Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Bioenergy Research: Advances and Applications
Bioenergy Research: Advances and Applications
Bioenergy Research: Advances and Applications
Ebook1,677 pages21 hours

Bioenergy Research: Advances and Applications

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Bioenergy Research: Advances and Applications brings biology and engineering together to address the challenges of future energy needs. The book consolidates the most recent research on current technologies, concepts, and commercial developments in various types of widely used biofuels and integrated biorefineries, across the disciplines of biochemistry, biotechnology, phytology, and microbiology.

All the chapters in the book are derived from international scientific experts in their respective research areas. They provide you with clear and concise information on both standard and more recent bioenergy production methods, including hydrolysis and microbial fermentation. Chapters are also designed to facilitate early stage researchers, and enables you to easily grasp the concepts, methodologies and application of bioenergy technologies. Each chapter in the book describes the merits and drawbacks of each technology as well as its usefulness.

The book provides information on recent approaches to graduates, post-graduates, researchers and practitioners studying and working in field of the bioenergy. It is an invaluable information resource on biomass-based biofuels for fundamental and applied research, catering to researchers in the areas of bio-hydrogen, bioethanol, bio-methane and biorefineries, and the use of microbial  processes in the conversion of biomass into biofuels.

  • Reviews all existing and promising technologies for production of advanced biofuels in addition to bioenergy policies and research funding
  • Cutting-edge research concepts for biofuels production using biological and biochemical routes, including microbial fuel cells
  • Includes production methods and conversion processes for all types of biofuels, including bioethanol and biohydrogen, and outlines the pros and cons of each
LanguageEnglish
Release dateDec 5, 2013
ISBN9780444595645
Bioenergy Research: Advances and Applications

Related to Bioenergy Research

Related ebooks

Power Resources For You

View More

Related articles

Reviews for Bioenergy Research

Rating: 5 out of 5 stars
5/5

2 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Bioenergy Research - Vijai G. Gupta

    1

    Current Bioenergy Researches

    Strengths and Future Challenges

    Naveen Kumar Mekala¹, Ravichandra Potumarthi²,∗, Rama Raju Baadhe¹ and Vijai K. Gupta³,    ¹Department of Biotechnology, National Institute of Technology, Warangal, Andhra Pradesh, India,    ²Department of Chemical Engineering, Monash University, Clayton, Victoria, Australia,    ³Molecular Glycobiotechnology Group, Department of Biochemistry, School of Natural Sciences, National University of Ireland Galway, Galway, Ireland,    ∗Corresponding author email: ravichandra.potumarthi@monash.edu; pravichandra@gmail.com

    Abstract

    Bioenergy is major stake holder in meeting global future energy needs. This contribution can be extended significantly in the near future, by reducing the greenhouse gas emission and saving environment, as well as improving trade balances, contributing to energy security, providing opportunities for socioeconomical development in rural areas. Bioenergy could sustainably offer a quarter and a third of global primary energy supply by 2050. Bioenergy is the only renewable source that can replace fossil fuels in all energy markets in the production of heat, electricity, and fuels for transport. Many bioenergy principles can be used to convert biomass feedstock into final bioenergy products. The technologies for producing heat and power from feedstock are already well defined and fully commercialized. A wide variety of conversion technologies are under construction, with improved competence, lower costs and improved environmental protection. However, the possible competition between raw materials for bioenergy with other biomass applications must be carefully answered. The output of biomass feedstock and food grains needs to be increased by good agricultural practices. Logistics and infrastructure issues should be spoken off, and there is need for further scientific innovations leading to more competent and cleaner conversion of more assorted feedstock.

    Keywords

    Anaerobic digestion; Biodiesel; Bioethanol; Biogas; Biofuel; Digesters; Fermentation; Transesterification

    Outline

    Introduction

    Different Forms of Bioenergy

    Biopellets

    Bioethanol

    Feedstock for Bioethanol

    Pretreatment of Lignocelluloses

    Biological Pretreatment

    Physical Pretreatment

    Chemical Pretreatment

    Bioethanol Fermentation

    Molecular Biology Trends in Bioethanol Production Development

    Bioreactors in Ethanol Production

    Immobilization of Cells for Ethanol Production

    Biodiesel

    Feedstocks for Biodiesel

    Biodiesel from Pure Vegetable Oil

    Biodiesel from Animal Fat Wastes

    Other Waste Cooking Oils

    Algae as a Biodiesel Source

    Bioreactors for Biodiesel Production

    Biogas

    Biogas Feedstock

    Household Digesters for Biogas

    Fixed Dome Digesters

    Floating Drum Digesters

    Social and Environmental Aspects of Biogas Digesters

    Conclusion

    References

    Introduction

    Modern world is facing two vital challenges, energy crisis and environmental pollution. Energy is a key component for all sectors of modern economy and plays an elementary role in improving the quality of life (US DOE, 2010). In current situations, approximately 80% of world energy supplies rely on rapidly exhausting nonrenewable fossil fuels. At the current rate of consumption, crude oil reserves, natural gas and liquid fuels were expected to last for around 60 and 120 years, respectively (British Petroleum Statistical Review, 2011). An additional challenge with fossil fuel consumption is emission of greenhouse gases (GHGs). In 2010, an average of 450 g of CO2 was emitted by production of 1 kWh of electricity from the coal (Figure 1.1) (International Energy Agency Statistics, 2012). It is also clear that coal's share of the global energy continues to rise, and by 2017 coal will come close to surpassing oil as the world's top energy source. China and India lead the growth in coal consumption over the next 5 years. Research says China will surpass the rest of the world in coal demand during the outlook period, while India will become the largest seaborne coal importer and second largest consumer, surpassing the United States (IEA, 2012).

    FIGURE 1.1 Global energy production chart signifies the growing demand for energy. Source: IEA, 2012. (For color version of this figure, the reader is referred to the online version of this book.)

    Growing global energy needs, release of environmental pollutants from fossil fuels and national security have finally tuned the attention in clean liquid fuel as a suitable alternative source of energy. The alternative bioenergy sources, not only cut the dependence on oil trade and reduce the doubts caused by the fluctuations in oil price, but also secure reductions in environmental pollution due to their high oxygen content (Huang et al., 2008; Boer et al., 2000).

    In this context, the availability of bioenergy in its two main appearances, wood and agro energy can offer cleaner energy services to meet basic energy requirements. This century could see a remarkable switchover from fossil fuel-based energy to bioenergy-based economy, with agriculture and forestry as the main sources of feedstock for biofuels such as wood pellets, fuel-wood, charcoal, bioethanol, and biodiesel (Agarwal, 2007). Moreover, energy crops can be part of highly specialized and various agricultural production chains and biorefineries, where a variety of bioproducts could be obtained besides bioenergy, which are important for their economic competitiveness (United Nations Environment Program, 2006).

    The exploitation of currently unused by-products and growing energy crops can address other existing environmental concerns. Perennial energy crops and plantations are generally characterized by higher biodiversity compared with conventional annual crops. By providing more continuous soil cover, they reduce the impact of rainfall and sediment transport, thereby preventing soil erosion. The introduction of annual energy crops into crop systems allows for diversification and expansion of crop rotations, replacing less favorable monocropping systems (Kheshgi et al., 1996). Deforested, degraded and marginal land can be rehabilitated with bioenergy plantations, thus helping to combat desertification and hopefully reducing market and geosocial pressures on high-quality arable land.

    Biofuels can be obtained in bulk when they are derived from agricultural crops, crop residues and processing wastes from agroindustries, forests, etc. Despite this immense potential, existing biofuel policies have been very costly; they produce slight reductions in fossil fuel use and increase, rather than decrease, in GHG emissions (Wuebbles and Jain, 2001). However, recent volatility and rise in international fossil fuel prices, make biomass increasingly competitive as energy feedstock.

    Current bioenergy research around the globe should direct us toward reduced production cost, higher energy conversion efficiency and greater cost-effectiveness of biofuels. After all we are aware of a fact use of biomass as a potentially large source of energy in the 21st century will have a significant impact in rural, agricultural and forestry development (UNEP, 2006).

    Different Forms of Bioenergy

    Organic matter holding bioenergy sources in side is known as biomass. We can utilize this biomass in many different ways, through something as simple as burning wood for heat, or as complex as growing genetically modified microbes to produce cellulosic ethanol (Adler et al., 2009). Since nearly entire bioenergy can be traced back to energy from sunlight, bioenergy has the key advantage of being a renewable energy source. Here, in this chapter we will discuss various forms of bioenergy and their application in detail.

    Biopellets

    Today, wood pellets are an imperative and well-accepted fuel in lots of different countries and the according markets are likely to rise even further in future. For these reasons, it is feared that the inadequate availability of cheap wood as a feedstock for pellets will limit this market increase (Marina et al., 2011; Larsson et al., 2008). As alternative, autumn leaves from urban areas, as a seasonal available waste material, are the possible substitutes for or additives to wood. In lot of Western countries, wood pellets become a more and more significant fuel for the use in small furnaces for household buildings or in industries as a climate-neutral alternative to crude oil or natural gas (Verma et al., 2012; Nielsen et al., 2009). This pelletized biomass has a number of advantages like tolerance against microbial degradation, high transport and storage density of bioenergy, and the process of pelletization is quite simpler (Figure 1.2).

    FIGURE 1.2 (a) Experimental flow sheet for pelletization of leaves; (b) leaf pellets. (For color version of this figure, the reader is referred to the online version of this book.)

    Bioethanol

    Bioethanol is the most common biofuel worldwide. It is produced by simple fermentation of sugars derived from wheat, corn, sugar beets, sugarcane, molasses and any sugar or starch sources that alcoholic beverages can be made from (Cara et al., 2008). Bioethanol can be used in petrol engines as a substitute for gasoline. Bioconversion of lignocellulosics into fermentable sugars is a biorefining area in which enormous research labors have been invested, as it is a prerequisite for the subsequent bioethanol production (Broder et al., 1992). Although extensive research has been carried out to meet the potential challenges of bioenergy generation, there is no self-sufficient process or technology available today to convert the lignocellulosic biomass to bioethanol (Tu et al., 2007).

    Use of bioethanol-blended fossil fuel for automobiles can significantly cut the petroleum use and exhaust GHG emission. Bioethanol can be produced from different kinds of raw materials and these raw materials are classified into three categories of agricultural raw materials: simple sugars, starch and lignocelluloses (Mustafa and Havva, 2009). Bioethanol from sugarcane, under proper conditions, is essentially a clean fuel and has several advantages over petroleum-derived gasoline in reducing GHG emissions and improving air quality in metropolitan cities. Conversion technologies for producing bioethanol from cellulosic biomass resources such as forest materials, agricultural residues and urban wastes are under development and have not yet been established commercially (Demirbas, 2008).

    Feedstock for Bioethanol

    Across the globe, there is a rising need to find out new and cheap carbohydrate sources for bioethanol production (Mohanty et al., 2009). Presently, a serious focus is on biofuels made from renewable energy crops such as sugarcane, corn, wheat, soybeans, etc. In a given production line, the comparison of the biomass includes several issues: (1) cultivation practices, (2) chemical composition of the biomass, (3) use of resources, (4) emission of GHGs, (5) availability of land and land use practices, (6) soil erosion, (7) energy balance, (8) price of the biomass, (9) contribution to biodiversity and landscape value losses, (10) direct economic value of the feedstock, (11) water requirements and water availability, (12) creation or maintain of employment, and (13) logistic cost (transport and storage of the biomass) (Gnansounou et al., 2005).

    Bioethanol feedstock can be divided into three major groups: (1) sugar-based feedstock (e.g. sugarcane, beet sugar, sorghum and fruits), (2) starchy feedstock (e.g. corn, sweet potato, rice, potatoes, cassava, wheat and barley), and (3) lignocellulosic feedstock (e.g. wood, straw, grasses, and corncob). In short term, production of bioethanol as a fuel is almost entirely dependent on starch and sugars from existing food crops (Smith, 2008; Potumarthi et al., 2012). The negative part in producing bioethanol from starch and sugar is that the feedstock tends to be costly and demanded by other applications as well (Enguidanos et al., 2002). Lignocellulosic biomass is envisaged to provide a major portion of the raw materials for bioethanol production in the long term due to its low cost and high availability (Gnansounou et al., 2005).

    Up to 2003, about 60% of global bioethanol was obtained from sugarcane and 40% from all other crops (Dufey, 2006). Brazil utilizes sugarcane for bioethanol production, while the United States and other western countries mainly use starch from corn, wheat and barley (Linde et al., 2008). Brazil is the largest producer of sugarcane with about 672,157,000 tons of global production followed by India, second largest producer with 285,029,000 ton production (Food & Agricultural Organization of United Nations, 2013).

    Bioethanol production in Brazil is less expensive than in the United States from corn or in Europe from sugar beet, because of lower labor costs, shorter processing time, lower transport costs, and other input costs. After sugarcane, starch is the high-yield feedstock for bioethanol production, but pretreatment is necessary to produce bioethanol by fermentation (Pongsawatmanit et al., 2007). Starch is a homopolymer consisting monomers of D-glucose and for bioethanol production it is necessary to break down this carbohydrate for obtaining glucose syrup, which can be further transformed into bioethanol by yeasts. Starch-based feedstock are the most utilized for bioethanol production in America and Europe.

    Biomass from agricultural waste (wheat straw, sugarcane bagasse, etc.), wood, and energy crops are attractive materials for bioethanol production since it is the most abundant reproducible assets on earth (Figure 1.3). The existing biomass from crops could produce up to 442 American billion liters per year of bioethanol (Bohlmann, 2006). Thus, the total possible bioethanol production from crop residues and wasted crops is 491 American billion liters per year, about 16 times higher than the existing world bioethanol production. Advantages of biofuels are as follows: (1) biofuels are easily available from common biomass sources, (2) biofuels have a considerable environmentally friendly potential, and (3) they are biodegradable and contribute to sustainability (Balat, 2007; Mekala et al., 2008). Although lignocellulosic biomass is the best alternative source, initial pretreatment is a must to attain simple sugars for simultaneous ethanol fermentation.

    FIGURE 1.3 Major lignocellulosic feedstock explored for bioethanol production. Source: Taherzadeh and Karimi, 2007.

    Pretreatment of Lignocelluloses

    Woody materials including bark, wood, and mixture of other residues from the forest contain cellulose, hemicelluloses, lignin and small amount of other biomass contents (Figure 1.4). Cellulose is the major component in plant biomass and it is made of anhydroglucopyranose or glucose residues, which can be converted to glucose and act as major source of hexoses in woody feedstock (Alvira et al., 2010). Due to the hydrogen bonds in it, cellulose is a highly crystalline structure and it is difficult to hydrolyze. Unlike cellulose, hemicelluloses are heteropolymers composed of both five-carbon sugars and six-carbon sugars, including glucose, mannose, arabinose, xylose and others (Bochek, 2003). Due to its amorphous structure, hemicellulose is easily breakable by dilute acid or base. Lignin is the third major part in wood and comprises the glue that guards woody biomass from pathogens. Lignin mainly consists of phenolic units and with available technology we cannot use lignin as a source of bioethanol. Pretreatment of these lignocelluloses separates the sugars and lignin and disrupts the crystalline portion of the biopolymers (Hu et al., 2008). Different pretreatment methods have been explored for achieving the optimistic situations with different biomass.

    FIGURE 1.4 Chemical composition of lignocellulosic biomass (SW, soft wood; HW, hard wood).

    In general, pretreatment methods can be divided into biological pretreatment, physical pretreatment, and chemical pretreatment according to the pretreatment procedure. Some pretreatments combine two or more of broadly explored methodologies. Table 1.1 recaps some of the broadly explored pretreatment methods as per the classification (Sun and Cheng, 2002).

    TABLE 1.1

    Pretreatment Methods of Lignocellulosic Feedstock

    Source: Moiser et al., 2005.

    Biological Pretreatment

    Most pretreatment technologies require selected and expensive amenities or equipment that have high power requirements, depending on the process. In particular, physical and chemical processes require rich energy for biomass conversion, whereas, biological treatment via microorganisms is a safe and environmentally friendly method and is increasingly being promoted as a process that does not require high energy, even for lignin removal from a lignocellulosic biomass (Okano et al., 2005; Potumarthi et al., 2013; Ravichandra et al., 2013).

    Phanerochaete chrysosporium is one among the best microbial models to study the lignin degradation by white rot fungi. Fungi breaks down lignin anaerobically through a family of extracellular enzymes collectively termed as lignases (Howard et al., 2003). Two families of lignolytic enzymes are generally considered to play vital role in the enzymatic degradation: peroxidases (lignin peroxidase) and phenol oxidase (Malherbe and Cloete, 2003). Other enzymes are not fully explored including glucose oxidase, methanol oxidase, glyoxal oxidase, and oxidoreductase (Eriksson, 2000). Another best example was Trichoderma reesei, which is a mesophilic cellulolytic fungus isolated in the mid-1950s. By the mid-1970s, an impressive collection of more than 14,000 cellulolytic fungi were isolated against cellulose and other insoluble fibers (Coyne et al., 2013). Trichoderma reesei, although a good producer of hemi and cellulolytic enzymes, is unable to degrade lignin (Mekala et al., 2008; Gupta et al., 2013).

    Actinomycetes are also best tested for their task in lignin biodegradation. Lignolytic enzymes like peroxidases, ligninase and manganese peroxidase were discovered in P. chrysosporium (Saritha et al., 2012). Based on this, P. chrysosporium was taken up for biological delignification of wood and paddy straw in ethanol production. But, the extent of delignification was inadequate to expose a significant portion of cellulose for enzymatic hydrolysis. Thus, from the reports available, it is evident that white rot fungi and actinomycete can be used jointly to remove lignin from lignocellulosic substrates, and further studies are required to shorten the incubation time and to optimize the delignification process.

    Physical Pretreatment

    Mechanical Comminution

    The objective is to cut the particle size and crystallinity of lignocellulosic biomass in order to increase the surface area and reduce the degree of polymerization. Methods like chipping, grinding and milling can be used to improve the further enzymatic hydrolysis. However, this process is not economically feasible due to the high energy requirement (Tassinari et al., 1980). During comminution, vibratory ball milling is found to be more efficient in breaking down the cellulose molecules of spruce and aspen chips and improving the digestibility of the biomass than ordinary ball milling (Sun and Cheng, 2002). The power requirement of mechanical comminution of agricultural materials depends on the final particle size and the waste biomass characteristics.

    Steam Explosion

    It is a hydrothermal pretreatment in which the lignocellulose is subjected to pressurized water vapors for few seconds to several minutes, and then suddenly depressurized. In this process, combination with the partial hydrolysis of hemicelluloses and solubilization, the lignin is redistributed and removed up to certain level from the material (Pan et al., 2005). Although this technique is cost-effective, it generates toxic by-products and the hemicelluloses degradation is partial (Saritha et al., 2012).

    Ultrasonic Pretreatment

    This technique is extensively used for the treatment of sludge from wastewater treatment plants. An experiment on carboxyl methyl cellulose with irradiation increased the rate of enzymatic hydrolysis up to 200% approximately (Imai et al., 2004). The mechanism of action in ultrasonic treatment remains unknown. One guess is that, the hydrogen bonds in the crystalline cellulose structures were broken due to irradiation energy, whose energy is higher than the hydrogen bond energy (Bochek, 2003).

    Extrusion

    This process disrupts the crystal structure of lignocellulose and increases the accessibility of carbohydrates to enzyme. In this case, materials are subjected to heating, mixing and shearing resulting in physical and chemical modifications in biomass structure (Karunanithy et al., 2008). However, the process is novel and not widely applied.

    Chemical Pretreatment

    Acid Hydrolysis

    During acid hydrolysis, concentrated acids like HCl and H2SO4 have been used to pretreat lignocellulosic biomass. Although acids are influential agents for cellulose hydrolysis, intense acids are poisonous, corrosive, and require chemical reactors that are resistant to corrosion. In addition, concentrated acid must be removed after hydrolysis of celluloses into simple sugars, which simultaneously enter into alcoholic fermentation (Potumarthi et al., 2013; Ravichandra et al., 2013). Hydrolysis using dilute acid has been industriously developed for pretreatment of lignocellulosic biomass (O'Donovan et al., 2013). The dilute sulfuric acid pretreatment can attain high reaction rates and drastically improve cellulose hydrolysis. Dilute acids at lower temperatures, saccharification suffered from low yields because of sugar decomposition (Chen et al., 2009). High temperatures, with dilute acids are favorable for cellulose hydrolysis. In recent times, dilute acid hydrolysis processes use less harsh environment and achieve high xylan to xylose conversion rates. Achieving high xylan to xylose conversion yields is required to achieve favorable economics, because xylan is the third most promising carbohydrate in many lignocellulosic feedstocks (Sun and Cheng, 2002). Primarily two types of dilute acid pretreatment processes are well studied: high-temperature (T > 160 °C), continuous flow process for low solids loading (5–15% (weight of biomass/weight of reaction mixture)) (Converse et al., 1989), and low-temperature (T < 160 °C), batch process for high solids loading (10–40%) (Esteghlalian et al., 1997). Although dilute acid hydrolysis can significantly improve the cellulose breakdown, overall cost is typically higher when compared with few other physicochemical pretreatment processes such as steam explosion.

    Alkaline Hydrolysis

    Usually alkaline hydrolysis was carried out at low temperature and pressure and it may be completed even at ambient conditions. The only drawback of this process is time; it might be hours or even days to complete the hydrolysis. During lime pretreatment, some calcium is tainted into nonrecoverable salts or included in the biomass (Chang et al., 2001). Other alkaline pretreatment methods include calcium, potassium, sodium and ammonium hydroxides as reactants based on biomass category. Among these reactants, sodium hydroxide receives the most attention followed by lime, due to its advantage of being low cost and secure to use, as well as it is easily recoverable from water as insoluble CaCO3 by reaction with CO2. Further delignification of feedstocks can be enhanced by supplying surplus air/oxygen (Hu et al., 2008). We can compare alkaline pretreatment of feedstocks to Kraft pulping, where lignin was removed efficiently, thus improving the reactivity of polysaccharides. Alkaline hydrolysis also effectively removes acetyl groups and uronic substitutions from hemicellulose; thus, the surface of hemicellulose becomes more accessible to the hydrolytic enzymes.

    Ammonia Hydrolysis

    Ammonia has abundant desirable characteristics as a pretreatment reagent. It is a valuable swelling reagent for lignocellulosic biomass, and ammonia has high selectivity for reactions with lignin over those with carbohydrates (Kim et al., 2003). It is one of the most extensively used commodity chemicals with about one-fourth the cost of sulfuric acid on molar basis. Its high volatility makes it easy to recover and recycle. It is a nonpolluting and noncorrosive chemical. One of the known reactions of aqueous ammonia with lignin is cleavage of ether (C–O–C) bonds in lignin as well as ester bonds in the lignin–carbohydrate complex (Lewin and Roldan, 1971). This above reaction indicates that ammonia pretreatment selectively cuts the lignin content in biomass. Lignin is believed to be a major hindrance in enzymatic hydrolysis and there are several advantages by removing lignin early in the conversion process before it faces the biological treatment.

    Ozonolysis

    Ozone is a leading oxidant that demonstrates high delignification efficiency. This ozonolysis is done at room temperature and at normal pressure. In this case we do not locate any inhibitory by-products, which affect the simultaneous fermentation steps (Saritha et al., 2012). An important drawback is ozone requirement in large quantities, which can make the process economically unapproachable (Sun and Cheng, 2002).

    Bioethanol Fermentation

    Once the lignocelluloses were hydrolyzed into simple sugars, they have to be fermented to ethanol. The hydrolyzate now contains various hexoses and pentoses, mainly glucose and xylose, depending on the substrate and the pretreatment method applied. Currently, fermentation of simple sugars is mostly done using yeast cultures (Saccharomyces cerevisiae), because of its well-known characteristics, toughness and high ethanol yield. However, S. cerevisiae can only ferment hexoses and not the pentoses. The pentose sugars can be fermented in an additional step by another microorganism or by S. cerevisiae itself through genetic engineering approaches, so that it is able to ferment pentoses as well (Van Zyl et al., 2007). List of most popular yeast strains used for ethanol fermentation are mentioned in Table 1.2. Besides a high yield, an important aspect with fermentation is alcohol tolerance in the fermenting organisms. A strategy to defeat this crisis is to have a system where the ethanol is recovered at regular intervals to keep the alcohol concentrations under control. Another problem is inhibitory compounds that

    TABLE 1.2

    Yeast Species That Produce Ethanol as the Main Fermentation Product

    Source: Lin and Tanaka, 2006.

    TABLE 1.3

    Comparison between Biodiesel and Petroleum Diesel

    are produced during the pretreatment. As mentioned above they can be reduced by an additional detoxification step, but this is an expensive operation (Van Maris et al., 2006).

    Molecular Biology Trends in Bioethanol Production Development

    In the last few years technologies breakthrough has compelled us for an alternative feedstock due to considerable shortage in agricultural land. In this sense, advances in metabolic pathway engineering/genetic engineering have led to the development of microbes skilled enough to convert biomass into ethanol (Das Neves et al., 2007). Generally, such development depends on expansion of the substrate range and inclusion of other biomass sources like arabinose or xylose in strains that cannot ferment sugars other than glucose. Examples of such microorganisms include genetically modified Escherichia coli, Saccharomyces sp., and Zymomonas mobilis, etc. (Davis et al., 2006).

    In cellulosic ethanol industry, aside from Pichia stipitis, natural xylose fermenting yeast, more efforts are being taken in obtaining recombinant bacterial and yeast strains that are able to ferment pentose sugars, such as arabinose and xylose. Figure 1.5 is one among the best examples depicting recombination process in microbes, where the tail end in E. coli and Klebsiella oxytoca or the front end of S. cerevisiae and Z. mobilis can be recombined for improved production of ethanol (Hagerdal et al., 2006).

    FIGURE 1.5 Strains that can be metabolically engineered for ethanol production. Source: Hagerdal et al., 2006. (For color version of this figure, the reader is referred to the online version of this book.)

    Moreover, genetic engineering of plants is another promising area, which most likely plays a key role in biofuel industry. The latest hybrid varieties have helped us considerably in improving starch yield from energy crops. For example, 25 kg of corn contains about 15 kg of starch. In the near future, that same 25 kg may contain as much as 17 kg of starch through hybrid corn. This would result in a gain of nearly $2 million in annual income by processing the same amount of corn in a 120 million liter per year ethanol production (DOE, 2007).

    Bioreactors in Ethanol Production

    A major commitment in cost-effective lignocellulosic bioethanol production is to employ reactor systems yielding the maximal cellulose conversion with the minimal enzyme. As a result, one of the most vital parameters for the fabrication and operation of bioreactors for lignocellulosic conversion is the efficient use of the enzymes to gain high specific rates of cellulose conversion (yield of glucose attained/amount of enzymes). The maximization of the product concentration, i.e. the amount of glucose obtained per liquid volume, is also a significant parameter as well as the optimization of the volumetric productivity.

    When hydrolysis is carried out with biomass comprised of high cellulose levels, the product concentration will drive up. For this reason, few researchers are attempting the enzymatic biomass conversion with high biomass loads (Jorgensen et al., 2007). The most imperative difficulty in high biomass loads is related to the viscosity of reaction mixture, which also influences the rheology of the mixture. In particular, mixing and mass transfer limitations and presumably increased inhibition by intermediates come into play. A variety of fed-batch strategies have been adopted with the scope of supplying the substrate without reaching excessive viscosities and unproductive enzyme binding to the substrate (Rudolf et al., 2005).

    General criteria in bioreactor design and in the choice of the operating conditions could be use of bioreactors or reaction regimes that allow a rapid decrease in the glucose concentration; running of the reactions at low to medium substrate concentrations in order to maintain higher conversion rates and thus obtain higher volumetric output of the reactor (Andric et al., 2010).

    The combination of the bioreactor with a separation unit has obtained prospective results with product inhibited or equilibrium limited enzyme-mediated conversions, because it potentially removes the products as they are accumulated (Gan et al., 2002). In this regard, membrane bioreactors could be a feasible process configuration. Unlike the Solid State Fermentation (SSF) approach in which the glucose consumption is carried out by the microbes simultaneously accessible in the hydrolyzate, the use of membrane bioreactors would finish the same function without any compromise in the reaction parameters. A membrane bioreactor (Figure 1.6) is a multitasking reactor that combines the reaction with a separation, namely, in this case the product was taken away by membrane separation, as one integrated unit (in situ removal) or alternatively in two or more separate units. The membrane bioreactors used for this separation processes are mainly ultra- and nanofiltration types (Pinelo et al., 2009). However, the use of this technology is restricted by the accumulation of unreacted lignocellulosics in large level and/or continuous processing (Andric et al., 2010). Already in the past, few scientists enhanced the efficiency of the continuous stirred tank bioreactor by incorporating membrane separation technologies during the reactor design.

    FIGURE 1.6 Schematic of membrane bioreactor integrated with membrane distillation (MD) process for alcohol distillation. Source: Gryta, 2012. (For color version of this figure, the reader is referred to the online version of this book.)

    Recently, an advanced reactor system was intended that removes the reducing sugars during the enzymatic hydrolysis of cellulose through a system consisting in a tubular reactor, in which the substrate was retained with a porous filter at the bottom and buffer entered at the top through a distributor (Yang et al., 2006). This hollow fiber ultrafiltration module with polysulfone membrane enabled the permeation and the separation of the sugars. To keep the volume constant in the tubular reactor, the entire buffer was recycled back from the ultrafiltration membrane and the makeup buffer was continuously supplied from the reservoir. In some applications an additional microfiltration unit has exceptionally been used to retain the unconverted lignin-rich solid fraction due to the presence of firmly bound enzymes or has been employed to remove the unconverted substrate from the reactor. These setups result in slightly complex process layouts for the hydrolysis (Knutsen and Davis, 2004).

    It is obvious that the optimization of the reactor designs will allow overcoming both the rheological and inhibition limit of the bioconversion and maximizing the enzymatic conversion. Therefore, the reactor design becomes more relevant for large-scale processing of cellulosic biomass.

    Immobilization of Cells for Ethanol Production

    For bioreactor application, immobilization of cells is a technique that has proved augmented ethanol productivity, operation stability and easier downstream processing, compared to processes using suspended cells (Das Neves et al., 2007). However, the specific advantages of immobilized cells depend on the nature of cells, reactor design and nature of the process. Entrapment of cells in natural polymers by ionic gelation (alginate) or by thermal precipitation (carrageenan and agar) is a method commonly used for cell immobilization (Ogbonna et al., 1991). Immobilization by passive adhesion to surfaces has great potential for industrial application since the immobilization method is relatively simple. The use of cheap carriers ensures that this method can be exploited with minimal increase in the overall production cost. Thus, one limiting factor of this technology is that it can only be adapted for practical industrial production if the expected increase in bioethanol productivity can overcome the increase in the production costs (cost of the carrier and immobilization) (Ogbonna et al., 1996).

    Biodiesel

    Biodiesel is a form of diesel fuel manufactured from vegetable oils, animal fats, or recycled restaurant greases. It is safe, biodegradable, and produces less air pollutants than petroleum-based diesel. Biodiesel can be used in its pure form (B100) or blended with petroleum diesel. Common blends include B2 (2% biodiesel), B5, and B20.

    Biodiesel is an ideal biofuel contender that eventually could replace petroleum based diesel. Currently, biodiesel production is still too costly to be commercialized. Due to the static cost associated with oil extraction and biodiesel processing and the variability in biomass production, future cost-saving efforts for biodiesel production should focus on the production of oil-rich feedstocks like microalgae, nonedible oils, etc.

    As discussed above, biodiesel is costlier than conventional diesel fuel, although it is rarely quoted as being competitive, as it will be if existing fluctuations in feedstocks/product prices are favorable. Using the distribution of these prices over the last 20 years, less than 5% of cost–benefit analyses based on fixed prices over the project life will show a positive result in producing biodiesel. If the feedstocks/product prices are varied each year, as will be the case in reality, biodiesel production will always be more expensive than conventional diesel (Duncan, 2003).

    Feedstocks for Biodiesel

    Biodiesel can be made from any oil/lipid source; the major components of these sources are triacylglycerol molecules. In general, biodiesel feedstocks can be categorized into three groups: pure vegetable oils, animal fats, and waste cooking oils.

    Biodiesel from Pure Vegetable Oil

    The first group is pure oils derived from various crops and plants such as soybean, canola (rapeseed), corn, cottonseed, flax, sunflower, peanut, and palm. These are the most widely used feedstocks by commercial biodiesel producers. The oil composition from vegetable crops is pure; this cuts down on preprocessing steps and makes for a more consistent quality of biodiesel product. However, there is an obvious disadvantage for vegetable oils as biodiesel feedstocks: wide scale production of crops for biodiesel feedstocks can cause an increase in worldwide food and commodity prices. Such a food vs fuels debate has reached national attention when using vegetable oils for biodiesel production. Alternative feedstocks usually arise out of necessity from regions of the world where the above materials are not locally available or as part of a concerted attempt to reduce reliance on imported petroleum.

    Jatropha curcas (Jatropha)

    The nonedible oil from Jatropha curcas (Jatropha) has recently attracted extensive attention as a feedstock for biodiesel production in India and other climatically parallel regions of the world (Kumartiwari et al., 2007; Kalbande et al., 2008). The Jatropha tree is a perennial shrub belonging to the Euphorbiaceae family whose seeds contain up to 30 wt% oil. This plant can be found in tropical and subtropical regions such as Africa, Indian subcontinent, Central America, and other countries of Asia. Since Jatropha oil contains a relatively elevated percentage of saturated fatty acids (Table 1.4), the corresponding methyl esters display relatively poor low temperature operability, as evidenced by pour point (PP) value of 2 °C (Kumartiwari et al., 2007).

    TABLE 1.4

    Biodiesel Production from Feedstocks High in Free Fatty Acids

    ∗Acid value (mg KOH/g) was given instead of FFA.

    ∗∗Conversion to esters (wt%) is provided instead of yield.

    Pongamia pinnata (Karanja)

    Another nonedible biomass originated in India is Pongamia pinnata (Karanja), which is a medium-sized deciduous plant that grows fast in damp and subtropical environments and matures in 5–7 years to tender fruit that contains two kidney-shaped kernels (Mohibbeazam et al., 2005). The oil content of Karanja kernels ranges between 25 wt% and 40 wt% (Karmee et al., 2005; Mohibbeazam et al., 2005). The primary fatty acid found in Karanja oil is oleic acid (45–70 wt%), followed by palmitic, linoleic, and stearic acids (Karmee et al., 2005; Naik et al., 2008). The low-temperature operability of the parallel methyl esters from karanja is superior to that of jatropha oil methyl esters as a result of the fairly high percentage of oleic acid in karanja oil, as evidenced by cloud point (CP) and PP values of −2 °C and −6 °C, respectively (Srivastava and Verma, 2008).

    Madhuca indica (Mahua)

    Madhuca indica, commonly known as Mahua, is a tropical plant found frequently in the central and northern plains and forests of India. It belongs to the family Sapotaceae and grows rapidly up to 20 m in height, possesses evergreen or semievergreen foliage, and is well adapted to dry environments (Ghadge and Raheman, 2006; Kumari et al., 2007). The fruit is nonedible, obtained from the tree in 4–7 years and contains one to two kidney-shaped kernels (Mohibbeazam et al., 2005). The oil content of dried Mahua seeds is about 50 wt%. Mahua oil is characterized by free fatty acid (FFA) content of around 20 wt% and a comparatively high percentage of saturated fatty acids such as stearic (14.0 wt%) and palmitic (17.8 wt%) acids (Ghadge and Raheman, 2006). The remaining fatty acids are mostly spread among unsaturated components such as linoleic (17.9 wt%) and oleic (46.3 wt%) acids (Singh and Singh, 1991). The relatively high percentage of saturated fatty acids (35.8 wt%) found in Mahua oil results in relatively poor low-temperature properties of the parallel methyl esters, as evidenced by PP value of 6 °C (Ghadge and Raheman, 2006).

    Nicotiana tabacum (Tobacco)

    Nicotiana tabacum, commonly referred as tobacco, is a commercial shrub with pink flowers and green capsules containing abundant small seeds grown in a large number of countries around the world. The foliage of the plant is the commercial product and used in the preparation of cigarettes and other tobacco-containing products. The oil content of the seeds, a by-product from tobacco, ranges from 36 wt% to 41 wt% (Usta, 2005). This tobacco seed oil contains more than 17 wt% FFAs (Veljkovic et al., 2006) and is high in linoleic acid (69.5 wt%), along with oleic (14.5 wt%) and palmitic (11.0 wt%) acid in significant amounts. Due to high linoleic acid content of tobacco seed oil, the corresponding methyl esters display relatively low kinematic viscosity (3.5 mm²/s) in comparison to most other biodiesel fuels (Usta, 2005).

    Biodiesel from Animal Fat Wastes

    The feedstock issues are very critical, which affect the economic potential of biodiesel production, since feedstock accounts around 75% of the biodiesel total cost (Figure 1.7). Recently, alternative lipid residues such as waste frying oil and nonedible animal fats have also received substantial attention from the biofuel sector. To take benefit of these low-cost and low-quality resources, a suitable act would be to reuse residues in order to integrate sustainable energy supply and waste management in food processing facilities. Animal fats are typically considered as waste by-products and less expensive than commodity vegetable oils, which make them attractive as feedstock for biodiesel production. These animal wastes are collected from chicken, cow, pork lard, and other animals such as fish and insects.

    FIGURE 1.7 Biodiesel production cost summary sheet. Source: Pruszko, 2007. (For color version of this figure, the reader is referred to the online version of this book.)

    Beef Tallow and Chicken Fat

    Animal fats like beef tallow and chicken fat are by-products from the meat industry and stand for cheap feedstock for biodiesel production. The key fatty acids found in beef tallow were oleic (47.2 wt%), palmitic (23.8 wt%), and stearic (12.7 wt%) acids. The prime fatty acids contained in chicken fat include oleic (40.9 wt%), palmitic (20.9 wt%), and linoleic (20.5 wt%) acids (Wyatt et al., 2005). Due to very low concentration of polyunsaturated fatty acid in beef tallow, the corresponding methyl esters illustrate excellent oxidative stability, as evidenced by an oil stability index (OSI) value of 69 h at 110 °C. In addition, other physical properties of beef tallow methyl esters include kinematic viscosity (40 °C) of 5.0 mm²/s, a flash point (FP) of 150 °C, and CP, PP and cold filter plugging point (CFPP) values of 11, 13, and 8 °C respectively (Moser, 2009). In chicken fat, due to high polyunsaturated fatty acid content, the corresponding methyl esters display poor oxidative stability, as evidenced by an OSI value of 3.5 h at 110 °C. Burning the B20 blends of beef tallow and chicken fat methyl esters results in NOx exhaust emissions of only 2.4% versus 6.2% of B20 blend of soybean methyl esters (SME) (Wyatt et al., 2005).

    Pork Lard

    Pork lard is a by-product of the food industry and symbolizes a low-cost feedstock for biodiesel production. The main fatty acids in pork lard includes stearic (121 wt%), linoleic (127 wt%), oleic (44.7 wt%), and palmitic (26.4 wt%) acids (Jeong et al., 2009). Due to high saturated fatty acid content in pork lard, the corresponding methyl esters exhibit quite high CFPP value of 8 °C and a relatively low iodine value (IV) of 72, along with a typical kinematic viscosity (40 °C) of 4.2 mm²/s. Another study determined that pork lard methyl esters have a kinematic viscosity (40 °C) of 4.8 mm²/s, FP of 160 °C, OSI value of 18.4 h at 110 °C, and CP, PP, and CFPP values of 11, 13, and 8 °C, respectively (Wyatt et al., 2005). Furthermore, combustion of B20 blends of pork lard methyl esters results in NOx exhaust emissions of only 3.0% versus 6.2% for a B20 blend of SME.

    Other Waste Cooking Oils

    Waste oils may include a variety of low-worth materials such as used cooking or frying oils, acid oils, tall oil, vegetable oil soapstocks, and other waste materials. Waste oils are usually characterized by relatively high FFA and water contents and potentially contain a variety of solid materials that must be removed by filtration prior to conversion to biodiesel (Moser, 2009). In the case of used cooking or frying oils, hydrogenation to increase the useful cooking lifetime of the oil may result in the introduction of relatively high-melting trans constituents, which influence the physical properties of the resulting biodiesel. Used frying or cooking oil is mainly acquired from restaurants and may cost between free to 50% less expensive than commodity vegetable oils, depending on the source and the availability (Predojevic, 2008). The physical properties of methyl esters prepared from used cooking or frying oils include kinematic viscosities (40 °C) of 4.23 (Meng et al., 2008), 4.79, and 4.89 mm²/s; FP of 171 °C; cetane number of 55, IV of 125, CFPP values of 1 and −6 °C (Cetinkaya and Karaosmanoglu, 2004), CP values of 9 and 3 °C, and PP values of −3 and 0 °C (Phan and Phan, 2008). The disparities in the physical property data among the various studies may be a result of feedstock origin or due to differences in product purity.

    Algae as a Biodiesel Source

    Algae can also be used to produce energy in a number of ways. One of the most competent ways is through exploitation of the algal oils to produce biodiesel. Algal biomass contains three major components: carbohydrates, proteins, and lipids/natural oils (Dunahay et al., 1996). Because the natural oil made by microalgae is in the form of triacylglycerol molecule, which is the right kind of oil for producing biodiesel, microalgae are the exclusive focus in the algae to biofuel arena. Actual biodiesel yield per hectare is about 80% of the yield of the parent crop oil given in Table 1.5.

    TABLE 1.5

    Comparison between Few Biodiesel Sources

    ∗About 70% oil (by wt) in biomass.

    ∗∗About 30% oil (by wt) in biomass.

    ∗∗∗For meeting 50% of all transport fuel needs of the United States.

    In view of Table 1.5, microalgae emerged to be the only source of biodiesel that has the potential to completely replace petroleum diesel. Unlike other oil crops, microalgae grow extremely rapidly and many are exceedingly rich in oil. Microalgae commonly double their biomass within 24 h. Biomass doubling times during exponential growth are commonly as short as 3–4 h. Oil content in microalgae can exceed 70% by weight of dry biomass (Metting, 1996; Spolaore et al., 2006). Oil levels up to 50% are quite common. Oil productivity, the mass of oil produced per unit volume of the microalgal broth per day, depends on the algal growth rate and the oil content of the biomass. Microalgae with high oil productivities are desired for producing biodiesel.

    Chemical Transesterification Process for Biodiesel Production

    The source oil used in making biodiesel consists of triglycerides (Figure 1.7), in which three fatty acid molecules are esterified with a molecule of glycerol. In biodiesel production, triglycerides are reacted with methanol in a reaction known as transesterification or alcoholysis. Transesterification produces methyl esters of fatty acids that are biodiesel and glycerol (Figure 1.7). The reaction occurs stepwise: triglycerides are first converted to diglycerides, then to monoglycerides and finally to glycerol.

    At equilibrium, transesterification needs 3 mol of alcohol for every mole of triglyceride to produce 1 mol of glycerol and 3 mol of methyl esters (Figure 1.8). Industrial processes use 6 mol of methanol for each mole of triglyceride (Fukuda et al., 2001). This large excess of methanol ensures that the reaction is driven in the direction of methyl esters, i.e. toward biodiesel. Yield of methyl esters exceeds 98% on a weight basis.

    FIGURE 1.8 Transesterification of oil to biodiesel.

    Transesterification is catalyzed by acids and alkalis (Fukuda et al., 2001). Alkali-catalyzed transesterification is about 4000 times quicker than the acid-catalyzed reaction. Thus, alkalis such as sodium and potassium hydroxide are frequently used as commercial catalysts at a concentration of about 1% by weight of oil. Alkoxides such as sodium methoxide (CH3ONa) act like better catalysts than sodium hydroxide and are being increasingly used. Use of lipases offers significant advantages, but it is currently not feasible because of the relatively high cost of the catalyst (Chisti, 2007). Alkali-catalyzed transesterification is carried out at about 60 °C under one atmospheric pressure, as methanol boils off at 65 °C at atmospheric pressure. Under these conditions, reaction takes about 90 min to complete (Meher et al., 2006). A higher temperature can be used in combination with higher pressure, but the process becomes expensive. During reaction, methanol and oil do not mix; hence, the reaction mixture shows two liquid phases. Other alcohols can be used, but methanol is the least expensive. To stop yield loss due to saponification reactions (soap formation), the oil and alcohol must be dry and the oil should have a least of FFAs. Biodiesel is recovered by repeated washing with water to remove glycerol and methanol.

    Enzymatic Transesterification Process for Biodiesel Production

    Lipases (triacylglycerol hydrolase, EC 3.1.1.3.) are enzymes that catalyze the breakdown of carboxylic ester link in the triacylglycerol molecule to form FFAs, di- and monoglycerides and glycerol. Although their purpose is to catalyze hydrolysis of ester links, they can also catalyze the esterification, the conception of this link between alcohol hydroxyl groups and carboxyl groups of carboxylic acids. Therefore, they can catalyze hydrolysis, alcoholysis, esterification and transesterification and they have a wide spectrum of biotechnological applications (Kirk et al., 2002). Lipases are also highly specific as regio, chemo and enantioselective catalysts. Thanks to protein engineering, it is possible to enhance catalytic potential of lipases and tailor them to exact application and process situation, enabling further expansion of their industrial applications (B van Beilen and Li, 2002). Among lipases from animal, plant and microbial origins, the most commonly used are microbial lipases. They have abundant advantages over lipases from animal and plant sources. Using microbes it is possible to achieve a higher yield of enzymes, and to genetically control the strain in obtaining a low-cost lipase with preferred properties for the conversion of fats and oils into biodiesel. In addition, the enzymatic yield is independent of potential seasonal variations and it is possible to achieve rapid growth of microbes in low-cost media (Gupta et al., 2004).

    Bioreactors for Biodiesel Production

    Microalgae are unicellular microscopic organisms, like simple plants with no leaves and roots that grow through photosynthesis process. They capture carbon dioxide during photosynthesis and convert it into feedstock that can be used as food, fertilizer, a source of medicine and biodiesel (Chojnacka and Marquez-Rochaet, 2004).

    Growing algae in open pond system raise several concerns such as impossibility to control growth settings and contamination threats. Algal cells in open ponds are exposed to the environment, light deficiency, subject to risk of contamination, and heterogenous medium depending upon the mixing mechanism, the shape of the ponds and the depth of the pond (Chojnacka and Marquez-Rochaet, 2004). On the other hand, closed ponds (photobioreactors) mitigate fluid culture contamination, and enhance full control over algal growth parameters such as homogenous culture, pH, light penetration, and carbon dioxide input. They would use less space with high algal biomass yield. However, they are costly to build and maintain (Mulumba and Farag, 2012).

    Design of tubular photobioreactor (TPBR) for algal cell growth was depicted in Figure 1.9. It has a main tank connected to two spiral tubes set in sequence. Both spiral parts were clear polyvinyl chloride (PVC) tubes of 1″ external diameter and 3/4″ internal diameter. The capacity of both spiral parts was 3.4 gallons. The main tank served as a feeding point of medium to the PVC tubes with a maximum capacity of 5 gallons (Chisti, 2007). Culture medium was pumped into the tubings at a fixed flow rate. These tubes provided an area of 20 ft² exposed to the fluorescence light. Air compressor supplies air to the system for aeration and to serve as a source of carbon dioxide (CO2). The air flow rate was set in the ranges of 190–210 gallons/h. In TPBR, selected algal strain was cultured using fresh medium with no modification. Algal growth and pH were measured over a period of time varying between 12 days and 14 days. A sample was taken every 2 days to quantify the turbidity using a spectrophotometer at 682 nm and cell counts were performed using a microscope. The pH of culture was measured using pH test strips.

    FIGURE 1.9 Schematic of tubular photobioreactor with airlift system. Source: Molina et al., 2001.

    The selected algal strain shows the typical growth curve of other microbes, which include lag, exponential or log, stationary and lytic phases. The length of each phase depends on light penetration, nutrients concentration, mixing mechanism, and the solubility of oxygen in medium. After reaching a stationary or lysis phase, algal culture was harvested by centrifugation followed by lyophilization to produce dry algal feedstock. Crude lipid from dried algal biomass was extracted using either modified Folch method (Cooksey et al., 1987) or Soxhlet extractor (Mulumba, 2010; Chojnacka and Marquez-Rochaet, 2004). In both methods, polar and nonpolar solvents such as methanol and chloroform/hexane were used (Table 1.6). The combination of polar and nonpolar solvents enhances the extraction of both polar and nonpolar lipid.

    TABLE 1.6

    Biodiesel Production with Various Lipases

    Biogas

    Biogas is obtained by anaerobic digestion (AD) of organic materials, which occurs inside the anaerobic biodigester. Chemical composition of this biogas depends on several parameters, such as type of digester employed, the kind of organic material and the constancy of the feeding process of the biodigester. The most significant biogas components are methane (CH4), carbon dioxide (CO2) and sulfuric components (H2S). The composition of biogas is a crucial parameter, because it allows identifying the suitable purification system, which aims to remove sulfuric gases and reduce the water volume, contributing to recover the combustion fuel conditions (Boe et al., 2007). Other important data collected from biogas analysis is referent to the low heat value, that combined to the efficiency and biogas consumption is important to estimate the electric generation potential. However, biogas production is much variable because it depends on several parameters, such as the kind of organic material (Liu et al., 2004). Biogas production involves three steps: fermentation, which includes hydrolysis and acid genesis, acetone genesis and methane genesis. In the fermentation process, during the hydrolysis the organic material is converted into smaller molecules and this material is transformed in soluble acids by acidogenese. Next step is acetanogenese process, transforming the products obtained in the first step into acetic acid, hydrogen and carbon dioxide. The last step is referent to metanogenese process, producing methane gas through anaerobic bacteria (Figure 1.10) (Seadi et al., 2008; Boe et al., 2007).

    FIGURE 1.10 Biochemical process in anaerobic digester. (For color version of this figure, the reader is referred to the online version of this book.)

    Biogas Feedstock

    A wide range of biomass types can be used as substrates for the production of biogas by AD. The most common biomass categories used in biogas production are listed below and in Table 1.7. Animal manure and slurry, agricultural residues and by-products, digestible organic wastes from food and agro industries, organic fraction of municipal waste and from catering, and sewage sludge, etc. are best study sources for biogas production.

    TABLE 1.7

    Characteristics of Some Digestible Feedstocks

    ∗VS - volatile solids

    Source: Seadi et al., 2008.

    Recently, various novel feed stocks has been tested and introduced for biogas synthesis in many countries, the dedicated energy crops (DECs), crops grown specifically for energy and biogas production. DECs can be herbaceous (grass, maize, and raps) and also woody crops (willow, poplar, and oak), although the woody crops need particular delignification treatment before AD.

    In AD, substrates can be classified according to the following criteria: methane yield, origin, dry matter (DM) content, etc. Table 1.7 gives a summary on the characteristics of some digestible feedstocks. Substrates with DM content less than 20% are used for what is called wet digestion (wet fermentation), which includes animal slurries and manure as well as various wet organic wastes from food industries. When the DM content is as high as 35%, it is called dry digestion (dry fermentation), and it is typical for energy crops and silages. The choice of types and amounts of feedstock for the AD substrate mixture depends on their DM content as well as the content of sugars, lipids and proteins.

    Household Digesters for Biogas

    It is difficult to accept one particular type of digester for household biogas production. The design of the digesters is diversified based on the availability of substrate, geographical location, and climatic conditions. For example, a digester designed in mountainous regions has less gas volume in order to avoid gas loss. For tropical countries, it is recommended to have digesters underground due to the geothermal energy (Bin, 1989). Of all the different digesters developed, the fixed dome model developed in China and the floating drum model developed in India sustained to perform well until today (Rajendran et al., 2012). Recently, plug flow digesters are gaining attention due to its portability and easy operation.

    Fixed Dome Digesters

    The fixed dome digesters (Figure 1.11) is also called hydraulic or Chinese digesters and it is the most frequent model developed and used in China for biogas production (Rajendran et al., 2012). In this case, digester is filled through the inlet pipe until the level reaches the base level of the expansion chamber. Biogas that is produced is accumulated at the upper part of the digester called storage part. The difference in the levels between the slurry inside the digester and the expansion chamber develops pressure inside due to accumulation of biogas. This accumulated biogas requires space and presses the substrate apart and enters into the expansion chamber. The slurry flows back into the digester straight away after gas is released (Adeoti et al., 2000). Fixed dome digesters are usually built underground and the size of the digester depends on the place, number of households, and the amount of substrate available every day. Generally size of these digesters normally varies between 5 m³ and 150 m³ in various parts of Asia (Tomar, 1994). Instead of having a digester for each home, a large-volume digester is used to produce biogas for 10 to 20 homes, and is called community-type biogas digesters. In countries where houses are clustered as in Africa, these types of biogas digesters are more viable (Adeoti et al., 2000).

    FIGURE 1.11 Schematic sketch of fixed dome digester. Source: GMI, India, 2013. (For color version of this figure, the reader is referred to the online version of this book.)

    Floating Drum Digesters

    The floating drum digester was first time constructed by Khadi and Village Industries Commission and this model was developed in 1962 (Figure 1.12). Although the model is old, it is one of the most extensively used designs for household purposes in India. This design includes a movable inverted drum placed on a well-shaped digester. The inverted steel drum acts as a storage tank, which can move up and down depending on the quantity of accumulated biogas at the top of the digester. The weight of this inverted drum applies the pressure needed for biogas flow through the pipeline (Singh and Sooch, 2004).

    FIGURE 1.12 Floating drum digester. Source: Working of biogas plants Working of biogas plants, 2013, www.tutorvista.com. (For color version of this figure, the reader is referred to the online version of this book.)

    Floating drum digesters manufacture biogas at a stable pressure with variable volume. In floating drum reactor, by position of the drum, the amount of biogas accumulated under the drum is easily noticeable. However, the floating drum needs to be coated with paint at regular intervals to avoid rusting. Additionally, fibrous materials in biomass will block the movement of the digester. Hence, their accumulation must be avoided if possible (Adeoti et al., 2000). In Thailand, the floating dome has been customized with two cement jars on each side of the floating drum. The average size of these digesters is around 1.2 m³ (Gosling, 1982). For small and medium-size farms the size varies from around 5 to 15 m³. Singh and Singh (1991) compared 14 different biogas plants with a floating drum model and optimized the various parameters for maximum biogas production.

    Social and Environmental Aspects of Biogas Digesters

    Change in the global climate is a major threat that the world is facing today. The nonrenewable energy consumption in the past has led to global warming that needs to be addressed (Bilen et al., 2008). The household digesters could reduce the pressure on the environment by dropping deforestation and GHG emissions followed by loss of cultivable land, and soil erosion (Gautam et al., 2009). Biogas production in rural areas can partly reduce global warming (Pei-dong et al., 2007). By using biogas in rural households economical, environmental, and social benefits were achieved (Yang et al., 2011). Even though both carbon dioxide and methane are major contributors to the greenhouse effect, the global warming effect of methane is 21 times greater than that of carbon dioxide (Dhingra et al., 2011). However, houses equipped with biogas systems exhibit leakage of gases in the biogas systems. Fortunately, the households with biogas plants have 48% less emissions compared to households without biogas systems (Pathak et al., 2009). It is worth talking about 10% of households, which had methane leakage (Yang et al., 2011). Research has already shown that by replacing firewood and coal with biogas, the emission of CO2 and SO2 would be reduced by 4193 thousand tons, and 62.0 thousand tons, respectively (Pei-dong et al., 2007).

    Conclusion

    We conclude that by accelerating research in areas of bioenergy, we can make significant contributions to sustainable development and use of feedstock. We must realize that by maximizing biomass conversion efficiency, we can minimize raw material requirements, while at the same time the financial position of various market sectors (e.g. energy, agriculture, and forestry) are strengthened. There is an international agreement on the fact that the feedstock accessibility is inadequate so that the raw materials should be used as competently as possible, i.e. expansion of multipurpose industries (biorefineries) that can utilize variable biomass sources as raw materials for bioenergy production. The main constraint in making this biorefinery a successful path is bringing the stakeholders together, who normally operate in different market sectors (e.g. energy, agriculture and forestry, fuel transportation, etc.). Above all, the government should make policies to help overcome the threshold by dropping production costs in the form of feedstock in tariffs, feedstock in premiums, tax exemptions, etc. These encouragements can be targeted at different parts of the supply chain like feedstock producers, energy producers, and distributors.

    References

    1. Adeoti O, Ilori MO, Oyebisi TO, Adekoya LO. Engineering design and economic evaluation of a family-sized biogas project in Nigeria. Technovation. 2000;20:103–108.

    2. Adler PR, Sanderson MA, Weimer PJ, Vogel KP. Plant species composition and biofuel yields of conservation grasslands. Ecol Appl. 2009;19:2202–2209.

    3. Agarwal AK. Biofuels (alcohols and biodiesel) applications as fuels for internal combustion engines. Prog Energy Combust Sci. 2007;33:233–271.

    4. Alvira P, Tomas PE, Ballesteros M, Negro MJ. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis: a review. Bioresour Technol. 2010;101:4851–4861.

    5. Andric

    Enjoying the preview?
    Page 1 of 1