Sei sulla pagina 1di 38

984101

ANAEROBIC PROCESSING OF PIGGERY WASTES: A REVIEW


By D. P. Chynoweth*, A. C. Wilkie**, and J. M. Owens* *Dept. of Agricultural and Biological Engineering **Soil and Water Science Department University of Florida, Gainesville, Florida Written for Presentation at the 1998 ASAE Annual International Meeting Sponsored by ASAE Orlando, Florida July 11-16, 1998

The swine industry is growing rapidly along with the world human population. The trend is toward more concentrated piggeries with numbers in herds in the thousands. Associated with these increased herds are large quantities of wastes, including organic matter, inorganic nutrients, and gaseous emissions. The trend in swine waste management is toward treatment of these wastes to minimize negative impact on the health and comfort of workers and animals and on the atmosphere, water, and soil environments. This review discusses the present and future role of anaerobic processes in piggery waste treatment with emphasis on reactor design, operating and performance parameters, and effluent processing. Piggery wastes, swine wastes, anaerobic treatment, anaerobic digestion, biogas

ANAEROBIC TREATMENT OF PIGGERY SLURRY


D. P. Chynoweth*, A. C. Wilkie**, and J. M. Owens* *Dept. of Agricultural and Biological Engineering **Soil and Water Science Department University of Florida, Gainesville, Florida

A paper presented at the pre-conference session: Management of Feed Resources and Animal Waste for Sustainable Animal Production in Asia-Pacific Region Beyond 2000 Eighth World Conference on Animal Production June 28 - July 4, 1998 Seoul, Korea

ABSTRACT The swine industry is growing rapidly along with the world human population. The trend is toward more concentrated piggeries with numbers of herds in the thousands. Associated with these increased herds are large quantities of wastes, including organic matter, inorganic nutrients, and gaseous emissions. The trend in swine waste management is toward treatment of these wastes to minimize negative impact on the health and comfort of workers and animals and on the atmosphere, water, and soil environments. Treatment of these wastes has traditionally involved land application, lagoons, oxidation ditches, and conventional batch and continuously stirred reactor designs. More sophisticated treatment systems are being implemented, involving advanced anaerobic digester designs, integrated with solids separation, aerobic polishing of digester effluents, and biological nutrient removal. This review discusses the present and future role of anaerobic processes in piggery waste treatment with emphasis on reactor design, operating and performance parameters, and effluent processing. INTRODUCTION The swine industry is growing rapidly as more people in developing countries can afford and acquire a taste for more meat in their diet, including pork. In the past 16 years, the annual global production of swine has increased from 790 million to 926 million with most of that increase occurring in China, India, and other emerging countries (FAO, 1990, 1991, 1995, 1996). In past years, piggeries were small (hundreds or less animals) and wastes were disposed of on the same land used to grow the feed, serving as fertilizer and soil conditioner. The increased demand for pork has resulted in establishment of larger centralized piggeries with herds frequently exceeding 1,000 and sometimes more than 10,000 head (Hatfield et. al, 1998). Wastes from these facilities exceed the capacity for direct land disposal without severe environmental impacts, including odor, attraction of rodents, insects and other pests, and release of animal pathogens, atmospheric methane and ammonia, nitrogen, phosphorus, and other nutrients into ground and surface waters. The characteristics of swine wastes vary with a number of factors, including the age and diet of the pigs, type of housing or confinement, and waste removal and pre-processing (Day & Funk, 1998; USDA, 1992; Zhang & Felmann, 1997). The wastes are either scraped or hydraulically flushed into a holding basin, after which they are treated directly or after solids separation. Commonly used systems for removal of organic matter include aerobic and anaerobic lagoons, oxidation ditches, and anaerobic digestion. More advanced systems include tertiary treatment operations, such as oxidation ponds, aquatic plants, wetlands, and denitrification units. This review addresses the current and future role of anaerobic processes for swine waste management. Anaerobic digestion has been applied in a variety of forms and scales to stabilize the organic matter in these wastes. This process results in effective organic matter and pathogen reduction with production of a useful fuel and compost. Properly operated anaerobic digesters result in reduction of odors associated with these wastes (Wilkie, 1998). Anaerobic treatment may also play an important role in nutrient removal. Following aerobic oxidation of ammonia to nitrate, nitrogen may be removed by anaerobic denitrification. Phosphorus removal may also be enhanced by anaerobic pretreatment, which results in formation of organic acids that enhance phosphorus uptake in aerobic processes. Algae ponds and wetlands have also been applied for effluent polishing and nutrient removal. The review summarizes the characteristics of piggery wastes, the role of anaerobic treatment of organic matter, the role of anaerobic digestion in reduction of pathogens and gaseous emissions, and anaerobic treatment of nitrogen and phosphorus. It also discusses integrated treatment systems that use anaerobic processes and future trends in utilization of anaerobic treatment of piggery wastes.

CHARACTERISTICS OF PIGGERY WASTES

Global Production Worldwide swine populations by region and select countries for the periods of 1979-81, 1989-91, and 1996 are presented in Table 1. The number of pigs in the world are about one billion, or one Table 1. World Population of Swine (1000 Head; FAO, 1990, 1991, 1995, 1996 1995, 1996) REGIONS World Asia Europe North and Cent. America South America Africa COUNTRIES China USA Brazil Germany Mexico Viet Nam France Canada India Denmark Korea Japan Philippines Italy 1979-81 779,506 368,702 173,384 97,327 51,722 10,155 313,660 64,045 34,102 34,468 16,895 -11,472 9,709 9,433 9,669 -9,851 7,712 8,885 1989-91 854,213 433,194 182,930 87,012 52,378 16,522 360,247 54,557 33,643 33,350 15,715 12,225 12,233 10,505 11,193 9,390 8,007 11,673 7,968 9,150 1996 927,354 535,844 166,963 96,644 56,122 21,652 452,198 60,190 35,350 24,698 18,000 17,200 14,523 12,043 11,900 10,709 10,300 10,200 8,941 7,984

per six persons. By region, the largest numbers are in Asia, followed by North and Central America, South America, and Africa. By country, the largest numbers are in China followed by the U.S., Brazil, and Germany. The most significant increases in numbers have occurred in certain emerging countries such as China, Vietnam, Korea, and India while numbers have decreased in several developed countries, including the U.S., Germany, and Italy. These trends may be attributed to a number of factors, such as improved economies resulting in higher meat consumption and export by emerging countries, and reduced meat consumption and environmental regulations in developed countries. Estimates of global production of swine wastes are presented in Table 2. Future Trends in Wastes Production The numbers of swine and associated quantities of wastes are likely to increase greatly over the next several decades due to the projected increase in human population and the trend of developing countries, with the highest rates of population increase, to shift to diets with a higher meat content. This will be offset to some extent by reduced red meat (including pork) consumption by the more developed countries. For example, in the U.S., the market share of red meat has decreased from 74% in 1970 to 59% in 1994 (Zhang & Felmann, 1997). The projected world human population increase of 27% by 2020 (US, 1997) should result in at least the same increase in numbers of swine.

Table 2. World Swine Waste Production by Region (Safley et al., 1992) Region Asia and Far East Eastern Europe Western Europe North America Latin America Africa Oceania Near East and Mediterranean Total . Swine Production Facilities The trend in developed and developing countries is to change from small piggeries, where the feed is grown and wastes are land applied locally, to fewer piggeries with greater numbers per facility and import of feed. This results in large quantities of wastes which must now be treated in order to prevent major environmental impact. Also, odors from larger facilities are objectionable to nearby communities. Swine operations are often separated into feeder pig producers (up to 18 kg; 60 days) and feeder pig finishing (98 kg or larger; 150-160 days). The traditional operators raise a pig from birth to death (farrow to finish). The trend in swine housing is confinement in open feedlots or slanted floor units (Day & Funk, 1998; Zhang & Felmann, 1997). Manure is collected by scraping and land applied (with or without prior treatment), or hydraulically flushed from slanted or slatted floor housing where the diluted waste is stored under the house or transported to storage tanks, lagoons, or other waste treatment systems. In some cases, treated wastewater is reused for flushing. Table 3 indicates that liquid flush systems prevail in developed countries, while dry storage and drylot systems are more common in emerging countries. Physical and Chemical Characteristics Typical values for swine waste characteristics as excreted and as collected in storage tanks and lagoons and of runoff water and sludges from feedlots are presented in Tables 4, 5, 6, and 7 (Day & Funk, 1998; USDA, 1992; Zhang & Felmann, 1997). These data are useful in predicting environmental impact and designing systems for waste treatment. Environmental Impact In past years, swine herds were small and wastes could be applied to land used to produce feed and other crops. In contemporary concentrated piggeries with large herds, wastes may exceed the carrying capacity of local ecosystems and are a potential cause of a number of pollution and health problems related to their organic matter, nutrients, pathogens, odors, dust, and airborne microorganisms (Zhang and Felmann, 1997). Table 3. Global Swine Waste Management System Usage (percent) (Safley et al., 1992) Region Asia Eastern Europe Western Europe Anaerobic Lagoons 1 8 0 Liquid Systemsa 38 39 77 5 Daily Spread 1 0 0 Dry Storage and Drylot 53 52 23 Other Systemsb 0 1 6 Total Manure, Mt/day (wet) 1,663,466 788,722 570,932 345,490 320,415 52,519 24,079 830 3,746,273 Volatile Solids, Mt/day 168,010 77,641 57,664 34,960 32,362 5,286 2,432 84 378,439

North America 25 Latin America 0 Africa 0 Oceania 55 Near East and Mediterranean 0 Global Average 5 a Includes liquid/slurry and pit storage b Includes deep pit stacks, litter, and other

50 8 7 0 32 42

0 2 0 0 0 1

18 51 93 0 68 45

6 40 0 28 0 5

Table 4. Typical Body Mass and Waste Production and Characteristics per day per 1000kg of Swine (Day & Funk, 1998) Parameter Live Weight, kg Total Manure, kg Urine, kg Density, kg/m3 Total Solids, kg Volatile Solids. kg BOD5, kg COD, kg pH TKN, kg* Ammonia-N, kg* Total P, kg Ortho-P, kg Mean 61 84 39 990 11 8.5 3.1 8.4 7.5 0.52 0.29 0.18 0.12 Std. Dev. -24 4.8 24 6.3 0.66 0.72 3.7 0.57 0.21 0.10 0.10 --

Table 5. Production and Characteristics of Fresh Manure by Pigs (Zhang & Felmann, 1997) Gestation Sow and Sow Litter Size, kg 15.9 29.5 68.1 125 170 Manure, kg/day 1.0 1.9 4.4 4.0 14.9 TS, kg/day 0.091 0.18 0.41 0.36 1.36 VS, kg/day 0.077 0.14 0.33 0.30 1.09 BOD5 0.032 0.059 0.14 0.12 0.45 N, kg/day 0.007 0.013 0.031 0.028 0.10 P2O5, kg/day 0.005 0.010 0.023 0.022 0.078 K2O, kg/day 0.005 0.11 0.024 0.022 0.082 Table 6. Swine Waste Characteristics From Storage Tanks Under Slats (USDA, 1992) Component Moisture, % TS, % w.b. VS, % w.b. FS, % w.b. N, g/L Farrow 96.5 3.50 2.28 1.22 3.6 Nursery 96.0 4.00 2.79 1.71 4.8 Finish 91.0 9.00 6.74 2.26 6.3 Breeding 97.0 3.00 1.80 1.20 3.0 Parameter Nursery Growing Finishing Boar 159 5.0 0.45 0.420 0.16 0.035 0.027 0.028

NH4-N, g/L P, g/L K, g/L C:N Ratio

2.8 1.8 2.8 4

4.0 1.6 1.6 3

2.7 2.2 6

1.2 2.1 3

Table 7. Swine Waste Characteristics From Storage/Treatment Facilities (USDA, 1992) Anaerobic Lagoon Feedlot* Component Sludge Supernatant Settled Sludge Moisture, % 92.4 99.8 88.8 TS, % w.b. 7.60 0.25 11.2 VS, % w.b. 4.68 0.12 90.7** FS, % w.b. 2.92 0.13 21.3** BOD5, g/L -0.40 -COD, g/L 64.6 1.2 -N, g/L 3.0 0.35 5.6** NH4-N, g/L 0.76 0.22 4.5** P, g/L 2.7 0.13 2.2** K, g/L 7.6 0.38 10.0** C:N Ratio 8 2 -*Semi-humid climate (76 cm annual rainfall); annual sludge removal **kg/d/1000kg animal weight Runoff Water 98.5 1.5 ----2.0** 1.2** 0.38** 1.10** --

Organic matter is concentrated and undergoes anaerobic decomposition producing odors related to hydrogen sulfide, ammonia, volatile acids, and other compounds. The highly biodegradable organic matter also attracts pests, including insects and rodents. Organic matter may also cause oxygen depletion in surface waters and other undesired effects related to color, turbidity, and taste and odor. When organic matter undergoes decomposition under highly anaerobic conditions, methane (a major greenhouse gas) is released into the atmosphere (Safley et al., 1992; USEPA, 1993). Nutrients each have their own impacts on surface and ground water. Nitrogen may be released as ammonia into the atmosphere, where it acts as a greenhouse gas and contributes to acid rain in its oxidized form. Ammonia may also react with nitrate in the atmosphere to form ammonium nitrate particles which contribute to smog and health problems. High ammonia levels in swine houses may also cause eye irritation, respiratory problems, and illness in workers and animals. The recommended maximum acceptable level for human and animal occupancy is 10 ppm (Morrison et al., 1991). In surface waters, nitrogen in the form of ammonia or nitrates causes blooms of algae and aquatic plants which contribute to eutrophication and their decomposition may lead to anaerobic conditions. These blooms, caused also by phosphorus, may consist of highly toxic algae (Pfiesteria) in brackish waters and have been implicated in kills of fish and other aquatic life, and as a cause of adverse health effects on humans and animals (Lusk, 1998). Ammonia is toxic to life in surface waters (Zhang & Felmann, 1997). Concentrations as low as 0.08 mg/L have been shown to cause trout kills. Runoff from swine raising operations and manure-fertilized fields commonly contains 200-200 mg/L ammonia which is well above the recommended USEPA standard of 0.02 mg/L (USDA, 1992). Nitrates in groundwater may cause significant health problems in human and animal development leading to methemoglobinemia, a disease causing oxygen starvation of developing tissues and possible death. The USEPA drinking water standard for nitrate-N is 10 mg/L (USDA, 1992). At elevated levels, nitrates are also toxic to fish and other aquatic organisms.

Sulfides are generated from degradation of protein and other sulfur-containing compounds in swine wastes. These may be toxic to aquatic organisms and cause odors and toxicity in swine-housing (Donham, 1991; Morrison et al., 1991). Hydrogen sulfide, detectable as an odor at concentrations as low as 0.005 ppm, causes loss of appetite, vomiting, and nausea at 50-500 ppm, and is lethal at 1000 ppm. Swine wastes contain pathogens and coliform bacteria and other microbial indicators of fecal pollution (CAST, 1996; Zhang & Felmann, 1997). Although the pathogens are mainly host specific, certain diseases such as salmonellosis, Q fever, Newcastle disease, histoplasmosis, cryptosporidiosis, and gardiasis may be transmitted by swine waste. Fecal indicator organisms originating from swine wastes make it impossible to distinguish the presence of animal or human fecal pollution. Pathogens in the wastewater may also result in cross-infection of the swine. Dust and bioaerosols from animal feed and manure may cause health problems in animals and workers if not controlled by ventilation and other means (Morrison et al., 1991; Zhang & Felmann, 1997). They may cause infectious and respiratory diseases, reduced immune response, allergies, and discomfort. Odors are a major environmental problem with large piggeries (Davidson et al., 1995; Wilkie et al., 1995a). These are caused by numerous volatile compounds such as ammonia, amines, volatile fatty acids, mercaptans, carbonyls, phenols, and indoles. Odor threshold concentrations are very low and are a major factor limiting location of these facilities. Laws, Regulations, Policy In response to the increased environmental impact of intensive rearing facilities for swine and other livestock, several laws, regulations, and policies designed to protect public health and the environment are being called into enforcement as they apply to these industries in the U.S. (Weitman, 1995). Control of methane, ammonia, and dust emissions may fall under the jurisdiction of agencies charged with enforcing the Clean Air Act which addresses the air quality of the nation. The Clean Water Act regulates animal feeding operations considered to be point sources of pollution by requiring permits issued by the National Pollution Discharge Elimination System (NDPES). Whether a piggery or other operation falls under the jurisdiction of this Act depends on the number of animal units (1000 animal units equals 2,500 swine), type of confinement, days of operation, and nature of the water receiving discharge. For example, Category 1 includes operations with over 1,000 animal units; confines, feeds, or maintains animals for a total of 45 days or more in a year; and does not sustain any crops, vegetative forage, or harvest residues. The NDPES permit for this category stipulates that there must be a storage facility to contain all of the manure plus processing water and runoff from a 25-year, 24-hour storm event. Monitoring is required at least once per year and a plan for nutrient management must be approved and implemented. Smaller facilities must submit plans based on Best Available Technology (BAT) that is economically achievable and Best Conventional Pollutant Control Technology (BCPCT) based on professional judgment. The Clean Water Act also regulates non-point source pollution by requiring states to devise a comprehensive plan that addresses contributors. This involves voluntary adoption of Best Management Practices (BMPs) which are encouraged by educational programs, training, financial and technical assistance, and demonstration projects. The Coastal Zone Act Reauthorization Amendments (CZARA) of 1990 regulates animal operations in 35 states which have coastal area watersheds. These facilities are regulated exclusive of those under the Clean Water Act. These regulations require storage and treatment of wastewater and stormwater, waste treatment, and nutrient management. Existing facilities in the U.S. under these regulations for swine have 100-200 head. Odor problems are becoming more prevalent due to the increased scale of intensive livestock operations and urban encroachment into rural areas; these problems are subject to the common laws of nuisance. Farmers are still protected to some extent from these complaints by the Right-to-Farm laws. Federal and state incentives are alternatives to regulation. For example, the Environmental Quality

Incentives Program (EQIP) provides cost sharing of up to 75% of the costs of pollution prevention practices. Ultimately, the tax payer, consumer, and producer must pay the price for maintenance of environmental quality. The cost of environmental pollution is difficult to assess as the effects are usually indirect and long-term. Whatever the case, human activities, including animal production must strive for sustainablity which will cost more. ANAEROBIC DIGESTION OF ORGANIC MATTER Anaerobic digestion may be defined as the engineered methanogenic anaerobic decomposition of organic matter. This process, occurring naturally in anaerobic environments such as sediments, soils, and animal guts, involves a mixed consortium of different species of anaerobic microoganisms that function in concert to degrade organic matter and complete the carbon cycle for a large fraction of organic matter (Chynoweth, 1996). Non-methanogenic populations depolymerize organic polymers and ferment them to acetate (sometimes via other acids and fermentation products), hydrogen, and carbon dioxide. Different methanogenic bacteria convert acetic acid, hydrogen, and carbon dioxide to methane (Boone et al., 1993; Smith & Frank, 1988). Most, but not all organic matter can be decomposed by this fermentation without chemical or physical pretreatment. Lignin is the major natural compound that is refractory to anaerobic decomposition. Other organics, such as cellulose, may be resistant to degradation when complexed tightly with lignin (e.g., in pine wood) or contained in biomass that contains methanogenic inhibitors (e.g., in eucalyptus wood). Anaerobic digestion has been applied for decades for treatment of domestic sludges, animal wastes, industrial wastes (McCarty, 1992), and more recently for the organic fraction of municipal solid wastes (Chynoweth & Isaacson, 1987). It has also been the subject of research for production of substitute natural gas (SNG) from wastes, energy crops, including terrestrial herbaceous and woody energy crops, and aquatic (freshwater and marine) energy crops (Chynoweth & Pullammanappallil, 1996; Legrand, 1993; Smith et al., 1988; Smith et al., 1992). Whatever the application, anaerobic digestion produces a useful energy form (methane) and a stabilized residue that can be subsequently applied to land as a soil amendment. Currently, its most common applications are treatment of domestic sludges, industrial wastes, and animal wastes. Its wider use has been hampered by the low cost of fossil-based energy, limited regulations on waste processing, a history of process instability, and greater knowledge and popularity of aerobic processes. However, the climate is changing for this technology with incentives to replace fossil fuels with renewable greener energy forms and stricter regulations on management of organic wastes that will require more costly in-vessel systems compared to land application, landfilling, or crude open lagoons.

A typical swine waste treatment system is shown in Figure 1. The wastes are transported directly or after

Figure 1. Illustration of Collection and Management Options for Piggery Wastes concentration into a digestion vessel which may vary in design from a lagoon to a mixed, non-mixed plug-flow, or attached film reactor. The operating temperature may be ambient, mesophilic (35oC), or thermophilic (55oC). The effluent is stored and land applied on a seasonal basis. The supernatant may be further treated for nutrient removal prior to discharge into receiving waters. The biogas is used either directly for heating or for operation of internal combustion engines to run equipment or generate electricity. Feed Characteristics Design and operating parameters of anaerobic treatment systems are largely dependent upon the influent total solids (TS) concentration (Chynoweth, 1987; Chynoweth & Isaacson, 1987). For swine wastes, the excreted concentration is about 10% and becomes diluted with urine and further diluted with flush water (in liquid systems), or concentrated when bedding is used in dry storage systems (Day & Funk, 1998; Sweeten et al., 1981; Zhang & Felmann, 1997). A typical concentration for tanks under slats is 3-4%. For a specific organic loading rate, the hydraulic retention time (HRT) of conventional stirred tank reactor

10

(CSTR) anaerobic digesters increases inversely with the total solids concentration. At very low concentrations (<1-2% TS), attached-film reactors can be used to substantially reduce the HRT by concentrating microorganisms on media to prevent their washout (Wilkie & Colleran, 1989). The characteristics of wastes are influenced more by dilution, storage, and separation practices rather than by diet or other factors (Hatfield et al., 1998). (separation) Wastes may also be separated into solids and liquid fractions by various techniques, including sedimentation, mechanical screening, centrifugation, and pressing. Separation not only segregates the wastes for optimization of digester design, but also facilitates manure handling with pumps, pipelines, and sprayers. As discussed below, it also provides for improved odor control. For digestion, the solids fraction can be treated in a conventional CSTR or plug-flow design and the liquid in a lagoon or a lowHRT attached-film design. Day (1998) and Zhang (1997) reviewed the design, performance, and economics of a variety of separations techniques. In general, the capital and operating costs are high and must be justified by the overall goals and economics of the piggery operation. The vibrating screen separator has been shown to be effective for separating flushed swine waste into solid and liquid fractions. Detailed mass balances are presented by Holmberg (1982) for the effect of screen mesh size and flow rate on separation of organic matter and nutrients. Analysis of the effect of an optimum screening regime (60 mesh and flow rates of 457-685 L/min) by Hill (1984), indicated that about 44% of the biodegradable organic matter and 64% of the total nitrogen passes through the screen. A model for predicting vibrating screen performance is presented by Ngoddy (1974) . (biodegradability) Ultimate biodegradability under anaerobic conditions is determined by long-term incubations and is measured in terms of methane yield and reduction in organic matter. Reported ultimate biodegradability (Bo) is useful in kinetic analyses and has been reported in the range of 0.32 to 0.48 m3 CH4/kg VS (Andreadakis, 1992; Hashimoto, 1984; Hill & Bolte, 1984; Iannotti et al., 1979; Safley & Westerman, 1990). This corresponds to volatile solids (VS) reductions in the range of 40-60%. These parameters are largely dependent upon the composition and digestibility of the feed ration. In full-scale digesters, the ultimate digestibility is seldom achieved because of retention time limitations. Iannotti (1979) reported a detailed analysis of digester feed and effluent properties including organic components (Table 8). Carbohydrates are the major components followed by proteins, lipids, and lignin. The waste composition would be significantly influenced by the animal diet. Lignin is not only refractory to anaerobic degradation, but also reduces availability of other components, especially cellulose. (feed blends and effect of diet blends) Blending of swine wastes with other organic wastes may be attractive, especially with high-solids feedstocks. Swine wastes provide excess nutrients and high-solids feedstocks serve as bulking agents, increasing the solids content of the blends. Studies have evaluated swine waste blended with wheat

Table 8. Digestibility of Organic Matter During Anaerobic Digestion of Piggery Wastes (Iannotti et al., 1979) (Finishing hogs fed 14% protein ration with corn or milo; mesophilic digester with retention time of 15 days).

11

Component TS, % VS, %TS COD, g/L total N, g/L protein, %TS hemicellulose, %TS cellulose, %TS lipids starch lignin

Influent 6.9 82.6 73.8 3.9 19.3 20.1 12.4 14.8 1.6 4.4

% Destroyed 52 60 58 -47 65 64 69 94 3

straw (Llabres-Luengo & Mata-Alvarez, 1987), corn stover (Fujita et al., 1980), algae and water hyacinth (Campos & d'Almeida Duarte, 1992), and sewage sludge (Wong, 1990). Wong (1989) investigated a variety of agro-industrial additives, including cardboard, newspaper, sawdust, and sugarcane wastes. Mixtures of swine waste with sawdust or cardboard gave the highest methane yields. Residues from sugarcane blends with pig wastes exhibited the highest fertilizer value. Wastes from pigs fed different sources of fiber (oat hulls, maize hulls, lupin hulls, maize cobs, soya bean hulls, pea hulls, wheat bran, lucerne stems, and lucerne leaves) exhibited different extents and rates of biodegradability during anaerobic digestion (Stanogias et al., 1985) . Volatile solids destruction ranged from 45% for wastes from the lucerne leaf diet to 80.4% from the maize hull diet. Reactor Designs The goals in selecting an appropriate anaerobic digester design are to maximize volatile solids (VS) conversion and associated methane yields, increase conversion rates and process stability, decrease process energy requirements, and ultimately achieve a reliable system with the lowest possible installation and operating costs. Odor control may also be a primary concern. No single reactor design is suitable for all applications in treatment of piggery wastes. Major factors influencing selection include: chemical characteristics of feed concentration of feed biodegradable matter concentration of feed particulate solids density of raw and digested feed scale of application continuity of feed availability desired products site

The designs most commonly used for treatment of swine wastes are lagoon, batch, fed-batch, completely-mixed, and plug-flow. Several new high-rate designs have been developed to retain solids and microorganisms and are particularly suitable for treatment of dilute wastes from flush systems or liquid fractions of separated wastes (Wilkie & Colleran, 1989). The principles of these various reactor designs are discussed below, and operating and performance data for several different reactor configurations are summarized in Table 9.

12

Table 9. Operating and performance data for different digester designs (CSTR) Reference Operational Data Reactor type Volume, m3 Temp, oC Type of waste Infl. TS, % w.b. Infl. VS, % w.b. Infl. COD, g L-1 OLR, kg VS m-3 d-1 OLR, kg COD m-3 d-1 HRT, d Performance Data MY, m3 kg VS added-1 MY, m3 kg COD added-1 MPR, m3 m-3 d-1 VS redn., % COD redn., % Effl. N, g L-1 Effl. P, g L-1 Gas Quality, % CH4 (Hashimoto, 1983) CSTR 0.004 55 whole 6.36% 5.04% 52.10 10.08 5.00 (Hashimoto, 1983) CSTR 0.004 35 whole 6.36% 5.04% 52.10 10.08 5.00 (Stevens & Schulte, 1979) CSTR 2.500 22.5 whole 5.48% 3.57% 74.30 1.80 20.00 (Mills, 1977) (Fischer et al., 1979) CSTR 140.000 35 flushed farrowfinish 1.50% 74.30 1.30 7.43 10.00 (Zhang et al., 1990) CSTR 199.000 35 Whole 3.00% 2.38% 41.35 1.70 14.00 (Iannotti et al., 1979) CSTR 0.420 35 finishing 6.88% 5.69% 0.07 3.78 15.00

CSTR 13.500 35

4.30%

0.31 3.12 50.2 3.34 61.1

0.26 2.69 43.7 3.29 64.2

0.29 0.14 0.52 22.4 35.7 3.66 63.0

0.79 60.0 51.0 2.24 0.38 60.0

0.57 66.0 73.0 2.25 64.0

0.52 60.0 58.0 3.97 59.0

Table 9. Operating and performance data for different digester designs (CSTR and two-phase) Reference (van Velsen, 1977) (van Velsen et al., (Petersen, 1982) 1979) 13 (Cavallero & Genon, 1984) (Maekawa et al., 1995)

Operational Data Reactor type Volume, m3 Temp, oC Type of waste Infl. TS, % w.b. Infl. VS, % w.b. Infl. COD, g L-1 OLR, kg VS m-3 d-1 OLR, kg COD m-3 d-1 HRT, d Performance Data MY, m3 kg VS added-1 MY, m3 kg COD added-1 MPR, m3 m-3 d-1 VS redn., % COD redn., % Effl. N, g L-1 Effl. P, g L-1 Gas Quality, % CH4

CSTR 0.240 32 whole 7.50 5.40 80.30 4.50 6.70 12.00

CSTR 6.000 30 whole 5.83 4.39 51.75 2.64

CSTR 40.000 37 whole 5.53 2.12 27.00

CSTR 0.050 35 supernatant 2.30 1.38 32.09 2.57 12.50

two-phase 1.300 38 whole diluted 4.00

4.74 6.75

0.35 0.90 38.3 40.3 2.32 76.3

0.29 0.61 46.0 0.72 34.2 1.41 41.2

63.3

78.5

60.1

Table 9. Operating and performance data for different digester designs (attached-film) Reference (Chou et al., 1997) (Bolte et al., 1986) (Bolte et al., 1986) (Nordstedt & Thomas, 1985) AF, oak wood blocks 0.005 (Hill & Bolte, 1988) AF, polyester felt 0.300 (Hill & Bolte, 1986) AF, nylon mesh 0.300

OPERATIONAL DATA Reactor type AF upflow immobile AF, nylon mesh & AF, nylon mesh & cells polyurethane foam polyurethane foam Volume, m3 0.014 0.005 0.005 14

Temp, oC Type of waste Infl. TS, % w.b. Infl. VS, % w.b. Infl. COD, g L-1 OLR, kg VS m-3 d-1 OLR, kg COD m-3 d-1 HRT, d PERFORMANCE DATA MY, m3 kg VS added-1 MY, m3 kg COD added-1 MPR, m3 m-3 d-1 VS redn., % COD redn., % Effl. N, g L-1 Effl. P, g L-1 Gas Quality, % CH4

37 screened, diluted

7.50 7.50 1.00

35 flushed, screened 1.33 1.03 16.11 3.44 3.00

55 31.1 35 flushed, screened settled supernatent flushed screened 1.42 2.02 1.88 1.13 1.43 1.50 17.80 33.47 19.69 11.34 7.22 7.50 1.00 2.00 2.00

35 flushed screened 1.89 1.56 21.10 7.80 2.00

0.30 0.12 0.90 61.0 1.03 46.5 49.7 0.97 71.0

0.22 2.43 40.6 34.5 0.98 67.7

0.23 1.68 38.4 51.7

0.37 0.29 2.80 51.0 46.2 0.87 62.5

0.27 0.20 2.15 42.9 42.1 0.82 60.5

77.0

79.2

Table 9. Operating and performance data for different digester designs(attached-film) Reference OPERATIONAL DATA Reactor type Volume, m3 Temp, oC Type of waste Infl. TS, % w.b. (Wilkie & Colleran, 1986) (Wilkie & Colleran, 1986) (Sorlini et al., 1990) (Sorlini et al., 1990) AF, PVC 0.015 30 whole 0.62 (Sorlini et al., 1990) AF, expanded clay 0.015 30 whole 0.62 (Hasheider & Sievers, 1983) AF, limestone 0.003 35 screened diluted

AF, polypropylene AF, polypropylene AF, wood chips rings rings 2.800 2.800 0.015 25 35 30 settled settled supernatent whole supernatent 1.33 1.45 0.62 15

Infl. VS, % w.b. Infl. COD, g L-1 OLR, kg VS m-3 d-1 OLR, kg COD m-3 d-1 HRT, d PERFORMANCE DATA MY, m3 kg VS added-1 MY, m3 kg COD added-1 MPR, m3 m-3 d-1 VS redn., % COD redn., % Effl. N, g L-1 Effl. P, g L-1 Gas Quality, % CH4

0.87 25.20 8.40 3.00

0.96 29.60 9.90 3.00

0.40 5.73 0.73 7.89

0.40 5.73 1.18 4.84

0.40 5.73 0.54 10.71

0.60 6.00 1.00

0.21 0.17 1.47 52.0 87.0 0.22 2.18 60.0 0.21 60.0

0.14

0.03

0.23

0.21 64.7

0.21 13.5

0.90 41.9 0.41

87.0

72.0

77.0

68.0

71.0

Table 9. Operating and performance data for different digester designs(attached-film and UASB) Reference OPERATIONAL DATA Reactor type Volume, m3 Temp, oC Type of waste Infl. TS, % w.b. Infl. VS, % w.b. Infl. COD, g L-1 OLR, kg VS m-3 d-1 OLR, kg COD m-3 d-1 (Ng & Chin, 1988) (Wilkie & Colleran, 1984) AF, clay 0.180 33 settled supernatent 1.51 30.05 5.00 16 (Lo et al., 1994) (Foresti & de Oliveira, 1995) UASB 0.011 25 screened diluted (Owens, 1988)

AF, activated carbon 0.004 na whole 0.40 7.80 14.20

UASB 0.015 25 screened diluted

UASB 0.002 20.7 flushed screened

12.00 3.58

3.73 4.50

8.84 4.40

HRT, d PERFORMANCE DATA MY, m3 kg VS added-1 MY, m3 kg COD added-1 MPR, m3 m-3 d-1 VS redn., % COD redn., % Effl. N, g L-1 Effl. P, g L-1 Gas Quality, % CH4

0.75

6.00

3.28

0.83

2.00

0.05 0.46 78.0 0.23 84.3 0.39 1.93 73.3 0.71 57.0 87.0 0.13 44.4 53.6

87.0

67.0

80.0

Table 9. Operating and performance data for different digester designs (plug-flow, baffle-flow, lagoons) Reference OPERATIONAL DATA Reactor type Volume, m3 Temp, oC Type of waste Infl. TS, % w.b. Infl. VS, % w.b. Infl. COD, g L-1 OLR, kg VS m-3 d-1 OLR, kg COD m-3 d-1 HRT, d [Yang, 1985a #89] (Floyd & Hawkes, 1986) (Boopathy & Sievers, 1991) (Yang & Chou, 1985) (Chandler et al., 1983) (Safley & Westerman, 1988)

plug-flow with recycle 11.500 26 whole diluted 0.79 5.27

Tubular 0.013 30 whole 5.30 3.88

ABR 0.010 35 whole 5.17 3.86 58.50 4.00

ABR 0.020 30 settled supernatent 0.09 1.77 1.75 3.53 0.50

Anaerobic Lagoon 19,000 ambient flushed 0.7

Anaerobic Lagoon

0.11

0.16

1.50

10.00

15.00 17

53-60

PERFORMANCE DATA MY, m3 kg VS 0.14 0.25 0.50 0.23 added-1 MY, m3 kg COD 0.07 0.04 0.11 added-1 MPR, m3 m-3 d-1 0.71 0.96 2.01 VS redn., % 88.7 61.0 COD redn., % 62.0 28.8 Effl. N, g L-1 1.10 0.14 -1 Effl. P, g L Gas Quality, % 67.5 63.0 62.0 CH4 Table 9. Operating and performance data for different digester designs (sequencing batch reactors) Reference OPERATIONAL DATA Reactor type Volume, m3 Temp, oC Type of waste Infl. TS, % w.b. Infl. VS, % w.b. Infl. COD, g L-1 OLR, kg VS m-3 d-1 OLR, kg COD m-3 d-1 HRT, d PERFORMANCE DATA MY, m3 kg VS added-1 MY, m3 kg COD added-1 MPR, m3 m-3 d-1 VS redn., % COD redn., % (Zhang et al., 1997) SBR 0.012 25 screened diluted 0.90% 4.50 2.00 (Zhang et al., 1997) SBR 0.012 25 screened diluted 3.30% 5.50 6.00 1.20 78.00 (Masse et al., 1993) (Hill et al., 1985)

0.04-0.05 75 0.24 69

0.03 (biogas)

SBR 0.025 20 screened 4.80% 3.00% 84.00

SBR 0.454 35 whole, scraped 12.77% 10.82% 0.02 60.00

0.24 1.08 39.0%

0.23 1.28 40.0%

0.76

0.33

56.0% 73.0% 18

46.4%

Effl. N, g L-1 Effl. P, g L-1 Gas Quality, % CH4

0.89 72.0%

2.47 61.0% 63.0% 59.8%

19

(anaerobic lagoon) An anaerobic lagoon is a deep pond (~5 meters) with steep sides which is not aerated and operates largely under anaerobic conditions. The surface may be covered (Chandler et al., 1983; Safley & Westerman, 1989) to facilitate biogas collection and odor reduction. Lagoons operate at ambient temperature, receive dilute swine waste slurries, and serve as reactors for treatment and reservoirs for storage. Biogas production rate from a swine waste lagoon was 0.05 m3/m3 per day and 0.13 m3/m2 per day. Production of biogas was found to be a function of VS loading and lagoon temperature (Cullimore et al., 1985; Safley & Westerman, 1989; Safley & Westerman, 1988). Lagoons may be inexpensive, but require large areas and are often associated with odors. (batch) Batch reactors consisting of large circular or rectangular tanks are fed waste along with an inoculum, and degradation is permitted to startup and proceed to completion. These reactors are often unstable and require careful attention to the inoculum-to-feed ratio; VS conversion is erratic. (completely mixed) The continuously-stirred tank reactor (CSTR) is the most common design used in wastewater and farm applications treating feeds with >3% solids. This design is usually heated, mixed constantly, and usually fed intermittently rather than continuously. The major disadvantage is the loss of inoculum and undigested solids at high loading rates. (anaerobic contact) This design employs a CSTR followed by a settling operation to concentrate washed out microorganisms and undigested solids for recycle back to the CSTR. This results in increased solids retention time (SRT) and reduced digester volume and is used for dilute wastewater applications. (plug-flow) The plug-flow reactor is non-mixed and the feed passes through a trench or cylinder (Floyd & Hawkes, 1986; Gorecki, 1993). These systems are often covered with a balloon plastic cover (Taiwan, 1993). The SRT may also be increased using sludge recycle (Yang and Nagano, 1985). Best performance was obtained at an HRT of 2 days and SRT of 3.25 days. In the baffled modification of this design, internal baffles facilitate mixing and result in extended retention of microorganisms and solids (Boopathy & Sievers, 1991; Yang & Chou, 1985; Yang & Moengangongo, 1987). These designs result in more conversion and higher conversion rates than CSTR or batch reactors. (continuously expanding/sequential batch)(fed-batch) Some batch reactors, such as the continuously expanding (Hill et al., 1985) or anaerobic sequential batch (Dague et al., 1992; Zhang et al., 1997), are intermittently fed, allowing the solids to settle, and supernatant is withdrawn between feeding intervals. Solids are also removed on an intermittent basis. These reactors may or may not be heated (Masse et al., 1993). They promote longer solids than liquid retention times and substantially improve process kinetics over batch and CSTR designs. (attached-film) Several designs use various inert media for attachment of bacteria, forming biofilms, and thus preventing their washout at high hydraulic loadings (Wilkie & Colleran, 1989). These designs are applicable for feedstocks with low suspended solids (<1-2%). Such high-rate systems have permitted reduction of hydraulic retention times to a few days or hours. In the case of anaerobic filters, the packing media may include stones, brick, plastic, tubes, etc. Flow direction may be up or down. Other designs (fluidized or expanded bed) use fine particles such as sand or silica. The waste stream is often recycled to maintain 20

an expanded or fluidized state of the biofilm coated medium. One of the most popular designs, the upflow anaerobic sludge blanket (UASB), takes advantage of the formation of granules consisting of dense consortia of microorganisms which are formed under carefully controlled conditions. Specialized gas/floc separators are employed to prevent washout of these granules. Lo (1994) showed that performance of UASB digesters treating screened swine waste could be improved by incorporation of a rope matrix for attachment of microorganisms in the mid-section of the reactor. Chou (1997) immobilized inoculum in cellulose triacetate for use in a packed bed reactor. Very low hydraulic retention times were achieved in this system without sacrifice in performance. Other successful media for microbial attachment have included nylon cuboid and polyurethane foam (Bolte et al., 1986; Bossier et al., 1986), wood blocks (Nordstedt & Thomas, 1985), nylon mesh (Hill & Bolte, 1986), polypropylene cascade mini-rings(Wilkie & Colleran, 1986), limestone (Hasheider & Sievers, 1983), sand and activated carbon (Ng & Chin, 1988). Wilkie and Colleran (1984) compared several media including clay, coral, mussel shell, and plastic pall-ring support materials, finding similar performance in upflow filters. (high solids) Swine wastes may be mixed with bedding or other wastes, such as the organic fraction of municipal solid wastes, to produce high-solids feedstocks exceeding 20% total solids. Several reactor designs have been recently developed to accommodate high-solids feeds without dilution. One group includes mixed digesters operated at solids concentrations as high as 35%. These include vertical reactors, with feed mixing with inoculum only during feeding, or intermittent mixing of reactor contents during operation as described by Chynoweth (1996). Some designs have horizontal reactors mixed by slow rpm mixers. The sequential batch anaerobic composting design (Chynoweth et al., 1991) uses batch leachbed reactors which are started up by interacting leachate between new and mature reactor solids beds to inoculate the new batch and convey fatty acids to a mature reactor during startup. Others interact leachbeds with methane-phase digesters during startup and operation. Advantages of high-solids systems include reduced odors, easier nutrient management, and reduced reactor size. (staged) Several designs involve one or more stages (usually two) where depolymerization and fermentation to organic acids occur in the first stage and degradation of acids and methanogenesis is accomplished in the second stage, in conventional or high-rate attached-film digesters. Actually in most piggeries, this first acid-forming stage occurs fortuitously during waste storage prior to treatment in anaerobic digesters. There are three major advantages to multi-phase designs (Chynoweth & Pullammanappallil, 1996). The first involves improved stability. In a single, combined-phase digester, overloading and inhibition result in accumulation of volatile organic acids for which resident populations are not present in sufficient numbers to metabolize. Enrichment for these organisms can take several weeks. In a two-phase system, formation of acids is encouraged in the first, or acid phase; therefore, the second methane phase is constantly receiving acids to encourage high populations of acid-utilizing organisms. In other words, the acid-phase is an intentionally imbalanced digester which is resistant to further imbalances resulting from overloading or inhibitors. The second advantage is that the slow-growing populations of microorganisms (acid utilizers and methanogens) can be concentrated in biofilms, thus permitting short retention times for the second-phase reactor. This reduces the overall reactor volume requirement, including both stages. The third advantage is that most of the biogas is produced in the methane-phase digester and the methane content of this gas is higher because of the prior release of much of the carbon dioxide in the acid phase. This advantage facilitates biogas utilization by localizing its production and increasing its methane content. Yang (1995) proposed a three-stage digester design for undiluted pig wastes (TS 8-10%). With an organic loading rate of 2.95 g VS/L per day, a methane yield of 0.42 L/g VS and VS reduction of 69.9% 21

was achieved. Tseng (1992) observed improved digestion of swine wastes employing the first stage reactor to perform solids sedimentation and acidification, and a second tank to perform methanogenesis. Operating Parameters (loading rate) The most meaningful parameter for describing the feed rate is loading rate which is the feed concentration divided by the HRT (Chynoweth & Pullammanappallil, 1996). Loading rate is expressed as weight of organic matter (VS or COD) per culture or bed volume per day (e.g., kg VS/m3/day). This parameter (corrected for head space) describes the reactor volume needed for a particular feed rate. Other parameters, such as solids concentration and retention time (hydraulic or solids), are misleading and do not provide a valid basis for comparison of digester costs. Aside from influencing digester size, solids concentration has a significant effect on digester design and performance, and on materials handling. Feeds with low concentrations of suspended solids (<1-2%) can be digested in high-rate attached-film reactors described above. The conventional method is to use lagoons with low loading rates or conventional digesters following solids separation. Feedstocks with medium solids concentrations (3-10%) require high hydraulic retention times (>15 days for mesophilic temperature) or some mechanism for retaining suspended solids, such as solids recycle or concentration of solids within the reactor, as in the sequential batch, upflow sludge blanket, or baffle-flow designs. In the case of feed blends (discussed below), feed solids may exceed 10%, allowing for the use of highsolids designs with reactor solids concentrations up to 35%. Advantages of these designs include higher loading rates, lower heating energy requirements, and less water as a waste product (Wujick & Jewell, 1980). It has also been shown that cellulolytic enzyme activity per unit reactor volume is higher in highsolids systems (Rivard et al., 1994). High solids systems have a unique set of advantages and limitations with respect to materials handling related to feed addition, mixing, and effluent removal. (start-up) Effective digester startup is dependent upon the quality and quantity of inoculum (Chynoweth & Pullammanappallil, 1996). In conventional CSTR digesters, the inoculum-to-feed ratio (VS basis) is typically greater than 10. In designs where washout of critical organisms is a concern, suspended solids in the effluent may be settled and recycled. With batch and plug-flow designs, inoculum is obtained from previous runs or by effluent recycle. Baffled systems trap inoculum throughout the reactor and inoculate the feed as it passes through. Attached-film reactors often take months to fully start up but have the advantage of inoculum retention during the course of operation. Under-inoculation of a digester results in imbalanced performance due to the more rapid growth of acid formers than methane formers leading to accumulation of organic acids and consequent pH reduction. (temperature) Biological methanogenesis has been reported at temperatures ranging from 2oC (in marine sediments) to over 100oC (in geothermal areas)(Zinder, 1993). Most applications of this fermentation have been performed under either ambient (15 to 25oC), mesophilic (30 to 40oC), or thermophilic (50 to 60oC) temperatures. In general, the overall process kinetics double for every 10oC increase in operating temperature up to some critical temperature (about 60oC) above which a rapid dropoff in microbial activityoccurs (Harmon et al., 1993). The populations operating in the thermophilic range are genetically unique (Zinder, 1993), do not survive at lower temperatures, and are more sensitive to temperature fluctuations outside of their optimum range. Digesters with lower temperatures are more stable and require less process energy, but require larger volumes. Thermophilic digesters have lower volume requirements but have higher energy requirements and are less stable. Ammonia is also more toxic in these digesters because the more toxic free ammonia is favored (Hansen et al., 1998). Typically, most digesters are operated at mesophilic or ambient temperatures. Several researchers have investigated psychrophilic anaerobic digestion of swine wastes (Safley & Westerman, 1990; Stevens & 22

Schulte, 1979; van Velsen et al., 1979; Zeeman et al., 1988). Digestion proceeds at temperatures as low as 10oC, requires longer retention times, and requires a low-temperature inoculum for effective startup. Mesophilic operation seems to be the most preferred because of the possibility for control of temperature fluctuations (not possible for ambient temperature operation) and the higher energy costs for thermophilic digestion. Thermophilic operation is practiced in circumstances when the reduced reactor sizes and the effective pathogen kill justify higher energy requirements and extra effort to ensure stable performance. (nutrients) Nitrogen and phosphorus are the major nutrients required for anaerobic digestion. These elements are building blocks for cell synthesis and are directly related to microbial growth requirements in anaerobic digesters. An average empirical formula for an anaerobic bacterium is C5H7O2NP0.06 (Speece, 1997). Thus, the nitrogen and phosphorus requirements for cell growth are 12% and 2%, respectively, of the volatile solids converted to cell biomass (about 10% of the total volatile solids converted); this would be equivalent to 1.2% and 0.024% of the biodegradable volatile solids, respectively, for nitrogen and phosphorus. Previous studies have identified critical feedstock C/N ratios of 15 for seaweed (Chynoweth et al., 1987) and 15-19 for swine waste (Sievers & Brune, 1978), above which nitrogen was limiting. In fact, nutrient limitations are better related to concentrations; e.g., a value of 700 mg/L was recently reported for the optimum NH4-N concentration in high-solids anaerobic digestion of the organic fraction of municipal solid waste (Kayhanian, 1994). Nutrients may also be concentrated by certain design and operating practices. For example, designs that concentrate solids (Chynoweth, 1987) or reuse supernatant or leachate from process effluent (Chen & Chynoweth, 1990; Chynoweth et al., 1991), concentrate nutrients extracted from the feedstock. Ammonia is also an important contributor to the buffering capacity in digesters but may also be toxic to processing in high solids digesters. Ammonia toxicity was exhibited from feeds that had normal C/N ratios because ammonia became concentrated in the supernatant as digestion proceeded (Jewell et al., 1993). Other nutrients needed in intermediate concentrations, include sodium, potassium, calcium, magnesium, chlorine, and sulfur. Requirements for several micronutrients have also been identified, including iron, copper, manganese, zinc, molybdenum, cobalt, nickel, selenium, and vanadium (Speece, 1997; Wilkie et al., 1986). Available forms of these nutrients may be limiting because of their ease of precipitation and removal by reactions with phosphate and sulfide. Limitations of these micronutrients have been demonstrated in reactors where the analytical procedures failed to distinguish between available and sequestered forms (Jewell et al., 1993). (mixing) Mixing is traditionally thought to be required for optimized digestion to enhance interaction between feed and cells (inoculation) and remove inhibitory metabolic products from the cells. Mixing is also practiced to break up floating scum and foam layers which are typical with some feedstocks, such as domestic sludge. Therefore, conventional digesters include mixing, which is accomplished by mechanical stirring, liquid recycle, or gas recycle. Mixed digesters have been referred to as microbial torture chambers based on research observations (J.G. Ferry, personal communication) that metabolism of certain compounds (e.g., benzoate) is inhibited by mixing and efficient consortia function well in UASB digesters employing granules or other biofilms (Switzenbaum, 1991). One explanation for this inhibition is that microbial consortia existing in clumps are disrupted from that optimum arrangement by mixing. Chynoweth (1987) also demonstrated that a nonmixed solids-concentrating reactor design exhibited more rapid kinetics, lower nutrient requirements, and greater stability than a CSTR design. This improved performance was attributed to reduction in washout of solids and critical organisms. The practice of mixing in swine waste digesters varies depending on the design and was previously addressed under the discussion of reactor options. However, the energy requirement related to mixing can require as much as 14 percent of the methane energy product in conventional low solids designs (see below under energy requirements). (inhibition) 23

Biomethanogenesis is sensitive to several groups of inhibitors, including alternate electron acceptors (oxygen, nitrate, and sulfate), sulfides, heavy metals, halogenated hydrocarbons, volatile organic acids, ammonia, and cations (Speece, 1996). The toxic effect of an inhibitory compound depends upon its concentration and the ability of the bacteria to acclimate to its effects. The inhibitory concentration depends upon different variables, including pH, HRT, temperature, and the ratio of the toxic substance concentration to the bacterial mass concentration. Antagonistic and synergistic effects are also common. Methanogenic populations are usually influenced by dramatic changes in their environment, but can be acclimated to otherwise toxic concentrations of many compounds. Inhibition in anaerobic digesters is reflected by accumulation of volatile acids (related to overloading or toxic feed components), high ammonia levels (related to nitrogenous feeds), or toxic feed components. Normal and inhibited turnover rates of volatile acids are discussed by Winter (1984). When the concentration of total volatile organic acids is in the range of 2,000 mg/L or higher, the onset of imbalance is indicated. Digesters, e.g., acid-phase systems, can acclimate to concentrations as high as 10,000 mg/L. This tolerance is related to alkalinity levels which are influenced by ammonia and bicarbonate. Swine wastes have a high nitrogen content, resulting in high ammonia concentrations during anaerobic digestion. Concentrations in the range of 3,000 mg/L or higher have traditionally been thought to be inhibitory to anaerobic digestion (Braun et al., 1981), especially in combination with high pH that favors the volatile NH3 form. Some researchers have found that acclimated digester cultures can function normally at much higher concentrations, even as high as 6,000 mg/L (Hansen et al., 1998; van Velsen, 1977). Cintoli (1995) evaluated use of zeolite to reduce ammonia in piggery wastes to sub-inhibitory levels (1500 to 300 mg/L) prior to treatment in a UASB digester. Swine and other animals are fed antibiotics and other drugs to prevent infectious diseases and promote growth. In general, results have shown that some antibiotics (e.g., monensin and chlorotetracycline) inhibit digestion (Varel & Hashimoto, 1981; Varel & Hashimoto, 1982) but acclimation is possible, and others (e.g. arsanilic acid, roxarsone, and avilamycin) either stimulated digestion or had no observable effect (Brumm & and Sutton, 1979; Brumm et al., 1980; Brumm et al., 1979; Brumm et al., 1977; Sutton et al., 1989). Camprubi (1988) evaluated different concentrations of antibiotics added to swine waste, including furazolidone, chloramphenicol, chlorotetracycline tylosin, erythromycin, carbadox, and copper sulfate. Chloramphenicol was the most potent inhibitor of anaerobic digestion. Performance Parameters (Gas and Methane Yields, Rates, and Reduction in Organic Matter) Total biogas and methane production, when related to organic matter, are directly influenced by the extent and rate of conversion. Biogas yields are related to organic matter fed which is expressed as volatile solids (VS) or chemical oxygen demand (COD). These data are typically reported as gas volume (m3) per weight (kg) VS or COD added. Methane yield is preferred over gas yield because pH changes in the reactor can cause changes in release or uptake of carbon dioxide that are unrelated to degradation. Use of VS permits calculation of a materials balance between the feed, effluent solids, and gas. Use of COD allows for calculation of an oxidation-reduction balance between the feeds and products. In the context of materials balances, the reduction in organic matter may be calculated as reduction in VS or COD. A typical methane yield for the organic fraction of swine waste is 0.3 m3/kg VS (Table 9) which corresponds to a volatile solids reduction of 50%. Methane production rate is a measure of process kinetics and is determined as volume of methane per volume of reactor per day (m3/m3/day). This parameter is the product of loading rate (kg/m3/day) and methane yield (m3/kg VS added). Values for piggery waste digestion have been reported in the range of 0.04 to 3.0 m3/m3/day. Methane content of the biogas is also a good indicator of stability. Under normal circumstances, this value is a function of the H/C ratio of the biodegradable fraction and is normally in the range of 50-60% 24

(Owens & Chynoweth, 1993). Since lowered methanogenic activity is the key factor leading to imbalance, a reduction of methane gas content is a key performance parameter and has been employed as an online control parameter (Chynoweth et al., 1994). The biochemical methane potential (BMP) assay (Chynoweth et al., 1993; Owen et al., 1979; Owens & Chynoweth, 1993) is useful for estimating the ultimate methane yield and relative conversion rates of feed samples, specific feed components, and remaining biodegradable matter in process residues. This assay may also be used to determine toxicity of feed components. In general, the test is conducted with miniature digesters (200 mL) which are optimized for conversion in terms of inoculum, feed concentration, nutrients, and buffer. These miniature batch digesters are incubated until no further gas production is observed. Measurements include gas production and composition of influent and effluent organic matter. (Organic Acids, pH, and Alkalinity) Organic acids, pH, and alkalinity are related parameters that influence digester performance (McCarty, 1964; WPCF, 1987). Under conditions of overloading and the presence of inhibitors, methanogenic activity cannot remove hydrogen and organic acids as fast as they are produced. The result is accumulation of acids, depletion of buffer, and depression of pH. If uncorrected via pH control and reduction in feeding, pH will drop to levels which stop the fermentation. Independent of pH, extremely high volatile acid levels (>10,000 mg/L) also inhibit performance. The major alkalis contributing to alkalinity are ammonia and bicarbonate. A normal healthy volatile acid-to-alkalinity ratio is 0.1. Increases to ratios of 0.5 indicate the onset of failure and a ratio of 1.0 or greater is associated with total failure. The most common chemicals for pH control are lime and sodium bicarbonate. Lime produces calcium bicarbonate up to the point of solubility of 1,000 mg/L. Sodium bicarbonate adds directly to the bicarbonate alkalinity without reaction with carbon dioxide. However, precautions must be taken not to add this chemical to a level of sodium toxicity (>3500 mg/L). The alkalinity needed to neutralize volatile acids (VFA) is calculated by multiplying 0.833 times VFA concentration (mg/L as acetic acid) (WPCF, 1987). Certain volatile fatty acids are particularly associated with the onset of digester failure, including propionic and higher numbered acids (Gourdon & Vermande, 1987; Hill, 1988; Wilkie and Smith, 1989). Accumulation of these acids results from a backup of hydrogen (or electron) flow and their formation as an alternative to methanogenesis for hydrogen utilization. Useful parameters based on this principle are the ratio of these acids to acetic acid (Hill et al., 1987) and the concentrations of iso-volatile acids (Hill & Holmbert, 1988).

Modeling Numerous models have been developed to provide a theoretical understanding of microbial populations and their interactions with the physical and chemical environment. Models use mathematical expressions to describe the interactions between various microbial populations involved in the process, including substrate utilization rates, microbial growth rates, product formation rates, and physico-chemical equilibrium relationships. More simplified anaerobic digestion models can be used for optimizing process design and operation and for process control. These usually incorporate four major steps, including depolymerization and solubilization, acidogenesis, methanogenesis, and inhibition. The following equations based on the Contois Model have been used to describe the kinetics of anaerobic digestion of swine waste at steady state (Chen, 1983; Hashimoto, 1984):

K B = Bo 1 m 1 + K
25

(1)

= Bo L1

Where:

K m 1 + K

(2)

= methane yield, m3/kg VS Bo = ultimate methane yield at infinite retention time, m3/kg VS K = kinetic parameter (inversely related to digester performance; values <0.6 indicate stability) m = maximum specific growth rate, d-1 v = volumetric methane production rate, m3 CH4/ m3 digester volume per day L = volumetric loading rate, kg VS/m3 digester volume per day

Reported values for ultimate anaerobic biodegradability (Bo) of swine manure in the range of 0.32 - 0.48 m3 CH4/kg VS were summarized from the literature by Chen (1983). Hashimoto (1984) evaluated the effect of temperature and influent substrate concentration on a kinetic constant (K) which is an indicator of digester stability. Temperature had no effect on (K) for influent substrate concentrations between 33.4 and 61.8 g/L. The constant (K) increased exponentially with influent substrate concentration. Chen (1983) further summarized (K) values from the literature which indicated that inhibition of digestion in CSTR digesters can be expected at feed volatile VS concentrations exceeding 6%. Hill (1996) presented a model showing that the ratio of TKN-N in influent-to-effluent is a good indicator of process steady state in livestock anaerobic digestion. Andreadakis (1992) summarized operation, performance, and kinetic parameters from the literature. Other models applicable to wastes (including swine wastes) have design and operating parameters (Hill, 1983), inhibition (Hill et al., 1983), and economics optimization of design and operation (Hill, 1984; Hill, 1985). Most of these models are limited to CSTR digesters and would have to be modified to accommodate other designs such as batch, sequential batch, plug flow, and attached-film digesters. Pathogen Reduction Destruction of human, animal, and plant pathogens during treatment of organic wastes (including swine wastes) is a major concern for any subsequent use of the effluent, such as for land application, recycle, or discharge. Diseases associated with animal manure include bacterial, ricksettial, viral, fungal, and parasitic infections. Most studies have shown that anaerobic digestion results in reduction in numbers of pathogens. This has specifically been shown for swine waste pathogens and indicator organisms in mesophilic digestion (Duarte et al., 1992). Bendixen (1994) has shown that most pathogens were killed under thermophilic conditions (55oC) and that mesophilic temperatures (35oC) did not result in effective reduction. Engeli (1993) has shown that plant pathogens not killed by aerobic treatment were significantly reduced by thermophilic anaerobic digestion. Use of Digester Effluent for Nutrients Digester effluent has been traditionally used as a soil conditioner or good source of inorganic nutrients (Zhang and Felmann, 1997; Day and Funk, 1998). Prior to such use, effluents should be cured or aged to free them of reduced products such as organic acids and reduced inorganic compounds such as ammonia and hydrogen sulfide. These reduced compounds may be toxic to some plant forms. In locations where high concentrations of animals are raised, the capacity of the wastes in terms of inorganic nutrients may exceed the fertilizer needs and local disposal is not possible on a sustained basis without contamination of surface and ground waters. Anaerobic Treatment Of Nutrients

26

A discussion of anaerobic treatment of swine wastes would not be complete without mentioning its role in biological nutrient removal. Per 1000 head in a finishing piggery, 18.6 tons per year (tpy) of nitrogen are ingested and 11.5 tpy excreted; the intake for phosphorus is 4.7 tpy, 4.5 tpy of which is excreted (Kephart, 1996). Anaerobic processes may play an important role in removal of nitrogen and phosphorus (Ekama & Wentzel, 1997). For nitrogen removal, ammonia nitrogen resulting from metabolism of nitrogenous organic compounds must be oxidized aerobically after which nitrogen may be removed anaerobically by denitrification. This is accomplished by recycle of aerobic effluent back through an anaerobic denitrification process. Biological removal of phosphorus is sometimes accomplished by the use of anaerobic prefermenters which produce volatile acids which then enhance uptake of phosphorus by bacteria in subsequent aerobic operations. This may be accomplished in a sequencing batch reactor that is alternated between aerobic and anaerobic conditions (Bortone et al., 1992). Algae, aquatic plants, and wetlands may also be used for nutrient removal from digester effluents (Lincoln & Earle, 1990; Yang & H., 1994; Zhang & Felmann, 1997). Gaseous Emissions Gaseous emissions from piggery wastes are of concern because of their potential health hazards to the animals and farm workers, their contribution to greenhouses gases related to global warming, and as the cause of odors which are objectionable to piggery workers and nearby residents. Over 75 compounds (including ammonia, hydrogen sulfide, volatile organic acids, amines, mercaptans, and heterocyclic nitrogen compounds) have been identified in animal waste emissions which contribute to odors and are largely a result of partial anaerobic decomposition of these wastes (Barth & Melvin, 1984). Under totally aerobic conditions, these compounds would not form as they would be converted to carbon dioxide, water, and oxides of sulfur and nitrogen, all of which are odorless. Under conditions of high concentrations of organic matter (such as animal wastes), oxygen is depleted and an imbalanced anaerobic decomposition occurs giving rise to these products. Many uncontrolled environments where manure is accumulated (manure piles, collection pits or tanks, etc.) have imbalanced anaerobic decomposition (Wilkie et al. 1995b). In a balanced methanogenic decomposition, the gases are limited primarily to methane, carbon dioxide, hydrogen sulfide, and ammonia. Levels of hydrogen sulfide and ammonia are high for swine wastes, compared to other animal wastes, because of the high protein content. As discussed above, ammonia and hydrogen sulfide cause discomfort and represent health risks to confined animals and workers. Methane is produced during the anaerobic decay of organic matter in manure. Worldwide emissions from this source range from 10 to 18 teragrams (Tg) per year, or 2 to 5% of global anthropogenic emissions(Safley et al., 1992; USEPA, 1993). Swine account for 40% of the emissions from animal wastes. No information could be found on the contribution of swine-waste ammonia emissions to atmospheric nitrous oxides. The influence of anaerobic digestion on odors from swine and other animal wastes can be significant (Powers et al., 1997; Wilkie, 1998). Anaerobic treatment is conducted in closed vessels and under conditions that lead to a balanced decomposition. Thus, volatile acids and other reduced compounds found under imbalanced conditions do not accumulate. Organic matter that would lead to production of odors is further decomposed leaving a stable residue that can be applied to fields without generating an odor nuisance. Although the biogas may be odorous due to hydrogen sulfide, the gas is usually enclosed until it is burned or treated for hydrogen sulfide removal prior to use. The problem with most aerobic processes treating animal wastes (e.g. composting and oxidation ditches), is that oxygen becomes limiting and the processes become partially anaerobic leading to volatile gases that are transported to the atmosphere by the aeration process. Other methods to control odors include, solids separation, aeration of anaerobic lagoon surfaces, and ventilation of housing gases through biofilters. Commercial Systems and Economics 27

Sweeten (1981) reviewed the technical and economic considerations for systems for production of methane from swine manure. Using 1980 prices, their construction costs ranged from $22-$36 per 68 kg hog. This is equivalent to $214-$357 per m3 digester volume. They concluded that treatment of concentrated wastes (8-10% TS) would be more economical than treatment of diluted flush wastes (2-3% TS). Chandler (1983) reported that a $89,000 75kw covered lagoon effectively treated wastes from a 1,000 sow (farrow to finish) piggery with an internal return of 34% and a payback period of 3 years. Yang (1995) used lab-scale data to determine the cost of three-stage anaerobic digestion of undiluted pig waste. The cost in Hawaii for treatment with this system was $3.73 per head per year for a 300-herd farm with a profit of $3.01 per head for a 1,000-herd farm. Oleszkiewicz (1983) compared nine full-scale waste treatment schemes for treatment of large-herd piggeries (>10,000 animals) with water-flushed slurry wastes in a developed country. It was found that some of the systems practiced, including extended aeration, chemical coagulation, series of lagoons, and systems featuring land disposal, are not cost effective. Systems with high-rate anaerobic and aerobic unit operations could treat the wastes more effectively at one-third to one/half the cost of more traditional systems. Anaerobic digestion in cost-effective schemes was used for secondary treatment of wastewater and settled sludges and for denitrification. Aerobic treatment was used to polish digester effluent and for biological nutrient removal. Durand (1988) presented the results of an economic system model for anaerobic treatment of confined swine wastes. The best configuration, of 12 evaluated, included flushing manure collection, thermophilic anaerobic digestion with effluent heat recovery, and direct combustion of gas. The economics were most sensitive to digester size, energy price, and efficiency of energy conversion. Table 10 summarizes data from several commercial swine waste digester systems in the U.S., including digester type and volume, gas production, electricity production (if applicable), and capital costs. A software package is available from a U.S. government program (AgSTAR) that facilitates determination of design and economics of different animal waste treatment systems based on anaerobic processes. Table 10. Summary of Economics of Anaerobic Digestion Treatment Systems for Swine Wastes Herd Size 1,150 Feed Type flush Digester Type nonmixed slurry tank covered lagoon plug-flow Dig. Vol., m3 207 Gas Prodn. Elect. Prodn., annual Cost $20,000 (materials); 1,960 m3/d 1,680 700,000k wh 1.0 million kwh 600,000 possible 175 estimated 625,000 $220,000 $525,000 Country US Ref. (Lusk, 1998) (Lusk, 1998) (Lusk, 1998) (Lusk, 1998) (Lusk, 1998) (Lusk, 1998) (Yang & Kuroshim a 1995)

16,500 13,000

flush flush plus dairy proc. wastes flush flush flush scraped (TS 810%)

29,400 1,302

US US

11,500 3,000 ? 1,000

covered lagoon covered lagoon covered lagoon threestage batch

26,180 10,892 30,000 100

980 (1960 in future) 336 40,000 131

$191,500 $85,000 $232,500 $67,558

US US US Hawaii

28

3,200

scraped

storage plus 2 CSTRs CSTR

88

197

supplies all farm energy requiremts .

$62,375

US

(Fischer et al., 1979)

The economics of swine waste treatment systems are highly site specific and dependent upon several factors, including land and labor costs, effluent discharge regulations, and energy prices. INTEGRATED SYSTEMS USING ANAEROBIC PROCESSES With over eight million hogs raised in Taiwan, the Taiwan Livestock Research Institute(Taiwan, 1993) developed a standardized waste treatment system that is used in over 90% of the piggeries. This system involves a complex combination of solids/liquid separation, composting, activated sludge, and anaerobic digestion operations. Redmud plastic-covered plug-flow digesters are employed to treat the overflow dilute fraction. The effluent is polished by activated sludge treatment. Solids are treated by composting. Chou (1995) evaluated automated control of this system. In Denmark, animal wastes are blended with the organic fraction of solid and industrial wastes for anaerobic digestion. The biogas is utilized primarily for heating and generation of electricity (Tafdrup, 1995).

Effluent Processing Fong and Yuen (1986) reported on a lab-scale process for concentrating piggery digester effluent as a potential animal feed. The effluent contained 14% protein. Yang (1994) evaluated a combined fixed-film and aquatic plant system for treatment of diluted piggery waste digester effluent. The system effected 90% COD reduction, 95% reduction of TKN-N, and 99% suspended solids reduction at an estimated cost $2.95 per pig per year for a farm with 1000 head. Camarero (1996) investigated final treatments for effluents from piggery digesters. Coarser fractions separated by flocculation were further digested as high-solids feeds; finer fractions were treated by aerobic digestion and chemical oxidation. De la Noue (1989), Lincoln et al. (1993) and Lincoln et al. (1996) investigated nutrient removal using various microalgae. Chlorella achieved the highest removal rates but Phormidium and Spirulina would be easier to harvest. Effective removal of nutrients by Phormidium was confirmed in laboratory experiments (Canizares-Villanueva et al., 1994; Lincoln & Earle, 1990) Biogas Utilization Figure 2 shows various options for utilization of biogas on farms (Ross & Drake, 1996)). The most efficient use of biogas is direct combustion for heat. Commonly, biogas is used directly with minimal cleanup (hydrogen sulfide and moisture removal) for electrical generation. The heat from generator engines may be captured with heat exchangers for digester heating or other uses. Biogas as produced is typically 60% methane and 40% carbon dioxide, but the methane content can be as high as 80% in attached-film digesters. The heating value is about 14.8-17.8 kJ/m3. The hydrogen sulfide is typically 1%. Iron sponges or iron-impregnated wood chips are often used for hydrogen sulfide removal. Upgrading biogas by removal of carbon dioxide is possible, and necessary for uses requiring compression, but is not economic for the small quantities generated by piggeries. Ross (1996) presents a detailed discussion of properties, conversion, handling and storage, instrumentation and controls, health, safety, and environmental considerations, and economics related to biogas use. . Net Energy Considerations 29

Process energy requirements should influence design and operating conditions selected for treatment of swine wastes and other feedstocks in digesters. For CSTR digesters, the components in order of their relative importance are feed heating, reactor heat losses, and mixing (Chen, 1983; Chen & Hashimoto, 1981; Srivastava, 1987). The total requirements can range from about 10% to over 100% of the methane product energy depending upon the design, feed solids concentration, loading rate, mixing regime, operating temperature, and ambient temperature. Since feed heating is the major requirement, the requirements are high for colder climates and higher digester operating temperatures. There is a strong incentive for ambient temperature operation for high-rate digesters receiving dilute waste streams. This is related to the high energy requirements to heat the dilute feed and the rapid kinetics of these designs at low temperatures. On the other hand, (Legrand, 1993) has calculated that high-solids digesters may be self-heating in tropical and sub-tropical climates. FUTURE OF ANAEROBIC PROCESSES IN SWINE WASTE MANAGEMENT Several demographic trends will influence swine waste management into the twenty-first century: increase in human population increase in swine meat consumption in developing countries decrease in swine meat consumption in developed countries centralization of swine production with herds in the tens of thousands

stricter local and global environmental regulations with respect to gaseous, liquid, and solids emissions public health regulations with respect to animal and worker comfort and health

Biogas Utilization on Farms


Anaerobic Digester Biogas Biogas Store
(low pressure)
2 Scrub HS 2 2 Scrub HS + C0

low pressure storage

high pressure storage

Heating: Stationary engine Transport engine Process heating Mechanical power (hot water,steam,etc.) Electrical power 30 Space heating Dom estic heating Dom estic cooking

Figure 2. Biogas Utilization Options

In developed nations like the United States and the European Union, piggeries will be treated like other industries, with emphasis on clean sustainable animal raising operations. Wastes will be rapidly removed from their site of production to minimize effects on animals and workers and treated with much the same objectives as for human and industrial wastes, i.e., removal of solids, organic matter, nutrients, and pathogens. The trend continues to be toward flushed systems which will be followed by combinations of operations for separation, anaerobic and aerobic organic matter reduction, nutrient removal, and disinfection. Anaerobic treatment should play an important role in future swine waste management for treatment of organics with its minimum biological sludge production, production of a useful methane fuel, emerging developments for biological removal of nitrogen and phosphorus, and its capacity to reduce pathogens. There are strong arguments in favor of minimizing water use in management of these wastes, and in fact to further increase solids by use of straw and other high-solids wastes for bedding. High-solids management not only reduces required reactor sizes, but also odors and water requirements. In developing countries, the trends will be the same, but piggeries will be smaller on the average and emphasis on environmental pollution control will be slower to develop. High level of treatment of piggery wastes will probably coincide with the emergence of effective treatment of human wastes. In general, it makes the most sense to grow swine in the vicinity of their feed production. This would facilitate a sustainable cycle for nutrient management. A modern approach to evaluation of systems involving anaerobic digestion of wastes from swine, or any feedstock, is life cycle assessment (LCA). LCA (IEA, 1997) involves the systematic identification of all materials, energy, and economic inputs and outputs of a system from cradle to grave, i.e., from the extraction of raw materials from the environment to their eventual assimilation back into the environment. The flows are assessed in terms of their potential to contribute to specific environmental impacts. For swine waste systems, this would involve assessment of feed production and processing, waste collection and storage, waste transport, air emissions, water emissions, net energy consumption, compost, and wastes. ACKNOWLEDGEMENTS The authors acknowledge several colleagues who promptly responded by sending recent reviews and papers on the subject matter, including R. Zhang, P. Lusk, and C.Y. Chou. They also gratefully acknowledge Gloria Chynoweth, who located references and entered them into a database. LITERATURE CITED Andreadakis, A. D. (1992). Anaerobic digestion of piggery wastes. Wat. Sci. Technol., 25(1), 9-16. Barth, C. L. & Melvin, S. W. (1984). Agriculture and the environment: an examination of critical issues for food policy - part 4: Odor. Ag. Eng., 65(4), 20-22. Bendixen, H. J. (1994). Safeguards against pathogens in Danish biogas plants. Wat. Sci. Tech., 30(12), 171-180. Bolte, J. P., Hill, D. T. & Wood, T. H. (1986). Anaerobic digestion of screened swine waste liquids in suspended particle-attachment growth reactors. Trans. ASAE, 29(2), 543-549. 31

Boone, D. R., Chynoweth, D. P., Mah, R. A., Smith, P. H. & Wilkie, A. C. (1993). Ecology and microbiology of biogasification. Biomass and Bioenergy, 5, 191-202. Boopathy, R. & Sievers, D. M. (1991). Performance of a modified anaerobic baffled reactor to treat swine waste. Trans. ASAE, 34(6), 2573-2578. Bortone, G., Gemelli, S., Rambaldi, A. & Tilche, A. (1992). Nitrification, denitrification and biologial phosphorus removal in sequencing batch reactors treating piggery wastewater. Wat. Sci. Technol., 26(5-6), 977-985. Bossier, P., Poels, J., Van Assche, P. & Verstraete, W. (1986). Influence of the dimensional characteristics of polyurethane foam on high rate anaerobic digestion of piggery manure. Biotechnol. Lett., 8(12), 901-906.Braun, R., Huber, P. & Meyrath, J. (1981). Ammonia toxicity in liquid piggery manure digestion. Biotechnol. Lett., 3(4), 159-164. Brumm, M. C. & Sutton, A. L. (1979). Effect of copper in swine diets on fresh waste composition and anaerobic decomposition. J. Anim. Sci., 49, 20-25. Brumm, M. C., Sutton, A.L. & Jones, D. D. (1980). Effect of dietary arsonic acids on perfromance characteristics of swine waste anaerobic digesters. J. Anim. Sci., 51, 544-549. Brumm, M. C., Sutton, A. L., Mayrose, V. B. & Krider, J. L. (1979). Effect of dietary arsonic acid on fresh swine waste composition and anaerobic decomposition. J. Anim. Sci., 48, 1305-1311. Brumm, M. C., Sutton, A. L., Mayrose, V. B., Nye, J. C. & Jones, H. W. (1977). Effect of arsanilic acid in swine diets on fresh waste production, composition and anaerobic decomposition. J. Anim. Sci., 44, 521-531. Brumm, M. C., Sutton, A. L., Mayrose, V. B., Nye, J. C. & Jones, H. W. (1977). Effect of arsanilic acid level in swine diets and waste loading rate on model anaerobic lagoon performance. Trans. ASAE, 20, 498-501. Camarero, L., Diaz, J. M. & Romero, F. (1996). Final treatments for anaerobically digested piggery slurry effluents. Biomass and Bioenergy, 11(6), 483-489. Campos, L. & d'Almeida Duarte, E. F. (1992). Increasing of biogas production rate by addition of algae and water hyacinth to pig manure. International Symposium on Anaerobic Digestion of Solid Waste, April 14-17, 1992, Venice, Italy. 469-471. Camprubi, M., Paris, J. M. & Casas, C. (1988). Effects of antimicrobial agents and feed additives on the performance of piggery waste anaerobic treatment. Anaerobic Digestion 1988 - 5th International Symposium on Anaerobic Digestion, May 22-26, 1988, Bologna, Italy. Pergamon Press. 239-248. Canizares-Villanueva, R. O., Corona, A. I., Monroy, O., De La Torre, M., Gomez-Lojero, C. & Travieso, L. (1994). Phormidium treatment of anaerobically treated swine wastewater. Wat. Res., 28(9), 18911895. CAST. (1996). Integrated Animal Waste Management. Council for Agricultural Science and Technology. #128. Cavallero, C. & Genon, G. (1984). Digestion tests of dilute piggery wastes. Ag. Wastes, 11, 227-233. Chandler, J. A., Hermes, S. K. & Smith, K. D. (1983). A low cost 75 kw covered lagoon biogas system. In Energy From Biomass and Wastes VII, ed. D. Klass, Institute of Gas Technology, Chicago, IL. Lake Buena Vista, FL, pp. 627-646. Chen, T., Chynoweth, D. P., & Biljetina, R. (1990). Anaerobic digestion of municipal solid waste in a nonmixed solids-concentrating digester. Appl. Biochem. Biotechnol., 24/25, 533-544. Chen, Y. R. (1983). Kinetic analysis of anaerobic digestion of pig manure and its design implications. Ag. Wastes, 8, 65-81. Chen, Y. R. & Hashimoto, A. G. (1981). Energy requirements for anaerobic fermentation of livestock wastes. In Livestock Waste: A Renewable Resource, ASAE. St. Joseph, MI, pp. 117-121. Chou, C. Y., Hu, Y. Y. & Huang, J. C. (1995). Automatic control for swine wastewater treatment systems. Seventh Intern. Symp. on Ag. and Food Processing Wastes (ISAFPW95), Chicago, Illinois. ASAE. 137-143. Chou, C. Y., Liu, A. C. & Yang, P. Y. (1997). Anaerobic digestion of swine wastewater by applying the immobilized cell. Proc. Internat. Symp. Ag.-Mech. and Automation, Taipai, Taiwan. Chinese Inst. of Agricultural Machinery. 287-292. Chynoweth, D. C., Bosch, G., Earle, J. F. K., Legrand, R. & Liu, K. (1991). A novel process for anaerobic composting of municipal solid waste. Appl. Biochem. Biotech., 28/29, 421-432.

32

Chynoweth, D. C., Svoronos, S. A., Lyberatos, G., Harmon, J. L., Pullammanappallil, P., Owens, J. M. & Peck, M. J. (1994). Real-time expert system control of anaerobic digestion. Water Sci. Tech., 30(12), 21-30. Chynoweth, D. P. (1987). Biomass Conversion Options. Magnolia Publishers, Orlando, Fl, 621-642. Chynoweth, D. P. (1996). Environmental impact of biomethanogenesis. Envir. Monitor. Assess. 42, 3-18. Chynoweth, D. P., Fannin, K. F. & Srivastava, V. J. (1987). Biological Gasification of Marine Algae. In Seaweed Cultivation for Renewable Resources, eds. K. Bird and P. Benson Elsevier Applied Science, New York, 285-303. Chynoweth, D. P. & Isaacson, H. R. eds. (1987). Anaerobic Digestion of Biomass. Elsevier Applied Science, New York. Chynoweth, D. P. & Pullammanappallil, P. (1996). Anaerobic Digestion of Solid Wastes. In Microbiology of Solid Waste, ed. A. C. Palmisano & M. A. Barlaz, CRC Press, Inc. Boca Raton, 71-113. Chynoweth, D. P., Turick, C. E., Owens, J. M., Jerger, D. E. & Peck, M. W. (1993). Biochemical methane potential of biomass and waste feedstocks. Biomass Bioenergy, 5, 95-111. Cintoli, R., Di Sabatino, B., Galeotti, L. & Bruno, G. (1995). Ammonium uptake by zeolite and treatment in UASB reactor of piggery wastewater. Wat. Sci. Tech., 32(12), 73-81. Cullimore, D. R., Maule, A. & Mansuy, N. (1985). Ambient temperature methanogenesis for pig manure waste lagoons: thermal gradient incubator studies. Ag. Wastes, 12(2), 147-57. Dague, R. R., Habben, C. E. & Pidaparti, S. R. (1992). Initial studies on the anaerobic sequencing batch reactor. Wat. Sci. Tech., 26(9-11), 2429-2432. Davidson, K. D., Westermann, P. W., Safley, L. M. & Barker, J. C. (1995). Relative odor levels from swine production facilities and lagoons. Seventh Intern. Symp. on Ag. and Food Processing Wastes (ISAFPW95), Chicago, IL. ASAE. 616-626. Day, D. L. & Funk, T. L. (1998). Processing manure: physical, chemical and biological treatment. In Animal waste utilization: effective use of manure as a soil resource, eds. J. L. Hatfield & B. A. Stewart, Ann Arbor Press. Chelsea, pp. 243-282. De la Noue, J. & Basseres, A. (1989). Biotreatment of anaerobically treated swine manure with microalgae. Biol. Waste, 29, 17-31. Donham, K. J. (1991). Health concerns from the air environment in intensive swine housing: where have we come from and where are we going? Making Swine Buildings a Better Place to Work. Nat. Pork Producers Council and Nat. Pork Board, Des Moines, IA. Duarte, E. A., Mendes, B. & Oliviera, J. S. (1992). Removal of salmonella, streptococci, and coliforms in pig breeding effluent by anaerobic mesophilic digestion. Wat. Sci. Tech., 26(9-11), 2169-2172. Durand, J. H., Fischer, J. R., Jannotti, E. L. & Miles, J. B. (1988). Development of an economic system model to optimize the performance of a swine waste digestion system. Biomass, 14, 45-65. Ekama, G. A. & Wentzel, M. C. (1997). Difficulties and developments in biological nutrient removal technology and modelling. IN Proc. 3rd AWWA/IAWQ BNR Conf, Brisbane, Aus, 3-13. Engeli, H., Edelmann, W., Fuchs, J. & Rottermann, K. (1993). Survival of plant pathogens and weeds during anaerobic digestion. Wat. Sci. Tech., 27, 69-76. FAO. (1990, 1991, 1995, 1996). FAO Food Production Yearbook. FAO. Fischer, J. R., Meador, N. F., Sievers, D. M., Fulhage, C. D. & Iannotti, E. L. (1979). Design and operation of a farm anaerobic digester for swine manure. Trans. ASAE, 22(5), 1129-1136. Fischer, J. R., Osborn, D. D., Meador, N. F. & Fulhage, C. D. (1979). Economics of a swine anaerobic digester. ASAE 1979 Winter Conference, December 11-14, 1979, New Orleans, LA. ASAE,13 pp. Floyd, J. R. S. & Hawkes, R. R. (1986). Operation of a laboratory tubular digester on piggery wastes. Ag. Wastes, 18, 36-60. Fong, J. C. L. S., Morrissette, D. & Yuen, R. T. H. (1986). Integrated production of biogas and protein biomass from piggery manure. Biotechnol. Lett., 8(4), 299-302. Foresti, E. & de Oliveira, R. A. (1995). Anaerobic treatment of piggery wastewater in UASB reactors. Seventh Intern. Symp. on Ag. and Food Processing Wastes (ISAFPW95), Chicago, Illinois. ASAE. 309-19. Fujita, M., Scharer, J. M. & Moo-Young, M. (1980). Effect of corn stover addition on the anaerobic digestion of swine manure. Ag. Wastes, 2(3), 177-184. Gorecki, J., Bortone, G., and Tilche, A. (1993). Anaerobic treatment of the centrifuged solid fraction of piggery wastewater in an inclined plug flow reactor. Wat. Sci. Tech., 28(2), 107-114. 33

Gourdon, R. & Vermande, P. (1987). Effects of propionic acid concentration on ananerobic digestion of pig manure. Biomass, 13, 1-12. Hansen, K. H., Angelidaki, I. & Ahring, B. K. (1998). Anaerobic digestion of swine manure: inhibition by ammonia. Wat. Res., 32, 5-11. Harmon, J. L., Svoronos, S. A., Lyberatos, G. & Chynoweth, D. P. (1993). Adaptive temperature optimization of continuous anaerobic digesters. Biomass and Bioenergy, 5, 279-288. Hasheider, R. J. & Sievers, D. M. (1983). Limestone bed anaerobic filter for swine manure. 1983 ASAE Summer Conference, June 26-29, 1983, Montana State University, Bozeman, MT. Hashimoto, A. G. (1984). Methane from swine manure: effect of temperature and influent substrate concentration on kinetic parameter (k). Ag. Wastes, 9, 299-308. Hashimoto, A. G. (1983). Thermophylic and mesophilic anaerobic fermentation of swine manure. Ag. Wastes, 6, 175-191. Hill, D. T. (1983). Design parameters and operating characteristics of animal waste anaerobic digestion systems - swine and poultry. Ag. Wastes, 5, 157-178. Hill, D. T. (1984). Economically optimized design of methane fermentation systems for swine production facilities. Trans. ASAE, 27(2), 525-529. Hill, D. T. (1988). Long chain volatile fatty acid relationships in anaerobic digestion of swine waste. Biol. Wastes, 23, 195-214. Hill, D. T. (1985). Practical and theoretical aspects of engineering modeling of anaerobic digestion for livestock waste utilization systems. Trans. ASAE, 28(1), 850-855. Hill, D. T. & Bolte, J. P. (1984). Characteristics of screened-flushed swine waste as a methane substrate. 1984 ASAE Conference, June 24-27, 1984, University of Tennessee, Knoxville, TN. ASAE. 1-20. Hill, D. T. & Bolte, J. P. (1986). Evaluation of suspended particle-attached growth ferments treating liquid swine waste. Trans. ASAE, 29(6), 1733-1788. Hill, D. T. & Bolte, J. P. (1988). Synthetic fixed media reactor performance treating screened swine waste liquids. Trans. ASAE, 31(5), 1525-1531. Hill, D. T. & Cobb, S. A. (1996). Simulation of process steady-state in livestock waste methanogenesis. Trans. ASAE, 39(2), 565-573. Hill, D. T., Cobb, S. A. & Bolte, J. B. (1987). Using volatile fatty acid relationships to predict anaerobic digester failure. ASAE Trans., 30, 496-501. Hill, D. T. & Holmbert, R. D. (1988). Long chain fatty acid relationships in anaerobic digestion of swine waste. Biol. Waste, 23, 195-214. Hill, D. T., Prince, T. J. & Holmberg, R. D. (1985). Continuously expanding digestor performance using a swine waste substrate. Trans. ASAE, 28(1), 216-221. Hill, D. T., Tollner, E. W. & Holmberg, R. D. (1983). The kinetics of inhibition in methane fermentation of swine manure. Ag. Wastes, 5, 105-123. Holmberg, R. D., Prince, T. J. & Van Dyke, N. J. (1982). Solid-liquid separation effect on physical properties of flushed swine waste. 1982 ASAE Conference, June 27-30, 1982, University of Wisconsin-Madison. ASAE. 1-26. Iannotti, E. L., Porter, J. H., Fischer, J. R. & Sievers, D. M. (1979). Changes in swine manure during anaerobic digestion. In Developments in Industrial Microbiology, 20, 519-529. IEA. (1997). Life Cycle Assessment of Anaerobic Digestion: A Literature Review. IEA Bioenergy; Energy Recovery from Municipal Solid Waste Task.; Anaerobic Digestion Activity, prepared by Ecobalance UK. Jewell, W. J., Cummins, R. J. & Richards, B. K. (1993). Methane fermentation of energy crops: maximum conversion kinetics and in in situ biogas purification. Biomass Bioenergy, 5, 261-278. Kayhanian, M. (1994). Performance of a high-solids anaerobic digestion process under various ammonia concentrations. J. Chem. Biotech., 59, 349-352. Kephart, K. (1996). Nutrients and swine production. Animal Agriculture and the Environment - North American Conference (Nutrients, pathogens, and community relations), December 11-13, 1996, Rochester, N.Y. Northeast Regional Agricultural Engineering Service. 239-248. Legrand, R. (1993). Methane from biomass systems analysis and CO2 abatement potential. Biomass Bioenergy, 5, 301-316. Lincoln, E. P., Crawford, J. J. W. & Wilkie, A. C. (1993). Spirulina in animal agriculture. Bull. Inst. Oceanogr. (Monaco), 12, 109-115. 34

Lincoln, E. P. & Earle, J. F. K. (1990). Wastewater treatment with microalgae. In Introd. Appl. Phycol., ed. I. Akatsuka, SPB Academic Publ. The Hague, The Netherlands, 429-456. Lincoln, E. P., Wilkie, A. C. & French, B. T. (1996). Cyanobacterial process for renovating dairy wastewater. Biomass and Bioenergy, 10, 63-68. Llabres-Luengo, P. & Mata-Alvarez, J. (1987). Kinetic study of the anaerobic digestion of straw-pig manure mixtures. Biomass, 14(4). Lo, K. V., Liao, P. H. & Gao, Y. C. (1994). Anaerobic treatment of swine wastewater using hybrid UASB reactors. Biores. Technol., 47(2), 153-157. Lusk, P. L. (1998). Methane Recovery from Animal Manures: The 1997 Opportunities Casebook. Prepared for National Renewable Energy Laboratory, Golden, CO. Maekawa, T., Liao, C. M. & Feng, X. D. (1995). Nitrogen and phosphorus removal for swine wastewater using intermittent aeration batch reactor followed by ammonium crystallization process. Wat. Res., 29(12), 2643-2650. Masse, D. I., Droste, R. L., Kennedy, K. J., Patni, N. K. & Munroe, J. A. (1993). Psychrophilic anaerobic treatment of swine manure in intermittently fed sequencing batch reactors. ASAE International Winter Conference, December 14-17, 1993, Chicago, IL. ASAE. 22 pp. McCarty, P., L. (1964). Anaerobic treatment fundamentals. Public Works Journal (Sept., Oct., Dec. Issues). McCarty, P. L. (1992). One-hundred years of anaerobic treatment. Anaerobic Digestion 1991. Elsevier, Amsterdam. 3-22. Mills, J. P. (1977). A comparison of an anaerobic digester and an aeration system treating piggery waste from the same source. Food, Fertilizer, and Agricultural Residues - Cornell Agricultural Waste Management Conference Proceedings, 1977. Ann Arbor Science. 415-422. Morrison, W. D., Braithwaite, L. A., DeBoer, S. & Smith, J. H. (1991). Air quality effects on pig health and performance and an overview of problem control methods studied at the University of Guelph. In Making Swine Buildings a Better Place to Work, ed. R. N. Goodwin, Nat. Pork Prod. Council and Nat. Pork Board, Des Moines, IA, 214-219. Ng, W. J. & Chin, K. K. (1988). Treatment of piggery wastewater by expanded-bed anaerobic filters. Biol. Wastes, 26, 215-228. Ngoddy, P. O., Harper, J. P. & Gerrish, J. B. (1974). Enhanced treatment of livestock wastewater. 1. Solid-liquid separation - estimation of vibratory screen performance on swine wastewater. J. Ag. Eng. Res., 19, 3113-3126. Nordstedt, R. A. & Thomas, M. V. (1985). Wood block media for anaerobic fixed bed reactor. Trans. ASAE, 28(6), 1990-1996. Oleszkiewicz, J. A. (1983). A comparison of treatments of low concentration of piggery waters. Ag. Wastes, 8(4), 215-231. Owen, W. F., Stuckey, D. C., Healy, J. B., Jr., Young, L. Y. & McCarty, P. L. (1979). Bioassay for Monitoring Biochemical Methane Potential and Anaerobic Toxicity. Wat Res., 13, 485-492. Owens, J. M. (1988). The use of granular sludge from upflow anaerobic sludge blanket digesters for treating flushed, screened swine wastewaters at ambient and mesophilic temperatures, North Carolina State. Raleigh, 149 p. Owens, J. M. & Chynoweth, D. P. (1993). Biochemical methane potential of MSW components. Wat. Sci. Tech., 27, 1-14. Petersen, G. (1982). Continuous mesophilic anaerobic digestion of swine manure. The Technological Institute of Denmark. #4. Powers, W.J., Wilkie, A.C., Van Horn, H.H. and Nordstedt, R.A. (1997). Effects of hydraulic retention time on performance and effluent odor of conventional and fixed-film anaerobic digesters fed dairy manure wastewaters. Trans. ASAE, 40:1449-1455. Rivard, C. J., Nieves, R. A., Nagle, N. J. & Himmel, M. E. (1994). Evaluation of discrete cellulase enzyme activities from anaerobic digestor sludge fed a municipal solid waste feedstock. Appl. Biochem. Biotechnol., 45/46, 453-460. Ross, C. C. & Drake, T. J. (1996). Handbook of biogas utilization. Vol 2. US DOE, SERBEP, TVA, Muscle Shoals, AL. Safley, L. M., Casada, M. E., Woodbury, J. W. & Roos, K. F. (1992). Global Methane Emissions From Livestock and Poultry Manure. EPA/400/1-91/048. USEPA, Washington D.C. 35

Safley, L. M., Jr. & Westerman, P. W. (1989). Anaerobic lagoon biogas recovery systems. Biol. Wastes, 27, 43-62. Safley, L. M., Jr. & Westerman, P. W. (1988). Biogas production from anaerobic lagoons. Biol. Wastes, 23, 181-193. Safley, L. M., Jr. & Westerman, P. W. (1990). Psychrophilic anaerobic digestion of animal manure: proposed design methodology. Biol. Wastes, 34, 133-148. Sievers, D. M. & Brune, D. E. (1978). Carbon/nitrogen ratio and anaerobic digestion of swine waste. Trans. ASAE, 21(3), 537-541, 549. Smith, P. H., Bordeaux, F. M., Goto, M., Shiralipour, A., Wilkie, A., Andrews, J. F., Ide, S. & Barnett, M. W. (1988). Biological production of methane from biomass. In Methane From Biomass: A Systems Approach, eds. W. H. Smith & J. R. Frank, Elsevier Applied Science Publishers, London, p.291-334. Smith, W. H. & Frank, J. R., eds. (1988). Methane from Biomass: A Systems Approach. Elsevier Applied Science Publishers, London. Smith, W.H., Wilkie, A.C. and Smith, P.H. (1992). Methane from biomass and waste - a program review. TIDE (TERI Information Digest on Energy), 2, 1-20. Sorlini, C., Ranalli, G. & Merlo, S. (1990). Microbiological aspects of anaerobic digestion of swine slurry in upflow fixed-bed digesters with different packing materials. Biol. Wastes, 31, 231-239. Speece, R. E. (1996). Anaerobic biotechnology for industrial wastewaters. Archae Press, Nashville, TN, 392p. Speece, R. F. (1997). Nutrient Requirements. In Anaerobic Digestion of Biomass, eds. D. P. Chynoweth & R. Isaacson, Elsevier Science. London, 109-128. Srivastava, V. (1987). Net energy. In Anaerobic Digestion of Biomass, eds. D. P. Chynoweth & R. Isaacson, Elsevier Science. London, 219-230. Stanogias, G., Tjandraatmadja, M. & Pearce, G. R. (1985). Effects of source and level of fibre in pig diets on methane production from pig faeces. Ag. Wastes, 12, 37-54. Stevens, M. A. & Schulte, D. D. (1979). Low temperature anaerobic digestion of swine manure. Journal of Envir. Eng. Division (ASCE), 105(Feb.), 33-42. Sutton, A. L., Nye, J. C., Patterson, J. A., Kelly, D. T. & Furumoto-Elkin, E. J. (1989). Effects of Avilamycin in swine and poultry wastes on methane production in anaerobic digesters. Biol. Wastes, 30, 35-45. Sweeten, J. M., Fulhage, C. & Humenik, F. J. (1981). Methane gas from swine manure. North Carolina Agricultural Extension Service. PHI-76. Switzenbaum, M. S. (1991). Anaerobic treatment technology for municipal and industrial wastes. Wat. Sci. Tech., 24, 281. Tafdrup, S. (1995). Viable energy production and waste recycling from anaerobic digestion of manure and other biomass materials. Biomass and Bioenergy, 9(1-5), 303-314. Taiwan. (1993). Engineering Design and Construction Manual for Swine Wastewater Treatment Facilities. Taiwan Livestock Research Institute. Tseng, S. (1992). An improved anaerobic digester for treatment of hog wastewater and sludge. International Symposium on Anaerobic Digestion of Solid Waste, April 14-17, 1992, Venice, Italy. 409-412. US, Bureau of Census. (1997). World POPClock, International Programs Center. http://www/census/gov/cgi-bin/ipc/popclockw. USDA. (1992). Agricultural Waste Management Field Handbook. USDA Soil Convservation Service. USDA. (1998). Agricultural Uses of Municipal, Animal, and Industrial Byproducts. Conservation Research Report No. 44. USDA-ARC. Hatfield, J. R., Brumm, M. C., and Melvin, S. W., Ch. 4. pp. 78-90. Swine Manure Management USEPA. (1993). Options for Reducing Methane Internationally, Vol. II: International Opportunities for Reducing Methane Emissions, Report to Congress. United States Environmental Protection Agency. EPA 430-R-93 006 B. van Velsen, A. F. M. (1977). Anaerobic digestion of piggery waste. 1. The influence of detention time and manure concentration. Neth. J. Agric. Sci., 25, 151-169. van Velsen, A. F. M., Lettinga, G. & den Ottelander, D. (1979). Anaerobic digestion of piggery waste. 3. Influence of temperature. Neth. J. Agric. Sci., 27, 255-267. Varel, V. H. & Hashimoto, A. G. (1981). Effect of dietary monensin or chlortetracycline on methane production from cattle waste. Appl. Environ. Microbiol., 41, 29-34. 36

Varel, V. H. & Hashimoto, A. G. (1982). Methane production from fermentor cultures acclimated to waste from cattle fed monensin, lasalocid, salinomycin, and avoparcin. Appl. Environ. Microbiol., 44, 14151420. Weitman, D. (1995). Federal water quality policy and animal agriculture. Seventh Intern. Symp. on Ag. and Food Processing Wastes (ISAFPW95), Chicago, IL. ASAE. 1-5. Wilkie, A.C. (1998). Anaerobic digestion of livestock wastes: A sustainable approach to odor abatement. In: "Proceedings of the North Carolina 1998 Pork Conference and Beef Symposium", pp.5-16. North Carolina Pork Council, Raleigh, North Carolina. Wilkie, A.C. and Smith, P.H. (1989). Methanogenesis in artificially created extreme environments. In: FEMS Symposium No. 49, "Microbiology of Extreme Environments and Its Potential for Biotechnology", M.S. da Costa, J.C. Duarte and R.A.D. Williams (eds.), Elsevier Science Publishers, London. pp.237-252. Wilkie, A., Goto, M., Bordeaux, F. M. & Smith, P. H. (1986). Enhancement of anaerobic methanogenesis from Napiergrass by addition of micronutrients. Biomass, 11, 135-146. Wilkie, A. & Colleran, E. (1989). The development of the anaerobic fixed-bed reactor and its application to the treatment of agricultural and industrial wastes. In International Biosystems III, ed. D. L. Wise, CRC Press, Boca Raton, FL. pp.183-226. Wilkie, A. & Colleran, E. (1986). Pilot-scale digestion of pig slurry supernatant using an upflow anaerobic filter. Envir. Tech. Lett., 7, 65-76. Wilkie, A. & Colleran, E. (1984). Start-up of anaerobic filters containing different support materials using pig slurry supernatant. Biotechnol. Lett., 6(11), 735-740. Wilkie, A. C., Riedesel, K. J. and Cubinski, K. R. (1995a). Anaerobic digestion for odor control. In: Nuisance Concerns in Animal Manure Management: Odors and Flies. H.H. Van Horn (ed.), Florida Cooperative Extension, University of Florida, Gainesville, FL, pp. 56-62. Wilkie, A. C., Van Horn, H. H., Powers, W. J., and Riedesel, K. J. (1995b). Anaerobic treatment technology - an integrated approach to controlling manure odors. In: International Livestock Odor Conference Proceedings '95, New Knowledge in Livestock Odor. Iowa State University College of Agriculture, Ames, IA, pp. 223-227. Winter, J. & Wildenauer, F. X. (1984). Comparison of volatile acid turnover during improved digestion of sewage sludge, cattle manure, and piggery waste. Third European Congress on Biotechnology, Munchen, Germany. Weinheim. 3, 81-87. Wong, M. H. (1990). Anaerobic digestion of pig manure mixed with sewage sludge. Biol. Wastes, 31, 223230. Wong, M. H. & Cheung, Y. H. (1989). Anaerobic digestion of pig manure with different agro-industrial wastes. Biol. Wastes, 28, 143-155. WPCF. (1987). Anaerobic Sludge Digestion. Manual Practice No. 19., sec. ed. Water Pollution Control Federation, Alexandria, VA. Wujick, W. J. & Jewell, W. J. (1980). Dry anaerobic fermentation. Biotech. Bioeng. Symp., 10, 43-65. Yang, P. Y. & Chou, C. Y. (1985). Horizontal-baffled anaerobic reactor for treating diluted swine wastewater. Ag. Wastes, 14, 221-239. Yang, P. Y. & Nagano, S. Y. (1985). A red mud plastic swine manure anaerobic reactor with sludge recycle in Hawaii. Trans. ASAE, 28(4), 1284-1288. Yang, P. Y. & Chen, H. (1994). A land-limited and energy and energy-saving treatment system for dilute swine wastewater. Biores. Tech., 49, 129-137. Yang, P. Y. & Kuroshima, M. (1995). A simple design and operation for the anaerobic treatment of highly concentrated swine waste in the tropics. Wat. Sci. Technol., 32(12), 91-97. Yang, P. Y. & Moengangongo, T. H. (1987). Operational stability of a horizontally baffled anaerobic reactor for diluted swine wastewater in the tropics. Trans. ASAE, 30(4), 1105-1110. Zeeman, G., Sutter, K., Koster, V. T., Koster, M. & Wellinger, A. (1988). Psychrophilic digestion of dairy cattle and pig manure: start-up procedures of batch, fed-batch and CSTR-type digesters. Biol. Wastes, 26, 15-31. Zhang, R. & Felmann, D. J. (1997). Animal manure management - agriculture scoping study. The EPRI Agricultural Technology Alliance - Electric Power Research Institute. C109139. Zhang, R. H., North, J. R. & D.L., D. (1990). Operation of a field scale anaerobic digester on a swine farm. Appl. Eng. Ag., 6(6), 771-776. 37

Zhang, R. H., Yin, Y., Sung, S. & Dague, R. R. (1997). Anaerobic treatment of swine waste by the anaerobic sequencing batch reactor. Trans. ASAE, 40(3), 761-767. Zinder, S. H. (1993). Physiological ecology of methanogens. In Methanogenesis: Ecology, Physiology, Biochemistry, and Genetics, ed. J. G. Ferry, Chapman and Hall. New York, NY, 128-206.

38

Potrebbero piacerti anche