Você está na página 1de 28

This article was downloaded by: [189.37.64.

11] On: 25 March 2012, At: 10:48 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

International Journal of Green Energy


Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/ljge20

Global Trends on the Processing of Biofuels


Mustafa Balat
a a

Sila Science, Trabzon, Turkey

Available online: 14 Jun 2008

To cite this article: Mustafa Balat (2008): Global Trends on the Processing of Bio-fuels, International Journal of Green Energy, 5:3, 212-238 To link to this article: http://dx.doi.org/10.1080/15435070802107322

PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

International Journal of Green Energy, 5: 212238, 2008 Copyright Taylor & Francis Group, LLC ISSN: 1543-5075 print / 1543-5083 online DOI: 10.1080/15435070802107322

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS Mustafa Balat


Sila Science, Trabzon, Turkey
The aim of the present paper is to investigate bio-fuels produced from biomass materials by thermochemical and biochemical methods and the utilization trends of the products in the world. Bio-fuels are liquid or gaseous fuels made from plant matter and residues, such as agricultural crops, municipal wastes and agricultural and forestry by-products. Liquid biofuels being considered world over fall into the following categories: (a) vegetable oils and biodiesels; (b) alcohols; and (c) biocrude and synthetic oils. Bioethanol can be produced from cellulose feedstocks such as corn stalks, rice straw, sugar cane bagasse, pulpwood, switchgrass, and municipal solid waste. 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 demonstrated commercially. Biodiesel fuel can be made from new or used vegetable oils and animal fats, which are non-toxic, biodegradable, renewable resources.The problems with substituting triglycerides for diesel fuels are mostly associated with their high viscosities, low volatilities and polyunsaturated character. Different ways have been considered to reduce the high viscosity of vegetable oils. Keywords: Bio-fuels; Global scenarios; Global production; Production processes

Downloaded by [189.37.64.11] at 10:48 25 March 2012

INTRODUCTION The majority of the worlds energy needs are supplied through petrochemical sources, such as coal and natural gases. With the exception of hydroelectricity and nuclear energy, all of these sources are finite and at current usage rates will be consumed shortly. The high energy demand in the industrialized world as well as in the domestic sector, and pollution problems caused due to the widespread use of fossil fuels make it increasingly necessary to develop renewable energy sources of limitless duration and smaller environmental impact than our traditional sources. This has stimulated recent interest in alternative sources for petroleum-based fuels (Meher et al., 2006). Petroleum based fuels became the primary source of energy for transportation needs in the 20th century. This has continued in the beginning of the 21st century with almost all vehicles running on gasoline, diesel or natural gas (Loppacher and Kerr, 2005). As time passed, the oil industry invested in infrastructure to enable long-distance diffusion of easily exploitable and thus cheaper petroleum resources, thereby halting the development of biofuels as alternatives to mineral oil-based petrol and diesel. After the oil crises of the

Address correspondence to Mustafa Balat, H. Osman Yucesan Cad. Zambak Sok. Polatoglu Ap. Kat 6, Besikduzu, Trabzon, Turkey. E-mail: mustafabalat@yahoo.com 212

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

213

Downloaded by [189.37.64.11] at 10:48 25 March 2012

1970s, bio-fuels briefly gained new momentum, but interest declined when oil prices returned to lower levels (Eikeland, 2006). The bio-fuels to be considered as relevant technologies by both developing and industrialized countries are due to a number of factors, including energy security reasons, environmental concerns, foreign exchange savings, and socioeconomic issues related to the rural sector. Increasing use of bio-fuels for energy generation purposes is of particular interest nowadays because they allow mitigation of greenhouse gases, provide means of energy independence, and may even offer new employment possibilities (Wierzbicka et al., 2005). Bio-fuels are thought to be the best hope in these and Powers are also ecofriendly. These fuels are non-toxic, biodegradable, and free of sulphur and carcinogenic compounds like benzene (Sastry et al., 2006). Bio-fuels are being investigated as potential substitutes for current high pollutant fuels obtained from conventional sources (Nwafor, 2004). Bio-fuels are liquid or gaseous fuels made from plant matter and residues, such as agricultural crops, municipal wastes and agricultural and forestry by-products. Liquid biofuels being considered the world over fall into the following categories: (a) vegetable oils and biodiesels; (b) alcohols; and (c) biocrude and synthetic oils. Liquid bio-fuels can be used as an alternative fuel for transport, as can other alternatives such as liquid natural gas (LNG), compressed natural gas (CNG), liquefied petroleum gas (LPG) and hydrogen. Figure 1 shows the resources of main liquid bio-fuels for automotives. Bioethanol and biodiesel can be mixed with the petroleum products (gasoline and diesel); they are substituting for and can be burned in traditional combustion engines with virtually no modifications needed. At present, most bio-fuel consumed is burnt as a mixturebioethanol/gasoline mixtures most commonly have five to ten percent bioethanol and biodiesel/diesel mixtures most commonly have five to twenty percent biodiesel. If pure bio-fuel is used, small modifications to the vehicle may be necessary, such as changing the material used for fuel lines. These are, however, quite minor. In addition, the distribution of liquid bio-fuels can easily be accommodated by the existing infrastructure for petroleum fuel distribution and retailing (Loppacher and Kerr, 2005). Energy outputs from bioethanol produced using corn, switchgrass, and wood biomass were each less than the respective fossil energy inputs. The same was true for producing biodiesel using soybeans and sunflower; however, the energy cost for producing soybean biodiesel was only slightly negative compared with bioethanol production.

Figure 1 Resources of main liquid bio-fuels for automotives (Demirbas, 2006a).

214

BALAT

Findings in terms of energy outputs compared with the energy inputs were (Pimentel and Patzek, 2005):

Bioethanol production using corn grain required 29% more fossil energy than the Bioethanol production using switchgrass required 50% more fossil energy than the Bioethanol production using wood biomass required 57% more fossil energy than the Biodiesel production using soybean required 27% more fossil energy than the biodiesel
fuel produced (Note, the energy yield from soy oil per hectare is far lower than the bioethanol yield from corn), Biodiesel production using sunflower required 118% more fossil energy than the biodiesel fuel produced. bioethanol fuel produced, bioethanol fuel produced, bioethanol fuel produced;

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Advantages of bio-fuels are the following (Puppan, 2002): (a) bio-fuels are easily available from common biomass sources, (b) they represent a carbon dioxide (CO2) cycle in combustion, (c) bio-fuels have considerable environmentally friendly potential, (d) there are many benefits for the environment, economy and consumers in using bio-fuels, and (e) they are biodegradable and contribute to sustainability. GLOBAL BIOFUEL SCENARIOS Worldwide interest reflects convergence of developed and developing country perspectives on the role bio-fuels can play in energy for sustainable development (Gururaja, 2005). The increase in bio-fuels utilization has also been accompanied over the past 34 years with policy decisions that encourage future growth of these fuels (Demirbas and Balat, 2006). The United States and several European Union (EU) member states already pursue successful bio-fuel policy (Puppan, 2002). The recent commitment by the United States government to increase bio-energy three-fold in 10 years has added impetus to the search for viable biofuels (Demirbas, 2006a). Future conditions for an international bio-fuel market in Europe will largely be decided by the EU policies on renewable energy and their interplay with national energy policies. So far, the Commission has indicated that biomass will play an important role in the future (Ericsson and Nilsson, 2004). The most important piece of legislation for biofuels in Europe is the Biofuels Directive (Directive 2003/30/EC). Adopted in May 2003, it aims to promote the use in transport of fuels made from biomass, as well as other renewable fuels. The directive sets a reference value of 5.75% for the market share of bio-fuels in 2010, measured in terms of energy content (Asher, 2006). In the Commissions view, mandating the use of bio-fuels will (a) improve energy supply security and (b) reduce greenhouse gas (GHG) emissions and (c) boost rural incomes and employment. Current regulations would preclude a notable negative impact on the rural environment (Jansen, 2003). In South America, Brazil also continued policies that mandate at least 22% bioethanol on motor fuels and encourage the use of vehicles that use hydrous ethanol to replace gasoline (Stevens et al., 2004). Figure 2 shows the shares of alternative fuels compared to the total automotive fuel consumption in the world as a futuristic view. Hydrogen is currently more expensive than conventional energy sources. There are different technologies presently being practiced to

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

215

Biofuels 20 Alternative fuel consunption, % 18 16 14 12 10 8 6 4 2 0 2000 2010

Natural gas

Hydrogen

Downloaded by [189.37.64.11] at 10:48 25 March 2012

2020 Years

2030

2040

2050

Figure 2 Shares of alternative fuels compared to the total automotive fuel consumption in the world (Demirbas, 2006a, 2007).

produce hydrogen economically from biomass. Biohydrogen technology will play a major role in the future because it can utilize the renewable sources of energy (Demirbas, 2006a). GLOBAL TRENDS ON THE BIO-FUEL PRODUCTION Brazil and the United States have the largest programs promoting bio-fuels in the world. The EU ranks third in bio-fuel production worldwide, behind Brazil and the United States (Demirbas and Balat, 2006). Bio-fuel production in the EU was concentrated in countries hosting the major transportation fuel markets, notably Germany, France, Italy and Spain, with substantial volumes produced also in the Czech Republic, Austria, Poland, Sweden and Denmark (Eikeland, 2006). Bioethanol Production Bioethanol is by far the most widely used bio-fuel for transportation worldwide. Global bioethanol production more than doubled between 2000 and 2005. Figure 3 shows global bioethanol production between 20002005. About 60% of global bioethanol production comes from sugarcane and 40% from other crops (Dufey, 2006). According to the United States Department of Agricultures 2006 figures (USDA, 2006), the United States produced 4 billion gallons (15.2 billion liters) of ethanol in 2005, up from 3.4 billion gallons in 2004. In 2005, Brazil, produced 4.2 billion gallons of ethanol, up from 4.0 billion gallons in 2004. The United States is predominantly a producer of bioethanol derived from corn, and production is concentrated in Midwestern states with abundant corn supplies. In 2005 the United States produced 4 billion gallons of bioethanol, which equates to about 3% of the country's total gasoline consumption (140 billion gallons per year) (Asher, 2006). Prompted by the increase in oil prices, Brazil began to produce bioethanol from sugarcane in the 1970s and is considered the most successful example of a commercial application of biomass for energy production and use. Extensive

216

BALAT

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Figure 3 Global bioethanol production between 20002005.

experience in bioethanol production, suitable natural conditions for sugarcane production and low labour costs have made Brazil the most efficient bioethanol-producing country (Dufey, 2006). Bioethanol represents approximately 1/3 of total vehicle fuels currently used in Brazil (Eikeland, 2006). The EU accounts for about 10% of global bioethanol production. EU bioethanol is generally produced using a combination of sugar beets and wheat (Schnepf, 2006). France is currently the front-runner in the EUs attempt to boost bioethanol use, accounting for 2% of global production, mainly from sugar beet and wheat. However, France is rapidly being overtaken by Germany and Spain as the EUs largest bioethanol producer. China accounts for about 9% of global bioethanol production, 80% of which is grain-basedmainly derived from corn, cassava and rice. India accounts for 4% of global bioethanol production. This is made from sugarcane (Dufey, 2006). With all of the new government programs in America, Asia, and Europe in place, total worldwide fuel bioethanol demand could grow to exceed 125 billion liters by 2020 (Bohlmann, 2006). Biodiesel Production Biodiesel production is highest in Europe, where more biodiesel is produced than ethanol, but total production of both fuels is fairly small compared to the production of bioethanol in Brazil and the United States. The main biodiesel-producing countries are Germany, France, and Italy, where the fuel is used as a diesel blend (5% or 20%) (Boyle, 2005). Biodiesel accounted for nearly 80% of EU bio-fuel production (Schnepf, 2006). Figure 4 shows global biodiesel production between 20002005.

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

217

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Figure 4 Global biodiesel production between 20002005.

According to the European Biodiesel Boards 2006 figures (EBB, 2006), EU-25 produced 3.2 million tons (3.3 billion liters) of biodiesel in 2005, up from 1.9 million tons in 2004. This represents an unprecedented 65% yearly growth for EU biodiesel production. Germany produced an estimated 1,669,000 tons in 2005, France produced 492,000 tons, and Italy produced 396,000 tons (EBB, 2006). In Germany, biodiesel is commonly sold neat (100%) (Boyle, 2005). BIOMASS CONVERSION PROCESSES TO BIO-FUELS Bio-fuels are made from biomass through thermochemical processes such as pyrolysis, gasification, liquefaction and supercritical fluid extraction or biochemical. Biochemical conversion of biomass is completed through alcoholic fermentation to produce liquid fuels and anaerobic digestion or fermentation, resulting in biogas (Demirbas, 2004a). Figure 5 shows the main biomass conversion processes. Nearly all bioethanol fuel is produced by fermentation of corn glucose in the United States or sucrose in Brazil, but any country with a significant agronomic-based economy can use current technology for bioethanol fermentation. This is possible because during the last two decades, technology for bioethanol production from nonfood-plant sources has been developed to the point at which large-scale production will be a reality in the next few years. Therefore, agronomic residues such as corn stover (corn cobs and stalks), sugarcane waste, wheat or rice straw, forestry and paper mill discards, the paper portion of municipal waste and dedicated energy cropscollectively termed biomasscan be converted into bioethanol fuel (Lin and Tanaka, 2006). Ethanol is most commonly used to increase octane and improve the emissions quality of gasoline. Ethanol is blended with

218

BALAT

Figure 5 Main biomass conversion processes.

Downloaded by [189.37.64.11] at 10:48 25 March 2012

gasoline to form an E10 blend (10% ethanol and 90% gasoline), but it can be used in higher concentrations such as E85 or E95. Original equipment manufacturers produce flexible-fuel vehicles that can run on E85 or any other combination of ethanol and gasoline (Balat, 2005a). Bioethanol has a higher octane number (107), broader flammability limits, higher flame speeds and higher heats of vaporization than gasoline. These properties allow for a higher compression ratio, shorter burn time and leaner burn engine, which lead to theoretical efficiency advantages over gasoline in an ICE. Table 1 shows some properties of alcohol fuels. Exploring new energy resources, such as biodiesel fuel, is of growing importance in recent years (Zhang et al., 2003). Biodiesel is often used as a blend B20 (20 vol.% biodiesel and 80 vol.% conventional diesel) rather than using B100 (100 vol.% biodiesel). It is asserted that 90% of air toxics can be eliminated by using B100 whereas 2040% are reduced using B20 (Joshi and Pegg, 2007). The properties of biodiesel and diesel fuels is given in Table 2. Among them, biodiesel produced from different vegetable oils (soybean, rapeseed and sunflower for example) seems very interesting for several reasons: it can replace diesel oil in boilers and internal combustion engines without major adjustments; only a small decrease in performances is reported; almost zero emissions of sulfates; a small net contribution of CO2 when the whole life-cycle is considered (including cultivation, production of oil and conversion to biodiesel); emission of pollutants comparable with that of diesel oil (Carraretto et al., 2004).

Table 1 Some properties of alcohol fuels. Fuel property Cetane number Octane number Auto-ignition temperature (K) Latent heat of vaporization (MJ/Kg) Lower heating value (MJ/Kg) Isoctane 100 530 0.26 44.4 Methanol 5 112 737 1.18 19.9 Ethanol 8 107 606 0.91 26.7

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS Table 2 Physical properties of biodisel and No 2. Diesel fuels. Property Specific gravity, kg/L Cetane number Cloud point, K Pour point, K Flasf point, K Sulfur, wt% Ash, wt% Iodine number Kinematic viscosity, 313 K Higher heating value, MJ/kg Biodiesel 0.87 to 0.89 46 to 70 262 to 289 258 to 286 408 to 423 0.0000 to 0.0024 0.002 to 0.01 60 to 135 3.7 to 5.8 39.3 to 39.8 No 2. Diesel 0.84 to 0.86 47 to 55 256 to 265 237 to 243 325 to 350 0.04 to 0.01 0.06 to 0.01 1.9 to 3.8 45.3 to 46.7

219

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Bioethanol from Biomass Bioethanol can be produced from different kinds of raw materials. Table 3 shows the bioethanol pathways from different raw materials. Bioethanol is used as a partial gasoline replacement. The raw materials are classified into three categories of agricultural raw materials: simple sugars, starch and lignocellulose. Figure 6 shows the flow chart for the production of bioethanol from lignocellulosic biomass materials. A large amount of ethanol can be produced from ethylene (a petroleum product). Catalytic hydration of ethylene produces synthetic ethanol.

C2 H 4 + H 2 O C2 H 5 OH Ethylene Steam Ethanol

(1)

The components of lignocellosic biomass include cellulose, hemicelluloses, lignin, extractives, ash, and other compounds. Cellulose is a remarkable pure organic polymer, consisting solely of units of anhydro glocose held together in a giant straight chain molecule. Bioethanol can be produced from cellulose feedstocks such as corn stalks, rice straw, sugar cane bagasse, pulpwood, switchgrass, and municipal solid waste. Cellulose must be

Table 3 Bioethanol patways from different raw materials. Raw material Wood Wood Straw Straw Wheat Sugar cane Sugar beet Corn grain Corn stalk Sweet sorghum Processing Acid hydrolysis + fermentation Enzymatic hydrolysis + fermentation Acid hydrolysis + fermentation Enzymatic hydrolysis + fermentation Malting + fermentation Fermentation Fermentation Fermentation Acid hydrolysis + fermentation Fermentation

Source: Sun and Cheng (2005).

220

BALAT

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Figure 6 Flow chart for the production of bioethanol from lignocellulosic biomass materials (Demirbas, 2006a).

hydrolyzed to glucose before fermentation to bioethanol. Hemicelluloses (arabinoglycuronoxylan and galactoglucomammans) are related to plant gums in composition and occur in much shorter molecule chains than cellulose. Hemicelluloses are derived mainly from chains of pentose sugars, and act as the cement material holding together the cellulose micells and fiber (Demirbas, 2007). Lignins are polymers of aromatic compounds. Their functions are to provide structural strength, provide sealing of water conducting system that links roots with leaves, and protect plants against degradation. Lignin is a macromolecule, which consists of alkylphenols and has a complex three-dimensional structure. Lignin is covalently linked with xylans in the case of hardwoods and with galactoglucomannans in softwoods. Even though mechanically cleavable to a relatively low molecular weight, lignin is not soluble in water. It is generally accepted that free phenoxyl radicals are formed by thermal decomposition of lignin above 525 K and that the radicals have a random tendency to form a solid residue through condensation or repolymerization (Demirbas, 2006a). Conversion of biomass to bioethanol is difficult due to: (1) the resistant nature of biomass to breakdown; (2) the variety of sugars that are released when the hemicellulose and cellulose polymers are broken and the need to find or genetically engineer organisms to efficiently ferment these sugars; (3) costs for collection and storage of low density biomass feedstocks.

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

221

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Bioethanol is produced by a process known as fermentation. In 1995, about 93% of the ethanol in the world was produced by the fermentation method and about 7% by the synthetic method. The fermentation method generally uses three steps: (1) the formation of a solution of fermentable sugars, (2) the fermentation of these sugars to ethanol, and (3) the separation and purification of the ethanol, usually by distillation. Fermentation involves microorganisms that use the fermentable sugars for food and in the process produces ethanol and other byproducts. These microorganisms can typically use the 6-carbon sugars, one of the most common being glucose. Therefore, lignocellulosic biomass materials containing high levels of glucose or precursors to glucose are the easiest to convert to ethanol. Microorganisms, termed ethanologens, presently convert an inadequate portion of the sugars from biomass to ethanol. Although fungi, bacteria, and yeast microorganisms can be used for fermentation, specific yeast (Saccharomyces cerevisiae also known as Bakers yeast) is frequently used to ferment glucose to bioethanol. Theoretically, 100 grams of glucose can produce 51.3 g of bioethanol and 48.7 g of CO2. However, in practice, the microorganisms use some of the glucose for growth and the actual yield is less than 100% (Badger, 2002). Cellulose can be broken down into glucose, a 6-carbon sugars, and hemicellulose into 5-carbon sugars. 5-carbon sugars are more difficult to convert into bioethanol than glucose. Lignin contains no sugars and encloses cellulose and hemicellulose, making it difficult to extract the latter two. There are three pathways for bioethanol production from cellulosic feedstock: acid hydrolysis, enzymatic hydrolysis, and biomass gasification, all followed by fermentation (Kojima and Johnson, 2005). In this case, the primary stoichiometric equations for the bioethanol production are as follows (Demirbas, 2005a): Pentosan to pentose:

n C4 H g O4 1g
Hexosan to hexose:

n H2 O 0.136 g

n C5 H10 O5 1.136 g

(2)

n C6 H10 O5 + n H 2 O 1g 0.111g

n C6 H12 O6 1.111g

(3)

Pentose and hexose to bioethanol, 0.511 grams per gram hexose or pentose:

3 C5 H10 O5 5 C2 H 5 OH 1g C6 H12 O6 1g 0.511 g 2 C2 H 5 OH 0.511 g

5 CO2 0.489 g

(4)

2 CO2 0.489 g

(5)

The most commonly used acid in acid hydrolysis is sulfuric acid. There are two types of processes: dilute acid and concentrated acid hydrolysis. Dilute acid hydrolysis occurs at higher temperature and pressure, and the sugar recovery efficiency is limited to

222

BALAT Table 4 Yields of bioethanol by concentrated sulfuric acid hydrolysis from cornstalks (% dry weight). Amount of cornstalk (kg) Cellulose content (kg) Cellulose conversion and recovery efficiency Ethanol stoichiometric yield Glucose fermentation efficiency Ethanol yield from glucose (kg) Amount of cornstalk (kg) Hemicelluloses content (kg) Hemicelluloses conversion and recovery efficiency Ethanol stoichiometric yield Xylose fermentation efficiency Ethanol yield from xylose (kg) Total ethanol yield from 1000 kg of cornstalks Source: Demirbas (2005a). 1000 430 0.76 0.51 0.75 130 1000 290 0.90 0.51 0.50 66 196 kg (225.7 L = 59 gallons)

Downloaded by [189.37.64.11] at 10:48 25 March 2012

about 50%, because the product sugar reacts further to form other chemicals under the same reaction conditions (Kojima and Johnson, 2005). Table 4 shows the yields of bioethanol by concentrated sulfuric acid hydrolysis from cornstalks. Biodiesel from Biomass Biodiesel, an alternative diesel fuel, is made from renewable biological sources such as vegetable oils and animal fats (Ma and Hanna, 1999). The basic constituent of vegetable oils is triglyceride. The problems with substituting triglycerides for diesel fuels are mostly associated with their high viscosities, low volatilities and polyunsaturated character (Srivastava and Prasad, 2000). Different ways have been considered to reduce the high viscosity of vegetable oils: (a) dilution, (b) microemulsions, (c) pyrolysis, (d) catalytic cracking and (e) transesterification (Demirbas, 2006b). The pyrolysis refers to chemical change caused by the application of thermal energy in the presence of an air or nitrogen sparge (Fukuda et al., 2001). Since then, several studies on vegetable oil pyrolysis as an alternative method to obtain chemicals and fuels have been reported in the literature (Chand and Wan, 1947; Grossley et al., 1962; Alencar et al., 1983; Pioch et al., 1993; Billaud et al., 1995; Schwab et al., 1988; Demirbas, 2002; Lima et al., 2004). Thermal decomposition of triglycerides produces the compounds of classes including alkanes, alkenes, alkadienes, aromatics and carboxylic acids. Different types of

Table 5 Viscosities of vegetable oils, methyl esters and ethyl esters from the vegetable oils by transeste-rificaion (measured at 311 K as mm2/s). Test sample Cottonseed Hazelnut kernel Poppyseed 42.4 3.5 5.1 Rapeseed 37.3 3.3 4.7 Safflowerseed 31.6 2.9 4.2 Sunflowerseed 34.4 3.2 4.6

Vegetable oil 33.7 24.0 Methyl ester 3.1 2.8 Ethyl ester 4.8 3.9 Viscosity of No. 2 diesel : 2.7 mm2/s Source: Balat (2005b).

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

223

vegetable oils produce large differences in composition when they are thermally decomposed. Pyrolyzed soybean oil, for instance, contains 79% carbon and 12% hydrogen (Fukuda et al., 2001). Transesterification is not a new process. The scientists Duffy and Patrick conducted it as early as 1853 (Stavarache, et al., 2005). The purpose of the transesterification of vegetable oils to their methyl esters (biodiesels) process is to lower the viscosity of the oil. The main factors affecting transesterification are molar ratio of glycerides to alcohol, catalyst, reaction temperature and pressure, reaction time and the contents of free fatty acids and water in oils. The commonly accepted molar ratios of alcohol to glycerides are 6:130:1 (Demirbas, 2003). Transesterification (also called alcoholysis) is the reaction of a fat or oil with an alcohol to form esters and glycerol. The reaction is shown in Eq. (7). A catalyst is usually used to improve the reaction rate and yield. Because the reaction is reversible, excess alcohol is used to shift the equilibrium to the products side (Ma and Hanna, 1999).

Downloaded by [189.37.64.11] at 10:48 25 March 2012

CH COOR I I CH COOR I CH COOR III Triglyceride Methanol


II

CH 2 OH + 3ROH
Catalyst

R I COOR + R COOR + R III COOR Biodiesel


II

I CHOH I CH 2 OH Glycerol

(6)

The overall process is a sequence of three consecutive and reversible reactions, in which diand monoglycerides are formed as intermediates. The stoichiometric reaction requires 1 mol of a triglyceride and 3 mol of the alcohol. However, an excess of the alcohol is used to increase the yields of the alkyl esters and to allow its phase separation from the glycerol formed (Schuchardt et al., 1998). The alcohols employed in the transesterification are generally short chain alcohols such as methanol, ethanol, propanol, and butanol. It was reported that when transesterification of soybean oil using methanol, ethanol and butanol was performed, 9698% of ester could be obtained after an hour of reaction (Dmytryshyn, et al., 2004). Catalysts used for the transesterification of triglycerides are classified as alkali, acid, enzyme or heterogeneous catalysts, among which alkali catalysts like sodium hydroxide, sodium methoxide, potassium hydroxide, potassium methoxide are more effective (Meher et al., 2006). Enzymes-catalyzed procedures, using lipase as catalyst, do not produce side reactions, but the lipases are very expensive for industrial scale production and a threestep process was required to achieve a 95% conversion. The transesterification process is catalyzed by phosphoric, hydrochloric, sulfuric and organic sulfonic acids. Acid-catalyzed process is useful when a high amount of free acids are present in the vegetable oil, but the reaction time is very long (4896 h), even at the boiling point of the alcohol, and a high molar ratio of alcohol was needed (20:1 wt/wt to the oil) (Stavarache, et al., 2005). The alkaline catalysts show high performance for obtaining vegetable oils with high quality, but a question often arises; that is, the oils contain significant amounts of free fatty acids which cannot be converted into biodiesels but to a lot of soap (Furuta et al., 2004). It often takes at least several hours to ensure the alkali (NaOH or KOH) catalytic

224

BALAT

Downloaded by [189.37.64.11] at 10:48 25 March 2012

transesterification reaction is complete. Moreover, removal of these catalysts is technically difficult and brings extra cost to the final product (Demirbas, 2003). Alkali-catalyzed transesterification is much faster than acid-catalyzed transesterification and is most often used commercially (Ma and Hanna, 1999). Triglycerides are readily transesterified batchwise in the presence of alkaline catalyst at atmospheric pressure and at a temperature of approximately 333 to 343 K with an excess of methanol. The mixture at the end of reaction is allowed to settle. The lower glycerine layer is drawn off while the upper methyl ester layer is washed to remove entrained glycerine and is then processed further. The excess methanol is recovered in the condenser, sent to a rectifying column for purification and recycled (Srivastava and Prasad, 2000). An alternative way of processing these vegetable oils is to use an acid catalyst. Three types of solid superacid catalysts were prepared and evaluated in the transesterification of soybean oil with methanol and the esterification of n-octanoic acid with methanol. Tungstated zirconia- alumina is a promising catalyst for the production of biodiesel fuels because of its activity for the transesterification as well as the esterification. Solid superacid catalysts of sulfated tin and zirconium oxides and tungstated zirconia were used in the transesterification of soybean oil with methanol at 475575 K and the esterification of n-octanoic acid with methanol at 450475 K (Furuta et al., 2004; Bala, 2005). Alkalicatalyzed transesterification is much faster than acid-catalyzed transesterification and is most often used commercially (Ma and Hanna, 1999). Water content is an important factor in the acid- and alkaline-catalyzed transesterification of vegetable oil. Figure 7 shows a direct comparison of the yield of methyl esters from various preparation methods as triglycerides with various water contents are

Figure 7 Yields of methyl esters as a function of water content in transesterification of triglycerides. Legend: (1) supercritical methanol; (2) alkaline-catalyzed; (3) acid-catalyzed (Kusdiana and Saka, 2004).

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

225

Downloaded by [189.37.64.11] at 10:48 25 March 2012

treated. The alkalinecatalyzed method used 1.5% sodium hydroxide as a catalyst in methanol, while in the acidcatalyzed method, 3% sulfuric acid in methanol is added to the reaction system. For the alkaline-catalyzed method, the conversion is slightly reduced when more water is added. In the acid-catalyzed method, however, only as little as 0.1% of water added led to some reduction of the yield of methyl esters and the conversion is significantly reduced to 6% when only 5% of water is added (Kusdiana and Saka, 2004). A non-catalytic biodiesel production route with supercritical methanol has been developed that allows a simple process and high yield because of simultaneous transesterification of triglycerides and methyl esterification of fatty acids (Demirbas, 2002). A few studies have been conducted via non-catalytic transesterification with supercritical methanol (Krammer and Vogel, 2000; Dunn, 2001; Kusdiana and Saka, 2001; Saka and Kusdiana 2001; Demirbas, 2002, 2003, 2005b). Supercritical methanol transesterification has several notable advantages over the conventional processes. For reactions in supercritical methanol, no catalyst is required and nearly complete conversions can be achieved in a very short time (24 minutes). This is primarily because supercritical methanol and oil exist in a single phase. Because the reaction is non-catalytic and fast, few investigations have explored the synthesis of biodiesel by transesterification of vegetable oils in supercritical methanol (Madras et al., 2004). In the earlier study (Balat, 2005b), the yields of ethyl esters from vegetable oils via transesterification in supercritical ethanol was investigated. Figure 8 shows the changes in yield percentage of ethyl esters as treated with subcritical and supercritical ethanol at different temperatures as a function of reaction time. The critical temperature and the critical pressure of ethanol are 516.2 K and 6.4 MPa, respectively. Biohydrogen from Biomass Producing hydrogen from woody biomass are mainly carried out via two thermochemical processes: (a) gasification followed by reforming of the syngas, and (b) fast pyrolysis followed by reforming of the carbohydrate fraction of the bio-oil. Production of hydrogen from renewable biomass has several advantages compared to that of fossil fuels. This is especially true if the hydrogen is manufactured from renewable resources (Demirbas, 2004b). The main gaseous products from biomass are the following:

Pyrolysis of biomass H 2 + CO2 + CO + Hydrocarbon gases Catalytic steam reforming of biomass H2 + CO2 + CO Gasification of biomass H2 + CO2 + CO + N 2

(7) (8) (9)

Hydrogen can be produced from biomass by pyrolysis (Demirbas, 2001; Moitra, 2004), gassification (Wang et al., 1997; Midilli et al., 2001; Merida et al., 2004), steam gasification (Hanaoka et al., 2005; Demirbas, 2004c; 2006c) steam-reforming of bio-oils, and enzymatic decomposition of sugars. Hydrogen is produced from pyroligneous oils produced from the pyrolysis of lignocellulosic biomass (Demirbas and Caglar, 1998).

226

BALAT

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Figure 8 Changes in yield percentage of ethyl esters as treated with subcritical and supercritical ethanol at different temperatures as a function of reaction time. Molar ratio of vegetable oil to ethyl alcohol: 1:40 (Balat, 2005b).

Figure 9 shows a process diagram of hydrogen production from biomass. The yield of hydrogen that can be produced from biomass is relatively low, 1618% based on dry biomass weight (Demirbas, 2001). Hydrogen yields and energy contents, compared, with biomass energy contents obtained from processes with biomass, are shown in Table 6. The strategy is based on producing hydrogen from biomass pyrolysis using a coproduct strategy to reduce the cost of hydrogen and concluded that only this strategy could compete with the cost of the commercial hydrocarbon-based technologies (Wang et al., 1998). This strategy will demonstrate how hydrogen and bio-fuel are economically feasible and can foster the development of rural areas when practiced on a larger scale. The process of biomass to activated carbon is an alternative route to hydrogen with a valuable co-product that is practiced commercially. The yield of hydrogen that can be produced from biomass is relatively low, 1214% based on the biomass weight (Demirbas, 2005c). In the proposed second process, fast pyrolysis of biomass to generate bio-oil and catalytic steam reforming of the bio-oil to hydrogen and carbon dioxide. Thermochemical gasification of biomass has been identified as a possible system for producing renewable hydrogen, which is beneficial to exploit biomass resources, develop a highly efficient clean way for large-scale hydrogen production, and has less dependence on insecure fossil energy sources. Most of the research spurred by this interest has been of

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

227

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Figure 9 Process diagram of hydrogen production from biomass (Czernik et al., 2000).

Table 6 Comparison of hydrogen yields were obtained by use of three different processes. Processes Hydrogen yield (w) Hydrogen energy contents/biomass energy content 91 83 124

Pyrolysis + catalytic reforming Gasification + shift reaction Biomass + steam + except heat (theoretical maximum)

12.6 11.5 17.1

Source: Wang et al (1997); Balat and Ozdemir (2005).

economic technology in nature, based on gasifier performance data acquired during system proof of conceptual tests. Less emphasis has been given to experimental investigation of hydrogen production via biomass gasification. Till now, all process equipment needed to produce hydrogen is well established in commercial use, except for the gasifiers. Comparison with other biomass thermochemical gasification such as air gasification or steam gasification, the supercritical water gasification can directly deal with the wet biomass without drying, and have high gasification efficiency at lower temperature (Hao et al., 2003).

228

BALAT Table 7 Biomass and process characteristics. Input data C, H, O amount Range of values Biomass composition C: 51%wdaf H: 6%wdaf O: 43%wdaf S/Ba: from 0.3 to 1 kg/kg Up to 12 mm 10731273 K Atmospheric Flash (> 500 C s1) 0.110 s
m

Biomass particle size Temperature Pressure Heating rate Gas residence time
a

Downloaded by [189.37.64.11] at 10:48 25 March 2012

S steam B = dry biomass =

H 2 Ogas + m H 2 O liquid mdry biomass

Source: Dupont et al (2007).

Modelling of biomass steam gasification to synthesis gas is a challenge because of the variability (composition, structure, reactivity, physical properties, etc.) of the raw material and because of the severe conditions (temperature, residence time, heating rate, etc.) required. This is well-illustrated in a fluidized gasification system, as shown in Table 7 (Dupont et al., 2007). The yield from steam gasification increases with increasing water-to-sample ratio. The yields of hydrogen from the pyrolysis and the steam gasification increase with increasing temperature. In general, the gasification temperature is higher than that of pyrolysis and the yield of hydrogen from the gasification is higher than that of the pyrolysis. The highest yields (% dry and ash free basis) were obtained from the pyrolysis (46%) and steam gasification (55%) of wheat straw while the lowest yields were from olive waste. The yield of hydrogen from supercritical water extraction was considerably high (49% by volume) at lower temperatures. The pyrolysis was carried out at the moderate temperatures and steam gasification at the highest temperatures. The pyrolysisbased technology, in particular, because it has coproduct opportunities, has the most favorable economics (Demirbas, 2006c). It is believed that in the future biomass can become an important sustainable source of hydrogen. Biomass has the advantage of low environmental impact compared with that for fossil fuels. The price of hydrogen obtained by direct gasification of lignocellulosic biomass, however, is about three times higher than that for hydrogen produced by steam reforming of natural gas (Spath et al., 2000). Steam reforming of natural gas is an endothermic, catalytic process carried out at about 1125 K and around 2.5 MPa according to the following reactions:

CH 4 + H 2 O

CO + 3H 2

(10) (11)

CO + H 2 O CO2 + H 2
The carbon dioxide is removed by absorption or membrane separation.

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

229

Steam reforming of hydrocarbons, partial oxidation of heavy oil residues, selected steam reforming of aromatic compounds, and gasification of coals and solid wastes to yield a mixture of H2 and CO (syngas), followed by water-gas shift conversion to produce H2 and CO2, are well-established processes. When the objective is to maximize the production of H2, the stoichiometry describing the overall process is

Cn H m + 2nH 2 O nCO2 + [2n + (m/2)]H 2

(12)

The simplicity of Eq. (12) hides the fact that, in a hydrocarbon reformer, the following reactions take place concurrently:

Cn H m + nH 2 O nCO + [2n + (m/2)]H 2


Downloaded by [189.37.64.11] at 10:48 25 March 2012

(13)

Partial oxidation of hydrocarbons is the exothermic reaction with oxygen and steam. The amounts of oxygen and water vapor are controlled so that the reaction proceeds without the need for external energy. An example reaction for this process is:

2C8 H18 + 2H 2 O + 9O2 12CO + 4CO2 + 20H 2

(14)

Another conventional process is Plasma Arc Process. The plasma arc processing of natural gas or oil uses electricity to produce pure carbon and hydrogen at temperatures near 1875 K. A pilot plant utilizing this technology produced 500 kg/h of carbon and 2000 m3/h of hydrogen from 1000 m3/h of natural gas and 2100 KW of electricity (Balat, 2006). Biomethanol from Biomass Methanol is produced on a very large scale ( 35 106 metric tons/year). At present, about 30% is used for formaldehyde production and 30% for MTBE (Methyl tertiary butyl ether). Only a few percent is used as fuel. In the future, the role of methanol as liquid fuel may increase considerably, e.g. in fuel cell cars, as intermediate for hydrogen production, etc. (van Swaaij et al., 2004). Methanol is mainly manufactured from natural gas, but biomass can also be gasified to methanol. Table 8 shows main production facilities of methanol and biomethanol. Methanol can be produced from hydrogen-carbon oxide mixtures by means of the catalytic reaction of carbon monoxide and some carbon dioxide with hydrogen. Biosynthesis gas (bio-syngas) is a gas rich in CO and H2 obtained by gasification of biomass.

Table 8 Main production facilities of methanol and biomethanol. Methanol Catalytic synthesis from CO and H2 Natural gas Petroleum gas Distillation of liquid from coal pyrolysis Source: Demirbas (2007). Biomethanol Catalytic synthesis from CO and H2 Distillation of liquid from wood pyrolysis Gaseous products from biomass gasification Synthetic gas from biomass and coal

230

BALAT

The requirements for syngas production from biomass for the subsequent methanol synthesis are not fulfilled by conventional gasification processes. In contrast to gasification processes for electricity production, the syngas for the methanol generation process is limited by inert gas components (CH4, N2), which are not converted during methanol synthesis. A second reqirement for the syngas composition is a high hydrogen content, because a main part of the biomass carbon is converted to CO2 in the gasification step (CO2 needs 3 moles of H2 for hydrogenation to methanol). The preferable H2/CO ratio in the gasifier raw gas has to be > 2 (Specht et al., 1998). In this case, a shift reactor and therefore the additional appliances are not required. The gasification of biomass always results in a gas containing a too low hydrogen portion, respective of a too high carbon portion (CO2) for the methanol synthesis, even if the requirement mentioned above is fulfilled. Since the reaction of the methanol synthesis is described by two equations (Specht and Bandi, 1999):

Downloaded by [189.37.64.11] at 10:48 25 March 2012

CO + 2H 2 CH3 OH CO2 + 3H 2 CH3 OH + H 2 O

(15) (16)

A variety of catalysts are capable of causing the conversion, including reduced NiO-based preparations, reduced Cu/ZnO shift preparations, Cu/SiO2 and Pd/SiO2, and Pd/ZnO (Takezawa et al., 1987; Iwasa et al., 1993). Typical synthesis conditions are a pressure of 50100 bar and a temperature of 220300C using a Cu/Zn/Al catalyst. An ideal synthesis gas should have a ratio H2/(2CO + 3CO2) at about 1.05, and a low CO2 content of approximately 3%. If the synthesis gas is produced through the conventional reforming of natural gas, the hydrogen concentration is in excess and the syntheses gas is too rich in hydrogen. Synthesis gas for methanol can also be derived from gasification of (residual) oil, coal and biomass (van Swaaij et al., 2004). The gases produced can be steam reformed to produce hydrogen and followed by watergas shift reaction to further enhance hydrogen production. When the moisture content of biomass is higher than 35%, it can be gasified in a supercritical water condition. The gas is converted to methanol in a conventional steam-reforming/ water-gas shift reaction followed by high-pressure catalytic methanol synthesis (Balat, 2006):

CH 4 + H 2 O CO + 3H 2 CO + H 2 O CO2 + H 2

(10) (11)

Eqs. (10 and 11) are called as gasification/shift reactions. Figure 10 shows production of biomethanol from carbohydrates by gasification and partial oxidation with O2 and H2O. Fuels from Bio-Syngas Via Fisher-Tropsch Synthesis The Fischer-Tropsch Synthesis (FTS) produces hydrocarbons of different length from a gas mixture of H2 and CO (syngas) from biomass gasification called bio-syngas (Prins et al., 2004). The FTS is a process by which gasoline, diesel oil, wax, and alcohols

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

231

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Figure 10 Biomethanol from carbohydrates by gasification and partial oxidation with O2 and H2O (Demirbas, 2007).

are produced from bio-syngas (Akgerman, 1984). The reactions of the FTS on iron catalysts can be simplified as a combination of the Fischer-Tropsch reaction and the water gas shift (WGS) reaction (van der Laan, 1999):

nCO + [2n + (m/2)]H 2 Cn H m + nH 2 O (FF ) H FT = 165 kJ/mol CO + H 2 O CO2 + H 2 (WGS ) HWGS = 41.3 kJ/mol

(13) (11)

232

BALAT

Downloaded by [189.37.64.11] at 10:48 25 March 2012

where n is the average carbon number and m is the average number of hydrogen atoms of the hydrocarbon products. The FTS is one of the key steps in the production of environmentally clean fuels. Production of hydrocarbons from syngas is possible with catalysts based on Ni, Fe and Co. The iron oxide catalyst working under pressure of 2.5 MPa in the temperature range of 603623 K is used in the SASOL (South African Coal, Oil, and Gas Corporation) process for production of diesel fuel and heavy hydrocarbons. Although a number of catalysts for Fischer-Tropsch process are developed, the new effective catalysts of various chemical compositions and geometric shapes are foreseen. The efficiency of a number of catalytic reactions and hence the catalytic performance, among other important factors, depend on the capability of the catalyst for heat transfer and diffusion (Balat, 2006). The products from FTS are mainly aliphatic straight-chain hydrocarbons (CxH y). Besides the CxHy also branched hydrocarbons, unsaturated hydrocarbons, and primary alcohols are formed in minor quantities. The product distribution obtained from FTS includes the light hydrocarbons methane (CH4), ethene (C2H 4) and ethane (C2H5), LPG (C3-C4, propane and butane), gasoline (C 5-C12), diesel fuel (C13-C22), and light and waxes (C23-C33) (Rapagna et al., 1998; Prins et al., 2004). The distribution of products is described by so-called SchulzFlory equation (Oukaci, 2002):

Wn = (1 a )2 n a n 1
where; n: carbon number Wn: weight fraction of product with carbon number n : chain growth probability Figure 11 shows the green diesel and other products from biomass via FisherTropsch synthesis. The design of a biomass gasifier integrated with a FTS reactor must be aimed at achieving a high yield of liquid hydrocarbons. For the gasifier, it is important to avoid methane formation as much as possible, and convert all carbon in the biomass to mainly carbon monoxide and carbon dioxide (Prins et al., 2004; Demirbas, 2007). The FT catalytic conversion process can be used to synthesize diesel fuels from a variety of feedstocks, including coal, NG and biomass. The production of diesel fuel from bio-syngas by FTS is given in Figure 12. The FT fuel can be used as alternative diesel fuel. Synthetic FT diesel fuels can have excellent autoignition characteristics. The FT diesel is composed of only straight chain hydrocarbons and has no aromatics or sulfur. Reaction parameters are temperature, pressure and H2/CO ratio. The synthetic FT diesel fuel can provide benefits in terms of both PM and NOx emissions (May, 2003). FT is most compatible with existing distribution for conventional diesel and only minimal adjustments are required to obtain optimal performance from existing Diesel engines. Properties of FT and No. 2 diesel fuels are given in Table 9. Physical properties of FT are very similar to No. 2 diesel fuel, and its chemical

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

233

Downloaded by [189.37.64.11] at 10:48 25 March 2012

Figure 11 Green diesel and other products from biomass via Fisher-Tropsch synthesis.

Figure 12 Production of diesel fuel from bio-syngas by Fisher-Tropsh synthesis (FTS).

properties are superior in that the FT process yields middle distillates that, if correctly processed (as through a cobalt-based catalyst), contain no aromatics or sulfur compounds.

234

BALAT Table 9 Properties of Fisher-Tropsch (F-T) diesel and No. 2 diesel fuels. Property Density, g/cm3 Higher heating value, MJ/kg Aromatics, % Cetane number Sulfur content, ppm Fisher-Tropsch diesel 0.7836 47.1 00.1 7680 00.1 No. 2 petroleum diesel 0.8320 46.2 816 4755 25125

CONCLUSIONS Bio-fuels are made from plant matter and residues, such as agricultural crops, municipal wastes and agricultural and forestry by-products. Bio-ethanol can be obtained from cellulose and hemicelluloses via hydrolysis and fermentation processes. Bio-ethanol will continue to be developed as a transport fuel produced in tropical latitudes and traded internationally, for use primarily as a gasoline additive. It is by far the most widely used bio-fuel for transportation worldwide. Biochemical conversion of biomass is completed through alcoholic fermentation to produce liquid fuels and anaerobic digestion or fermentation, resulting in biogas. Bio-fuels are also made from biomass through thermochemical processes, such as pyrolysis, gasification, liquefaction and supercritical fluid extraction, or biochemical processes. Fischer-Tropsch processing of biomass into liquid fuels is not a realistic proposition. Biodiesel is made from renewable biological sources such as vegetable oils and animal fats. ACKNOWLEDGMENT
I would like to thank the Sila Science for the financial support and Prof. Ayhan Demirbas for his helpful comments on this paper.

Downloaded by [189.37.64.11] at 10:48 25 March 2012

REFERENCES
Akgerman, A. (1984). Diffusivities of Synthesis Gas and Fischer-Tropsch Products in Slurry Media. DOE Report- DOE/PC/70032-T1, the US Department of Energy-USDOE Assistant Secretary for Fossil Energy, Washington, DC, USA. Alencar, J.W., Alves, P.B., Craveiro, A.A. (1983). Pyrolysis of tropical vegetable oils. J. Agric. Food Chem. 31:12681270. Asher, A. (2006). Opportunities in Biofuels Creating Competitive Biofuels Markets. Biofuels Australasia 2006 Conference, Sydney, Australia, November 2022. Badger, P.C. (2002). Ethanol From Cellulose: A General Review. In trends in new crops and new uses. J. Janick and A. Whipkey (eds.), ASHS Press, Alexandria, VA, pp.1721. Bala, B.K. (2005). Studies on biodiesels from transformation of vegetable oils for diesel engines. Energy Education Science and Technology 15:145. Balat, M. (2005a). Current alternative engine fuels. Energy Sources 27:569577. Balat, M. (2005b). Biodiesel from vegetable oils via transesterification in supercritical ethanol. Energy Education Science and Technology 16:4552. Balat, M., Ozdemir, N. (2005). New and Renewable Hydrogen Production Processes. Energy Sources 27:12851298. Balat, M. (2006). Sustainable transportation fuels from biomass materials. Energy Education Science and Technology 17: 83103.

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

235

Billaud, F., Dominguez, V., Broutin, P., Busson, C. (1995). Production of hydrocarbons by pyrolysis of methyl esters from rapeseed oil. J. Am. Oil Chem. Soc. 72:11491154. Bohlmann, G.M. (2006). Process economic considerations for production of ethanol from biomass feedstocks. Industrial Biotechnology 2:1420. Boyle, G. (2005). An Overview of Alternative Transport Fuels in Developing Countries: Drivers, Status, and Factors Influencing Market Deployment. Hydrogen Fuel Cells and Alternatives in the Transport Sector: Issues for Developing Countries, United Nations University International Conference, Maastricht, The Netherlands, November 79. Carraretto, C., Macor, A., Mirandola, A., Stoppato, A., Tonon, S. (2004). Biodiesel as alternative fuel: Experimental analysis and energetic evaluations. Energy 29: 21952211. Chand, C.C., Wan, S.W. (1947). China's motor fuels from tung oil. Ind. Eng. Chem. 39:15431548. Chen, C., Spliethoff, H., Yang, L.B., Andries, J. (2003). Hydrogen production from gasificationPyrolsis of biomass. III. International Slovak Biomass Forum, Bratislava, February 34, pp.3640. Czernik, S., French, R., Feik, C., Chornet, E. (2000). Production of Hydrogen from BiomassDerived Liquids. Hydrogen, Fuel Cells & Infrastructure Technologies Program, Proceedings of the 2000 DOE Hydrogen Program Review, Vol. 1, Washington, DC, USA. Demirbas, A., Caglar, A. (1998). Catalytic reforming of biomass and heavy oil residues to hydrogen. Energy Educ Sci Technol 1:4552. Demirbas, A. (2001). Yields of hydrogen of gaseous products via pyrolysis from selected biomass samples. Fuel 80:18851891. Demirbas, A. (2002). Diesel fuel from vegetable oil via transesterification and soap pyrolysis. Energy Sources 24:835841. Demirbas, A. (2003). Biodiesel fuels from vegetable oils via catalytic and non-catalytic supercritical alcohol transesterifications and other methods: a survey. Energy Convers Mgmt. 44:20932109. Demirbas, A. (2004a). Bioenergy, global warming, and environmental impacts. Energy Sources 26:22536. Demirbas, M.F. (2004b). Producing Hydrogen from Biomass via Non-conventional Processes. Energy Explor. Exploit. 22:225233. Demirbas, A. (2004c). Hydrogen rich gas from fruit shells via supercritical water extraction. International Journal of Hydrogen Energy 29:12371243. Demirbas, A. (2005a). Bioethanol from cellulosic materials: A renewable motor fuel from biomass. Energy Sources 27:327337. Demirbas, A. (2005b). Biodiesel production from vegetable oils via catalytic and non-catalytic supercritical methanol transesterification methods. Progress in Energy and Combustion Science 31:466487. Demirbas, A. (2005c). Hydrogen production from biomass via supercritical water extraction. Energy Sources 27:14091417. Demirbas, A. (2006a). Global biofuel strategies. Energy Education Science and Technology 17:3363. Demirbas, A. (2006b). Biodiesel production via non-catalytic SCF method and biodiesel fuel characteristics. Energy Convers Mgmt 47: 22712282. Demirbas, M.F. (2006c). Hydrogen from Various Biomass Species via Pyrolysis and Steam Gasification Processes. Energy Sources 28:245252. Demirbas, A. (2007). Progress and recent trends in biofuels. Progress in Energy and Combustion Science 33:118 Demirbas, M.F., Balat, M. (2006). Recent advances on the production and utilization trends of biofuels: A global perspective. Energy Conversion and Management 47:23712381. Dmytryshyn, S.L., Dalai, A.K., Chaudhari, S.T., Mishra, H.K., Reaney, M.J. (2004). Synthesis and characterization of vegetable oil derived esters: evaluation for their diesel additive properties. Bioresource Technology 92:5564. Dufey, A. (2006). Biofuels production, trade and sustainable development: emerging issues. International Institute for Environment and Development (IIET), London, November, 60 p.

Downloaded by [189.37.64.11] at 10:48 25 March 2012

236

BALAT

Dunn, R.O. (2001). Alternative jet fuels from vegetable-oils. Trans ASAE 44:11511157. Dupont, C., Boissonnet, G., Seiler, J.M., Gauthier, P., Schweich, D. (2007). Study about the kinetic processes of biomass steam gasification. Fuel 86:3240. EBB (European Biodiesel Board). (2006). EU biodiesel production growth hits record high in 2005. EBB publishes annual biodiesel production statistics, 164/COM/06, Bruxelles, April 25. Eikeland, P.O. (2006). Biofuels the new oil for the petroleum industry? FNI Report 15/2005, The Fridtjof Nansen Institute, Lysaker, Norway, January, 39 p. Ericsson, K., Nilsson, L.J. (2004). International biofuel tradea study of the Swedish import. Biomass Bioenergy 26:20520. Fukuda, H., Konda, A., Noda, H. (2001). Biodiesel fuel production by transestirification of oils. Journal of Bioscience and Bioengineering 92:405416. Furuta, S., Matsuhasbi, H., Arata, K. (2004). Biodiesel fuel production with solid superacid catalysis in fixed bed reactor under atmospheric pressure. Catalysis Communications 5: 721723. Grossley, T.D., Heyes, T.D., Hudson, B.J.F. (1962). The effect of heat on pure triglycerides. J. Am. Oil Chem. Soc. 39:914. Gururaja, J. (2005). Biofuels-issues, challenges and options. Enhancing International Cooperationon Biomass, 5th Global Forum on Sustainable Energy, Vienna, May 1113. Hanaoka, T.,Yoshida, T., Fujimoto, S., Kamei, K., Harada, M., Suzuki, Y., Hatano, H., Yokoyama, S., Minowa, T. (2005). Hydrogen production from woody biomass by steam gasification using a CO2 sorbent. Biomass and Bioenergy 28:6368. Hao, X.H., Guo, L.J., Mao, X., Zhang, X.M., Chen, X.J. (2003). Hydrogen production from glucose used as a model compound of biomass gasified in supercritical water. Int. J. Hydrogen Energy 28:5564. Iwasa, N., Kudo, S., Takahashi, H., Masuda, S., Takezawa, N. (1993). Highly selective supported Pd catalysts for steam reforming of methanol. Catal Lett 19:211216. Jansen, J.C. (2003). Policy Support for Renewable Energy in the European Union. Energy Research Centre of the Netherlands (ECN). (Available from: www.ecn.nl/docs/library/report/2003/ C03113. pdf). Joshi, R.M., Pegg, M.J. (2007). Flow properties of biodiesel fuel blends at low temperatures. Fuel 86:143151. Kojima, M., Johnson, T. (2005). Potential for Biofuels for Transport in Developing Countries. Energy Sector Management Assistance Programme (ESMAP), Energy and Water Department, The World Bank Group, Washington, D.C., USA, October. Krammer, P., Vogel, H. (2000). Hydrolysis of esters in subcritical and supercritical water. Supercrit Fluids 16:189206. Kusdiana, D., Saka, S. (2001). Kinetics of transesterification in rapeseed oil to biodiesel fuels as treated in supercritical methanol. Fuel 80:693698. Kusdiana, D., Saka, S. (2004). Effects of water on biodiesel fuel production by supercritical methanol treatment. Bioresource Technol 91:289295. Lima, D.G.; Soares, V.C.D., Ribeiro, B.E., Carvalho, D.A., Cardoso, E.C.V., Rassi, F.C., Mundim, K.C., Rubima, J.C., Suarez, P.A.Z. (2004). Diesel-like fuel obtained by pyrolysis of vegetable oils. J. Anal. Appl. Pyrolysis 71:987996. Lin, Y., Tanaka, S. (2006). Ethanol fermentation from biomass resources: current state and prospects. Appl Microbiol Biotechnol 69: 627642. Loppacher, L.J., Kerr, W.A. (2005). Can biofuels become a global industry?: Government policies and trade constraints. Energy Politics 5:727. Ma, F., Hanna, M.A. 1999. Biodiesel production: A review. Biores Technol 70:115. Madras, G., Kolluru, C., Kumar, R. (2004). Synthesis of biodiesel in supercritical fluids. Fuel 83:20292033. May, M. (2003). Development and demonstration of Fischer-Tropsch fueled heavy-duty vehicles with control technologies for reduced diesel exhaust emissions. 9th Diesel Engine Emissions Reduction Conference. Newport, Rhode Island, August 2428.

Downloaded by [189.37.64.11] at 10:48 25 March 2012

GLOBAL TRENDS ON THE PROCESSING OF BIO-FUELS

237

Meher, L.C., Sagar, D.V., Naik, S.N. (2006). Technical aspects of biodiesel production by transesterificationa review. Renewable and Sustainable Energy Reviews 10:248268. Merida, W., Maness, P.C., Brown, R.C., Levin, D.B. (2004). Enhanced hydrogen production from indirectly heated, gasified biomass, and removal of carbon gas emissions using a novel biological gas reformer. International Journal of Hydrogen Energy 29:283290. Midilli, A., Dogru, M., Howarth, C.R., Ayhan, T. (2001). Hydrogen production from hazelnut shell by applying air-blown downdraft gasification technique. International Journal of Hydrogen Energy 26;2937. Moitra, N. (2004). An insight into the oil prone precursors of fossil fuels for selection of an exinerich, high hydrogen biomass, its pyrolysis, and product evaluation as fuel. Energy Sources 26:13631368. Nwafor, O.M.I. (2004). Emission characteristics of diesel engine operating on rapeseed methyl ester. Renewable Energy 29:119129. Oukaci, R. (2002). Fischer-Tropsch Synthesis. 2nd Annual Global GTL Summit Executive Briefing, London, UK, May 2830. Pimentel, D., Patzek, T.W. (2005). Ethanol production using corn, switchgrass, and wood; biodiesel production using soybean and sunflower. Natural Resources Research 14:6576. Pioch, D., Lozano, P., Rasoanatoandro, M.C., Grailla, J., Geneste, P., Guida, A. (1993). Biofuels from catalytic cracking of tropical vegetable oils. Oleagineux 48:289291. Prins, M.J., Ptasinski, K.J., Janssen, F.J.J.G. (2004). Exergetic optimisation of a production process of FischerTropsch fuels from biomass. Fuel Proc Technol 86:375389. Puppan, D. (2002). Environmental evaluation of biofuels. Period Polytech Ser Soc Man Sci 10:95116. Rapagna, S., Jand, N., Foscolo, P.U. 1998. Catalytic gasification of biomass to produce hydrogen rich gas. Int. J. Hydrogen Energy 23:551557. Saka, S., Kusdiana, D. (2001). Biodiesel fuel from rapeseed oil as prepared in supercritical methanol. Fuel 80:225231. Sastry, G.S.R., Krishna Murthy, A.S.R., Raviprasad, P., Bhuvaneswari, K., Ravi, P.V. (2006). Identification and determination of bio-diesel in diesel. Energy Sources, Part A 28:13371342. Schnepf, R. (2006). European Union Biofuels Policy and Agriculture: An Overview. CRS Report for Congress, March 16. Schuchardt, U., Sercheli, R., Vargas, R.M. (1998). Transesterification of vegetable oils: A review. J. Braz. Chem. Soc. 9:199210. Schwab, A.W., Dykstra, G.J., Selke, E., Sorenson, S.C., Pryde, E.H. (1988). Diesel fuel from thermal decomposition of soybean oil. J. Am. Oil Chem. Soc. 65:17811786. Spath, P., Jane, J., Mann, M., Amos, W. (2000). Update of Hydrogen from Biomass: Determination of the Delivered Cost of Hydrogen. NREL Milestone Report, April. Specht, M., Bandi, A., Baumgart, F., Murray, C.N., Gretz, J. (1998). Synthesis of Methanol from Biomass/CO2 Resources. 4th International Conference on Greenhouse Gas Control Technologies, Interlaken, Switzerland, August 30 September 2. Specht, M., Bandi, A. (1999). The Methanol-Cycle Sustainable Supply of Liquid Fuels. Center of Solar Energy and Hydrogen Research (ZSW), Hessbruehlstr. 21C, 70565. Stuttgart. (www.methanol.org/pdf/ZSWMethanolCycle.pdf). Srivastava, A., Prasad, R. (2000). Triglycerides-based diesel fuels. Renewable and Sustainable Energy Reviews 4:111133. Stavarache, C., Vinatoru, M., Nishimura, R., Maed, Y. (2005). Fatty acids methyl esters from vegetable oil by means of ultrasonic energy. Ultrasonics Sonochemistry 12:367372. Stevens, D.J., Wrgetter, M., Saddler, J. (2004). Biofuels for Transportation: An Examination of Policy and Technical Issues. IEA Bioenergy Task 39, Liquid Biofuels Final Report 20012003. Sun, Y., Cheng, J.J. (2005). Dilute acid pretreatment of rye straw and bermudagrass for ethanol. Bioresource Technology 96:15991606. Takezawa, N., Shimokawabe, M., Hiramatsu, H., Sugiura, H., Asakawa, T., Kobayashi, H. (1987). Steam reforming of methanol over Cu/ZrO2. Role of ZrO2 support. React Kinet Catal Lett 33:191196.

Downloaded by [189.37.64.11] at 10:48 25 March 2012

238

BALAT

Downloaded by [189.37.64.11] at 10:48 25 March 2012

USDA (United States Department of Agriculture). (2006). The Economic Feasibility of Ethanol Production from Sugar in the United States. Washington, D.C., July. (www.usda.gov/oce/ EthanolSugarFeasibilityReport3.pdf). van der Laan, G.P. (1999). Kinetics, selectivity and scale up of the Fischer-Tropsch synthesis. Thesis University of Groningen, Groningen, the Netherlands, April. (dissertations.ub.rug.nl/ faculties/ science/1999/g.p.van.der.laan/ 23k ). van Swaaij, W.P.M., Kersten, S.R.A., van den Aarsen, F.G. (2004). Routes for methanol from biomass. International 2-Day Business Conference on Sustainable Industrial Developments, Delfzijl, The Netherlands, April. Wang, D., Czernik, S., Montana, D., Mann, M., Chaornet, E. (1997). Biomass to hydrogen via fast pyrolysis and catalytic steam reforming of the pyrolysis oil or its fractions. Ind. Eng. Chem. Res. 36:15071518. Wang, D., Czernik, S., Chornet, E. (1998). Production of hydrogen from biomass by catalytic steam reforming of fast pyrolysis oils. Energy Fuels 12:1924. Wierzbicka, A., Lillieblad, L., Pagels, J., Strand, M., Gudmundsson, A., Gharibi, A., Swietlicki, E., Sanati M., Bohgard, M. (2005). Particle emissions from district heating units operating on three commonly used biofuels. Atmospheric Environment 39:139150. Zhang, Y., Dube, M.A., McLean, D.D., Kates, M. (2003). Biodiesel production from waste cooking oil: 2. Economic assessment and sensitivity analysis. Bioresource Technol 90:229240.

Você também pode gostar