Вы находитесь на странице: 1из 50

Pytheas Business Guides

February 2008

Treatment of Municipal Solid Waste Anaerobic Digestion Technologies

www.pytheas.net

Pytheas Business Guides

Anaerobic Digestion (AD) technologies have the potential to immensely reduce the environmental impact of waste disposal while capturing biogas energy; they also complement other organic waste diversion technologies such as composting

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

Index Glossary of Terms 1.0 2.0 2.1 2.2 3.0 3.1 4.0 4.1 4.1.1 4.1.2 4.1.3 4.2 4.2.1 4.2.2 4.2.3 5.0 5.1 5.2 5.3 5.4 6.0 6.1 6.2 6.3 7.0 7.1 7.2 8.0 8.1 8.2 8.3 8.4 8.5 8.5.1 8.5.2 8.5.3 9.0 9.1 Introduction Background Digestion Process Description State of MSW disposal in the U.S. and Europe Categories of Engineered Anaerobic Digestion Systems Material Handling Systems A typical sorting line Review of Commercial AD Technologies for MSW treatment Single-stage Wet Systems The Wabio Process The Waasa Process The BIMA Digester Single-stage Dry Systems The Dranco Process The Valorga Process The Kompogas Process Multi-stage Digesters The BTA Process The Linde-KCA Process The ArrowBio Process The Biopercolat Process Batch Digesters The Biocel System The SEBAC System The Anaerobic Phased Solids (APS) Digester Digester Performance Biogas Yield Life Cycle Analysis Environmental Impacts Municipal Solid Waste Anaerobic Digestion Emissions Anaerobic Digestion Emissions from use of Biogas VOC Emissions from Composting and Digestate Landfill Gas Emissions Solid Residues Compost Quality Heavy Metals Dioxin in Trash and Compost Pesticides and Herbicides in Compost and Composted Digestate Economics Costs Sources 4 7 9 9 11 13 15 18 18 18 20 21 21 21 22 23 25 25 27 28 30 31 31 32 33 35 35 37 39 39 40 40 41 41 41 42 43 44 44 46

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

Glossary of Terms
Aerobic Alternative daily cover Anaerobic Anaerobic digester With or in the presence of oxygen. Material other than soil used to cover the surface of active landfills at the end of each day to control diseases, fires, odors, etc. Without oxygen. A dedicated unit process for controlling the anaerobic decomposition of organic material. Typically consists of one or more enclosed, temperature controlled tanks with material handling equipment designed to prevent the introduction of oxygen from the atmosphere. Decomposition of biomass into methane by anaerobic bacteria. Also called biomethanization. A rotating drum often with a trommel screen used for size reduction and pretreatment of the organic fraction in mixed MSW for sorting. Can be aerated to encourage biological breakdown. Can be operated at retention times from several hours to several days. A landfill operated as a bioreactor using leachate recycling (or other management schemes) to increase the rate of organic decomposition and biogas production. Not to be confused with anaerobic digester. Biochemical oxygen demand is the amount of oxygen required for complete (aerobic) biological decomposition of a material. The standard laboratory method (BOD5) tests the amount of dissolved oxygen consumed in a closed aqueous system over a five-day period. It is a fairly direct but time-consuming measure of biodegradability of liquid streams. Biothermal Energy is the concept of utilizing composting material to generate heat and biogas for use as an alternative energy source. Organisms which assist in the breaking down waste material also create heat, methane, carbon dioxide and other gases. These byproducts can be used in a variety of applications. Most prominently, greenhouses can use the biothermal energy and carbon dioxide to promote the growth of plants. The simultaneous digestion of a mixture of two or more feedstocks. Codigestion is sometimes employed to balance nutrient requirements of the process, improve gas production rate, as well as to process wastes from multiple sources. A rapid conversion of chemical energy into thermal energy. The reaction is exothermic. Organic matter is oxidized with sufficient air (or oxygen) for reactions to go to completion. The carbon and hydrogen are oxidized to carbon dioxide and water, respectively. Compost here refers to stabilized and screened organic material ready for horticultural or agricultural use. If anaerobically digested material is used as compost, it must be biologically stabilized, typically through aeration and maturation. A digester configuration in which the entire digester contents are mixed to create homogeneous slurry. Either in the presence of oxygen (aerobic) or in an oxygen-depleted atmosphere (anaerobic), the process in which microbes digest biogenic carbonaceous materials and emit any number of energetic, inert gases and liquids. The use of microorganisms such as yeast, bacteria, and fungi to convert substrates such as sugar into products. In the absence of oxygen, these products can include ethanol, methane, and carbon dioxide plus some increase in cell mass. When oxygen is present, the increase in cell mass is generally much greater with water and carbon dioxide usually the primary products.

Biogasification Biomixer

Bioreactor-landfill

Biochemical oxygen demand (BOD)

Biothermal energy

Co-digestion

Combustion

Compost

Continuously stirred tank reactor Digestion

Fermentation

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

Gasification

Production of energetic gases from solid or liquid organic feedstocks usually by partial oxidation. Primary energetic gases produced are hydrogen, carbon monoxide, and methane, along with an inorganic ash residue. The material left over after separation of recyclables and putrescible material from the mixed waste stream. Composed mostly of inorganic material, grey waste usually contains a significant amount of organic material. Depending on its composition, grey waste and can be treated biologically or burned prior to final disposal. The average time a volume element of fluid resides in a reactor. It is computed from liquid-filled volume of an anaerobic digester divided by the volumetric flow rate of liquid medium. Increasing the HRT allows more contact time between substrate and bacteria but requires slower feeding and/or larger reactor volume. Gasification using hydrogen gas to react with the carbon in organic materials to produce a methane rich gas effluent, and provide heat for the process. Any pyrolytic products present are usually converted into methane. Steam pyrolysis is often used as a precursor process that can enhance the hydrogen reaction kinetics, despite the presence of water in the feed. Since oxygen is not intentionally introduced, carbon oxides are reduced and methane increased as the hydrogen pressure is increased. Toxic hydrocarbons, like furans and dioxins, are chemically reduced by hydrogasification to less hazardous chemical compounds. A chemical or biological process in which water is added to other molecules (the conditions are wide ranging and many molecules can be hydrolyzed). Hydrolysis is a pre-treatment or preliminary step in fermentation processes that ultimately yield biogas or ethanol. For cellulose and hemicellulose, a variety of hydrolysis methods can be used to break down the long chain polymer into simple glucose molecules, Efficiencies of hydrolysis vary among methods and feedstocks. A waste processing system that combines a sorting facility for materials recovery (the mechanical portion) with biological treatment, either aerobic or anaerobic, for stabilizing the organic fraction before landfilling.

Grey waste

Hydraulic retention time (HRT)

Hydrogasification

Hydrolysis

Mechanical biological treatment

Materials recovery facility A facility where mixed MSW is sorted in order to recover material for reuse or recycling. In California, the post MRF fraction is typically landfilled. Mechanically separated OFMSW Mesophilic Municipal solid waste (MSW) Organic material separated from the mixed waste stream by mechanical means (i.e., trommel screens, shredders, magnets, density dependent mechanisms). One of two optimal temperature ranges for microorganisms involved in anaerobic digestion. Mesophilic temperature is around 35 C (95 F). MSW includes all of the solid wastes that are generated from residential sources, commercial and business establishments, institutional facilities, construction and demolition activities, municipal services, and treatment plant sites. Hazardous wastes are generally not considered MSW. Note that some regions or countries consider only residential solid waste as MSW. Material containing carbon and hydrogen. Organic material in MSW includes the biomass components of the waste stream as well as hydrocarbons usually derived from fossil sources (e.g., most plastics, polymers, the majority of waste tire components, and petroleum residues). The biogenic fraction of MSW. OFMSW can be either removed from the waste stream at the source (source-separation) or downstream by mechanical separation, picking lines a combination of the two. A digester in which materials enter at one end and push older materials toward the opposite end. Plug flow digesters do not usually have internal mixers, and the breakdown of organic matter naturally segregates itself along the length of the digester.

Organic

Organic fraction of municipal solid waste (OFMSW) Plug flow digester

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

Pre-treatment

In reference to municipal solid waste, pre-treatment can refer to any process used to treat the raw MSW stream before disposal or management. This includes separation, drying, comminuting, hydrolysis, biological treatment, heating, pyrolysis, and others. A thermochemical decomposition of organic material at elevated temperatures without the participation of oxygen. It involves the simultaneous change of chemical composition and physical phase, and is irreversible. Pyrolytic oils, fuel gas, chars and ash are produced in quantities that are highly dependent on temperature, residence time and the amount of heat applied. The average length of time solid material remains in a reactor. SRT and HRT are equal for complete mix and plug flow reactors. Some two-stage reactor concepts and UASB reactors decouple HRT from the SRT allowing the solids to have longer contact time with microbes while maintaining smaller reactor volume and higher throughput.

Pyrolysis

Solids retention time

Source-separated OFMSW Organic solid waste separated at the source (i.e., not mixed in with the other (SS-OFMSW) solid wastes). Often comes from municipal curbside recycling programs in which yard waste and sometimes kitchen scraps are collected separately from the rest of the MSW stream. Thermophilic One of two optimal temperature ranges for microorganisms (bacteria) involved in anaerobic digestion. Thermophilic optimum temperature range is between 45 and 70 C (104 and 160 F). The amount of solid material (or dry matter) remaining after removing moisture from a sample. Usually expressed as a percentage of the as-received or wet weight. Moisture content plus TS (both expressed as percentage of wet weight) equals 100%. A trommel is a screened cylinder used to separate materials by size - for example, separating the biodegradable fraction of mixed municipal waste or separating different sizes of crushed stone. This is a standard laboratory technique used to measure the anaerobic biodegradability and associated methane yield from a given substrate. The test is run until no further gas production is detected and can last up to 100 days. The results can be influenced by the substrate concentration and particle size, the inoculum source, the food to microorganism ratio, and the presence or build-up of inhibitory compounds among others (also known as ultimate biomethane potential, BMP, and Bo). UASB digestion is a generic, mature technology specifically designed for the treatment of high strength wastewaters, such as in dairy and candy manufacture and other industries. VOCs are organic chemicals that have a high vapor pressure at ordinary, roomtemperature conditions. Their high vapor pressure results from a low boiling point, which causes large numbers of molecules to evaporate or sublimate from the liquid or solid form of the compound and enter the surrounding air. The amount of combustible material in a sample (the remainder is ash). The value is usually reported as a percentage of the TS, but may occasionally be given as a fraction of the wet weight. VS is used as an indicator or proxy for the biodegradability of a material, though recalcitrant biomass (i.e., lignin) which is part of the VS is less digestible.

Total solids (TS)

Trommel

Ultimate methane potential

Upflow anaerobic sludge basket (UASB) Volatile organic compounds (VOCs)

Volatile solids (VS)

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

1.0 Introduction Biodegradation of organic material occurs in nature through the action of both aerobic and anaerobic microorganisms. In aerobic systems, partial oxidation of the ingested organic material is the result yielding carbon dioxide and water and undigested residue. Anaerobic bacteria will also degrade organic matter, including some abiogenic forms, in the absence of oxygen with ultimate products being nonreactive residues, carbon dioxide, and methane. These bacteria naturally occur in the environment in anaerobic niches such as marshes, sediments, wetlands, and the digestive tracts of ruminants and certain species of insects. Facultative organisms are capable of persisting in either environment. Anaerobic digestion (AD) is a bacterial fermentation process that operates without free oxygen and results in a biogas containing mostly methane and carbon dioxide. It occurs naturally in anaerobic niches such as marshes, sediments, wetlands, and the digestive tracts of ruminants and certain species of insects. AD is also the principal decomposition process occurring in landfills. AD systems are employed in many wastewater treatment facilities for sludge degradation and stabilization, and are used in engineered anaerobic digesters to treat high-strength industrial and food processing wastewaters prior to discharge. There are also many instances of AD applied at animal feeding operations and dairies to mitigate some of the impacts of manure and for energy production. AD of municipal solid waste (MSW) is used in different regions worldwide to: Reduce the amount of material being landfilled; Stabilize organic material before disposal in order to reduce future environmental impacts from air and water emissions; and Recover energy. Over the past 30 years, AD of MSW technology has advanced in Europe because of waste management policies enacted to reduce the long-term health and environmental impacts of landfill disposal. This has led to relatively high landfill tipping fees, which, in combination with generous prices paid for renewable energy, has created an active commercial market for AD and other MSW treatment technologies in Europe. Installed AD capacity for the treatment of MSW in Europe is estimated to be more than 6 million tons per year. In some parts of Europe, source separation of the organic fraction of municipal solid waste (OFMSW) is common and even mandatory, which contributes to the growth of biological treatment industries. Regions outside of Europe are also enacting more stringent waste disposal regulations, leading to the development of new AD and other MSW conversion plants.
Investors and city planners will be more likely to adopt AD of MSW if additional revenues are provided initially; these revenues can come from supports for the energy produced, (i.e., tax credits and guaranteed markets), increased tipping fees and, potentially, green or carbon credits

Generally in the U.S. landfills continue to be the lowest-cost option for managing MSW, since unlike Europe and Japan, space for new landfills is not as scarce, waste management policies are less rigorous, and full life-cycle costs and impacts are not accounted for. Furthermore, the energy market and regulatory mechanisms for licensing MSW AD and other conversion facilities in most of the U.S. have not been developed to easily accommodate commercial systems. Composting of the OFMSW has increased significantly over the past 20 years, particularly for source-separated wastes, but by far the majority of the yard and food waste generated still goes to landfills. AD facilities are capable of producing energy and reducing the biodegradable content of the organic waste prior to composting, which reduces emissions of pollutants and greenhouse gases. However, in many parts of the world these environmental and public

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

health benefits have not been adequately internalized economically, especially considering the lack of familiarity with the technology. Investors and city planners will be more likely to adopt AD of MSW if additional revenues are provided initially. These revenues can come from supports for the energy produced (i.e. tax credits and guaranteed markets), increased tipping fees and, potentially, green or carbon credits. Many European countries have passed laws mandating that utility More states have to companies purchase green energy, whereas in the U.S. few of the adopt laws (that are farms or wastewater treatment facilities that produce excess enforced), electricity from biogas have secured contracts with the utilities. mandating, that Additionally, while European Union directives have called for utility companies mandatory pre-treatment and decreased disposal of biodegradable purchase green material in landfills, no equivalent regulations exist in U.S. federal or energy; requiring state codes. However, waste diversion requirements or targets exist also the pretreatment and in many states in the U.S., and reducing OFMSW disposal has been decreased disposal a focus of waste managers and municipalities attempting to achieve of biodegradable the targets. Nonetheless, interest in AD of MSW is growing, and material in landfills several U.S. jurisdictions are investigating landfill alternatives that include AD. The technologies have been used successfully for over twenty five years in Europe where the industry continues to expand. Facilities were also built recently in Canada, Japan, Australia and several other countries. The European market has shown a large preference for single-stage over two-stage digesters and a slight preference for dry digestion systems over wet systems. However, the choice of AD technology depends on the composition of the waste stream, co-product markets, and other site-specific requirements. The design of any new digester facility should be based on a thorough feasibility study, and special attention should be paid to all aspects of the treatment process, including waste collection and transportation, pre-treatment processing (i.e. pulping, grinding, and sieving), material handling, post-treatment processing (i.e. aeration and wastewater treatment), public education, and strategic location of the system to be sited. Landfill bioreactors may merit further consideration in their own right, but special attention should be paid to their performance and air/water emissions. In addition to electricity, other value-added product streams from AD systems could provide revenue to help improve the economic viability of organic waste treatment technologies. For example, technologies for upgrading biogas to natural-gas quality biomethane are available, as are technologies that utilize lignocellulosic materials which include residues from digesters. However, regulatory and definitional barriers need to be minimized in order to fully capitalize on these technologies and product streams. The public desire for change in waste management practices will lead to a reduction in landfill availability. AD and other conversion technologies have the potential to minimize the environmental impact of waste disposal by reducing the amount of biodegradable materials in landfills. Public policies that encourage organic solid waste disposal reduction will help to facilitate the adoption of such technologies. In addition, as the technologies advance, their installation costs should decrease. Equally important, AD technology developers need to work closely with waste collection and management companies in order to develop and implement appropriate digester system designs and material handling strategies and achieve successful enterprises.

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

2.0 Background AD is a biological process typically employed in many wastewater treatment facilities for sludge degradation and stabilization, and it is the principal biological process occurring in landfills. Internationally, AD has been used for decades, primarily in rural areas, for the production of biogas for use as a cooking and lighting fuel. Many household-scale digesters are employed in rural China and India for waste treatment and gas production. Over the past twenty plus years, Europe has developed large-scale centralized systems for municipal solid waste treatment with electricity generation as a co-product. Other industrialized countries have followed the European model. Biodegradation of organic material occurs in nature principally through the action of aerobic microorganisms. Ultimately, complete oxidation of the carbonaceous organic materials results in the production of carbon dioxide (CO2) and water (H2O). Anaerobic microorganisms degrade the organic matter in the absence of oxygen with ultimate products being CO2 and methane (CH4), although lignin and lignin-encased biomass degrade very slowly. Anaerobic microorganisms occur naturally in low-oxygen niches such as marshes, sediments, wetlands, and in the digestive tract of ruminant animals and certain species of insects. 2.1 Digestion Process Description The anaerobic digestion of organic material is accomplished by a consortium of microorganisms working synergistically. Digestion occurs in a four-step process: Hydrolysis, Acidogenesis, Acetogenesis, and Methanogenesis (Figure 1): 1. Step 1. Large protein macromolecules, fats and carbohydrate polymers (such as cellulose and starch) are broken down through hydrolysis to amino acids, long-chain fatty acids, and sugars. Step 2. These products are then fermented during acidogenesis to form three, four, and five-carbon volatile fatty acids, such as lactic, butyric, propionic, and valeric acid. Step 3. In acetogenesis, bacteria consume these fermentation products and generate acetic acid, carbon dioxide, and hydrogen. Step 4. Finally, methanogenic organisms Figure 1. Anaerobic Digestion Four-step Process consume the acetate, Source: Soil Water & Life Solutions hydrogen, and some of the carbon dioxide to produce methane. Three biochemical pathways are used by methanogens to produce methane gas. The pathways along with the stoichiometries of the overall chemical reactions are:

2.

3.

4.

Copyright 2008 Pytheas Limited

21 February 2008

Pytheas Business Guides

a. b. c.

Acetotrophic methanogenesis: 4 CH3COOH 4 CO2 + 4 CH4 Hydrogenotrophic methanogenesis: CO2 + 4 H2 CH4 + 2 H2O Methylotrophic methanogenesis: 4 CH3OH + 6 H2 3 CH4 + 2 H2O

Methanol is shown as the substrate for the methylotrophic pathway, although other methylated substrates can be converted. Sugars and sugar-containing polymers such as starch and cellulose yield one mole of acetate per mole of sugar degraded. Since acetotrophic methanogenesis is the primary pathway used, theoretical yield calculations are often made using this pathway alone. From the stoichiometry above, it can be seen that the biogas produced would theoretically contain 50% methane and 50% carbon dioxide. However, acetogenesis typically produces some hydrogen, and for every four moles of hydrogen consumed by hydrogenotrophic methanogens a mole of carbon dioxide is converted to methane. Substrates other than sugar, such as fats and proteins, can yield larger amounts of hydrogen leading to higher typical methane content for these substrates. Furthermore, hydrogen and acetate can be biochemical substrates for a number of other products as well. Therefore, the overall biogas yield and methane content will vary for different substrates, biological consortia and digester conditions. Typically, the methane content of biogas ranges from 40-70% (by volume).
AD digestion requires attention to the nutritional needs of the bacteria degrading the waste substrates; the most important nutrients for bacteria are carbon and nitrogen, but these two elements must be provided in the proper ratio

Anaerobic conditions are required for healthy methanogenesis to occur. This means that the reactors used must be well sealed which allows the biogas to be collected for energy conversion and eliminates methane emissions during the anaerobic digestion process. In addition to methane and carbon dioxide, semi-harmful contaminants such as hydrogen sulfide and ammonia are produced, albeit in much smaller amounts (<1% by volume). The production of these trace gases in the biogas depends on the sulfur and nitrogen contents of the feedstock. However, these elements are also nutrients required by the bacteria, so they cannot be eliminated completely. In fact, anaerobic digestion requires attention to the nutritional needs of the bacteria degrading the waste substrates. The most important nutrients for bacteria are carbon and nitrogen, but these two elements must be provided in the proper ratio. Otherwise, ammonia can build up to levels that can inhibit the microorganisms. The appropriate carbon/nitrogen (C/N) ratio depends on the digestibility of the carbon and nitrogen sources; therefore, the appropriate C/N ratio for organic MSW may be different from that for other feedstock such as manure or wastewater sludge. In general, the optimal conditions for anaerobic digestion of organic matter are near-neutral pH, constant temperature (thermophilic or mesophilic), and a relatively consistent feeding rate. Imbalances among the different microorganisms can develop if conditions are not maintained near optimum. The most common result of imbalance is the buildup of organic acids which suppresses the methanogenic organisms adding to even more buildup of acidity. Acid buildup is usually controlled naturally by inherent chemical buffers and by the methanogens themselves as they consume acids to produce methane. These natural controls can break down if too much feed is added and organic acids are produced faster than they are consumed, if inhibitory compounds accumulate, or if the feed stream lacks natural pH buffers such as carbonate and ammonium. Solid concentrations higher than about 40% TS can also result in process inhibition, likely due to the reduced contact area available to the AD microorganisms. The TS content of OFMSW typically ranges from 30-60%, thus some water may need to be added. Process water can be used, but this may also result in the buildup of inhibitory compounds. Thus, low-solids digesters require the addition of fresh water. Higher temperatures result in faster reaction

Copyright 2008 Pytheas Limited

21 February 2008

10

Pytheas Business Guides

kinetics which, in practice, translates to smaller reactors needed to process a given waste stream. However, the micro-organisms themselves are adapted to relatively narrow temperature ranges. Mesophilic and thermophilic microbes are adapted to roughly 30-40 C (86-104 F) and 50-60 C (122-140 F) respectively. 2.2 State of Municipal Solid Waste disposal in the U.S. and Europe In 2007, the U.S. generated about 255 million tons of MSW; about, 63 million tons (24.7%) were recycled, 22 million tons composted (8.5%) and 32 million tons combusted with energy recovery (12.5%). The remaining about 138 million tons (54.3%) were discarded in landfills. Handling and treatment of OFMSW is more difficult than treating wastewater or manure. As such, the AD of OFMSW requires a larger amount of investment and technological experience. Furthermore, capital and operating costs are higher for AD than for composting or landfilling. The low tipping fees charged by landfills in the U.S. and relatively low energy prices compared to those in Europe make it difficult for AD and other conversion technologies to be costcompetitive. However, life-cycle analyses (LCA) have shown that AD of MSW reduces the environmental impact and is more cost-effective (in Europe) on a whole-system basis than landfilling or composting over the life of the project. The European Union (EU) passed a regulation in 2002 to standardize reporting and timing of data collection for waste disposal, and the first year of data was included in the most recent Eurostat Yearbook (2006-2007); for comparison with past years, only average per capita statistics were reported. As of 2004, Europeans disposed of an average of 0.6 MT (0.7 tons) of MSW; however, unlike U.S. statistics on MSW production, this waste did not include construction and demolition debris which makes up 30 % of the reported MSW in the U.S. Nonetheless, the Western European average per capita disposal was almost half of the U.S. average.

Figure 2. U.S. Total MSW generation


Source: United States Environmental Protection Agency

Figure 3. U.S. Per Capita MSW generation


Source: United States Environmental Protection Agency

In Europe, the per capita MSW production increased over the past ten years, but landfill disposal declined slightly. Per capita combustion with energy recovery has remained relatively constant while composting, recycling, and other treatments almost doubled since 1995. OFMSW typically comprises 50-60% by weight of the solid waste stream collected by municipalities in Europe that do not practice source separation. In 2004, this totaled some 200 million MT (220 million tons). In 1999, the EU adopted the Landfill Directive (Council Directive 99/31/EC) which became enforceable in 2001. It required the biodegradable portion of MSW to be reduced by 25% of that disposed in 1995 within five years, 50% within eight years, and 65% within 15 years. Furthermore, Article 6(a) required that all waste that gets landfilled must be treated, with the exception of inert materials for which treatment is not technically feasible. Each country in the EU is held to this standard as a minimum requirement, but in practice Germany, Austria, Denmark, Luxembourg, the Netherlands, and Belgium had already imposed such restrictions and now have even stricter requirements while France, Italy, Sweden, England and Finland
Copyright 2008 Pytheas Limited 21 February 2008

11

Pytheas Business Guides

converted their facilities subsequent to adopting the law. Greece, Ireland, Portugal and Spain were in the process of converting their facilities as of 2004. As a consequence, installed AD capacity in Europe has increased sharply and now stands at more than 4 million tons annual capacity (Figures 4 and 5). Most notably, Spain recently installed several large-scale AD facilities and now processes over 1 million tons OFMSW per year which accounts for over 50% of the organic waste produced there. In Germany the Recycling and Waste Law, the Directive on Residential Waste Disposal, the Federal Ordinance on Handling of Biowaste, and the Ordinance for Environmentally Sound Landfilling set stringent limits for the composition of treated MSW prior to disposal in landfills. For example, in Germany the upper limits on total organic carbon and energy content of material going to landfill were set at 18% and 6,000 kJ/kg (2,580 BTU/lb). Furthermore, energy prices in Europe are generally higher than in the U.S. and many European countries provide financial incentives to renewable energy producers. For example, in Germany the Renewable Energy Act guaranteed renewable electricity producers a high percentage of the retail electricity price (75-90%) with biomass earning 0.15-0.25 $/kWh (converted from Euros to PPP-adjusted U.S. dollars). Tariffs with prescribed annual reductions were guaranteed for up to 20 years. The regulation also required utilities to connect renewable producers to the grid. In addition many AD of MSW facilities in the EU also sell green certificates and carbon credits. Direct subsidies and soft loans are also used to support new renewable energy producers.

Copyright 2008 Pytheas Limited

21 February 2008

12

Pytheas Business Guides

3.0 Categories of Engineered Anaerobic Digestion Systems The most common MSW Anaerobic Digestion technologies are: One-stage Continuous Systems Low-solids or Wet High-solids or Dry Two-stage Continuous Systems Dry-Wet Wet-Wet Batch Systems One Stage Two Stage Single-stage digesters are simple to design, build, and operate and are generally less expensive. The organic loading rate (OLR) of single-stage digesters is limited by the ability of methanogenic organisms to tolerate the sudden decline in pH that results from rapid acid production during hydrolysis. Two-stage digesters separate the initial hydrolysis and acidproducing fermentation from methanogenesis, which allows for higher loading rates but requires additional reactors and handling systems. In Europe, about 90% of the installed AD capacity is from single-stage systems and about 10% is from two-stage systems (Figure 4).

Figure 4. Growth of MSW AD technology by number of stages in Europe


Source: Soil Water & Life Solutions

Another important design parameter is the total solids (TS) concentration in the reactor, expressed as a fraction of the wet mass of the prepared feedstock. The remainder of the wet mass is water by definition. The classification scheme for solids content is usually described as being either high-solids or low-solids. High-solids systems are also called dry systems and low-solids systems may be referred to as wet systems. A prepared feedstock stream with less than 15% TS is considered wet and feedstock with TS greater than 15-20% are considered dry (although there is no established standard for the cutoff point). Feedstock is typically diluted with process water to achieve the desirable solids content during the preparation stages. Before AD became an accepted technology for treating MSW, single-stage wet digesters were used for treating agricultural and municipal wastewater. However, MSW slurry behaves differently than wastewater sludge. Because of the heterogeneous nature of MSW, the slurry tends to separate and form a scum layer which prevents the bacteria from degrading these organics. The scum layer tends to evade the pump outlets and can clog pumps and pipes when it is removed from the reactors. To prevent this, pretreatment to remove inert solids and homogenize the waste is required. Solids can also short circuit to the effluent pipe before they have broken down completely, therefore design modifications were made to allow longer contact time between bacteria and dense, recalcitrant material. Furthermore, MSW tends to contain a higher percentage of toxic and inhibitory compounds than wastewater. In diluted slurry, these compounds diffuse quickly and evenly throughout the reactor. In high enough

Copyright 2008 Pytheas Limited

21 February 2008

13

Pytheas Business Guides

concentrations, this can shock the microorganisms, whereas in a dry system the lower diffusion rate protects the microbes. Because of these constraints, dry systems have become prevalent in Europe (Figure 5), making up 60% of the singlestage digester capacity installed to date. Dry digesters treat waste streams with 20-40% total solids without adding dilution water. However, these systems may retain some process water or add some water either as liquid or in the form of steam used to heat the incoming feedstock. Furthermore, as organic matter breaks down, the internal MC of the digester will increase (from 64-72%). Nonetheless, heavy duty pumps, Figure 5. Growth of MSW anaerobic digester technology by solids content (<5% conveyors, and augers are TS = Wet, >20% TS = Dry) in Europe Source: Soil Water & Life Solutions required for handling the waste, which adds to the systems capital costs. Some of this additional cost is offset by the reduction in pretreatment equipment required. Most dry digesters operate as plug flow digesters, but due to the viscosity of the feed, the incoming waste does not mix with the contents of the digester. This prevents inoculation of the incoming waste which can lead to local overloading. Therefore, most of the digester designs include an inoculation loop in which the incoming OFMSW is mixed with some of the exiting digestate paste prior to loading. Multi-stage systems are designed to take advantage of the fact that different portions of the overall biochemical process have different optimal conditions. By optimizing each stage separately, the overall rate can be increased. Typically, two-stage processes attempt to optimize the hydrolysis and fermentative acidification reactions in the first stage where the rate is limited by hydrolysis of complex carbohydrates. The second stage is optimized for methanogenesis where the rate in this stage is limited by microbial growth kinetics. Since methanogenic archaea prefer pH in the range of 78.5 while acidogenic bacteria prefer lower pH, the organic acids are diluted into the second stage at a controlled rate. Often a closed recirculation loop is provided to allow greater contact time for the unhydrolyzed organic matter. Some multi-stage systems apply a microaerophilic process in an attempt to increase the oxidation of lignin and make more cellulose available for hydrolysis. Although adding oxygen to an anaerobic environment seems counterintuitive, sludge granules can shield the obligate anaerobes from oxygen poisoning and the practice has been shown to increase biogas yield in some situations. In two-stage systems, because methanogens are more sensitive to oxygen exposure than fermentative bacteria, the air may preferentially inhibit methanogens, which could help maintain a low pH in the hydrolysis stage. However, if the oxygen is not completely consumed and the biogas contains a mixture of oxygen and hydrogen and/or methane, hazardous conditions could be created. Process flexibility is one of the advantages of multi-stage systems. However, this flexibility also increases cost and complexity by requiring additional reactors, material handling and process control systems. On the opposite end of the spectrum, batch or sequential batch systems aim to reduce complexity and material handling requirements. As opposed to continuous wet and dry systems, the feedstock does not need to be carefully metered into a batch reactor, thereby eliminating the need for complex material handling equipment. The

Copyright 2008 Pytheas Limited

21 February 2008

14

Pytheas Business Guides

primary disadvantage of batch digesters is uneven gas production and lack of stability in the microbial population. To surmount these issues, batch systems can also be combined with multi-stage configurations. 3.1 Material Handling Systems A typical sorting line

Figure 6. A Municipal Waste Treatment Plant Diagram Material Handling Systems


Source: Ecoenergy Gesellschaft fur Energie- und Umwelttechnik mbh

Extensive pre- and post-digestion processing units, regardless of the waste source or digester type are used. Pre-sorting is necessary to prevent clogging of the pumps and to reduce the amount of reactor volume occupied by inert material. Even source-separated waste inevitably contains metal and plastic contaminants and must be pre-sorted. A typical sorting line includes the following components (also seen in Figure 6 above and Table 1 below): Receiving Can include some visual (manual or robotic) sorting and removal of bulky or potentially harmful items; Provides a buffer for inflow rate fluctuations. Particle size reduction Can be mechanical and/or biological; Relies on the relative ease of reducing the particle size of the organic fraction. Separation Can be based on magnetism, density, and size. The receiving area allows for unloading of raw MSW and isolation of MSW from different sources. Some receiving areas use robotics to minimize human contact with the waste. Others

Copyright 2008 Pytheas Limited

21 February 2008

15

Pytheas Business Guides

incorporate a sorting line for workers to manually remove the most obvious inorganic materials. Once the MSW has been loaded into the mechanical separation system, human contact is minimal as biological and mechanical processes prepare the MSW for density and/or size separation. Density separation requires wetting the MSW; therefore it is more commonly applied when using low-solids digesters. Organic material breaks into smaller particles more easily than inorganic material, therefore a mechanical macerator or agitator is often employed prior to screening. In addition, some aerobic treatment can help break down the organic matter. This may also be accompanied by a loss of digestible organic matter; therefore short retention times are used. Between several hours and one or two days is typical for Figure 7. By-products and extracted inert material of an MSW Treatment Plant Source: Ecoenergy Gesellschaft fur Energie- und Umwelttechnik mbh rotating drums, or biomixers, which combine agitation with aerobic treatment. Organic materials have high biogas and methane yields even when the MSW had spent only 24 hours in the rotating drum. In a rotating drum system, a sieve may line the sides of the drum allowing undersized particles to pass to the dosing unit while expelling oversized, primarily inorganic, particles. Alternatively, the waste may pass through one or more trommel screens after the drum for sieving. Dosing units store mixed waste to even out fluctuations in the content and volume of MSW going to the digester. They can also be used for heating and inoculating the digester feed. Heat may be added as steam, which can be produced using waste heat from engine generators. Some systems have a separate feed mixer which combines the sorted MSW with digester paste in order to inoculate the new feed and bring it to the appropriate MC. In Bassano, Italy a Valorga digester accepts source-separated waste and grey waste. As can be seen from the diagram below (Figure 8), even source-separated waste passes through a primary sieve and a magnetic metals removal unit.

Figure 8. MSW Treatment Plant at Bassano, Italy Pre-processing Diagram


Source: Bolzonella, D., P. Pavan, S. Mace, and F. Cecchi. Dry Anaerobic digestion of differently sorted organic municipal waste: A full scale experience. Water Science & Technology. 2006. 53(8): p. 23-32 Modified by SWLS

The grey waste which is the inorganic fraction of the source-separated waste consists primarily of inorganic materials. (In fact, organics make up only 10-16% of this material, and paper makes up an additional 34-50%).

Copyright 2008 Pytheas Limited

21 February 2008

16

Pytheas Business Guides

The grey waste passes through an additional drum screen and densimetric separator which suspends the waste in water, removing the floating layer as well as the heavy particles that sink to the bottom. The Treviso wastewater treatment facility found its anaerobic digesters to be too large for processing waste activated sludge (WAS) only, so they built a separation unit to remove the organic fraction of MSW for co-digestion with the sludge. As can be seen in Figure 9, the waste passes through a shredder and magnetic separator, then a second shredder and trommel screeners, and finally a density separator. The emerging waste is 96% organics and paper as compared with 76% for the incoming waste and 24% of the incoming organic and paper materials are lost during the sorting process. Metals are reduced by 100%, plastics are reduced by 93%, and glass is reduced by 98%. The digestate that exits an anaerobic digester contains undigested organics that will continue to break down if not treated further. This can lead to methane emissions typically not accounted for when analyzing the environmental impact of AD. In the EU and particularly in Germany, where the composition of OFMSW entering a landfill is tightly regulated, extensive posttreatment processing is incorporated into the AD facility. This eliminates transportation costs which could be quite high considering the relatively high MC (40-50%) of the exiting digestate.

Figure 9. Mass Balance of the Treviso Wastewater Treatment Digester


Source: Soil Water & Life Solutions

It should be noted, however, that the inorganic materials separated from the incoming MSW stream still have to be transported to a processing facility, typically a material recovery facility or landfill. Dewatering units allow for the re-capture of process water which can provide inoculants and reduce the cost of adding water to the digester. Novel digesters can subject digester paste to a steam treatment step followed by a second digester in order to produce a high quality peat for use as a planting medium.
# 1 2 3 4 5 6 7 8 Separation Technique Trommels and screens Manual separation Magnetic separation Eddy current separation Wet separation technology Air classification Ballistic separation Optical separation Separation property Size Visual examination Magnetic properties Electrical conductivity Differential densities Weight Density and elasticity Diffraction Materials targeted Oversize (paper, plastic); Small (organics, glass, fines) Plastics, contaminants, oversize Ferrous metals Non ferrous metals Floats (plastics, organics); Sinks (Stones, glass) Light (plastics, paper); Heavy (stones, glass) Light (plastics, paper); Heavy (stones, glass) Specific plastic polymers Key concerns Air containment and cleaning Ethic of role, health and safety issues Proven technique Proven technique Produces wet waste streams Air cleaning Rates of throughput Rates of throughput

Table 1. Municipal waste separation techniques


Source: Soil Water & Life Solutions

Copyright 2008 Pytheas Limited

21 February 2008

17

Pytheas Business Guides

4.0 Review of Commercial Anaerobic Digestion Technologies for MSW Treatment A number of commercial vendors have designed a variety of digesters for the global market (Table 2). These commercial systems span the full range of categories of engineered AD systems. The following review attempts to summarize the research reported in the literature for a number of the existing and emerging systems, with special attention paid to the most commercially successful and innovative systems.
Number Of Plants 8 4 23 1 1 13 4 17 2 4 38 8 2 3 22 10 Capacity range (tons/annum) 3,000 55,000 90,000 180,000 1,000 150,000 35,000 100,000 10,000 90,000 6,000 60,000 3,000 120,000 40,000 150,000 50,000 200,000 1,000 110,000 15,000 150,000 24,000 30,000 25,000 87,6000 10,000 270,000 3,000 230,000 x x x x x x x x x x x x x x x x x x x x x x x x x X X x x x x x x x x x x x Number of stages 1 2 X x x Total solids content <20% >20% x x x x x Operating Temperature 35C 55C x x x x x x x

Process System Name AAT ArrowBio BTA Biocel Biopercolat Biostab DBA-Wabio Dranco Entec Haase Kompogas Linde-KCA/BRV Preseco Schwarting-Uhde Valorga Waasa

Table 2. Commercial anaerobic digester technologies with large scale reference plans in Europe
Source: Nichols, C.E., Overview of anaerobic digestion technologies in Europe. BioCycle. 2004. 45(1): p. 47-53 and SWLS

4.1 Single-stage Wet Systems Single-stage wet systems have been built by a number of different companies throughout Europe. Since this was the most familiar configuration from wastewater treatment, it was one of the first systems tested on OFMSW. Below, the Wabio and Waasa systems are described in more detail, but other companies have also provided components and full scale systems to many wet OFMSW digesters, most notably Biotechnische Abfallverwertung GmbH & Co. KG (BTA), Ecoenergy Gesellschaft fr Energieund Umwelttechnik mbh (Schubio), and Linde-KCA-Dresden GmbH. 4.1.1 The Wabio Process For the first time fermentation of municipal solid waste on industrial scale has been implemented in 1989 in Vaasa, Finland with the Wabio-Process. The Wabio-Process was developed by Outokumpu EcoEnergy Oy and the first plant has been in continuous operation since then. Outokumpu EcoEnergy also licensed its Wabio to Deutsche Babcock Anlagen (DBA) which erected in 1992 in Bottrop, Germanys first fermentation plant for biowaste. After DBA merged with the company Steinmller in 1999, the DBA-Wabio-Process was abandoned. The Wabio-Process is an anaerobic digestion process that operates at higher feed rates than other digestion systems. In the processors, inert material and both lighter and heavier fractions can be separated from the reactor feed, thus improving the quality of the humus. In the biothermal method waste material is pretreated so that the biologically gooddecomposable material, the so called compost fraction, and the material suitable for combustion, the refuse-derived fuel, are separated. In the pretreatment also magnetic metals are removed from the waste material. The compost fraction is decomposed biologically. The raw material is the compost fraction of municipal solid waste material for the compost and the

Copyright 2008 Pytheas Limited

21 February 2008

18

Pytheas Business Guides

sludge from the sewage disposal works and the final product is humic earth and a considerable amount of biogas. The refuse-derived fuel (RDF) of the biothermal energy method is a combustion material which is homogenized in the pretreatment and of which heat content is about 20% higher than that of wooden chips. The refuse-derived fuel can advantageously be burned for example in the combination of a fluidized bed boiler and an after-burning chamber (Figure 10). In order that the combustion of the municipal solid waste materials should be safe in regard to the environment, on one hand a sufficient low combustion temperature is required so that nitrogen oxides are not simultaneously created when the concentration of carbon monoxide is small, and on the other hand the temperature must be sufficiently high that the organic compounds are fully decomposed. In the Wabio-Process (see Figures 11 and 12), when using biogas as a burning material in the after-burning chamber of the combustion gases of the refuse-derived fuel, the temperature can be adjusted advantageous for the essentially complete burning and the decomposing of the harmful components, such as the polyaromatic hydrocarbons (PAH) and the polychlorated dibenzo-p-dioxines (PCDD) and the dibenzo-furans (PCDF), which are possible in the municipal solid waste material. The after-burning chamber used is shaped so that it forms an essentially closed system with the fluidized bed boiler used in the treatment of the refuse-derived fuel and that after the connection between the fluidized bed boiler and the afterburning chamber the gas flowing is essentially going downwards. The biogas used as a burning Figure 11. Wabio - Fluidized bed boiler and aftermaterial in the after-burning chamber consists 60- burning chamber Source: United States Patent US4934285 65% methane and the balance carbon dioxide and is fed into the after-burning chamber through burners positioned perpendicular to the direction of the gas flow in the after-burning chamber. The highest point at which biogas used as a burning material is fed into the afterburning chamber is advantageously sufficiently low down that the temperature at the connection between the fluidized bed chamber and the after-burning chamber is not increased so that ash will not be fused at the level of the connection between the fluidized bed chamber and the afterburning chamber.

Figure 10. RDF pellets


Source: SWLS

Figure 12. The Wabio Process Basic Flow Diagram


Source: Outokumpu EcoEnergy Oy Modified by SWLS

When the gas flow is essentially downwards in the after-burning chamber, the ash that is at least partly fused under the treatment of the combustion gases can be removed from the lower part of the chamber while

Copyright 2008 Pytheas Limited

21 February 2008

19

Pytheas Business Guides

the gases that are essentially free of the components harmful for the environment can be conducted to a further treatment. 4.1.2 The Waasa Process (by Deutsche Babcock Anlagen) The Waasa system, built in early 1990s was one of the original MSW digesters. Today there are at least ten operational Waasa plants in Europe (see Table 2). The Waasa system consists of a vertical pulper that homogenizes the incoming MSW and removes floating debris from the surface and sunken grit from the bottom of the pulper. Densityfractionated MSW is then pumped to the prechamber of a continuously stirred tank reactor (see Figures 13 and 14). The pre-chamber helps alleviate short circuiting and an inoculation loop ensures that incoming waste is exposed to microorganisms in order to minimize acid buildup.

Figure 13. The Waasa System Flow of methane fermentation process


Source: Deutsche Babcock Anlagen

For the organic fraction of MSW to be used in this system, it must undergo pretreatment in a pulper which shreds, homogenizes, and dilutes the material to the desired concentration of total solids (10 to 15% TS). Recycled process water and some fresh makeup water are used in the dilution. The slurry is then digested in large complete mix (completely stirred) reactors. The pretreatment required to obtain adequate slurry quality while removing coarse or heavy contaminants is complex and inevitably incur a 15 to 25% loss of volatile solids from the portion reaching the digester. Mechanical impellers and injection of a portion of the biogas into the bottom of the reactor tank are used to keep the material continuously stirred and as homogenous as possible. To reduce shortcircuiting of the feed (i.e., passage of a portion of the feed through the reactor with a shorter retention time than that for the average bulk material), a prechamber within the main reactor tank is used. Fresh material from the pulper Figure 14. The Waasa Process Basic Flow Diagram enters the prechamber along Source: Soil Water & Life Solutions with some of the biomass from the main tank for inoculation. The prechamber operates in plug flow taking a day or two before the material makes its way into the main reactor, thus ensuring all material entering the process has a guaranteed few days retention time. Even with the prechamber arrangement, enough shortcircuiting occurs that all pathogens are not eliminated requiring a pasteurization step in the pretreatment. Steam is injected in the pulper to maintain feed at 70C for one hour. The process can be operated at both thermophilic and mesophilic temperatures and both types can run in parallel (the thermophilic process has a retention time of 10 days while 20 days is required in the mesophilic design). The operational performance indicates that gas production is in the range 100 to 150 m/ton of biowaste added, volume reduction of 60%, weight reduction 50 to 60%, and a 20 to 30% internal consumption (heat) of biogas. The

Copyright 2008 Pytheas Limited

21 February 2008

20

Pytheas Business Guides

digestate can be further treated by aerobic composting, but this depends on the waste quality. The Waasa plant located in Groningen, Netherlands, has four 2,740 m (725,000 gal) tanks treat 92,000 MT/y (101,000 tons/y) of OFMSW out of an initial 250,000 MT/y (275,000 tons/y) of raw MSW. This system produces 0.10-0.15 m/kg (3.2-4.8 scf/lb) biogas from wet source-separated waste, with a weight reduction of 50-60%. This is a relatively high biogas yield, indicating high digestibility of the feedstock and good conversion efficiency in the digester. The typical OLR for a single-stage wet system is 4-8 kg VS/m p.a. (0.033-0.066 lbs VS/gal/y). Assuming 15% of the reactor volume is gas head space, the working volume would be 9,350m (2,470,000 gal), thus the wet loading rate would be 27 kg/d (59 lbs/d) and the resulting VS content would be 20-40%. Assuming 30% VS content and a biogas yield per wet ton of 125 m (4,410 scf), the average specific biogas yield would be 0.417 m/kg VS (13.4 scf/lb). 4.1.3 The BIMA Digester (by Entec Biogas GmbH) Entec Biogas GmbH of Austria builds digesters that treat primarily agricultural, industrial, and municipal wastewater. One system designed for Schaalsee Biogas & Recycling GmbH in Kogel, Germany treats food and restaurant waste from Hamburg and Mecklenburg Vorpommern in two 2,600 m (690,000 gal) constantly stirred tank reactors. The operation of the system mirrors that of the Waasa digester. The company also designed a self-mixing system known as the BIMA digester which eliminates mechanical mixing by utilizing the pressure differential between two chambers within the reactor. The two-chamber system uses the produced biogas to create a level difference in the chambers and in this way builds up a mixing pressure of up to 500 mbar. The turbulent mixing occurs against the biogas production in intervals of 4-10 times a day. Ideal applications of this system are high solid sludge and waste, such as in the sewage sludge treatment, treatment of organic solid wastewater, manure, organic household and industrial waste, etc. 4.2 Single-stage Dry Systems In dry, or high-solids, systems, the digester contents are kept at a solids content of 20-40% TS (equivalent to 60-80% MC). Handling material at high solids concentration requires different pre-treatment and transfer equipment (i.e., conveyor belts, screws, and special pumps for the highly viscous streams). Research in the 1980s indicated that biogas yields and production rates for single-stage dry systems were as high as or greater than that of wet systems. The challenge of dry systems is handling, mixing, and pumping the high-solids streams rather than maintaining the biochemical reactions. Although some of the handling equipment (such as pumps capable of handling high-solids slurries) may be more expensive than those for wet systems, the dry systems are more robust and flexible regarding acceptance of rocks, glass, metals, plastics, and wood pieces in the reactor. These materials are not biodegradable and will not contribute to biogas production but they generally can pass through the reactor without affecting conversion of the biomass components. The only pretreatment required is removal of the larger pieces (greater than 5 cm [2 in]), and minimal dilution with water to keep the solids content in the desired range. This allows for reduced sorting equipment costs which can offset some of the additional material handling expenses. Because of their high viscosity, loading rate, and rapid hydrolysis, materials in dry reactors move via plug flow (materials added on one end of the digester push older materials toward the opposite end), and the incoming feedstock needs to be inoculated or mixed to avoid

Copyright 2008 Pytheas Limited

21 February 2008

21

Pytheas Business Guides

localized acid buildup. Two of the most commonly used commercial-scale designs inoculate the feedstock by mixing it with a portion of the digested material, while another incorporates mixing via high-pressure biogas injection (see Figures 15, 16, 17). All three systems operate as plug-flow digesters. 4.2.1 The Dranco Process (by Organic Waste Systems nv) The Dranco process was developed in the late 1980s. It is a high-solids, single-stage anaerobic digestion system that operates at thermophilic temperatures (Figure 15). Feed is introduced into the top of the reactor and moves downward to the conical bottom where an auger removes digestate. A fraction of the digestate is transferred to the mixing pump where it is blended with fresh feed to inoculate the material and steam to bring the feed to the working temperature. The rest of the digestate is dewatered to produce process water and press cake.

Figure 15. The Dranco Process Schematic & Mass Balance


Source: Organic Waste Systems nv Modified by SWLS

There is no mixing within the reactor, other than that brought about by the downward, plugflow movement of the waste and some biogenic gas that bubbles upwards. The press cake contains active bacteria, some ammonia and undigested solids and must be aerobically stabilized for use as agricultural compost. Source separated household and industrial wastes are preferred in order to maintain the quality of the compost. In the Dranco digester of a plant in Brecht, Belgium a high average loading rate of 15 kg VS/m/d (0.13 lbs/gal/d) was maintained over a one year period. The conditions inside the reactor were 35% TS and 14 day hydraulic retention time (HRT). The performance of the Brecht plant was reported as 65% VS destruction with a 0.103 m/kg (1.65 scf/lb) wet weight biogas yield. The TS content in the feedstock was reported at 40% and the VS content (as a percentage of TS) was 55%. By inference, the specific biogas yield for the system was 0.468 m/kg VS (7.50 scf/lb VS). This relatively low yield along with the relatively low VS destruction may indicate that a large portion of the VS loaded was recalcitrant which explains how such a high loading rate was achieved. To support this theory, it was reported that the waste composition was 15% kitchen waste, 75% garden waste, and 10% paper, whereas a Dranco system in Salzburg, Austria treating 80% kitchen waste and 20% garden waste achieved a biogas yield of 0.622 m/kg VS (9.96 scf/lb). The VS destruction and OLR were not reported, but elsewhere it has been stated that the typical Dranco system is designed for 12 Kg/VS/m/d (0.1 lbs/gal/d). As of February 2008, 18 commercial Dranco system plants exist in Belgium, Spain, Germany, Italy, Switzerland and Japan. 4.2.2 The Valorga Process (by Valorga International S.A.S.) The Valorga process was developed in 1981 to treat organic solid waste and accepts MSW after appropriate separation of the recalcitrant fraction. A high-solids digester is fed with

Copyright 2008 Pytheas Limited

21 February 2008

22

Pytheas Business Guides

OFMSW that has 25-30% TS content adjusted using steam for heating and process water for diluting the incoming feed as needed. Mesophilic or thermophilic systems are used depending on feedstock and economics (Figure 16). The reactor is a continuous single-stage modified plug-flow reactor. Typical plug-flow reactors involve only natural mixing, but the Valorga digester uses pressurized biogas for mixing. This eliminates the need for an inoculation loop. The reactor consists of a vertical outer cylinder with an inner wall extending to about 2/3 of the diameter of the tank (see Figure 16). Material enters at the bottom on one side of the inner wall and must flow around the wall before it can exit. The retention time is on the order of three weeks. Biogas is injected in the base of the reactor and the bubbles serve as a means for mixing and keeping solids suspended. The digestate is dewatered and typically composted. Feedstock, however, Figure 16. The Valorga Process - Basic Flow Diagram Source: Valorga International S.A.S. Modified by SWLS with less than 20% percent TS do not always perform well in the Valorga system because dense grit particles settle out too quickly and clog the gas recirculation vents. Biogas yields have been reported in the range of 0.220.27 m/kg VS (7.058.65 scf/lb VS) which corresponds to 0.800.16 m/wet kg (2.6 5.1 scf/wet lb), indicating a VS content in the OFMSW of 35-60%. The solid retention time is 18-23 days and post digestion solids composting takes about two weeks. In the Varennes-Jarcy Valorga facility, the plant treats a total of 100 MT/y (110 tons/y) of MSW: 70 MT/y (77 tons/y) mechanically sorted OFMSW (MS-OFMSW) and 30 MT/y (33 tons/y) source sorted OFMSW (SS-OFMSW). A remotely operated mechanical claw loads the waste into a 50 m (160 ft) rotating drum with a 2-3 day retention time. Although the drum is not aerated, the temperature of the MSW increases indicating the possibility of biological activity to help break down the organic fraction into smaller particles which are screened out. These fines are sent to a dosing unit for storage and steam heating prior to being pumped into one of the three 4,000 m (1 million gal) reactors. Under typical operating conditions the SS-OFMSW and MS-OFMSW are loaded into separate tanks and the digestate is treated separately, allowing plant operators to control the quality of the compost produced. Dewatering occurs in three steps resulting in process water containing less than 3% TS. The solids are transferred to an enclosed aeration bed where air heated by waste generator heat is blown through the curing piles and sucked through vents in the roof to a scrubber and biofilter. Large automated mechanisms turn the compost and transfer it to a maturation bed after 2-3 weeks. As of February 2008 21 Valorga plants are operating in Spain, Germany, Italy, Switzerland, and the Netherlands. 4.2.3 The Kompogas Process (by Kompogas AG) Unlike the Dranco and Valorga single-stage dry digesters, the Kompogas system utilizes a horizontal plug flow digester with internal rotors to assist in degassing and homogenizing the waste (see Figure 17 below).

Copyright 2008 Pytheas Limited

21 February 2008

23

Pytheas Business Guides

The system is prefabricated in two sizes: 15,000 or 25,000 MT/y (16,500 or 27,600 tons/y). Larger capacities can be acquired by combining the units in parallel. The internal MC has to be carefully maintained at 7277% in order for the system to flow properly; therefore some of the process water and/or digestate is mixed with incoming OFMSW. This also ensures that incoming feed is inoculated in order to prevent excessive acid buildup near the front end of the digester.

Figure 17. The Kompogas Process - Basic Process Diagram


Source: Kompogas AG Modified by SWLS

In more detail, in order to produce energy from garden and kitchen waste, organic waste is collected separately and freed of foreign matter by screening and manual sorting. The waste is then conditioned (shredding, etc.) and transported to an intermediate storage facility. The storage capacity of the intermediate facility guarantees continued and fully automated operation of the fermentation process. During this period of storage, hydrolytic and selfheating processes take place, which favor the ensuing fermentation. Transportation to the fermenter is done by a powerful hydraulic pump. In the enclosed reactor, micro-organisms transform the organic substance in the material into compost and biogas. The process takes place at a temperature of 55 to 60 C. The horizontal design of the Kompogas digester with the plug-flow forward-motion principle ensures the desired retention time of about 20 days is achieved. The green waste is transformed into compost and gas in the fermenter. Depending on the specific composition of the organic waste, between 100 and 140 m (Nm) or at about 0.11 0.14 m/kg (3.44.2 scf/tons) wet weight of biogas are produced per ton with a methane content of approximately 60%. This corresponds to about 70 liters of petrol. Another hydraulic pump extracts the fermentation residue and feeds it to the dewatering system. Presses then separate the fermentation residue into fresh compost and press water. In order to neutralize any odors, the fresh compost is aerated in a compost maturing hall thereby passing from an anaerobic to an aerobic state, using odor control systems with biofilters. As of February 2008 30 systems are operating in Europe, 2 in the U.S., 1 in Africa, 2 in Russia, 11 in Asia, and 1 in Australia.

Copyright 2008 Pytheas Limited

21 February 2008

24

Pytheas Business Guides

5.0 Multi-stage Digesters Digesters that are operated in parallel are not multi-stage digesters; when each reactor is a separate single-stage digester. This may be done because of tank size limitations, to simplify management, or to expand capacity of an existing plant. A true multi-stage digester applies different conditions to the reactors in each stage. The difference can be in the organic loading rate (OLR) of each stage, the presence or absence of oxygen, the introduction of an intermediate treatment, or the overall reactor configuration. Many different combinations or factors are possible. There are relatively few commercial, operational multi-stage AD units. Although it was expected that more of the multi-stage systems would be in operation by now due to their higher loading rates, improved process stability, and flexibility, the added complexity and presumed expense of building and operating commercial multi-stage systems have so far negated the yield and rate enhancements. Nonetheless, the potential of multi-stage digesters to improve performance has prompted much research, and a few notable commercial multi-stage digesters have been successful. Some of these use multiple stages for reasons other than separating acidogenesis from methanogenesis. 5.1 The BTA Process (by Biotechnische Abfallverwertung GmbH & Co.) Developed in Germany and applied (via several licensing companies) throughout Western Europe and in select locations in Canada and Japan, the BTA system is one of the oldest and most successful in terms of the number of existing operational digesters. Although small units are single-stage, the majority of the BTA digesters are large (>100,000 MT/y [110,000 tons/y]) multi-stage, wet-wet units. The multistage BTA digester (Figure 18) utilizes a pulper and hydrocyclone much like those employed by the Wabio and Waasa single-stage digesters.

Figure 18. The BTA Process Basic Flow Diagram


Source: Biotechnische Abfallverwertung GmbH & Co. Modified by SWLS

Pulped and density-fractionated MSW passes through a solid/liquid separation unit and leachate is passed directly to a methanogenesis reactor. Solid extract is mixed with process water to bring the MC to 75% and then pumped into a hydrolysis reactor with a residence time of 4 days. Hydrolysis leachate is then transferred into the methanogenesis reactor which has a 2d HRT. Dewatered digestate is then either treated aerobically or disposed. Installations

Copyright 2008 Pytheas Limited

21 February 2008

25

Pytheas Business Guides

with a designed capacity of less than 100,000 MT/y (110,000 tons/y) often utilize the pulper as the hydrolysis tank, eliminating one step in the process. In more detail, the BTA Process comprises two central steps: (a) the hydromechanical pretreatment, and (b) the subsequent biological step towards anaerobic digestion.

Figure 19. Process scheme of the BTA Multi-stage wet fermentation


Source: Biotechnische Abfallverwertung GmbH & Co.

The hydromechanical pre-treatment facilitates efficient removal of impurities as well as complete separation of digestible organic components into an organic suspension. This happens within two core components, the Waste Pulper and the Grit Removal System. Within the Waste Pulper the feedstock is added to pre-filled process water in order to separate the waste mixture into fractions by taking advantage of natural buoyancy and sedimentation forces. Moreover, non-soluble organic components are reduced to fibers by shearing forces and brought into suspension. Thus, heavy materials are fed aside and light materials are skimmed off. In total, three fractions are effectively separated into (a) organic materials, (b) light materials (plastics, foil, textile, wood, etc.), and (c) heavy materials (stone, bones, batteries, etc.). Having passed the Waste Pulper, the organic suspension still contains sand and fine impurities, which are removed by the Grit Removal System. This reliably ensures protection of downstream plant components from wear, silting up, sediments, and obstruction. The cleared bio-suspension is temporarily stored in a suspension tank. This way, the processing stage is decoupled from the digestion itself, keeping the latter independent from the working cycle of the waste reception unit. The organic fraction is digested within the fermenter, generally under mesophilic conditions between 35C and 38C. Fermenter layout and operation mode depend on the kind of feedstock as well as on projectspecific requirements concerning microbiological breakdown and biogas yield. In practice,

Copyright 2008 Pytheas Limited

21 February 2008

26

Pytheas Business Guides

stirring of the fermenter is often hampered with great problems for many systems. The Waste Pre-Treatment ensures a virtually impurity-free organic suspension at high homogeneity. Therefore, by pressing in biogas, it is possible to keep the fermenter load homogenous, i.e., Safe agitation of the entire fermenter volume, high utilisation of the energy input, low susceptibility to faults due to the lack of moving parts within the fermenter, no maintenance work required inside the fermenter, no loss of fermenter volume by sedimentation. Further treatment of the digested substrate can be adjusted according to the respective project. Generally, a decanter centrifuge for continuous separation of solids from liquids is employed. The solid material (ca. 30 % dry matter) is perfectly suitable for the stabilisation and production of quality compost. In most cases the supernatant is utilised as process water and is retained within the cycle. This results in decreased freshwater consumption. 5.2 The Linde-KCA Process (by Linde-KCA-Dresden GmbH) Linde-KCA has built low and high solids (wet and dry) digestion systems and mechanicalbiological treatment systems (MBT) for separated MSW since 1985 and currently has eight digesters operating in Germany, Portugal, Spain, and Luxembourg, mesophilic and thermophilic. MBT systems include aerobic composting systems with Figure 20. Linde-KCA Two-stage Wet Digestion System mechanical manipulation of the Source: Linde-KCA-Dresden GmbH feedstock and intensive aeration. Some systems include intensive aerobic digestion as a pre-process for a feedstock that is later anaerobically digested. The typical dry digester is operated in two stages. The first stage is aerobic and the hydrolysis product is transported via conveyor to a horizontal plug-flow digester with internal rotors for mixing (see Figure 21) and transporting solids to the dewatering unit (although this is a twostage system, the first stage could also be considered an aerobic pretreatment stage apart from the anaerobic digester). Nevertheless, the digester is capable of handling 15-45% TS and generates roughly 0.10 m/wet kg (3.2 scf/wet lb) of biogas. Linde-KCA has designed, built, and operated a facility in Radeberg, Germany, which codigests source separated biogenic wastes from household and industrial sources along with sewage sludge from waste water treatment. The company reports that this codigestion concept enhances degradation of the sewage sludge component of the Figure 21. Linde-KCA Dry Digestion Plug Flow System feedstock (increases biogas Source: Linde-KCA-Dresden GmbH Modified by SWLS production from the sewage sludge) and decreased capital and operating costs compared to those for two separate facilities. Another codigestion facility designed and built by Linde-KCA is located on a dairy farm in Behringen, Germany. The plant takes the low solids manure and co-digests, with grease from restaurant grease traps, solids from pig manure, and a range of other food processing waste

Copyright 2008 Pytheas Limited

21 February 2008

27

Pytheas Business Guides

including brewery sludge, rape seed press cake, and low grade seed potatoes. The facility processes about 100 t/day (wet) of which 75% is dairy cow manure. The facility also produces about 650 kW from two Jenbacher 450 kW gensets; 30% of the power production is used on site for operating the plant and the dairy with the balance sold to the grid. 5.3 The ArrowBio Process (by ArrowBio, ArrowEcology & Engineering Overseas Ltd.) The ArrowBio process integrates separation and preparation (preprocessing) and advanced anaerobic digestion (Upflow Anaerobic Sludge Blanket digestion or UASB) via the medium of water. The water is derived from the moisture content of the waste. Clean recyclables are recovered from mixed waste via gravitational separation in water, and biodegradable organics are extensively converted to methane-rich biogas. It achieves system integration by Source: ArrowBio, ArrowEcology & Engineering Overseas Ltd. exploiting the moisture content of MSW; moisture content levels are typically around 30% by weight. Thus a ton of MSW consists of approximately 1,400 pounds of material and 600 pounds of water (or, 70 gallons of water per ton). In the course of anaerobic digestion, moisture in MSW is liberated as liquid water. The mechanism is that, as organics are converted to methane and carbon dioxide, and the gases bubble out of solution, water originally imparting moistness is left behind in liquid form (Figures 22 and 23). A portion of the water liberated at the back-end biological stage (UASB) is exchanged with vat water at the front-end physical separation/preparation stage. As such, the bioreactors receive fresh substrate, and the vat receives makeup water. Among UASBs benefits are extensive conversion of organics to methane-rich biogas, production of a relatively small amount of clean, stabilized, digestate, and a modest facility footprint. These characteristics stem from integration of the two stages. The front-end physical stage (a) removes grit and other landfillables to recover traditional non-biodegradable recyclables (e.g., plastic bottles and jugs), (b) other secondary material, and (c) to isolate and prepare the biodegradables for UASB digestion at the back-end stage. These functions are performed in unison and are inextricable. The non- biodegradable and biodegradable fractions are separated gravitationally in water as abetted by filter screens and size-reduction devices. Separation in water is far more efficient than in air, owing to the comparative densities (relative buoyancies) of the two fluids. Thus, depending on their solubility or specific gravity and tendency to absorb water, items dissolve, sink, float, or become suspended in the water. Incidental benefits of tipping into water include the suppression of dust, and the neutralization of odors. Neutralization is immediate because odorous compounds are soluble in water. Their biodegradation soon follows as vat water is continuously pumped to the backend enclosed digesters. Also, because the system is watery throughout, any input surges are evened out, contributing to the systems overall resiliency. The load of mixed waste is tipped onto a walking floor, from which it falls into the water vat immediately upstream of a partially submerged rotating paddle. The paddle urges floaters and buoyancy-neutral items forward into the main body of water. Sinkers are diverted to the left and passed sequentially to a bag breaker, magnetic pickup, eddy current device, and a
Figure 22. The ArrowBio Process logic

Copyright 2008 Pytheas Limited

21 February 2008

28

Pytheas Business Guides

pneumatic (vacuum/forced draft) station from which film plastic is swept into ductwork. Ducts from several such stations converge on the cyclone. Thereby, metals and film plastic are removed. Items that escape this processing train the first time around reenter the water vat for another chance to dissolve, float or sink or, if buoyancy-neutral, be suspended in the forward-moving water column. Overflow from the water vat, screened to exclude large items, passes though smaller enclosed trommel screens and thence, according to partitioning criteria, to large and small settlers. In the settlers grit is separated from organics and removed from the system. Meanwhile, larger floaters and buoyancy-neutral items are lifted to a slow speed shredder and thence to the large trommel screen. The overs from this trommel consist mostly of film plastic and are removed at a pneumatic station. The unders (material that passed through screen) are washed into a nonmechanical device for further solubilization and size reduction. Non-soluble substances are thus reduced to a suspension of fine particles whose surfaces are roughened to favor microbial colonization. Thus non-biodegradables are Source: ArrowBio, ArrowEcology & Engineering Overseas Ltd. Modified by SWLS recovered for recycling as secondary material commodities, and soluble and particulate organics come into solution or fine suspension, including food sticking to containers and the contents of unopened diapers. The latter are disrupted in the processing train, freeing the feces and urine-soaked cottony absorbent. Insoluble biodegradable organics (e.g., non-source-separated food-tainted paper products, tough fruit rinds) get increasingly soggy and fragmented, ultimately to the point of passing screens of selected sizes. The organics, now in watery isolation, are pumped to the biological stage. Reciprocally, return-water from the digester system refreshes the separation/preparation water vat. Within half an hour after tipping the last load of the day, the work of the physical separation/preparation element is complete. This part of the plant is then shut down until deliveries resume the next working day. The organic flow (Figure 23) first enters acidogenic bioreactors for several hours of preliminary treatment. There, readily metabolized substances already in solution are fermented (e.g., sugars fermented to alcohols), while certain complex molecules are biologically hydrolyzed to their simpler components (cellulose to sugar, fats to acetic acid). The overflow, rich in such intermediate metabolites, then enters the UASB digestion bioreactor. Operationally, excess biological granules suspended in similarly excess water (both excesses represent growth at the expense of the waste) are transferred to a settling tank. Supernatant is pumped to the physical separation/preparation element as needed for makeup water, or to an aerobic tank for polishing if necessary. Water may be stored or used immediately as in irrigation. The solids are dewatered for use as a stabilized organic soil amendment. When needed to maintain UASB digestion at its optimum temperature of ~35C (95F), some of the biogas is used to fire boilers. Most of the gas is used to fuel a generator,
Figure 23. Waste digestion in the ArrowBio Process Basic Flow Diagram

Copyright 2008 Pytheas Limited

21 February 2008

29

Pytheas Business Guides

via storage in a reservoir. Waste heat from the generator usually suffices to maintain digestion temperature. 5.4 The Biopercolat Process (by WEHRLE Umwelt GmbH) The Biopercolat process by WEHRLE Umwelt GmbH is a dry-wet, two-stage process. The first hydrolysis stage is carried out under partial aerobic conditions with high solids content (see Figure 24).

Figure 24. Schematic waste-flow diagram of the two-stage Biopercolat Process


Source: WEHRLE Umwelt GmbH

Process water is continually percolated through the hydrolysis reactor a horizontal tunnel that slowly rotates in order to slightly aerate the mixture and prevent clogging and channeling. The leachate passes on to the second-stage fermentation reactor an anaerobic plug flow filter filled with support material operating at mesophilic temperature. After two to three days in the percolator, the solids are separated and transferred to an enclosed tunnel composter. The liquid fraction is transferred to the fermentation reactor, and displaced liquid is partly recirculated back through the percolator and partly aerated for disposal as wastewater.

Copyright 2008 Pytheas Limited

21 February 2008

30

Pytheas Business Guides

6.0 Batch Digesters Some of the first dry digesters were envisioned as modified landfills. This resulted in the creation of batch systems that recycled leachate in a manner similar to landfill bioreactors. However, unlike landfill bioreactors the batch digester conditions were more carefully controlled and as a result biogas production rates were higher and retention times were lower. In batch systems, digesters are filled once with fresh wastes, with or without addition of seed material, and allowed to go through all degradation steps sequentially in the dry mode, i.e. at 30-40 % TS. Though batch systems may appear as nothing more than a landfill-in-a-box, they in fact achieve 50- to 100-fold higher biogas production rates than those observed in landfills because of two basic features: (a) Leachate is continuously re-circulated, which allows the dispersion of inoculants, nutrients, and acids, and in fact is the equivalent of partial mixing and (b) batch systems are run at higher temperatures than that normally observed in landfills. Batch systems have up to now not succeeded in taking a substantial market share. However the specific features of batch processes, such as a simple design and process control, robustness towards coarse and heavy contaminants, and lower investment cost (ca. 40%) make them particularly attractive for developing countries. Note, however, that the land area required by batch processes is considerably larger than that for continuously-fed dry systems, since the height of batch reactors is about five-fold less and their OLR two-fold less, resulting in a ten-fold larger required footprint per Ton treated wastes. Operational costs are comparable to those of other systems. The hallmark of batch systems is the clear separation between a first phase where acidification proceeds much faster than methanogenesis and a second phase where acids are transformed into biogas. Three basic batch designs may be recognized, which differ in the respective locations of the acidification and methanogenesis phases. The primary disadvantage of batch digesters is uneven gas production and lack of stability in the microbial population. Sequential and phased batch digesters attempt to surmount these disadvantages, and preliminary lab experiments have revealed complex population dynamics in these systems resulting in the ability to separate useful fermentation products such as hydrogen and organic acids. 6.1 The Biocel System (by Orgaworld bv) The Biocel system was developed in the 1980s and 1990s in Holland at the Wageningen University as a part of the early research on high-solids digestion of MSW. The initial goal of the system was to reduce cost by simplifying material handling and eliminating the need for mixing while simultaneously achieving relatively high loading and conversion rates. Success with the lab-scale system led to construction of a pilot 5 m (1,000 gal) reactor by the early 1990s which was used for more extensive testing of start-up, heating, and leachate recycling. By 1997 a full-scale 50,000 MT/y (55,000 tons/y) plant consisting of a digester and enclosed post-digestion aeration beds had been built in Holland to treat SS-OFMSW. Currently the Dutch company Orgaworld owns and operates the Biocel plant along with several tunnel composting facilities and one AD facility also in the Netherlands. The company plans to increase electricity production to >10 million kWh/y by treating up to their permitted 85,000 MT/y (94,000 tons/y). While batch systems may simplify material handling, they sacrifice control over the biological processes. Because a batch is loaded all at once, the internal conditions change as microbial populations shift in response to the consumption of waste and production of intermediate metabolites. A lag phase occurs as organic polymers break down followed by a sudden drop in pH as organic acids are produced from the hydrolysate. If this pH drop is too severe,

Copyright 2008 Pytheas Limited

21 February 2008

31

Pytheas Business Guides

methanogenesis cannot occur. In a lab study, even after 100 days, only hydrogen and CO2 were produced. The initial lab and pilot studies attempted to mitigate this effect by mixing the incoming feed with digestate from a previous batch, using aerobic pre-treatment, adding buffers, changing the inoculation rate, and altering the leachate recycling rate. At the pilot scale the maximum achievable OLR was 7 kg VS/m/d (0.058 lbs/gal/d) which is similar to continuous high-solids digesters. Published biogas yields for full-scale digesters treating a variety of wet OFMSW types), but at full-scale the average OLR was 3.6 kg VS/m/d (0.030 lbs/gal/d) which is closer to the OLR of low-solids systems. Another advantage of batch systems is the low water input requirement. However, during lab and pilot-scale studies it was found that no leachate drained from the waste when the TS content was higher than 35%, and methanogenesis was inhibited by lack of contact between bacteria and substrate. In comparison, continuous high-solids systems achieve stabile digestion at 25-35% TS. However, the relatively low moisture content of the feedstock makes it more difficult to heat. It was found that loading cold feed and allowing it to slowly reach the target temperature resulted in doubling the digestion time required at the pilot scale. Therefore cold feed has to be preheated. The full-scale Biocel system is comprised of fourteen 720 m (190,000 gal) leach bed reactors, each loaded in 480 m (130,000 gal) batches keeping the pile 4 m (13 ft) high to avoid excessive compaction. Reactor temperature is kept at 35-40C (95-104F) by heating leachate which is sprayed over the pile. The digester retention time is 21 days and the posttreatment aeration bed retention time is 1-3 weeks. At full-scale, a complex system of vacuums and pumps from the generator exhaust system flushes oxygen out of the headspace and captures odors when opening digester doors for loading and unloading. Fresh MSW is sorted manually and loaded by shovel without any pretreatment for size reduction or screening. For each MT (1.1 tons) of MSW loaded, the system produces 70 kg (150 lbs) of biogas, 120 kg (265 lbs) of water vapor, 500 kg (1100 lbs) of compost, and 230 kg (510 lbs) of wastewater. At the pilot scale, the biogas yield was 0.70 m/kg wet waste (2.2 scf/lb) which is lower than typical but no comparison has been made with continuous systems using the same feedstock. 6.2 The SEBAC System (Sequential Batch Anaerobic Composting) The SEBAC system was developed in the early 1990s at the University of Florida with the goal of eliminating mixing and minimizing handling while maintaining high conversion rates and system stability. Similar to the Biocel process, the SEBAC system consists of twoor three-batch, leach-bed reactors with leachate recirculation by sprayer, but unlike the Biocel system, the SEBAC digesters are loaded in sequence such that leachate can be transferred between reactors. OFMSW is roughly chopped to 10 cm and placed in a batch reactor. Leachate from a mature reactor is sprayed onto the fresh Figure 25. The SEBAC System Process Diagram Source: Chynoweth, D.P., J. Owens, D. Okeefe, J.F.K. Earle, G. Bosch, and R. Legrand, material and recycled to the top Sequential batch anaerobic composting of the organic fraction of municipal solid-waste. of the pile until methanogenesis Water Science and Technology. 1992. 25(7): p. 327-339. stabilizes. The reactor (see Figure 25) is then switched over to internal recirculation until methane production slows as

Copyright 2008 Pytheas Limited

21 February 2008

32

Pytheas Business Guides

the batch matures. In theory this allows organic acids to be applied to mature reactors with active methanogenic populations, and it allows the microbes from mature reactors to be sprayed on fresh waste as an additional inoculant. In practice, the dynamics of leaching are not well understood, thus the system appears to be difficult to control. At the lab scale, the SEBAC process had difficulty starting when loaded with pure food waste. Bulking agents were required to prevent compaction and allow leachate to drain through the pile. However, even with the most successful startup scheme, the highest biogas production rate was not achieved until 50-60 days after loading. It was not clear from this study what the overall methane yield was, but an earlier pilot scale study reported yields of 0.16 and 0.19 m CH4/kg VS (2.6 and 3.0 scf/lb) with retention times of 21 and 42 days respectively. This is much lower than published thermophilic methane yields for continuous digesters treating OFMSW. The waste stream contained 60% paper and cardboard, 10% plastic, and 6% yard waste, and yields represented 8090% of the ultimate methane potential. However, it appears that large quantities of paper and plastic are required to allow proper leaching. No known full-scale SEBAC systems have been built but research on the system continues. A new design (SEABAC II) tailored for the low gravity environment of manned space missions reduced the volume requirement by 60% by eliminating the headspace and intentionally wetting and compacting the feedstock, which also allowed for forced leachate pumping. The prototype SEABAC II digester operating at 35C (95F) was able to ultimately achieve a methane yield of 0.3 m CH4/kg VS (4.8 scf/lb VS) in about 14 days from a mix of rice, paper, and dog food. 6.3. The Anaerobic Phased Solids (APS) Digester Like the SEBAC system, the APS Digester uses batch loading to stimulate rapid organic acid production in a two-stage digester system. However, the APS Digester system avoids the problems caused by using leach bed reactors by combining highsolids reactors for the first stage with a low-solids mixed biofilm reactor in the second stage. The high-solids reactors are loaded in phased batches, and the leachate from the batch reactors is continuously circulated through a single low-solids digester. In theory, batch loading simplifies material handling, and because the hydrolysis reactors are highsolids digesters, they can handle Figure 26. Schematic of APS Digester System Source: University of California, Davis relatively large inorganic contaminants. Leachate recirculation prevents solids from fouling the wet methanogenesis reactor. Because the batches are phased, the leachate contains a relatively constant concentration of organic acids. A pilot demonstration plant for the APS Digester system with a capacity of about 1-2 tons (dry) per day of organic waste has been under development at the University of California, Davis (Figure 26).

Copyright 2008 Pytheas Limited

21 February 2008

33

Pytheas Business Guides

The pilot plant consists of five 38 m (10,000 gal) vertical steel cylindrical tanks. The four hydrolysis tanks possess hot water jackets for heating the reactor contents. The methanogenesis tank is heated via thermal heat exchange with a natural gas/biogas powered boiler. The system was designed to operate at both mesophilic and thermophilic temperatures. Fresh feedstock is loaded via a chopper pump and hydraulic ram system and mixing is accomplished using high-velocity liquid jets. The gas collection system was designed to separately collect hydrogen-rich biogas from the hydrolysis reactors. System operators can monitor and control the digesters via remote computer system. The overall design objective was to build a low maintenance, yet flexible, two-stage system with no internal moving or custom-built parts. In laboratory studies, the system was able to digest rice straw with a lingo-cellulosic content (lignin, cellulose and hemicellulose) of 85% and achieve 40-60% solids reduction with a biogas yield of 0.4-0.5 m/kg VS (6.4-8.0 scf/lb VS) which is on par with yields seen for much more highly degradable substrates. Laboratory studies have been conducted with other substrates as well such as food waste, OFMSW, food processing wastes, and animal manure. The biogas yields from the food waste collected from restaurants and green waste (grass clippings) were 0.60 and 0.44 m/kg VS (9.6 and 7.0 scf/lb VS), respectively with 12 day solids retention in the digesters.

Copyright 2008 Pytheas Limited

21 February 2008

34

Pytheas Business Guides

7.0 Digester Performance 7.1 Biogas yield Digester performance depends greatly on reactor configuration and OFMSW source. Many scientific reports indicate performance of MSW digesters in terms of biogas yield per wet weight of MSW treated (see Table 3). Full scale plants typically achieve biogas yields of 0.10 0.15 m/wet kg (3.2 to 4.8 scf/wet lb). However, comparisons of systems based on yield per wet weight MSW assume consistency of MSW and biogas composition. Biogas can contain from 50-70% methane by volume, too wide a range for accurately estimating energy potential. Methane yield is more useful than biogas yield but requires accurate CO2 or CH4 detectors or expensive lab tests (i.e. gas chromatography). Also MSW can vary widely in MC and digestibility, Methane yield is based largely on the amount of paper, grass, wood, and other more useful than lignocellulosic material contained. Therefore the scientific literature biogas yield but requires accurate typically reports yield in terms of methane yield per dry weight of CO2 or CH4 detectors volatile solids. This assumes that volatility is a proxy for or expensive lab biodegradability, but lignocellulosic material tends to be less tests biodegradable than other volatile compounds. A better proxy for biodegradability is the five-day biological oxygen demand (BOD-5), but the standard method for measuring the BOD-5 content of a feedstock takes too long to be used for measuring the ongoing biogas yield of a digester, thus it is rarely reported in the literature. Therefore, when comparing systems treating different MSW streams, one must be careful to take note of compositional differences. In addition, biogas yield says nothing about the rate of methane production. Reactor efficiency is more important than overall yield for determining financial performance of a system. The overall biogas production rate and the MSW throughput rate, or organic loading rate, are important determinants of a systems efficiency. For comparing system efficiencies, compositional feedstock differences are important. The maximum achievable OLR, however, is highly dependent on reactor configuration. High temperature reactors are commonly referred to as highrate reactors because of the increased reaction rate.
Table 3. Published biogas yields for full scale digesters treating a variety of wet OFMSW types
Source 1: Saint-Joly, C., S. Desbois, and J.P. Lotti, Determinant impact of waste collection and composition on anaerobic digestion performance: industrial results. Water Science and Technology. 2000. 41(3): p. 291-297 Source 2: Bolzonella, D., P. Pavan, S. Mace, and F. Cecchi, Dry anaerobic digestion of differently sorted organic municipal solid waste: a full-scale experience. Water Science and Technology. 2006. 53(8): p. 23-32 Source 3: Soil Water & Life Solutions Source 4: Gallert, C., A. Henning, and J. Winter, Scale-up of anaerobic digestion of the biowaste fraction from domestic wastes. Water Research. 2003. 37(6): p. 1433-1441 Source 5: Beck, R.W., Final report: Anaerobic digestion feasibility study for the Bluestem Solid Waste Agency and Iowa Department of Natural Resources., in Final Report: Anaerobic Digestion Feasibility Study for the Bluestem Solid Waste Agency and Iowa Department of Natural Resources. 2004. Bluestem Solid Waste Agency

Two-stage reactor configurations were developed in order to increase the achievable OLR. Increasing the OLR often leads to disproportionate increases in organic acid production due to biological growth rate and pH tolerance differences between acid producing and acid consuming microbes. An upper limit on OLR seems to exist at around 15 kg VS/m (0.125 lbs VS/gal), but the achievable OLR can be greatly affected by the overall digestibility of the waste. The biogas or methane production rate in itself is not very useful because it depends on the loading rate, but by combining the OLR and biogas yield the reactor efficiency can be determined in terms of biogas production rate per unit of reactor volume.

Copyright 2008 Pytheas Limited

21 February 2008

35

Pytheas Business Guides

Comparing the performance of industrial scale OFMSW digesters treating different waste streams is difficult, especially since companies tend to protect performance data. The efficiency of a digester in terms of gas production per unit digester volume can be calculated by multiplying the OLR by the biogas yield. Hence it can be seen that even though wet digestion of SS-OFMSW at 55C (130F) achieved a biogas yield of about 0.8 m/kg VS (12.8 scf/lb VS), the OLR was only 2 kg VS/m/d (0.02 lbs VS/gal/d) and the biogas production rate was only 1.6 m/m/d (0.21 scf/gal/d). For comparison, the mesophilic digestion of food waste resulted in much lower biogas yields of 0.45 and 0.3 m/kg VS (7.2 and 4.8 scf/lb VS), but at OLR of 6 and 9 kg VS/m/d (0.05 and 0.075 lbs VS/gal/d) the biogas production rate was 2.7 m/m/d (0.36 scf/gal/d) which is 70% higher. Based on this analysis, dry digestion of kitchen waste + paper had the highest biogas yield per unit reactor volume closely followed by dry thermophilic digestion of OFMSW from a pilot scale study using simulated OFMSW composed of paper, yard and food waste. As per Table 4 below, HRT ranges from 9-30 days, with the average of thermophilic reactors being 66% that of mesophilic digesters (10-16 vs. 15-25 days). No clear difference exists between wet and dry digesters in terms of HRT, but the obtainable OLR is about three times higher for dry digesters and two times higher for thermophilic digesters. The achievable OLR for SS-OFMSW is about 65% that of MS-OFMSW, which is to be expected since the higher digestibility of SS-OFMSW leads to greater acidification of the digester.
Temperature Mesophilic Mechanically sorted Dry Semi-dry Wet Food industry (Source separated) Dry Semi-dry Residential (Source separated) Dry Wet Thermophilic Mechanically sorted Dry Semi-dry Food industry (Source separated) Dry Residential (Source separated) Dry Semi-dry Grand Total Table 4. Reactor conditions typical of single-stage OFMSW digesters
Source: Cecchi, F., P. Traverso, P. Pavan, D. Bolzonella, and L. Innocenti, Characteristics of the OFMSW and behaviour of the anaerobic digestion process., in Biomethanization of the organic fraction of municipal solid waste, J. Mata-Alvarez, Editor. 2002. IWA: Cornwall, U.K. p. 141-180 Modified by Soil Water & Life Solutions

Substrate

TS

Min HRT 15.1 15.3 17.0 15.0 14.0 14.5 17.0 12.0 15.5 17.0 14.0 10.8 9.0 12.0 6.0 12.0 12.0 12.0 12.0 12.0 13.3

Max HRT 24.9 26.7 30.0 20.0 30.0 19.5 25.0 14.0 27.5 25.0 30.0 16.2 17.5 20.0 15.0 16.0 16.0 15.0 16.0 14.0 21.3

Min OLR 3.8 4.9 6.0 6.0 2.6 3.5 4.0 3.0 2.5 4.0 1.0 6.6 7.5 9.0 6.0 6.0 6.0 6.0 4.0 8.0 5.0

Max OLR 5.9 7.0 9.0 8.0 4.0 5.0 6.0 4.0 5.0 6.0 4.0 12.4 17.5 15.0 20.0 9.0 9.0 9.0 6.0 12.0 8.6

A BTA (wet process) full-scale wet digester (see Table 3) was reported to reach OLRs comparable to those typical of high-solids dry systems. The digester was designed to accommodate 8,000 MT/y (8,800 tons/y) but was initially accepting 7,200 MT/y (8,000 tons/y). After a study demonstrated that a lab-scale digester could sustain OLR as high as 15 kg COD/m/d (0.125 lbs COD/gal/d), the full-scale digesters loading rate was increased to 12,000 MT/y (13,000 tons/y). The lab study initiated the digester at an HRT of 20 days and reduced the HRT to 5.7 days in 5 steps (20-13-10-8-7-5.7) of roughly 15-20 days each. Whenever HRT was reduced, OLR was increased by adjusting the amount of extra water added to the feed. A maximum OLR of 15 kg COD/m/d (0.125 lbs/gal) was sustained in the

Copyright 2008 Pytheas Limited

21 February 2008

36

Pytheas Business Guides

lab for 45 days with consistent biogas production of 50-60 L/d (1.8-2.1 scf/d). In addition to biogas production, VFA, VS reduction, and pH were monitored to ensure that digestion was stable. COD reduction and biogas yield both increased linearly with OLR at the lab scale. Upon successful completion of the lab study, loading rate of the operating full scale digester was increased from 8 to 15 kg COD/m/d (0.067-0.125 lbs COD/gal/d) stepwise over two months. The digester was run for five months at 15 kg COD/m/d (0.125 lbs COD/gal/d) with an average biogas production rate of 4.5 m/m/d (0.6 scf/gal/d). The study demonstrated that a low-solids system could be loaded at higher OLR than typical for such systems, although these results may not necessarily generalize to other systems. 7.2 Life Cycle Analysis Anaerobic digesters are environmentally friendly but often cost more than alternative treatment technologies. However, wise management decisions take whole-system evaluations of technology options into account. The standard method for performing such analyses is the life cycle assessment (LCA) methodology outlined in ISO 14040 of the International Standards Organization. The LCA methodology begins with defining the scope and boundaries of the system under consideration followed by an inventory analysis and impact assessment. The outcome of an LCA is specific to the system under consideration, but if many analyses have similar outcomes generalizations can be made with caution. Decision models have been developed based on LCA results for making solid waste management decisions.
AD produces less air and water pollution than aerobic composting or landfilling of OFMSW; also AD has a positive net energy balance while other technologies, including landfilling with gas collection, consume net energy

Several scientific studies have investigated the overall environmental and economic impacts of anaerobic digestion using the LCA methodology. All of the studies found that AD produced less air and water pollution than aerobic composting or landfilling of OFMSW. A Canadian LCA compared AD, open windrow composting, and landfilling of MSW where landfills with and without energy production were included. The report found that AD produced less air and water pollution than any of the other technologies. The study also found that over the life of the project, AD had a positive net energy balance, while the other technologies including landfilling with gas collection consumed net energy (Table 5). However, the study did not seem to account for the embodied energy of construction or transportation. For example, additional transportation would be required for an AD facility located at a centralized site some distance from the landfill. A centralized digester, however, would serve multiple landfills. The costs and benefits of centralized OFMSW treatment would have to be evaluated for an entire region.
AD vs. LF Energy consumption (GJ/y and mmBTU/y) GHG emissions (MT/y and tons/y CO2 eq.) NOx (MT/y and tons/y) SOx (MT/y and tons/y) PM-10 (MT/y and tons/y) VOC (MT/y and tons/y) Lead (kg/y and lbs/y) -400,000 -121,908 -48.8 -68.4 -58.4 -8.6 -88.3 Metric Units AD vs. WC -430,370 -84,795 -50.3 -74.6 -50.8 -3.8 -93.0 WC vs. LF +32,228 +38,170 +1.5 +6.21 -7.6 -4.7 +4.72 AD vs. LF -380,000 -134,379 -53.8 -75.4 -64.4 -9.5 -194.7 Standard Units AD vs. WC WC vs. LF -407,910 -93,470 -55.4 -82.2 -56.0 -4.2 -205.0 +30,546 +42,075 +1.7 +6.83 -8.4 -5.2 +10.4

Table 5. Comparison of the energy use and emissions from anaerobic digestion (AD), open windrow composting (WC), and landfilling without energy recovery (LF)

Source: Haight, M., Assessing the environmental burdens of anaerobic digestion in comparison to alternative options for managing the biodegradable fraction of municipal solid wastes. Water Science and Technology. 2005. 52(1-2): p. 553-559 Modified by SWLS

The model also assumed that excess electricity could be sold to the local power grid. The study did include emissions from post-digestion treatment of residuals, and the reductions in emissions due to AD were high (see Table 5 above). Interestingly, open windrow composting

Copyright 2008 Pytheas Limited

21 February 2008

37

Pytheas Business Guides

led to an increase in air and water pollution for most pollutants as compared with landfilling. This would most likely change if in-vessel composting were considered.

Copyright 2008 Pytheas Limited

21 February 2008

38

Pytheas Business Guides

8.0 Environmental Impacts Environmental impacts from biomass power generation include those due to gaseous, liquid, and solid process emissions or byproducts, noise, nuisance and aesthetic factors, and indirect or lifecycle impacts. Indirect emissions can include fugitive emissions from fuel yard and feedstock operations such as dust, volatile organic compounds and odors from handling and fuel storage, and vehicle and equipment emissions from fuel, feedstock, residue, and other material transport and processing. One of the most important environmental impacts related to biomass conversion technologies are air emissions, especially NOx. Groundlevel, or tropospheric, ozone is rarely emitted by human activity but is formed when natural and anthropogenic ozone precursors, primarily NOx and VOCs, react in the presence of sunlight, hence the name photochemical smog; tropospheric ozone is controlled by managing precursor emissions. Oxides of nitrogen are created by high temperature reactions of oxygen and nitrogen. The source of nitrogen for these reactions is primarily the combustion air but fuelborne nitrogen can be significant. Though NOx is a criteria pollutant, it is not emitted at levels high enough to raise the ambient concentration above state or federal limits. Of the ozone precursors, the control, or reduction of NOx emissions are the more difficult and expensive. For the same fuel and power output, intermittent combustion systems such as reciprocating engines have higher uncontrolled NOx emissions than continuous combustion devices such as flares, boilers, and gas turbines. The other important ozone precursors, VOCs, are usually emitted from combustion sources as products of incomplete combustion. They can be controlled with well designed and operated combustion systems for all types of fuels. Noncombustion anthropogenic VOC sources include evaporated hydrocarbon fuels, fugitive emissions from oil refineries and petrochemical plants, fermented beverage manufacturing, large animal feeding operations and feed ensiling. Carbon monoxide emissions are also controlled by following good combustion techniques. Particulate matter is currently controlled by standard methods that include cyclone separators, filtering, and electrostatic precipitators, often with multiple particle control devices used in combination. Control of sulfur and ammonia emissions are important for atmospheric PM reduction as well because of secondary reactions leading to the formation of sulfates and nitrates. Both sulfur oxides (SOx) and NOx also contribute to acid deposition in the environment (Acid Rain). Liquid emissions from gasifier systems (condensed tars and/or spent scrubber liquids) may be toxic and should not be released untreated. There are a variety of means to treat these liquids that are generally technically viable and are commercially applied in handling industrial waste liquids. Whether the techniques can be economically applied to biomass fueled gasification systems remains uncertain. Depending on the treatment, there may be toxic solids remaining which will require disposal as hazardous waste. The latter is potentially true for any biomass conversion technology handling un-separated municipal wastes. Solid fuel combustion facilities such as biomass fired stoker or fluidizedbed combustors generally have low or zero liquid emissions. Usually, liquid emissions from these systems are associated with boiler blowdown or cooling tower discharges. These types of discharges are intermittent and/or of small volume and can be disposed through municipal sewer systems or kept on site in retention ponds where the water eventually evaporates. Moreover, liquid emissions from AD systems can be substantial and may need treatment depending on the source of feedstock. 8.1 Municipal Solid Waste Anaerobic Digestion Emissions Emissions from MSW AD systems include air emissions from the use of biogas (usually stationary reciprocating engine, vehicle, gas boiler, gas turbine, or flare), air emissions from

Copyright 2008 Pytheas Limited

21 February 2008

39

Pytheas Business Guides

post treatment of solids/digestate (compost and/or land application), liquid emissions from excess process water and dewatering operations, and land emissions depending on how the solid digestates are used or disposed. In addition, there can be fugitive gas and dust emissions that depend on control strategies, operational practices, and level of maintenance at a particular facility (e.g., enclosed receiving buildings which may have exhaust air treatment to minimize VOC and dust emissions from unloading and feedstock storage). 8.2 Air Emissions from use of Biogas Emissions from the use of biogas are typical of combustion processes burning methane with the exception of sulfur. Gaseous emissions include NOx, SOx, CO, unburned hydrocarbons, and particulate matter. If the biogas is consumed in applications other than simple flaring, the emissions should receive an offset credit by the amount that would have been created from the displaced fuel. Very large landfills and waste water treatment facilities can produce enough biogas to justify relatively high capital cost prime movers such as gas turbines with heat recovery (combined cycle), or steam boilers for large scale combined heat and power. In cases where NOx emission limits are very low, micro or midsized gas turbines may be attractive. The conversion efficiency of turbines at these scales (less than 2 MW) are not as high as reciprocating engine efficiencies. Turbines generally are more expensive as well, but tend to have superior emissions control performance because of the greater control achievable with continuous combustion engines as contrasted with intermittent reciprocatingtype engines and higher excess air. The choice of prime mover type depends on the value of power, heat, emission reductions, etc. Depending on SOx emission limits and/or concentration of H2S in the biogas, gas cleaning to lower sulfur and moisture content prior to use in reciprocating engines may be necessary. Wet scrubbing is typical and iron catalyzed oxidation and biological fixed film reactor gas scrubbing methods are also used for sulfur removal. In some cases, simple compression and cooling of the gas, a common moisture removal technique, is sufficient to lower sulfur content to a suitable level. Elemental sulfur is eventually deposited and can be safely disposed. NOx is always produced to some degree when fuel is burned with air and can be emitted at high levels from uncontrolled reciprocating gas engines. Because of impurities in biogas such as sulfur, and in some cases siloxanes (principally associated with landfill gas and sewage digester gas), catalytic converters for control of NOx and unburned hydrocarbons are not often used. Flaring biogas generally emits lower NOx because temperatures with high excess air are lower than those encountered in combustion chambers, but there is no recoverable energy. Lean burn, prestratified charge, and exhaust gas recirculation technologies are used for reducing NOx emissions from this class of engine, but are currently not available for power sets below about 500 kW. Carbon monoxide emissions can become an issue especially when implementing some NOx control strategies for reciprocating engines. For example, very lean burn conditions can lead to incomplete combustion or misfiring which will increase CO and unburned hydrocarbons in the exhaust. The additional impacts of waste treatment by AD due to energy production are almost entirely due to products of combustion of the biogas and disposal or use of digester residues (digestates). 8.3 VOC Emissions from Composting of Digestate Volatile organic compounds (VOCs) can be potential air pollutants, due to their malodorous and hazardous properties and contribution to tropospheric ozone. In addition, VOCs can contribute to global warming and stratospheric ozone depletion. VOCs are usually defined as the organic compounds (except methane) with boiling points less than 79C (175F), while

Copyright 2008 Pytheas Limited

21 February 2008

40

Pytheas Business Guides

semivolatile compounds are the organic compounds with boiling points between 79C (175F)and 177C (350F). Methane also participates with NOx in the formation of ozone, but the rates are slower and the effect is more important on a global atmospheric scale than on a local scale of diurnal variation. VOCs are emitted from decomposing biogenic material including MSW and green waste compost facilities as well as posttreatment of digestate from MSW AD systems. The amount of VOC emissions are significant and need to be considered for installations in ozone nonattainment or transport air basins. 8.4 Landfill Gas Emissions The conversion of biodegradable material in AD systems will reduce eventual landfill gas production and emission and should receive emission credits that can be applied to offset emissions from the power production component. The bacterial decomposition of biogenic landfilled material produces significant quantities of landfill gas, which is composed of approximately 50% methane and 50% carbon dioxide. The methane emissions from landfills are particularly important, since methane is a more potent greenhouse gas than carbon dioxide and since landfills represent the second largest source category of anthropogenic methane emissions behind the energy industry. 8.5 Solid Residues Lignin components of biomass cannot be degraded anaerobically. With practical AD systems, size, capital, and time constraints limit the portion of biodegradable material that is actually converted to something less than theoretical. This portion, combined with the lignin components, microbial cell biomass, and the inorganic (ash) components, result in substantial solid residue that may or may not have as a whole commercial use. A potential product from AD of MSW is a compost or soil amendment resulting from aerobic stabilization of the solid residue or digestate. If a compost market does not exist or adequate quality cannot be achieved, then the solid residual may be used in thermochemical conversion or sent to landfill. Compost quality consists of at least four aspects: Content of toxic compounds such as heavy metals; Absence of pathogens; Content of undesirable goods such as plastic, metal, glass etc.; Plant nutritional value, i.e. inorganic nutrients as well as content of organic compounds for improving the structure and humus content of the soil. 8.5.1 Compost Quality Heavy Metals The quality of composted digestate depends heavily on the quality or composition of the digester feedstock; source separation of household and yard biogenic wastes from the gray or rest fraction for use in biochemical treatment for both improved digester performance and high quality or useable compost from the solid residue must be strongly encouraged.
Copyright 2008 Pytheas Limited 21 February 2008
Treatment A Emission Compound Total VOC NH3 H 2S Total Volatiles Aerobic Composting Emission (mg/ton) 590 152 nd 742 Treatment B Anaerobic Digestion Emission (mg/ton) 217 1.8 17 235.8
Modified by SWLS

Aerobic Stabilization Emission (mg/ton) 3 41 nd 44

Table 6. Emissions for different biochemical treatment methods


Source: University of California, Davis

41

Pytheas Business Guides

Heavy metals and other contaminants present in digester feedstock predominately end up in the solid digestate. Scientific studies reported that 80% of the heavy metals introduced into the digester (after pulping and separation of heavy and light fraction) were discharged to the solid digestate. The balance was not reported but presumably most of the remaining heavy metals were in solution in the process water. Metals concentration is generally higher in the digestate compared to that of the feedstock because of biomass conversion to biogas. Table 7 below displays heavy metals data of finished compost from two operating commercial facilities in Belgium. Both use the Dranco AD process. The Brecht Plant feeds sourceseparated household biogenic and yard wastes to a thermophilic high solids digester; the digestate is dewatered and composted in a covered facility with forced aeration. The Brecht Plant meets Flemish Compost Standards. The Bassum Plant, however, does not meet Flemish Compost Standards (especially zinc and copper). About 60% of the feedstock used for digestion is from separation of gray waste; heavy metals content of the fiber and sludge fractions from composted digestate of a portion of the gray waste fraction is higher than that for source separated feedstock.
Flemish Element Lead (Pb) Nickel (Ni) Zinc (Zn) Copper (Cu) Cadmium (Cd) Chromium (Cr) Units ppm TS ppm TS ppm TS ppm TS ppm TS ppm TS Compost Standards <120 <20 <300 <90 <1.5 <70 Brecht Plant Average 97 8 180 32 1 23 Bassum Plant Fiber Material Untreated Treated 80 25 350 100 0.9 60 60 18 180 55 0.5 40 Sludge Fraction Untreated Treated 220 30 900 240 1.9 40 90 18 250 80 1.4 36

Table 7: Compost analysis from the Brecht Plant (source-separated) and the Bassum Plant (gray waste digestion)

Source: (1) Organic Waste Systems nv, (2) De Baere, L. and Boelens J. Rest or mixed waste sortingdigestionseparation for the recovery of recyclables and energy.1999. Modified by SWLS

Other studies that examined compost in several regions within Germany, observed that heavy metals content in compost of source separated biogenic fraction averaged 25% that of mixed MSW compost (Table 8). Most, if not all, of these compost samples were from raw or nonAD treated feedstock.

Element Lead (Pb) Nickel (Ni) Zinc (Zn) Copper (Cu) Cadmium (Cd) Chromium (Cr) Mercury (Hg)

Mixed MSW Compost (ppm) 420 84 919 222 2.8 107 1.9

Source-separate Compost (ppm) 83 26 224 41 0.4 61 <0.2

Table 8. Heavy metal content in compost from MSW in Germany

Source: Kraus, P., and Grammel U. Die Relevanz der Schadstoffdiskussion bei der Bioabfallkompostierung. (Relevance of contaminant discussion for bio waste composting). Abfallwirtschaft 9, 1992. MIC BaezaVerlag Kassel

8.5.2 Dioxin in Trash and Compost Both biochemical and thermochemical processing strategies of MSW involve potentially hazardous emissions and exposure routes need careful inspection. Dibenzopdioxin and dibenzofuran (PCDD/F or dioxin/furans) compounds are generated by many industrial processes. Dioxin/furans are present in detectable amounts in the environment. The compounds are persistent and fatsoluble and therefore tend to bioaccumulate. Studies showed that PCDD/F in raw household mixed waste is present in the amount of 57 ng/kg toxic equivalent. Composted mixed MSW had PCDD/F levels of 38 ng/kg TEQ , followed by compost of source separated household waste with 14 ng/kg TEQ and about 10 ng/kg TEQ in green and garden waste compost.

Copyright 2008 Pytheas Limited

21 February 2008

42

Pytheas Business Guides

8.5.3 Pesticides and Herbicides in Compost and Composted Digestate Certain synthetic pesticides or herbicides are known or suspected to persist in the environment after application. Clopyralid is a persistent herbicide used to control broadleaf weeds such as dandelions, clover, and thistle. It is harmless to humans or animals but survives animal digestion and compost processes. There is potential for detectable amounts of pesticides and herbicides in MSW AD feedstocks. Studies have shown presence of chemicals and pesticides in household organic or Biowaste. Typical household and yard wastes in Germany over a fourseason period distributed between rural and urban sources were examined for more than 50 pesticide/herbicide chemicals. All samples contained CPP residues. Tropical fruit peels and residues had significant concentration of thiabendazole and methidathion, which are typically applied postharvest for protection in transport and storage. Dimethoat, dodemorph, endosulfan, and other fungicides were found in vegetable and garden wastes. Laboratory AD was carried out on raw samples and showed that some CPPs could be metabolized, modified, or otherwise stabilized to varying degrees. However, many are persistent (such as dodemorph) and can negatively affect plants and crops if the compost contains high enough levels.

Copyright 2008 Pytheas Limited

21 February 2008

43

Pytheas Business Guides

9.0 Economics When comparing AD systems, costs only make sense if the same process steps are included. Therefore, when interpreting economic studies of AD one must consider which of the following cost items are included in the analysis: Pre-development costs Site preparation, site establishment and permitting; Land acquisition; Environmental impact assessment; Engineering planning and design; Hydro-geological investigation. Construction costs Infrastructure (access roads, piping, utility connections); Cleaning and excavation; Buildings and construction; Equipment (tanks, machinery, electronics); Labor. Operating costs Maintenance fees; Labor; Materials; Water and energy; Supervision and training; Insurance; Overheads; Wastewater disposal; Solid residuals disposal; Regulatory fees. Digesters can be incorporated into existing waste treatment facilities or they may be operated as stand-alone units. This can affect ongoing expenses as well as the need for capital. For example, material handling equipment, land, and transportation equipment at a landfill could be shared with the digester. A composting operation that installs a digester would not need to build additional aeration beds. The problems that arise in defining which costs and revenues to include in a financial analysis become especially apparent when comparing AD with other waste treatment technologies. Digesters are commonly compared with landfills, but AD is a waste pre-treatment process, whereas landfilling is a waste disposal option. Only a portion of the typical MSW stream is suitable for AD, and the remaining portion must be disposed of in some other way, such as at a landfill or an incineration facility. A proper economic evaluation of an AD system would assign economic values to the landfill space reduction, energy and other revenues, and environmental protections provided by the system and then compare this against the cost of building and running the system. 9.1 Costs The costs provided in Table 9 below are predominantly based on European examples for the purposes of indicative comparisons rather than accurate reflections of actual costs. There are a wide range of costs dependent upon the complexity of the technology and the degree of mechanization and automation employed. Costs will involve differing site specific issues such as permitting, labor, emission controls and other requirements. It should also be noted that MSW treatment technologies are sensitive to the markets and outlets for recycled materials, RDF and soil conditioners that are produced by the different processes. It is likely that material outputs from a particular technology may
Copyright 2008 Pytheas Limited 21 February 2008

44

Pytheas Business Guides

have a negative value and these are not included in the above costs. The impact of the markets/outlets for these materials is also not reflected in the costs provided, nor is the cost associated with the landfill of any residues should a market fail to emerge.
Technology Pyrolysis Gasification Incineration Plasma Gasification Anaerobic Digestion In-vessel Composting Sanitary Landfill Bioreactor Landfill Plant Capacity (Tons/Day) 70-270 900 1,300 900 300 500 500 500 Capital Cost (Million US$) 16-90 15-170 30-180 50-80 20-80 50-80 5-10 10-15 O&M Cost (US$/Ton) 80-150 80-150 80-120 80-150 60-100 30-60 10-20 15-30 Planning to commissioning (Months) 12-30 12-30 54-96 12-30 12-24 9-15 9-15 12-18

Table 9. Indicative comparison of Municipal Solid Waste Treatment Technologies


Source: Soil Water & Life Solutions (2008)

Undeveloped markets/outlets (and risks associated with loss of markets) for products/outputs of any of these processes needs to be reflected in cost models. Any building should have sufficient capacity to house new separation equipment. The above costs are also not the same as the price a local authority may pay for a treatment service, which will also include other factors such as finance costs and profit margins.

Copyright 2008 Pytheas Limited

21 February 2008

45

Pytheas Business Guides

Sources (Alphabetically) Agnolucci, P., Use of economic instruments in the German renewable electricity policy. Energy Policy, 2006. 34(18): p. 3538-3548. Angelidaki, I., Ahring, B., Thermophilic anaerobic digestion of livestock waste: effect of ammonia. Appl. Microbiol, 1993. Biotechnol. 38, 560564. Angelidaki, I., Ellegaard, L., Codigestion of manure and organic wastes in centralized biogas plants , 2003. Appl. Biochem. Biotechnol. 109, 95105. Angelidaki, I., Sanders, W., Assessment of the anaerobic biodegradability of macropollutants, 2004 . Rev. Environ. Sci. Bio/Technol. 3, 117129. Archer, E., Baddeley, A., Klein, A., Schwager, J., and Whiting, K., Mechanical-Biological-Treatment: A Guide for Decion Makers Processes, Policies & Markets. 2005. Juniper Consultancy Services Ltd, March 2005, Version 1.0. Babbitt, H.E., B.J. Leland, and F.H. Whitley, The Biologic Digestion of Garbage with Sewage Sludge. University of Illinois. Engineering Experiment Station. Bulletin No. 287. 1936. Urbana: University of Illinois. Babbitt, H.E., Disposal of Garbage with Sewage Sludge. Sewage Works Journal, 1934. 6(6): p. 11031108. Barry, M., E. Colleran, and A. Wilkie, Two-stage anaerobic digestion of organic residues and energy crops, in Energy Conservation and Use of Renewable Energies in the Bio-Industries, F. Vogt, Editor. 1982. Pergamon: Oxford. Beck, R.W., Final report: Anaerobic digestion feasibility study for the Bluestem Solid Waste Agency and Iowa Department of Natural Resources, in Final Report: Anaerobic Digestion Feasibility Study for the Bluestem Solid Waste Agency and Iowa Department of Natural Resources. 2004. Bluestem Solid Waste Agency. Beggs, R.D., R. Konwinski, R.H. Zhang, and P. Shaffer. Anaerobic Phased Solids Digestion of Mixed Wastes. in Proceedings of Water Environment Federations Annual Technical Exhibition and Conference. 2007. Bidlingmaier, W., J.-M. Sidaine, and E.K. Papadimitriou, Separate collection and biological waste treatment in the European Community. Reviews in Environmental Science and Biotechnology, 2004. 3(4): p. 307-320. Biljetina, R., V.J. Srivastava, and H.R. Isaacson, Conversion of MSW (municipal solids waste) to methane in the SOLCON (solids-concentrating) digester, in Energy from biomass and wastes. 1988. Institute of Gas Technology: New Orleans, LA. Bjorklund, A., M. Dalemo, and U. Sonesson, Evaluating a municipal waste management plan using ORWARE. Journal of Cleaner Production, 1999. 7(4): p. 271-280. Blischke, J., Combining Anaerobic Digestion with enclosed tunnel composting. BioCycle, 2004. 45(4): p. 49-68. Bolzonella, D., P. Battistoni, C. Susini, and F. Cecchi, Anaerobic codigestion of waste activated sludge and OFMSW: the experiences of Viareggio and Treviso plants (Italy). Water Science and Technology, 2006. 53(8): p. 203-211. Bolzonella, D., P. Pavan, S. Mace, and F. Cecchi, Dry anaerobic digestion of differently sorted organic municipal solid waste: a full-scale experience. Water Science and Technology, 2006. 53(8): p. 23-32. Brummeler-Ten, E. and I.W. Koster, Enhancement of dry anaerobic batch digestion of the organic fraction of municipal solid-waste by an aerobic pretreatment step. Biological Wastes. 1990. 31(3): p. 199-210. Brummeler-Ten, E., H. Horbach, and I.W. Koster, Dry anaerobic batch digestion of the organic fraction of municipal solid-waste. Journal of Chemical Technology and Biotechnology. 1991. 50(2): p. 191-209. Brummeler-Ten, E., M.M.J. Aarnink, and I.W. Koster, Dry anaerobic-digestion of solid organic waste in a BIOCEL reactor at pilot-plant scale. Water Science and Technology. 1992. 25(7): p. 301-310. Brummeler-Ten, E., Full scale experience with the BIOCEL process. Water Science and Technology. 2000. 41(3): p. 299-304. Brummeler-Ten, E., I.W. Koster, and J.A. Zeevalkink, Dry anaerobic digestion of the organic fraction of municipal solid waste, in Advances in Water Pollution Control, E.R. Hall and P.N. Hobson, Editors. 1988. Pergamon Press: Oxford. p. 335-344. BTA, BTA-Technology for the utilization of biodegradable waste, in Short Information. 2007, BTA. Cecchi, F., P. Traverso, P. Pavan, D. Bolzonella, and L. Innocenti, Characteristics of the OFMSW and behaviour of the anaerobic digestion process., in Biomethanization of the organic fraction of municipal solid waste, J. Mata-Alvarez, Editor. 2002. IWA: Cornwall, U.K. p. 141-180. Chynoweth, D.P. and R. Legrand, Anaerobic digestion as an integral part of municipal waste management, in Proc. Landfill Gas and Anaerobic Digestion of Solid Waste. 1988. Chester, England. p. 467-480. Chynoweth, D.P., G. Bosch, J.F.K. Earle, R. Legrand, and K.X. Liu, A novel process for anaerobic composting of municipal solid-waste. Applied Biochemistry and Biotechnology, 1991. 28-9: p. 421-432. Chynoweth, D.P., J.M. Owens, A.A. Teixeira, P. Pullammanappallil, and S.S. Luniya, Anaerobic digestion of space mission wastes. Water Science and Technology, 2006. 53(8): p. 177-185. Chynoweth, D.P., J. Owens, D. Okeefe, J.F.K. Earle, G. Bosch, and R. Legrand, Sequential batch anaerobic composting of the organic fraction of municipal solid-waste. Water Science and Technology, 1992. 25(7): p. 327-339.

Copyright 2008 Pytheas Limited

21 February 2008

46

Pytheas Business Guides

Clarke, W.P., Cost-benefit analysis of introducing technology to rapidly degrade municipal solid waste. Waste Management & Research. 2000. 18(6): p. 510-524. Corum, L., Biogas Digesters to Recycle Green Waste. BioCycle. 2004. 45(5): p. 54-58. Costi, P., R. Minciardi, M. Robba, M. Rovatti, and R. Sacile, An environmentally sustainable decision model for urban solid waste management. Waste Management, 2004. 24(3): p. 277-295. De Baere, L., Anaerobic digestion of solid waste: State-of-the-art. Water Science and Technology, 2000. 41(3): p. 283-290. De Baere, L., and Boelens J. Rest or mixed waste sortingdigestionseparation for the recovery of recyclables and energy.1999. De Baere, L. Stateoftheart of anaerobic digestion of solid waste. 2003. Ninth International Waste Management and Landfill Symposium. Cagliari, Italy. De Baere, L. and Boelens J., The treatment of grey and mixed solid waste by means of anaerobic digestion: Future developments. In: presentations (volume II) of Il International symposium on anaerobic digestion of solid waste, Barcelona, Spain, 1999. De Baere, L., Will anaerobic digestion of solid waste survive in the future? Water Science and Technology, 2006. 53(8): p. 187-194. DeLaclos, H.F., S. Desbois, and C. Saint-Joly, Anaerobic disgestion of municipal solid organic waste: Valorga full-scale plant in Tilburg, the Netherlands. Water Science & Technology, 1997. 36(6-7): p. 457462. Diaz, L.F., G.M. Savage, G.J. Trezek, and C.G. Golueke, Biogasification of Municipal Solid-Wastes. Journal of Energy Resources Technology-Transactions of the Asme, 1981. 103(2): p. 180-185. Edelmann, W., K. Schleiss, and A. Joss, Ecological, energetic and economic comparison of anaerobic digestion with different competing technologies to treat biogenic wastes. Water Science and Technology. 2000. 41(3): p. 263-273. Edelmann, W., U. Baier, and H. Engeli, Environmental aspects of the anaerobic digestion of the organic fraction of municipal solid wastes of solid agricultural wastes. Water Science and Technology. 2005. 52(1-2): p. 203-208. EPA, Municipal Solid Waste in the United States: 2005 Facts and Figures. 2005. US Environmental Protection Agency, Office of Solid Waste: Washington, DC. Farneti, A., C. Cozzolino, D. Bolzonella, L. Innocenti, and C. Cecchi, Semi-dry anaerobic digestion of OFMSW: the new full-scale plant of Verona (Italy), in II Int. Symp. Anaerobic Dig. Solid Waste, Barcelona, June 15-17 (eds. J. Mata-Alvarez, A. Tilche and F. Cecchi). 1999. Int. Assoc. Wat. Qual. p. 330-333, Int. Assoc. Wat. Qual. Finstein, M.S., Operational Full-Scale ArrowBio Plant Integrates Separation and Anaerobic Digestion in Watery Processing, With Near-Zero Landfilling. 2003. Proceedings of WasteCon 2003, SWANAs 41st Annual International Solid Waste Exposition, October 14-16 2003 St. Louis, Missouri, p. 290-296. Forster-Carneiro, T., L.A. Fernandez, M. Perez, L.I. Romero, and C.J. Alvarez, Optimization of SEBAC start up phase of municipal solid waste anaerobic digestion. Chemical and Biochemical Engineering Quarterly. 2004. 18(4): p. 429-439. Fricke, K., H. Santen, and R. Wallmann, Comparison of selected aerobic and anaerobic procedures for MSW treatment. Waste Management. 2005. 25(8): p. 799-810. Frigon, J.C. and S.R. Guiot, Anaerobic digestion as a sustainable solution for biosolids management by the Montreal Metropolitan Community. Water Science and Technology. 2005. 52(1-2): p. 561-566. Gallert, C., A. Henning, and J. Winter, Scale-up of anaerobic digestion of the biowaste fraction from domestic wastes. Water Research. 2003. 37(6): p. 1433-1441. Gelegenis, J., Georgakakis, D., Angelidaki, I., Mavris, V., Optimization of biogas production by codigesting whey with diluted poultry manure. 2007. Renew, Energy 32, 21472160. Gelegenis, J., Georgakakis, D., Angelidaki, I., Christopoulou, N., Goumenaki, M., Optimization of biogas production from olive-oil mill wastewater, by codigesting with diluted poultry-manure. 2007. Appl. Energy 84, 646663. Ghosh, S., M.P. Henry, A. Sajjad, M.C. Mensinger, and J.L. Arora, Pilot-scale gasification of municipal solid wastes by high-rate and two-phase anaerobic digestion (TPAD). Water Science and Technology. 2000. 41(3): p. 101-110. Goldstein, N., Source separated organics as feedstock for digesters. BioCycle. 2005. 46(8): p. 42-46. Haight, M., Assessing the environmental burdens of anaerobic digestion in comparison to alternative options for managing the biodegradable fraction of municipal solid wastes. Water Science and Technology. 2005. 52(1-2): p. 553-559. Hall, S., A. Thomas, F. Hawkes, and D. Hawkes, Operation of linked percolating packed bed anaerobic digesters, in Fifth International Symposium on Anaerobic Digestion. 1988. Bologna, Italy. Hamzawi, N., K.J. Kennedy, and D.D. McLean, Anaerobic digestion of co-mingled municipal solid waste and sewage sludge. Water Science and Technology, 1998. 38(2): p. 127-132. Hansen, K. H., Angelidaki, I., Ahring, B.K., Anaerobic digestion of swine manure: inhibition by ammonia. 1998. Water Res. 38, 512. Hartmann, H., Ahring, B.K., Anaerobic digestion of the organic fraction of municipal solid waste: influence of co-digestion with manure. 2005. Water Res. 39, 15431552. Hartmann, H., Angelidaki, I., Arhing, B.K., Co-digestion of the organic fraction of municipal waste with other waste types, 2003 In: Mata-Alvarez, J. (Ed.), Biomethanization of the Organic Fraction of Municipal Solid Wastes, IWA Publishing, UK. Hartmann, H. and B.K. Ahring, Strategies for the anaerobic digestion of the organic fraction of municipal solid waste: an overview. Water Science and Technology. 2006. 53(8): p. 7-22.

Copyright 2008 Pytheas Limited

21 February 2008

47

Pytheas Business Guides

Hartman, K., An APS-Digester for the treatment of organic solid waste and power generation: UC Davis Digester Design, in Biological and Agricultural Engineering. 2004. University of California, Davis: Davis, Calif. p. 54. Hellweg, S., G. Doka, G. Finnveden, and K. Hungerbuhler, Assessing the eco-efficiency of end-of-pipe technologies with the environmental cost efficiency indicator - A case study of solid waste management. Journal of Industrial Ecology. 2005. 9(4): p. 189-203. Jenkins, B.M., Biomass in California: challenges, opportunities, and potentials for sustainable management and development, in PIER Collaborative Report. 2005. CEC: Sacramento. p. 500-01. Johansen, J.E. and R. Bakke, Enhancing hydrolysis with microaeration. Water Science and Technology. 2006. 53(8): p. 43-50. Juanga, J.P., C. Visvanathan, and J. Trankler, Optimization of anaerobic digestion of municipal solid waste in combined process and sequential staging. Waste Management & Research. 2007. 25(1): p. 3038. Kayhanian, M. and D. Rich, Pilot-scale high solids thermophilic anaerobic digestion of municipal solid waste with an emphasis on nutrient requirements. Biomass & Bioenergy. 1995. 8(6): p. 433-444. Kayhanian, M. and G. Tchobanoglous, Innovative 2-stage process for the recovery of energy and compost from the organic fraction of municipal solid-waste (MSW). Water Science and Technology. 1993. 27(2): p. 133-143. Kayhanian, M., Biodegradability of the organic fraction of municipal solid-waste in a high-solids anaerobic digester. Waste Management & Research. 1995. 13(2): p. 123-136. Kelleher, M., Anaerobic digestion outlook for MSW streams. BioCycle. 2007. 48(8): p. 51. Kelleher, M., Feasibility of Generating Green Power Through Anaerobic Digestion of Garden Refuse from Sacramento Area, in Report to SMUD Advanced Renewable and Distributed Generation Program. 2005. RIS International Ltd.: Toronto. Komilis, D. P., Ham R. K., and Park J. K. Emission of volatile organic compounds during composting of municipal solid wastes. 2004. Water Research 38(7): 17071714. Kraus, P., and Grammel U. Die Relevanz der Schadstoffdiskussion bei der Bioabfallkompostierung. (Relevance of contaminant discussion for bio waste composting). Abfallwirtschaft 9, 1992. MIC Baeza Verlag Kassel. Kubler, H., Hoppenheidt K., Hirsch P., Kottmair A., Nimmrichter R., Nordsieck H., Mucke W., and Swerev M. 2000. Full scale codigestion of organic waste. Water Science & Technology 41(3): 195202. Legrand, R. and D.P. Chynoweth, Anaerobic digestion as a solid waste management tool, in Proc. Twelfth Annual Madison Waste Conf. 1989. Madison, WI. p. 165-185. Lissens, G., P. Vandevivere, L. De Baere, E.M. Biey, and W. Verstraete, Solid waste digestors: process performance and practice for municipal solid waste digestion. Water Science and Technology. 2001. 44(8): p. 91-102. Liu, T.C. and S. Ghosh, Phase separation during anaerobic fermentation of solid substrates in an innovative plug-flow reactor. Water Science and Technology. 1997. 36(6-7): p. 303-310. Liu, H.W., H.K. Walter, G.M. Vogt, H.S. Vogt, and B.E. Holbein, Steam pressure disruption of municipal solid waste enhances anaerobic digestion kinetics and biogas yield. Biotechnology and Bioengineering. 2002. 77(2): p. 121-130. Loikala J., Jatteenkasittely Wabioprosessissa; vaihtoehto perinteiselle kaatopaikkakasittelylle, The treatment of waste material in the Wabio process; an alternative for a traditional treatment in a dumping ground, Kemia-Kemi 15. 1988. 3, pp. 280-282. Ludwig, C., S. Hellweg, and S. Stucki, eds. Municipal Solid Waste Management-Strategies and Technologies for Sustainable Solutions. 2003. Springer-Verlag: Berlin. 534. Lusk, P.D., COMPARATIVE ECONOMIC-ANALYSIS - ANAEROBIC DIGESTER CASE-STUDY. Bioresource Technology. 1991. 36(3): p. 223-228. Lyum, M., Life and Times of an Organics Recycling Company. BioCycle. 1999. 40(11): p. 34. Nichols, C.E., Overview of anaerobic digestion technologies in Europe. BioCycle. 2004. 45(1): p. 47-53 Okeefe, D.M., D.P. Chynoweth, A.W. Barkdoll, R.A. Nordstedt, J.M. Owens, and J. Sifontes, Sequential batch anaerobic composting of municipal solid-waste (MSW) and yard waste. Water Science and Technology. 1993. 27(2): p. 77-86. Oleszkiewicz, J.A. and H.M. Poggi-Varaldo, High-solids anaerobic digestion of mixed municipal and industrial waste. Journal of Environmental Engineering-ASCE. 1998. 124(10): p. 1032-1032. Pfeffer, J.T., Methane from Urban Solid-Wastes - Refcom Project. Process Biochemistry. 1978. 13(6): p. 8-11. Pierce, J. L. Landfill gas electric power generation: A costeffective and environmentally sound renewable energy source. 1997. PowerGen 97 Asia. Pirt, S.J. and Y.K. Lee, Enhancement of methanogenesis by traces of oxygen in bacterial digestion of biomass. FEMS Microbiology Letters. 1983. 18(1-2): p. 61-63. Repa, E.W., NSWMAs 2005 Tip Fee Survey, in NSWMA Research Bulletin. 2005. National Solid Wastes Management Association. p. 05-3. Rijkens, B.A., A novel two-step process for the anaerobic digestion of solid waste, in Energy from Biomass and Wastes V. 1981. Institute of Gas Technology. p. 463-475. Rijkens, B.A., Two-phase process for the anaerobic digestion of solid wastes - First results of a pilotscale experiment, in Proceedings: Energy from Biomass, 2nd E.C. Conference. 1982. Berlin. Rivard, C.J., B.W. Duff, J.H. Dickow, C.C. Wiles, N.J. Nagle, J.L. Gaddy, and E.C. Clausen, Demonstration-scale evaluation of a novel high-solids anaerobic digestion process for converting organic wastes to fuel gas and compost. Applied Biochemistry and Biotechnology. 1998. 70-2: p. 687-695.

Copyright 2008 Pytheas Limited

21 February 2008

48

Pytheas Business Guides

Rivard, C.J., Intermediate-scale high-solids anaerobic digestion system operational development, in Other Information: PBD: Feb 1995. 1995. Department of Energy: United States. p. Size: 34 p. Rivard, C., Recycling and Energy Recovery Pilot Project: Project Report and Future Efforts , in Other Information: PBD: 19 May 1999. 1999. National Renewable Energy Laboratory: Golden, CO. p. Size: vp. Saint-Joly, C., S. Desbois, and J.P. Lotti, Determinant impact of waste collection and composition on anaerobic digestion performance: industrial results. Water Science and Technology. 2000. 41(3): p. 291297. Schfer, G., Europe in figures--Eurostat yearbook 2006-07, M. Feith, et al., Editors. 2007. Eurostat: Luxembourg. Schu, K.; Schu, R., Sand im Getriebe der Vergrung? In: Khle-Weidemeier (ed.): Internationale Tagung MBA. 2007. Cuvillier Verlag, Gttingen, p. 494-506 Seo, S., T. Aramaki, Y.W. Hwang, and K. Hanaki, Environmental impact of solid waste treatment methods in Korea. Journal of Environmental Engineering-Asce. 2004. 130(1): p. 81-89. Shanmugam, P., Horan, N.J., Simple and rapid methods to evaluate methane potential and biomass yield for a range of mixed solid wastes. 2009. Bioresour Technol. 100, 471474. Smet, E., Van Langenhove H., and De Bo I. The emission of volatile compounds during the aerobic and the combined anaerobic/aerobic composting of biowaste. 1999. Atmospheric Environment 33(8): 12951303. Srivastava, V.J., Biljetina R., Isaacson H.R., and Hayes T.D. Biogasification of community-derived biomass and solid wastes in a pilot-scale SOLCON reactor, in 10th Symposium on Biotechnology for Fuels and Chemicals. 1988. Institute of Gas Technology: Gatlinburg, TN. p. Pages: 23. Srivastava, V., Two-phase anaerobic digestion of carbonaceous organic materials., USPTO, Patent No. 5,500,123. 1996. USA. Svensson, L.M., K. Christensson, and L. Bjornsson, Biogas production from crop residues on a farmscale level: is it economically feasible under conditions in Sweden? Bioprocess and Biosystems Engineering. 2005. 28(3): p. 139-148. Taube, J., K. Vorkamp, Forster M., and Herrmann R. Pesticide residues in biological waste. 2002. Chemosphere 49(10): 13571365. Teixeira, A.A., D.P. Chynoweth, P.J. Haley, and J. Sifontes, Commercialization of SEBAC Solid Waste Management Technology, in 03ICES-037. 2003. SAE International. Teixeira, A.A., D.P. Chynoweth, J.M. Owens, E. Rich, A.L. Dedrick, and P.J. Haley, Prototype Space Mission SEBAC Biological Solid Waste Management System, in 34th Internat. Conf. on Environ. Systems (ICES). Paper 2004-ICES-098. 2004. SAE International. Themelis, N.J. and P.A. Ulloa, Methane generation in landfills. Renewable Energy. 2007. 32(7): p. 1243-1257. Tsilemou, K. and D. Panagiotakopoulos, Approximate cost functions for solid waste treatment facilities. Waste Management & Research. 2006. 24(4): p. 310-322. Van den Broek, B. and J. Cook, A Plan to Achieve 50% Waste Diversion in Sydney. 1994. Waste Service NSW, Sydney. Vandevivere, P., L. De Baere, and W. Verstraete, Types of anaerobic digesters for solid wastes, in Biomethanization of the Organic Fraction of Municipal Solid Wastes, J. Mata-Alvarez, Editor. 2002. IWA Publishing: Barcelona. p. 111-140. Verma, S., Anaerobic digestion of biodegradable organics in municipal solid wastes. 2002. Columbia University. Vogt, G.M., H.W. Liu, K.J. Kennedy, H.S. Vogt, and B.E. Holbein, Super blue box recycling (SUBBOR) enhanced two-stage anaerobic digestion process for recycling municipal solid waste: laboratory pilot studies. Bioresource Technology. 2002. 85(3): p. 291-299. Vorkamp, K., E. Kellner, J. Taube, K. D. Moller, and R. Herrmann. Fate of methidathion residues in biological waste during anaerobic digestion. 2002. Chemosphere 48(3): 287297. Vorkamp, K., J. Taube, J. Dilling, E. Kellner, and R. Herrmann. Fate of the fungicide dodemorph during anaerobic digestion of biological waste. 2003. Chemosphere 53(5): 505514. Walter, D.K. and C. Brooks, Refuse Conversion to Methane (Ref COM): A Proof-of-Concept Anaerobic Digestion Facility. IEEE Transactions on Power Apparatus and Systems. 1980. PAS-99(6): p. 2363-2368. Whyte, R. and G. Perry, A rough guide to anaerobic digestion costs and MSW diversion. Biocycle, 2001. 42(10): p. 30-33. Wichert, B., L. Wittrup, and R. Robel, Biogas, compost and fuel cells. BioCycle, August. 1994. 35(8): p. 34-36. Williams, R.B., B.M. Jenkins, and D. Nguyen, Solid Waste Conversion: A review and database of current and emerging technologies. 2003. California Integrated Waste Management Board. Williams, R.B. and B.M. Jenkins, Management and conversion of organic waste and biomass in California, in 2nd World Conference and Technology Exhibition on Biomass for Energy, Industry and Climate Protection., Rome, Italy. 2004. Williams, R.B., UC Davis Technology Assessment for Advanced Biomass Power Generation, in PIER Consultation Report. 2005. California Energy Commission: Sacramento, California. Zhang, R.H. and Z.Q. Zhang, Biogasification of rice straw with an anaerobic-phased solids digester system. Bioresource Technology. 1999. 68(3): p. 235-245. Zhang, R. and Z. Zhang, Biogasification of solid waste with an anaerobic-phased solids-digester system, USPTO, Patent No. 6,342,378. 2002. US. Zitomer, D.H. and J.D. Shrout, Feasibility and benefits of methanogenesis under oxygen-limited conditions. Waste Management. 1998. 18(2): p. 107-116.

Copyright 2008 Pytheas Limited

21 February 2008

49

Pytheas Business Guides

Disclaimer The above notes have been compiled to assist you; however, actions taken as a result of this document are at the discretion of the reader and not PYTHEAS or SWLS. All rights reserved. The material in this publication may not be copied, stored or transmitted without the prior permission of the publishers. Short extracts may be quoted, provided the source is fully acknowledged.

Copyright 2008 Pytheas Limited

21 February 2008

50

Вам также может понравиться