Вы находитесь на странице: 1из 313

SIMULATION MODELING AND PROCESS CONTROL OF COMPOSTING SYSTEMS UNDER COMPLEXITY AND UNCERTAINTY

A Thesis

Submitted to the Faculty of Graduate Studies and Research in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in Engineering University of Regina

by Yupeng Lin Regina, Saskatchewan January, 2006

Copyright 2006: Y. P. Lin

ABSTRACT
In this dissertation research, methodologies of modeling and control for composting processes that are associated with a variety of complexities and uncertainties have been developed. They include (i) a multi-component modeling system for simulating aerobic composting process, (ii) a simulation-based factorial analysis approach to characterize the interactive effects of multiple factors on composting process, (iii) a fractional fuzzy vertex method combined with interval analysis to tackle the effects of uncertainties on composting process, (iv) a model-based real-time neural predictive control system for managing composting process, and (v) a stepwise-inference-based real-time predictive control system for composting process. In addition, a series of composting experiments have been undertaken to verify the developed model and justify the real-time process control systems. The multi-component modeling system has been developed to simulate composting process based on the principles of microbiology, mass transfer, and biotechnology. The developed model was verified through a series of composting experiments, and the simulation results were generally consistent with the experimental outputs with acceptable accuracy levels. The developed model represented a unique contribution to the fields of composting process modeling and solid waste management. The simulation-based factorial analysis method has been developed to characterize the interactive effects of multiple factors on a composting process. It is based on the developed composting-process model and the two-level factorial analysis method. In addition, normal probability plot has been produced to support the system analysis. The study results could be useful for guiding composting-process operation and
II

management. Based on the developed composting-process model, the effects of uncertainties on a composting process have been characterized through a fractional fuzzy vertex method combined with interval analysis. The effects of uncertain control factors and input parameters as well as their interactions have been studied. In addition, the effects of uncertain control factors have been quantified through a series of sensitivity analyses. The model-based real-time neural predictive process control system (MNPC) has been developed for supporting the management of composting process. A series of composting experiments have been undertaken to train and justify the developed MNPC. The testing results showed that the predicted values were generally consistent with the experimental ones, indicating that the developed MNPC was satisfactory. The real-time composting process control system based on stepwise cluster analysis and GA optimization technology (SIPC) has been proposed as an advanced alternative to the MNPC. The stepwise cluster analysis was used to establish a stepwise-inference (SI) controller and an SI emulator based on a series of experimental studies. The genetic algorithm technique was employed to identify a desired control action. The justification results showed that the developed SIPC could effectively tackle the dynamics and complexity of composting process and support the relevant management and control practices.

III

ACKNOWLEDGEMENTS
I want to express my sincere appreciation to my supervisor, Dr. G. H. Huang, for his extremely helpful guidance, support and encouragement with his wisdom, knowledge and experience. His countless hours spent for guiding both my writing and research efforts have been extremely helpful for the successful completion of this dissertation. My appreciation also extends to his family. I would like to express my gratitude to my committee members, Dr. Rene Mayorga, Dr. Mingzhe Dong, Dr. Maynard Chen and Dr. Dianliang Deng, for their insightful suggestions, which were helpful for improving this thesis works. I also gratefully acknowledge the Faculty of Graduate Studies and Research and the Faculty of Engineering at the University of Regina for providing research scholarships and assistantships for my Ph.D. study. I would also like to send my special thanks to Dr. Guangming Zeng, Chair of the Department of Environmental Science and Engineering at the Hunan University. Dr. Zengs persistent encouragement and help are always invaluable assets to my endeavors. My further appreciation goes to Dr. Jianbing Li for his kind helps. Many thanks are also due to my friends in the EVSE research group for their assistance in various aspects. They include Xiaosheng Qin, Li He, Xianghui Nie, Bing Chen, Baiyu Zhang, Yimei Zhang, Kai An, Xueling Sun, Hongwei Lu, Renfei Liao, Wei Sun and many others. I appreciate discussions and interactions with them during my Ph.D. study. I also want to express my special thanks to many classmates and friends in China and North America. Their continuous support and encouragement have meant the world to me. Finally, I would particularly like to acknowledge my father, mother, sister, brother, and many other family members for their affection and moral support. I will always be in their debt.
IV

TABLE CONTENTS

ABSTRACT ACKNOWLEDGEMENTS LIST OF TABLES LIST OF FIGURES

II IV X XII

CHAPTER 1 INTRODUCTION ....1


1.1. Background.....1 1.2. Challenges in Modeling Composting Processes.2 1.3. Challenges in Composting Process Control...2 1.4. Objectives.....4

CHAPTER 2 LITERATURE REVIEW..6


2.1. Background....................................................................................................6 2.2. Modeling of Composting Processes.................................................................8 2.3. Studies in Composting Process Control....11 2.3.1. Composting as a Controlled Process...................................................11 2.3.2. Parameters Considered for Process Control........................................13 2.3.3. Experimental Studies of Composting Process Control...21 2.3.4. Composting Process Control for Field-Scale Experiments.............23 2.4. Process Control Studies in Other Areas.26 2.4.1. Model Predictive Control....26
V

2.4.2. Intelligent Process Control..................................................................28 2.5. Summary..........40

CHAPTER 3 A MULTI-COMPONENT MODELING SYSTEM FOR SIMULATING AEROBIC COMPOSTING PROCESS....43


3.1. Statement of Problems....43 3.2. Model Development.....................................................................................44 3.2.1. State Variables.....................................................................................46 3.2.2. Conversion Reactions..46 3.2.3. Modeling Formulation.49 3.3. Modeling Inputs and Outputs........................................................53 3.4. Model Solution................55 3.5. Model Verification...........59 3.6. Effects of Modeling Parameters............66 3.7. Summary......74

CHAPTER 4 SIMULATION-BASED FACTORIAL ANALYSIS FOR CHARACTERIZING INTERACTIVE EFFECTS FROM MULTIPLE FACTORS..................86
4.1. Statement of Problems86 4.2. Factorial Analysis of Interactive Effects in Composting Process........87 4.3. Case Study...91 4.3.1. Main Effects and Normal Probability Plot for Composting Substrates ............93

VI

4.3.2. Discussion......116 4.4. Summary117

CHAPTER 5 SIMULATION-BASED FRACTIONAL FUZZY VERTEX METHOD COMBINED WITH INTERVAL ANALYSIS FOR CHARACTERIZING EFFECTS OF UNCERTAINTIES.....................................................119
5.1. Statement of Problems..............................................................................119 5.2. Fuzzy Characterization on the Effects of Uncertainties....121 5.3. Effects of Control-Factor Uncertainties..130 5.4. Effects of Input-Parameter Uncertainties ..150 5.5. Combined Effects of Uncertainties..153 5.6. Discussion ..167 5.7. Summary167

CHAPTER 6 MNPC: MODEL-BASED REAL-TIME NEURAL PREDICTIVE CONTROL FOR WASTE COMPOSTING PROCESSES.................169
6.1. Statement of Problems..169 6.2. Neural Predictive Control for Composting Processes...170 6.2.1. Design of Neural Predictive Control for Composting Processes..171 6.2.2. Neural Network Modeling174 6.2.3. Nonlinear Optimization through Simulated Annealing.179 6.3. Application.........................................................................186 6.3.1. Experimental Studies.........................................................................186

VII

6.3.2. Training for the NN Emulator and Controller ..187 6.3.3. Result and Discussion...188 6.4. Summary....201

CHAPTER 7 SIPC: STEPWISE-INFERENCE-BASED REAL-TIME PREDICTIVE CONTROL FOR WASTE COMPOSTING PROCESSES..203
7.1. Statement of Problems..203 7.2. Research Framework....204 7.3. Stepwise Cluster Analysis.................205 7.3.1. Clustering Principles.207 7.3.2. Tests of Optimal Cutting Points209 7.3.3. Mergence of Clusters211 7.3.4. Prediction..211 7.4. Discrete and Nonlinear Optimization......212 7.4.1. Optimization for Optimal Control Conditions..212 7.4.2. Characteristics of Genetic Algorithm ...214 7.4.3. Procedure of Genetic Algorithm ..215 7.5. Application.....223 7.5.1. Data Collection .223 7.5.2. Stepwise Cluster Analysis 225 7.5.3. Result and Discussion ......226 7.5.4. Comparisons of SIPC and MNPC 231 7.6. Summary....241

VIII

CHAPTER 8 CONCLUSIONS.251
8.1. Summary....251 8.2. Research Achievements.253 8.3. Recommendations for Future Research..253

REFERENCES..255

IX

LIST OF TABLES

Table 3.1 Table 3.2 Table 4.1 Table 4.2 Table 4.3 Table 4.4 Table 4.5 Table 4.6 Table 4.7 Table 4.8 Table 4.9 Table 4.10 Table 4.11 Table 5.1 Table 5.2 Table 6.1 Table 6.2 Table 6.3 Table 6.4 Table 7.1

Nominal parameters values...54 Error analysis for model verification........................72 Algebraic signs for calculating effects in the 24 design90 The 24 factorial design..............92 Simulation results for the 24 design............................................................94 Main effects for soluble substrate concentration..........................................95 Data for normal probability plots (for the effects of soluble substrate concentration).......99 Main effects for insoluble substrate concentration.............102 Data for normal probability plot (for the effects of insoluble substrate concentration).......................................................................................103 Main effects for active biomass concentration....106 Data for normal probability plots (for the effects of active biomass concentration).109 Main effects for total substrate concentration.112 Data for normal probability plots (for the effects of total substrate concentration).........................................................................................113 Variations in interval widths....152 Variations in interval widths.................................................................166 Input and output variables for NN controller and emulator........................182 Ranges of data used to train the NN controller and emulator.189 Error analysis for controlled experiment conditions...193 Error analysis for the predicted results from MNPC......200 Input and output variables for the SI controller and emulator....................224
X

Table 7.2 Table 7.3 Table 7.4 Table 7.5

Error analysis for the SI controller and emulator........................................229 Error analysis for controlled experiment conditions...................................235 Error analysis for the predicted results of SIPC..........................................240 Error analyses for results from SIPC and MNPC.......................................243

XI

LIST OF FIGURES
Figure 3.1 Figure 3.2 Figure 3.3 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 Figure 3.9 Variations of OUA over time (200 hours)...60 Variations of OUR over time (200 hours).................................................61 Variations of soluble substrate concentration over time (120 hours).............62 Variations of active biomass concentration (200 hours)...63 Variations of insoluble substrate concentration over time (200 hours).64 Schematic diagram of the experimental reactor................................65 Predicted and observed OUA67 Predicted and observed OURs..........68 Predicted and observed soluble substrate concentrations.............................69

Figure 3.10 Predicted and observed insoluble substrate concentrations..........................70 Figure 3.11 Predicted and observed active biomass concentrations................................71 Figure 3.12 Effect of m on OUA................................................................................76 Figure 3.13 Effect of m on OUR................................................................................77 Figure 3.14 Effect of m on soluble substrate concentration.......................................78 Figure 3.15 Effect of m on biomass concentration.....................................................79 Figure 3.16 Effect of m on insoluble substrate concentration....................................80 Figure 3.17 Effect of b on OUA...................................................................................81 Figure 3.18 Effect of b on OUR....................................................................................82 Figure 3.19 Effect of b on soluble substrate concentration...........................................83 Figure 3.20 Effect of b on biomass concentration........................................................84 Figure 3.21 Effect of b on insoluble substrate concentration........................................85

XII

Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.8 Figure 5.1 Figure 5.2 Figure 5.3 Figure 5.4 Figure 5.5 Figure 5.6 Figure 5.7 Figure 5.8 Figure 5.9

Normal probability plot for soluble substrate concentration (the 72nd hour).97 Normal probability plot for soluble substrate concentration (the 144th hour)............................................................................................98 Normal probability plot for insoluble substrate concentration (the 72nd hour)...104 Normal probability plot for insoluble substrate concentration (the 144th hour)...105 Normal probability plot for active biomass concentration (the 72nd hour)...110 Normal probability plot for active biomass concentration (the 144th hour)...............................................................................................111 Normal probability plot for total substrate concentration (the 72nd hour)...................................................................................................114 Normal probability plot for total substrate concentration (the 144th hour)...............................................................................................115 Selection schemes of interior points in FV and FFV methods...127 Framework of the proposed FFV method..128 Membership functions of temperature...132 Membership functions of moisture133 Membership functions of oxygen content..........................134 Fuzzy modeling output of soluble substrate concentration (the 24th hour)...135 Fuzzy modeling output of soluble substrate concentration (the 72nd hour)...136 Fuzzy modeling output of insoluble substrate concentration (the 24th hour)...137 Fuzzy modeling output of insoluble substrate concentration (the 72nd hour)...138

XIII

Figure 5.10 Fuzzy modeling output of insoluble substrate concentration (the 144th hour)...139 Figure 5.11 Fuzzy modeling output of active biomass concentration (the 24th hour)...140 Figure 5.12 Fuzzy modeling output of active biomass concentration (the 72nd hour)...141 Figure 5.13 Fuzzy modeling output of active biomass concentration (the 144th hour)...142 Figure 5.14 Fuzzy modeling output of OUA (the 24th hour).143 Figure 5.15 Fuzzy modeling output of OUA (the 72nd hour)....144 Figure 5.16 Fuzzy modeling output of OUA (the 144th hour)...145 Figure 5.17 Effect of temperature on modeling outputs.147 Figure 5.18 Effect of oxygen content on modeling outputs...........148 Figure 5.19 Effect of moisture on modeling outputs..149 Figure 5.20 Modeling output of soluble substrate concentration (the 24th hour)..........154 Figure 5.21 Modeling output of soluble substrate concentration (the 72nd hour).........155 Figure 5.22 Modeling output of insoluble substrate concentration (the 24th hour).......156 Figure 5.23 Model output of insoluble substrate concentration (the 72nd hour)...157 Figure 5.24 Modeling output of insoluble substrate concentration (the 144th hour)158 Figure 5.25 Model output of active biomass concentration (the 24th hour)..159 Figure 5.26 Modeling output of active biomass concentration (the 72nd hour)160 Figure 5.27 Modeling output of active biomass concentration (the 144th hour)..161 Figure 5.28 Modeling output of OUA (the 24th hour)..........................................162 Figure 5.29 Modeling output of OUA (the 72nd hour)..163 Figure 5.30 Modeling output of OUA (the 144th hour)164 Figure 6.1 Neural predictive control system..172
XIV

Figure 6.2 Figure 6.3 Figure 6.4 Figure 6.5 Figure 6.6 Figure 6.7 Figure 6.8 Figure 6.9

NPC design for waste composting..........................................173 Structural representation of the BP network...176 The BP Network architecture of the NN controller180 The BP Network architecture of the NN emulator..181 Comparison of the controlled and predicted results for temperature190 Comparison of the controlled and predicted results for moisture..191 Comparison of the controlled and predicted results for oxygen content192 Predicted and observed results for soluble substrate concentration...........196

Figure 6.10 Predicted and controlled results for insoluble substrate concentration..197 Figure 6.11 Predicted and observed results for active biomass concentration..............198 Figure 6.12 Predicted and observed results for OUA...................................................199 Figure 7.1 Figure 7.2 Figure 7.3 Figure 7.4 Figure 7.5 Figure 7.6 Figure 7.7 Figure 7.8 Figure 7.9 Stepwise-inference-based process control.206 Procedure of a simple genetic algorithm...216 Decision variables encoding..220 Two points crossover operation.221 Cluster tree of temperature in the SI controller.227 Cluster tree of active biomass concentration in the SI emulator...228 Comparison of the controlled and predicted results for temperature.........232 Comparison of the controlled and predicted results for moisture..233 Comparison of the controlled and predicted results for oxygen content234

Figure 7.10 Predicted and observed results for soluble substrate concentration...236 Figure 7.11 Predicted and observed results for insoluble substrate concentration..........................237 Figure 7.12 Predicted and observed results for active biomass concentration..........................238
XV

Figure 7.13 Predicted and observed results for OUA...............................................239 Figure 7.14 Predicted temperature levels from SIPC and MNPC............................244 Figure 7.15 Predicted moisture levels from SIPC and MNPC.245 Figure 7.16 Predicted oxygen content levels from SIPC and MNPC.......246 Figure 7.17 Predicted soluble substrate concentrations from SIPC and MNPC...247 Figure 7.18 Predicted insoluble substrate concentrations from SIPC and MNPC........248 Figure 7.19 Predicted active biomass concentrations from SIPC and MNPC..249 Figure 7.20 Predicted OUAs from SIPC and MNPC250

XVI

CHAPTER 1 INTRODUCTION

1.1. Background The conservation and recycling of natural resources are significant

socio-economic and environmental concerns (Faucette, 2004). Composting provides a valuable means for mitigating pollution and recovering nutrients from organic solid wastes (Ekinci, 2001). However, an efficient performance for composting is generally hard to achieve as the whole process is affected by a variety of biochemical and physical factors and complicated with many nonlinearities and uncertainties (Hall, 1998). Process control in composting can significantly enhance process efficiency, pathogen destruction, and temperature maintenance (Park et al., 2004). Effective process control should be based on the insight into complicated multi-phase multi-component processes, as well as the capability to simulate the relevant process dynamics (Szmidt and Fox, 2001). Previously, many efforts on comprehending various aspects of the kinetics and dynamics of composting processes have been made, and a number of mathematical models have been developed to simulate the process (Kishimoto et al., 1987; Stombaugh and Nokes, 1996; Hall, 1998; Hamelers, 2001; Ekinci, 2001; Ryckeboer and Mergaert, 2003; Tirumalasetty, 2004). However, complete insight and effective control of composting processes have not been achieved. Consequently, studies into comprehensive simulation and real-time control of composting processes are desired.

1.2. Challenges in Modeling Composting Processes Composting is an effective approach to convert organic substrate to a stable humus-like product via biological degradation (Das and Keener, 1996). It can accelerate the biodegradation of organic wastes and produce stable products that can be used to improve soil properties (Haug, 1993). Effective modeling of composting processes would facilitate the operation and control of the system as well as the generation of reliable composting products. Previously, many studies were conducted to simulate composting processes (Kishimoto et al., 1987; Stombaugh and Nokes, 1996; Hall, 1998; Hamelers, 2001; Alpert, 2002; Ryckeboer and Mergaert, 2003). Many impact factors were considered, including temperature, moisture content, oxygen concentration, carbon dioxide content, energy flow, and substrate properties (Haug, 1993, 1996; Hall, 1998; Ekinci, 2001; Hamelelers, 2001; Trefry and Franzmann, 2003). However, a composting system is extremely complex and dynamic, involving a variety of biological, physical and chemical reactions (Cundiff and Mankin, 2003; Nakasaki et al., 2005). A number of attempts were undertaken in modeling composting processes in the past decades, with most of the developed models being empirical. Even though bio-kinetics was taken into account in a few models, they were too simple to thoroughly explore into composting processes. Therefore, development of a modeling system that incorporates the principles of microbiology, mass conversion, and biotechnology for composting processes is desired. 1.3. Challenges in Composting Process Control Process control is crucial for the management and operation of composting

systems (Christensen and Carlsb, 2002). Effective process control can improve compost quality and enhance process efficiency in composting systems (Hall, 1998). Previously, a number of studies on the composting process-control methodologies were reported, as shown in Jeris and Regan (1973a, b, c), MacGregor et al. (1981), Kishimoto et al. (1987), Keener et al. (1992), Robinson and Stentiford (1993), Kumar (1993), Haug (1993, 1996), Das and Keener (1996), Stentiford (1996), VanderGheynst et al. (1996), Oppenheimer (1997), and Patni and Kinsman (1997). A description of both theoretical and practical composting processes was given by Hall (1998). Ekinci (2001) provided an update on process control studies for composting systems, including experimental, modeling and field-scale work. More recent work related to composting process control included Nobuyuki et al. (2001), Nelson, et al. (2003), Ryckeboer and Mergaert (2003), and Tirumalasetty (2004). However, the previous studies of composting process control mainly focused on system dynamics and empirical control. There has been scarce of research on real-time process control of composting systems. In fact, a composting system is extremely complex and dynamic, changing over time as microbial populations grow, evolve, die, and/or are replaced by new populations (Cundiff and Mankin, 2003). Moreover, uncertainties exist in a variety of the system components, forming an additional category of complexities. These facts form a barrier in applying traditional control theories to composting systems. Consequently, approaches for real-time process control are desired. The dynamics, discreteness, and nonlinearity of composting systems warrant the demand for advanced process-control technologies. Fortunately, technologies of artificial neural networks (ANN) and stepwise cluster analysis (SCA) are capable of dealing with

complicated nonlinear and discrete problems (Zhang and Stanley, 1999; Sarimimveis and Bafas, 2003; Huang, 2004). Therefore, development of a real-time process control system based on the ANN and SCA technologies is desired. 1.4 Objectives As an extension of the previous efforts, this research emphasizes on (a) development of a multi-component simulation system that is able to characterize the dynamic interactive and uncertain effects of various factors that exist in composting systems, and (b) advancement of a real-time process control technique to support the composting operation. These objectives entail the following: (i) To develop a multi-component modeling system for simulating the aerobic composting process based on a series of experimental studies. (ii) To develop a simulation-based factorial analysis approach combined with normal probability plot to characterize the interactive effects from multiple factors. (iii) To develop a simulation-based fractional fuzzy vertex method combined with interval analysis to tackle the effects of uncertainties on composting processes. The influences of the uncertainties will be quantified. (iv) To develop a real-time process control system based on the multi-component simulation and the neural predictive control methodology and then justify the developed system through a series of composting experiments. (v) To further develop a real-time process control system based on stepwise inference and nonlinear optimization and then compare the two developed process-control systems based on the experimental work. This thesis is structured as follows. Chapter 2 reviews the previous studies of
4

waste composting, system simulation, and process control. Chapter 3 presents the development of a multi-component modeling system for simulating the aerobic composting process based on a series of experimental studies. Chapter 4 presents a simulation-based factorial analysis approach to characterize the interactive effects of multiple factors. Chapter 5 illustrates the development of a fractional fuzzy vertex method combined with interval analysis to tackle the effects of uncertainties on composting processes. Chapter 6 focuses on the development of a model-based neural predictive control system for composting processes, as well as the justification of the developed system based on a series of experimental studies. Chapter 7 presents a real-time process control system based on stepwise inference and nonlinear optimization. Chapter 8 presents conclusions of this dissertation research.

CHAPTER 2 LITERATURE REVIEW


2.1. Background Waste management has become a critical area of practice and research due to the increasing concerns of environmental pollution and resources shortage (Brewer, 2001). Most of solid waste management professionals recognize that there is no single, simple solution to solid waste problems. Instead, an integrated approach, combining the elements of multiple techniques, is used in an increasing number of cases (Uif, 1998; Fromme, 1999). Integrated solid waste management is a comprehensive strategy involving four key elements applied in hierarchical manner (Skinner, 1996): a) reduction of the volume and toxicity of solid waste; b) recycling or reuse for as much as possible of the generated waste; c) recovery of energy from the remaining waste through combustion systems equipped with the best available pollution control technology; d) utilization of landfills with adequate environmental controls. Composting can meet multiple objectives, including reduction of odor, recycling of nutrients to soils, and management of livestock wastes through an integrated approach (Haug, 1980, 1986; Miller, 1991; Hansen et al., 1993; Alpert et al., 2002; Park et al., 2004). Thus, composting offers a potential for producing a usable product from the organic fraction of solid waste and is thus a form of recycling. Composting can also produce combustible fuel for energy recovery (Brodie, 1996). By itself, composting is not a solution to all waste problems; however it can serve as a valuable component in a waste management plan (Lasoff, 2000). It advances the goals of decomposing putrescible material, decreasing volume, weight and water content, producing a stabilized process
6

residue, and inactivating pathogenic organisms (Godden, 1983; Hoitink and Keener, 1992; Finstein and Hogan, 1992; Szmidt and Fox, 2001). Composting may be defined as part of a sustainable resource management strategy (Richard, 1992; Golueke and Diaz, 1996; Faucette, 2004). It is essential to encourage recycling, the only sustainable waste management practice which avoids the existence itself of wastes by transforming possible waste materials into a series of products (Campbell, 1990; Sequi, 1996). With sustainable transformation of wastes into organic fertilizers, composting would complement sustainable agriculture (Cathcart et al., 1986). The sustainable agriculture and the use of compost can be considered as essential activities for a sustainable society (Sinha and Heart, 2002). Hence, sustainability considerations are major driving forces for composting technologies. Improvements in composting process control will help increase the efficiency and economic viability of the related technologies, and thus contributing to agricultural and societal sustainability (Molla, 2005). Composting is a significant method for recycling nutrients, reducing pollution and conserving resources (Ghiorse, 1996; Ryckeboer and Mergaert, 2003). It has been used for centuries as a mode of nutrient recycling in agricultural systems (Jacobowitz and Steenhuis, 1984). The use of composting products has been shown to sustain crop production at levels of quantity and quality similar to those obtained from chemical fertilizer (Finstein and Morris, 1975). Composted organic wastes can supply nutrients to plants in a balanced way, providing high yields with low risks of soil and ground water contamination (Rodrigues et al., 1996).

2.2. Modeling of Composting Processes Previously, many studies were undertaken in modeling composting processes (Whang and Meenaghan, 1980; Heijnen and Roele, 1981; Hamelers, 1992; Marugg, 1993; Hauhs, 1996; Stombaugh and Nokes, 1996; Mysliwiec et al., 2001; Nelson et al., 2003; Trefry and Franzmann, 2003; Hamelers, 2004; Johannessen et al., 2005). Jeris and Regan (1973a, b, c) explored various facets of composting processes. Haug (1993) discussed a variety of possible ways for modeling the dynamics of composting systems. A number of composting process models were proposed with some of them being verified through experiments (Finger, 1976; Suler and Finstein, 1977; Beck, 1979; De Bertoldi et al., 1983; Nakasaki, 1987; Das and Keener, 1996; VanderGheynst et al., 1996; Oppenheimer, 1997; Nakaya, 1999; Ekinci, 2001; Hamelers, 2004; Johannessen et al., 2005). The main modeling parameters include temperature, moisture, oxygen content, carbon dioxide, energy content, and material type (Haug, 1993, 1996; Richard, 1997; VanderGheynst et al., 1996, 1997; Hall, 1998; Hamelers, 2001; Ekinci, 2001; Johannessen et al., 2005). Many factors can affect composting processes, including temperature, moisture, gas flow and many other physical, chemical and biological variables (Beck, 1984; Ljung and Glad 1994; Ekinci, 2001). Various portions of composting processes were modeled, such as respiratory activities (Haug, 1993, 1996; Nobuyuki, 2001) and process kinetics (Richard, 1997; Bari, 2000). Respiration is typically a function of temperature, species concentration, nutrient level, and oxygen content. Conduction, convection and phase change were also considered. Conduction could be negligible under many situations (MacGregor et al., 1981; Haug, 1993, 1996; Ekinci, 2001; Trefry and Franzmann, 2003), especially in forced air systems with relatively high air flow rates (Hall, 1998). Air flow
8

rates will affect temperature and thus change moisture content. Therefore, aeration rate, moisture level, and temperature are interrelated. Smith and Eilers (1980) proposed a two dimensional finite difference model for forced aerated windrow composting of biosolids. They studied spatial and temporal solutions of airflow, substrate degradation, and heat, water and oxygen balances. In their model, it was assumed that the flow patterns and quantities of air were independent of time. The model accounted for spatial and temporal changes in dry matter and, therefore, microbial degradation rates. Field validation of the model proved that dry matter, volatile matter and moisture content were predicted accurately. Nakasaki (1987) developed a process model describing the composting of sewage sludge cake. Their approach was based on a heat and mass balance in a complete batch reactor. Forced aeration was applied in this system. The carbon dioxide release, the volatile-matter conversion, water content, and system temperature were calculated. Three experiments were performed to assess the validity of the model. A continuous feed complete mix reactor was presented by Haug (1993). The composting mass was schematically represented by gas, liquid and solid phases. Organic matter degradation was modeled as a first-order reaction with respect to the substrate concentration. The rate constant in the equation was a function of temperature. Rate limitations due to suboptimal values of the air filled volume fraction, moisture content and oxygen concentration in the gas phase were taken into account. Forced convection was used as a starting point in Haugs model. Free convection was only studied theoretically using a pore tube model, which limited the application of the model. Van Lier et al. (1994) studied phase-II composting of the mushroom substrate in

bulk fermentation tunnels. A mathematical model of the composting process was developed based on mass and heat transfer. Differential equations were solved through time-dependent analysis using a continuous simulation and modeling program. The substrate in the tunnel system was divided into equal layers that differed in density and porosity. The simulation results included oxygen demand, water and dry-matter losses, temperatures in various layers, and conductive heat loss through the walls of the containers of the tunnel. The simulation results were compared with experimental data. Stombaugh and Nokes (1996) started with microbial growth and death kinetics as the basis for their dynamic composting model. Substrate degradation rates were established using conversion factors based on stoichiometry of degradation and parameterization for reasonable values. The model was validated through comparison with a laboratory composting reactor while composting a mixture of cracked corn and palletized corncobs. The limited number of experiments conducted for validation of the model showed that the model could predict temperature fluctuation, oxygen uptake rate, moisture exchange, and substrate degradation for a readily composted input mixture. Das and Keener (1996) simulated air recycling as part of a numerical dynamic model in a large scale composting system, assuming that the inlet air temperature equal to 90% of that in the compost bed, and that the relative humidity was 98%. The oxygen depletion in the air was not computed, as the model did not directly relate degradation kinetics to oxygen levels. Results of the study showed that temperature differences between top and bottom layers of the bed were lower when air was recirculated. In addition, the highest total degradation within the cross-section was at the bottom of the bed (12% of the initial dry matter).

10

Hall (1998) constructed a numerical model using four primary variables: substrate, oxygen, moisture, and energy. Effects of changes in variables on system dynamics were explored. Open- and closed-loop simulations were performed with a primary focus on using aeration rate to control temperature. The aeration rates varied between 0.02 and 0.50 kgair/kgsubstrate/hour, with a constant flow of 0.04 kgair/kgsubstrate/hour providing the quickest heating profile. Because Hall (1998) mainly focused on the control of temperature and did not take the comprehensive effects of composting parameters into account, this composting control system is too simple to fully reflect the dynamics of the composting process. Hamelers (2001) developed a particle-level mathematical model based on biofilm theory by using polymeric substrate, microbial biomass, oxygen, and water concentrations as state variables. Reaction rates describing conversion of substrate and biomass growth were modeled using first-order-kinetic and Mechaelis-Menten multiple-substrate-dependent rates. The model was validated for the oxygen uptake rate of compost material using a 0.07 m diameter and 0.3 m long cylinder reactor. However, this study mainly focused on one composting particle, and the composting process was based on controlling the temperature of composting materials within a certain range. More recent studies related to composting-process model included Mysliwiec et al. (2001), Nelson et al. (2003), Trefry and Franzmann (2003), Hamelers (2004), and Johannessen et al. (2005). 2.3. Studies in Composting Process Control 2.3.1. Composting as a Controlled Process Many researchers studied composting as a controlled process. For example, a
11

number of studies on composting processes and their control were reported in Schulze (1960, 1961, and 1962), Jeris and Regan (1973a,b,c), Kishimoto et al. (1987), Keener et al. (1992), Walker, 1993; Robinson and Stentiford (1993), Koelsch (1994), Haug (1993, 1996), Das and Keener (1996), Stentiford (1996), Richard and Choi (1996), Hall and Aneshansley (1997), Oppenheimer (1997), Richard (1997), and VanderGheynst et al. (1997). An overall description of both theoretical and practical composting processes was given by Hall (1998). Understanding composting process via modeling and experimentation could aid in improving process control (VanderCheynst, 1997; Nakaya, 1999). Ekinci (2001) gave a recent update on process control studies in waste composting, including experimental, modeling and field-scale studies. More recent work related to composting process control included Szmidt and Fox (2001), Nobuyuki et al. (2001), Nelson, et al. (2003), Ryckeboer and Mergaert (2003), Rosenfeld et al. (2004), Tirumalasetty (2004), and Nakasaki et al. (2005). However, the previous studies mainly focused on system dynamics and empirical control as well as a few mathematical models. However, little work was undertaken in terms of real-time process control for composting systems, especially for model-based real-time process control. The challenge for process control involved the complex and inherent difficulty in prediction behaviour of biological systems (Harremoes and Madsen, 1999; Cundiff and Mankin, 2003). Process modelling and controllability studies were done for not only composting (Poincelot, 1977; Robinson and Stentiford, 1993; Haug 1993, 1996; Sanchez-Monedero et al., 1996; Hall, 1998; Nakaya, 1999; Ekinci, 2001) but also other biological systems, including food processing and fermentation (Griffin and Luard, 1979; Dochain and Bastin, 1984; Shi and Shimizu, 1992; Dahhou et al. 1992; Cook et al., 1997;

12

Rodrigo, et al., 1999; Thibault et al, 2000). Biological activities posed particular challenges in control, as they were non-linear, time varying, and often difficult to model (Nelson et al., 2003). In addition, problems associated with optimizing biological control systems included the difficulties in predicting the behaviours of system dynamics (Cundiff and Mankin, 2003). A composting system, for example, was extremely complex and dynamic, changing over time as microbial populations grow, evolve, die, and/or are replaced by new populations (McCarty, 1971; McKinley et al., 1985). The number of variables required to truly describe this system was unknown. General models (Heij and Willems, 1989) could not predict the system behaviour accurately. Hall (1998) proposed that a very simple and robust traditional system feedback scheme might work adequately for composting process control. However, there were significant variations in start-up time, ramp-up in temperature, and other major variables, which made the system highly nonlinear. Therefore, an attempt at a simple, robust traditional system control scheme could not address the dynamics and complexity of composting process completely. Advanced process control technologies were desired for such biological systems. 2.3.2. Parameters Considered for Process Control The success of a composting process depends on the management of oxygen, moisture, and energy transport within the system (Hoitink and Keener, 1993; VanderGheynst et al., 1997). Most of the previous studies used feedback control via temperature, moisture and oxygen measurement, either on-line or through analysis of experimental or modeling data (Faucette, 2004). The controlled parameters included air flow rate (Jeris and Regan, 1973a;
13

MacGregor, et al., 1981; Roig and Bernal, 1996; Ekinci, 2001; Iriarte and Ciria, 2001), substrate considerations (Jeris and Regan, 1973b; Stombaugh and Nokes, 1996; Trefry and Franzmann, 2003), density and other physical or geometric parameters (Hall, 1998; Hamelers, 2001), and moisture addition (Robinson and Stentiford, 1993; Stentiford, 1996; Zhang, 2000). Haug (1993, 1996) modeled the controlled process through analyzing air flow and variations in substrate type. Agitation and mixing were also investigated (VanderGheynst et al., 1996; Richard, 1997; Christensen and Carlsb, 2002). Fixed parameters such as density, substrate, and initial moisture should be optimized for a given material and composting system. Three main parameters were studied by many researchers, including oxygen or aeration rate, moisture, and temperature (Stombaugh and Nokes, 1996; Hall, 1998; Zhang, 2000; Trefry and Franzmann, 2003). Parameter I: Oxygen/Aeration Gas exchange supplies oxygen and removes carbon dioxide, heat, and water vapor. The significance of supplying oxygen is that aerobic respiration generates heat at a rate sufficient for self-heating (Cooney et al., 1968; Atchley and Clark, 1979; Rynk, 1991; Sesay, 1998). Furthermore, for temperature control which is accomplished by varying aeration rate, the relationship between heating due to high oxygen conditions and cooling due to convective and phase-change phenomena is critical. A number of researchers considered oxygen or carbon dioxide levels as feedback variables (Jeris and Regan, 1973a, b; Nakasaki and Shoda, 1987; DeBertoldi et al., 1988; Haug, 1993, 1996; Vinci, 1996; Bodelier and Laanbroek, 1997; Richard et al., 1999; Zhang, 2000; Ekinci, 2001). With a known flow rate, measurement of carbon dioxide or
14

oxygen should be sufficient to define the rate of degradation (Jeris and Regan, 1973a). It was generally accepted that oxygen levels below 5% might limit aerobic activities (Haug, 1993). However, temperature could also be a critical variable which indirectly reflected degradation rate, and might give a more direct method for controlling degradation rate as well as other functions such as pathogen destruction (Nakasaki et al., 2005). In addition, the relationship between moisture and oxygen availability is important. Excess moisture in the compost can reduce the air flow and hence reduce degradation rate. Jeris and Regan (1973b) considered the effects of the moisture level, the free air space within the pile, and the recycling of innocuous from previous composting into new compost mixes. Moisture and free air space were also closely related, as increased moisture content would generally decrease air space and hence decrease oxygen availability and degradation rate. Parameter II: Moisture/Water Content Water is a critical ecological as well as physical factor in substrate dense matrix ecosystems of which composting systems are an example (Miller, 1989; Baker and Allmaras, 1990). Sufficient moisture in the solid waste is required for maximum efficiency of microbial stabilization (Dirksen and Dasberg, 1993; Iriarte and Ciria, 2001). Water is both required for and produced by microbial activities. A major part of the composting process is the loss of water, as a result of evaporation as the process progresses (Stentiford, 1996). Moisture also affects porosity and gas diffusivity and is removed via vaporization (evaporative cooling), as driven by microbial heat generation (Dean et al., 1987; Oppenheimer, 1997). Finally, at low moisture contents (below 30%), microbial activity may be greatly reduced by the shortage of water (Nakaya, 1999; Iriarte
15

and Ciria, 2001). Conversely, when the moisture content is excessive, the interstices within the organic mass become filled with water and aeration is restricted. As the oxygen concentration is depleted within the material, an anaerobic condition develops which results in odors and a significant decrease in the rate of decomposition (Jeris and Regan, 1973b). Jeris and Regan (1973b) found that 95% of the maximum oxygen consumption rate was obtained when the moisture content was between 63 and 79%. Das and Keener (1996) studied moisture levels and compactability with regard to aeration. They found that wet materials, when compacted, could dramatically reduce aeration of compost. Avoiding excessive compaction, especially with wet materials, would help reduce fan power needed to supply demanded aeration in static beds. Other objectives of moisture management may include (Zhang, 2000): a) maintenance of optimal composting conditions, b) decreasing the amount of amendment required in agricultural applications, and c) elimination of excess water. In addition, moisture in compost may be increased by respiration and decreased by evaporation (with energy loss as well). When moisture decreases to a low level enough, additional moisture may help restart the process (Bakshi, 1987; Richard and Choi, 1996). The moisture content between 50 and 70% is most suitable for composting and should be maintained during the periods of active bacterial reactions, i.e. mesophilic and thermophilic growth (Polprasert, 1989; Robinson and Stentiford, 1993; Zhang, 2000). The moisture content of the compost affects the structural properties of the materials, the thermal properties of the materials as well as the rate of biodegradation (Stentiford, 1996; Nakasaki et al., 2004, 2005). Too little moisture slows down decomposition and prevents the pile from heating up. Too much moisture, signalled by a foul odor and a drop in temperature, drives out air,

16

drowns the pile and washes away nutrients (Inbar, 1990). Moisture content is typically reduced during the maturing process from about 60% to about 30%, which provides a stable compost product. The maximum rate of transfer of nutrients and waste products takes place in a liquid environment (100% moisture) (Zhang, 2000). However, it is not possible to operate a composting system on this basis where solid substrate is being used. For moisture content between 50 and 65% on a wet basis, degradation rate is not generally inhibited (Jeris and Regan, 1973a). Above 65%, anaerobic conditions may occur, and below 50% moisture, bacterial growth becomes a rate limiting factor (Keener et al., 1992). The optimum moisture level varies with materials because of their specific water holding properties (Zhang, 2000). However, it is clear that maintaining sufficient moisture (>30-40%) but avoiding excess (> 65-75%) for a given material is necessary to maintain effective degradation.

Parameter III: Temperature The temperature of compost is a function of the accumulation of heat generated metabolically and, simultaneously, the temperature is a determinant of metabolic activities (MacGregor et al., 1981). It is a critical variable in the composting process and a number of studies have addressed heat and mass transport in composting processes (Finger et al., 1976; Characklis and Gujer, 1979; Luong and Volesky, 1983; Incropera and DeWitt, 1985; Macky and Derrick, 1986; Kishimoto et al., 1987; Bach et al., 1987; Keener et al., 1992;

Nakasaki and Akiyama, 1988; Richard and Walker, 1989;

VanderGheynst et al., 1997; Hall, 1998; Ekinci, 2001; Nakasaki et al., 2004, 2005).

17

Temperature control is important for pathogen destruction, respiration-rate optimizing, moisture removal, and compost stabilization (Soares et al., 1995). The thermophilic temperatures reached during the composting process could reduce pathogen concentrations dramatically and provide for a safe compost product (Cooney and Wise, 1975; Polprasert, 1989). Haug (1993) considered the attainment of thermophilic conditions via heat from biological activity to be a necessary part of the composting process. Attaining and maintaining a sufficient temperature for a specified period of time is necessary for consistent pathogen destruction (Pereira-Neto et al., 1987). In addition, maintaining an optimal temperature (specific temperature as yet to be determined) in the thermophilic range may provide conditions for quicker and more complete composting. At extremely high temperatures, proteins may be denatured and biological activity may decline as populations die, thus ultimately slowing the composting process. Maintaining a consistent optimal temperature may provide for faster and more complete composting, pathogen destruction, and production of stable final compost. Use of a blower to force air through a composting pile greatly increases ventilative heat removal, leaving conduction to play only a minor role (Sartai, et al., 1995). Calculation based on such a pile indicates that approximately 98% of the heat removal is through ventilative mechanisms and the remainder through conduction (Haug, 1993). In forced aeration systems, more air is required to remove heat than to supply oxygen (Haug, 1993). Schulze (1962) proposed a log-linear fit of data and simultaneously suggested that higher temperatures produced exponentially higher respiration rates. Jeris and Regan

18

(1973a) hypothesized that temperatures above 70C would produce slower respiration rates, and proposed a second-order curve fit to their data which peaked near 60C. For newsprint and stabilized refuse, the peak rates occurred at temperatures between 30 and 50C. Unfortunately, neither Jeris and Regan (1973a, b, c) nor Schulze (1961) did experiments above 70C in their studies. MacGregor et al. (1981) discussed both temperature self-limiting composting systems as well as those that were not. The self-limiting system reached inhibitive temperatures (>60C) which debilitated the microbial community, suppressing decomposition, heat output and water removal. In contrast, non-self-limiting temperature (< 60C) supported a robust community, promoting decomposition, heat output and water removal. The thermophilic community in self-heating organic masses was most active at approximately 55 to 60C, and higher temperatures tended to slow respiration (Hall, 1998; and Ekinci, 2001). Thus, they concluded that maintaining temperature at approximately 60C was optimal for process control. Snell (1957) suggested that the optimum temperature for composting systems was 45C, while Jeris and Regan (1973a, b) found a peak near 60C in their bench scale studies. Atchley and Clark (1979), suggested 70C, and Haug (1993) suggested 72C as optimal temperature levels. Variations in material, composting method, aeration rate and other factors could affect optimal composting temperature (Hall, 1998). In addition, several researchers (Haug, 1993; Ekinci, 2001) used empirical data from various substrates to estimate temperature dependence curves. It was clear that different materials have varied characteristics, and that the optimal composting conditions might vary significantly depending upon substrate composition.

19

Other Related Parameters In addition to oxygen, moisture and temperature, the other parameters chemical, physical and biological parameters may also be important (Golueke and Diaz, 1987; Miller, 1992; Bosma et al., 1993; Deng and Cliver, 1995). Jeris and Regan (1973c) found that pH, C:N, and specific substrates are to be relevant parameters. Starting with an appropriate C:N ratio (15 to 40:1) can significantly improve the composting efficiency. Similarly, pH far from neutral (outside the pH range of 5.5 to 8.5) can significantly limit degradation rate (Rynk et al., 1991; Hall, 1998). Various material properties could also have significant effects. Material properties which were studied include density, bulking agent, recycling ratio and particle size (Mitscherlich and Marth, 1984; Kashmanian, 1995; Zhang, 2000). Finger et al. (1976) found that density variations over the range (320-448 kg/m-3) did not significantly affect aerobic microbial growth in semisolid piles. However, very deep piles might cause compaction which could reduce air flow and oxygen availability, and also require larger fans to supply the needed air. Strombaugh and Nokes (1996) considered the reduction of flow through increased height and density, and calculated the optimal depth for an aerated static pile at 2 to 3 meters. Furthermore, nitrogen losses during composting were also demonstrated to be relevant (Morisaki, 1989; Sanchez-Monedero et al., 1996; Biasiak, 1997; Ryckeboer and Mergaert, 2003), with loss levels varying between 40 to 80% depending upon details of the composting process (Bernal et al., 1994). However, these parameters are different from temperature, oxygen and moisture because they are not significantly affected by outer factors. For example, for a given composting system (with a given substrate), if the initial C:N ratio is suitable through
20

pre-treating processes, the final C:N ratio will be at an optimal level if the entire composting process is done under optimal conditions. In comparison, the temperature, oxygen and moisture are affected significantly by outer factors such as weather and aeration. Thus, even if the initial conditions are optimal, their values will be varying during the composting process if there are no additional control measures. Therefore, temperature, moisture and oxygen are the most important parameters that should be controlled in the entire composting process. Understanding the relationships between these factors and the composting substrate are main issues in studying the composting process control (Zheng, 2001). 2.3.3. Experimental Studies of Composting Process Control Many experimental studies were undertaken in controlling composting processes (Jeris and Regan, 1973a, b; MacGregor et al., 1981; Stentiford, 1996; Ekinci, 2001). However, few studies on real-time process control for composting systems were conducted (Das and Keener, 1996). Ekinci (2001) gave a review of work to date in this area. Moreover, a fair amount of empirical control work were undertaken, however, little theoretical or experimental effort was made to assess controllability of composting processes from a system point of view. Complexities in composting systems present an opportunity for developing advanced process control technologies to enhance the process efficiency (Dochain and Bastin, 1984; Ogata, 1990; Franklin et al., 1991; Nakaya, 1999). There are three principal factors that need to be controlled during composting processes (Xi, 2002): aeration rate, temperature, and moisture content. Many researchers studied composting process energetics and transport (Finger et al., 1976; Kishimoto et al., 1987; Saucedo-Castaneda et al., 1990; Stombaugh and Nokes, 1996; VanderGheynst et al.,
21

1997; Nakaya 1999; Ekinci, 2001; and Nakasaki et al., 2004, 2005), providing bases for improved effectiveness. Heat and mass transfer may limit the overall growth in a large compost pile. The important mechanisms of heat removal are vaporization and dry-air convection, both of which are dependent on ventilation (MacGregor et al., 1981; Ekinci, 2001). Temperature was frequently used as a yardstick in composting to judge the efficiency and degree of stabilization of the treatment process. The generation of heat by microorganisms results in elevated temperature of the compost which may exceed 70C (Jeris and Regan, 1973a). Maximum stabilization rates have been reported under thermophilic conditions (Nakasaki et al., 2004, 2005). Aeration in composting performs three main functions. Firstly, it provides the oxygen necessary for microorganisms to aerobically decompose the substrate. Secondly, it removes water vapor and other gaseous products. Thirdly, it removes heat from the composting materials in order to maintain desired temperatures. Due to these important functions, many researchers paid attention to it (Ebeling, 1995; Iriarte and Ciria, 2001). As discussed before, water is a critical factor in composting processes (Miller, 1989). Sufficient moisture in the solid waste is required for maximum efficiency of microbial stabilization (Jeris and Regan, 1973b). Water is both required for and produced by microbial activities. A major part of the composting process is the loss of water, as a result of evaporation as the process progresses (Stentiford, 1996). Because of this, moisture is always a key control factor in composting process control (Oppenheimer, 1997; Nakaya, 1999; Zhang, 2000; Ekinci, 2001). Reactor choice is also important in experimental studies. Vertical packed bed

22

reactors with forced aeration have been used by a variety of researchers for exploring composting processes (Ashbolt and Line, 1982; Bach et al., 1985; Nakasaki et al., 1985; Bach et al., 1987; Hogan et al., 1990; Kubota and Nakasaki, 1991; Keener et al., 1992; Lynch and Cherry, 1995; VanderGheynst et al., 1997; Nakaya, 1999; Zhang, 2000; Hamelers, 2001). Forced aeration systems are intended to supply air to the composting mass using pressurized air systems. Laboratory reactors from 250 ml to 770 litres have been used in experimental work (Van Durme et al., 1992; Doeblin, 1993; Richard, 1997; VanderGheynst, 1997; Xi, 2002). Common reactor sizes for bench scale experimental work range from 4 to 50 litres (Ashbolt and Line, 1982; Kumar, 1993; Biasiak, 1997; Hamelers, 2001). 2.3.4. Composting Process Control for Field-Scale Experiments In addition to laboratory and pilot-scale work, a number of scientific field studies of composting control have been made. Hall (1998) gave a review about these studies. Jeris and Regan (1973a, b, c) published their seminal articles more than three decades ago, being used by a number of researchers with their Rutgers Strategy to control temperature and moisture under given different composting substrates. A number of studies for field-scale systems suggested that, with some form of distribution such as a plenum chamber or multiple-hole spacing, aeration could be maintained effectively (MacGregor et al., 1981; Haug, 1993, 1996; VanderGheynst et al., 1997; Hall, 1998; Zhang, 2000; Xi, 2002). The Rutgers Strategy was a simple temperature feedback mechanism which used forced aeration to cool the pile, maintaining relatively constant temperatures in an aerated, usually unmixed compost pile. In general, the Rutgers Strategy involved only on-off control based on a given set-point for
23

temperature, or simplistic additions of water in the case of moisture control. Haug (1993, 1996), Bernal et al. (1994), Paredes et al. (1996), and Hall (1998) all referred to these simple control strategies in their work. Excellent work was done by Nakasaki et al. (1985) where temperature feedback control using municipal sludge was studied. Their conclusions suggested that thermodynamic behavior was a direct consequence of the heat output-temperature interaction and called for further study of this interaction as the focal point of the analysis. MacGregor et al. (1981) devised a practical means of controlling temperature in field-scale composting and evaluated the effects of such an approach on process control. Michel et al. (1996) studied various yard trimmings as substrates and looked at pile size and turning regime in naturally aerated piles. They recognized that additional turning (and thus aeration) would improve process efficiency. Roig and Bernal (1996) adopted the Rutgers system to compost various agricultural wastes, and used the produced compost as agricultural soil amendments. They concluded that controlling temperature to a desired setpoint provided a superior compost product in terms of soil amendment characteristics. Natural aeration proved insufficient for effective composting, but with the addition of forced aeration, composting effectiveness improved. In field-scale experiments, temperatures in excess of 70C (and in some cases exceeding 80C) have been observed and held for several days (Nakasaki et al., 2004). Other than control, the major concerns in the design of a composting facility were the relationships among bed depth, airflow and fan size (Bari, 2000a, b). Substrate composition, sifting, and materials handling were also significant concerns. The potential use of compost in agriculture was directly dependent upon its quality, which was a

24

function of the feedstock (substrate) material, composting process and compost maturity. It was important to recognize that without a good feedstock, process control limited application. However, process control in field situations could address both the composting process and compost maturity. Field work was limited in terms of scientific study. Practical concerns might overshadow interest in collecting additional scientific data (Kuter et al., 1985). Furthermore, the ability to control parameters could be limited in field situations. Practical concerns included odors, pathogen reduction, cost and safety of working conditions. These interests indicated that temperature was a major control variable in field situations, both for purposes of optimizing reaction rates, and also for destroying pathogens (Hall, 1998). In general, almost all pathogenic microorganisms were killed at 65C (Kishimoto et al., 1987; Rynk, 1991; Hall, 1998; Ekinci, 2001; and Nakasaki et al., 2004, 2005). NYSDEC regulations suggested 55C for 3 to 7 days, depending upon the technology used (Krause, 2002). VanderGheynst et al. (1997) discussed a field-scale agricultural composting application where the desire to reuse compost as dairy bedding necessitated concern for pathogen destruction. Lack of controllability of temperature limited this objective. Modeling may be applicable to field scale situations (Bari, 2000a, b). Haug (1996) discussed three field scale facilities in operation in the United States. In these cases, modeling aided in process design and control. When using laboratory-scale experimental data to design a full-size composting system, it was important to address scaling factors. Although total aeration required for respiration per unit substrate mass might be similar, greater aeration might be required to control temperature, because normalized conductive

25

losses tended to decrease as reactor size increases. 2.4. Process Control Studies in Other Areas 2.4.1. Model Predictive Control Model predictive control (MPC) is an optimization-based strategy that uses a process model to predict the effect of potential control actions on the evolving state of the process (Huang, 2004). In the late 1970s, MPC was initiated in the chemical process industries (Garcia et al., 1989; Ricker, 1990; Eaton and Rawlings, 1992). In MPC, the control objective was optimized on-line subject to the constraints (Antwerp and Braatz, 2000). The previous studies contributed to the development of MPC for the real industrial processes (Egardt, 1979; Ricker, 1990; Muske and Rawlings, 1993; Tan and Keyser, 1994; Braatz and VanAntwerp, 1997; Rao et al., 1997; Rao and Rawlings, 2000; Diehl et al., 2002; Das and Potra, 2003; Leskens et al. 2005; Mjalli, 2005). Balasubramhanya and Doyle (2000) developed low order nonlinear models to capture the essential nonlinear dynamic transport behaviors. Tight control of the multi-component reactive distillation column was obtained with the use of the developed reduced order nonlinear models in a model predictive control algorithm. The inherent trade off between modeling accuracy and computational tractability was addressed. Zhu et al. (2000) made a first-step contribution in developing a comprehensive methodology for plant-wide control via integration of linear and Nonlinear MPC. A simple controller coordination strategy that counteracted interaction effects was proposed for the cases of a linear subsystem and a nonlinear one. The hybrid method was applicable to plants that could be decomposed into approximately linear and highly nonlinear subsystems that interacted via mass and energy flows.
26

Lee et al. (2001) developed a model-based predictive control (MPC) method with the aim of applying it to periodic process control problems through introducing concepts from repetitive control. The method used a linear time-varying system description of a periodic process and could handle constraints on inputs and outputs. McAvoy (2002) discussed a model predictive control approach to statistic process control in chemical plants. The methodology promised a means of reducing product quality variability in plants where online quality measurements were not feasible. Rather than holding setpoints in a plant control system constant, a small number was varied in response to the effect of process disturbances on selected measurements. Cano and Odloak (2003) studied the control of integrated systems in the presence of uncertainties. To deal with unknown steady states, the controller incorporated a state-space model in the incremental form, which was a framework frequently adopted by MPC packages. For integrating systems, minimizing the integrating states at steady state was not sufficient to guarantee the stability of the uncertain plant. A modified cost function was proposed to allow the controller to stabilize a family of plants. Sarimveis and Bafas (2003) proposed a fuzzy model predictive control (FMPC) methodology, which was based on a dynamic non-linear fuzzy model for the plant and belonged to the family of indirect fuzzy control methodologies (Wang, 1994). The method was based on a dynamic fuzzy model of the process to be controlled, which was used for predicting the future behaviors of the output variables. A nonlinear optimization problem was then formulated, which minimized (a) the difference between the modeled predictions and the desired trajectory over the prediction horizon and (b) the control energy over a shorter control horizon. The method was illustrated via the application to a

27

nonlinear single-input single-output reactor. Leskens et al. (2005) introduced a model predictive control system for improving the process operation of municipal solid waste (MSW) combustion plants. The developed system aimed at tackling a typical MSW combustion control problem. Using the proposed control system, an assessment of the improvement in performance was made through comparison between the developed MPC-based and the conventional MSW combustion control systems. The result showed that an MPC-based combustion-control system was capable of improving control/operation performance compared with conventional combustion-control system. 2.4.2. Intelligent Process Control An alternative approach to classical model-based control is to design controllers whose structure and consequent outputs in response to external commands are determined by experiential evidence (i.e. the observed input/output behaviors of the process, rather than by reference to a mathematical or model-based description of the controller (Abu-Rub et al., 2004). The controller is called intelligent controller. Intelligent or self-organizing control is applicable to processes with characteristics being ill defined, complex, nonlinear, time varying and stochastic. Intelligent controllers are inherently nonlinear (Murray et al., 1992). Intelligent control was originally proposed by Fu (1971) and was defined as an approach to generate control actions by employing aspects of artificial intelligence. Harris (1994) illustrated the inter-relationships between the various techniques that were utilized in intelligent control and the functionality and infrastructure that they attempted to incorporate.
28

The purpose of intelligent control is to incorporate the positive intelligent, flexible and creative attributes of human controllers, whilst avoiding the characteristics of inconsistency, unreliability, temporal instability and fatigue, associated with the human condition. One of the key features of intelligent control is the ability to change the behaviour to adapt to new circumstances (Moran and Zafiriou, 1988; Astrom and Wittenmark, 1989). There are three basic approaches for intelligent control: (i) knowledge based systems (KBS) including expert systems (Weiss and Kulikowski, 1991; Hunt, 1992; Koyama et al., 2003), (ii) fuzzy logic (Oishi et al., 1991; Harris et al., 1993), and (iii) artificial neural networks (Miller et al., 1990). These categories are not distinct as there exist strong interrelationships between fuzzy and expert systems, as well as between fuzzy and some associative memory neural networks (Lin and Lee, 1991). Jamishidi (2002) presented a general introduction related to the tools for intelligent control including fuzzy controllers and neural networks controller. Two different approaches including (i) artificial neural networks (ANNs) with associated neuro-controllers and (ii) fuzzy-logic controllers, dominate the field of intelligent control. (1) Fuzzy Logic Control Fuzzy logic was suggested by Zadeh (1968) as a method for mimicking the ability of the human reasoning process to use a relatively small number of rules and still produce a smooth output through interpolating. A fuzzy system is one that has at least one system component that uses fuzzy logic for its internal knowledge representation. Fuzzy logic is used internally for knowledge representation and, externally, can be considered as a system component that describes various uncertain relations. Fuzzy logic forms a bridge
29

between two areas of qualitative and quantitative modeling. The models formed using fuzzy logic have been applied to many types of information processing, including static control systems, adaptive control systems, process modeling, signal estimation, image processing, temporal planning and decision making (Sugeno, 1985; Keller et al., 1992a, b; Huang and Fan, 1993; Harris, 1994; Ghomshei and Meech, 2000). Moore and Harris (1994) showed a typical static controller. The measured inputs were fuzzified and the composition with rule bases was analyzed. The inferred fuzzy set output was then defuzzified. This type of controller, encapsulating human operator knowledge, was an example of fuzzy logic used in qualitative systems. The adaptive fuzzy controller can produce a complete rule base to replace inadequate or faulty rules. The self-organizing fuzzy controller is one of the most widely used techniques. A self-organizing fuzzy logic controller, as with any learning system, requires a performance measure/index or objective function against which the achieved response can be evaluated (Procyk and Mamdani, 1979; Sutton and Jess, 1991; Tanaka and Sugeno 1992; Ferrer et al., 1998). This can then be used to generate a feedback signal to alter the control policy. A self-organizing controller is therefore able to improve its capability based purely on experiential means, with little or no priori knowledge of the process being necessary. The controller was illustrated in Moore and Harris (1994). The system consisted of two levels: control and adaptation. The control level was identical to that used in the static fuzzy controller, with the rule-base supplied from the adaptation level. The adaptation (or rule modification) process involved identifying the rule that caused the poor performance and applying the modification required by adaptation signals.

30

A review conducted by Maiers and Sherif (1985) covered over 450 papers which addressed fuzzy-logic applications to automatic control and decision making in industrial and environmental systems. Procyk and Mamdani (1979) developed a self-organizing fuzzy controller (SOFC) which had major advantages over a static rule-based controller, particularly in the small amount of priori process knowledge required and the ability to adapt to changes in process parameters. Sutton and Jess (1991) subsequently improved the SOFC algorithm by Procyk and Mamdani (1979). Umbers and King (1980) developed a fuzzy logic controller (FLC) that reduced fuel consumption, since it initiated control actions earlier and could respond to smaller variations in measured signals relative to the corresponding human operator. Tsai et al. (1993) utilized the fuzzy control theory to forecast and control effluent suspended-solid concentration and to further predict the mixed liquor suspended solids (MLSS) concentration in a wastewater treatment system. Boscolo et al. (1993) applied the fuzzy approximate reasoning to the management and control of a pilot-scale anaerobic digester. Raju et al. (1994) developed a hierarchical fuzzy control algorithm and applied it to the control of feed-water flow in a steam generator of a power plant. Muller and Libelli (1997) described a real-time process control scheme to cope with the problem of input disturbances in wastewater treatment processes, based on a fuzzy inferential control system. Ferrer et al. (1998) developed a fuzzy-logic-based control system and tested it in the main aerobic reactor of a BARDENPHO process pilot plant. The system was compared with two ordinary aeration process controllers: one- and two-aeration-level on/off controllers. Peres et al. (1999) described in a general form a hierarchical structure

31

of fuzzy control and fuzzy model used in an end-milling process. These modules represented possible solutions to complex problems: optimization and supervision of the milling process. Li and Chang (2000) presented a hybrid fuzzy logic proportional plus conventional integral-derivative (Fuzzy PID) controller to improve the control performance of the conventional PID-type controller. The proposed Fuzzy PID controller was constructed by using an incremental fuzzy logic controller in place of the proportional term in the conventional PID controller. Lin and Lai (2002) introduced the TakagiSugeno (TS) fuzzy logic theory and the gain scheduling technique to design guidance and control (GC) parameters over coupled flight conditions. A genetic algorithm was used as the computing device to determine the consequent function parameters of the fuzzy rules. Kohn-Rich and Flashner (2002) designed fuzzy control laws for tracking control of a large class of mechanical systems. They employed the framework of Lyapunovs stability theory to formulate a class of control laws that guaranteed convergence of the tracking errors to within specification limits in presence of bounded parameter uncertainties and input disturbances. The proposed control laws possessed a large number of parameters and functional relationships to be chosen according to a methodology developed in the paper. The large number of design degrees of freedom made the approach suitable for fuzzy logic implementation. Demiroren and Yesil (2004) presented a method based on fuzzy logic controllers (FLCs) for automatic generation control (AGC) of power systems. The technique was applied to control systems with three units including two steam turbines and one hydro

32

turbine tied together through power lines. The AGC based on fuzzy PI-type controller was proposed. Controller parameters could be changed very quickly by the system dynamics because no parameter estimation was required in designing controller for nonlinear systems. Guclu (2005) developed a Fuzzy logic control system of seat vibrations based on a non-linear full vehicle model. In his study, the dynamic behaviors of (a) a non-linear eight degrees of freedom vehicle model having active suspensions and (b) a fuzzy logic (FL) controlled passenger seat were examined. Three cases of control strategies were taken into account. The time responses of the non-linear vehicle model due to road disturbance and the frequency responses were obtained for each control strategy. At the end, the performances of these strategies were compared. (2) Neural Networks Control Dayhoff (1990) provided details about artificial and biological neural networks. Since neural networks had the ability to learn the complex dynamic behavior of a physical system, they offered a cost-effective approach for developing useful process models. Cybenko (1989) and Hornik et al. (1989) proved that continuous functions could be approximated to an arbitrary degree of exactness on a compact set by a feed-forward neural network comprising two hidden layers. Different artificial neural networks (ANNs) exist for various purposes. The most common ANNs for engineering applications are the back-propagation networks (BPN) (Hagan et al., 1996). Syu and Chen (1998) applied back-propagation neural networks to a continuous wastewater treatment process. In this case, neural networks adaptive control was demonstrated for its ability in carrying out on-line operations successfully.
33

Neural networks were successfully used for a number of chemical engineering process controls, as shown in Narendra and Parthasarathy (1990), Ydstie (1990), Hernandez and Arkun (1990), Psichogios and Ungar (1991), Willis (1992), and Song and Park (1993). Hussain (1999) provided an extensive review of various applications utilizing neural networks for chemical process control in both simulation and online implementations. Integrations of neural networks into model-predictive-control, inverse-model-based and adaptive-control techniques were reported. Numerous related references could also be found in the studies of ANNs advantages and limitations (Widrow and Lehr, 1990; Haykin, 1994). There is no clear advantage of one network over the other as well as of one activation function over the other (Alexander and Michael, 2001; Ma, 2001). Many studies on neural network control were reported in Vigie et al. (1990), Psichogios and Ungar (1991), Montague et al. (1991), Donat (1991), Gnaebe and Goodwin (1992), Takahashi, (1993), Ortega and Camacho (1994), Ramchandran and Rhinehart (1995), Gokhale (1995), Megan and Cooper (1995), Ramchandran and Rhinehart (1995), Hsu et al. (1995), Emmanouilides and Petrou (1997), Yu (1998), Zhang and Stanley (1999), Tay and Zhang (1999, 2000), Sohn et al. (2000), Wang and Gao (2000), Kolehmainen et al. (2001), and Kulkarni (2004). Tement (1995) discussed an industrial application of multivariate nonlinear feed-forward/feedback model predictive control, where the model was given by a dynamic neural network. A multi-pass packed bed reactor temperature profile was modeled via recurrent neural networks using the back-propagation through time training algorithm. This model was then used in conjunction with an optimizer to build a

34

nonlinear model predictive controller. Tan (1996) presented three optimizing methods for the design of an external recurrent neural network based predictive controller to compensate for large-time-delays in nonlinear processes. These techniques included the gradient descent method, the Newton-Raphson algorithm, and the method of Levenberg Marquardt. An application of these algorithms to a simulated digester process was also presented. Syu and Chang (1997) proposed a recurrent back propagation neural network for the on-line adaptive pH control of penicillin analyze. Kasparian and Batur (1998) presented a model-reference-based neural network structure that could be used for adaptive control of linear and nonlinear processes. The proposed neural network controller was tested on several simulated non-linear systems. Also, a fast algorithm was introduced for training the proposed neural network controller. This algorithm was based on Davidons least squares minimization technique. Cannas et al. (2001) introduced a Universal Tabu search meta-heuristic to optimally design a recurrent neural network architecture. The problem of choosing the number of hidden neurons and the number of taps and delays in the network synapses was formalized as an optimization problem whose cost function to be minimized was the network error calculated on a validation data set. Guh (2003) developed a hybrid artificial intelligence technique for building a real-time statistical control system. An artificial neural network based control chart monitoring sub-system and an expert system based control chart alarm interpretation sub-system were integrated for automatically implementing the statistic process control tasks comprehensively. An example was provided to demonstrate that hybrid intelligence

35

can be usefully applied for solving the problems in a real time statistic process control system. Hoffmann (2005) proposed a numerical control of Kohonen neural network for scattered data approximation. Surface reconstruction from scattered data was presented by using Kohonen neural network. The network produced a topologically predefined grid from the unordered data which could be applied as a rough approximation of the input set or as a base surface for further process. The quality and computing time of the approximation could be controlled by numerical parameters. In the application, ruled surface was produced from a set of unordered lines by the network. More recently, a new process control technology neural network predictive control has been developed. The traditional MPC system enhanced by feed-forward neural networks is defined as neural predictive control (NPC), which has been effectively applied in process control associated with non-linear optimization problems (Zamarreno and Vega, 1999). Donald and Pamaela (1997) presented an efficient implementation of the generalized predictive control using a multi-layer feed forward neural network. In using the Newton-Raphson as the optimization algorithm, the number of iterations needed for convergence was significantly reduced. The simulation results showed convergence to a feasible solution within two iterations, and the timing data showed that the real-time control was possible. Gu and Hu (2002) presented a new path-tracking scheme for a car-like mobile robot based on neural predictive control. A multi-layer back-propagation neural network was employed to model non-linear kinematics of the robot instead of a linear regression

36

estimator in order to adapt the robot to a large operation range. The neural predictive control for path tracking was a model-based predictive control desired path. The results showed that the NPC was one of the solutions for time-variable nonlinear systems. Mjalli (2005) developed a neural network model-based predictive control for liquidliquid extraction contactors. In his study, neural network-based control algorithms were applied to control the product compositions of a Scheibel agitated extractor of type I. Model predictive control algorithm was implemented to control the extractor. The extractor hydrodynamics and mass transfer behavior were modeled using the non-equilibrium backflow mixing cell model. The results showed that the model predictive control was capable of solving the servo control problem with minimum controller moves. (3) Fuzzy Neural Networks Control The fuzzy-logic control techniques suffer from the following problems (Horikawa et al., 1992): (1) the derivation of fuzzy control rules is often time consuming and difficult, (2) the system performance relies significantly on so-called process experts who may not be able to transcribe their knowledge into the requisite rule form, (3) there exists no formal framework for the choice of the parameters of a fuzzy control system, and (4) the static fuzzy controller has no mechanisms for adapting to real-time plant change. These difficulties may affect applicability of the fuzzy-logic control under stringent conditions (Stephanopoulos and Han, 1996; Huang and Wang, 1999). To overcome the above-mentioned drawbacks, learning abilities of the neural networks were introduced to automate and realize the designs of fuzzy logic control systems (Jang, 1992; Yager, 1992; Lin, 1994). The combination of techniques from these
37

two fields took the benefits of both neural networks and fuzzy logical systems. The neural networks provided the connectionist structure (fault tolerance and distributed representation properties) and learning ability to the fuzzy logical systems; the fuzzy logical systems provided a structural framework with high-level fuzzy IF-THEN rule thinking and reasoning to the neural networks (Chen and Peng, 1999). Undoubtedly, the hybrid neurofuzzy of intelligent control offered substantial advantages over single neural/fuzzy-based algorithms (Kumar and Garg, 2005). Enbutsu (1993) proposed an automatic fuzzy rule extraction method using an ANN. The simulation results of the proposed method using full-scale plant data demonstrated that a fuzzy system whose rulebase was modified automatically with extracted rules had better performance than a conventional fuzzy system whose rulebase included only operators heuristics. Khalid (1993) developed an adaptive fuzzy-neural control scheme by integrating two neural models within a basic fuzzy logic controller. The performance of the adaptive fuzzy-neural controller was compared to the basic fuzzy logic controller and a conventional digital-PI controller under identical conditions of varying complexities in the process. The results showed that the adaptive fuzzy-neural control scheme was superior in performance to the other two controllers. Huang and Wang (1999) introduced a graphical model which was the extended fuzzy causal network and applied it to a case study of waste water treatment plants. The structure of the network was developed using parameter sensitivity studies and the relationships between connected parameters were obtained using a learning approach adapted from fuzzy neural networks.

38

Das et al. (2001) developed hybrid fuzzy logic committee neural networks and applied it for the recognition of swallow acceleration signals from artifacts to improve the reliability of the recognition or automated diagnostic systems. Two sets of fuzzy logic-committee networks (FCN) consisting of seven member networks were developed, trained and evaluated. The hybrid intelligent system consisting of fuzzy logic and committee networks provided a reliable tool for recognition and classification of acceleration signals due to swallowing. Zhou et al. (2002) presented a fuzzy neural network (FNN) for manufacturing process control. The developed FNN utilized the input and output layer to on-line fine-tune scaling factors, and the hidden layers to realize the fuzzification, fuzzy inference defuzzification and tune parameters such as membership functions, fuzzy control rules dynamically. A new combining learning algorithm which combined the gradient-based error back-propagation algorithm with similar Newton algorithm was proposed to improve the convergence speed and release computational burden during the learning process. An adaptive fuzzy-neural control (AFNC) scheme for multi-input multi-output uncertain robotic systems was proposed by Yu (2004) for suppressing the effects caused by multiple time-delayed state uncertainties, unmodeled dynamics, and disturbances. Each delayed uncertainty was assumed to be bounded by an unknown gain. A reference model with the desired amplitude and phase properties was given to construct an error model. A fuzzy-neural (FN) system was used to approximate an unknown controlled system from the strategic manipulation of the model following tracking errors. The proposed AFNC scheme used two on-line estimations, which allowed for the inclusion of

39

identifying the gains of the delayed state uncertainties and training the weights of the FN system simultaneously. 2.5. Summary In the past decades, many research efforts were made in composting process modeling, technology development, and process control. However, more advanced studies in these areas are desired for improving system efficiency and reliability: (1) A composting system is extremely complex and dynamic, involving a variety of biological, physical and chemical factors. Although a number of attempts were made in modeling composting processes, most of the developed models were empirical and could not effectively reflect interactions among various processes. Even though bio-kinetics was taken into account in a few models, they were significantly simplified to address the complex composting process. Thus, development of a modelling system that incorporates the principles of microbiology, mass transfer, and biotechnology is desired. Such a model could simulate the entire composting process and thus provide sound bases for the operation and control of composting practices. (2) The literatures review indicated that many impact factors exited with varied effects on the composting process. However, in most of the previous studies, only effects from individual factor were highlighted, while interactions among these factors were seldom addressed. In the modeling studies, analyses of the impact factors were mainly based on sensitivity analysis in terms of the effects from variations of one factor (or a limited number of factors); consequently, the joint effects from interactions among multiple factors were not considered. Thus, development of methodologies that can deal with such interactive effects is desirable.
40

(3) A composting system is extremely complex and dynamic, changing over time as microbial populations grow, evolve, die, and/or are replaced by new populations; moreover, uncertainties exist in a variety of the system components, forming an additional category of complexities. Effective composting process control and operation need to be based on characterization for the effects of such complexities on process behaviors. Literature review showed that a number of efforts in both modeling and experimental work were made in studying the complexities of composting processes. However, few studies were conducted with regards to uncertainties in composting processes. Thus, it is desired that the uncertainties be addressed in more advanced composting-process studies. (4) Process control is crucial for the management of composting systems. Effective process control can help improve compost quality and enhance process efficiency. A number of efforts were made to determine dynamic process-control policies for composting systems. However, previous studies of composting process control mainly focused on systems dynamics and empirical control; there has been no research focusing on real-time control for composting processes. In fact, a composting system is complicated with not only dynamic, but also cumulative effects from a variety of impact factors and their interactions. These complexities result in barriers for applying traditional control methodologies to composting processes. As a result, real-time process control technologies that can deal with the cumulative effects as well as the discrete and nonlinear interrelationships are desired. Consequently, studies of more effective methodologies for simulation modeling and process control of composting systems will be of contribution to the literature of

41

environmental systems engineering. In this dissertation research, issues of modeling and control for composting processes as well as advanced methodologies for systems analysis under complexities and uncertainties will be tackled. The developed methodologies will then be justified through a series of composting experiments.

42

CHAPTER 3 A MULTI-COMPONENT MODELING SYSTEM FOR SIMULATING AEROBIC COMPOSTING PROCESS


3.1. Statement of Problems Composting is a safe and effective way of treating biodegradable fraction of waste streams (Das and Keener, 1996). It can accelerate biodegradation of organic wastes and produce stable products that can be used for improving soil properties (Haug, 1993). Improved process control in composting can enhance process efficiency, pathogen destruction, and temperature maintenance (Hall, 1998; Park et al., 2004). However, effective process control should be based on the insight into the complicated multiphase multi-component composting process, as well as the capability to simulate the relevant process dynamics. Such simulation efforts are essential for supporting the prediction of system performance under different operating conditions, and thus quantifying interactive dose-response relationships between control actions and process efficiencies. Consequently, real-time management of composting process control with desired settings could be realized (Stentiford, 1996). Previously, a number of studies were conducted to simulate composting processes (Kishimoto et al., 1987; Stombaugh and Nokes, 1996; Hall, 1998; Alpert, 2002; Ryckeboer and Mergaert, 2003; Hamelers, 2004). Several parameters, processes, and reaction conditions were considered, including temperature, moisture, oxygen content, carbon dioxide content, energy flow, and substrate properties (Haug, 1993, 1996; Hall, 1998; Ekinci, 2001; Hamelelers, 2001; Trefry and Franzmann, 2003). A number of attempts were undertaken in modeling composting processes in the past decades, with
43

most of the developed models being empirical. Only a few models were based on microbial kinetics. In addition, these models were mainly related to external control factors such as temperature, moisture, and oxygen content, while the substrate constitution and the related impact factors were seldom considered. Even though bio-kinetics was taken into account in a few models, they were significantly simplified to address the complex composting process. The objective of this section, therefore, is to develop a multi-component modeling system for simulating the aerobic composting process. The model is based on the principles of microbiology, mass transfer, and biotechnology in relation to the composting process. Levels of soluble substrate, insoluble substrate, active biomass, inert material, moisture, temperature, and oxygen concentration are considered as state variables. The relationships among these variables are also incorporated within the model. The model can thus be used for predicting the oxygen uptake rate (OUR) of the composting substrate. Moreover, it can simulate processes with various substrates based on different settings of the related parameters. 3.2. Model Development A composting model is a tool used to predict behaviours within a composting process (Hall, 1998). In the compositing process, OUR is the amount of oxygen that is taken up by a unit sample of waste in a unit period of time (Richard, 1999). It is the most important composting process-rate indicator, because it is directly linked to composting reaction and linearly related to heat production. OUR is also independent of molecular substrate composition, and is thus a direct measure of composting stability (Luong and Volesky, 1983). OUR depends strongly on the state of waste (e.g. it is influenced by
44

temperature and water content of the waste). To predict OUR, a kinetic model is required. In this study, a multi-component composting model is proposed based upon the principles of microbiology, mass transfer, and biotechnology. Generally, composting substrate is considered to be made up of particles, and the composting reaction occurs within these particles. A waste particle is made up of four components: water, soluble substrate, insoluble substrate, and inert matter. These components are not pure; however, they have distinct physical, chemical, and biological properties. Except for water, these components can be considered as solids that exist in tiny lumps. These solids interact with the water. The particles together will then form a porous matrix, of which the pores are filled with water. The water contains soluble substrate, part of which is produced by microbial hydrolysis from insoluble substrate. The soluble substrate is oxidized by aerobic micro-organisms, transforming the soluble organic matter into carbon dioxide and new micro-organisms while producing heat. This reaction is the basis of the entire composting process (Hamelers, 2001). However, a composting model based on the assumption of one waste particle is not realistic because composting substrate is a kind of heterogeneous matter and the particles within the substrate are usually too small to have the same characteristics. Based on this consideration, one unit of composting substrate is considered as a research object. It is assumed that this unit is large enough such that the substrate within the unit can be divided into many sub-units with the same characteristics as the entire composting substrate; thus, the composting reactions in all of the units will have the identical result. It is also assumed that the volume of composting unit is unchanged during the composting process, such that the obtained concentration levels will be based on the same initial

45

volume. 3.2.1. State Variables State variables describe the properties of composting substrates under different conditions. In this model, the considered variables include aerobic biomass concentration ( X ), soluble substrate concentration ( S s ), insoluble substrate concentration ( S i ), inert material concentration ( I ), oxygen concentration ( O2 ), water content ( W ), and temperature ( T ). 3.2.2. Conversion Reactions Conversion reactions describe the transform relations between different constitutes of composting substrate. Three conversion reactions considered within the simulation are: growth of aerobic biomass, decay of aerobic biomass, and solubilisation of insoluble substrate. The aerobic biomass growth is modeled with a Monod-type growth model with an explicit dependence on the levels of oxygen and soluble substrate (Stombaugh and Nokes, 1996; Hamelers, 2001):

RG = m

K s w + S s K O2 w + O2

Ss

O2

X
(3.1) is biomass concentration

where RG is growth rate of biomass [molm-3s-1]; X

[molm-3]; O2 is oxygen concentration [molm-3]; S s is soluble substrate concentration [molm-3]; m is maximum conditional growth rate constant [s-1]; w is free water fraction related to moisture [-]; K O2 is half saturation constant of oxygen [molm-3]; K S is half saturation constant of soluble substrate [molm-3].
46

Temperature is an important factor that varies during the entire composting process. It is associated with Equation (3.1) as follows:

RG = m

K s w + S s K O2 w + O2

Ss

O2

X f1 (T )
(3.2)

where f 1 (T ) is a temperature function used to express the effect of temperature on the growth rate of biomass; RG is growth rate of biomass at T C [molm-3s-1]. Biomass decay rate is given by:
R d = b X f 2 (T )

(3.3)

where Rd is biomass decay rate at T C [molm-3s-1]; b is biomass decay rate constant [s-1]; f 2 (T ) is temperature function used to express the effect of temperature on

biomass decay. This means that the net production rate of biomass can be written as follows:
R X = RG Rd

(3.4)

where RX is net biomass production rate at T C [molm-3s-1]. The consumption rate of oxygen is linearly related to the production rate of aerobic biomass, and is expressed as follows (Hamelers, 2001):

RO2 =

RG YO2

(3.5)

where RO2 is the oxygen uptake rate [molm-3s-1], and YO2 is biomass yield on oxygen [O2-1]. The hydrolysis rate of insoluble substrate is described by:
Rh = K h S i f 3 (T )

(3.6)

where Rh is hydrolysis rate [molm-3s-1]; K h is hydrolysis rate constant [s-1]; S i is

47

insoluble substrate concentration [molm-3]; f 3 (T ) is temperature function used to express the effect of temperature on the hydrolysis rate. The net production of soluble substrate is influenced by both the hydrolysis rate and the soluble substrate consumption rate. Thus, the net production of soluble substrate can be written as follows (Stombaugh and Nokes, 1996):

RS s = Rh

RG YSs

(3.7)

where RSs is net production rate of soluble substrate [molm-3s-1]; YSs is biomass yield on soluble substrate [molm-3Ss-1]. The net production rate of insoluble substrate is given by:

RSi = Rh +

Rd YSi

(3.8)

where RSi is net production rate of insoluble substrate [mols-1]; YSi is biomass yield on insoluble substrate [molm-3Si-1]. The second term describes the conversion of dead biomass into insoluble substrate. During the oxidation of soluble substrate, water is formed, yet it is also needed for the construction of new cells (i.e. the largest part of a cell is water). Generally, microbial culturing is performed in aqueous culture. Though the amount of water needed for new cells is lower than the amount present, and the amount of water needed for hydrolysis is negligible, water still should be considered in the model. Therefore, the net water production rate may be developed as follows:
Rw = R x + RG Yw

(3.9)
48

where Rw is net production rate of water at TC [mol s-1]; is biomass water content [molmol-1]; Yw is biomass yield expressed in water [molmol-1].

3.2.3. Modeling Formulation Based on the above consideration of conversion reactions, it can be assumed that the convective mass transport between composting units could be counteracted. The balance equations for all state variables are proposed as follows. Oxygen

OOUA Ss O2 1 = RO2 = m X f1 (T ) t S s + K s w K O2 w + O2 YO2

(3.10)

t = 0, t > 0, t > 0,

m = m0 , m = m0 ,

O2= 0 ;
O2 = O2,i ;
OOUA = 0. t

m = mf ,

where O2 ,i is initial oxygen concentration [molm-3]; O2 is oxygen concentration [molm-3]; OOUA is oxygen uptake accumulation (OUA) during time period t [molm-3];
m0 is initial substrate mass of unit volume [Kg.m-3]; m f is final substrate mass of unit

volume [Kg.m-3]. Soluble substrate

S s Ss O2 1 = RSS = m X f1 (T ) + k h Si f 3 (T ) t S s + K s w K O2 w + O2 YSS
(3.11)

49

t =0 t >0 t >0

m = m0 m = m0

S s = S s ,0 ; S s = S s ,0 ;
S s =0. t

m = mf

Where S s , 0 is initial soluble substrate concentration [molm-3].

Biomass

Ss O2 X = RX = m X f1 (T ) b X f 2 (T ) t S s + K s w K O2 w + O2

(3.12)

t =0 t >0 t >0

m = m0
m = m0

X = X0;
X = X0;

m = mf

X =0. t

where X 0 is initial biomass concentration [molm-3].

Insoluble substrate

Si bX = RSi = k h Si f 3 (T ) + f 2 (T ) t YSi

(3.13)

t =0 t >0

m = m0
m = m0

S i = S i .0 ;
S i = S i .0 ;

t > 0 m = mf

S i =0. t

where S i , 0 is initial insoluble substrate concentration [molm-3].

50

Inert material As inert material will not be converted and the substrate volume is assumed to be fixed, the average concentration of inert material will not change over time. Water Water content can be derived from the water balance over the entire unit substrate and runs as follows:
R W = RW rH 2O = Rx + G rH 2O t Yw

(3.14)

t = 0 m = m0 t >0 t >0
m = m0

W = W0 ;
W = W0 ;

m = mf

W = 0. t

where rH 2O represents rate of water loss from substrate [mols-1], which is related to air flow rate, substrate mass, and substrate temperature. Volume The actual volume of the substrate decreases rapidly over time. To reflect this change, it is assumed that the change of substrate volume is directly related to that of substrate mass. This relationship can be expressed as follows:

dm V ' = Dv dt t

(3.15)

t =0

m = m0

V ' = V0 ;
51

t >0 t >0

m = m0

V ' = V0 ;
V ' =0. t

m = mf

where V ' is actual composting unit volume at time t [m3]; D v is a function mainly related to substrate category [kgm-3]. For a given substrate, D v can be determined based on a series of composting experiments (Xi, 2002). Substrate mass equations The substrate mass can be obtained though the following relationship:

m = (O2 PO2 + S s PSs + X PX + S i PSi + I PI + W PW ) V

(3.16)

where m is unit substrate mass at time t [kg]; V is initial volume of unit substrate mass [m3]; PO2 is molecular weight of oxygen [kgmol-1]; PS s is nominal molecular weight of soluble substrate [kgmol-1]; PX is nominal molecular weight of biomass [kgmol-1]; PSi is nominal molecular weight of insoluble substrate [kgmol-1]; PI is nominal molecular weight of inert material [kgmol-1]; PW is molecular weight of water [kgmol-1]. Equation (3.16) is based on an assumption that the substrate volume is unchanged during the composting process. However, the substrate volume varies significantly during the entire composting process. Therefore, a modification is needed for the concentrations of all substrate components. The modified equation is:
m = (O2 PO2 + S s PS s + X ' PX + S i PSi + I ' PI + W ' PW ) V '
' ' '

(3.17)

' ' where O2 , Ss , X ' , Si' , I ' , and W ' are corrected concentrations based on the actual substrate

52

volume. The proposed model contains six differential equations (Equations 3.10 to 3.15) and a number of general equations (Equations 3.1 to 3.9, 3.16, and 3.17). Several equations contain nonlinear kinetic expressions (Equations 3.1, 3.2, 3.4, 3.5, 3.7, and 3.9). The model could be solved numerically through discretization in time. Before solving the model, the nominal parameters and functions need to be determined as discussed in the following section.

3.3. Modeling Inputs and Outputs The input and output parameters are essential components of the model. The modeling inputs are those that characterize the composting materials and the modeling outputs are the OUR and part of the state variables. OUR is the rate at which a sample of waste consumes oxygen. It is defined based on the initial volume, since the measured values of the OUR levels are for the initial composting material. The state variables, i.e. X (t ) , S s (t ) , S i (t ) , and V ' , are functions of time and the unit composting substrate. They can be obtained from the numerical solutions after all the nominal parameters and the effect coefficients are determined. In detail, the model inputs mainly include temperature, moisture, oxygen concentration, and initial conditions of the state variables. They are referred to as waste characteristics, which are important for plant design and process control. To solve the model, values of the nominal parameters should also be determined. The reference values extracted from the literature are shown in Table 3.1 (Stombaugh and Nokes, 1996; Hamelers, 2001). In addition, the three temperature-effect functions should

53

be determined. For f 1 (T ) , it is assumed that the temperature effects on microbes follow

54

Table 3.1 Nominal parameters values Symbol Name


max

Value

Unit s-1 s-1 s-1 molm-3

Maximum biomass growth rate constant 1.0 10-4 Biomass decay constant Hydrolysis coefficient Substrate saturation constant 7.6 10-5 4.9 10-7 0.31

b
Kh
Ks

KO

Oxygen saturation constant Biomass yield on oxygen Biomass yield on soluble substrate Biomass yield on water Biomass yield on insoluble substrate Free water fraction

3.40 10-4 molm-3 1.12 0.53 1.34 1.02 0.83 molmol-1 molmol-1 molmol-1 molmol-1 -

YO2
Ys YW

YSi

55

the general rule: (a) the activity of biomass is minor at low temperature (below 50C); (b) the activity is high at moderate temperature (50 to 60C); (c) thermal kill happens at high temperature (over 70C), leading to rapid decrease of the activity. Based on this rule, it is assumed that, at 55C, composting microbes will reach the highest activity level ( f 1 (T ) = 1). Using quadratic curve fitting, the obtained function is:
f1 (T ) = 3.11 10 4 T 2 + 3.48 10 2 T + 0.0265 (5C < T < 75C)

(3.18)

For f 2 (T ) , it is assumed that the effect of temperature on the decay of biomass can be expressed as a quadratic function, and it has the minimum at 55C. Thus we have:
f 2 (T ) = 2.142 10 4 T 2 2.356 10 2 T + 1.348 (5C < T < 75C)

(3.19)

Similarly, it is assumed that the effect of temperature on the hydrolysis rate is (assume f 3 (T ) = 1 at 55C):
f 3 (T ) = 0.0182T

(5C < T < 75C)

(3.20)

These three functions can be used as a reference set. For a given substrate, the detailed functions will be determined based on the corresponding experiments. Here, the standard temperature is set at 55C because this temperature is used for composting reaction in most studies (Hall, 1998; Ekinci, 2001). 3.4. Model Solution Consider a forced aerobic composting system, wherein moisture, oxygen concentration, and temperature are controllable. The moisture was kept within a certain range that is suitable for composting reaction. The oxygen concentration was assumed constant for the entire system. However, due to variations in biomass concentrations, the oxygen concentrations for microbes are different. It is assumed that the relationship
56

between the concentrations of biomass and oxygen is linear. Thus, Equations (3.10) to (3.13) can be reformulated as:

OOUA Ss O2 1 = m f1 (T ) t S s + K s w K O2 w + O2 / X YO2 S s Ss O2 1 = m f1 (T ) + k h Si f 3 (T ) t S s + K s w K O2 w + O2 / X YSS Si bX = k h Si f 3 (T ) + f 2 (T ) t YSi Ss O2 X = m f1 (T ) b X f 2 (T ) t S s + K s w K O2 w + O2 / X

(3.21)

(3.22)

(3.23)

(3.24)

Solutions of Equations (3.21) to (3.24) provide the curves of OUA, S s , S i , and


X over time, and then the curve of OUR can be obtained. The methodology used to solve

these balance equations is based on the presence of the two distinct stages: substrate-saturated stage and substrate-limited one. In the first stage, sufficient soluble substrate is present and the OUR is determined by the active biomass, i.e. the soluble substrate is sufficient. It is assumed that the switch time is very short without any intergrade period between the two stages. The substrate-limited stage starts when there is insufficient substrate for biomass within the aerobic region. This situation begins when the soluble substrate concentration becomes very low (nearly zero). The solution for the saturated stage will be based on Equations (3.21) to (3.24). When the average soluble substrate concentration is under a steady state with a low value (assume S s = 0, and thus
S s = 0 ), the substrate-limited stage is considered. The t

biomass production rate will then be solely dependent on the soluble substrate production
57

rate, i.e. the hydrolysis rate. Under such circumstances, a new set of equations can be obtained:

Si bX = k h Si f 3 (T ) + f 2 (T ) t YSi
X = k h S i f 3 (T ) YSs b X f 2 (T ) t

(3.25)

(3.26) (3.27)

OOUA 1 = k h S i f 3 (T ) YS s f1 (T ) t YO2

Solutions of Equations (3.25) to (3.27) provide the curves of OUA, OUR, S s , S i , and
X over time within the substrate-limited stage.

The initial conditions for this modeling study are set as follows: the initial concentration of soluble substrate, insoluble substrate, active biomass and inert material are 2200, 3600, 2 and 400 molm-3, respectively. The system temperature is mainly affected by a number of external factors. It is assumed that temperature is controlled based on a composting temperature curve (Hall, 1998). At the first 24 hours, the average temperature is 35C; then, the temperature is maintained at approximately 55C. After the concentration of soluble substrate becomes very low, the composting reaction enters the limited stage when the temperature will be kept at approximately 45C. Therefore, the entire composting process is separated into three periods with temperature levels of 35, 55, and 45C, respectively. Based on Equations (3.18) to (3.20), we have: If T = 35C, f1 = 0.864, f 2 = 0.786, and f 3 = 0.637. If T = 45C, f1 = 0.963, f 2 = 0.722, and f 3 = 0.819.

58

If T = 55C, f1 = 1.000, f 2 = 0.700, and f 3 = 1.000. The Runge-kutta algorithm is employed to numerically solve the model, since it is difficult to get analytical solutions for Equations (3.21) to (3.27). The obtained results are shown in Figures 3.1 to 3.5. Figure 3.1 presents the dynamic variations of OUA over time. At the beginning, the OUA increased slowly due to low biomass concentration; with time going, the OUA increased rapidly. Approximately 100 hours later, the increasing rate of OUA dropped down because the OUR decreased sharply as the composting reaction entered the substrate limited stage. Then, about 200 hours later, when the composting reaction approached the end, the OUA became stable. Figure 3.2 shows the dynamic variations of OUR over time. Initially, the OUR increased slowly because the biomass concentration was low; 48 hours later, OUR increased sharply with the growth of biomass; 96 hours later, the OUR reached the peak. Then, the OUR decreased quickly when the soluble substrate concentration became low. Several hours later, the OUR turned into a stable state. Figure 3.3 provides the dynamic variations of soluble substrate over time. Initially, the soluble substrate concentration increased slowly since its production rate through hydrolysis was faster than its consumption rate. About 30 hours later, it reached the peak; then, it decreased quickly due to microbe growth. Approximately 100 hours later, the concentration of soluble substrate became very low. Since the soluble substrate concentration was assumed to be zero after the saturated stage, the subsequent variations of soluble substrate concentration are not shown in Figure 3.4. Figure 3.4 presents the dynamic variations of biomass concentration over time,

59

which were similar to those of OUR, since the OUR is directly related to the biomass concentration. At the start, the biomass concentration increased slowly; about 48 hours later, the biomass began to grow quickly. Around 96 hours later, it reached the peak; then the biomass concentration decreased rapidly due to the shortage of soluble substrate. Figure 3.5 shows the dynamic variations of insoluble substrate concentration over time. Initially, the concentration decreased slowly, since the biomass concentration was low and the hydrolysis rate was slightly faster than the production rate of insoluble material. Approximately 72 hours later, the concentration of insoluble substrate increased rapidly since a large amount of insoluble substrate was produced through microbe metabolism. About 100 hours later, the soluble substrate was consumed totally, and the composting reaction entered the substrate limited stage, i.e. the hydrolysis of insoluble substrate played the main role in the composting process. As a result, the insoluble substrate concentration kept decreasing slowly till the end of composting process. Figures 3.1 to 3.5 show the simulated variations of OUA, OUR, S s , S i , and X over time. These variations are similar to the experimental results as presented in the previous studies (Stombaugh and Nokes, 1996; Hamelers, 2001), indicating reasonable simulated results through this study. 3.5. Model Verification To verify the developed model, a series of composting experiments were undertaken. A bench-scale reactor was constructed by using a cylindrical PVC tube with an effective volume of approximately 30 liters (Figure 3.6). The air feed system was flowing upward through a fine mesh screen near the bottom of the reactor. The sensors

60

were used for online detection of system temperature.

61

1600 1400 1200 1000 800 600 400 200 0 0 24 48 72 96 120 144 168 192 Time (hr)

OUA (mol O 2 m )

-3

Figure 3.1 Variations of OUA over time (200 hours)

62

OUR (molm h )

-1 -3

40 30 20 10 0 0 24 48 72 96 120 144 168 192 Time (hr)

Figure 3.2 Variations of OUR over time (200 hours)

63

2400 2000 Ss (molm ) 1600 1200 800 400 0 0 24 48 72 Time (hr) 96 120
-3

Figure 3.3 Variations of soluble substrate concentration over time (120 hours)

64

270 240 210 180 150 120 90 60 30 0 0 24 48 72 96 120 144 168 192 Time (hr)

X (molm )

-3

Figure 3.4 Variations of active biomass concentration (200 hours)

65

4600 4400 Si (molm ) 4200 4000 3800 3600 3400 0 24 48 72 96 Time (hr) 120 144 168 192
-3

Figure 3.5 Variations of insoluble substrate concentration over time (200 hours)

66

6 1 2 3 4 5

9 10 11

12 13 14

Figure 3.6 Schematic diagram of the experimental reactor


1. Air vent-pipe; 2. Reactor body; 3. Heat preservation coat; 4. Air feed system; 5. Air flowrate meter; 6. Gas sampling pipe; 7. Cap of the reactor; 8. Sampling hole; 9. Airproof rubber band; 10. Composting substrate; 11. Temperature sensor; 12. Air feed pipe; 13. Plastic plate with small holes; 14. Wastewater pipe

67

The raw materials included apple, potato, rice, carrot, leaves, meat, sawdust, soy bean, soil, and coal ash. They were pre-treated for moisture and size control. During the experimental process, conditions of temperature, moisture and airflow rate were controlled; about 30 g of solid waste samples at three locations were collected twice a day; the concentrations of soluble, insoluble substrate, inert material, and active biomass, and moisture were measured with standard methods (Eaton et al., 1998). An oxygen analyzer was used for detecting oxygen concentration at the reactors inlet and outlet, such that the OUA and OUR can be quantified. Figures 3.7 to 3.11 show comparisons of the predicted and experimental results, with the related error analysis being shown in Table 3.2. It can be seen that the predicted results of Figures 3.9 and 3.10 well matched the experimental ones (the average relative error was 12.56 and 5.68%, respectively); in comparison, the predicted results of Figures 3.7, 3.8 and 3.11 had some differences from the experimental ones (the average relative error was 18.91, 25.98, and 19.82%, respectively). However, such results are generally acceptable considering the complexities and dynamics of the composting process. Thus, the developed model has generally acceptable accuracy levels. 3.6. Effects of Modeling Parameters Several parameters have effects on the simulation process. Particularly, the maximum conditional growth rate ( m ) and the biomass decay rate constant (b) have most significant effects, since they are directly related to microbial growth and death, as well as indirectly associated with the OUR of substrate. In this study, the effects of these two parameters were investigated in detail.

68

1800 1600 Predicted Value (mol O2m ) 1400 1200 1000 800 600 400 200 0 0 200 400 600 800 1000 1200 1400 1600 1800
-3 -3

Observed Value (mol O2m )

Figure 3.7 Predicted and observed OUAs

69

35 30 Predicted Value (mol O 2 m h )


-3 -1

25 20 15 10 5 0 0 5 10 15 20
-3

25
-1

30

35

Observed Value (mol O2 m h )

Figure 3.8 Predicted and observed OURs

70

3000 2500 2000 1500 1000 500 0 0 500 1000 1500 2000
-3

Predicted Value (mol Ssm)

-3

2500

3000

Observed Value (mol Ssm )

Figure 3.9 Predicted and observed soluble substrate concentrations

71

4600 4400 Predicted Value (mol Sim)


-3

4200 4000 3800 3600 3400 3200 3200 3400 3600 3800 4000 4200
-3

4400

4600

Observed Value (mol Sim )

Figure 3.10 Predicted and observed insoluble substrate concentrations

72

350 Predicted Value (mol Xm ) 300 250 200 150 100 50 0 0 50 100 150 200 250-3 300 Observed Value (mol Xm ) 350
-3

Figure 3.11 Predicted and observed active biomass concentrations

73

Table 3.2 Error analysis for model verification Average relative error (%) 18.91 25.98 12.56 5.68 19.82 Minimum relative error (%) 1.56 2.89 1.02 0.67 2.45 Maximum relative error (%) 105.42 150.08 38.36 15.12 135.25

Variable OUA OUR Ss Si X

74

Figure 3.12 shows the variations of OUA over time under three levels of maximum growth rate (0.9, 1.0, and 1.1 10-4 s-1, respectively). The OUA increased rapidly when m = 1.1 10-4 s-1, moderately when m = 0.9 10-4, and slowly when

m = 1 10-4. This implies that an increase m level would result in a raised OUA.
However, when the process entered the curing stage, the three curves would have almost identical trend, indicating that the total OUA levels were almost the same to each other. Figure 3.13 offers the OUR levels under different maximum growth rates. It is shown that although the peaks were different from each other, the overall trend were similar. First, the peak of OUR curve with m = 1.1 10-4 s-1 appeared; secondly, the peak with m = 1.0 10-4 s-1 took place; and lastly, the curve with m = 0.9 10-4 s-1 occurred. The peak of OUR under m = 1.1 10-4 was higher than that under the other two m conditions, being consistent with the results shown in Figure 3.13. This indicated that an increased m would result in a raised OUR level. Figure 3.14 presents the effect of m on soluble substrate concentration. The soluble substrate would be consumed rapidly under m = 1.1 10-4 s-1, but slowly under m = 0.9 10-4 s-1. This implies that a raised m level would result in an increased consumption rate of soluble substrate (which is related to microbial activities) as well as a raised OUR level. Figure 3.15 shows the effect of m on biomass concentration. The general trends are similar to those presented in Figure 3.14, since the biomass concentration is related to the OUR, i.e., a raised m level would increase the growth speed of microbes.

75

Figure 3.16 shows the effect of m level on insoluble substrate concentration. The three curves were of insignificant difference on the first 24 hours under all m levels. However, during the 48th to 120th hour, the insoluble substrate concentration increased significantly with an increased m : firstly, the peak of insoluble substrate with

m = 1.1 10-4 s-1 appeared; secondly, the peak with m = 1.0 10-4 s-1 took place; and
lastly, the curve with m = 0.9 10-4 s-1 occurred. After the curves entered the curing stage, the change of m did not show much effect on the composting process. The effects of biomass decay rate constant (b) are analyzed by setting b = 6.84, 7.60, and 8.36 10-5s-1, respectively. The results are presented in Figures 3.18 to 3.22. The OUR and soluble substrate consumption rate would be faster under b = 6.84 10-5 s-1 than that under b = 8.36 10-5 s-1. Generally, from Figures 3.12 to 3.16, an increased m levels would result in raised OUA, OUR, soluble substrate consumption rate, and microbe growth rate; from Figures 3.17 to 3.21, a decreased b level would help enhance the composting process. Moreover, variations in maximum growth rate would affect the composting process more significantly than those of biomass decay rate constant.

3.7. Summary A multi-component modeling system has been developed to simulate the composting process based on the principles of microbiology, mass transfer, and biotechnology. In the developed model, the state variables included levels of soluble substrate, insoluble substrate, active biomass, inert material, oxygen content, temperature,
76

and moisture; moreover, processes of microbial growth, decay, and hydrolysis were incorporated within the modelling framework. The effectiveness of the proposed model was demonstrated through its application to a case study. The simulated levels of OUA, soluble substrate concentration, insoluble substrate concentration, and biomass concentration over time were obtained through the developed model. The developed model was verified through the composting experiment and the results showed that the simulation results were generally consistent with the experimental outputs with acceptable accuracy levels. Moreover, further sensitivity analyses were conducted with the results demonstrating that an increased growth rate or decreased decay rate of biomass would lead to a raised OUR and thus an enhanced composting process. In general, the developed modelling system could help supply a solid basis for further research of composting process control. Based on the developed model, the interactive and uncertain effects of multiple factors on composting process could be further examined, where more advanced modeling techniques would be developed.

77

1600 1400 1200 1000 800 600 400 200 0 0 24 48 72 96 120 144 Time (hr)

OUA (mol O2 m-3)

m = 1.010-4s-1 m = 0.910-4s-1 m = 1.110-4s-1


168 192

Figure 3.12 Effect of m on OUA

78

OUR (mol O2m-3h-1)

40 30 20 10 0 0 24 48 72 96 120 144 168 192


Time (hr)

m = 1.010-4s-1 m = 0.910-4s-1 m = 1.110-4s-1

Figure 3.13 Effect of m on OUR

79

2400 2000 1600 1200 800 400 0 0 24 48 72 96 Time (hr)

Ss (molm-3)

m = 1.010-4s-1 m = 0.910-4s-1 m = 1.110-4s-1

120

144

Figure 3.14 Effect of m on soluble substrate concentration

80

X (mol m-3)

300 200 100 0 0 24 48 72 96 120 144

m = 1.010-4s-1 m = 0.910-4s-1 m = 1.110-4s-1

168

192

Time (hr) Figure 3.15 Effect of m on biomass concentration

81

4600 4400 4200 4000 3800 3600 3400 0 24 48 72 96 120 144 Time (hr)

Si (molm-3)

m = 1.010-4s-1 m = 0.910-4s-1 m = 1.110-4s-1


168 192

Figure 3.16 Effect of m on insoluble substrate concentration

82

OUA (mol O2m )

-3

1500 1200 900 600 300 0 0 24 48 72 96 120 144 168 192 Time (hr)
b = 7.610-5s-1 b = 6.8410-5s-1 b = 8.3610-5s-1

Figure 3.17 Effect of b on OUA

83

OUR (mol O2m h )

40 30 20 10 0 0 24 48 72 96 120 144 168 192 Time (hr)


b = 7.610-5s-1 b = 6.8410-5s-1 b = 8.3610-5s-1

-3

-1

Figure 3.18 Effect of b on OUR

84

2400 Ss (mol m ) 2000 1600 1200 800 400 0 0 24 48 72 Time (hr) 96 120 144
b = 7.610-5s-1 b = 6.8410-5s-1 b = 8.3610-5s-1
-3

Figure 3.19 Effect of b on soluble substrate concentration

85

400 X (mol m ) 300 200 100 0 0 24 48 72 96 120 144 168 192 Time (hr)
-3

b = 7.610-5s-1 b = 6.8410-5s-1 b = 8.3610-5s-1

Figure 3.20 Effect of b on active biomass concentration

86

4900 4600 4300 4000 3700 3400 0

Si (molm )

b = 7.610-5s-1 b = 6.8410-5s-1 b = 8.3610-5s-1

-3

24

48

72

96

120 144 168 192

Time (hr)

Figure 3.21 Effect of b on insoluble substrate concentration

87

CHAPTER 4 SIMULATION-BASED FACTORIAL ANALYSIS FOR CHARACTERIZING INTERACTIVE EFFECTS FROM MULTIPLE FACTORS
4.1. Statement of Problems In chapter 3, a multi-component modeling system has been developed to simulate the composting process based on the principles of microbiology, mass transfer, and biotechnology. Based on such a model, an advanced process control system could be developed. To achieve this objective, however, it is necessary to characterize the effects of various factors and their interactions on the system behaviors. Previously, the effects of various factors on composting processes were studied. The factors included airflow rate (Hall, 1998), substrate composition (Stentiford, 1996), reactor configuration (Haug, 1993, 1996), and moisture level (Robinson and Stentiford, 1993; Zhang, 2000). In addition, several studies for modelling composting processes were undertaken (Hall, 1998; Hamelelers, 2001). A number of parameters, processes, and reaction conditions were investigated, including temperature, moisture, oxygen content, carbon dioxide content, energy flow, and substrate properties (Kumar, 1993; Hall, 1998; Hamelelers, 2001; Kamil, 2001; Ryckeboer and Mergaert, 2003; Nelson, et al. 2003; Nakasaki et al., 2004, 2005). The above studies involved both experimental and modelling work. Experimental approaches have been useful in identifying the pertinent factors and their critical impacts over a range of factorial interactions. Yet, modeling system dynamics holds the potential for analyzing, not only contributions from numerous impact factors and their interactions,
88

but also the associated mechanism of reaction kinetics (Hall, 1998; Hamelelers, 2001). This is made possible through sensitivity analyses within a multi-level and multi-component context, wherein crucial factors and/or factor combinations can be identified, and their effects on the composting processes determined. However, in the previous modeling attempts, only effects from individual impact factors were highlighted, while the interactions among the factors were ignored (Hall, 1998). Examinations of the influences of multiple factors were typically determined through sensitivity analysis that varied a single factor (or a limited number of factors) at a time, while holding the others constant; consequently, the joint effects among multiple factors were ignored. The methodology of factorial analysis provides a viable means of studying such interactive effects (Box, 1978). Though not as yet applied to the composting studies, the method is amenable for being integrated within a composting process model to support multivariate inference of the interactive effects. The objective of this section, therefore, is to characterize the interactive effects of multiple factors on the composting process through a simulation-based factorial analysis methodology. In detail, A two-level factorial design combined with the developed composting process model will be introduced to analyze the influences of various factors on the composting process, with special attention paid to the interactive effects among the factors. To screen the most important factors when high-order interactions occur, normal probability plot will be employed for supporting result analysis. 4.2. Factorial Analysis of Interactive Effects in Composting Processes Factorial designs are widely used in experiments involving multiple factors where the joint effects of the factors on a response need to be analyzed. There are several
89

special cases of general factorial design that are important because they are widely used and form the bases for many designs of experimental work (Montgomery, 2000). The most important is the design of k factors, each at only two levels. These levels may be quantitative, such as two values of temperature, pressure, or time; or they may be qualitative, such as two machines, two operators, the high and low levels of a factor, or perhaps the presence and absence of a factor. A complete replicate of such a design requires 2 2 2 (Montgomery, 2000). Generally, the statistical model for a 2 K factorial design would include k main effects, two-factor interactions, three-factor interactions, , and one k-factor interaction. That is to say, for a 2 K design the complete model would contain 2 K -1 effects. The definition and notation for treatment combinations is based on the literature (Montgomery, 2000). The treatment combinations can be written in a standard order by introducing the factor one at a time, with each new factor being successively combined with those that precede it. For example, the standard order for a 24 design is (1), a, b, ab,
2 = 2 K observations, and is called a 2k

factorial design

c, ac, bc, abc, d, ad, bd, abd, cd, acd, bcd, and abcd (Table 4.1).
To estimate an effect or to compute the sum of squares for an effect, the contrast associated with that effects must be determined first. This can always be done by using a table of plus and minus signs, such as Table 4.1. However, this is awkward for large values of K; thus an alternate method can be used (Box, 1976). In general, the contrast for effect AB...K can be determined by expanding the right-hand side of

ContrastAB

= (a 1)(b 1)

(k 1)

(4.1)

In expanding Equation (4.1), the ordinary algebra is used with 1 being replaced
90

by (1) in the final expression. The sign in each set of parentheses is negative if the factor is included in the effect and positive if the factor is not included. To illustrate the use of Equation (4.1), consider a 25 factorial design. The contrast for ABCD would be:
Contrast ABCD = (a 1)(b 1)(c 1)(d 1)(e 1)

= abcde + cde + bde + ade + bce + ace + abe + e + abcd + cd + bd + ad + bc + ac + ab + (1) a b c abc d abd acd bcd ae be ce abce de abde acde bcde

(4.2)

Once the contrasts for the effects have been computed, we may estimate the effects and compute the sums of squares according to:
AB K= 2 (Contrast AB n2 k
K

(4.3)

and
SS AB
K

1 (Contrast AB n2 k

)2

(4.4)

respectively, where n denotes the number of replicates (Montgomery, 2000). There is also a tabular algorithm due to Dr. Yates (1959) that is useful for manual calculation of the effect estimates and the sums of squares (Box, 1976; Montgomery, 2000). For even a moderate number of factors, the total number of treatment combinations in a 2k factorial design is large. For example, a 25 design has 32 treatment combinations, a 26 design has 64 treatment combinations, and so on. However, in some cases, the available resources would only allow a single replicate of the design to be run. In such a situation, two problems arise in the assessment of effects from unreplicated factorials: (1) occasionally real and meaningful high-order interactions occur, and (2) it is necessary to allow for selection. A method (Daniel, 1959) by which effects are plotted on normal probability often provides an effective way to solve these two difficulties.
91

Table 4.1 Algebraic signs for calculating effects in the 24 design


Treatment
Combination (1) a b ab c ac bc abc d ad bd abd cd acd bcd abcd I + + + + + + + + + + + + + + + + A + + + + + + + + B + + + + + + + + AB + + + + + + + + C + + + + + + + AC + + + + + + + + BC + + + + + + + +

Factorial Effect
ABC + + + + + + + + D + + + + + + + + AD + + + + + + + + BD + + + + + + + + ABD + + + + + + + + CD + + + + + + + + ACD + + + + + + + + BCD + + + + + + + + ABCD + + + + + + + +

92

For a composting system, the factors such as temperature, moisture and oxygen content can be controlled and the initial biomass concentration can be adjusted through incubation. To improve the operation and control of composting processes, it is necessary to determine the effects of these impact factors, especially in terms of the interactive effects. To achieve this goal, a two level factorial design combined with the developed model will be introduced to analyze the effects of various factors on the composting process, especially for the interactive effects among multiple factors. To screen the significant factors, normal probability plot will be generated to support significance analysis. Temperature (A), moisture (B), oxygen content (C) and initial biomass concentration (D) were each set at two levels. The maximum and minimum values of these factors were determined based on literature review (Haug 1993; Hall 1998; Hamelers 2001). Therefore, there were 16 sets of experimental tests. Through the composting model, the predicted concentrations of soluble substrate, insoluble substrate, active biomass and total substrates for each experimental set could be obtained. Then, the main effects of each parameter could be calculated with the normal probability levels for the main effects being plotted. 4.3. Case Study Consider a reactor-based composting system where the interactive effects of various factors are to be examined. It is assumed that the composting materials are pre-treated before the reaction process. The initial concentrations of soluble substrate, insoluble substrate, and inert material are set as 2200, 3600, and 400 molm-3, respectively. Thus, the ranges of parameters for the 24 factorial design are listed in Table 4.2.
93

Table 4.2 The 24 factorial design


Temperature (C) (A) 35 55 35 55 35 55 35 55 35 55 35 55 35 55 35 55 Moisture (%) (B) 40 40 60 60 40 40 60 60 40 40 60 60 40 40 60 60 Oxygen (molm ) (C) 0.05 0.05 0.05 0.05 0.15 0.15 0.15 0.15 0.05 0.05 0.05 0.05 0.15 0.15 0.15 0.15
-3

Run Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

Run Label (1) a b ab c ac bc abc d ad bd abd cd acd bcd abcd

Biomass (molm-3) (D) 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0

94

In this study, the temporal points of the 24th, 72nd and 144thhr were analyzed, which mostly characterised the composting process at three different stages: initial, active, and curing stage. Based on the developed composting-process model, the concentrations of soluble substrate (Ss), insoluble substrate (Si), active biomass (X) and total substrates (TS) can be calculated (Table 4.3). 4.3.1. Main Effects and Normal Probability Plot for Composting Substrates (1) Soluble substrate Based on Table 4.3, the main effects on soluble substrate concentration can be derived as shown in Table 4.4. At the 24th hour, factors D, A and C had effects on the soluble substrate concentration. At the same time, the interaction A-D also had an important effect. This implies that the initial biomass concentration was the most critical factor that affected the soluble substrate concentration in the initial stage; this was followed by factors of temperature and oxygen content. Among the four factors, only moisture was positively correlated to soluble substrate, while the other three factors were negatively correlated. Therefore, it could be concluded that the effects of initial biomass concentration and temperature are more significant than those of oxygen concentration and moisture at the initial stage; in addition, the effect from the interaction A-D could not be ignored. However, due to the slow composting process in the initial stage, the effect levels were relatively low. At the 72nd hour, factors A and C and the interaction A-C had significant effects on the soluble substrate concentration. It is indicated that when reaching the active stage, temperature would become the most crucial factor, followed by oxygen content (instead of the initial biomass concentration). Moreover, the effect of interaction A-C also became significant. Similarly, only moisture was positively correlated to soluble substrate.
95

Table 4.3 Simulation results for the 24 design


Run Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 Run Label (1) a b ab c ac bc abc d ad bd abd cd acd bcd abcd 24th hr 2243 2250 2243 2251 2242 2247 2242 2247 2223 2207 2224 2211 2221 2196 2221 2198 Ss (molm-3) 72nd hr 2043 1217 2076 1441 1917 1945 250.9 1869 955.1 1933 1238 1597 1662 144th hr 620.30 995.83 339.54 780.56 Si (molm-3) 24th hr 3544 3514 3543 3514 3543 3515 3543 3514 3551 3527 3551 3526 3551 3529 3551 3529 72nd hr 3550 3820 3538 3736 3591 4318 3582 4166 3628 3947 3604 3835 3723 4467 3702 4416 144th hr 4155 4432 3980 4434 4556 4423 4554 4424 4309 4428 4098 4432 4556 4418 4555 4417 24th hr 4.4101 13.139 4.345 12.714 4.5585 14.191 4.5343 14.013 8.4167 23.750 8.1889 22.463 8.9656 27.280 8.8742 26.649 X (molm-3) 72nd hr 44.072 158.96 38.022 125.21 68.936 305.16 63.216 322.41 58.506 170.33 48.745 132.98 107.05 151.67 94.474 201.83 144th hr 96.761 22.522 74.515 27.045 15.962 21.733 21.464 21.747 86.753 22.044 66.937 23.880 13.330 21.690 14.225 21.691 TS (molm-3) 24th hr 6190 6177 6191 6177 6190 6175 6190 6175 6182 6157 6182 6159 6181 6152 6181 6153 72nd hr 6037 5596 6053 5702 5978 5024 5991 5140 5956 5472 5986 5606 5827 5018 5858 5017 144th hr 5272 4854 5450 4861 4972 4845 4975 4845 5136 4850 5346 4856 4970 4839 4970 4839

96

Table 4.4 Main effects for soluble substrate concentration


24th hr Effects A B C D AB AC AD BC BD CD ABC ABD ACD BCD ABCD -6.725 1.15 -4.525 -33.1 0.75 -3.075 -12.65 -0.45 0.625 -2.45 -0.3 0.375 -1.65 -0.225 -0.125 Sum of squares 180.90 5.29 81.902 4382.4 2.25 37.822 640.09 0.81 1.5625 24.01 0.36 0.5625 10.89 0.2025 0.0625 Percent contribution 0.0337 9.8526e-004 0.0153 0.8162 4.1906e-004 0.0070 0.1192 1.5086e-004 2.9101e-004 0.0045 6.7050e-005 1.0477e-004 0.0020 3.7715e-005 1.1641e-005 72nd hr Effects -1242.7 118.307 -674.64 -204.78 70.9575 -475.24 25.8175 -32.540 -15.484 -9.2349 -31.240 -32.484 62.7151 -37.632 -39.882 Sum of squares 6177600 55987 1820600 167740 20140 903420 2666.2 4235.4 959.1301 341.1344 3903.8 4221.1 15733 5664.8 6362.4 Percent contribution 0.6722 0.0061 0.1981 0.0183 0.0022 0.0983 2.9013e-004 4.6090e-004 1.0437e-004 3.7122e-005 4.2481e-004 4.5933e-004 0.0017 6.1644e-004 6.9236e-004 144th hr Effects -342.02 102.068 -342.02 -62.004 -102.06 342.02 62.004 -102.06 8.1853 62.004 102.06 -8.185 -62.00 -8.185 8.1853 Sum of squares 467940 41672 467940 15378 41672 467940 15378 41672 267.9974 15378 41672 267.9974 15378 267.9974 267.9974 Percent contribution 0.2865 0.0255 0.2865 0.0094 0.0255 0.2865 0.0094 0.0255 1.6411e-004 0.0094 0.0255 1.6411e-004 0.0094 1.6411e-004 1.6411e-004

97

At the 144th hour, the main effects of A, C and B, and interaction A-C were significant. This implied that temperature and oxygen content were the most important factors, instead of the initial biomass concentration and moisture in the curing stage. In addition, temperature, oxygen content and the initial biomass concentration all had the negative effects. At the 72nd and 144th hour, a number of high-order interactions occurred in Table 4.4. Therefore, a normal probability plot could be produced to screen the important factors. The information for normal probability plot is provided in Table 4.5 and Figures 4.1 and 4.2. It is indicated that all of the effects that lie along the line are negligible, whereas significant effects should be far from the line. In Figure 4.1 (the 72nd hour), the points of A, C, D and A-C were away from the reference line; thus, temperature, oxygen content, the initial biomass concentration, and the interaction of temperature and oxygen content had significant effects, which were consistent with the result as shown in Table 4.4. In Figure 4.2 (the 144th hour), the important effects that emerged from this analysis were those of A, C and A-C. Although the factor B, interactions A-B and A-D were also significant in Table 4.4, they lied along the reference line of Figure 4.2, demonstrating that their effects were negligible. The above analysis shows that the order of significant factors related to soluble substrate concentration is the initial biomass concentration, temperature, oxygen content and moisture at the 24th hour; in addition, interaction A-D can not be neglected. At the 72nd hour, the order becomes temperature, oxygen content, the initial biomass concentration and moisture, and interaction A-C also has important effect. At the 144 th

98

3 2 Z-Score 1 0 -1500 -1000 -500 -1 0 -2 -3 Main effect


Figure 4.1 Normal probability plot for soluble substrate concentration (the 72nd hour)

500

99

3 2 Z-Score 1 0 -400 -200 -1 -2 -3 Main effect


Figure 4.2 Normal probability plot for soluble substrate concentration (the 144th hour)

200

400

100

Table 4.5 Data for normal probability plots (for the effects of soluble substrate concentration)
Number (N) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Ss Main effects 72nd hr -1.2427e+003 -674.6425 -475.2425 -204.7825 -39.8825 -37.6325 -32.5401 -32.4849 -31.2401 -15.4849 -9.2349 25.8175 62.7151 70.9575 118.3075 144th hr -342.0290 -342.0290 -102.0687 -102.0687 -62.0046 -62.0046 -8.1853 -8.1853 8.1853 8.1853 62.0046 62.0046 102.0687 102.0687 342.0290 Rank (R) 1 (1.5) 2 (1.5) 3 (3.5) 4 (3.5) 5 (5.5) 6 (5.5) 7 (7.5) 8 (7.5) 9 (9.5) 10 (9.5) 11 (11.5) 12 (11.5) 13 (13.5) 14 (13.5) 15 Normal Probability Plot (R-0.5)/N 0.033 (0.067) 0.100 (0.067) 0.167 (0.200) 0.233 (0.200) 0.300 (0.333) 0.367 (0.333) 0.433 (0.467) 0.500 (0.467) 0.567 (0.600) 0.633 (0.600) 0.700 (0.733) 0.767 (0.733) 0.833 (0.867) 0.900 (0.867) 0.967 Z-score -1.84 (-1.5) -1.28 (-1.5) -0.97 (-0.842) -0.73 (-0.842) -0.525 (-0.435) -0.34 (-0.435) -0.17 (-0.085) 0 (-0.085) 0.17 (0.255) 0.34 (0.255) 0.525 (0.627) 0.73 (0.627) 0.97 (1.125) 1.28 (1.125) 1.84

101

hour, the order is temperature, oxygen content, moisture and the initial biomass concentration; furthermore, the interaction A-C should be considered. (2) Insoluble substrate Based on Table 4.3, the main effects on insoluble substrate concentration can be calculated and listed in Table 4.6. At the 24th hour, factors A and D had important effects on the insoluble substrate; interaction A-D also showed the significant effect. This implies that temperature was the critical factor that affected the insoluble substrate concentration at the initial stage; this was followed by factors of initial biomass concentration, oxygen content and moisture. Among the four factors, only moisture was positively correlated to soluble substrate concentration. Similarly, due to the slow composting process at the initial stage, the effect levels were relatively low. At the 72nd hour, factors A, C and D and interaction A-C had the significant effects on the insoluble substrate concentration. This implies that when composting reaching the active stage, temperature and oxygen content would become the most important factors on the insoluble substrate while not the initial biomass concentration and moisture. In addition, the effect of interaction A-C could not be overlooked. Temperature, the initial active biomass and oxygen content all had the plus sign, i.e., they had the positive effects on the insoluble substrate concentration. At the 144th hour, factors C, A and B had the significant effects on the insoluble substrate concentration. In addition, interactions A-B and B-C had the similar effects. This implies that when composting process reached the curing stage, oxygen content was the most crucial factor, followed by temperature and moisture; while the initial biomass concentration only had a little effect. Among all the factors, only moisture was negatively
102

correlated to the insoluble substrate concentration. Similarly, the information for normal probability plot is provided in Table 4.7 and Figures 4.3 and 4.4. In Figure 4.3 (the 72nd hour), factors of A, C, D and interaction A-C were far away the reference line. This implied that temperature, oxygen content and the initial biomass concentration had the significant effects on the insoluble substrate concentration at the active stage, which was consistent with the results of Table 4.6. In Figure 4.4 (the 144th hour), the important effects that emerged from this analysis were the effects of C, A, and interaction A-C. Oxygen content was more significant than temperature because the point of A was more close to the reference line. The main effects of factor B and interactions A-B and A-D could be ignored. The above analysis shows that the order of effect factors for the insoluble substrate concentration is the initial biomass concentration, temperature, oxygen content and moisture at the 24th hour; in addition, interaction A-D is not negligible. At the 72nd hour, the order becomes temperature, oxygen content, the initial biomass concentration, and moisture; interaction A-C is also significant. At the 144th hour, the order of effect factors is oxygen content, temperature, moisture, and the initial biomass concentration. (3) Active biomass Based on Table 4.3, the main effects on active biomass concentration can be calculated and listed in Table 4.8. At the 24th hour, factors A and D had the effects on active biomass concentration. This implies that the effect of temperature was more important than that of the initial biomass concentration, while factors B and C did not show much effect because of the low activity of microorganism. In addition, only moisture had the negative effect.
103

Table 4.6 Main effects for insoluble substrate concentration


24th hr Effects A B C D AB AC AD BC BD CD ABC ABD ACD BCD ABCD -26.22 -0.25 0.925 140 -0.15 0.575 3 0.1 -0.125 0.5 0.05 -0.075 0.3 0.025 0.025 Sum of squares 2751 0.25 3.4225 453.69 0.09 1.3225 36 0.04 0.0625 1 0.01 0.0225 0.36 0.0025 0.0025 Percent contribution 0.8472 7.6988e-005 0.0011 0.1397 2.7716e-005 4.0726e-004 0.0111 1.2318e-005 1.9247e-005 3.0795e-004 3.0795e-006 6.9289e-006 1.1086e-004 7.6988e-007 7.6988e-007 Effects 473.22 -58.07 219 127.35 -41.62 219 28.45 -0.3 6 34.925 -1.6 12.35 7.975 15.825 16.225 72nd hr Sum of Squares 895770 13491 332580 64872 6930.6 191840 3237.6 0.36 144 4879 10.24 610.09 254.4025 1001.7 1053 Percent contribution 0.5906 0.0089 0.2193 0.0428 0.0046 0.1265 0.0021 2.3736e-007 9.4944e-005 0.0032 6.7516e-006 4.0225e-004 1.6774e-004 6.6047e-004 6.9428e-004 Effects 80.4625 -47.887 204.487 32.1125 49.4375 -215.48 -36.562 46.9125 -4.3125 -34.737 -48.662 4.4125 32.9875 4.3875 -5.0875 144th hr Sum of squares 25897 9172.9 167260 4124.9 9776.3 185740 5347.3 8803.1 74.3906 4826.8 9472.2 77.8806 4352.7 77.0006 103.5306 Percent contribution 0.0595 0.0211 0.3844 0.0095 0.0225 0.4269 0.0123 0.0202 1.7097e-004 0.0111 0.0218 1.7899e-004 0.0100 1.7697e-004 2.3794e-004

104

Table 4.7 Data for normal probability plot (for the effects of insoluble substrate concentration)
Number (N) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Si Main effects 72nd hr -58.0750 -41.6250 -1.6000 -0.3000 6.0000 7.9750 12.3500 15.8250 16.2250 28.4500 34.9250 127.3500 219.0000 219.0000 473.2250 144th hr -215.4875 -48.6625 -47.8875 -36.5625 -34.7375 -5.0875 -4.3125 4.3875 4.4125 32.1125 32.9875 46.9125 49.4375 80.4625 204.4875 Normal Probability Plot Rank (R) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 (R-0.5)/N 0.033 0.1 0.167 0.233 0.3 0.367 0.433 0.5 0.567 0.633 0.7 0.767 0.833 0.9 0.967 Z-score -1.84 -1.28 -0.97 -0.73 -0.525 -0.34 -0.17 0 0.17 0.34 0.525 0.73 0.97 1.28 1.84

105

3 2 Z-Score 1 0 -200 -100 -1 -2 -3 Main effect


Figure 4.3 Normal probability plot for insoluble substrate concentration (the 72nd hour)

100

200

300

400

500

600

106

3 2 Z-Score 1 0 -300 -200 -100 -1 0 -2 -3 Main effect


Figure 4.4 Normal probability plot for insoluble substrate concentration (the 144th hour)

100

200

300

107

Table 4.8 Main effects for active biomass concentration


24th hr Effects A B C D AB AC AD BC BD CD ABC ABD ACD BCD ABCD 12.738 -0.366 1.455 7.835 -0.264 1.062 3.685 0.135 -0.193 0.7827 0.0908 -0.1357 0.5586 0.063 0.0391 Sum of squares 649.1005 0.5366 8.4681 245.5614 0.2790 4.5116 54.3412 0.0731 0.1492 2.4505 0.0330 0.0736 1.2480 0.0159 0.0061 Percent contribution 0.6714 5.5495e-004 0.0088 0.2540 2.8862e-004 0.0047 0.0562 7.5567e-005 1.5434e-004 0.0025 3.4128e-005 7.6156e-005 0.0013 1.6420e-005 6.3330e-006 Effects 130.69 -4.725 67.239 -20.04 3.8026 31.155 -43.68 17.001 2.3414 -31.12 17.624 4.9846 -42.17 4.1699 4.9572 72nd hr Sum of squares 68326 89.3153 18085 1607.8 57.8387 3882.6 7632.8 1156.2 21.9293 3874.9 1242.4 99.3854 7116.3 69.5535 98.2958 Percent contribution 0.6027 7.8789e-004 0.1595 0.0142 5.1022e-004 0.0343 0.0673 0.0102 1.9345e-004 0.0342 0.0110 8.7672e-004 0.0628 6.1356e-004 8.6711e-004 Effects -25.9495 -3.6615 -33.5768 -3.9 5.255 31.4196 2.9645 5.2646 -0.6098 1.4075 -6.8504 -0.0652 -0.5213 -0.5456 1.2138 144th hr Sum of squares 2693.5 53.626 4509.6 60.84 110.46 3948.8 35.153 110.86 1.4873 7.9239 187.71 0.017 1.0869 1.1907 5.8932 Percent contribution 0.2297 0.0046 0.3845 0.0052 0.0094 0.3367 0.0030 0.0095 1.2682e-004 6.7564e-004 0.0160 1.4510e-006 9.2676e-005 1.0153e-004 5.0249e-004

108

At the 72nd hour, factors A and C and interactions A-D and A-C had the significant effects on the active biomass concentration. This implies that when the composting process reached the active stage, temperature was the most critical factor that affected active biomass concentration at the initial stage; this was followed by factors of oxygen content and initial biomass concentration. Also, interactions A-D and A-C could not be neglected. Among the factors, temperature and oxygen content were positively correlated to active biomass concentration. At the 144th hour, factors C and A and interaction A-C had the significant effects on the active biomass concentration. This implies that oxygen content was the most crucial factor; this was followed by temperature and moisture, when the composting process reached the curing stage. Besides, they all were negatively correlated to the active biomass concentration. The information for normal probability plots is provided in Table 4.9 and Figures 4.5 and 4.6. In Figure 4.5 (the 72nd hour), factors of A and C were away the reference

line. This implies that only temperature and oxygen content had the significant effects on the active biomass concentration at the active stage. In Figure 4.6 (the 144th hour), the important effects that emerged from this analysis were the main effects of C, A and interaction A-C. Oxygen content was more important than temperature because the point of A was more close to the reference line in Figure 4.6. This result is consistent with the result shown in Table 4.8. The above analysis shows that the order of effect factors for the active biomass concentration is temperature, the initial biomass concentration, oxygen content and moisture at the 24th hour. At the 72nd hour, the order becomes temperature, oxygen

109

content, moisture and the initial biomass concentration. At the 144th hour, the order of factors is oxygen content, temperature, the initial biomass concentration, and moisture; the interactive effect of temperature and oxygen content is also significant. (4) Total substrates Based on Table 4.3, the main effects on the total substrates can be calculated and listed in Table 4.10. At the 24th hour, factors A and D had the effects on the concentration of total substrates. This implies that the effect of temperature was more significant than that of initial biomass concentration; while factors B and C did not show much effect because of the low microbial activities. In addition, only moisture had the positive effect. At the 72nd hour, factors A and C and interaction A-C had the important effects on the total substrates concentration. This implies that temperature was the most critical factor that affected the total substrates at the initial stage; this was followed by factors of oxygen content and the initial biomass concentration. Among the four factors, only moisture was positively correlated to the total substrate concentration. Similarly, at the 144th hour, factors A and C and interaction A-C had the significant effects on the total substrate concentration. Also, only moisture had the positive effect. The information for normal probability plots is provided in Table 4.11 and Figures 4.7 and 4.8. In Figure 4.7 (the 72nd hour), factors A and C and the interaction A-C were far away the reference line. This implies that temperature and oxygen had the significant effects on total substrate concentration at the active stage; also, their interaction could not be neglected. This result is consistent with the result shown in Table 4.10.
110

Table 4.9 Data for normal probability plots (for the effects of active biomass concentration)
Number (N) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 X Main effects 72nd hr -43.6828 -42.1792 -31.1243 -20.0490 -4.7253 2.3414 3.8026 4.1699 4.9572 4.9846 17.0018 17.6240 31.1554 67.2395 130.6965 144th hr -33.5768 -25.9495 -6.8504 -3.9000 -3.6615 -0.6098 -0.5456 -0.5213 -0.0652 1.2138 1.4075 2.9645 5.2550 5.2646 31.4196 Normal Probability Plot Rank (R) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 (R-0.5)/N 0.033 0.1 0.167 0.233 0.3 0.367 0.433 0.5 0.567 0.633 0.7 0.767 0.833 0.9 0.967 Z-score -1.84 -1.28 -0.97 -0.73 -0.525 -0.34 -0.17 0 0.17 0.34 0.525 0.73 0.97 1.28 1.84

111

3 2 Z-Score 1 0 -100 -50 -1 0 -2 -3 Main effect


Figure 4.5 Normal probability plot for active biomass concentration (the 72nd hour)

50

100

150

112

3 2 Z-Score 1 0 -40 -20 -1 0 -2 -3 Main effect 20 40

Figure 4.6 Normal probability plot for active biomass concentration (the 144th hour)

113

Table 4.10 Main effects for total substrate concentration


24th hr Effects A B C D AB AC AD BC BD CD ABC ABD ACD BCD ABCD -20.20 0.5250 -2.100 -14.60 0.3500 -1.475 -5.97 -0.200 0.3000 -1.175 -0.125 0.1750 -0.800 -0.075 -0.050 Sum of squares 1632 1.1025 17.64 852.64 0.4900 8.7025 142.80 0.1600 0.3600 5.5225 0.0625 0.1225 2.5600 0.0225 0.0100 Percent contribution 0.6126 4.1380e-004 0.0066 0.3200 1.8391e-004 0.0033 0.0536 6.0052e-005 1.3512e-004 0.0021 2.3458e-005 4.5977e-005 9.6083e-004 8.4448e-006 3.7533e-006 Effects -638.81 55.461 -319.08 -97.53 33.162 -225.0 10.612 -15.86 -7.112 -5.412 -15.26 -15.16 28.537 -17.63 -18.68 72nd hr Sum of squares 1.6323e+006 1.2304e+004 4.0727e+005 3.8054e+004 4.3990e+003 2.0266e+005 450.5006 1.0065e+003 202.3506 117.1806 931.7756 919.6056 3.2576e+003 1.2443e+003 1.3969e+003 Percent contribution 0.7077 0.0053 0.1766 0.0165 0.0019 0.0879 1.9531e-004 4.3636e-004 8.7729e-005 5.0804e-005 4.0397e-004 3.9870e-004 0.0014 5.3948e-004 6.0562e-004 Effects -287.55 50.500 -171.12 -33.800 -47.375 158 28.425 -49.90 3.2750 28.700 46.575 -3.850 -29.525 -4.3250 4.3000 144th hr Sum of squares 3.3074e+005 10201 1.1714e+005 4.5698e+003 8.9776e+003 99856 3.2319e+003 9.9600e+003 42.9025 3.2948e+003 8.6769e+003 59.2900 3.4869e+003 74.8225 73.9600 Percent contribution 0.5509 0.0170 0.1951 0.0076 0.0150 0.1663 0.0054 0.0166 7.1459e-005 0.0055 0.0145 9.8754e-005 0.0058 1.2463e-004 1.2319e-004

114

Table 4.11 Data for normal probability plots (for the effects of total substrate concentration)
Number (N) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 TS Main effects 144th hr -287.5500 -171.1250 -49.9000 -47.3750 -33.8000 -29.5250 -4.3250 -3.8500 3.2750 4.3000 28.4250 28.7000 46.5750 50.5000 158 Normal Probability Plot Rank (R) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 (R-0.5)/N 0.033 0.1 0.167 0.233 0.3 0.367 0.433 0.5 0.567 0.633 0.7 0.767 0.833 0.9 0.967 Z-score -1.84 -1.28 -0.97 -0.73 -0.525 -0.34 -0.17 0 0.17 0.34 0.525 0.73 0.97 1.28 1.84

72nd hr -319.0875 -225.0875 -97.5375 -18.6875 -17.6375 -15.8625 -15.2625 -15.1625 -7.1125 -5.4125 10.6125 28.5375 33.1625 55.4625 55.4625

115

3 2 Z-Score 1 0 -400 -300 -200 -100 -1 0 -2 -3 Main effect


Figure 4.7 Normal probability plot for total substrate concentration (the 72nd hour)

100

200

116

3 2 Z-Score 1 0 -400 -300 -200 -100 -1 -2 -3 Main effect


Figure 4.8 Normal probability plot for total substrate concentration (the 144th hour)

100

200

117

In Figure 4.8 (the 144th hour), the important effects that emerged from this analysis were the main effects of A, C and interaction A-C. This result was also consistent with the result as shown in Table 4.11. The above analysis shows that the order of effect factors is temperature, the initial biomass concentration, oxygen content, and moisture at the 24th hour. At the 72nd hour, the order becomes temperature, oxygen content, the initial biomass concentration, and moisture. At the 144th hour, the order is temperature, oxygen content, moisture, and the initial biomass concentration. In addition, interaction A-C has a significant main effect at the both 72nd and 144th hours.

4.3.2. Discussion A two-level factorial design method combined with the developed

composting-process model has been introduced to analyze the effects of various factors on the composting process, especially for the interactive effects among multiple factors. At the 24th hour, the initial biomass concentration showed a significant effect on the soluble substrate concentration; this was followed by temperature and oxygen content. Temperature showed the significant effects on the concentrations of insoluble substrate, active biomass and total substrates, which was followed by the initial biomass concentration. For the soluble and insoluble substrate, the interactive effect of temperature and initial biomass concentration could not be neglected. Due to the low activity of microorganism, most of the levels were relatively low at the initial composting stage. At the 72nd hour, temperature became the most crucial factor of the composting

118

process, and then was oxygen content. For the concentrations of soluble, insoluble substrate and total substrates, the interactive effect of temperature and oxygen content could not be ignored. At the 144th hour, temperature was the most important factor for the concentrations of soluble substrate and total substrates, followed by oxygen content; while oxygen content was the most important factor for the concentrations of insoluble substrate and active biomass, and then was temperature. Moisture and the initial biomass concentration did not show significant effects. For the soluble and total substrates, the interaction of temperature and oxygen content was also significant. To identify the most significant factors when high-order interactions occurred, normal probability plots were produced. In these plots, not all of the main effects of factors well fitted the normal probability distributions. In fact, due to the process dynamics, it was impossible that the effects be strictly linear. For those that were approximately linear, the normal probability plots would be helpful for determining the most important impact factors. 4.4. Summary In this section, a simulation-based factorial analysis approach has been developed to characterize the interactive effects of multiple factors on composting processes. It is based on the developed composting-process model and the two level factorial analysis method. To identify significant factors when high-order interactions occur, normal probability plots have been produced to support the systems analysis. The results showed that the study factors had various effects in different stages of the composting process in different stage. At the 24th hour, the factors did not significant
119

effects. At the 72nd hour, when the composting process reached the active stage, many factors became significant, especially for temperature and oxygen content; their interactive effects were also significant. At the 144th hour, temperature and oxygen content still had significant effects on the composting process, but not as significant as those in the active stage. The above results could be useful for guiding composting-process operation and management. Varied control strategies could be implemented in different composting stages. Moreover, the results could be verified through composting experiments and thus provide guidance for further improving the developed composting-process model.

120

CHAPTER 5 SIMULATION-BASED FRACTIONAL FUZZY VERTEX METHOD COMBINED WITH INTERVAL ANALYSIS FOR CHARACTERIZING EFFECTS OF UNCERTAINTIES

5.1. Statement of Problems In Chapter 3, a multi-component composting model has been developed. Based on the developed model, the effects of multiple impact factors on the composting process have been studied in Chapter 4. These studies are helpful for supporting the control and operation of composting systems. However, composting systems are complex with dynamic and uncertain features. The system behaviours would change over time as microbial populations grow, evolve, die, and/or are replaced by new populations (McCarty, 1971); moreover, uncertainties exist in a variety of system components, forming an additional category of complexities. Thus, effective composting process control and operation would have to be based on the sound insight of such complexities. Previously, a number of studies were conducted to characterize composting processes (Stombaugh and Nokes, 1996; Hall, 1998; Hamelers, 2001; Alpert, et la. 2002; Ryckeboer and Mergaert, 2003). Multiple process parameters were considered, including temperature, moisture, oxygen content, carbon dioxide content, energy flow, and substrate properties (Haug, 1993, 1996; Ekinci, 2001; Hamelelers, 2001; Nakasaki et al., 2004, 2005). For example, Jeris and Regan (1973a, b) studied the effects of oxygen and carbon dioxide levels on composting processes. The results showed that measurement of carbon dioxide or oxygen should be sufficient to define the rate of degradation under a known flow rate. MacGregor et al. (1981) studied the effect of temperature on the
121

composting process, and concluded that the temperature of compost was a function of the accumulation of heat generated metabolically and, simultaneously, the temperature is also a determinant of metabolic activities. Miller (1989) thought water was a critical ecological as well as physical factor in substrate-dense matrix ecosystems of which composting systems were an example; sufficient moisture in the solid waste was required for maximum efficiency of microbial stabilization. At low moisture content (below 30%), microbial activity might be greatly reduced by the shortage of water (Iriarte and Ciria, 2001). Generally, these studies were done to characterize composting processes in two ways: one was only through composting experiment to determine the effects of various factors; the other was based on composting-process models combined with experimental studies and sensitivity analyses. However, almost all of these studies mainly focused on deterministic composting conditions, and while the effects of uncertain factors were seldom taken into account. Furthermore, a number of attempts were made in modeling composting processes, with most of the developed models being empirical; even though a few modes were taken into account microbial kinetics, they were significantly simplified to address the complex composting process. In fact, the purpose of characterizing a composting process is to control the process to obtain satisfied composting products. A model-based process control system can help reduce the cost of system operation and generate desired final products. Therefore, it is desirable to build a model-based process control system for solid waste composting. However, in a model-based process control system, two kinds of uncertainties will

122

significantly affect the simulation and process-control efforts. Firstly, composting substrates involve multiple components, varying with garbage sources. As a result, the parameters of composting-material composition will be associated with a number of uncertainties. Secondly, the controlled factors need to be maintained within targeted ranges during the composting process, which can be challenging in practical cases. For example, when the operator wants to control the composting temperature at 55C, the actual temperature can then be 54 to 56C since many ambient factors will affect the control effort. Existence of these uncertainties leads to difficulties when a model-based process control system is developed. Unfortunately, no previous study was conducted in this research field. As a result, the objective of this section is to develop a simulation-based fractional fuzzy vertex method combined with interval analysis for characterizing effects of uncertainties. The study will be based on the developed composting-process model as shown in Chapter 3. The fractional fuzzy vertex method will be used to characterize the effects of uncertain control factors, and the interval analysis will be to tackle the effects of uncertain input parameters. In addition, the degrees of influence from the uncertain control factors will be quantified through a series of sensitivity analyses.

5.2. Fuzzy Characterization on the Effects of Uncertainties Among methods for characterizing the effects of uncertainties, there are three main approaches: probabilistic analysis, interval analysis, and fuzzy set theory. The probabilistic analysis is a main measure to address uncertainties in modeling studies. It has been widely applied to environmental studies during the past decades. In such an

123

approach, uncertainties are described by probability distributions, with the modeling outputs being characterized in probabilistic terms. The most popular technique of this methodology is Monte Carlo simulation (Lahkim and Garcia, 1999) which involves repeated generations of pseudo-values for uncertain parameters drawn from known probability distributions to produce probability and cumulative distribution. Thus, a cumulative probability under which a specific event will happen can be derived. Such probabilistic information can then be interpreted into an indicator of uncertainties. However, in engineering practices, it is often difficult to build a probability distribution due to the lack of data or the high cost of getting the data (Wood et al., 1990; Liu, et, al, 2003). In the interval analysis approach, each uncertain parameter is assumed to have upper and lower limits without a probability structure. It is especially useful when the uncertainties are highly extensive. This method has been widely used in optimization problems of water pollution control planning and solid waste management (Huang and Moore, 1993; Luo et al., 2004). Another approach based on fuzzy set theory has been proved to be more practical. This method facilitates the analysis of systems with uncertainties being derived from vagueness or fuzziness rather than randomness alone. Thus, uncertainties can be handled in a direct way without a large number of realizations. This approach has been successfully applied to many fields, such as risk assessment, decision analysis, engineering design, and optimization (Bosma et al., 1993; Dou et al., 1995; Saikuman and Mujumdar, 1998). In fact, many applications of fuzzy arithmetic were mostly based on the fuzzy vertex method. With this approach, uncertain parameters are numerically

124

implemented as fuzzy sets, and arithmetical operations are carried out by discretization on the membership domain. Uncertainties can thus be represented in the outputs as membership functions. However, the number of simulation runs could become extremely high when many uncertain parameters and/or -cut levels were investigated; this could lead to high computational requirements and hinder applicability of such a method. Therefore, a modified fuzzy vertex method is desirable for facilitating its practical applicability. In the fuzzy set theory, a membership function [ A (x ) ] that is continuous and has a range of [0, 1] can be used to represent the degree to which an element belongs to the set in question (Zadeh, 1965). The -cut level is defined as the set of elements that belong to fuzzy set A at least to degree , and this degree is also called the degree of confidence or the degree of plausibility. The -cut level can be described as:

A = {x A ( x ) }
(5.1) where A is the -cut level of A and consists of all components of the universe set X whose membership grade is greater than . The support of fuzzy set A is defined by a classical set as sup p( A) = {x A (x ) > 0}. The fuzzy vertex (FV) method was first proposed by Dong and Shah (1987). The basic idea is to divide input membership domain into a series of equally spaced -cuts; for each -cut, the upper and lower points, as well as the interior points of each fuzzy variable, are selected. In a simulation process, with different combinations of these points being used as inputs, the output membership domain can be given by the maximum and minimum values of the modeling results at each -cut level.
125

Dong and Shah (1987) noted that the maximum or minimum value of the response function might not occur at the bounds of the interval. The extremum of the function within the interval must be included or an incorrect result will be obtained. In the FV method, all possible combinations of the interior points from all the -cuts are evaluated. However, the number of evaluations functions will increase with the numbers of both -cuts and variables. The total number of functions that need to be evaluated for n variables and m -cuts can be estimated based on the following equation:
p = (2 j 1)
j =1 m n

(5.2)

Obviously, the number of functions will become unmanageable for problems with large numbers of uncertain variables and/or -cuts. Meanwhile, if insufficient number of

-cuts is used, the interior extremum might be missed. Therefore, a method that can
reduce the number of functions but still ensure the solution containing the extremum within the bounds of the fuzzy variables is desirable. Based on the concept of fuzzy transformation as proposed by Hanss (2002), the interior extremum can also be obtained based on a more efficient selection of internal points. In order to address this concept, the following elements are considered: (a) a set of

m -cuts; (b) a set of n independent fuzzy-valued input parameters ( x i ) with the


membership function j ( xi ) at the j th -cut level, where i = 1, 2, , n and j = 1, 2, , m , and (c) evaluation functions of y = f ( x1 , x 2 ,
, xn ) .

In a fuzzy-parameterized simulation model, each fuzzy parameter (xi) is first divided into m -cuts, leading to the following set:

Pi = {xi(1) , xi(2) , ..., xi( m ) }


126

(5.3) where xi( j ) (j = 1, 2, , m) represents a fuzzy interval at the jth -cut level. It can be given as:

X i( j ) = [ai( j ) , bi( j ) ]

(5.4)

where ai( j ) and bi( j ) are the lower and upper bounds of input values at the jth -cut level, respectively. For j = 1, ai(1) = bi(1) . Under this condition, X i(1) converges into a single point (peak value of parameter i). Obviously, -axis is subdivided into m - 1 segments, with a distance of = 1/(m 1) . The membership grade (j) at -cut j can be given by:

j =

j 1 , m 1

j = 1, 2, , m.

(5.5)

For the traditional FV method, the input vector

cij (i = 1, 2, , n; j = 1, 2, ,

m) consisted of interior points at the jth -cut level and the ith input variable can be given by:

cij = (Ti1 , Ti 2 ,...., Ti k ,..., Ti 2 j1 ) , (k = 1, 2, , 2j - 1)

(5.6)

where Ti k is the kth element of cij , representing the value for the kth interior point. The

Ti k is defined by the following equations:


ai( j k +1) Ti k = ai(1) or bi(1) b ( k j +1) i 1 k j 1 k= j j +1 k 2 j 1 ,

(5.7)

For the modified selection scheme, the input vector cij (i = 1, 2, , n; j = 1, 2, , m) consisting of interior points at the jth -cut level and ith input variable can be given by:

127

cij = ( Ii1, j , Ii2, j ,...., Iir, j ,..., Ii ,j j ) , (r = 1, 2, , j)


where I ir, j is the rth element of cij , defined by the following equations:
ai( j ) = ( I ir, 11 + I ir, j 1 ) / 2 j bi( j ) r =1 2 r j 1 r= j ,

(5.8)

r i, j

(5.9)

Such a modified approach can be denoted as the Fractional Fuzzy Vertex (FFV) method. Figure 5.1 shows the selection schemes of interior points in both FV and FFV methods. Based on Equation (5.9), the number of interior points at the jth -cut level is j in the FFV method. Thus the number of evaluation functions at the jth -cut level becomes j n in the FFV method. Checking with Equation (5.2), the calculation amount has been greatly reduced. With a reduced number of interior-point combinations, the FFV method has an improved applicability. Figure 5.2 shows a general framework of the FFV method. The detailed algorithm can be described as follows: a) Build fuzzy sets A1, A2, , An for uncertain inputs X1, X2, , Xn, and denote an element of Ai by x i , where i =1, 2, , n , and n is the total number of uncertain inputs. The output ( y ) can be related to the inputs ( x1 , x2 , , x n ) through the simulation model [i.e., mapping y = f ( x1 , x 2 ,
, x n ) ].

b) Discretize the range of membership grade [0, 1] into a finite number of -cuts ( 1 , 2 , ...., m ), which are evenly distributed. c) Arrange combinations of the m -cut levels for n input parameters. The interior points on fuzzy intervals at each -cut level can be defined based on Equations (5.8) and (5.9).
128

d) Run the simulation model for each of the combinations, and obtain all possible values for the output y, which are denoted as y k ( k = 1, 2, , p ).

j 1

j 1 1

j-1 j
Ti k

j-1 2j-1 j

I i1, j 1

I i2, j 1

j-1
bi j

aij
I i2, j

2m-1 xi

m xi

FV method

FFV method

Figure 5.1 Selection schemes of interior points in FV and FFV methods

129

Build fuzzy sets (A1, A2, , An) for uncertain inputs (X1, X2, , Xn)

Select membership grades (1, 2, , m) based m -cut levels

( ai1 , bi1 ) for Ai at 1

( aij , bi j ) for Ai at j

( aim , bim ) for Ai at m


cim for Ai at m

c1 for Ai at 1 i
Combinations of c1 i (Total number = 1n)

cij for Ai at j
Combinations of cij (Total number = jn)

Combinations of cim (Total number = mn)

Simulation Output y1(k), k = 1n Output y1= [min (y1), max (y1)] at membership grade = 1

Simulation

Simulation

Output yj(k), k = 1,2,, jn Output yj(k), k = 1,2,, mn Output Output ym= [min (ym), max (ym)] yj= [min (yj), max (yj)] at membership grade = m at membership grade = j

Generation of fuzzy outputs

Figure 5.2 Framework of the proposed FFV method

130

e) Analyze the results ( y k ) regarding the same output membership value based on the same -cut levels ( 1 , 2 , ...., m ) of the input parameters. The upper and lower bounds of the outputs at each -cut level can be given by the maximum and minimum values of yq ( q = 1, 2, , j ) (which were obtained in step (d) under the same -cut level; j is the number of intervals and boundary points at the j th -cut level). f) The final solution to the fuzzy set B can be obtained based on the analysis of the upper and lower bounds of the outputs at different -cut levels. Generally, evaluation of the influences is critical for quantifying the proportion to which the uncertainty of single parameter contributes to the overall uncertainties of the modeling output (Hanss, 2002). If the modeling function ( f ) is available in analytical form, its total differential (df) can be easily determined and be used to compute the degrees of influence from each fuzzy input parameter (Hanss, 2002). Let xi ( i = 1, 2, .., n) be one of the fuzzy inputs, with its peak being at x i . Based on the differential calculus, the total differential at point p = ( x1 , x 2 , described as:
, x n ) can be

df =
i =1

f ( p )dxi xi

(5.10)

where

df is the total differential, which can be interpreted as an approximation of the when the inputs x i are changed by dxi around

overall change rate of function f ( p )

x i . Each input change ( dxi ) individually contributes ( df i ) to the overall change rate ( df ).

If the change rate ( dxi ) of the i th input is assumed to be a constant percentage ( c ) of its
131

corresponding peak, then dxi can be replaced by cxi . Thus Equation (5.10) can be given as:
n f ( p ) df = c xi xi i =1

(5.11)

Based on Equation (5.11), normalized change rate ( i ) for each input parameter can be defined as follows:

i =

xi f / xi
n i =1 i

x= p

[x f / x ]

(5.12)

i x= p

where i is interpreted as the degree of influence for input parameter i , satisfying the following condition:

i =1

=1

(5.13)

The value of i can be used for quantifying the influence of the i th parameter ( x i ) on the overall variation ( df ) of the modeling outputs [ f ( x1 , x 2 ,
, x n ) ].

5.3. Effects of Control-Factor Uncertainties

A reactor composting system is considered to examine the effects of uncertain control factors. Assume that the composting materials were pre-treated. The initial concentrations of soluble substrate (Ss), insoluble substrate (Si), biomass (X) and inert material (I0) were set as 2500, 3500, 2 and 500 molm-3, respectively. Based on the developed model and FFV techniques, impacts of uncertainties that exist in multiple control factors were analyzed. Such uncertainties exist in system temperature, moisture and oxygen content, which could be represented by symmetric
132

fuzzy membership functions. Assume that the composting process could be divided into three distinct stages, and the various control conditions could be set in different stages. Thus, five cut levels (0, 0.25, 0.5, 0.75, and 1) at each time point (the 24th, 72nd and 144th hour) were examined for each fuzzy input. Based on Haug (1993, 1996) and Hall (1998), the system temperature was set as 31 to 39C at the 24th hour, 51 to 59C at the 72nd hour, and 26 to 34C at the 144th hour; the moisture was 61 to 69% at the 24th hour, 51 to 59% at the 72nd hour, and 31 to 39% at the 144th hour; the oxygen content was 0.08 to 0.12 mol O2.m-3 at the 24th hour, 0.13 to 0.17 mol O2.m-3 at the 72nd hour, and 0.03 to 0.07 mol O2.m-3 at the 144th hour. The resulting fuzzy membership functions are presented in Figures 5.3 to 5.5. After the membership functions were determined, we could run the simulation model based on the FFV approach. Consequently, fuzzy membership functions of the four outputs (i.e. the concentrations of soluble substrate, insoluble substrate, and active biomass, and OUA) at three time points (i.e. the 24th, 72nd, and 144th hour) could be obtained (Figures 5.6 to 5.16). As shown in Figures 5.6 to 5.16, the outputs under different -cut levels at any given time point could be generated based on the given fuzzy membership functions. For example, at the 72nd hour (Figure 5.7), the soluble substrate concentration varied from 616 to 937 molm-3 under = 0.25, and from 693 to 802 molm-3 under = 0.75; it converged to 744 molm-3 when = 1. These facts demonstrated that the influences of input uncertainties could be effectively reflected in the fuzzy outputs as generated through the proposed FFV technique. The outputs under = 1 were deterministic when the uncertainties are simplified into mean values.

133

Membership function of temperature (at the 24th hr) 1 0.8 0.6 0.4 0.2 0 30 32 34 C 36 38 40 -cut level

Membership function of temperature (at the 72nd hr) 1 0.8 0.6 0.4 0.2 0 50 52 54 C
Membership function of temperature (at the 144th hr) 1 0.8 0.6 0.4 0.2 0 25 27 29 C 31 33 35 -cut level

-cut level

56

58

60

Figure 5.3 Membership functions of temperature

134

Membership function of moisture (at the 24th hr) 1 0.8 0.6 0.4 0.2 0 60 62 64 % 66 68 70 -cut level

Membership function of moisture (at the 72nd hr) 1 0.8 0.6 0.4 0.2 0 50 52 54 %
Membership function of moisture (at the 144th hr) 1 0.8 0.6 0.4 0.2 0 30 32 34 % 36 38 40

-cut level

56

58

60

-cut level

Figure 5.4 Membership functions of moisture


135

Membership function of oxygen content (at the 24th hr)

1 0.8 0.6 0.4 0.2 0 0.07

-cut level

0.08

0.09

0.1
-3

0.11

0.12

0.13

mol O2m

Membership function of oxygen content (at the 72nd hr) 1 0.8 0.6 0.4 0.2 0 0.12 -cut level

0.13

0.14

0.15
-3

0.16

0.17

0.18

mol O2m

Membership function of oxygen content (at the 144th hr) 1 0.8 0.6 0.4 0.2 0 0.03 -cut level

0.04

0.05

0.06
-3

0.07

0.08

mol O2 m

Figure 5.5 Membership functions of oxygen content


136

1 -cut level 0.8 0.6 0.4 0.2 0 2510 2512 2514 2516 2518 2520
-3

2522

2524

2526

2528

Ss (molm )

Figure 5.6 Fuzzy modeling output of soluble substrate concentration (the 24th hour)

137

1 -cut level 0.8 0.6 0.4 0.2 0 500 600 700 800
-3 Ss (molm )

900

1000

1100

Figure 5.7 Fuzzy modeling output of soluble substrate concentration (the 72nd hour)

138

1 -cut level 0.8 0.6 0.4 0.2 0 3450 3452 3454 Si (molm )
Figure 5.8 Fuzzy modeling output of insoluble substrate concentration (the 24th hour)
-3

3456

3458

139

1 -cut level 0.8 0.6 0.4 0.2 0 3900 3920 3940 3960 3980
-3

4000

4020

4040

Si (molm )

Figure 5.9 Fuzzy modeling output of insoluble substrate concentration (the 72nd hour)

140

1 -cut level 0.8 0.6 0.4 0.2 0 4575 4580 4585 4590 4595
-3

4600

4605

4610

Si (molm )

Figure 5.10 Fuzzy modeling output of insoluble substrate concentration (the 144th hour)

141

1 0.8 0.6 0.4 0.2 0 5 7 9 X (molm )


-3

-cut level

11

13

15

Figure 5.11 Fuzzy modeling output of active biomass concentration (the 24th hour)

142

1 -cut level 0.8 0.6 0.4 0.2 0 250 270 290 310 X (molm )
Figure 5.12 Fuzzy modeling output of active biomass concentration (the 72nd hour)
-3

330

350

370

143

1 0.8 0.6 0.4 0.2 0 8 9 10


-3

-cut level

11

12

13

X (molm )
Figure 5.13 Fuzzy modeling output of active biomass concentration (the 144th hour)

144

-cut level

1 0.8 0.6 0.4 0.2 0 13 18 23 OUA (mol O2m )


Figure 5.14 Fuzzy modeling output of OUA (the24th hour)
-3

28

145

1 -cut level 0.8 0.6 0.4 0.2 0 800 850 900 950
-3

1000

1050

OUA (mol O2m )


Figure 5.15 Fuzzy modeling output of OUA (the 72nd hour)

146

-cut level

1 0.8 0.6 0.4 0.2 0 1380 1385 1390 1395 1400 1405 1410 1415 1420 OUA (mol O2 m )
-3

Figure 5.16 Fuzzy modeling output of OUA (the 144th hour)

147

To quantify the effects of uncertain control factors, the degrees of influence from the input parameters on the modeling outputs could be obtained based on the above derivations (Equations 5.12 and 5.13). As the modelling outputs changed, the degrees of influence on the modeling results would vary dynamically. Figures 5.17 to 5.19 show the results of influences from temperature, oxygen content and moisture on the modeling outputs at the 24th, 72nd, 144th hr. It is indicated that the degrees of influence had significant deviations among different stages. As shown in Figure 5.17, temperature had a significant effect on active biomass concentration (54.9%) at the 24th hr, which was followed by OUA (41.7%). In comparison, at the 72nd hr, the effect of temperature on the soluble substrate concentration was 55.2%, followed with active biomass concentration and OUA (21.7% and 19.6%, respectively). At the 144th hour, the effect of temperature on the OUA was 92.9%, while those on the others were insignificant. Figure 5.18 shows the results for the influences of oxygen concentration at the 24th, 72nd, and 144th hr. Similarly, the oxygen concentration had varied effects at different stages. At the 24th hour, the oxygen concentration had a significant effect on the active biomass concentration (55.8%), followed by OUA (41.3%). At the 72nd hr, the effect of oxygen concentration on the soluble substrate concentration was 53.7%, followed by the biomass concentration and OUA (25.9 and 19.0%, respectively). At the 144th hour, the effect of oxygen concentration on the OUA became 64.9% and that on active biomass concentration was 22.8%, while the effect on the insoluble substrate concentration was negligible.

148

100% 90% 80% 70% 60% 50% 40% 30% 20% 10% 0%
1 2 3

Si Ss X OUA

Stage

Figure 5.17 Effect of temperature on modeling outputs Note: Stages 1, 2 and 3 stand for the 24th, 72nd, and 144th hour.

149

70% 60% 50% 40% 30% 20% 10% 0% 1 2 Stage 3

Si Ss X OUA

Figure 5.18 Effect of oxygen content on modeling outputs Note: Stages 1, 2 and 3 stand for the 24th, 72nd, and 144th hour.

150

70% 60% 50% 40% 30% 20% 10% 0% 1 2 Stage 3

Si Ss X OUA

Figure 5.19 Effect of moisture on modeling outputs Note: Stages 1, 2 and 3 stand for the 24th, 72nd, and 144th hour.

151

Figure 5.19 shows the influences of moisture at the 24th, 72nd and 144th hours. At the 24th hr, the effect on the active biomass concentration was 56.3%, followed by that on OUA (41.7%). At the 72nd hr, the effect on the soluble substrate concentration was 52.2%, followed by those on active biomass concentration and OUA (25.2 and 19.3%, respectively). At the 144th hour, moisture had a significant effect on OUA (60.9%), while those on the concentrations of active biomass and insoluble substrate were 20.5 and 18.5%, respectively. The results of Figures 5.17 to 5.19 demonstrate that the proportion to which each single uncertain parameter contributes (to the overall uncertainties of the modeling outputs) can be quantified through a series of sensitivity analyses. As a result, sensitive parameters that may lead to prediction errors can be identified. Practical arrangements can be made to efficiently reduce the uncertainties by targeting at parameters that are of high effects.

5.4. Effects of Input-Parameter Uncertainties

Another type of uncertainties was from the modeling inputs. In composting processes, variations in the composition of composting substrates usually exist, leading to uncertainties in the related modeling inputs. To determine the effect of such uncertainties, these related parameters can be considered as intervals. Thus, the effects of the interval parameters on the modeling outputs can be analyzed through interval analysis. Suppose there are n interval parameters, denoted by x i , xi X , i = 1, 2, ..., n . For each x , it can be defined by xi = [ a i , bi ] , where a i is the low bound of interval and bi is the upper bound of interval. If take one bound (low or upper bounds) from
152

each interval parameter, different combinations can be obtained, which can be denoted by

C j , j = 1, 2, ..., m . Thus, there will be 2n combinations (i.e., C1 = [ a1 , a 2 ,


C 2 = [b1 , a 2 , , a n ] , , C m = [b1 , b2 , , bn ] ).

, an ] ,

To study the effects of interval parameters on the composting process, the initial interval concentrations of soluble substrate (Ss), insoluble substrate (Si), active biomass (X) and inert material (I0) were set as [2400 2500], [3400 3500], [1 and [400 2], 500] molm-3. As a result, there were 16 combinations. For each combination, the control factors were assumed to be stable. Through running the developed simulation model, the results for all combinations could be obtained. Then maximize and minimize the results, the interval solutions can be acquired (Table 5.1). At the 24th hr, the interval widths for the concentrations of soluble substrate, insoluble substrate, and active biomass and OUA were 81.2, 105.9, 4.485, and 10.276, respectively. Compared with the initial interval widths, only minor changes occurred, indicating that the initial intervals of inputs had insignificant effects on the modeling outputs at the initial stage. At the 72nd hr, the interval widths for the concentrations of soluble substrate, insoluble substrate, and active biomass, and OUA became as 423.7, 315.9, 57.19, and 253.4, respectively. At this stage, the composting reactions were intensive due to the high microbial activities. As a result, the interval widths were enlarged greatly. For example, the interval width of soluble substrate concentration (Ss) changed from 81.2 to 423.7, and the insoluble substrate concentration (Si) was from 105.9 to 315.9. Thus, the effects of interval inputs would be intensified when the composting process entered active stage.

153

Table 5.1 Variations in interval widths


Parameter Ss (molm ) Si (molm-3) X (molm- ) OUA (mol O2m-3)
3 -3

Initial interval width 100 100 1 -

The 24th hour 2439 3347 4.485 10.276 + 2520 3453 8.691 20.073

Interval width 81.2 105.9 4.205 9.797

The 72nd hour 743.8 3673 262.71 702.304 + 1168 3989 319.90 955.74

Interval width 423.7 315.9 57.190 253.4 -

The 144th hour 4591 10.3136 1402 4444 10.015 1350 +

Interval width 147.3 0.299 51.5

154

At the 144th hr, the interval widths for the concentrations of insoluble substrate and active biomass, and OUA became as 147.3, 0.299, and 51.5, respectively. At this stage, the soluble substrate concentration was assumed to be zero. As shown in Table 5.1, the interval widths reduced significantly compared with the conditions in the active stage, indicating that the effects of intervals became negligible when composting process entered the curing stage. The above results showed that the interval inputs had some minor effects at the initial composting stage. However, when the composting process reached the active stage, the effects became significant; these effects got decreased at the curing stage.
5.5. Combined Effects of Uncertainties

To study the combined effects of uncertainties, the same initial conditions as those for interval analysis were considered. Similarly as before, different combinations for those interval parameters could be set. For each combination, the control factors were set as fuzzy variables. Based on the FFV method, run the composting model for each combination, the corresponding results related to the different combinations could be obtained. Then, maximize and minimize these results under different -cut levels, and the results were presented in Figures 5.20 to 5.30. The resulting intervals under different -cut levels at any time point can be obtained from Figures 5.20 to 5.30. For example, at the 72nd hr (Figure 5.21), the soluble substrate concentration varied from 515 to 1407 molm-3 under = 0.25, from 593 to 1313 molm-3 under = 0.75, and from 644 to 1275 molm-3 when = 1. Through analyzing the resulting -cut curves and the associated intervals, the effects of the uncertain input parameters and control factors could be evaluated.
155

1 0.8 0.6 0.4 0.2 0 2400

-cut level

2420

2440

2460

2480
-3

2500

2520

2540

Ss (molm )
Figure 5.20 Modeling output of soluble substrate concentration (the 24th hour)

156

1 0.8 0.6 0.4 0.2 0 450 550 650 750 850 950 1050 1150 1250 1350 1450
-3

-cut level

Ss (molm )

Figure 5.21 Modeling output of soluble substrate concentration (the 72nd hour)

157

1 -cut level 0.8 0.6 0.4 0.2 0 3340 3360 3380 3400 Si (molm )
Figure 5.22 Modeling output of insoluble substrate concentration (the 24th hour)
-3

3420

3440

3460

158

1 -cut level 0.8 0.6 0.4 0.2 0 3600 3650 3700 3750 3800 3850
-3

3900

3950

4000

4050

Si(molm )

Figure 5.23 Modeling output of insoluble substrate concentration (the 72nd hour)

159

1 -cut level 0.8 0.6 0.4 0.2 0 4400 4450 4500 4550
-3

4600

4650

Si (molm )

Figure 5.24 Modeling output of insoluble substrate concentration (the 144th hour)

160

1 0.8 0.6 0.4 0.2 0 2 4 6 8 X (molm )


-3

-cut level

10

12

14

Figure 5.25 Modeling output of active biomass concentration (the 24th hour)

161

1 -cut level 0.8 0.6 0.4 0.2 0 200 250 300 X (molm )
Figure 5.26 Modeling output of active biomass concentration (the 72nd hour)
-3

350

400

162

1 0.8 0.6 0.4 0.2 0 8 9 10 X (molm )


-3

-cut level

11

12

13

Figure 5.27 Modeling output of active biomass concentration (the 144th hour)

163

1 -cut level 0.8 0.6 0.4 0.2 0 5 10 15 20


-3

25

30

OUA (mol O2 m )

Figure 5.28 Modeling output of OUA (the 24th hour)

164

1 -cut level 0.8 0.6 0.4 0.2 0 600 700 800 900
-3

1000

1100

OUA (mol O2 m )

Figure 5.29 Modeling output of OUA (the 72nd hour)

165

1 -cut level 0.8 0.6 0.4 0.2 0 1320 1340 1360 1380 1400
-3

1420

1440

OUA (mol O2 m )

Figure 5.30 Modeling output of OUA (the 144th hour)

166

167

In Figures 5.20 to 5.30, the flat peak intervals present the effects of interval inputs; while the difference between the base interval and the peak interval denotes the effects of fuzzy control factors. Therefore, the percentage of peak-interval width to the base-interval width can be analyzed to quantify the effects from the uncertainties of the fuzzy control factors (Table 5.2). As shown in Table 5.2, at the 24th hr, the effect of fuzzy control factors was 7.5% for the insoluble substrate concentration, and 38.1% for the soluble substrate concentration. For the active biomass and OUA, the effects from the fuzzy control factors were 58.9 and 51.5%, respectively. These results indicated that, at the initial stage, uncertainties in the control factors had significant effects on the active biomass concentration and OUA, but only minor effects on the soluble and insoluble substrate concentrations. At the 72nd hour, uncertainties in the control factors had the most significant effect on the active biomass concentration (58.1%), followed by the soluble substrate concentration (56.3%); for the insoluble substrate concentration, the effect was 21.5%, and for OUA was 40.7%. These results indicated that the fuzzy control factors had the significant effects on the concentrations of active biomass and soluble substrate in the active stage. At the 144th hr, when the composting process reached the curing stage, the effects from uncertainties in the fuzzy control factors were 12.3% for the insoluble substrate concentration, and 92.6% for the active biomass concentration. However, the differences between base-interval width and peak interval width were very little for active biomass (only 0.299 and 4.031, respectively), indicating that microbial activities were very weak at this stage.
168

Table 5.2 Comparison of interval-width variations


The 24th hr
parameters Peak interval + Interval width
Difference in interval width (DIW)

The 72nd hr
DIW/BI % Interval size Difference in interval width (DIW) DIW/BI %

The 144th hr
Interval size Difference in interval width (DIW) DIW/BI %

3347

3453

105.9 8.6 7.5

3673

3989

315.9 86.4 21.5

4444

4591

147.3 20.6 12.3

Si

Base interval (BI) Peak interval

3343

3457

114.5

3624

4027

402.3

4432.7

4600.6

167.9

2439

2520

81.2 49.9 38.1

743.8

1168

423.7 546.3 56.3

Ss

Base interval (BI) Peak interval Base interval Peak interval Base interval (BI)

2410

2541

131.1

491.4

1460

969.99

4.485 2.745 10.28

8.691 12.986 20.073

4.205 6.036 10.24 9.797 10.392 51.5 58.9

262.71 219.6 702.3

319.9 356.2 955.7

57.190 78.8 135.59 253.4 174.0 40.7 58.1

10.02 8.216 1350

10.314 12.248 1402

0.299 3.732 4.031 51.5 35.5 40.8 92.6

OUA

7.135

27.323

20.18

607.6

1035

427.33

1332.9

1419.9

87

169

5.6. Discussion

The above results showed that uncertainties in input parameters and control factors have various effects on the modeling outputs in different stages. These results would be useful for guiding further process control. For example, temperature has significant effects on the soluble substrate and active biomass concentrations at the active composting stage (Figure 5.17), which means that any minor change in temperature would greatly affect the composting process; such effects should be considered in model-based composting process studies. Therefore, it is critical to mitigate the uncertainties in temperature when controlling the operation of a composting process. Similarly, if a modeling output is sensitive to inputs presented as intervals, a wide interval will be generated. This will lead to difficulties in establishing a model-based process control system. For example, the active biomass concentration has significant changes in the resulting interval widths at the active stage. If the intervals are wide (e.g. 200 to 300), difficulties will arise in determining the related control actions since the optimal control conditions can vary with active biomass concentrations. In this study, two kinds of uncertainties were studied. The results show that the uncertainties of interval input parameters and fuzzy control factors could be effectively handled through the proposed FFV method combined with interval analysis.

5.7. Summary

Based on the developed composting-process model, the effects of uncertainties on the composting process have been characterized through a fractional fuzzy vertex method combined with interval analysis. The effects of uncertain control factors and input
170

parameters as well as their combined effects have been studied. In addition, the effects of uncertain control factors have been quantified through a series of sensitivity analyses. The results showed that the uncertainties caused by interval input parameters and fuzzy control factors had various effects in different stages. They were useful for guiding practical process operation and control.

171

CHAPTER 6 MNPC: MODEL-BASED REAL-TIME NEURAL PREDICTIVE CONTROL FOR WASTE COMPOSTING PROCESSES

6.1. Statement of Problems

Process control is crucial for the management and operation of composting systems (Christensen and Carlsb, 2002). Effective process control can improve compost quality and enhance process efficiency (Hall, 1998). Previously, a number of studies on composting processes and possible control methodologies were reported in Jeris and Regan (1973a,b,c), MacGregor et al. (1981), Kishimoto et al. (1987), Keener et al. (1992), Robinson and Stentiford (1993), Kumar (1993), Haug (1993, 1996), Das and Keener (1996), Stentiford (1996), VanderGheynst et al. (1996), Oppenheimer (1997), and Patni and Kinsman (1997). An overall description of both theoretical and practical composting processes was given by Hall (1998). Understanding composting processes via modeling and experimentation can aid in improving process control (VanderCheynst, 1997; Nakaya, 1999). Ekinci (2001) gave a recent update on process control research for composting systems, including experimental, modeling and field-scale studies. More recent work related to composting process control included Nobuyuki et al. (2001), Nelson, et al. (2003), Ryckeboer and Mergaert (2003), and Tirumalasetty (2004). However, previous studies of composting process control mainly focused on system dynamics and empirical control, and there has been little research in relation to real-time process control. In fact, a composting system is extremely complex and dynamic, changing over time as microbial populations grow, evolve, die, and/or are replaced by new populations (Cundiff and Mankin, 2003). Moreover, uncertainties exist in a variety of system components, forming
172

an additional category of complexities. These facts form a barrier in applying traditional control theories to composting systems. As a result, real-time process control for composting systems has not been achieved. Traditional model predictive control (MPC) system enhanced by feed-forward neural networks is called neural predictive control (NPC) (Song and Kiovo, 1998). The NPC has been widely applied to process control through incorporations with non-linear optimization techniques such as manipulator control (Sanner and Slotine, 1995; Song and Kiovo, 1998), chemical process control (Zamarreno and Vega, 1999), mobile robot path tracking (Ortega and Camacho, 1994; Gu and Hu, 2002), and other methdologies (Tay and Zhang, 1999, 2000; Sohn et al., 2000, Wang and Gao, 2000; Kolehmainen et al., 2001; Kulkarni, 2004). However, application of such a technology to composting research has never been attempted. Thus, neural network predictive control technology will be applied to composting process as a new attempt in this research. The objective of this section, therefore, is to develop a model-based real-time neural predictive control system (MNPC) based on the developed composting-process model (Chapter 3). The system will include a neural network composting process model (i.e. NN emulator), a neural network controller (i.e. NN controller), an optimization process to identify desired control actions, and a composting process to be controlled. A series of composting experiments will be conducted for training and justifying the developed MNPC.

6.2. Neural Predictive Control for Composting Processes

NPC can be considered as a basic type of model-based predictive control in which

173

the prediction model is a trained neural network. The basic system structure can be seen in Figure 6.1. It usually consists of three components: the process to be controlled; a neural network that models the process; an optimization process that determines the optimal control action for the process (Soloway and Haley, 1997). The NPC normally optimizes the reaction response over a specified time horizon (Hunt, 1992). However, composting process is highly dynamic and complex, which makes the general NPC system inapplicable. Therefore, it is necessary to design a suitable NPC process control system for the composting process. Before a NPC system is set up, the control factors should be firstly identified. For a composting process, a number of impact factors exist, such as pH, C/N, and porosity (Zhang, 2000; Xi, 2002). However, many studies showed that temperature, moisture and oxygen content are the most important control factors (Hall, 1998). Therefore, they are considered as the key control factors in this study.

6.2.1. Design of Neural Predictive Control for Composting Processes

To acquire desired compost product, the conditions of control factors should be kept optimal. To achieve this goal, a composting NPC system was designed as shown in Figure 6.2. As shown in Figure 6.2, the system contains a NN emulator, a NN controller, an optimization process to identify desired control actions, and a composting process to be controlled. The optimal control output at each running cycle (obtained from the optimization block) can be used for the composting process.

174

Optimization Loop
r (t ) y n (t )

Optimization

u ' (t )

NN M odel of the Reaction Process

u * (t )

Reactor
y (t )

Figure 6.1 Neural predictive control system

175

Optimization Loop

R(t +1)
Model

U ' (t )

X p ( t + 1)

Optimization

NN Emulator

X ' (t + 1)

U * (t )

X (t)
NN Controller
U (t )

Composting Process

X (t + 1)

Figure 6.2 NPC design for waste composting

176

In Figure 6.2, X (t ) consists of the composting substrate constitutes and OUA;

R(t + 1) is the reference value from the composting model; X p (t + 1) is the output of
NN emulator; U (t ) is the control action coming from the NN controller; U (t ) is the optimal control action after the optimization process; U ' (t ) is the tentative control signal. The reference value comes from the developed composting-process (Chapter 3), which can be expressed as follows:

X (t + 1) = f [X (t ),U (t ), P]

(6.1)

where P is the constant parameters in the model; X (t ) , U (t ) , and X (t + 1) are the same as before. Based on this model, the state of composting substrate can be determined if the conditions of the previous substrate situation and control actions are known, i.e., the model can predict the changes of substrate components over time and thus reflect the reaction dynamics. As demonstrated in Figure 6.2, the NN emulator and NN controller are the main components of the developed system. They will be introduced in detail in the following sections.

6.2.2. Neural Network Modeling

Over the past decades, a number of neural network models have been developed, with the back propagation one (i.e. BP network) being the most popular. Thus, a BP network will be used in this study. Generally, operation of neural network is divided into

177

two stages: training and application. The first stage is mainly to determine the weight values, and the other one (also called working stage) (Hagan, et al. 1996) is to solve practical problems through using the neural network with the generated weight values. The structure of a typical BP network is shown in Figure 6.3. Suppose that there is a neural network with K+1 layers, and the numbers of nodes in every layer are n0-1 n1-1and nk-1 . For convenience, the bias k is incorporated in the weight vector of layer k as Wnkk = , and the first layer vector element xn0 = 1 . From layer 0 to layer 1, the original input vector ( X ), weight matrix ( W1 ), state vector ( Z1 ), output vector ( U1 ) and their relations could be listed as follows:
1 1 u1 z1 x1 1 1 x u2 z2 2 T 1 X = Z1 = = W1 X U1 = = f (Z1 ) W1 = ( wij ) n0 n1 1 1 xn0 un1 zn1

From layer k-1 to layer k, the weight matrix ( Wk ), state vector ( Z k ), output vector ( U k ) and their relations could be written as:

z1k u1k k k z2 = W T Y u2 = f ( Z ) k Wk = (wij ) nk 1nk Z k = k k k 1 U k = k k znk unk


where uik = f ( zik ) with f being a transfer function which is normally defined in a log-sigmoid style:

f (s) =

1 1 + exp ( s )
178

(6.2)

Output Variables

Output Layer

Hidden Layer

Input Layer

Input Variables Figure 6.3 Structural representation of the BP network

179

The back-propagation algorithm for multi-layer networks is a gradient descent optimization program where the objective function is to minimize the mean square error of performance index. The algorithm is provided with a set of proper network behavior examples: { X '1 , R1 }, { X '2 , R2 }, , { X 'Q , RQ } where X 'i (i = 1, 2, ..., Q ) is an input vector to the network, and Ri is the corresponding target output. As each input is applied to the network, the network output will be compared to the target. The network parameters should be adjusted in order to minimize the sum squared error:

F ( w) =

1 Q T 1 Q ei ei = ( Ri U 'i )T ( Ri U 'i ) 2 i=1 2 i=1

(6.3)

where w is a vector containing all of network weights and biases. The steepest descent algorithm for the approximate mean squared error could be written as:

Wk +1 = Wk + (

e ) |W =Wk W

(6.4)

where is the learning rate. In this study, the NN controller is expressed as the following general function:

U (t ) = f ( X (t ))
(6.5) The NN emulator is expressed as:

X (t + 1) = f ( X (t ),U (t ))

(6.6)

Suppose there is a BP network with one hidden layer containing h nodes and

ki input nodes; the composting substrate situation is denoted as X (t ) and the control
180

action is denoted as U (t ) . The dynamic function of BP network for NN emulator can be expressed as follows:

x (k ) = wk , j f j ( z ( j )) + bk
j =1

(6.7)

and

z ( j ) = w j ,ki xu (ki) + b j
ki =1

(6.8)

where x(k ) is the output at the k th node of the network [ x(k ) X (t ) ] ; z ( j ) is the activation level of the j th nodes output function; h is the node number in the hidden layer; bk is the bias on the output node; xu(ki ) is the XU at the ki th node of the network ( XU = [x1

x2

xk

u1 u2

ui ] ); ki = k + i ; p is the number of
T

input nodes; wk , j is the weight connecting the j th hidden node to the output node k ;

w j ,ki is the weight connecting the ki th input node to the j th hidden node; b j is the
bias on the j th hidden node. Similarly, the NN controller can be expressed as the following BP network:

u (i ) = wi , j f j ( z ' ( j )) + bi
j =1

h'

(6.9)

and

z ( j ) = w j ,k x ( k ) + b j
' k =1

p'

(6.10)

where u (i ) is the output at the i th node of the network [ u(i) U (t ) ]; f () is the output function for the j th node of the hidden layer; z ' ( j ) is the activation level of the j th nodes output function; h ' is the node number in the hidden layer; p ' is the number of
181

input nodes; wi , j is the weight connecting the j th hidden node to the output node i ;

w j ,k is the weight connecting the i th input node to the j th hidden node; x(k ) is the
input at the k th node of the network [ x(k ) X (t ) ]; b j is the bias on the j th hidden node; and bi is the bias on the output node. In this study, the control factors consist of temperature, moisture, and oxygen content; the input variables include the concentrations of soluble substrate, insoluble substrate and active biomass, as well as the OUA. Therefore, U (t ) includes three elements, and X (t ) includes four elements. As a result, the NN controller and emulator can be designed as shown in Figures 6.4 and 6.5. The input and output variables for the NN controller and emulator are listed in Table 6.1.

6.2.3. Nonlinear Optimization through Simulated Annealing

Generally, the optimization process is to minimize a cost function over a defined prediction horizon under a number of constraints. The general form is:

Min J = Min{ [ xn (t ), u (t ), t ]dt}


u (t ) u (t ) tk

tk + P

(6.11a)

Subject to:

Eq( xn (t ), u (t )) = 0 IEq( xn (t ), u (t )) 0

(6.11b)

where u (t ) is a set of control factors; xn (t ) contains the system outputs that are related to both control factors u (t ) and system states x (t ) ; Eq and IEq are the equality and inequality constraints, respectively; t k is the discrete time at point k ; P
182

is the prediction horizon.

183

x1 (t ) u 1 (t + 1)

x 2 (t )

u 2 (t + 1) x3 (t) u 3 (t + 1)

x 4 (t )

I n p u t la y e r w ith 4 n o d e s

H id d e n la y e r

O u tp u t la y e r w ith 3 n o d e s

Figure 6.4 The BP network architecture of the NN controller

184

u 1 ( t + 1) u 2 (t + 1) x1 (t + 1) u 3 (t + 1) x1 ( t ) x 2 (t ) x 3 (t ) x 4 (t ) x 2 (t + 1) x 3 (t + 1) x 4 (t + 1)

I n p u t la y e r w ith 7 n o d e s

H id d e n la y e r

O u tp u t la y e r w ith 4 n o d e s

Figure 6.5 The BP network architecture of the NN emulator

185

Table 6.1 Input and output variables for the NN controller and emulator
Input (I) or Output (O) I I I I O O O Input (I) or Output (O) I I I I I I I O O O O

NN controller Insoluble substrate Soluble substrate Active Biomass Oxygen Uptake Accumulation Temperature Moisture Oxygen content

Symbol Si Ss X OUA T M O2

NN emulator Insoluble substrate Soluble substrate Active biomass Oxygen Uptake Accumulation Temperature Moisture Oxygen content Insoluble substrate Soluble substrate Active Biomass Oxygen Uptake Accumulation

Symbol Si Ss X OUA T M O2 Si Ss X OUA

186

The objective function ( J ) could have different forms in terms of different control requirements. If the constraints are linear, the non-linear programming problems may be simplified into a quadratic programming (QP) formulation. The study problem for composting-process control can be formulated as:

J = [rn (t + i ) xn (t + i ) ] + m [u m (t + i ) ]
2 n =1 i = N1 m =1 i =1

N2

Nu

(6.12a)

subject to:
min max u m u m (t + i ) u m (i = 1, 2, ..., N ) min max (i = N1 , ..., N 2 ) xn xn (t + i) xn max (i = 1, 2, ..., N ) u m (t + i ) u m u m (t + i) = 0 (i > N 1)

(6.12b)

Where:
rn (t + i ) = the n th output from the reference model at time t + i ; x n (t + i ) = the n th predicted output from the NN emulator;
u m (t + i ) = the m th control increment at time t + i , which can be defined as

u m (t + i ) u m (t + i 1) ;
m = control input weight factor;
N1 = minimum output horizons; N 2 = maximum output horizons;
N = control horizon;

N = number of output variables;


M = number of control factors.

The minimization of the objective function ( J ) could be achieved through the


187

simulated annealing algorithm (SA). The SA was originally proposed by Metropolis et al. (1958) as a means of finding the equilibrium configuration of a collection of atoms under a given temperature. One of the significant advantages over the other traditional optimization methods lies in its ability to avoid becoming trapped at local minima. Due to its remarkable advantages, SA has drawn significant attention in many engineering applications (Dougherty and Marryott, 1991). The algorithm utilizes a random search which accepts not only variations that decrease the objective value, but also occasionally changes that increase it. To determine whether the latter change is acceptable or not, a Boltzmann probability distribution is used:
p ( Ei ) = exp( Ei / T )

(6.13)

where E i is a energy state representing the objective function value for the i th atom; T is the temperature; p ( E i ) is the probability of the system with energy E i . Then, an accepted probability for any energy change between the two atoms is:
p ( Ei ) = exp( E / T )

(6.14)

where E is the variation of energy state representing the change in the objective function. In an annealing process, the generation of optimal solution and the annealing mechanism are both critical. An alternative strategy suggested by Parks (1990) is to generate solutions according to the following formula:
X i +1 = X i +

(6.15)

where X i +1 is the (i + 1) th control vector satisfying all of the constraints, X i is the i th control vector, is a vector of random numbers in the range of (-1, 1) and is a step

188

factor which defines the maximum change allowed for each variable. This tunes the maximum step size associated with each control vector towards a value that gives acceptable changes. The annealing mechanism determines the degree of uphill movement permitted during the search, and thus is critical to the algorithm efficiency. For an annealing mechanism, four parameters should be specified, i.e. initial temperature (T0), final temperature (Tf), length of the Markov chains (L), and a rule for decrementing the temperature. The initial temperature should be high enough to melt the system completely and should be reduced towards its freezing point as the search progresses (Press et al., 1986). The final temperature should ensure the search progress can be halted when it ceases to make progress. The length of Markov chain is introduced to maintain a stationary, equilibrium state for an optimization system under each temperature. This length should be long enough to guarantee an optimal solution being found and short enough to ensure computational feasibility. The rule for decrementing the temperature represents how the initial temperature gradually transits to the final one. It should avoid occurrence of negative temperatures. The simplest and most common temperature decrement rule is:
Tk +1 = k aTk

(6.16)

where k a is annealing constant close to, but smaller than, 1. Based on Figure 6.2, the procedures for implementing the MNPC can be listed as follows: a) Define X (t ) and denote an element of X (t ) as x i , where i = 1, 2, , n , and n is the total number of variables characterizing the composting substrate
189

situation; X (t ) is the input of the controller. b) For the given X (t ) , the controller can produce control action U (t ) , and denote an element of U (t ) as u j , where j =1, 2, , m , and m is the total number of control variables. The control variables combining the substrate components, i.e. X (t ) + U (t ) , become an input for the emulator and the emulator can produce output X p (t + 1) . c) Also, X (t ) + U (t ) will be an input to the reference model and produce output

X r (t + 1) .
d)

X p (t + 1) will be compared with X r (t + 1) . If the objective is satisfied, control


action U (t ) will be considered the optimal; then it will be applied to the practical composting process. If not, the control action will be adjusted with a new input for the emulator; then the emulator will produce a new output to be optimized, till the optimal result, i.e. U (t ) , is achieved.
*

e)

After the optimal control action is applied to the practical composting process, the output will become the next input for the NN controller; then repeat the steps

a to d . Information of the implemented composting process could be used for


further training the NN controller and emulator.

6.3. Application 6.3.1. Experimental Studies

To train the MNPC, a series of composting experiments were undertaken. Eight bench-scale reactors were constructed by using a cylindrical PVC tube, with an effective
190

volume of approximately 30 liters (see Figure 3.7 in Chapter 3). The air-feed system was flowing upward through a fine mesh screen near the bottom of the reactor. The sensors were used for online detection of system temperature. The raw materials included apple, potato, rice, carrot, leaves, meat, sawdust, soy bean, soil and coal ash. They were pre-treated for moisture and size control. During the experimental process, conditions of temperature, moisture and airflow rate were controlled; about 30 g of solid waste samples at three locations were collected twice a day; the concentrations of soluble substrate, insoluble substrate, inert materials, and active biomass, and moisture were measured with standard methods (Eaton et al., 1998). An oxygen analyzer was used for detecting oxygen concentrations at the reactors inlet and outlet, such that the oxygen content and OUA can be quantified. Two batches of composting experiments were conducted (total 16 composting experiments; each has 16 sets of data), and a total of 256 sets of data were obtained. These data were used to train the NN emulator and controller. The ranges of the data used to train the neural networks and the initial experimental conditions are listed in Table 6.2. Then, a single composting experiment under controlled conditions was undertaken to justify the developed MNPC.

6.3.2. Training for the NN Emulator and Controller

The input and output parameters for the NN emulator and controller are already listed in Table 6.1. During training the emulator and controller, the weighting matrix and bias vector in each layer were adjusted to fit the NN outputs with the experimental results. A comparison of the network outputs and the experimental results was conducted through

191

analyzing the mean squared errors based on the steepest descent algorithm. The course of convergence was a continuous descent. The training process would not stop until a defined threshold value ( ) is achieved (stop if MSE lastrun MSE actualrun ). After the training process, the NPC system for composting process could then be established.

6.3.3. Results and Discussion

The results from a single controlled composting reactor were used for verifying the developed MNPC. The initial concentrations of soluble substrate (Ss), insoluble substrate (Si), active biomass (X) and inert material (I0) were 2503.5, 3410.9, 2.315, and 409.5 molm-3, respectively. During the composting experiment, the composting conditions were controlled based on the optimization results of MNPC. Figures 6.6 to 6.8 show the controlled results compared with the predicted results of MNPC. An error analysis was conducted as shown in Table 6.3. The correlation coefficient between the observed and predicted values ( R ) can be calculated by (Gontarski et al., 2000):
N N (x pi x p )( xoi xo ) i =1
2 2 N N N 2 N 2 N x pi x pi N xoi xoi i =1 i =1 i =1 i =1

R=

(6.17)

where N is the number of data points, x pi is the predicted value, xoi is the observed value, x p is average of the values predicted by the network, and x o is the average of observed values. The average relative error (ARE) is defined by:

192

Table 6.2 Ranges of data used to train the NN controller and emulator
Variable Unit Range Ss molm-3 0~2851 Si molm-3 2895~4543 X molm-3 1.58~362.99 I molm-3 418~589 OUA molm-3 0~2180 T C 22~71 M % 38~72 O2 molm-3 0.038~0.167

Initial conditions of different experiments 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 2161 2030 2226 2342 2467 2528 2648 2756 2848 2765 2654 2518 2498 2396 2265 2163 3728 3856 3652 3598 3475 3362 3215 3173 3026 2895 3124 3268 3395 3457 3692 3812 1.58 3.21 2.35 1.88 2.96 1.77 2.15 2.63 2.92 3.01 1.87 2.68 2.61 2.01 2.97 2.22 436 456 501 418 488 464 569 467 571 498 467 584 521 429 491 589 22 22 22 22 22 22 22 22 22 22 22 22 22 22 22 22 71 70 69 70 70 71 69 68.5 71 70 71.5 71 71 72 70 69 0.049 0.052 0.051 0.053 0.049 0.049 0.051 0.048 0.051 0.053 0.048 0.049 0.049 0.051 0.049 0.054

193

70 60 50 40 30 20 10 0 0 24 48 72 96 Time (hr)

Temperature (C)

Controlled Predicted 120 144 168 192

Figure 6.6 Comparison of the controlled and predicted results for temperature

194

80 Moisgure (%) 70 Controlled 60 50 40 30 0 24 48 72 96 120 144 168 192 Time (hr) Predicted

Figure 6.7 Comparison of the controlled and predicted results for moisture

195

0.18 0.15 0.12 0.09 0.06 0.03 0 0 24 48 72 96 120 144 Time (hr)

O2 (molm )

Controlled Predicted

-3

168

192

Figure 6.8 Comparison of the controlled and predicted results for oxygen content

196

Table 6.3 Error analysis for controlled experiment conditions Control Variable Temperature Moisture Oxygen content Average relative error (%) 3.23 2.82 8.28 Minimum relative error (%) 1.21 1.44 3.23 Maximum relative error (%) 8.71 7.63 30.51 Correlation coefficient 0.993 0.989 0.985

197

ARE = (
i =1

xoi x pi xoi

) / n 100%

(6.18)

The correlation coefficient represents the consistency of trends between the controlled and predicted results while the ARE characterizes the accuracy of controlled results. As mentioned in Chapter 4, the composting process can be divided into three stages: initial, active, and curing stage. Comparisons of the controlled and predicted results are shown in Figures 6.6 to 6.8: At the initial stage, the controlled values of temperature and moisture levels well matched with the predicted results of MNPC, while the oxygen content had a few variations. At the active stage, the controlled values of temperature and moisture still matched the predicted ones, but with increased errors. When entering the curing stage, all the controlled values well matched the predicted ones. If ARE and correlation coefficient were used to evaluate the accuracy and consistency between the controlled and predicted results, it could be seen that the AREs of Figures 6.6 and 6.7 were 3.23% and 2.82%, respectively; the correlation coefficients were 0.993 and 0.989, respectively. These results indicated that the observed and predicted values well matched each other in terms of both accuracy and consistency. In Figure 6.8, the controlled values of oxygen content had higher error levels with the ARE being 8.28% and the correlation coefficient being 0.985. The oxygen content was harder to be controlled and measured accurately; in comparison, temperature and moisture were relatively easy to be controlled. However, considering the dynamics and complexity of composting system, all of the controlled results were generally acceptable. Therefore, it can be concluded that the controlled results of composting conditions in Figures 6.6 to 6.8 were basically consistent with the predicted results of MNPC, and

198

could meet the requirements of system justification. Based on the controlled composting conditions, the predicted results of MNPC compared with the observed ones are shown in Figures 6.9 to 6.12, with the related error analysis being shown in Table 6.4. From Figures 6.9 to 6.12, it can be seen that the predicted results from the MNPC were basically consistent with the experimental ones. The AREs of Figures 6.9 and 6.10 were 9.24 and 2.52%, respectively; the correlation coefficients were 0.978 and 0.992, respectively. These results indicated that the predicted and observed results well matched in terms of accuracy and consistency. In Figures 6.11 and 6.12, the AREs were 24.47 and 22.56%, respectively; the correlation coefficients were 0.972 and 0.986, respectively. These demonstrated that a consistency existed between the predicted and observed results but the accuracy was not satisfactory. The main reason is that the concentration of biomass was sensitive to variations of environmental factors during the composting process, and thus was more difficult to be predicted; the OUA had the similar situation since it was directly related to the concentration of biomass. However, these results were acceptable for the MNPC system considering the dynamics and complexity of composting process. In fact, composting was a slow biological process, such that the concentrations of insoluble substrate and soluble substrate would not change significantly in a relative short period of time. In addition, both the predicted and observed results were based on the same initial conditions, i.e., each time point in the figures had the same initial conditions. Consequently, there were no accumulation errors in Figures 6.9 to 6.12. The main advantage of the developed MNPC was associated with neural networks learning ability. With such a learning ability and based on the objective function (6.12), the predicted results from the MNPC would approach those from the simulation model if

199

3000 Predicted Value (mol Ssm) 2500 2000 1500 1000 500 0 0 500 1000 1500 2000
-3

-3

2500

3000

Observed Value (mol Ssm )

Figure 6.9 Predicted and observed results for soluble substrate concentration

200

4600 4400 Predicted Value (mol Sim)


-3

4200 4000 3800 3600 3400 3400 3600 3800 4000 4200
-3

4400

4600

Observed Value (mol Sim )

Figure 6.10 Predicted and observed results for insoluble substrate concentration

201

350 300 Predicted Value (mol Xm )


-3

250 200 150 100 50 0 0 50 100 150 200 250


-3

300

350

Observed Value (mol Xm )

Figure 6.11 Predicted and observed results for active biomass concentration

202

1800 1600 Predicted Value (mol Om ) 2 1400 1200 1000 800 600 400 200 0 0 200 400 600 800 1000 1200 1400 1600 1800
-3 -3

Observed Value (mol O2 m )

Figure 6.12 Predicted and observed results for OUA

203

Table 6.4 Error analysis for the predicted results from MNPC Average relative error (%) 9.24 2.52 24.47 22.56 Minimum Maximum relative error relative (%) error (%) 1.73 0.34 2.24 1.36 93.42 6.63 133.51 140.65 Correlation coefficient 0.978 0.992 0.972 0.986

Variable Ss Si X OUA

204

a sufficient number of iterative training operations were conducted. However, the above results could hardly be considered ideal, since the reference data set was from the simulation model with a limited scope in reflecting the complete spectra of the impact factors under a variety of system conditions. To deal with such a situation, it is necessary to adjust the objective function after the MNPC runs for an extensive period of time. Thus, we have:

J = [OUA(t + i + 1) OUA(t + i )] + m [u m (t + i ) ]
2 i = N1 m =1 i =1

N2

Nu

(6.19a)

subject to:

(i = 1,..., N u ) u min u (t + i ) u max OUA OUA(t + i ) OUA (i = N1 ,..., N 2 ) min max (i = 1,..., N u ) u (t + i ) u max u (t + i ) = 0 (i > N 1) u

(6.19b)

where OUA is an increasing function of time, and can thus be considered as an indicator of composting-process progression (Hamelers, 2001). Therefore, the function (6.19) seeks to maximize the OUA increasing rate (i.e. maximize composting reaction rate) under minimized actions of controlling interference. Based on such an objective function, the MNPC would generate results with progressively improved accuracy along with the ongoing training process. This is also the main advantage of the MNPC compared with the other process control technologies.

6.4. Summary

In this section, a model-based real-time neural predictive process control system for waste composting processes has been developed. The system includes a neural

205

network emulator, a neural network controller, an optimization procedure, and a composting reactor. To train and justify the MNPC, a series of composting experiments have been undertaken. The data from two batches of composting experiments have been used to train the MNPC, and a single composting experiment under strictly controlled conditions has been used to justify the developed system. The testing results showed that the predicted values were basically consistent with the experimental ones, indicating that the developed MNPC was generally satisfactory. The advantage of the developed MNPC is that the neural networks have a learning ability, which makes the system different from the general MPC system. Through adjusting the objective function, the MNPC could generate results with progressively improved accuracy along with the on-going training process.

206

CHAPTER 7 SIPC: STEPWISE-INFERENCE-BASED REAL-TIME PREDICTIVE CONTROL FOR WASTE COMPOSTING PROCESSES

7.1. Statement of Problems

In chapter 6, a model-based real-time neural predictive control system (MNPC) has been developed, which capitalizes the advantages of both model predictive control and neural networks. The neural predictive control has become one of the most popular process control technologies and been widely applied to many research fields. The neural networks have learning ability and are capable of dealing with both linear and nonlinear systems (Guh, 2003). However, under many circumstances, information of the related assumptions and confidence levels are needed to justify the generated decision supports, while such information is unavailable in the MNPC. In addition, process control for composting systems involves discreteness and non-linearity, which can only be tackled through advanced technologies. The technique of stepwise cluster analysis could tackle the issues of the related assumptions and confidence levels as well as the difficulties of discreteness and nonlinearity (Huang, 1994; Huang, 2004); the genetic algorithm (GA) is capable of dealing with nonlinear optimization and has been widely applied to various research fields (Diamantis and Tsahalis, 2002; Walker et al., 2003). However, there has been no research attempt to relate the stepwise cluster analysis with the GA (for nonlinear optimization) in the field of process control. Therefore, development of a real-time predictive control system based on a hybrid of the stepwise cluster analysis and the GA based optimization would help generate a more robust process control system. As a
207

result, the objective of this section is to develop a real-time composting process control system based on such a hybrid. The developed system includes a stepwise-inference emulator (SI emulator), a stepwise-inference controller (SI controller), a GA-based optimization procedure, and a composting reactor. The specific tasks include: (a) Establishing a bridge between control actions and substrate characteristics through stepwise cluster analysis (i.e. SI controller). The controller could produce a control action based on any given substrate characteristics. (b) Establishing a statistical model for simulating the composting process through the stepwise cluster analysis (i.e. SI emulator). The emulator could predict the next substrate characteristics based on the given substrate characteristics and control action. (c) Optimizing the predicted results from the SI controller and emulator to identify the desired control action through GA-based optimization; the results would be applied to real-world composting process. (d) Initiating and justifying the developed stepwise-inference-based real-time predictive control system (SIPC) based on a series of composting experiments. (e) Comparing the developed SIPC with MNPC.

7.2. Research Framework

Figure 7.1 shows the framework of the stepwise-inference-based process control scheme. The architecture includes four main components: SI controller, SI emulator, optimization procedure, and composting reactor. The SI emulator is a statistical model for simulating the composting process. The SI controller is for establishing a bridge between

208

substrate situations and control actions. The optimization procedure is to identify desired control actions; the optimization results at each running cycle could be directly used for guiding the composting process. The composting reactor could be operated under controlled conditions. In Figure 7.1, X (t ) consists of the substrate constitutes and oxygen uptake accumulation (OUA); X p (t + 1) is the output of SI emulator; U (t ) is the control action generated from the SI controller; U (t ) is the optimal control action after the optimization procedure; U ' (t ) is the tentative control signal.

7.3. Stepwise Cluster Analysis

Considering the high complexities and dynamics of the composting system, it is inevitable that some important information could be missed or ignored during the process of establishing a composting-process model, because almost all models are selective abstractions of the reality. Under situations when a large number of experimental data exist, statistical relationships could be developed to substitute the general simplified models. For composting systems, many variables can be either continuous or discrete, and the relations among them can be either linear or nonlinear. Therefore, the conventional continuous and linear methods cannot effectively reflect such complexities. A stepwise cluster analysis method was advanced for tackling such discreteness and nonlinearity (Huang, 2004). In this study, the SI controller could be expressed with the following general function:

209

Optimization Loop
U ' (t )

X p ( t + 1)

Optimization

SI Emulator

X ' (t + 1)

U * (t )

X (t)
SI Controller
U (t )

Composting Reactor

X (t + 1)

Figure 7.1 Stepwise-inference-based process control

210

U (t ) = f ( X (t ))
(7.1) The SI emulator could be expressed as:

X (t + 1) = f ( X (t ),U (t ))

(7.2)

The functions of SI emulator and controller could be realized through stepwise cluster analysis based on a series of composting experiments. For the SI controller, the operating conditions to substrate characteristics were considered as dependent variables; the substrate characteristics were independent variables. For the SI emulator, the substrate characteristics to operating conditions and previous substrate characteristics were considered as dependent variables; the operating conditions and previous substrate situations were independent variables. Assume that there are n scenarios of system conditions (experimental data). There would then be n sets of such independent and dependent variables. Similarly, assume that there are m independent variables denoted as x = ( x1 , x 2 , dependent variables denoted as y = ( y1 , y 2 ,
, x m ) , and p

, y p ) . Thus, all data can be given by


, m , and i = 1, 2, , p.

matrices X = ( x tr ) nm and Y = ( yti ) n p , where, r = 1, 2,

7.3.1. Clustering Principles

In the stepwise cluster analysis, sample sets of dependent variables will be cut or merged into new sets, and values of independent variables will be used as references to judge into which new set a sample in the parent set will enter. After the completion of the cut and merge processes, cluster trees could be developed and further used for the

211

prediction of new dependent variables according to the new independent variables values. The essence of this method is, based on given criteria, to cut one sample set of dependent variables into two, and to merge two sets into one, step by step, in order to classify samples and sieve variables. Let cluster h , which contains n h samples, be cut into two sub-clusters e and f ( e and f contain ne and n f samples, respectively, i.e., ne + n f = nh ). According to Wilks likelihood-ratio criterion, if the cutting point is optimal, the value of Wilks ( = W / T ) should be the minimum (Wilks, 1960, 1962, 1963; Kennedy and Gentle, 1981), where T and W are the total-sample SSCP matrix

{t }
ij

and the within-group SSCP matrix

{ w },
ij

respectively; T and W mean the

determinants of matrixes

{t }
ij

and

{ w },
ij

respectively. When the value is very

large, clusters e and f cannot be cut, but must be merged into greater cluster h . By Raos F-approximation (R-Statistic), we have:

R=

1 1 / S Z S P ( K 1) / 2 + 1 P ( K 1) 1 / S

(7.3) (7.4) (7.5)

Z = nh 1 ( P + K ) / 2

S=

P 2 ( K 1) 2 4 P 2 + ( K 1) 2 5

where statistic R is distributed approximately as an F - variate with v1 = P ( K 1) and v2 = P ( K 1) / 2 + 1 degrees of freedom; K is the number of groups; P is the number of dependent variables. The R -statistics will reduce to an exact F -variate when P = 1 or 2 , or when K = 2 or 3 . Since the number of groups is two ( K = 2 ) in this study, an exact F -test is possible based on the Wilks criterion. Thus we have:
212

F ( P, nh P 1) =

1 nh P 1 P

(7.6)

Therefore, the criteria of cutting and merging clusters become to make a number of F tests (Rao, 1952, 1965; Tatsuoka, 1971).

7.3.2. Tests of Optimal Cutting Points

To determine the optimal cutting point, n h samples in cluster h are sequenced according to the values of x r( ,hk) in

{xr } ,

i.e. x r( ,h1)r x r( ,h2)r

xr( hn)r . Then the ,


h

total-sample SSCP matrix and within-group SSCP matrix of dependant variable y are calculated based on the following sequence statistic {k r }:
r bij (k r , nh ) = r r r nh k r { Bi( h ) (k r ) Bi( h ) (nh ) B (j h ) (k r ) B (j h ) (nh )

][
r

]}

n k
r h

(7.7) (7.8) (7.9)

r ( r r r r t ij (nh ) = Aijh ) (nh ) nh Bih (nh ) B h (nh ) j r r r wij (k r , nh ) = t ij (nh ) bij (k r , nh )

where
h Bi(or) j (u ) =

1 u (h) yi or j ,k u k =1
u

(7.10)

( Aijh ) (u ) = y i(,h ) y (jhk) , k k =1

k = 1r , 2 r ,
r

r , ( n h 1), r

(7.11)
,m

i, j = 1, 2,

, p, and r = 1, 2,
r

For each xr , a cutting point ( k * ) is derived, which satisfies:


r ( k , n ) = min { ( k r , nh )} r r
*r

r h

r ( nh 1)

k =1

(7.12)

213

For each independent variable, an index of independent variable which will be used for cutting judgments ( r * ) is derived, which satisfies:
r r ( k *r * , nh ) = min ( k r , nh ) r =1 m

}
r*

(7.13)

Thus, the optimal cutting point of cluster h is k * , and the relevant value of independent variable (which will be used as the reference for new sample prediction) is
x (*h ) *r * . Then the F-test can be undertaken. If
r ,k

r r 1 (k *r* , nh* ) nh* P' F ( P' , n P'1) = F1 r (k *r* , nh* ) P' r* h

(7.14)

is satisfied, cluster h can be cut into two sub-clusters according to the distribution of

xr* :
(a) data in the dependent-variable set with k r k * are allocated into sub-cluster
* r*

e ( < f ),
(b) data in the dependent-variable set with k r > k * are allocated into sub-cluster
* r*

f .
where P ' is the number of dependent variables under consideration. Among these independent variables, x r * is the most important one which significantly affects the values of dependent variables. Conversely, if Equation (7.14) is not satisfied, cluster h cannot be cut. Then the other clusters will be tested to decide whether to cut or not, i.e. to test the confidence of

h = 1, 2,

, H ( H is the total number of clusters at the current stage). When no cluster

can be cut, then the next step is to undertake the mergence of clusters.

214

7.3.3. Mergence of Clusters

To test the mergence of clusters e and f among the existing H clusters, the following total-sample SSCP matrix and within-group SSCP matrix should be calculated firstly:
( ( t ij ( ne , n f ) = Aije ) ( ne ) + Aij f ) ( n f ) ne Bi( e ) ( ne ) + n f Bi( f ) ( n f )

ne B (j e ) (ne ) + n f B (j f ) ( n f ) /( ne + n f )
bij (ne , n f ) = ne n f Bi( e ) ( ne ) Bi( f ) ( n f ) B (j e ) ( ne ) B (j f ) ( n f ) ne + n f

(7.15)

][

(7.16) (7.17)

wij (ne , n f ) = t ij (ne , n f ) bij (ne , n f )

where Aij and Bi or j have the same formulations as Equations (7.10) and (7.11); and

i, j = 1, 2,

, p . Then the F-test can be undertaken. If


1 (ne , n f ) (ne + n f ) P '1
(ne , n f ) P' < F2

F (P ' , ne + n f P '1) =

(7.18)

is satisfied, clusters e and f can be merged into a new cluster h . Otherwise, it should be similarly tested whether other clusters can be merged for e = 1, 2, and f = 2, 3, ,H . , H 1

7.3.4. Prediction

After all calculations and tests have been completed when all hypotheses of further cut or mergence are rejected, a cluster tree can be derived for each dependent variable.
215

Each cutting point, which leads to two branches, corresponds to a value x ( h ) r * of an * *


r ,k

independent variable. When a new sample set of independent variables

{xr }

is

examined, its x r * values can be compared with x ( h ) r * at the cutting points, and * *
r ,k

classified into the relevant branches. Step-by-step, the sample will finally enter a tip cluster which cannot be either cut or merged further. The criterion to classify a new sample to the relevant branches is: (a) sample data with x r * x ( h ) R* are merged into cluster e ( < f ), * *
r ,k

(b) sample data with x r * > x ( h ) R* are merged into cluster f . * *


r ,k

Let e' be the tip cluster where the new sample enters. Then the predicted dependent variables

{y i }

are: (7.19)

yi = yi( e ') Ri( e ')

Where yie ' is the mean of dependent variable i in sub-cluster e' , and Rie ' is the radius of y i in cluster e' :

yi( e ') =

1 ne ' ( e ') yi,k i ne ' k =1

(7.20)

ne ' ne ' Ri( e') = max( yi(,ek') ) min( y i(,ek') / 2, i k =1 k =1

(7.21)

7.4. Discrete and Nonlinear Optimization 7.4.1. Optimization for Optimal Control Conditions

The functions of SI emulator and controller could be realized through the stepwise

216

cluster analysis based on a large numbers of experimental inputs and outputs. For the SI emulator with given composting substrate characteristics, a pseudo equation, X = G (U ) , can be used to describe its function, where X = ( x1 , x 2 , (substrate characteristics), and U = (u1 , u 2 , conditions). The U = (u1 , u 2 ,
, u m ) from the SI controller cannot be considered optimal , x n ) are named state variables

, u m ) are control variables (operating

because there are no standard optimization curves of composting process that can be used as a reference for composting experiments. Therefore, the U = (u1 , u 2 ,
, u m ) must be

optimized before it is applied to real-world composting processes. This could be realized through adjusting the control conditions. Thus, the optimization effort is to identify the desired control actions that lead to the most satisfied reaction efficiency based on the predicted results from the SI emulator. Generally, for a given substrate in an aerobic composting process, the final OUA only has minor fluctuation (Hamelers, 2001). In this study, it is assumed that the final OUA is constant for the given substrate. As a result, the objective for the study system could be formulated as:

min J = [OUA(t + i + 1) OUA(t + i )] + m [u m (t + i )] (7.22a)


2 2 i = N1 m =1 i =1

N2

Nu

subject to

(i = 1, 2, ..., N u ) u min u (t + i ) u max OUA OUA(t + i ) OUA (i = N1 , ..., N 2 ) min max (i = 1, 2, ..., N u ) u (t + i ) u (t + i ) u (t + i ) = 0 (i > N 1) u
217

(7.22b)

where
u m (t + i ) = the

m th control increment at time t + i , which is defined as

u m (t + i ) u m (t + i 1) ;
m = control input weight factor;
N1 = minimum output horizons; N 2 = maximum output horizons;
N = control horizon;
M = number of control factors;

= constant used to determine the change ranges of control variables at time t + i .


The OUA is an increasing function of time, and thus can be considered as an indicator of composting-process progression (Hamelers, 2001). Therefore, the function (7.22) seeks to maximize the OUA increasing rate (i.e. maximize composting reaction rate) under minimized actions of controlling interference.

7.4.2. Characteristics of Genetic Algorithm

The genetic algorithm (GA) was developed through imitating the principle of natural evolution (Holland, 1975). It differs from the other optimization and search procedures in four main ways: (i) it uses a coding of the decision variable set rather than the decision variables themselves, (ii) it searches for optimal solutions in a population of decision variable sets rather than a single decision variable sets, (iii) it uses the objective function itself rather than derivative information of the objective function and constraints, and (iv) it uses probabilistic transition rules for the realization perturbation rather than

218

deterministic rules (Goldberg, 1989; Davis, 1991). Thus, the advantage of using GA is that the implementation procedure is independent of the cluster tree (that defines the interrelationship between system performance and operating conditions) and the objective function. The objective function could be highly nonlinear. With these characteristics, GA is computationally simple, but powerful in its optimal-solution search procedure in large and complex decision spaces. GA could also effectively deal with problems with non-convexities (Borkowski and Grabska, 2001); therefore, it is applicable for solving the discrete and nonlinear optimization model with high nonlinearity (Holland, 1975, Chipperfield et al., 1994). The details of GA technique were described in Holland (1975), Goldberg (1989), Davis (1991) and Rawlings (1991).

7.4.3. Procedure of Genetic Algorithm

A genetic algorithm mainly includes five components: (i) encoding, (ii) initialization of the population, (iii) fitness evaluation, (iv) evolution performance, and (v) working parameters. Figure 7.2 shows the general procedure for a genetic algorithm. The procedure includes the following steps (Rawlings, 1991):
Step 1: Encode the model parameters; Step 2: Predetermine the working parameters; Step 3: Randomly generate an initial population of chromosomes (binary strings) as

current generation;
Step 4: Decode the binary strings into primary decision variables for each design in the

current population;

219

Step 5: Evaluate fitness for each chromosome in the current population. This step

consists of computing the related objective function value and mapping the objective function value into non-negative fitness using appropriate scaling method;

220

Start

Encoding of Model parameters

Population initiation

Decision variables decoding

Fitness evaluation

Parent chromosomes selection

Crossover and mutation to produce a new generation

Using the new generation as the current generation

No

Meeting the stopping criteria? Ye Stop

Figure 7.2 Procedure of a simple genetic algorithm

221

Step 6: From the current population, select chromosomes with the highest fitness values

as parent chromosomes;
Step 7: Randomly select mating pairs from the parent chromosomes and perform

crossover operation to create new generation of children chromosomes at the predetermined crossover rate;
Step 8: Perform mutation operation to the children chromosomes at the mutation rate.

This step helps to prevent the algorithm form converging to a local optimum;
Step 9: Replace the entire population by the children chromosomes as the current

generation;
Step 10: Check the stopping criteria. If the stopping criterion is met, stop; otherwise, go

to step 3. In the applications, these steps would be repeated from generation to generation for a specified number of generations or until given stopping criterion (e.g. convergence of the population to a single solution) is achieved. Many applications of GA proved its effectiveness and robustness (Davis, 1991; Koumosis and Georgion, 1994; Natsuaki et al., 1995; Nagendra et al., 1996; Bennet et al., 2000; Rafiq et al., 2001).
Encoding and decoding

Encoding is to convert a decision variable into a string which is named chromosome (Holland, 1975; Reeves, 1995). The bits of the encoded variables are named genes. If there are n decision variables in an optimization problem and each decision variable is encoded as a k-bit binary number, then a chromosome is a string of n k binary digits as shown in Figure 7.3 (Walker et al., 2003). In an encoding process, the

222

length of the string depends on the required precision. For example, if the domain of variable xj is [ x j ,min , x j , max ] and the required precision is 2 places after the decimal point, then the required length (demoted with k) for a variable can be calculated as follows:

2 k 1 < ( x j ,max x j ,min ) 102 2 k 1


Obviously, accuracy of the computing can be improved by increasing k.

(7.23)

Correspondingly, in decoding process, the value of the j-th decision variable xj can be computed using the following equations:

x j = x j ,min +

( x j ,max x j ,min ) 2 1
k

i = ( j 1)k +1

a(i) 2

jk

i ( j 1)k 1

(7.24)

Where a(i) = (0, 1), is the i-th binary digit value in the corresponding chromosome (Davis, 1991).
Fitness evaluation

The evaluating criterion is a crucial aspect of GA. It is usually referred to as the fitness or objective function in conventional optimization methods (Holland, 1975; Rawlings, 1991; Ugwu and Tah, 2002). The fitness is used to select parent chromosomes to generate new chromosomes. The fitness of a member is a non-negative function which is determined by the objective function and constraints. The members are ranked according to their fitness in an increasing order. If a negative value exists among the calculated fitness then every fitness value is reduced by the minimum fitness in order to have all fitness values positive (Turkcan and Akturk, 2003).
Evolution performance

In general, there are three procedures to perform evolution and generate the new
223

population, including selection, crossover and mutation. The GA uses the random selection techniques to choose the chromosomes (strings) as parent ones for generating a new generation (Holland, 1975; Goldberg, 1989). The roulette-wheel selection was used in this section (Diamantis and Tsahalis, 2002). The selected probability (Pselect) for a single chromosome depends on the proportion of its fitness to the total fitness of the population. The selected chromosomes are called parent chromosomes or strings. Parent chromosomes are chosen with this method until the mating pool is full. The crossover operator is the most important search operator of genetic algorithms. Useful segments of different parents are selected and combined to generate a new individual that benefits from advantageous bits combinations (Holland, 1975). The process of crossover is applied to recombine two chromosomes (parents strings) and generate two new chromosomes (offspring string), when a random value associated to this pair is greater than a predefined crossover rate (Pcross) (Diamantis and Tsahalis, 2002). A two-point crossover is shown in Figure 7.4. It chooses two crossover locations within the digits of the strings randomly, and exchanges the digits between the two parent strings. Mutation maintains variability in the population, and reduces the chance that the population will prematurely converge on one possible sub-optimal solution (Davis, 1991). The operation of one-point mutation simply alters one bit at randomly predetermined location in the chromosome when a (0, 1) random value associated to that bit is smaller than a predefined mutation rate (Pmutat) (Chelouah and Slarry, 2000). For binary encoding, the bit value of 0 is change to 1 and vice versa. For standard genetic algorithms, the probability of any member in the population to be selected to undergo mutation is: Pi = 1/M (7.25)

224

A sample chromosome

Decision variables

Decision Variable 1

Decision Variable 2

Decision Variable N

Figure 7.3

Decision variables encoding

225

Parent Chromosome


Crossover

Old string 1

Old string 2


Children Chromosome Figure 7.4 Two points crossover operation

New string 1

New string 2

226

Working parameters

A set of parameters are needed to be predefined for guiding the genetic algorithm (Holland, 1975; Davis, 1991; Rawlings, 1991; Nagendra et al., 1996; Bennet et al., 2000; Rafiq et al., 2001): (i) chromosome length L which is the product of the number of decision variables (n) and the length of a string (k), (ii) population size M which is usually within the range of 30 to 200, (iii) crossover rate Pcross which is usually within the range of 0.6 to 0.95, (iv) mutation rate Pmutat which is usually within the range of 0.001 to 0.05, and (v) stopping criterion which is used to judge whether stop the search process. After a predetermined generation number N is reached or when there are no significant differences among the best solutions, the process will be stopped. Based on Figure 7.1, the procedure for implementing SIPC can be listed as follows: a) Define X (t ) and denote an element of X (t ) as x i , where i =1, 2, , n , and n is the total number of variables characterizing the composting substrate characteristics; X (t ) is the input of the controller. b) For the given X (t ) , the controller can produce a control action U (t ) , and denote an element of U (t ) as u j , where j =1, 2, , m , and m is the total number of control variables. The control variables combining the substrate characteristics, i.e. X (t ) + U (t ) , become an input for the emulator and the emulator can produce an output X p (t + 1) . c) The X p (t + 1) will be compared with X (t ) . If the objective is satisfied, the

227

control action U (t ) will be considered the optimal; then it could be applied to practical composting process. If not, the control action will be adjusted with a new input for the emulator; then the emulator will produce a new output to be optimized, till the optimal result, i.e. U (t ) , is achieved.
*

d) After the optimal control action is applied to the practical process, the output will become the next input for the SI controller; then repeat the steps a to c . Information of the implemented composting process could be used for further updating the SI controller and SI emulator.

7.5. Application 7.5.1. Data Collection

The data used to initiate the SIPC are the same as those used to train the developed MNPC in Chapter 6. For each composting experiment based on the measured initial substrate characteristics, there were 16 sets of data (samples were taken twice a day). Hence, there were 256 sets of data from 16 experiments. These data are about the relationships between the composting substrate characteristics (Ss, Si, X and OUA) and the operating conditions (T, M and O2) in a composting process, and can be used to initiate the SI emulator and controller. For the SI controller, the inputs include the concentrations of soluble substrate, insoluble substrate, and active biomass, and OUA; the outputs are temperature, moisture, and oxygen content. For the SI emulator, the inputs include temperature, moisture, oxygen content, OUA, and the concentrations of soluble substrate, insoluble substrate, and active biomass; the outputs are the concentrations of soluble substrate, insoluble

228

substrate, and active biomass, and OUA. The input and output variables for the SI controller and emulator are listed in Table 7.1.

229

Table 7.1 Input and output variables for the SI controller and emulator

SI Controller Insoluble substrate Soluble substrate Biomass Oxygen Uptake Accumulation Temperature Moisture Oxygen content

Input (I) or Output (O) I I I I O O O

Symbol Si Ss X OUA T M O2

SI Emulator Insoluble substrate Soluble substrate Active biomass Oxygen Uptake Accumulation Temperature Moisture Oxygen content Insoluble substrate Soluble substrate Biomass Oxygen Uptake Accumulation

Input (I) or Output (O) I I I I I I I O O O O

Symbol Si Ss X OUA T M O2 Si Ss X OUA

230

7.5.2. Stepwise Cluster Analysis

Significant correlations existed between the input and output varialbes of the SI controller and emulator. Therefore, the techniques of stepwise cluster analysis was applicaple. Totally, 3 cluster trees for the SI controller and 4 for the SI emulator were obtained, forming a set of forecasting systems. Figures 7.5 and 7.6 present two cluster trees as examples to illustrate the modeling results. Figure 7.5 is the tree of temperature in the SI controller. Figure 7.6 is the tree of the active biomass concentration in SI emulator. In these cluster trees, the criteria for cutting and merging clusters are: (a) cut a cluster when p < 0.05 , (b) merge clusters when p > 0.05 . where the p values shown at cutting and merging knots are significance levels of the
F-test; p > 0.05 suggests no statistical difference between two clusters. The cluster

trees could help clearly show the significance levels of different independent variables. Based on these trees, the control factors can be predicted through the SI controller. For example, set Si (X1) = 3631.1, Ss (X2) = 2087.4, X (X3) = 53.3, and OUA (X4) = 119.6 as the inputs for the SI controller. To predict the temperature (T) for the next time period, we have: X3 > 21.82 for the first branch knot so that the sample enters cluster 3 (Figure 7.5); X3 < 88.38, so that it enters cluster 6; X3 > 31.13, so that it enters cluster 13; X4 < 1343.4, so that it enters cluster 22; X1 < 4133.5, so that it finally enters cluster 50 with a prediction value of T = 59.21 2.85C. Similarly, we can obtain the predicted results of moisture and oxygen content. Based on the predicted results from SI controller, the control variables could be
231

optimized through the predictive function of the SI emulator. For example, let the substrate characteristics the same as the above: Si (X1) = 3631.1, Ss (X2) = 2087.4, X (X3) = 53.3, and OUA (X4) = 119.6. Then the predicted results from the SI controller are Temperature (X5) = 59.21, Moisture (X6) = 0.632, and Oxygen content (X7) = 0.119. Thus, the prediction process based on the SI emulator is (Figure 7.6): X5 > 59 for the first branch knot so that the sample enters cluster 3; X7 < 0.138, so that it enters cluster 6; X1< 4377.5, so that it enters cluster 12; X7 < 0.131, so that it enters cluster 23; X4 < 152.77, so that it finally enters into cluster 47 with a prediction value of X = 136.12 9.47 molm-3. Similarly, Si, Ss and OUA can be predicted. The predicted results from the SI emulator will be optimized based on objective function (7.22). The optimization task could be completed through the GA based on the predicted results from the SI controller. Performance of the stepwise cluster analysis method was analyzed through comparisons of the predicted results and the data. The error analysis for the SI controller and emulator is shown in Table 7.2. It can be seen that the maximum average relative error is 1.83% (O2) for the controller and 4.24% (X) for the emulator; the maximum relative error is 10.62% (O2) for the controller and 45.71% (OUA) for the emulator. The results of Table 7.2 demonstrated a satisfactory performance of the proposed method.

7.5.3. Result and Discussion

After the initiation of the SIPC, a single composting experiment under controlled conditions was undertaken to justify the developed system. The initial concentrations of soluble substrate (Ss), insoluble substrate (Si), active biomass (X), and inert material (I) were 2542.7, 3362.3, 1.771, and 426.3 molm -3 , respectively. During the composting

232

1 2
x4 0 x3 21.82 p < 0.05

3 6
x3 88.38 p < 0.05 p < 0.05 x2 26.70

4 9

x4 5.76

p < 0.05

5
x3 17.94 p < 0.05

7
p < 0.05 x2 26.50 p < 0.05

p < 0.05 x3 12.25 p < 0.05

10 16 30 46
(32.44 3.70)
x3 5.12 p < 0.05

8
(22 .00 0.00)

11
x1 3866.2

x3 31.13

x2 25.80 p < 0.05

12 20 33
p > 0.05 x2 25.9 p < 0.05 p > 0.05

13 21 22
x1 4133.5 p < 0.05

14 23

15 26

17 31

x1 3792.4 p < 0.05

18

p < 0.05

x3 194.55 p < 0.05

19

28

29
x2 25.5 p < 0.05

32

x3 18.61 p < 0.05

x4 1343.4 p < 0.05

24 25
(62.71 2.70)
p > 0.05

27
x2 28.10 p < 0.05 x3 331.05 p < 0.05

x3 125.39 p < 0.05

41
p > 0.05

x3 14.10 p < 0.05

34 49

35
p > 0.05

37 38

39 52 51

40 53
(66.5 0.51)

42 43

47
x3 16.05 56 p < 0.05 p > 0.05

48
(44.73 0.40)

(30.25 (33.17 3.40) 0.55)

36
(56.92 2.60)

(52.75 0.90)

55

58

(59.21 2.85)

50

(63.26 1.45) (63.30 2.00)

(65.25 0.68)

44 45
p > 0.05

(40.06 0. 60)

54 57
(44.98 4.85)

(36.73 0.50)

Figure 7.5 Cluster tree of temperature in the SI controller

233

2 4
x3 3.31 p < 0.05

x5 59

p < 0.05

3 6
x7 0.138 p < 0.05 p < 0.05 x5 65.0

x7 0.096

p < 0.05

5
p < 0.05

7
p < 0.05

x5 37.4 p < 0.05

p < 0.05 x6 0.66

9
p < 0.05

10 20 40

x5 58.0
P > 0.05

11

x1 4377.50

17
x4 4.75 p < 0.05 x4 0.07

x3 13.08 p < 0.05 x3 45

18

19

16
p < 0.05 x6 0.51 p < 0.05

12
x7 0.131 p < 0.05

14 13x1 4204.10 25 41

p < 0.05 X2 122.96 p < 0.05

15

p < 0.05 x4 23.71 p < 0.05

x6 0.50

33

34
x7 0.61

35 36 p < 0.05 p < 0.05 x4 11.9 x5 36.4 p < 0.05 57 58


(8.91 0.85)

37 38

39

31

32
x7 0.14

21 22
p < 0.05

p > 0.05

23 24
p > 0.05

x7 0.145

p < 0.05

26

27 43

28
p < 0.05

42
p > 0.05 x6 0.58

x4 1079.20

51 52
(2.98 0.34)
x4 3.66

53 54 55 56
P > 0.05 p < 0.05

x4 152.77

(28.70 2.20)

49 50
p < 0.05

29 < 0.05 p 30 47 48

p > 0.05 x4 375.09 p < 0.05

44
p < 0.05

(6.86 0.73)

p < 0.05 (10.88 0.46) x3 26.99 p < 0.05

45

p < 0.05

x1 3914.40

(136.12 9.47)(156.99 8.34)


x5 62.10

x3 13.7

61 62
(19.51 0.40)
p < 0.05 p > 0.05

73 74
p > 0.05

63

64

46

65
p < 0.05 p < 0.05

66

75
(2.98 0.34)

76

77

78
x4 1410.60

(3.89 0.37) (11.68 0.21) (12.24 0.26)

69
p > 0.05

70

71
(250.50 4.84)
p < 0.05

72 (354.88 5.84) 88 89 (340.40 5.13) 92


(328.94 9.20)

x4 730.11

93 68
(21.53 2.48)
x4 57.05

67
(4.75 0.39)

79
(13.97 0.20)

80

(26.63 0.47)
p < 0.05

85

x4 1275.30

P > 0.05

(13.35 0.28)
x7 0.07 p < 0.05

90 60
p < 0.05 x2 26.50

91
x7 0.133 p < 0.05 p>0.05

(231.26 10.76) (185.37 5.14)

86

87
p < 0.05

x4 1358.50

59
x1 4397.60

96 97
p < 0.05

(316.56 9.20)

94

95

p < 0.05 x1 4423.0

98
(190.28 0.43) (192.65 1.46)
P > 0.05

83

84
p > 0.05

99 100
x4 66.70 p < 0.05

(17.57 0.56) (18.19 0.42)

101 81 82
x4 49.54 p < 0.05

102 103

(15.08 0.50) (15.82 0.26)

104 105
(57.47 3.17) (71.78 2.65) (80.01 10.41) 106 (64.98 5.72)

Figure 7.6 Cluster tree of active biomass concentration in the SI emulator

234

Table 7.2 Error analyses for SI controller and emulator


Average relative error (%) Minimum relative error (%) Maximum relative error (%)

T SI controller M O2 Ss SI emulator Si X OUA

1.26 1.02 1.83 2.16 1.59 4.24 3.65

0.01 0.02 0.04 0.21 0.08 0.04 0.02

9.58 7.23 10.62 22.54 4.86 37.45 45.71

235

experiment, the conditions were controlled based on the optimization results of SIPC. Figures 7.7 to 7.9 show the controlled results compared with the predicted ones of SIPC. The error-analysis results for experimental control conditions are shown in Table 7.3. It can be seen from Figures 7.7 to 7.9 that all of the controlled values well matched the predicted ones. If ARE and correlation coefficient were used to represent the accuracy and consistency between the controlled and predicted results, it can be seen that the AREs of Figures 7.7 and 7.8 were 2.63 and 2.01%, respectively; the correlation coefficients (R) were 0.995 and 0.990, respectively. The results showed that the controlled temperature and moisture levels were of satisfactory in both accuracy and consistency (Figures 7 and 8). In Figure 7.9, the controlled values of oxygen contents had minor difference with the ARE and correlation coefficient being 5.86 and 0.987, respectively. However, these results were also satisfactory although they were not as good as those of temperature and moisture. Therefore, the controlled results of composting conditions in Figures 7.7 to 7.9 were basically consistent with the predicted results of SIPC and could meet the requirements of system justification. Under the controlled composting conditions, a comparison of the predicted results from the SIPC with the observed ones are shown in Figures 7.10 to 7.13, and the related error analyses are provided in Table 7.4. In Figures 7.10 to 7.13, the predicted results of SIPC were basically consistent with the observed results. The ARE levels of Figures 7.10 and 7.11 were 8.01 and 2.41%, respectively; the correlation coefficients were 0.984 and 0.995, respectively. These results demonstrated that the predicted results well matched the observed ones with satisfactory accuracy and consistency. In Figures 7.12 and 7.13, the AREs were

236

16.16 and 18.43%, respectively; the correlation coefficients were 0.982 and 0.989, respectively. Such results showed a satisfactory consistency but a less satisfactory accuracy. Combined with the result analyses of Chapter 6, it could be concluded that the concentration of active biomass was more difficult to predict than those of the others no matter what kinds of technology were adopted; while the OUA is directly related to the concentration of active biomass, and thus had a similar situation. Of course, these results were generally satisfactory considering the dynamics and complexities of composting systems. In fact, as discussed in Chapter 6, composting was a slow biological process, such that the concentrations of insoluble substrate and soluble substrate could not vary significantly in a relative short period of time. Furthermore, both the predicted and observed results were based on the same initial conditions, i.e., each time point in the figures had the same initial conditions. Consequently, there were no accumulated errors in Figures 7.10 to 7.13.
7.5.4. Comparison of SIPC and MNPC

The basic concept of the developed SIPC is similar to that of MNPC, and the produced results are both satisfactory. In the MNPC experiment, the average relative errors between controlled and predicted temperature, moisture, and oxygen content were 3.23, 2.82, and 8.28%, respectively; In the SIPC experiment, the average relative errors between predicted and observed temperature, moisture and oxygen content were 2.63, 2.01, and 5.86%, respectively. These results showed that the controlled conditions in these two experiments had a similar accuracy level and should not have significant effects on the justification results.
237

80 70 60 50 40 30 20 10 0 0 24 48 72 96 120 144 Time (hr)

Controlled Predicted

Temperature (C)

168

192

Figure 7.7 Comparison of the controlled and predicted results for temperature

238

80 Moisgure (%) 70 60 50 40 30 0 24 48 72 96 120 144 168 192 Time (hr) Controlled Predicted

Figure 7.8 Comparison of the controlled and predicted results for moisture

239

0.18 0.15 0.12 0.09 0.06 0.03 0 0 24 48 72 96 120 144 Time (hr)

O2 (molm )

Controlled Predicted

-3

168

192

Figure7.9 Comparison of the controlled and predicted results for oxygen content

240

Table 7.3 Error analysis for controlled experiment conditions Control variable Temperature Moisture Oxygen content Average relative error (%) 2.63 2.01 5.86 Minimum relative error (%) 1.04 0.87 1.11 Maximum relative error (%) 6.59 6.41 15.42 Correlation coefficient 0.995 0.990 0.987

241

3000 Predicted Value (mol Ssm ) 2500 2000 1500 1000 500 0 0 500 1000 1500 2000
-3

-3

2500

3000

Observed Value (mol Ssm )

Figure 7.10 Predicted and observed results for soluble substrate concentration

242

4500 4300 4100 3900 3700 3500 3300 3300

Predicted Value (mol Sim )

-3

3500

3700

3900

4100
-3

4300

4500

Observed Value (mol Sim )

Figure 7.11 Predicted and observed results for insoluble substrate concentration

243

400 350 Predicted Value (mol Xm ) 300 250 200 150 100 50 0 0 50 100 150 200 250
-3 -3

300

350

400

Observed Value (mol Xm )

Figure 7.12 Predicted and observed results for active biomass concentration

244

1800 1600 Predicted Value (mol Om ) 2 1400 1200 1000 800 600 400 200 0 0 200 400 600 800 1000 1200 1400 1600 1800
-3 -3

Observed Value (mol O2m )

Figure 7.13 Predicted and observed results for OUA

245

Table 7.4 Error analysis for the predicted results of SIPC Average relative error (%) 8.01 2.41 16.16 18.43 Minimum relative error (%) 1.21 0.31 2.12 1.09 Maximum relative error (%) 70.42 6.08 73.21 85.68 Correlation coefficient 0.984 0.995 0.982 0.989

Variable Ss Si X OUA

246

However, as shown in Table 7.5, the average relative errors of the predicted results for MNPC were 9.24 (Ss), 2.52 (Si), 24.47 (X), and 22.56% (OUA), respectively; in comparison, the errors from the SIPC were 8.01 (Ss), 2.41 (Si), 16.16 (X), and 18.43% (OUA), respectively. It can be seen that the error levels from the SIPC were lower than those from the MNPC. Similarly, the correlative coefficients from the SIPC were higher than those from the MNPC: Ss (0.984 vs. 0.978), Si (0.995 vs. 0.992), X (0.982 vs. 0.972), and OUA (0.989 vs. 0.986). Based on the above facts, we could conclude that the developed SIPC was of higher accuracy and consistence levels than the MNPC. Figures 7.14 to 7.20 show comparisons of the two developed systems based on the same initial conditions where the initial concentrations of soluble substrate, insoluble substrate, active biomass and inert material were 2500, 3500, 2 and 400 molm-3, respectively. In these figures, the correlation coefficients were 0.991 (T), 0.986 (M), 0.984 (O2), 0.981(Ss), 0.985 (Si), 0.967 (X), and 0.993 (OUA). These results indicated that the predicted values from the two developed systems had a similar trends and a significant consistency level. However, it was difficult to tell which curve was closer to the optimum since there were no standard curves for composting processes.

7.6. Summary

In this section, a real-time composting process control system based on stepwise cluster analysis and GA optimization technology has been developed. The system includes a SI emulator, a SI controller, an optimization procedure, and a composting reactor. The stepwise cluster analysis was used to establish the SI controller and emulator

247

based on a series of experimental data. The SI controller could produce a control action based on a substrate characteristic, and the SI emulator could predict the next stages substrate characteristic based on the given substrate characteristic and control action. To identify a desired control action, the GA technique was applied. The developed system has been justified based on a series of composting experiments. The results showed that the developed SIPC could effectively tackle the dynamics and complexity of composting processes and support the control for the composting process. The developed SIPC was compared with the MNPC, with the results showing that the predicted outputs of the SIPC were more accurate than those of the MNPC.

248

Table 7.5 Error analyses for results from SIPC and MNPC Average relative error (%) 8.01 2.41 16.16 18.43 9.24 2.52 24.47 22.56 Minimum relative error (%) 1.21 0.31 2.12 1.09 1.73 0.34 2.24 1.36 Maximum relative error (%) 70.42 6.08 73.21 85.68 93.42 6.63 133.51 140.65 Correlation coefficient 0.984 0.995 0.982 0.989 0.978 0.992 0.972 0.986

System

Variable Ss Si

SIPC

X OUA Ss Si

MNPC

X OUA

249

70 60 Temperature (C) 50 40 30 20 10 0 0 24

SIPC R = 0.991 MNPC

48

72

96 Time (hr)

120

144

168

192

Figure 7.14 Predicted temperature levels from SIPC and MNPC

250

80 Moisgure (%) 70 60 50 40 30 0 24 48 72 96 120 144 168 192 Time (hr) R = 0.986


SIPC MNPC

Figure 7.15 Predicted moisture levels from SIPC and MNPC

251

0.18 0.15 0.12 0.09 0.06 0.03 0 0

R = 0.984

SIPC MNPC

O2 (molm )

-3

24

48

72

96

120

144

168

192

Time (hr)

Figure 7.16 Predicted oxygen content levels from SIPC and MNPC

252

Ss (molm )

3000 2500 2000 1500 1000 500 0 0

R = 0.981

SIPC MNPC

-3

24

48

72

96

120

Time (hr)

Figure 7.17 Predicted soluble substrate concentrations from SIPC and MNPC

253

5000 Si (molm )
-3

R = 0.985

SIPC MNPC

4500 4000 3500 3000 0 24 48 72 96 120 144 168 192 Time (hr)

Figure 7.18 Predicted insoluble substrate concentrations from SIPC and MNPC

254

350 300 250 200 150 100 50 0 0

R = 0.967

SIPC MNPC

X (molm )

-3

24

48

72

96

120

144

168

192

Time (hr)

Figure 7.19 Predicted active biomass concentrations from SIPC and MNPC

255

1800 1500 1200 900 600 300 0 0

OUA (mol O2 m )

-3

R = 0.993 SIPC MNPC

24

48

72

96 Time (hr)

120

144

168

192

Figure 7.20 Predicted OUAs from SIPC and MNPC

256

CHAPTER 8 CONCLUSIONS

8.1. Summary

(1) In this study, methodologies of modeling and process control for composting system that are associated with a variety of complexities and uncertainties have been developed. They include (i) a multi-component modeling system for the aerobic composting process, (ii) a simulation-based factorial analysis approach to characterize the interactive effects of multiple factors on the composting process, (iii) a simulation-based fractional fuzzy vertex method combined with interval analysis to tackle the effects of uncertainties on the composting process, (iv) a model-based real-time neural predictive control system (MNPC) for the composting process, and (v) a stepwise-inference-based real-time predictive control system (SIPC) for the composting process. In addition, a series of composting experiments have been undertaken to verity the developed model and justify the real-time process control systems. (2) A multi-component modeling system has been developed to simulate the composting process based on the principles of microbiology, mass transfer, and biotechnology. The developed model was verified through the composting experiment and the results showed that the simulation results were generally consistent with the experimental outputs with acceptable accuracy levels. Moreover, further sensitivity analyses were conducted with the results demonstrating that an increased growth rate or decreased decay rate of biomass would lead to a raised oxygen uptake rate and thus an enhanced composting process.
257

(3) A simulation-based factorial analysis approach has been developed to characterize the interactive effects of multiple factors on the composting process. It is based on the developed composting-process model and the two-level factorial analysis method. In the study, normal probability plots have been produced to support the systems analysis when high-order interactions occur. The study results could be useful for guiding composting-process operation and management. (4) Based on the developed composting-process model, the effects of uncertainties on the composting process have been characterized through a fractional fuzzy vertex method combined with interval analysis. The effects of uncertain control factors and input parameters as well as their combined effects have been studied. In addition, the effects of uncertain control factors have been quantified through a series of sensitivity analyses. The study results were useful for guiding practical process operation and control. (5) A model-based real-time neural predictive process control system for waste composting process has been developed. A series of composting experiments have been undertaken to train and justify the MNPC. The testing results showed that the predicted values were basically consistent with the experimental ones, indicating that the developed MNPC was generally satisfactory. (6) A real-time composting process control system based on stepwise cluster analysis and GA optimization technology has been developed. The system includes a SI emulator, a SI controller, an optimization procedure, and a composting reactor. The developed system was justified based on a series of composting experiments. The results showed that the developed SIPC could effectively tackle the dynamics and complexity of

258

composting process and support the control for the composting process. In addition, the developed SIPC was compared with the MNPC, with the results showing that the predicted outputs of the SIPC were more accurate than those of the MNPC.

8.2. Research Achievements

(1) The developed multi-component composting-process simulation model improved upon the previous studies of composting-process simulation. In the developed model, the processes of microbial growth, decay, and hydrolysis were incorporated. The model could simulate composting processes with multiple substrates based on varied settings of the model parameters. It could help provide a basis for further developing model-based process control systems. (2) The developed simulation-based factorial analysis method and the fractional fuzzy vertex technique combined with interval analysis were new attempts for characterizing the interactive and uncertain effects from various impact factors and their interrelations. (3) The developed MNPC and SIPC were both the real-time process control systems, which were both the first attempt in the field of composting-process control. These developed systems could be applied to real-world composting process and thus help improve the efficiency of the composting process and the quality of the generated products.

8.3. Recommendations for Future Research

(1) Due to the complexities of composting processes, further improvements for

259

the developed models and process-control techniques are desired, particularly in the reflection of system uncertainties and dynamics. (2) In the developed simulation model, seven state variables have been taken into account. To make the model more applicable, it is desired that more factors be involved. Furthermore, the empirical equations in the model could be improved through additional experimental work, in order to upgrade the prediction accuracy. (3) To characterize the interactive effects of the composting process, a simulation-based factorial analysis method was developed. A series of composting experiments could be conducted for verifying the research results. (4) In this study, only two kinds of uncertainties were taken into account in developing the fuzzy vertex method combined with interval analysis. Further studies by considering more uncertain factors are desired for supporting more robust management and operation of composting systems. (5) The developed MNPC and SIPC were both based on the same set of experimental data, and the justification experiments were also conducted under similar control conditions. Actually, the composting systems could have different styles, and the substrate compositions could also be of various settings. If the developed systems were applied to other composting processes with significantly different substrate compositions, then more persuasive conclusions could be made.

260

References
Abu-Rub, H., J. Guzinski, Z. Krzeminski, and H. Toliyat (2004). Advanced control of induction motor based on load angle estimation. IEEE Trans. Ind. Electron., 51(1): 414. Alexander, S. and R. Michael (2001). Robustness of a neural network model for differencing. Journal of Computational Neuroscience, 11(2): 165-173. Alpert, J. E., M. Hernandez, and Y. Soto (2002). Process and structural upgrades improving a state-of-the-art composting facility. BioCycle, 43(5): 40-42. Antwerp, J. G. V. and R. D. Braatz (2000). Model predictive control of large scale processes. Journal of Process Control, 10: 1-8. Ashbolt, N. J. and M. A. Line (1982). A bench-scale system to study the composting of organic wastes. Journal of Environmental Quality, 11(3): 405-408. Astrom, K. J. and B. Wittenmark (1989). Adaptive Control. Addison-Wesley, New York, USA. Atchley, S. H. and J. B. Clark (1979). Variability of temperature, pH, and moisture in an aerobic composting process. Applied and Environmental Microbiology, 38(6): 1040-1044. Atkinson, C. F., D. D. Jones, and J. J. Gauthier (1996). Biodegradabilities and microbial activities during composting of Municipal Solid Waste in bench-scale reactors. Compost Science and Utilization, 4(4): 14-23. Bach, D. B., M. Shoda, and H. Kubota (1985). Composting reaction rate of sewage sludge in an autothermal packed bed reactor. Journal of Fermentation Technology, 63(3): 271-278.
261

Bach, P. D., K. Nakasaki, M. Shoda, and H. Kubota (1987). Thermal balance in composting operations. Journal of Fermentation Technology, 65(3): 199-209. Baker, J. M. and R. R. Allmaras (1990). System for automating and multiplexing soil moisture measurement by time-domain reflectometry. Soil Science Society of America Journal, 54: 1-6. Bakshi, M., V. K. Gupta, and P. N. Langar (1987). Effect of moisture level on the chemical composition and nutritive value of fermented straw. Biological Wastes, 21: 283-290. Balasubramhanya, L. S. and F. J. Doyle (2000). Nonlinear model-based control of a batch reactive distillation column. Journal of Process Control, 10: 209-218. Bari, Q. H. (2000a). Kinetic analysis of forced aeration composting: Reaction rates and temperature. Waste Management and Research, 18(4): 303-312. Bari, Q. H. (2000b). Kinetic analysis of forced aeration composting: Application of multilayer analysis for the prediction of biological degradation. Waste Management and Research, 18(4): 313-319. Beck, M. B. (1979). Model structure identification from experimental data. In: Theoretical Systems Ecology: Advances and Case Studies, ed. E. Halfon, Academic Press, London, UK, pp. 259-289. Beck, M. B. (1984). Understanding environmental systems. In: Predictability and Nonlinear Modeling in Natural Sciences and Economics, eds. J. Grassman and G van Straten, Kluwer Academic Publishers, Kluwer, the Netherlands, pp. 294-311. Beck, M. B. (1993). Construction and evaluation of models of environmental systems. In: Modeling Change in Environmental systems, eds. A. J. Jakeman, M. B. Beck, M.

262

McAleer, John Wiley & Sons, New York, USA, pp. 3-35. Bennet, L. D., M. J. Mawdesley, and M. K. Ford (2000). Investigating a genetic algorithm based decision support system for the location of new major housing allocations within the local process. In: Second International Conference on Decision Making in Civil and Urban Engineering, Lyon, France, pp. 897908. Bernal, M. P., A. F. Navarro, A. Roig, J. Cegarra, and D. Garcia (1994). Carbon and nitrogen transformation during composting of sweet sorghum bagasse. Biological Fertility of Soils, 22: 141-148. Bernal, M. P., J. M. Lopez-Real, and K. M. Scott (1993). Application of natural zeolites for reduction of ammonia emissions during the composting of organic wastes in a laboratory composting simulator. Bioresource Technology, 43: 35-39. Biasiak, A. E. (1997). Development and assessment of methods to quantify ammonia volatilization during aerobic degradation of high-solids organic waste. M.S. Thesis, Cornell University, Ithaca, USA. Bodelier, P. L., and H. J. Laanbroek (1997). Oxygen uptake kinetics of pseudomonas chlororaphis grown in glucose- or glatamate-limited continuous cultures. Archives Microbiology, 167: 392-395. Borkowski, A. and E. Grabska (2001). Graphs in layout optimization. In: Artificial Intelligence in Construction and Structural Engineering. Proceedings of the 8th International Workshop of the European Group of Structural Engineering Applications of Artificial Intelligent, eds, C.J. Anumba, O.O. Ugwu and Z. Ren, Loughborough University, Leicestershire, UK, pp. 114121. Boscolo, A., C. Mangiavacchi, F. Drius, F. Rongione, P. Pavan, and F. Cecchi (1993).

263

Fuzzy control of an anaerobic digester for the treatment of the organic fraction of municipal solid waste. Water Science and Technology, 27(2): 57-68. Bosma, W. J. P., A. Bellin, and A. Rinaldo (1993). Linear equilibrium adsorbing solute transport in physically and chemically heterogeneous porous formations (II): numerical results. Water Resource Research, 29(12): 4013-4043. Box, G. E. P., W. G. Hunter, and J. S. Hunter (1978). Statistics for Experimenters. John Wiley & Sons, New York, USA. Brewer, L. J. (2001). Maturity and stability evaluation of composted yard debris. M.S. Thesis, Oregon State University, Corvallis, USA. Brodie, H. L. (1996). Exporting nutrients as compost. In: Animal Agriculture and the Environment: Nutrients, Pathogens and Community Relations, ed. M. Sailus, NRAES-96, Rochester, USA, pp. 277-284. Burge, W. D., N. K. Enkiri, and D. Hussong (1987). Salmonella regrowth in compost as influenced by substrate. Microbial Ecology, 14: 243-253. Campbell, J. E. (1990). Dielectric properties and influence of conductivity in soils at one to fifty megahertz. Soil Science Society of America Journal, 54: 332-341. Cannas, B., G. Celli, A. Fanni, and F. Pilo (2001). Automated recurrent neural network design of a neural controller in a custom power device. Journal of Intelligent and Robotic Systems, 31: 229251. Cano, R. A. R. and D. Odloak (2003). Robust model predictive control of integrating processes. Journal of Process Control, 13: 101-114. Cathcart, T. P., F. W. Wheaton, and R. B. Brinsfiled (1986). Optimizing variables affecting composting of blue crab scrap. Agricultural Wastes, 15: 269-286.

264

Chan, W. T., D. K. H. Chua, and G. Kannan (1996). Construction resource scheduling with genetic algorithms. Journal of Construction Engineering and Management, ASCE, 122: 125132. Characklis, W. G., and W. Gujer (1979). Temperature dependency of microbial reactions. Prog. Water Technol. Suppl. 1:111-130. Chelouah, R. and P. Slarry (2000). A continuous genetic algorithm designed for the global optimization of multimodal functions. Journal of Heuristics, 6: 191-213. Chen, C. T. and S. T. Peng (1999). Intelligent process control using neural fuzzy techniques. Journal of Process Control, 9: 493-503. Chipperfield, A. J., P. J. Fleming, H. Pohleim, and C. M. Fonseca (1994). Genetic Algorithm Toolbox Users Guide, ACSE Research Report No. 512, University of Sheffield, Sheffield, UK. Christensen, K. K. and M. Carlsb (2002). Strategies for evaluating the sanitary quality of composting. Journal of Applied Microbiology, 92(6): 1143-1158. Cooney, C. L. and D. L. Wise (1975). Thermophilic anaerobic digestion of solid waste for fuel gas production. Biotechnology and Bioengineering, 17:1119-1135. Cooney, C. L., D. I. Wang, and R. I. Materles (1968). Measurement of heat evolution and correlation with oxygen consumption during microbial growth. Biotechnology and Bioengineering, 11: 269-281. Cundiff, J. S. and K. R. Mankin (2003). Dynamics of biological systems. American Society of Agricultural Engineering, ASAE publication, St. Joseph, USA. Cybenko, G. (1989). Approximation by superpositions of a sigmoidal function. Math. Control Signal Systems, 2: 303-314.

265

Dahhou, V., G. Roux, and C. Chamilothoris (1992). Modeling and adaptive predictive control of a continuous fermentation process. Applied Mathematical Modeling, 16: 545-552. Das, A., N. P. Reddy, and J. Narayanan (2001). Hybrid fuzzy logic committee neural networks for recognition of swallow acceleration signals. Computer Methods and Programs in Biomedicine, 64: 8799. Das, I. and F. A. Potra (2003). Subsequent convergence of iterative methods with applications to real-time model-predictive control. Journal of Optimization Theory and Applications, 119(1): 37-47. Das, K. and H. M. Keener (1996). Process control based on dynamic properties in composting: moisture and compaction considerations. In: The Science of Composting, eds. M. DeBertoldi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 116-125. Davis, L. (1991). Handbook of genetic algorithms. Van Norstrand Reinhold, New York, USA. Dayhoff, J. E. (1990). Neural Network Architecture. Van Norstrand Reinhold, New York, USA. Dean, T. J., J. P. Bell, and A. J. B. Baty (1987). Soil moisture measurement by an improved capacitance technique (I): sensor design and performance. Journal of Hydrology, 93: 67-78. DeBertoldi, M., A. Rutili, B. Citterio, and M. Civilini (1988). Composting management a new process control through O2 feedback. Waste Management Research, 6: 239-259.

266

Demiroren, A. and E. Yesil (2004). Automatic generation control with fuzzy logic controllers in the power system including SMES units. Electrical Power and Energy Systems, 26: 291305. Deng, M. Y. and D. O. Cliver (1995). Antiviral effects of bacteria isolated from manure. Microbial Ecology, 30: 43-54. Diamantis, Z. G. and D. T. Tsahalis (2002). Optimization of an active noise control system inside an aircraft, based on the simultaneous optimal positioning of microphones and speakers, with the use of a genetic algorithm. Computational Optimization and Applications, 23: 65-76. Diehl, M, H. G. Bock, and J. P. Schloder (2002). Real-time optimization and nonlinear model predictive control of processes governed by differential-algebraic equations. Journal of Process Control, 12(4): 577-585. Dirksen, C. and S. Dasberg (1993). Improved calibration of time domain reflectometry soil water content measurements. Soil Science Society of America Journal, 57: 258-264. Dochain, D. and C. Bastin (1984). Adaptive identification and control algorithms for nonlinear bacterial growth systems. Automatica, 20 (5): 621-634. Doeblin, E. O. (1990). Measurement systems application and design. McGraw Hill Publisher, New York, USA. Donald, S. and L. H. Pamaela (1997). Neural generalized predictive control: A Newton-Raphson implementation. National Aeronautics and Space

Administration, Langley Research Center, Hampton, USA. Donat, S. A. (1991). Neural net based model predictive control. International Journal of

267

Control, 54: 1453-1468. Dong, W. and H. C. Shah (1987). Vertex method for computing functions of fuzzy variables. Fuzzy Sets and Systems, 24: 65-78. Dou, C., W. Woldt, I. Bogardi, and I. Dahab (1995). Steady state groundwater flow simulation with imprecise parameters. Water Resource Res., 31(11): 2709-2719. Dougherty, D. E. and R. A. Marryott (1991). Optimal groundwater management (I): Simulated Annealing. Water Resources Research, 27(10): 2493-2508. Eaton, A. D., L. S. Clesceri, A. E. Greenberg, and M. A. H. Franson (1998). Standard methods for the examination of water and wastewater (20th ed). American Public Health Association, Washington D. C., USA. Eaton, J. W. and J. B. Rawlings (1992). Model predictive control of chemical processes. Chem. Eng. Sci., 47:705-720. Ebeling, J. M (1995). Engineering design and construction details of distributed monitoring and control systems for aquaculture. In: Proceedings from the Aquaculture Expo VIII and Aquaculture in the Mid-Atlantic Conference, NRAES-90, Washington D.C., USA, pp. 1-22. Egardt, B. (1979). Stability of adaptive controllers. Springer-Verlag, New York, USA. Ekinci, K. (2001). Theoretical and experimental studies on the effects of aeration strategies on the composting process. Ph.D. Dissertation, the Ohio State University, Columbus, USA. Emmanouilides, C. and L. Petrou (1997). Identification and control of anaerobic digesters using adaptive, online trained neural networks. Comput. Chem. Eng., 21(1): 113-143.

268

Enbutsu, I. (1993). Integration of multi AI paradigms for intelligent operation support systemsfuzzy rule extraction from a neural net work. Water Science and Technology, 28(11&12): 333-340. Faucette, B. (2004). Evaluation of environmental benefits and impacts of compost and industry standard erosion and sediment control measures used in construction activities. Ph.D. Dissertation, University of Georgia, Atlanta, USA. Ferrer, J., M. A. Rodrigo, A. Seco, and J. M. Penya-roja (1998). Energy saving in the aeration process by fuzzy logic control. Water Science and Technology, 38(3): 209-217. Ferrero, G. L. (1996). The thermie programme and composting projects in the energy from biomass and waste sector. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 15-22. Finger, S. M., R. T. Hatch, and T. M. Regan (1976). Aerobic microbial growth in semisolid matrices: heat and mass transfer limitation. Biotechnology and Bioengineering, 18: 1193-1218. Finstein, M. S. and J. A. Hogan (1992). Integration of composting process microbiology, facility structure and decision-making. In: Proceedings of the International Composting Research Symposium, Renaissance Publications, Columbus, USA. Finstein, M. S. and M. L. Morris (1975). Microbiology of municipal solid waste composting. Applied Microbiology, 19: 113-151. Franklin, G. F., J. D. Powell, and A. Emami-Naeini (1991). Feedback control of dynamic systems (2nd Ed.). Addison Wesley, New York, USA.

269

Fromme, W. R. (1999). Characterization of changes occurring in natural organic matter during the composting of a synthetic compost and a municipal solid waste. Ph.D. Dissertation, University of Cincinnati, Cincinnati, USA. Fu, K. S. (1971). Learning control systems and intelligent control systems: an intersection of artificial intelligence and automatic control. IEEE Transactions on Automatic Control, 16: 70-72. Garcia, C. E., D. M. Prett, and M. Moran (1989). Model predictive control: theory and practice - a survey. Automatica, 25: 335-348. Ghiorse, W. C. (1996). Cryptosponidium and Giardia: Update. In: Animal Agriculture and the Environment: Nutrients, Pathogens and Community Relations, ed. M. Sailus, NRAES -96,Rochester, USA, pp. 56-59. Ghomshei, M. M., and J. A. Meech (2000). Application of fuzzy logic in environmental risk assessment: some thoughts on fuzzy sets. Cybernetics and Systems: An International Journal, 31 (3): 317-332. Gnaebe, S. F. and G. C. Goodwin (1992). Adaptive PW design exploiting partial prior information. In: IFAC Adaptive Systems in Control and Signal Processing, Grenoble, France, pp. 371-376. Godden, B. (1983). Evolution of enzyme activities and microbial populations during composting of cattle manure. European Journal of Applied Microbiology and Biotechnology, 17: 306-310. Gokhale, V. (1995). A comparison of advanced distillation control techniques for a propylene-propane splitter. Industrial & Engineering Chemistry Research, 34: 4413-4419.

270

Goldberg, D. E. (1989). Genetic Algorithms in Search, Optimization, and Machine Learning. Addison Wesley, New York, USA. Goldstein, J. (1996). State of the art and perspectives for composting in the United States of America. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 714-721. Golueke, C. G. and L. F. Diaz (1987). Composting and the limiting factor principle. BioCycle, 28: 22-25. Golueke, C. G. and L. F. Diaz (1996). Historical review of composting and its role in municipal waste management. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi,B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 3-13. Gontarski, C. A., P. R. Rodrigues, M. Mori, and L. F. Prenem (2000). Simulation of an industrial wastewater treatment plant using artificial neural networks. Computers and Chemical Engineering, 24: 1719-1723. Gray, K. R., K. Sherman, and A. J. Biddlestone (1971). Review of composting part I. Process Biochemistry, 6(6): 32-36. Gray, K. R., K. Sherman, and A. J. Biddlestone (1971). Review of composting part II. Process Biochemistry, 6(10): 22-28. Griffin, D. M. and E. J. Luard (1979). Water stress and microbial ecology. In: Strategies of Microbial Life in Extreme Environments, ed. M. Shilo, Dahlem Konferenzen, Berlin, Germany, pp. 49-63. Gu, D. and H. Hu (2002). Neural predictive control for a car-like mobile robot.

271

International Journal of Robotics and Autonomous Systems, 39(2&3): 1-15. Guclu, R. (2005). Fuzzy logic control of seat vibrations of a non-linear full vehicle model. Nonlinear Dynamics, 40: 2134. Guh, R. S. (2003). Integrating artificial intelligence into on-line statistical process control. Quality and Reliability Engineering International, 19: 1-20. Hagan, M., H. Demuth, and M. Beale (1996). Neural network design. PWS Publishing, Boston, USA. Hall, S. C. and M. Aneshansley (1997). Use of a programmable logic controller to optimize milk flow rate in an on-farm instant milk cooler. In: Proceedings of Northeast Agricultural and Biological Engineering Conference, Maryland, USA, pp. 13-16. Hall, S. G. (1998). Temperature feedback and control via aeration rate regulation in biological composting systems. Ph.D. Dissertation, Cornell University, Ithaca, USA. Hamelers, H. V. M. (2004). Modeling composting kinetics: A review of approaches. Reviews in Environmental Science and Biotechnology, 3(4): 331-342. Hamelers, H. V. M. (1992). A theoretical model of composting kinetics. In: Proceedings of the International Composting Research Symposium, eds. Hoitinck, H.A.J. & Keener, H.M., Renaissance Publications, Columbus, USA, pp. 36-58. Hamelers, H. V. M. (2001). A mathematical model for composting kinetics. Ph.D. Dissertation, Wageningen University, Wageningen, the Netherlands. Hansen, R. C., H. M. Keener, C. Marugg, C. A. Dick, and H. A. J. Hoitink (1993). Composting of poultry manure. In: Science and Engineering of Composting:

272

Design, Environmental, Microbiological and Utilization Aspects, ed. Hoitink, H.A.J., Renaissance Publications, Columbus, USA, pp. 131-153. Hanss, M. (2002). The transformation method for the simulation and analysis of systems with uncertain parameters. Fuzzy Sets and Systems, 130: 277-289. Harremoes, P. and H. Madsen (1999). Fiction and reality in the modeling world- balance between simplicity and complexity, calibration and identifiability, verification and falsification. Water Science and Technology, 39(9): 1-8. Harris, C. J. (1994). Advances in intelligent control. Taylor & Francis Inc, Philadelphia, USA. Harris, C. J., M. Brown, and C. G. Moore (1993). Intelligent control: aspects of fuzzy logic and neural networks. World Scientific Press, London, UK. Haug, R. T. (1980). Compost engineering: principles and practice. Ann Arbor Science, Ann Arbor, USA. Haug, R. T. (1986). Composting process design criteria part I: feed conditioning. BioCycle, 27: 38-43. Haug, R. T. (1993). The practical handbook of compost engineering. Lewis Publishers, Boca Raton, USA. Haug, R. T. (1996). Composting plant design and process management. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 60-70. Hauhs, M. (1996). Summary of a workshop on ecosystem modeling: The end of an era. The Science of the Total Environment, 183(1-2): 1-5. Haykin, S. (1994). Neural networksa comprehensive foundation. Macmillan College

273

Publishing Company, New York, USA. Heij, C. and J. C. Willems (1989). A deterministic approach to approximate modeling. In: From data to model, ed. J. C. Willems, Springer Verlag, Berlin, Germany, pp. 49-134. Heijnen, J. J. and J. A. Roele (1981). A macroscopic model describing yield and maintenance relationship in aerobic fermentation processes. Biotechnology and Bioengineering, 23: 739-763. Hernandez, E. and Y. Arkun (1990). Neural network modeling and an extended DMC algorithm to control nonlinear systems. AM Contr. Conf., 1: 2454-2459. Hoffmann, M. (2005). Numerical control of kohonen neural network for scattered data approximation. Numerical Algorithms, 39: 175186. Hogan, J. S., K. L. Smith, D. A. Todhunter, and P. S. Schoenberger (1990). Bacterial counts associated with recycled newspaper bedding. Journal of Dairy Science, 73: 1756-1761. Hoitink, H. A. J. and H. M. Keener (1993). Science and engineering of composting: design, environmental microbiological and utilization aspects. Renaissance Publications, Columbus, USA. Hoitink, H. A. J., M. J. Boehm, and Y. Hadar (1992). Mechanisms of suppression of soilborn plant pathogens in compost amended substrates. In: Proceedings of the International Composting Research Symposium, Renaissance Publications, Columbus, USA, pp. 154-167. Holland, J. H. (1975). Adaptation in Natural and Artificial Systems. University of Michigan Press, Ann Arbor, USA.

274

Horikawa, S., T. Furuhashi, and Y. Uchikawa (1992). On fuzzy modeling using fuzzy neural networks with the back-propagation algorithm. IEEE Trans. Neural Networks, 3(5): 801-806. Hornik, K., M. Stinchcombe, and H. White (1989). Multi-layer feed-forward networks are universal Approximators. Neural Networks, 2, 359-366. Hsu, K., H. V. Gupta, and S. Sorooshian (1995). Artificial neural network modeling of the rainfall runoff process. Water Resources Research, 31(10): 2517-2530. Huang, G. H. (1994). Grey mathematical programming and its application to municipal solid waste management planning. Ph.D. Dissertation, McMaster University, Hamilton, Canada. Huang, G. H., and R. D. Moore (1993). Grey linear programming, its solving approach, and its application. International Journal to Systems Science, 24: 159-172. Huang, Y. C. and X. Z. Wang (1999). Application of fuzzy causal networks to waste water treatment plants. Chemical Engineering Science, 54: 2731-2738. Huang, Y. F. (2004). Development of environmental modeling methodologies for supporting system simulation, optimization and process control in petroleum waste management. Ph.D. Dissertation, University of Regina, Regina, Canada. Huang, Y. L. and L. T. Fan (1993). A fuzzy-logic based approach to building efficient fuzzy rule based expert system. Computers Chem. Eng., 17: 181-192. Hunt, K. J., D. Sbarbaro, R. Zbikowski, and P. J. Gawthrop (1992). Neural networks for control system- a survey. Automatica, 28: 1083-1112. Hussain, M. A. (1999). Review of the applications of neural networks in chemical process controlsimulation and online implementation. Artificial Intelligence in

275

Engineering, 13: 55-68. Inbar, Y. (1990). New Approaches to compost maturity. BioCycle, 31: 64-69. Incropera, F. P., and D. P. DeWitt (1985). Fundamentals of heat and mass transfer. John Wiley and Sons, New York, USA. Iriarte, M. L. and P. Ciria (2001). Performance characteristics of three aeration systems in the composting of sheep manure and straw. Journal of agricultural engineering research, 79 (3): 317-330. Jacobowitz, L. A., and T. S. Steenhuis (1984). Compost impact on soil moisture and temperature. BioCycle, 25: 56-60. Jamishidi, M. (2003). Tools for intelligent control: Fuzzy controllers, neural networks and genetic algorithms. Philosophical Transactions Mathematical, Physical & Engineering Science, 361: 1781-1808. Jang, J. S. (1992). Self-learning fuzzy controllers based on temporal back propagation. Neural Network, 3: 723-741. Jeris, J. S. and R.W. Regan (1973a). Controlling environmental parameters for optimum composting (I): experimental procedures and temperature. Compost Science, 14 (1): 10-15. Jeris, J. S. and R. W. Regan (1973b). Controlling environmental parameters for optimum composting (II): moisture, free air space and recycle. Compost Science, 14(2): 8-15. Jeris, J. S. and R. W. Regan (1973c). Controlling environmental parameters for optimum composting (III). Compost Science, 14(3): 16-22. Johannessen, G. S., C. E. James, H. E. Allison, D. L. Smith, J. R. Saunders, and A. J.

276

Mccarthy (2005). Survival of a shiga toxin-encoding bacteriophage in a compost model. FEMS Microbiology Letters, 245 (2): 369-377. Kashmanian, R. M. (1995). Poultry industry finds added value in composting. BioCycle, 36(1): 55-57. Kashmanian, R. M, and R. F. Rynk (1996). Agricultural composting in the United States: trends and driving forces. In: The Science of composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 648-659. Kasparian, V. and C. Batur (1998). Model reference based neural network adaptive controller. ISA Transactions, 37: 21-39. Keener, H. M., C. Marugg, R. C. Hansen, and H. A. J. Hoitink (1992). Optimizing the efficiency of the composting process. In: Proceedings of the International Composting Research Symposium, eds. H. A. J. Hoitink and H. Keener, Renaissance Publications, Columbus, USA, pp. 59-94. Keller, J. M., R. Krishnapuram, and F. C. H. Rhee (1992a). Evidence aggregation networks for fuzzy logic interface. Neural Networks, 3(5), 761-769. Keller, J. M., R. R. Yager, and H. Tahani (1992b). Neural network implementation of fuzzy logic rules with neural networks. Int. J. Approximate Reasoning, 6: 221-240. Khalid, M. (1993). Adaptive fuzzy control of a water bath process with neural networks. Engg. Applic. Artif. Intell., 7(1): 39-52. Kishimoto, M., C. Preechaphan, T. Yoshida, and H. Taguchi. (1987). Simulation of an aerobic composting of activated sludge using a statistical procedure. MIRCEN

277

Journal of Applied Microbiology and Biotechnology, 3: 113-124. Koelsch, R. (1994). Planning and managing for odor control. In: Proceedings from the Liquid Manure Application Systems Conference, ed. R. Koelsch, NRAES-79, Rochester, USA, pp. 95-108. Kohn-Rich, S. and H. Flashner (2003). Robust fuzzy logic control of mechanical systems. Fuzzy Sets and Systems, 133: 77-108. Kolehmainen, M., H. Martikainen, and J. Ruuskanen (2001). Neural networks and periodic components used in air quality forecasting. Atmospheric Environment 35: 815-825. Koumosis, V. K. and P. Georgiou (1994). Genetic algorithms in discrete optimization of steel truss roofs. Journal of Computing in Civil Engineering, 8: 309325. Koyama, H., K. I. Manabe, and S. Yoshihara (2003). A database oriented process control design algorithm for improving deep-drawing performance. Journal of Materials Processing Technology, 138: 343-348. Krause, T. (2002). Comparing compost standards. Resource Recycling, 21(10): 28-33. Kubota, H. and K. Nakasaki (1991). Accelerated thermophilic composting of garbage. BioCycle, 32: 66-68. Kulkarni, M. S. (2004). A single neural network controller approach for optimal guidance of an agile missile. M.S. Thesis, University of Alabama at Birmingham, Birmingham, USA. Kumar, C. A. (1993). A comparison of control strategies for an aerobic compost reactor. M.S. Thesis, Florida Institute of Technology, Melbourne, USA. Kumar, M., and D. P. Garg (2005). Neuro-fuzzy control applied to multiple cooperating

278

robots. Industrial Robot: An International Journal, 32(3): 234-239. Kuter, G. F., H. A. J. Hoitink, and L. A. Rossman (1985). Effects of aeration and temperature on composting of municipal sludge in a full-scale system. Journal of Water Pollution Control Federation, 57(4): 309-3 15. Lahkim, M. B. and L. A. Garcia (1999). Stochastic modeling of exposure and risk in a contaminated heterogeneous aquifer (I): Monte Carlo uncertainty analysis. Enviro. Eng. Sci., 16(5): 315-328. Lasoff, M. A. (2000). Composting industry. Waste Age, 31(8): 44-53. Lee, J. H., S. Natarajan, and K. S. Lee (2001). A model-based predictive control approach to repetitive control of continuous processes with periodic operations. Journal of Process Control, 11: 195-207. Leskens, M., L. B. M. van Kessel, and O. H. Bosgra (2005). Model predictive control as a tool for improving the process operation of MSW combustion plants. Waste Management, 25: 788798. Leton, T. G. and G. I. Stentiford (1990). Control of aeration in static pile composting. Waste Management and Research, 8(4): 299-306. Li, W. and X. G. Chang (2000). Application of hybrid fuzzy logic proportional plus conventional integral-derivative controller to combustion control of stoker-fired boilers. Fuzzy Sets and Systems, 111: 267-284. Lin, C. L. and R. M. Lai (2002). A novel approach to guidance and control system design using genetic-based fuzzy logic model. IEEE Transactions on Control Systems Technology, 10(4): 600-610. Lin, C. T. (1994). Reinforcement structure/parameter learning for neural-network-based

279

fuzzy logic control systems. IEEE Transactions on Fuzzy System, 2: 46-63. Lin, C. T. and C. S. G. Lee (1991). Neural-network-based fuzzy logic control and decision system. IEEE Transactions on Computers, 40(12): 1320-1336. Liu, L., R. X. Hao, and S. Y. Cheng (2003). A possibilistic analysis approach for assessing environmental risks from drinking groundwater at

petroleum-contaminated sites. Journal of Environmental Informatics, 2(1): 31-37. Ljung, L. and T. Glad (1994). Modeling of dynamic systems. Prentice Hall, New Yersey, USA. Luo, B., I. Maqsood, Y. Y. Yin, G. H. Huang, and S. J. Cohen (2004). Adaptation to climate change through water trading under uncertainty an inexact two-stage nonlinear programming approach. Journal of Environmental Informatics. 2(2): 58-68. Luong, J. H. T. and B. Volesky (1983). Heat evolution during microbial process estimation, measurement, and applications. Advances in Biochemical

Engineering/Biotechnology, 28: 1-40. Lynch, N. J. and R. S. Cherry (1995). Design of passively aerated compost piles: Vertical air velocities between pipes. In: The Science of Composting, eds. M. De Bertoldi, P. Sequi, B. Lemmes and T. Papi, Blackie Academic and Professional, London, UK, pp. 973-982. Ma, J. W. (2001). A neural network approach to real-time pattern recognition. International Journal of Pattern Recognition and Artificial Intelligence, 15(6): 937-947. MacGregor, S. T., F. C. Miller, K. M. Psarianos, and M. S. Finstein (1981). Composting

280

process control based on interaction between microbial heat output and temperature. Applied and Environmental Microbiology, 41 (6): 1321-1330. Macky, B. M. and C. M. Derrick (1986). Elevation of the heat resistance of salmonella typhirmurium by sublethal heat shock. Journal of Applied Bacteriology, 66: 389-393. Maiers, J. and Y. S. Sherif (1985). Applications of fuzzy set theory. IEEE Transactions on Systems Man and Cybernetics, 15: 175-189. Marugg, C. (1993). A kinetic model of the yard waste composting process. Compost Science & Utilization, 1(1): 38-51. McCarty, P. L. (1971). Energetics and bacterial growth. In: Organic Compounds in Aquatic Environments. eds. S. C. Faust and J. V. Hunter, Marcel Dekker, New York, USA, pp. 495-512. McKinley, V. L., J. R.Vestal, and A. E. Eralp (1985). Microbial activity in composting. BioCycle, 26 (7): 47-50. Megan, L. and D. J. Cooper (1995). A neural network strategy for disturbance pattern classification and adaptive multivariable control. Comput. Chem. Eng., 19(2): 171-186. Metropolis, N., A. Rosenbluth, and M. Rosenbluth (1958). Equation of state calculations by fast computing machines. J. Chem. Phys., 21 (6): 1087-1092. Michel, F. C., J. F. Huang, L. J. Forney, and C. A. Reddy (1996). Field scale study of the effect of pile size, turning regime and leaf to grass mix ratio on the composting of yard trimmings. In: The Science of Composting, eds. M. De Bertoldi, P. Sequi, B. Lemmes and T. Papi, Blackie Academic and Professional, London, UK, pp.

281

577-584. Miller W. T., R. S. Sutton, and P. J. Werbos (1990). Neural networks for control. MIT Press, Cambridge, USA. Miller, F. C. (1989). Matric water potential as an ecological determinant in compost a substrate dense system. Microbial Ecology, 18: 59-71. Miller, F. C. (1991). Biodegradation of solid wastes by composting. In: Biological Degradation of Wastes, ed. Martin, A. M., Elsevier Applied Science, London, UK, pp. 1-31. Miller, F. C. (1992). Composting as a process based on ecologically selective factors. In: Soil Microbial Ecology, ed. Metting, F. B., Marcel Dekker, New York, USA, pp. 515-543. Mitscherlich, F. and E. Marth (1984). Microbial survival in the environment: bacteria and rickettsiae important in human and animal health. Springer Verlag, Berlin, Germany. Mjalli, F. S. (2005). Neural network model-based predictive control of liquidliquid extraction contactors. Chemical Engineering Science, 60: 239-253. Molla, A., Fakhru'l-Razi, A., Hanafi, M., and M. Alam (2005). Compost produced by solid state bioconversion of biosolids: a potential resource for plant growth and environmental friendly disposal. Communications in Soil Science and Plant Analysis, 36(11&12): 1435-1447. Montague, G. A., M. J. Willis, M. T. Tham, and A. J. Morris (1991). Artificial neural network based control. International conference on control, 1: 266-271 Montgomery, D. C. (2000). Design and analysis of experiments (5th Ed.). John Wiley &

282

Sons, New York, USA. Moore, C. G. and C. J. Harris (1994). Aspects of fuzzy control and estimation, Advances in Intelligent Control. Taylor & Francis Inc., Philadelphia, USA. Moran, M. and F. Zafiriou (1988). Robust Process Control. Prentice Hall, Englewood Cliffs, USA. Morisaki, N. (1989). Nitrogen transformation during thermophilic composting. Journal of Fermentation and Bioengineering, 67(1): 57-61. Muller, A. and S. M. Libelli (1997). Fuzzy control of disturbances in a wastewater treatment process. Water Research, 31(12): 3157-3167. Murray, R., D. Neumerkel, and D. Sbarbaro (1992). Neural networks for modeling and control of a non-linear dynamic system. In: Proceedings of the 1992 IEEE International Symposium on Intelligent Control, Scotland, UK, pp. 404-409. Muske, K. R. and J. B. Rawlings (1993). Linear model predictive control of unstable processes. Journal of Process Control, 3: 85-96. Mysliwiec, M. J., J. S. VanderGheynst, and M. Rashid (2001). Dynamic volume averaged model of heat and mass transport within a compost biofilter (I): Model development. Biotechnology and bioengineering, 73 (4): 282-295. Nagendra, S., D. Jestin, R. T. Haftka, and L. T. Watson (1996). Improved genetic algorithms for the design of stiffened composite panels. Computers and Structures, 58: 543555. Nakasaki, K. (1987). A new composting model and assessment of optimum operation for effective drying of composting material. Journal of Fermentation Technology, 65(4): 441-447.

283

Nakasaki, K. and I. Akiyama (1988). Effects of seeding on thermophilic composting of household organic waste. Journal of Fermentation Technology, 66(1): 37-42. Nakasaki, K., and M. Shoda (1987). Oxygen diffusion and microbial activity in the composting of dehydrated sewage sludge cakes. Journal of Fermentation Technology, 56(1): 43-48. Nakasaki, K., K. Nag, and O. Ariga (2004). Degradation of fats during thermophilic composting of organic waste. Waste Management & Research, 22(4): 276-282. Nakasaki, K., K. Nag, and S. Karita (2005). Microbial succession associated with organic matter decomposition during thermophilic composting of organic waste. Waste Management & Research, 23 (1): 48-56. Nakasaki, K., M. Shoda, and H. Kubota (1985). Effect of temperature on composting of sewage sludge. Applied and Environmental Microbiology, 50(6): 1526-1530. Nakasaki, K., M. Shoda, and H. Kubota. (1985). Effect of seeding during thermophilic composting of sewage sludge. Applied and Environmental Microbiology, 49: 724-726. Nakaya, I. (1999). Laboratory study of optimizing moisture control in composting by using a water holding amendment. M.S. Thesis, University of Dayton, Dayton, USA. Narendra K. S. and K. Parthasarathy (1990). Identification and control of dynamical systems using neural networks. IEEE Transactions on Neural Networks, 1(1): 4-27. Natsuaki, Y., S. Mukandai, K. Yasuda, and H. Furuta (1995). Application of genetic algorithm for the problem of determining the laying sequence for a continuous

284

girder reinforced concrete floor system. In: Building for the 21st Century, ed. Y.C. Leo, EASEC-5, Gold Coast, Australia, pp. 811816. Nelson, M. I., E. Balakrishnan, and X. D. Chen (2003). A Semenov model of self-heating in compost piles. Process Safety and Environmental Protection, 81 (5): 375-383. Nobuyuki Y., H. Junko, and M. Takashi (2001). Monitoring of the composting process using a mediator-type biochemical oxygen demand sensor. The Analyst, 126(10): 1751-1755. Ogata, K. (1990). Modern control engineering. Prentice Hall, Englewood Cliffs, USA. Oishi, K., M. Tominaga, A. Kawato, Y. Abe, S. Imayasu, and A. Nanba (1991). Application of fuzzy control theory to sake brewing process. Journal of Fermentation and Bioengineering, 72(2): 115-121. Oppenheimer, J. R. (1997). Compost process model development, validation, and simulation to assess moisture and energy management. M.S. Thesis, Cornell University, Ithaca, USA. Ortega, J. G. and E. F. Camacho (1994). Neural network GPC for mobile robot path tracking. Robotics and Computer-Integrated manufacturing, 11: 271-278. Paredes, C., M. P. Bernal, J. Cegarra, A. Roig, and A. F. Navarro (1996). Nitrogen transformation during the composting of different organic wastes. In: Progress in Nitrogen Cycling Studies, eds. O. V. Cleemput et al., Kluwer Academic Publishers, the Netherlands, pp. 121-125. Park, M., O. Singvilay, W. Shin, E. Kim, J. Chung, and T. Sa (2004). Effects of long-term compost and fertilizer application on soil phosphorus status under paddy cropping system. Communications in Soil Science and Plant Analysis, 35 (11&12):

285

1635-1648. Parks, G. T. (1990). An intelligent stochastic optimization routine for nuclear fuel cycle design. Nuclear Technology, 89: 233-246. Patni, N. K. and R. G. Kinsman (1997). Composting of dilute manure slurries to reduce bulk by water evaporation. In: Proceedings of ASAE International Meeting, ASAE, St Joseph, USA. Pereira-Neto, J. T, E. I. Stentiford, and D. D. Mara (1986). Pathogen survival in a refuse sludge forced aeration compost system. In: Effluent Treatment and Disposal, Pengamon Press, Oxford, UK, pp. 373-391. Peres, C. R., R. E. H. Guerra, and R. H. Haber (1999). Fuzzy model and hierarchical fuzzy control integration: an approach for milling process optimization. Computers in Industry, 39: 199-207. Poincelot, R. P. (1977). The biochemistry of composting. In: National conference on Composting of Municipal Residues and Sludges, Information Transfer Inc., Maryland, USA, pp. 348-362. Polprasert, C. (1989). Organic Waste Recycling. John Wiley & Sons, New York, USA. Press, Wm. H., B. Flannery, S. Teukolsky, and Wm. Vettering (1986). Numerical recipes. Cambridge University Press, New York, USA. Procyk, T. J. and E. H. Mamdani (1979). A linguistic self-organizing process controller. Automatica, 15: 15-30. Psichogios, D. M. and L. H. Ungar (1991). Direct and indirect model based control using artificial neural networks. Ind. Eng. Chem. Res., 30: 2564-2576. Rafiq, M. Y., J. D. Mathews, and P. Jgodzinski (2001). Interactive role of the human

286

computer interaction in design. In: Artificial Intelligence in Construction and Structural Engineering. Proceedings of the 8th International Workshop of the European Group of Structural Engineering Applications of Artificial Intelligent, eds. C. J. Anumba, O. O. Ugwu and Z. Ren, Loughborough University, Leicestershire, UK, pp. 3142. Raju, G. V. S., J. Zhou, and R. A. Kisner (1991). Hierarchical fuzzy control. Int. J. Control, 54(5): 1201-1216. Ramchandran, S. and R. R. Rhinehart (1995). A very simple structure for neural network control of distillation. Journal of Process Control, 5(2): 115-128. Rao, C. R. (1952). Advanced statistical methods in biometric research, Wiley, New York, USA, pp. 106-207. Rao, C. R. (1965). Linear statistical inference and its applications, Wiley, New York, USA, pp. 239-301. Rao, C. V. and J. B. Rawlings (2000). Linear programming and model predictive control. Journal of Process Control, 10: 283-289. Rao, C. V., J. C. Campbell, J. B. Rawlings, and S. J. Wright (1997). Efficient implementation of model predictive control of large sheet and film forming processes. In: Proceeding of the American Control Conference, IEEE Press, Piscataway, USA, pp. 2940-2944. Rawlings, G. J. E. (1991). Foundations of Genetic Algorithm. Morgan Kaufman Publisher, San Mateo, USA. Reeves, C. R. (1995). Modern Heuristic Techniques for Combinatorial Problems. McGraw-Hill Book Company, London, UK.

287

Richard, T. L. (1992). Municipal solid waste composting: physical and biological processing. Biomass and Bioenergy, 3(3-4): 163-180. Richard, T. L. (1997). The kinetics of solid state aerobic biodegradation. Ph.D. Dissertation, Cornell University, Ithaca, USA. Richard, T. L. and H. L. Choi (1996). Optimizing the composting process for moisture removal: theoretical analysis and experimental results. In: Proceedings of ASAE International Meeting, ASAE, St. Joseph, USA. Richard, T. L. and L. P. Walker (1989). Temperature kinetics of aerobic solid-state biodegradation. Proceedings of the IBE, 1: A10-A30. Richard, T. L., L. P. Walker, and J. M. Gosset (1999). The effects of oxygen on solid-state biodegradation kinetics. Proceedings of the IBE, 2: A22-A39. Ricker, N. L. (1990). Model predictive control with state estimation. Ind. Eng. Chem. Res., 29: 374-382. Robinson, J., and E. I. Stentiford (1993). Improving the aerated static pile composting method by the incorporation of moisture control. Compost Science and Utilization, 1: 53-69. Rodrigo, M. A., A. Seco, and J. Ferrer (1999). Nonlinear control of an activated sludge aeration process: use of fuzzy techniques for tuning PID controllers. ISA Transactions, 38: 231-241. Rodrigues, M.S., J. M. Lopez-Real, and H. C. Lee (1996). Use of composted societal organic wastes for sustainable crop production. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 447-456.

288

Roig, A. and M. P. Bernal (1996). Effectiveness of the Rutgers system in composting several different wastes for agricultural uses. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 663-672. Rosenfeld, P. E., M. A. Grey, and I. H. Suffet (2004). Compost odor control using high carbon wood ash. Water science and technology: a journal of the International Association on Water Pollution Research, 49(9): 171-178. Ryckeboer, J. and J. Mergaert (2003). Microbiological aspects of biowaste during composting in a monitored compost bin. Journal of Applied Microbiology, 94(1): 127-137. Rynk, R. F., E. A. Johnson, and L. F. Whitney (1991). An expert system assisted control system for a composting process. In: Automated Agriculture for the 21st Century, Proceedings of the December 1991 Symposium, ASAE, St Joseph, USA. Sanchez-Monedero, M. A., M. P. Bernal, A. Roig, J. Cegarna, and D. Garcia (1996). The Effectiveness of the Rutgers system and the addition of bulking agent in reducing N-losses during composting. In: Progress in Nitrogen Cycling Studies, eds. O. van Cleemput, G. Hofman, and A. Vermoesen. Kluwer Academic Publishers, the Netherlands, pp. 133-139. Sanner, R. M. and J. E. Slotine (1995). Structurally dynamic wavelet networks for the adaptive control of uncertain robotic systems. In: Proceedings of the 34th Conf. on Decision & Control, IEEE Press, New Orland, USA, pp. 2460-2467. Sarimimveis, H. and G. Bafas (2003). Fuzzy model predictive control of non-linear processes using genetic algorithms. Fuzzy Sets and Systems, 139: 59-80.

289

Sartai, M., L. Fernandes, and N. K. Patni (1995). Influence zone of aeration pipes and temperature variations in passively aerated composting of manure slurries. Transactions of the ASAE, 38(6): 1835-1843. Sasikuman, K., and P. P. Mujumdar (1998). Fuzzy optimization model for water quality management of a river system. J. Water Resour. Plng. Mgmt., 124(2): 79-88. Schulze, K. L. (1960). Rate of oxygen consumption and respiratory quotients during the aerobic decomposition of a synthetic garbage. Compost Science, 1: 36-40. Schulze, K. L. (1961). Relationship between moisture content and activity of finished compost. Compost Science, 2: 32-34. Schulze, K. L. (1962). Continuous thermophilic composting. Applied Microbiology, 10: 108-122. Sequi, P. (1996). The role of composting in sustainable agriculture. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 23-29. Sesay, A. A. (1998). Aerated static pile composting of municipal solid waste (MSW): A comparison of positive pressure aeration with hybrid positive and negative aeration. Waste Management and Research, 16(3): 264-272. Shi, S. and K. Shimizu (1992). Neuro-fuzzy control of bioreactor systems with pattern recognition. Journal of Fermentation and Bioengineering, 74 (1): 3945-3951. Sinha, R. K. and S. Heart (2002). A cost-effective microbial slurry technology for rapid composting of municipal solid wastes in waste dump sites in India and its feasibility for use in Australia. The Environmentalist, 22 (1): 12-19. Skinner, J. H. (1996). I.S.W.A. Policy in the regard of composting as an integrated system

290

of waste management. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 30-40. Smith, R. and R. G. Eilers (1980). Numerical simulation of aerated sludge composting report to USEPA, Report No. EPA-600/2-80-191. Snell, J. R. (1957). Some engineering aspects of high rate composting of garbage and reuse. Journal of the Sanitary Engineering Division, ASCE, 83(SA1), Proceedings Paper 1178:1-36. Soares, H. M., B. Cardenas, D. Weir, and M. S. Switzenbaum (1995). Evaluating pathogen regrowth in biosolids compost. BioCycle, 36: 70-75. Sohn, S. H., S. C. Oh, and B. W. Jo (2000). Prediction of ozone formation based on neural network. Journal of Environmental Engineering, 126(8): 688-696. Soloway, D. and P. J. Haley (1997). Neural generalized predictive control: A Newton-Raphson implementation. NASA Technical Memorandum, Hampton, USA. Song, B. and A. Koivo (1998). Neural network model based control of a flexible link manipulator. In: Proceedings of the IEEE Int. Conf. on Robotics & Automation, ICRA-98, Leuven, Belgium, pp. 812-817. Song, J. J. and S. Park (1993). Neural model-predictive control for nonlinear chemical processes. Journal of chemical engineering of Japan, 26(4): 347-354. Stentiford, E. I. (1996). Composing control: principles and practice. In: The Science of Composting, eds. M. DeBertoldi, P. Sequi, B. Lemmes, and T. Papi, Blackie Academic and Professional, London, UK, pp. 49-59.

291

Stephanopoulos, G. and C. Han (1996). Intelligent systems in process engineering: A review. Computers Chem. Eng., 20(6&7): 743-791. Stombaugh, D. P. and S. E. Nokes (1996). Development of a biologically based aerobic composting simulation model. Transactions of the ASAE, 39(1): 239-250. Sugeno, M. (1985). Industrial applications of fuzzy control. Elsevier Science, New York, USA. Suler, D. J. and M. S. Finstein (1977). Effect of temperature, aeration, and moisture on CO2 formation in bench-scale, continuously thermophilic composting of solid waste. Applied and Environmental Microbiology, 33(2): 345-350. Sutton, R. and I. M. Jess (1991). A design study of a self-organizing fuzzy autopilot for ship control. Proceedings Institute of Mechanical Engineers, 205: 35-47. Syu, M. J. and B. C. Chen (1998). Back-propagation neural network adaptive control of a continuous wastewater treatment process. Ind. Eng. Chem. Res., 37(9): 3625-3630. Syu, M. J. and J. B. Chang (1997). Recurrent backpropagation neural network adaptive control of penicillin acylase fermentation by arthrobacter viscosus. Ind. Eng. Chem. Res., 36: 3756-3761. Szmidt, R. and C. Fox (2001). Compost processesInterdependencies for process control and compost quality. Acta horticulture, 549: 55-61. Takahashi, Y. (1993). Adaptive predictive control of nonlinear time-varying system using neural network. IEEE international symposium on neural networks, 3: 1464-1468. Tan, Y. (1996). Optimization techniques for the design of a neural predictive controller. Neurocomputing, 10: 83-96.

292

Tan, Y. and R. Keyser (1994). Neural network based adaptive predictive control. In: Advances in Model Based Predictive Control, ed. D. Clarke, Oxford University Press, Oxford, UK, pp. 358-369. Tanaka, K. and M. Sugeno (1992). Stability analysis and design of fuzzy control systems. Fuzzy Sets and Systems, 45: 135-156. Tatsuoka, M. M. (1971). Multivariate Analysis, Wiley, New York, USA, pp. 38-197. Tay, J. H., and X. Zhang (1999). Neural fuzzy modeling of anaerobic biological wastewater treatment systems. Journal of Environmental Engineering, 125(12): 1149-1159. Tay, J. H., and X. Zhang (2000). A fast predicting neural fuzzy model for high-rate anaerobic wastewater treatment systems. Water Research, 34(11): 2849-2860. Tement, K. O. (1995). Model predictive control of an industrial packed bed reactor using neural networks. Journal of Process Control, 5(1): 19-27. Thibault, J., V. VanBreusegent, and A. Cheruy (1990). On-line prediction of fermentation variables using neural networks. Biotechnology and Bioengineering, 36: 1041-1048. Tirumalasetty, S. (2004). Effects of process control strategies on the co-composting of municipal solid waste (MSW) and residual digested. M.S. Thesis, South Dakota School of Mines and Technology, Rapid City, USA. Trefry, M. G., and P. D. Franzmann (2003). An extended kinetic model accounting for nonideal microbial substrate mineralization in environmental samples.

Geomicrobiology Journal, 20(2): 113-129. Tsai, Y. P., C. F. Ouyang, M.Y. Wu, and W. L. Chiang (1993). Fuzzy control of a dynamic

293

activated sludge process for the forecast and control of effluent suspended solid concentration. Water Science Technology, 38(11&12): 355-367. Turkcan, A and M. S. Akturk (2003). A problem space genetic algorithm in multiobjective optimization. Journal of Intelligent Manufacturing, 14: 363-378. Ugwu, O. O. and J. H. M. Tah, (2002). Development and application of a hybrid genetic algorithm for resource optimization and management. Engineering, Construction and Architectural Management, 9(4): 304-317. Uif, S. (1998). Systems analysis of waste management-the ORWARE model, transport and compost sub-models. Ph.D. Dissertation, Swedish university of agricultural sciences, Uppsala, Sweden. Umbers, I. G. and P. J. King (1980). Analysis of human decision making in cement kiln control and implications for automation. International Journal of Man Machine Studies, 12: 11-23. Van Durme, G. P., B. F. McNamara, and C. M. McGinley (1992). Bench-scale removal of odor and volatile organic compounds at a composting facility. Water Environment Research, 64(1): 19-27. Van Lier, J. J. C., J. T. Van Ginkel, G. Straatsma, J. P. G. Gerrits, and L. J. L. D. Van Griensven (1994). Composting of mushroom substrate in a fermentation tunnel: compost parameters and a mathematical model. Netherlands Journal of Agricultural Science, 42(4): 271-292. VanderGheynst, J. S. (1997). Experimentation, modeling and analysis of a high-solids aerobic decomposition process. Ph.D. Dissertation, Cornell University, Ithaca, USA.

294

VanderGheynst, J. S., J. M. Gossett, and L. P. Walker (1996). High-solids aerobic decomposition: pilot-scale reactor development and experimentation. Process Biochemistry, 32(5): 361-375. VanderGheynst J. S., L. P. Walker, and J. Y. Parlange (1997). Energy transport in a high-solids aerobic degradation process: mathematical modeling and analysis. Biotechnology Progress, 13 (3): 238-248. Vigie, P., C. Coma, P. Y. Renaud, C. Chamilothoris, B. Dahhou, and J. B. Pourciel (1990). Adaptive predictive control of a multistage fermentation process. Biotechnology and Bioengineering, 35: 217-223. Vinci, B. J., S. T. Summerfelt, M. B. Timmons, and B. J. Watten (1996). Carbon dioxide control in intensive aquaculture: design tool development. Proceedings of the Aquacultural Engineering Society, 2: 399-419. Walker, J. M. (1993). Control of composting odors. In: Science and Engineering of Composting: Design, Environmental, Microbiological and Utilization Aspects, eds. H. A. J. Hoitink and H. M. Keener, Renaissance Publications, Columbus, USA, pp. 185-218. Walker, R. F., P. T. Jackway, and D. Longstaff (2003). Genetic algorithm optimization of adaptive multi-scale GLCM features. International Journal of Pattern Recognition and Artificial Intelligence, 17(1): 17-39. Wang, C. H. and X. Gao (2000). The research on modeling water-processing system by neural networks. Control and Decision, 15(3): 378-380. Wang, L. X. (1994). Adaptive fuzzy systems and control. Prentice-Hall, Englewood Cliffs, USA.

295

Weiss, S. M. and C. A. Kulikowsii (1991). Computer systems that learn: classification and prediction methods from statistics, neural networks, machine learning and expert systems. Morgan Kaufman Publisher, San Mateo, USA. Whang, D. S. and G. F. Meenaghan (1980). Kinetic model of composting process. Compost Science, 21: 44-46. Widrow, B. and M. Lehr (1990). Thirty years of adaptive neural networks, perceptrons, madaline and backpropagation. Proceeding IEEE, 78(9): 1415-1442. Wilks, S. S. (1960). A combinatorial test for the problem of two samples from continuous distributions. In Contributions to Mathematical Statistics, Edited by T. W. Anderson. Wiley, New York, USA, pp. 137-153. Wilks, S. S. (1962). Mathematical Statistics, Wiley, New York, USA, pp. 20-209. Wilks, S. S. (1963). Statistical inference in geology. In Contributions to Mathematical Statistics, edited by T. W. Anderson. Wiley, New York, USA, pp.112-128. Willis, M. J. (1992). Artificial neural networks in process estimation and control. Automatica, 28(60): 1181-1187. Wood, K. L., K. N. Otto, and E. K Antonsson (1992). Engineering design calculations with fuzzy parameters. Fuzzy sets and systems, 52:1-20. Xi, B. D. (2002). Study of composting technology of municipal solid waste with complex microbial community. Ph.D. Dissertation, Tsinghua University, Beijing, P. R. of China. Yager, R. R. (1992). Implementing fuzzy logic controller using neural network framework. Fuzzy Sets and Systems, 48: 53-64. Ydstie, B. E. (1990). Forecasting and control using adaptive connectionist networks.

296

Comput. Chem. Eng., 14: 583-599. Yu, T. (1998). Comparative of BP and RBF artificial neural networks function on model building of ozonation and biological activated carbon water purification system. China Environmental Science, 5: 393-397 Yu, W. S. (2004). H Tracking-based adaptive fuzzy-neural control for MIMO uncertain robotic systems with time delays. Fuzzy Sets and Systems, 146: 375401. Zadeh, L. A. (1968). Fuzzy algorithms. Information and Control, 12: 94-102. Zamarreno, J. M. and P. Vega (1999). Neural predictive control: application to a highly non-linear system. Artificial Intelligence, 12: 149-158. Zhang, Q. and S. J. Stanley (1999). Real-time water treatment process control with artificial neural networks. Journal of Environmental Engineering, 125(2): 153-160. Zhang, Z. (2000). The effects of moisture and free air space on composting rates. M.S. Thesis, Iowa state university, Ames, USA. Zheng, S. (2001). Development of a decision support system for on-site composting for small farm. M.S. Thesis, University of Delaware, Newark, USA. Zhou, Y. F., S. J. Li, and R. C. Jin (2002). A new fuzzy neural network with fast learning algorithm and guaranteed stability for manufacturing process control. Fuzzy Sets and Systems, 132: 201-216. Zhu, J. Y., M. A. Henson, and B. A. Ogunnaike (2000). A hybrid model predictive control strategy for nonlinear plant-wide control. Journal of Process Control, 10(5): 449-458.

297

Вам также может понравиться