilic H2/C02 and mesophilic glucose conversions. Bioresource Technology, 2000. 71: p. 195-209. Kettunen, RH. and J .A. Rintala, The effect of low temperature (5-29 degrees C) and adaptation on the methanogenic activity of biomass. Applied Microbiology and Biotechnology, 1997. 48(4): p. 570-576. Muralidharan, V., K.D. Rinker, I.S. Hirsh, E.J. Bouwer, and RM. Kelly, Hydrogen Transfer Between Methanogens and F errnentive Heterotrophs in Hyperthermophilic Cultures. Biotechnology and Bioengineering, 1997. 56(3): p. 268-278. Schmidt, IE. and BK. Ahring, Review: Granular Sludge Formation in Upflow Anaerobic Sludge Blanket (UASB) Reactors. Biotechnology and Bioengineering, 1996. 49: p. 229-246. Dinopoulour, G., R.M. Sterritt, and J.N. Lester, Anaerobic Acidogenesis of a Complex Wastewater: 11. Kinetics of Growth, Inhibition, and Product Formation. Biotechnology and Bioengineering, 1987. 31: p. 969-978. Cord-Ruwisch, R., T.I. Mercz, C.-Y. Hoh, and GE. Strong, Dissolved Hydrogen Concentration as an On-Line Control Parameter for the Automated Operation and Optimization of Anaerobic Digesters. Biotechnology and Bioengineering, 1997. 56(6): p. 627-634. Fukuzaki, S., N. Nishio, M. Shobayashi, and S. Nagai, Inhibition of the Fermentation of Propionate to Methane by Hydrogen, Acetate, and Propionate. Applied and Environmental Microbiology, 1990. 56(3): p. 719-723. Hill, D.T. and SA. Cobb, Modeling Predictive Indicators in Animal Waste Methanogenesis. Transactions of the Asae, 1993. 36(3): p. 879-885. 184 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. Ahring, B.K., P. Westermann, and RA. Mah, Hydrogen inhibition of acetate metabolism and kinetics of hydrogen consumption by Methanosarcina therrnophila TM-l. Archives of Microbiology, 1991. 157: p. 38-43. Angelidaki, I., L. Ellegaard, and BK. Ahring, A Comprehensive Model of Anaerobic Bioconversion of Complex Substrates to Biogas. Biotechnology and Bioengineering, 1999. 63(3): p. 363-372. Zhao, X., S. Maher, R. Musleh, M. Khire, T. Voice, and S. Hashsham. Bioreactor landfill research and demonstration project at the Northern Oaks Landfill, MI. in Waste Tech. 2003. New Orleans, Louisiana: NSWMA WASTEC. Lide, D.R., ed. CRC handbook of chemistry and physics. 72 ed. 1991/ 1992, CRC Press, Inc. Montgomery, J .M., Water Treatment Principles and Design. 1985, New York: Wiley-Interscience. Snoeyink, V.L. and D. Jenkins, Water Chemistry. 1980, New York: John Wiley & Sons. Sponza, D.T. and M. Isik, Decolorization and inhibition kinetic of Direct Black 39 azo dye with granulated anaerobic sludge. Enzyme and Microbial Technology, 2004. 34: p. 147-158. Magbanua, B.S.J., Y.-T. LU, and C.P.L.G. JR, A technique for obtaining representative biokinetic parameter values from replicate sets of parameter estimates. Water Research, 1998. 32(3): p. 849-855. Ribes, J ., K. Keesman, and H. Spanjers, Modelling anaerobic biomass grth kinetics with a substrate threshold concentration. Water research, 2004. 38: p. 45 02-45 10. Goudar, CT. and T.G. Ellis, Explicit oxygen concentration expression for estimating extant biodegradation kinetics fiom respirometric experiments. Biotechnology and Bioengineering, 2001. 75(1). Spear, J .R., L.A. Figueroa, and 3D. Honeyman, Modeling the removal of uranium U(VI) from aqueous solutions in the presence of sulfate reducing bacteria environmental Science and Technology, 1999. 33‘(15): p. 2667-2675. 185 118. 119. 120. 121. 122. 123. 124. Knightes, CD. and CA. Peters, Statistical Analysis of Nonlinear parameter estimation for Monod biodegradtion kinetics using bivariate data. Biotechnology and Bioengineering, 2000. 69(2). Smith, L.H., P.L. McCarty, and PK. Kitanidis, Spreadsheet method for evaluation of biochemical reaction rate coefficients and their uncertainties by weighted nonlinear least-squares analysis of the integrated monod equation Applied and Environmental Microbiology, 1998. 64(6): p. 2044-2050. Liu, L., Z. Wang, J. Yao, X. Sun, and W. Cai, Investigation on the properties and kinetics of glucose-fed aerobic granular sludge. Enzyme and microbial technology, 2005. 36: p. 307-313. Fu, W. and AP. Mathews, Lactic acid production from lactose by Lactobacillus plantarum: kinetic model and effects of pH, substrate, and oxygen Biochemical Engineering Journal, 1999. 3: p. 163-170. Alonso, C., X. Zhu, M.T. Suidan, B.R. Kim, and B.J. Kim, Parameter estimation in biofilter systems. environmental Science and Technology, 2000. 34: p. 2318- 2323. Tien, TH. and A. Katayama, Biodegradation kinetics of volatile hydrophobic organic compounds in cultures with variable fractional volumes. Biotechnology and Bioengineering, 2004. 85(6). Grannemann, N.G., R.J. Hunt, J .R. Nicholas, T.E. Reilly, and TC. Winter, The Importance of Ground Water in the Great Lakes Region. 2000, US. GEOLOGICAL SURVEY: Water-Resources 125. 126. 127. 128. Investigations: Lansing, Michigan. p. 19. Gass, T.E., Geothermal Heat Pumps. Geothermal Resources Council Bulletin, 1982. 11(11): p. 3-8. Aichele, S.S., Ground-Water Quality Atlas of Oakland County, Michigan. 2000, US Geological Survey Water-Resources Investigations. Todd, D.K., Groundwater Hydrology. 1980, New York: John Wiley and Sons. 516. Dubrunfaut, M., CR. Hebd. Seances Acad. Sci., 1856. 42: p. 945. 186 129. Association, A.W.W., Water quality and treatment : a handbook of community water supplies, ed. R.D. Letterman. 1999, New York: McGraw-Hill. 130. Speece, R.E., Anaerobic biotechnology for industrial wastewaters. 1996, Nashville: Archae Press. 187