1

Introduction

An overview of biofuels and production technologies

C. Du1, X. Zhao2, D. Liu2, C.S.K. Lin3, K. Wilson4, R. Luque5,  and J. Clark6     1University of Huddersfield, West Yorkshire, United Kingdom     2Tsinghua University, Beijing, China     3City University of Hong Kong, Hong Kong, China     4Aston University, Birmingham, United Kingdom     5University of Cordoba, Cordoba, Spain     6University of York, York, United Kingdom

Abstract

Biofuel is a rapidly growing research field and fast-moving industry. Since the publication of the first edition of Handbook of Biofuels Production in 2011, significant research progresses in biofuel production technology have been made, improved understanding of biofuel production processes has been acquired, and the industrial production of biofuels has moved forward. With this in mind, the second edition of the handbook keeps the underlying principles of various biofuel production technologies, and covers the latest progress in biofuel-related fields.

Keywords

Biodiesel; Bioethanol; Biofuel; Biomass; Biopyrolysis; Fatty acid methyl ester; Fossil fuel; Greenhouse gas emission; Lignocellulosic

1.1. Introduction

The increasing demand of renewable energy and the growing concern of global warming are still considered to be key challenges for the society worldwide. The sustainable development of the energy industry needs a continuous supply of renewable, sustainable energy. In 2013, over 90 million barrels of crude oil were consumed globally each day (US Energy Information Administration, 2014). Along with economic and population growth, the demand of energy will surge as well. Currently, 80% global energy consumption came from fossil resources, namely crude oil, natural gas, and coal. These fossil fuels are generated from organic materials that synthesized on Earth millions of years ago, and are unable to be regenerated within a short period, eg, it takes over hundreds of years for regeneration. Although the recent booming of shale gas releases the tension of the fossil fuel shortage and drags down the fossil fuel price, the finite nature of fossil fuel does not change. Based on the current daily fossil-usage data, the fossil regeneration rate (even the fossil discovery rate) will never match the consumption rate. A decade ago, some scientists warned that the fossil fuel would run out in 40 years. Our fossil fuel reserves might last for 40 or 100 years, depending upon the conditions that are put on our fossil fuel use (Dunlap, 2015). Optimists even consider that with the increasing fossil exploration, the fossil fuel would last longer than our current estimation. However, even if fossil fuel could last 300 years, this is just a short spell in human history. The exploration of new, renewable energy resources cannot wait until the depletion of fossil fuel.
On the other hand, the appeal of the reduction of greenhouse gas (GHG) emission has been the hottest topic in every recent United Nations Climate Change Conference. In 1997, the Kyoto Protocol was signed by most of the industrialized countries with the aim of reducing the global GHG emission. After the Kyoto Protocol's first commitment period expired on 2012, 37 countries, including 28 members of the European Union, agreed to a second commitment period of GHG emission reduction in Doha. Although the two largest GHG emission countries did not participate in the Kyoto Protocol, they both set their own CO2 emission targets (UNFCCC, 2015). According to the latest report of the Intergovernmental Panel on Climate Change (IPCC, 2014), the GHG concentration in the atmosphere could reach from 750 to 1300-ppm CO2 equivalents. As a consequence, the global average surface temperature could increase by 3.7–4.8°C. If we would like to control the temperature change within 3°C in 2100 compared to that of preindustrial levels, the GHG concentration in the atmosphere should be controlled to lower than 650-ppm CO2 equivalents. This means a change of GHG emission should at least not exceed 24% of the 2010 emission level (IPCC, 2014). Since 78% of GHG emissions in recent decades came from fossil fuel combustion and industrial processes, the development of a low-carbon economic system to replace the fossil fuel–based system is urgent.
Along with several other renewable technologies, biofuel has made and will continuously make a significant contribution to meet targets on the usage of renewable energy resources and the reduction of GHG emission. Besides the above-mentioned major reasons, the advantages of development and application of biofuels also include: improving national energy security, utilizing existing transportation system, utilizing existing fuel distribution system, and facilitating rural development.
Currently, in the first generation of bioethanol, food crops such as corn, sugar cane, and wheat are used for the production of energy. These are starch- or sucrose-rich feedstocks that are readily fermented by microorganisms. However, these crops are also used for food and feed production, resulting in competition. At present, commercial production of the first-generation biomass utilizes readily-available sugars from these food plants for the fermentation process of biofuel production.
However, the second generation of bioethanol uses lignocellulosic raw materials as the main substrate, which has a more complex composition as compared to the first-generation feedstocks. Lignocellulosic feedstocks are high in cellulose, hemicellulose, and lignin. Second-generation feedstocks avoid competition with food and feed products. Examples are waste streams from food- or feed-crops such as wheat straw or corn stover, also municipal or industrial waste streams, or energy crops that grow on marginal lands that are unsuitable for regular agriculture. To use the preferred second-generation feedstocks, further advances in technological development are needed to unlock the more hidden sugars in the crop residues or woody plant materials. Significant research efforts and investment have been spent to improve the technology in order to enable the commercial use of the second-generation feedstocks.
Different generations of biofuels also differ in other characteristics. While the food part of the food crops is made of easily digestible sugars, the sugars captured in lignocellulosic compositions of the second-generation feedstocks are more difficult to utilize. So why do we want to use these more challenging second-generation feedstocks? This is to reduce competition with food, arable land, and water. Using residues can help to avoid land-use changes, and energy crops can be genetically engineered to reduce water usage. It can also bring in extra income for farmers. In the future, water-based feedstocks such as algae may become as important as the third-generation feedstocks. The third-generation feedstock is used for processes where CO2 is utilized as one of the substrates. A common example would be photosynthetic algae that use sunlight and CO2 to produce useful organic molecules. These third-generation systems would completely eliminate the need for agricultural land.
This book aims to provide an overview of the latest progresses in various technologies for biofuel production. The special emphasis has been focused on the advanced generation of biofuels, which produce biofuels from nonfood materials. We keep the same the classification method, dividing different technologies into three main sections: chemical, biological, and thermochemical conversions.
In the first few introductory chapters, details on policies, socioeconomic, and environmental implications of the implementation of biofuels (chapter: Multiple objectives policies for biofuels production: environmental, socioeconomic and regulatory issues), life-cycle assessment (LCA) (chapter: Life cycle sustainability assessment of biofuels), techno-economic assessment (chapter: Techno-economic studies of biofuels), environmental concern (chapter: Multiple objectives policies for biofuels production: environmental, socio-economic and regulatory issues), and the different biofuel feedstocks (chapter: Feedstocks and challenges to biofuel development) will be presented. The rest of the book is aimed to give a detailed and balanced overview on key technologies and processes for the production of various type of biofuels, including but not limited to, bioethanol, biodiesel, biohydrogen, biogas from anaerobic digestion, biosyngas from gasification, and bio-oil from pyrolysis.

1.2. Development of (bio)chemical conversion technologies

The utilization of “biofuels” in transportation has a long history. In 1900, the Paris Exposition Universelle, a small version of the diesel engine, was shown, which runs on peanut oil. Using vegetable oil in diesel engine began in the 1920s and continued through the early 1940s. With the booming of oil industry, together with the shortcomings of directly using vegetable oil, eg, high viscosity, petroleum diesel has been predominately used in the diesel engine. In the 1970s, the oil crisis sparked interests in biofuels. Austria started biodiesel research in 1974, and in 1985, a pilot plant producing 500 tons/year of biodiesel was built in Styria, Austria, using rapeseed oil as the starting material. Fig. 1.1 shows the biodiesel production trend in Europe. European countries had set its own policy to blend biodiesel to petroleum diesel. In the United Kingdom, around 3.4% of the total diesel used in 2014 was biodiesel.
Biodiesel is a mixture of long-chain fatty acid methyl ester (FAME) that is produced from biomaterials through transesterification of triacylglycerol (TAG, ie, plant oil and animal fats) with methanol. In chemistry, the biodiesel synthesis could be expressed by the reaction as shown in Fig. 1.2. In principle, any triacylglycerol could be used for biodiesel production. In fact, the first generation of biodiesel was mainly produced from edible plant oil, such as soybean, rapeseed, and palm oil. The low price of plant oil before 2008 and high diesel price in the European countries allowed enough profit for the biodiesel production. For example, during 2004–2005, the rapeseed oil price was less than €600/ton, while the diesel price was €1.11/L. However, in 2008, the plant oil price increased sharply, and the margin of biodiesel production from plant oil was low. During the soaring of raw material cost, together with the concern of food shortage, production of biodiesel from nonedible oil, eg, waste cooking oil, grease, Jatropha oil, and microalgae attracted increasing interest. It was reported that in the United Kingdom between April 2014 and 2015, over 50% of the biodiesel would be derived from waste cooking oil (Biofuel statistics, UK government, 2014).
image
Figure 1.1 Biodiesel production trend in Europe in the period of 2003–2013.
image
Figure 1.2 The chemical equation of transesterification reaction.
In terms of biodiesel conversion processes, chemical conversion using alkali and acid-based catalysts is still the most favorite approach. Various investigations have been carried out to develop novel catalysts and/or novel processes for efficient conversion of TAG to FAME. This part was reviewed in the chapter “Production of biodiesel via catalytic upgrading & refining of sustainable oleageneous feedstocks.” The chapter “Biochemical catalytic production of biodiesel” introduced a promising alternative way of biodiesel production via enzyme-catalyzed processes. Recently, microalgae has been demonstrated to have the potential for biodiesel production. Significant progress has been made since the concept was first introduced. The character of microalgae oil showed its great potential of dominating biodiesel production. Since the publication of the first edition, further intensive investigation on this field has been carried out. The advantages and limitations of microalgae oil production are discussed in the chapters “Production of fuels from microbial oil using oleaginous microorganisms” and “Utilization of biofuels in diesel engines.”

1.3. Development of biological conversion technologies

A wide range of biofuels could be produced from fermentative and biological processes, with bioethanol that dominated the liquid biofuel production. The first generation of bioethanol production has already been fully developed, and its products have been utilized in many countries across continents. The second generation of bioethanol is still the center of bioethanol research, with only a few pilot plants and demonstration plants on operation. However, Italy made a commitment of blending 0.6% advanced biofuel by 2018, and 1% by 2022. Also, it built up the world's first second-generation of bioethanol plant “Beta Renewables” at Crescentino near Turin. This plant was officially opened in October 2013, which was designed to produce 75 million liters of bioethanol per year. This will definitely boost the research and commercialization of the second generation of bioethanol.
Besides bioethanol (chapter: Biochemical production of bioalcohols), anaerobic digestion of organic waste materials for the biogas production progressed rapidly worldwide in the past 5 years. For example, the annual biogas production in China increased dramatically from 10.5 billion m3 to 248 billion m3 from 2007 to 2010 (Deng et al., 2014; Wellinger, 2011). The chapters “Production of biogas via anaerobic digestion,” “Biological and fermentative production of hydrogen,” and “Biological and fermentative conversion of syngas” reviewed biogas, biohydrogen, and fermentative conversion of syngas (synthesis gas), respectively.

1.4. Thermochemical conversion technologies

Direct combustion of biomass is one of the first types of energy that ancient people could manage. Burning biomass for cooking, keeping warm, and safety have been used by humans for thousands of years. Until now, biomass combustion still supplies around 11% of world energy. With the increasing incentive to utilize renewable materials for fuels and chemicals generation, various thermochemical conversion of biomass technologies emerged.
Biopyrolysis is a typical thermochemical process, which converts biomass into biosyngas, bio-oil, and biochar at elevated temperature with limited supply of air (oxygen). Bio-oil is the target product of biopyrolysis. A wood chip–derived bio-oil normally has a density of around 1.2 kg/L with an energy density of around 18.0 MJ/kg, which is around 2–4 times higher than those of the wood chips. However, the complex composition of the pyrolysis oil, the high water content, and the high acidity prevent wide applications of bio-oil. Various investigations have been carried out with the hope of upgrading bio-oils into a replacement of fossil transportation fuels, including optimizing operation conditions, pretreating biomass before pyrolysis, introducing catalysts, designing novel reactors, and many others, as reviewed in the chapters “Catalytic fast pyrolysis for improved liquid quality,” “Production of biofuels via hydrothermal conversion,” and “Production of biofuels via bio-oil upgrading and refining.”
Further increase to the reaction temperature in the biopyrolysis process would lead the thermochemical reactions to shift toward biosyngas production. Such a process is then termed gasification. The main components of biosyngas are CO and H2. The ratio of CO and H2 is depended on the type of substrate and gasification conditions, eg, whether steam is used in the gasification process. The resultant biosyngas could be burnt directly for energy generation, or to be used for the synthesis of biofuels via the Fischer-Tropsch process or a newly-emerged gas fermentation process. These contents were reviewed in the chapters “Production of biosyngas and bio-hydrogen via gasification,” “Production of bioalcohols via gasification,” “Production of biofuels via Fischer-Tropsch synthesis: biomass-to-liquids,” and “Production of biofuels via bio-oil upgrading & refining”. Besides these, the chapter “Chemical routes for the conversion of cellulosic platform molecules into high-energy density biofuels” discussed alternative approaches that could be used for the high-energy-density biofuels production via chemical conversion routes.

1.5. Process integration and biorefinery

The successful development of a biofuel production process, especially the second generation of biofuel, required knowledge from biotechnology, engineering, chemistry, plant science, and other relevant fields. Process integration is required to improve the mass and energy flow efficiency within a biofuel production process (Sadhukhan et al., 2014). Furthermore, production integration is also required, that is, to fully utilize the potential of the biomass raw materials and to generate a range of products. This concept is designated as “Biorefinery,” which is analogous to petroleum refineries (Clark and Deswarte, 2008). These products include high-volume, low-value products, such as transportation fuels (eg, bioethanol, biodiesel), medium-volume, medium-value products, such as platform chemicals and materials (eg, succinic acid, lactic acid, polyhydroxybutyrate [PHB]), as well as low-volume, high-value products, such as pharmaceuticals (eg, arteannuin, antioxidants).
Actually, the biorefinery concept has already been applied in the first generation of biofuels. Along with bioethanol production, Distiller's Dried Grains with Solubles (DDGS) is generated. DDGS is sold as an animal feed, and is an important income stream for a bioethanol company. Even more, companies consider themselves to be animal feed–producing or commodity food–producing companies—the biofuel production is just to utilize the low nutritional parts of the biomass or the organic waste to generate another product. Similarly, glycerol is the principle by-product of biodiesel, which is produced from transesterification of TAG with a glycerol to biodiesel mass ratio of 1:10. At the earlier stage in a biodiesel business model, glycerol is normally refined to pure glycerol and is sold as an income stream. However, due to the soaring of biodiesel production, the glycerol market was quickly saturated. As a consequence, the glycerol price dropped significantly. Therefore, various research has been carried out to convert glycerol, or crude glycerol into other value-added products, such as 1,3-propanediol, succinic acid, PHB, and biogas via anaerobic digestion (Koutinas et al., 2014).
Biofuel production actually plays a major role in the economics of biorefineries. The chapter “Biofuel production from food wastes” reviewed the topic of biofuel-driven biorefineries, and the chapters “Biochar in thermal and thermochemical biorefineries—production of biochar as a coproduct,” “Algae for biofuels: an emerging feedstock,” and “Utilization of biofuels in diesel engines” focused on the biofuels and other value-added production formation from the following interesting raw biomass: food waste, lignocellulose, and algae. Last but not least, engine tests are of utmost importance to test the feasibility of biofuels implementation and are still on-going activities. Chapter “Utilization of biofuels in diesel engines” summarized some experimental results on the implementation of biofuels in engine tests.

1.6. Future trends

In the past 5 years, the biofuel industry continuously grew with an average increase in annual biofuel production of 6.4% (BP, 2015 annual report). Currently, around 3% of world transportation fuel is provided by biofuel. According to a recent article published by the International Energy Agency, this figure could potentially grow up to 27% in 2050 (IEA, 2011).
In the past, most biofuel companies received government subsidies and tax reduction. These policies stimulated the rapid growth of bioenergy industry. Nowadays, some governments have started to withdraw this kind of support to biofuel producers. On one hand, the profit of biofuel production dropped significantly, and some companies had to shut down or reduce their activities. On the other hand, this change enables only the highly efficient, highly competitive technologies to survive, and therefore increases the competitiveness of the whole biofuel industry in a nonprotective energy market.
The next 5–10 years will be a crucial period for the development of bioenergy technologies. Most attention has been put in the industrialization progress of lignocellulosic bioethanol production. Various life-cycle assessments have demonstrated that utilizing lignocellulosic biomass for bioethanol fermentation would lead to a significant reduction of GHG. Identification of effective pretreatment methods, production of low-cost cellulolytic enzymes and enhanced fermentation yield and productivity using the hydrolysate from biomass into bioethanol are the most feasible fields to advance the technology.
Alternatively, lignocellulosic raw materials could be used for biofuel and biochemical production via thermochemical processes, such as fast pyrolysis, catalytic pyrolysis, and gasification. Fast pyrolysis actually has been commercialized, and the main product is bio-oil that can be readily stored, transported, and used for the production of liquid fuels and various chemicals (Bridgwater, 2012). Bio-oils have been successfully tested as fuels in engines, turbines, and boilers, and upgraded to high-quality hydrocarbon fuels (Czernik and Bridgwater, 2004). However, upgrading of the bio-oils to a quality of transport liquid fuel still faces several technical challenges due to the very complex compositions, and the process is not currently economically feasible. Catalytic pyrolysis refers to the pyrolysis process of biomass using various catalysts with aims of elimination and substitution of oxygen and oxygen-containing functionalities, in addition to increasing the hydrogen to carbon ratio of the final products (Dickerson and Soria, 2013). However, robust and highly-selective catalysts have to be further developed, and the cost of the process has to be reduced for commercial application. Gasification is one of the most promising technologies to produce gas fuels from lignocellulosic biomass. It is a thermochemical partial-oxidation process converting carbonaceous substances such as biomass into gas in the presence of a gasifying agent such as air, steam, oxygen, CO2, or a mixture of these. Syngas (synthesis gas) is the main product generated by biomass gasification, which consists mainly of H2, CO, CO2, N2, small particles of char, ashes, tars, and oils. However, in most markets, biomass gasification has yet to become consolidated as a mature technology to compete with other methods of energy conversion (Ruiz et al., 2013).
Utilizing various waste materials and by-products for biofuel and biochemical production not only reduced the burden of waste treatment but also provided an alternative way to generate green fuels and chemicals. Such waste feedstocks include various organic wastewater and residues from food processing plants, pulp mills, sugar mills, ethanol or biodiesel plants, and other biorefinery plants. The main components of the waste materials including starch, sugars, glycerol, etc. could be used as carbon sources of various microorganisms for producing bioethanol, biochemicals, and biodiesel feedstocks such as microbial lipids. Sugarcane molasses is a by-product of sugar processing, and has been successfully used for bioethanol production (Dasgupta et al., 2014). Organic effluents from different plants could be well-converted to intracellular lipid by oleaginous microorganisms, which can be used as a feedstock for biodiesel production (Marjakangas et al., 2015; Sun et al., 2015). The by-product glycerol from biodiesel production has many applications for producing chemicals and intermediates. A promising way to utilize this glycerol is to produce 1,3-propanediol, a monomer for producing polytrimethylene terephthalate (PTT). Actually, biological conversion of biodiesel by-product glycerol to 1,3-propanediol has been successfully industrialized in China (Liu et al., 2010). However, the efficiency of conversion of various waste materials to fuels and chemical needs yet to be enhanced. The impurities and inhibitors present in the waste might exert inhibition to the enzymes and microorganisms during the biological conversion. The economic feasibility of the processes still needs comprehensive evaluations.
Marine biomass, including microalgae and macroalgae (seaweeds), would still be one of the centers of bioenergy research. The potential of microalgae for the biodiesel production has been well-recognized. However, the high energy input in microalgae cultivation and microalgae processing limited its application in biofuel production. One of the most important challenges for autotrophic microalgae cultivation is the low growth rate, biomass density, and oil content. Another challenge refers to the high energy consumption for oil extraction because most microalgae has rigid cell wall structure and it is usually energy-intensive to disrupt the wall and release the intracellular oils. Therefore, currently the production of algal oil is primarily confined to high-value specialty oils with nutritional value such as polyunsaturated fatty acid, rather than commodity oils for biofuels (Hu et al., 2008). Modification of the microalgae by genetic engineering might improve the efficacy of CO2 to oil conversion and increase biomass density. However, more work should be done to extract the intracellular oils in a lower cost and increase the economic competiveness of the microalgae oil-based biofuel system.
The significant progress of the bioenergy industry encourages further exploration on low-carbon technologies for the production of advanced-generation biofuels (and biochemicals) from low-value waste biomass. Collective efforts from various aspects surrounding bioenergy technologies, including politicians, economists, environmentalists, scientists, and engineers, are needed to come up with alternatives, policies, and choices to advance the key technologies for a more sustainable future.

Acknowledgment

C.S.K. Lin, R. Luque, and J. Clark gratefully acknowledge the contribution of the COST Action TD1203-EUBis.

References

IEA (Ed.), 2011. Biofuels Can Provide up to 27% of World Transportation Fuel by 2050. IEA, Washington DC, USA. https://www.iea.org/newsroomandevents/pressreleases/2011/april/biofuels-can-provide-up-to-27-of-world-transportation-fuel-by-2050-iea-report-.html.

Bridgwater A.V. Review of fast pyrolysis of biomass and product upgrading. Biomass and Bioenergy. 2012;38:68.

Biofuel Statistics: Year 7 (Renewable transport fuel obligation statistics: year 7, report 1 data tables), UK government, https://www.gov.uk/government/statistics/biofuel-statistics-year-7-2014-to-2015-report-1 (published 06.11.14., last accessed 06.03.16).

BP Strategic Report, 2015. http://www.bp.com/content/dam/bp/pdf/investors/bp-strategic-report-2015.pdf (last accessed 06.03.16.).

Clark J.H, Deswarte F.E.I. The biorefinery concept – an integrated approach. In: Clark J.H, Deswarte F.E.I, eds. Introduction to Chemicals from Biomass. Chichester, West Sussex: Wiley; 2008.

Czernik S, Bridgwater A.V. Overview of applications of biomass fast pyrolysis oil. Energy and Fuels. 2004;18:590.

Dasgupta D, Ghosh P, Ghosh D, Suman S, Khan R, Agrawal D, et al. Ethanol fermentation from molasses at high temperature by thermotolerant yeast Kluyveromyces sp. IIPE453 and energy assessment for recovery. Bioprocess and Biosystems Engineering. 2014;37:2019.

Deng Y, Xu J, Liu Y, Mancl K. Biogas as a sustainable energy source in China: regional development strategy application and decision making. Renewable and Sustainable Energy Reviews. 2014;35:294.

Dickerson T, Soria J. Catalytic fast pyrolysis: a review. Energies. 2013;6:514.

Dunlap R.A. Sustainable Energy. SI ed. Stamford, Connecticut, USA: Cengage Learning; 2015.

EIA. Petroleum statistics. In: U.S. Energy Information Administration, pp. Oil: Crude and Petroleum Products Explained. 2014. http://www.eia.gov/energyexplained/index.cfm?page=oil_home#tab3.

Hu Q, Sommerfeld M, Jarvis E, Ghirardi M, Posewitz M, Seibert M, et al. Microalgal triacylglycerols as feedstocks for biofuel production: perspectives and advances. The Plant Journal. 2008;54:621.

IPCC. Summary for policymakers. In: Field C.B, Barros V.R, Dokken D.J, Mach K.J, Mastrandrea M.D, Bilir T.E, et al., eds. Climate Change 2014: Impacts, Adaptation, and Vulnerability Part A: Global and Sectoral Aspects Fifth Assessment Report of the Intergovernmental Panel on Climate Change Ed. Cambridge, UK and New York, USA: IPCC; 2014:1.

Koutinas A.A, Vlysidis A, Pleissner D, Kopsahelis N, Lopez Garcia I, Kookos I.K, et al. Valorization of industrial waste and by-product streams via fermentation for the production of chemicals and biopolymers. Chemical Society Reviews. 2014;43:2587.

Liu H, Xu Y, Zheng Z, Liu D. 1,3-Propanediol and its copolymers: research, development and industrialization. Biotechnology Journal. 2010;5:1137.

Marjakangas J.M, Lakaniemi A.-M, Koskinen P.E.P, Chang J.-S, Puhakka J.A. Lipid production by eukaryotic microorganisms isolated from palm oil mill effluent. Biochemical Engineering Journal. 2015;99:48.

Ruiz J.A, Juárez M.C, Morales M.P, Muñoz P, Mendívil M.A. Biomass gasification for electricity generation: review of current technology barriers. Renewable and Sustainable Energy Reviews. 2013;18:174.

Sun Q, Li A, Li M, Hou B, Shi W. Advantageous production of biodiesel from activated sludge fed with glucose-based wastewater. Acta Scientiae Circumstantiae. 2015;35:819.

Sadhukhan J, Ng K.S, Hernandez E.M. Biorefineries and Chemical Processes: Design, Integration and Sustainability Analysis. first ed. Malaysia: John Wiley & Sons; 2014.

UNFCCC, ed. Doha Amendment to the Kyoto Protocol. 2015 Doha.

Wellinger A. Biogas: Simply the Best. Brussels, Belgium: European Biogas Association; 2011.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset