Sunteți pe pagina 1din 12

Applied Energy 88 (2011) 1728

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Competitive liquid biofuels from biomass


Ayhan Demirbas *
Sirnak University, Dean of Engineering Faculty, Department of Mechanical Engineering, Sirnak, Turkey

a r t i c l e

i n f o

a b s t r a c t
The cost of biodiesels varies depending on the feedstock, geographic area, methanol prices, and seasonal variability in crop production. Most of the biodiesel is currently made from soybean, rapeseed, and palm oils. However, there are large amounts of low-cost oils and fats (e.g., restaurant waste, beef tallow, pork lard, and yellow grease) that could be converted to biodiesel. The crop types, agricultural practices, land and labor costs, plant sizes, processing technologies and government policies in different regions considerably vary ethanol production costs and prices by region. The cost of producing bioethanol in a dry mill plant currently totals US$1.65/galon. The largest ethanol cost component is the plant feedstock. It has been showed that plant size has a major effect on cost. The plant size can reduce operating costs by 1520%, saving another $0.02$0.03 per liter. Thus, a large plant with production costs of $0.29 per liter may be saving $0.05$0.06 per liter over a smaller plant. Viscosity of biofuel and biocrude varies greatly with the liquefaction conditions. The high and increasing viscosity indicates a poor ow characteristic and stability. The increase in the viscosity can be attributed to the continuing polymerization and oxidative coupling reactions in the biocrude upon storage. Although stability of biocrude is typically better than that of bio-oil, the viscosity of biocrude is much higher. The bio-oil produced by ash pyrolysis is a highly oxygenated mixture of carbonyls, carboxyls, phenolics and water. It is acidic and potentially corrosive. Bio-oil can also be potentially upgraded by hydrodeoxygenation. The liquid, termed biocrude, contains 60% carbon, 1020 wt.% oxygen and 3036 MJ/kg heating value as opposed to <1 wt.% and 4246 MJ/kg for petroleum. 2010 Elsevier Ltd. All rights reserved.

Article history: Received 22 February 2010 Received in revised form 13 July 2010 Accepted 18 July 2010 Available online 17 August 2010 Keywords: Biomass Liquid biofuel Bio-oil Biocrude Biodiesel Bioethanol

Contents 1. 2. 3. 4. 5. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bioethanol. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Current industrial applications of bioethanol and biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bio-oil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Effect of catalysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Yields and fuel properties of bio-oil. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Bio-oil upgrading. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biocrude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1. Liquefaction mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2. Properties of biocrude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3. Refinement and upgrading of biocrude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 19 19 20 21 22 22 23 24 25 25 26 26 26

6.

7.

1. Introduction Biofuel is a renewable energy source produced from natural (plant) materials, which can be used as a substitute for petroleum
* Tel.: +90 462 230 7831; fax: +90 462 248 8508. E-mail address: ayhandemirbas@hotmail.com 0306-2619/$ - see front matter 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.apenergy.2010.07.016

fuels. The most common biofuels, such as ethanol from corn, wheat or sugar beet and biodiesel from oil seeds, are produced from classic food crops that require high-quality agricultural land for growth [1,2]. Serious problems face the world food supply today. Food versus fuel is the dilemma regarding the risk of diverting farmland or crops for liquid biofuels production in detriment of the food supply

18

A. Demirbas / Applied Energy 88 (2011) 1728 Table 2 Estimated costs of biofuels compared with the price of oil (biofuels exclusive of taxes), (US cents/liter). Biofuel Corresponding pre-tax price of petroleum products Bioethanol from sugar cane Bioethanol from corn Bioethanol from beet Bioethanol from wheat Bioethanol from lignocellulose Biodiesel from animal fats Biodiesel from vegetable oil FischerTropsch synthesis liquids 2006 3560 2550 6080 6080 7095 80110 4055 70110 90110 Long-term about 2030 2535 3555 4060 4565 2565 4050 4075 7085

on a global scale. There is disagreement about how signicant this is, what is causing it, what the impact is, and what can or should be done about it. Biofuel production has increased in recent years. The rise in world oil prices led to a sharp increase in biofuels production around the world. Some commodities such as corn, sugar cane, and vegetable oil can be used either as food, feed or to make biofuels. Vegetable oils are a renewable and potentially inexhaustible source of energy with energy content close to diesel fuel. On the other hand, extensive use of vegetable oils may cause other significant problems such as starvation in developing countries. Forest and agricultural education, science and modern technology leads in the solving the problems of global food resources [310]. Biofuels can be classied based on their production technologies: First-generation biofuels; second generation biofuels; third generation biofuels; and fourth generation biofuels. Classication of biofuels based on their generation technologies is shown in Table 1. The rst-generation biofuels appear unsustainable because of the potential stress that their production places on food commodities. Second generation biofuels need to build on the need for sustainable liquid fuels through processing including pyrolysis and hydrothermal liquefaction. FischerTropsch and other catalytic processes in order to make more complex molecules and materials on which a future sustainable society will be based [1118]. Energy prices inuence consumer choices and behavior and can affect economic development and growth. Energy prices, which include taxes, must be clearly distinguished from costs, average costs from marginal costs, and contract markets from spot markets. Fuel from biomass comprises the largest exploited renewable energy globally. Low cost, high capacity processes for the conversion of abundantly available biomass into liquid biofuels are essential for reducing dependence on petroleum sources, expanding the utilization of carbon neutral processes, and increasing rural income [1923]. Production of grain-based ethanol and vegetable-oilbased biodiesel are being practiced today with difculties due to the competition with the food supply. Biofuels production costs can vary widely by feedstock, conversion process, scale of production and region. For biofuels, the cost of feedstock (crops) is a major component of overall costs. Total biofuel costs should also include a component representing the impact of biofuels production on related markets, such as food. In particular, the cost of producing oil-seed-derived biodiesel is dominated by the cost of the oil and by competition from high-value uses like cooking [14,24,25]. Table 2 shows estimates of the costs of biofuels [26]. The objective is to produce inexpensive biomass streams that can be used to make a range of fuels, chemicals and other materials that are cost competitive with conventional commodities. The term liquid biofuel is referred to biomass-to-liquid fuel (BTLF). Liquid biofuels may offer a promising alternative. Liquid biofuels are substitute fuel sources to petroleum; however some still include a small amount of petroleum in the mixture. The biggest difference between biofuels and petroleum feedstocks is oxygen content [2730].
Table 1 Classication of biofuels based on their generation technologies. Generation First-generation biofuels Second generation biofuels Third generation biofuels Fourth generation biofuels Feedstocks Sugar, starch, vegetable oils, or animal fats Non food crops, wheat straw, corn, wood, solid waste, energy crop Algae Vegetable oil, biodiesel Examples Bioalcohols, vegetable oil, biodiesel, biosyngas, biogas Bioalcohols, bio-oil, bioDMF, biohydrogen, bioFischerTropsch diesel Vegetable oil, biodiesel Biogasoline

There are two global liquid transportation biofuels that might replace gasoline and diesel fuel; these are bioethanol and biodiesel, respectively. It is assumed that biodiesel is used as a petroleum diesel replacement and that bioethanol is used as a gasoline replacement [15,31,32]. Ethanol derived from bio-matter such as corn, sugar cane or grain is frequent, this will often be referred to as bioethanol. Ethanol demand is expected to more than double in the next 10 years. For the supply to be available to meet this demand, new technologies must be moved from the laboratories to commercial reality [30]. The world ethanol production is about 60% from feedstock from sugar crops. Biodiesel refers to any diesel-equivalent biofuel usually made from vegetable oils or animal fats. It can be used directly as fuel, which requires some engine modications, or blended with petroleum diesel and used in diesel engines with few or no modications. Biodiesel has become more attractive recently because of its environmental benets [3336]. The cost of biodiesel, however, is the main obstacle to commercialization of the product. With cooking oils used as raw material, the viability of a continuous transesterication process and recovery of high quality glycerol as a biodiesel by-product are primary options to be considered to lower the cost of biodiesel [37,38]. Biomass, when subjected to high temperature in the absence of oxygen (i.e., pyrolysis), converts into gas, solid char, and liquid products. The liquid product, called bio-oil or pyrolysis oil, is typically brown, dark red, or black in color with a density of about 1.2 kg/L. Bio-oil has water content of typically 1433 wt.%, which cannot be easily removed by conventional methods (e.g., distillation); in fact, bio-oil may phase separate above certain water content. The higher heating value of bio-oil is typically 1522 MJ/kg, which is lower than that for conventional fuel oil (4346 MJ/kg), mainly due to the presence of oxygenated compounds in bio-oil [39]. In search of renewable fuels, as early as the mid-20th century, researchers started to convert biomass into petroleum-like liquids. For example, Berl [40] treated biomass using alkaline water at 500 K to produce a viscous liquid that contained 60% carbon and 75% heating value of the starting material. The liquid, termed biocrude, contains 1020 wt.% oxygen and 3036 MJ/kg heating value as opposed to <1 wt.% and 4246 MJ/kg for petroleum [41]. The high oxygen content imparts lower energy content, poor thermal stability, lower volatility, higher corrosivity, and tendency to polymerize over time [42]. Compared with bio-oil from fast pyrolysis, biocrude produced from hydrothermal liquefaction has higher energy value and lower moisture content but requires longer residence time and higher capital costs. Typical hydrothermal liquefaction conditions range from 550 to 650 K, 70 to 20 MPa, with liquid water present and reaction occurring for 1060 min. The biomass thermochemical conversion processes are globally endothermic, and the heat required can be supplied by concentrated solar energy in such a way that the energy evolved from

A. Demirbas / Applied Energy 88 (2011) 1728

19

the fuel produced ideally represents the sum of energy stored during the photosynthesis and the direct thermal collection [43]. Nonetheless, if extra energy is supplied from solar and wind resources, then more of the biomass carbon can be converted into the liquid fuel [4446]. Solar energy has the potential to provide many of the thermal needs of biofuel production plants. The use of direct concentrated radiation in biomass reactors has been tested, with some challenges. A variety of solar systems are available that can produce and store energy at the temperature levels needed in biomass processing. A working uid can then transfer the heat to the biofuel process when needed. With supplementing energy from solar and wind sources, a higher amount of carbon in biomass can be converted to liquid fuels [4355]. Biofuels could be a peaceful energy carrier for all countries. Biofuels have received increased attention as a renewable and environmentally-friendly option to help meet todays energy needs [56]. Policy-makers will need to pay more attention to the implications for the transition to biofuel economy. The generation of biofuel depends on the availability of production necessities both of which are increasingly rare and may become political issues, as much as oil and natural gas are today [57]. Due to some technological and economic consequences, practical experiences of biofuel-energy do have wide applications neither in the richest countries nor in the poorest countries at present. For the developed countries, active involvement in biofuel research and development, especially through collaborative international programs, could facilitate the introduction of new biofuel technologies as they become competitive. A major dilemma now faced by the developing countries is how to invest in biofuel research and development for the transition to biofuel economy. Most developing countries will probably rather than developers of cutting-edge technologies. Developing countries have at least as much to gain from a move towards the biofuel economy as industrialized ones, since they generally suffer more from urban pollution and their economies tend to be more energy intensive. International organizations have an important role to play in assisting countries in creating a market-based policy relating to biofuel and other clean energy systems [58]. International organizations should support the developing countries for the transition to a biofuel economy to provide capital both national and foreign [5962]. Biofuels share in the energy market is increasing with the implementation of fuel cell systems for sustainable energy supply. The concept of sustainable development embodies the idea of the inter-linkage and the balance between economic, social and environmental concerns. Current European Union (EU) policies on alternative motor fuels focus on the promotion of biofuels [63 65]. The denition of the marginal producer depends on the policy stance on biofuels. Biofuel pricing policy should not be employed as an anti inationary instrument. It should be applied in such a way, that it does not create cross subsidies between classes of consumers. In a proposed biofuels directive the introduction of a mandatory share scheme for biofuels, including as from 2009 minimum blending shares [66].

Table 3 General properties of biodiesels. Chemical name Chemical formula range Kinematic viscosity range Density range Boiling point range Flash point range Distillation range Vapor pressure Solubility in water Physical appearance Odor Biodegradability Reactivity Fatty acid (m)ethyl ester C14C24 methyl esters or C1525 H28-48O2 3.35.2 mm2/s, at 40 C 860894 kg/m3, at 15 C 200 C 155180 C 195325 C <5 mm Hg, at 22 C Insoluble in water; however, biodiesel can absorb up to 1500 ppm water Light to dark yellow, clear liquid Light musty/soapy odor More biodegradable than petrodiesel Stable, but reacts with strong oxidizers

2. Biodiesel Biodiesel refers to a renewable fuel for diesel engines that is derived from animal fats or vegetable oils (e.g., rapeseed oil, canola oil, soybean oil, sunower oil, palm oil, used cooking oil, beef tallow, sheep tallow, and poultry oil). Biodiesel is a clear amberyellow liquid with a viscosity similar to petroleum diesel (petrodiesel, diesel). With the ash point of 420 K, biodiesel is nonammable and nonexplosive, in contrast to petrodiesel, which has a ash point of 337 K. This property makes biodiesel-fueled vehicles much

safer in accidents than those powered by diesel or gasoline. Unlike petrodiesel, biodiesel is biodegradable and nontoxic and signicantly reduces toxic and other emissions when burned as a fuel [6775]. Technically, biodiesel is a diesel engine fuel comprised of monoalkyl esters of long-chain fatty acids derived from animal fats or vegetable oils, designated B100, and meeting the requirements of the ASTM D-6751 standard. Some of its technical properties are listed in Table 3 [15]. Chemically, biodiesel is referred to as a monoalkyl ester, especially (m)ethyl ester, of long-chain fatty acids derived from natural lipids via the transesterication process. Biodiesel is typically produced by reacting a vegetable oil or animal fat with methanol or ethanol in the presence of a catalyst to yield methyl or ethyl esters (biodiesel) and glycerin [20]. The purpose of the transesterication process is to lower the viscosity and oxygen content of the vegetable oil. In this process, an alcohol (e.g., methanol, ethanol, butanol) is reacted with the vegetable oil (fatty acid) in the presence of an alkali catalyst (e.g., KOH, NaOH) to produce biodiesel and glycerol. Being immiscible, biodiesel is easily separated from glycerol. Transesterication is an inexpensive way of transforming the large, branched molecular structure of the vegetable oils into smaller, straight-chain molecules of the type required in regular diesel combustion engines. Generally, methanol is preferred for transesterication, because it is less expensive than ethanol. Biodiesel produces slightly lower power and torque, which results in a higher consumption than No. 2 diesel fuel for driving. However, biodiesel is better than diesel in terms of sulfur content, ash point, aromatic content, and biodegradability. Precautions should be taken in very cold climates, where biodiesel may gel earlier than diesel upon cooling. The cost of biodiesels varies depending on the feedstock, geographic area, methanol prices, and seasonal variability in crop production [7683]. Most of the biodiesel is currently made from soybean, rapeseed, and palm oils. The high value of soybean oil as a food product makes production of a cost-effective biodiesel very challenging. However, there are large amounts of low-cost oils and fats (e.g., restaurant waste, beef tallow, pork lard, and yellow grease) that could be converted to biodiesel. Biodiesel produces slightly lower power and torque; hence, the fuel consumption is higher than No. 2 diesel. However, biodiesel is better than diesel in terms of sulfur content, ash point, aromatic content, and biodegradability [15,20]. 3. Bioethanol Ethanol from carbohydrates by fermentation is a historical industry with very early application in beverage making. The

20

A. Demirbas / Applied Energy 88 (2011) 1728

recent use of ethanol as fuel has increased its production. Most ethanol is currently being produced from sugar cane or corn [23]. Yeast is used to ferment sugars into ethanol. In the case of carbohydrates (such as corn), a pretreatment step of converting carbohydrate into sugars is needed. Currently, the corn ethanol industry uses either a dry-milling or a wet-milling process. Upon fermentation, ethanol content is only about 10%, which requires a signicant effort in separation to produce the pure ethanol needed for fuel use. Distillation can concentrate ethanol to just below the azeotropic concentration (95 mol%), after that, specialized separations (e.g., molecular sieve, azeotropic distillation, lime drying) are needed. Efforts are underway to improve the microorganism and separation processes to reduce the overall cost of ethanol production. With the current trend in ethanol use, demand is likely to increase signicantly in the near future [8486]. Starch or sugar-based ethanol production has been blamed for the rise in the food prices. To satisfy current and future demands, ethanol production from lignocellulosic biomass fermentation is a viable option that does not compete with the food supply. The process includes the following key steps: pretreatment, hydrolysis of cellulose, hydrolysis of hemicelluloses, fermentation of ve-carbon and six-carbon sugars, separation of lignin residue, and recovery and concentration of ethanol. Currently, the ethanol produced from lignocellulosics is more expensive than that from starches. Hence, production costs need to be reduced. Requirements for this cost reduction include effective pretreatment to reduce cellulase use, hydrolysis of hemicelluloses and cellulose to sugars, use of both six-carbon and ve-carbon sugars, and process integration for reducing capital and energy costs [13,15,23]. Bioethanol can be used directly in cars designed to run on pure ethanol or blended with gasoline to make gasohol. Anhydrous ethanol is required for blending with gasoline. No engine modication is typically needed to use the blend. Ethanol can be used as an octane-boosting, pollution-reducing additive in unleaded gasoline. Wholesale US prices for fuels in 2008 and 2009 are shown in Table 4. The cost of feedstock is a major economic factor in the viability of biofuel production. Renewable alcohols are at present more expensive of synthesis-ethanol from ethylene and of methanol from natural gas. The simultaneous production of biomethanol (from sugar juice) in parallel to the production of bioethanol appears economically attractive. The crop types, agricultural practices, land and labor costs, plant sizes, processing technologies and government policies in different regions considerably vary ethanol production costs and prices by region. The cost of producing bioethanol in a dry mill plant currently totals US$1.65/gallon [87]. Ethanol from sugar cane, produced mainly in developing countries with warm climates, is generally much cheaper to produce than ethanol from grain or sugar beet in IEA countries. For this rea-

son, in countries like Brazil and India, where sugar cane is produced in substantial volumes, sugar cane-based ethanol is becoming an increasingly cost-effective alternative to petroleum fuels. Ethanol derived from cellulosic feedstock using enzymatic hydrolysis requires much greater processing than from starch or sugar-based feedstock, but feedstock costs for grasses and trees are generally lower than for grain and sugar crops. If targeted reductions in conversion costs can be achieved, the total cost of producing cellulosic ethanol in OECD countries could fall below that of grain ethanol [87]. Estimates show that bioethanol in the EU becomes competitive when the oil price reaches US$70 a barrel while in the United States it becomes competitive at US$5060 a barrel. For Brazil the threshold is much lower between US$25 and US$30 a barrel. Other efcient sugar producing countries such as Pakistan, Swaziland and Zimbabwe have production costs similar to Brazils. Anhydrous ethanol, blendable with gasoline, is still somewhat more expensive. Prices in India have declined and are approaching the price of gasoline [87]. The generally larger US conversion plants produce biofuels, particularly ethanol, at lower cost than plants in Europe. Production costs for ethanol are much lower in countries with a warm climate, with Brazil probably the lowest-cost producer in the world. Production costs in Brazil, using sugar cane as the feedstock, have recently been recorded at less than half the costs in Europe. Production of sugar cane ethanol in developing countries could provide a low-cost source for substantial displacement of oil worldwide over the next 20 years [88]. For biofuels, the cost of feedstock (crops) is a major component of overall costs. In particular, the cost of producing oil-seedderived biodiesel is dominated by the cost of the oil and by competition from high-value uses like cooking. The largest ethanol cost component is the plant feedstock. Operating costs, such as feedstock cost, co-product credit, chemicals, labor, maintenance, insurance and taxes, represent about one third of total cost per liter, of which the energy needed to run the conversion facility is an important (and in some cases quite variable) component. Capital cost recovery represents about one-sixth of total cost per liter. It has been showed that plant size has a major effect on cost. The plant size can reduce operating costs by 1520%, saving another $0.02 $0.03 per liter. Thus, a large plant with production costs of $0.29 per liter may be saving $0.05$0.06 per liter over a smaller plant [88]. 4. Current industrial applications of bioethanol and biodiesel During the OPEC oil embargo of the 1970s, ethanol fuel reemerged as a fuel extender during the gasoline shortage. Later, ethanol was used as a replacement for lead, as a cleaner burning

Table 4 Wholesale US prices for fuels in 2008 and 2009. Fuel Gasoline Jet fuel Diesel Natural gas Heating oil Coal Soybean oil Corn Ethanol Methanol Bio-diesel Wood (biomass) Heating value (lower) 121.7 MJ/gallon 116.8 MJ/gallon 126.0 MJ/gallon 38.1 MJ/kg 133.6 MJ/gallon 33.3 MJ/kg 32.5 MJ/kg 15 MJ/kg 79.8 MJ/gallon 18.0 MJ/kg 123.5 GJ/gallon 15 MJ/kg July 2008 price $/GJ $3.1/gallon $2.9/gallon $4.2/gallon $9.2/Million BTU $3.6/gallon $100/ton $0.6/lb $5.8/bushel; $0.23/kg $2.5/gallon $1.4/gallon or $526/ton $5.5/gallon $100/dry-ton ($50/green ton) 25.2 24.6 33.3 8.7 26.6 3.0 40.8 15.3 31.1 29.2 44.5 6.6 $1.3/gallon $1.2/gallon $1.2/gallon $4.3/Million BTU $1.2/gallon $66/ton $0.3/lb $3.4/bushel; $0.13/kg $1.6/gallon $0.65/gallon or $216/ton $2.6/gallon $50/ton ($25/green ton) March 2009 price $/GJ 10.7 10.3 9.5 4.1 9.0 2.0 20.4 8.9 20.1 13.8 21.0 3.3

A. Demirbas / Applied Energy 88 (2011) 1728

21

octane enhancer. In 1995, about 93% of the ethanol in the world was produced by the fermentation method and about 7% by the synthetic method. As a result of energy security needs, farmer incentives, and clean air regulations, US ethanol demand has grown from 175 million gallons in 1980 to about 9000 million gallons in 2008 [8991]. The global ethanol production has increased from 10.7 billion gallons in 2004 to 20.4 billion gallons in 2008. US ethanol production has exceeded that of Brazil (Table 5). Ethanol demand is expected to more than double in the next 10 years. At present, ethanol accounts for about 94% of the global biofuel production, with the majority being produced from sugarcane and corn. Brazil and the United States are the world leaders in ethanol production; together, they account for 78% of the worlds ethanol production. Ethanol is a well-established biofuel used for transport and in the industry sectors in several countries, notably in Brazil. Brazil has been using ethanol as a fuel since 1925; at that time, the production of ethanol was 70 times more than the production and consumption of petrol. Currently, the petrol sold in Brazil contains 25% ethanol. The United States has used ethanol produced from corn since 1980, and the current petrol contains 10% ethanol [92,93]. The fermentation process can use any sugar-containing material to produce ethanol. The sugar-containing agricultural raw materials can be classied into three categories: (1) sugar, (2) starch, and (3) cellulose. Sugars (e.g., from sugarcanes, molasses, sugar beats, and fruits) can be directly fermented using yeast to produce ethanol. Starch (e.g., from corn and other grains, potatoes, root crops) and cellulose (e.g., from wood, agricultural residue, grasses) are rst converted to sugar by hydrolysis/pretreatment and then fermented. Due to the sturdy nature of cellulose molecules and biomass structure, hydrolysis to sugar is a difcult process. The course of transesterication is affected by several factors, including the type of catalyst (alkaline, acid, or enzyme), molar ratio of alcohol-to-vegetable oil, temperature, purity of the reactants (mainly water content), and free fatty acid content. In conventional transesterication, free fatty acid and water always produce negative effects, as the presence of these compounds cause soap formation, consumes catalyst, and reduces catalyst effectiveness, leading to a low conversion [94,95]. The alcohol-to-vegetable oil molar ratio plays a major role in transesterication. For example, an excess of alcohol favors the formation of ester products. However, a further increase in the amount of alcohol makes the recovery of the glycerol difcult. Hence, an optimum alcohol-to-oil ratio has to be established for a given process. Base-catalyzed transesterication is faster than acid-catalyzed. The base catalyst rst reacts with the alcohol, producing an alkox-

ide and the protonated catalyst. The alkoxide attacks the carbonyl group of the triglyceride, generating a tetrahedral intermediate, from which the alkyl ester and the corresponding anion of the diglyceride are formed. The anion deprotonates the catalyst, thus regenerating it for another cycle. Diglycerides and monoglycerides are converted by the same mechanism to alkyl esters and glycerol. Alkaline metal alkoxides (e.g., CH3ONa and CH3CH2ONa) are the most active catalysts because they give very high yields (>98%) in short reaction times ($30 min), even when applied at low molar concentrations (0.5 mol%). However, these alkoxides require the absence of water, which makes them inappropriate for typical industrial processes because commercial vegetable oils do contain some amount of moisture. On the other hand, alkaline metal hydroxides (e.g., KOH and NaOH) are cheaper than metal alkoxides, but are less active. Water present in alcohol or oil can hydrolyse some of the produced ester, with consequent soap formation. This undesirable saponication reduces the biodiesel yields and causes difculty in glycerol recovery due to the formation of emulsions. Some remediations have been developed. For example, the use of 2 3 mol% potassium carbonate can increase the yield and reduce soap formation. Most of the biodiesel produced today is made via the base-catalyzed reaction for several reasons. It requires low temperature and pressure. It provides a high yield (98% conversion) with minimal side reactions and reaction time. It directly converts oil to biodiesel with no intermediate compounds. The reactors can be made from simple materials (e.g., stainless steel). The process includes the following key steps: (1) mixing of alcohol and catalyst, (2) transesterication reaction, (3) separation, (4) biodiesel washing, (5) unreacted alcohol removal, (6) glycerin neutralization, and (7) the nal product renement.

5. Bio-oil Pyrolysis presents an attractive option to convert solid biomass into liquid bio-oil, which is easier to transport, store, and upgrade. Pyrolysis is the thermal decomposition of biomass, which occurs in the absence of oxygen or when signicantly less oxygen is supplied than needed for complete combustion [90]. Conventional or slow pyrolysis is dened as the pyrolysis that occurs under a slow heating rate. Fast pyrolysis (also called thermolysis) is a process in which biomass is rapidly heated to high temperatures in the absence of oxygen. A high bio-oil production requires very low vapor residence time of typically 1 s to minimize secondary reactions; although, reasonable yields can be obtained at residence times of up to 5 s if the vapor temperature is kept below 675 K. In contrast to slow pyrolysis, fast pyrolysis is an advanced process, which needs to be carefully controlled to produce high yields of bio-oil. A typical reaction temperature of around 775 K is employed. In fact, fast pyrolysis has now been adopted for the production of food avors to replace the traditional slow pyrolysis, which had much lower yields [96,97]. Bio-oil is a viscous, corrosive, and unstable mixture of a large number of oxygenated molecules, depending on the pyrolysis process and biomass feedstock. Due to the high oxygen content, the heating value is less than half that of petroleum liquid. Bio-oil needs to be upgraded before use as liquid fuel. Various methods available include solvent fractionation, hydroprocessing, and catalytic cracking. Bio-oil can be produced from a variety of dried forest

Table 5 Global ethanol production, in billion gallons per year [93]. Country United States Brazil China India France Canada Germany Thailand Russia Spain South Africa United Kingdom Remaining countries World Year 2004 3.40 3.87 0.92 0.32 0.22 0.06 0.06 0.65 0.20 0.09 0.10 0.08 1.35 10.75 Year 2008 8.93 6.90 1.02 0.61 0.40 0.26 0.22 0.15 0.15 0.13 0.11 0.11 1.40 20.37 Share year 2008 (%) 43.8 33.9 5.0 3.0 1.9 1.3 1.1 0.7 0.7 0.6 0.5 0.5 6.9 100.0

22

A. Demirbas / Applied Energy 88 (2011) 1728

residue and agricultural wastes. Biomass feedstock with good potential include bagasse (from sugarcane), rice straw, rice hulls, peanut hulls, oat hulls, switchgrass, wheat straw, coconut bers, and wood. If the purpose is to maximize the yield of liquid products resulting from biomass pyrolysis, a low temperature, high heating rate, short gas residence time process would be required. For a high char production, a low temperature, low heating rate process would be chosen. If the purpose were to maximize the yield of fuel gas resulting from pyrolysis, a high temperature, low heating rate, long gas residence time process would be preferred. Biomass is a mixture of structural constituents (hemicelluloses, cellulose and lignin) and minor amounts of extractives which each pyrolyse at different rates and by different mechanisms and pathways. It is believed that as the reaction progresses the carbon becomes less reactive and forms stable chemical structures, and consequently the activation energy increases as the conversion level of biomass increases [98100]. Fuel properties of diesel, biodiesel and pyrolysis oil (bio-oil) from biomass are given in Table 6. A large-scale pyrolysis test was performed in a pilot-plant reactor [101], where air-dried bagasse (20 kg; 8 wt.% moisture) was placed in four pans and introduced into a cylindrical horizontal reactor 3 m long and 0.6 m in diameter (Fig. 1). The bagasse bed height was 140 mm. The sample was heated at a slow rate (over a period of 200 min) until the feed attained a temperature of 800 K. The pilot reactor was operated in batch mode at 800 K for 1 h. The solid residue was cooled to room temperature under a ow of nitrogen. The average total pressure during the pyrolysis was $12 kPa. Four traps that were maintained at 298, 273, 243, and 193 K were used as a condensation train [102]. 5.1. Effect of catalysis Several biomasses inherently contain metal ions and salts. Potassium and calcium are the major metal ions present, along with minor amounts of sodium, magnesium, and other elements. For others biomasses, metal ions can be added externally to alter or tailor the products from biomass pyrolysis. For example, the presence of alkaline cations is known to affect the mechanism of thermal decomposition. These cations cause fragmentation of the monomers from the natural polymer chains, rather than the predominant depolymerization that occurs in their absence. Due to the complex and heterogeneous nature of the system, these catalysts are not typically recycled back as in the case of conventional catalysts. Catalyzed biomass pyrolysis offers a great potential for future development as a route to modify bio-oil properties. Bio-oil is not stable and is not miscible with conventional liquid fuels. Hence, it needs to be upgraded, especially to reduce its oxygen and water content. The use of catalyst during the pyrolysis can greatly improve the composition of bio-oil. A variety of catalysts including pure zeolite (HZSM-5), FCC catalysts, aluminas (a,

Fig. 1. Schematic of the circulating large-scale pyrolysis uidized bed reactor.

c-Al2O3), transition metal catalysts (Fe/Cr), and Al-MCM-41 have


been tested [97,103]. The bio-oil obtained by catalytic pyrolysis does not require costly preupgradation processing involving condensation and reevaporation. The catalyst acidity and pore size affect the formation of aromatic and aliphatic hydrocarbons. Hydrogen transfer reactions, essential for hydrocarbon formation, are known to increase with catalyst acidity. The high acid density of ZnCl2 catalysts contribute greatly to high amounts of hydrocarbons in the liquid product. Zinc chloride catalyst, as a Lewis acid, contributed to hydrogen transfer reactions and the formation of hydrocarbons in the liquid phase [104]. 5.2. Yields and fuel properties of bio-oil Typical bio-oil yields are in the 6080 wt.% range, depending on the biomass composition [97]. For example, the bio-oil yields from wood are in the range of 7280 wt.%, depending on the relative content of cellulose and lignin in the feedstock. A lower bio-oil yield (6065 wt.%) is obtained from high lignin contents, such as that found in bark. However, the bio-oil obtained from a higher lignin containing biomass does have a higher energy density/content. Mixed-paper feedstock gives a yield in excess of 75 wt.%. Bio-oil differs signicantly from petroleum fuels due to its very high viscosity, moisture content, and oxygen content, and a lower

Table 6 Fuel properties of diesel, biodiesel and pyrolysis oil (bio-oil) from biomass. Property Flash point Water and sediment Kinematic viscosity (at 313 K) Sulfated ash Ash Sulfur Sulfur Copper strip corrosion Cetane number Aromaticity Carbon residue Carbon residue Test method D D D D D D D D D D D D 93 2709 445 874 482 5453 2622/129 130 613 1319 4530 524 ASTM D975 (diesel) 325 K min 0.05 max%vol 1.34.1 mm2/s 0.01 max%wt 0.05 max%wt No 3 max 40 min 0.35 max%mass ASTM D-6751 (biodiesel) 403 K 0.05 max%vol 1.96.0 mm2/s 0.02 max%wt 0.05 max%wt No 3 max 47 min 35 max%vol 0.05 max%mass Pyrolysis oil (bio-oil) 0.010.04 251000 0.050.01%wt 0.0010.02%wt 0.0010.02%wt

A. Demirbas / Applied Energy 88 (2011) 1728

23

heating value. The viscosity is comparable to heavy fuel oil, which depends on the biomass feedstock, pyrolysis temperature prole, degree of thermal degradation, degree of catalytic cracking, and the content of light components and water. The high water content (1530 wt.%) of bio-oil can be easily removed by conventional methods, such as distillation. Also, the high water content is responsible for low energy density, low ame temperature, and ignition difculties. Addition of more water to bio-oil can lead to phase separation. In contrast to diesel and gasoline, which are nonpolar and do not absorb water, bio-oil is highly polar and can readily absorb water up to 35 wt.% [105]. The higher heating value of bio-oil is 1619 MJ/kg compared with 4045 MJ/kg for conventional petroleum fuels. Fuel properties of diesel, biodiesel and pyrolysis oil (bio-oil) from biomass are given in Table 7. Typical properties and characteristics of wood-derived bio-oil are presented in Table 8. Bio-oil has many special features and characteristics. These require consideration before any application, storage, transport, upgrading or utilization is attempted. The biooil formed at 725 K contains a high concentration of compounds, such as acetic acid, 1-hydroxy-2-butanone, 1-hydroxy-2-propanone, methanol, 2,6-dimethoxyphenol, 4-methyl-2,6-dimetoxyphenol and 2-cyclopenten-1-one, and so on, with a high percentage of methyl derivatives. As the temperature is increased, some of these compounds transform via hydrolysis. The formation of unsaturated compounds generally involves a variety of reaction pathways, such as dehydration, cyclization, DielsAlder cycloaddition reactions, and ring rearrangement. For example, 2,5-hexandione can undergo cyclization under hydrothermal conditions to produce 3-methyl-2-cyclopenten-1-one with very high selectivity of up to 81% [105,106]. 5.3. Bio-oil upgrading Fig. 2 shows biomass upgrading by pyrolysis. Upgrading of condense liquid from biomass include the three stages. There are physical upgrading (differential condensation, liquid ltration and solvent addition), catalytic upgrading (deoxygenating and reforming), and chemical upgrading (new fuel and chemicals synthesis). The bio-oil obtained from the fast pyrolysis of biomass has high oxygen content. Ketones and aldehydes, carboxylic acids and esters, aliphatic and aromatic alcohols, and ethers have been detected in signicant quantities. Because of the reactivity of oxygenated groups, the main problems of the oil are instability. Therefore study of the deoxygenation of bio-oil is needed. In the

Table 8 Upgrading processes of ash pyrolysis liquid or bio-oil. Upgrading process Hydro-deoxygenation (HDO) Homogenous catalyst hydrogenation Alcoholysis and reactive distillation Reactive extraction Products Fuel, chemicals (phenolics) Stabilized oil Fuel Chemicals (organic acids)

Fig. 2. Biomass upgrading by pyrolysis.

previously work the mechanism of hydrodeoxygenation (HDO) of bio-oil in the presence of a cobalt molybdate catalyst was studied [107].

Table 7 Typical properties and characteristics of wood-derived bio-oil. Property Appearance Miscibility Characteristics From almost black or dark red-brown to dark green, depending on the initial feedstock and the mode of fast pyrolysis Varying quantities of water exist, ranging from $15 wt.% to an upper limit of $3050 wt.% water, depending on production and collection Pyrolysis liquids can tolerate the addition of some water before phase separation occurs Bio-oil cannot be dissolved in water Miscible with polar solvents such as methanol and acetone, but totally immiscible with petroleum-derived fuels Bio-oil density is $1.2 kg/L, compared to $0.85 kg/L for light fuel oil Viscosity of bio-oil varies from as low as 25 cSt to as high as 1000 cSt (measured at 313 K) depending on the feedstock, the water content of the oil, the amount of light ends that have collected, the pyrolysis process used, and the extent to which the oil has been aged It cannot be completely vaporized after initial condensation from the vapor phase at 373 K or more, it rapidly reacts and eventually produces a solid residue from $50 wt.% of the original liquid It is chemically unstable, and the instability increases with heating It is always preferable to store the liquid at or below room temperature; changes do occur at room temperature, but much more slowly and they can be accommodated in a commercial application Causes unusual time-dependent behavior Properties such as viscosity increases, volatility decreases, phase separation, and deposition of gums change with time

Density Viscosity Distillation

Ageing of pyrolysis liquid

24 Table 9 Properties of the raw and upgraded bio-oil. Property Density Elemental analysis (wt.%) C H O N Higher heating value (MJ/kg) Raw bio-oil 1.12 60.4 6.9 41.8 0.9 21.3

A. Demirbas / Applied Energy 88 (2011) 1728

Upgraded bio-oil 0.93 87.7 8.9 3.0 0.4 41.4

to form water and an n-parafn, decarboxylation removes oxygen in the biomass carboxy groups to form carbon dioxide and a shorter n-parafn [109]. Properties of the raw and upgraded bio-oil are given in Table 10. As seen Table 10, properties of the raw and upgraded bio-oil are highly different.

6. Biocrude Biocrude has an oxygen content of 1020 wt.% and heating value of about 35 MJ/kg, which can be further improved by hydrodeoxygenation (HDO) to produce liquids similar to diesel and jet fuel. The process has a high heating requirement; hence, proper energy integration is needed for the commercial plants. Hydrothermal treatment can be used to liquefy biomass and increase energy density to produce biocrude. Hydrothermal processing offers a number of potential advantages over other biofuel production methods, including high throughputs, high energy and separation efciency, the ability to use mixed feedstocks like wastes and lignocellulose, the production of direct replacements for existing fuels, and no need to maintain specialized microbial cultures or enzymes [109]. Typical hydrothermal liquefaction conditions range from 550 to 650 K, 7 to 20 MPa with liquid water present and the reaction occurs for 1060 min. Both batch and continuous reactor schemes are available, with some challenges in continuous feeding of biomass. Recent advancements (e.g., twinscrew feeding) are likely to address some of the challenge. Compared with petroleum, the biocrude has its own characteristics, such as higher oxygen content, consisting of various types of molecules of wide-ranging molecular weight, and being in solid state at ambient temperature. Generally, hydrothermal medium refers to water that has been heated and compressed simultaneously. As water is heated, it continues to acquire an interesting set of properties. For example, density decreases, dielectric constant decreases, and ionic product rst increases and then decreases. The dielectric constant can play a key role in inuencing biomass reactions, as described by the Kirkwood relation. During a reaction, the transition state may be of lower or higher polarity than the initial state. A high dielectric medium lowers the activation energy of a reaction for a transition state of higher polarity than the initial state. As a consequence, many reactions have a high activation volume. By variation of the relative dielectric constant with temperature and pressure, reaction rates can be controlled [110,111]. Although ambient water solubilizes a limited number of polar compounds, near critical water (475575 K) dissolves both nonpolar organic molecules and inorganic salts and is comparable to that of the popular organic solvent acetone, mainly due to a reduced dielectric constant in the 2030 range [112], from that of 80 for ambient water, while keeping density high enough (0.70.8 g/ cm3). As the solubility of organics in water increases with the increase in temperature, the behavior is different with respect to inorganic salts. For example, the solubility of salts reaches a maximum at 575675 K; after that, the solubility drops very rapidly with an increase in temperature (e.g., the solubility of NaCl water is 40 wt.% at 575 K and 100 ppm at 725 K, 25.3 MPa). At supercritical temperatures, water practically has no dissolving power for salts due to a very low (near 2) dielectric constant. In addition, due to the dramatic drop in the dissociation constant, dissolved salts behave like weak electrolytes [113]. Additional useful properties of hydrothermal medium are the high molecular diffusion and low viscosity. Both give rise to efcient heat transfer and mass transfer for solid/liquid biomass liquefaction reactions. Diffusivity of molecules is not as high as in the

Table 10 Comparison of liquefaction and pyrolysis. Process Liquefaction Pyrolysis Temperature (K) 525600 650800 Pressure (MPa) 520 0.10.5 Drying Unnecessary Necessary

The main hydrodeoxygenation (HDO) reaction is represented in the following equation:

ACH2 OA H2 ! ACH2 A H2 O

This is the most important route of chemical upgrading. The reaction (Eq. (1)) has strong analogies with typical renery hydrogenations like hydrodesulfurization and hydrodenitrication. In general, most of the hydrodeoxygenation studies have been performed using existing hydrodesulfurization catalysts (NiMo and CoMo on suitable carriers). Such catalyst need activation using a suitable sulfur source and this is a major drawback when using nearly sulfur-free resources like bio-oil. Table 9 shows the upgrading processes of ash pyrolysis liquid or bio-oil. Hydroprocessing is the term used for the catalytic reactions of hydrogen with process streams to remove the heteroatoms sulfur, nitrogen, oxygen, and metals, and for the saturation of aromatic structures. Hydroprocessing can also refer to hydrogenation of olens or other unsaturated species. Hydroprocessing is carried out in xed-bed reactors at high pressures, with the severity of the reaction conditions generally increasing with the heaviness of the feed. Reactors used with the heavier feeds are trickle beds, in which the liquid oil trickles through the catalyst bed, with hydrogen concentrated in the gas phase. Slurry reactors have also been used for heavy feedstocks. The primary objective catalytic partial hydrodeoxygenation is to increase the energetic value of the oil by removing bound oxygen in the form of water. Hydrodeoxygenation of bio-oils involves treating the oils at moderate temperatures with high-pressure hydrogen in the presence of heterogeneous catalysts. The process was carried out in two distinct stages, a rst stage at relatively low temperatures (525575 K), aimed to stabilize the bio-oil and a second stage at higher temperatures (575675 K) to de-oxygenate the intermediate product. Different types of catalysts were screened, ranging from conventional sulded catalysts used in the HDS process (i.e. NiMo/Al2O3, CoMo/Al2O3) to novel non-sulded catalyst based on noble metal catalysts (i.e. Ru/Al2O3). Operating conditions were optimized to obtain the highest yield of a hydrocarbon like liquid product [108]. The hydrogen can be used to convert biomass to fuels, which requires a higher level of hydrodeoxygenation compared to the conversion of fossil feedstocks to fuels. Green diesel refers to an acceptable diesel pool blend component produced from a suitable biofeedstock. Green diesel can be produced by either hydrodeoxygenation or decarboxylation of plant oils and greases with propane as a co-product. Although hydrodeoxygenation removes oxygen from a triglyceride or free fatty acid by reaction with hydrogen

A. Demirbas / Applied Energy 88 (2011) 1728

25

gas phase but is much higher than in ambient water. Both of these properties can be used to reduce heating and mixing needs in the hydrothermal reactors. A combination of high diffusivity, low viscosity, and high miscibility can accelerate chemical reactions and improve reaction efciency. 6.1. Liquefaction mechanism Hydrothermal liquefaction of biomass using subcritical water has some potential advantages over other processes. Because water serves as both reaction medium and reactant, the process can use wet biomass without the need for energy-intensive drying. Moreover, biomass residue generated from many varieties of operations can conditions. Hence, supercritical water in this high temperature, low-density region is more suitable for free radical reactions and a poor medium for ionic reactions. The pyrolysis and direct liquefaction with water processes are sometimes confused with each other, and a simplied comparison of the two follows. Both are thermochemical processes in which feedstock organic compounds are converted into liquid products. In the case of liquefaction, feedstock macro-molecule compounds are decomposed into fragments of light molecules in the presence of a suitable catalyst. At the same time, these fragments, which are unstable and reactive, repolymerize into oily compounds having appropriate molecular weights [114]. With pyrolysis, on the other hand, a catalyst is usually unnecessary, and the light decomposed fragments are converted to oily compounds through homogeneous reactions in the gas phase. The differences in operating conditions for liquefaction and pyrolysis are shown in Table 9. Liquefaction processes result in liquid product, which can be easily stored and transported and require lower process temperatures. Due to these advantages, it is becoming increasingly evident that liquid products offer more potential for the production of biobased products than gas products and this is reected in the rapid development of these processes and the large amount of research in this area [115]. Appell et al. proposed the following mechanism for liquefaction of carbohydrate in the presence of sodium carbonate and carbon monoxide [116,117]. The liquefaction mechanism is summarized in the following four steps: 1. reaction of sodium carbonate and water with carbon monoxide, to yield sodium formate, 2. dehydration of vicinal hydroxyl groups in a carbohydrate to an enol, followed by isomerization to ketone, 3. reduction of newly formed carbonyl group to the corresponding alcohol with formate ion and water, and 4. the hydroxyl ion reacts with additional carbon monoxide to regenerate the formate ion. According to this mechanism, deoxygenation occurs through decarboxylation from ester formed by the hydroxyl group and formate ion derived from the carbonate. Liquefaction of biomass in subcritical water proceeds through a series of structural and chemical transformations [118120] that involve the following: 1. solvolysis of biomass resulting in micellar-like structure; 2. depolymerization of cellulose, hemicelluloses, and lignin; and 3. chemical and thermal decomposition of monomers to smaller molecules. Due to their low costs and high activity, alkali salts are the most used catalyst in hydrothermal liquefaction. Formate ions derived from the alkali carbonates react with the hydroxyl group of biomass to cause decarboxylation, which forms esters. Thus, alkali

carbonates catalyze hydrolysis of biomass macro-molecules into smaller fragments. The micellar-like, broken-down fragments produced are then degraded to smaller compounds by dehydration, dehydrogenation, deoxygenation, and decarboxylation. These compounds further rearrange through condensation, cyclization, and polymerization, leading to newer compounds [121]. However, with the choice of an alkali, there are some differences in activity. For example, Karagoz et al. [122] examined the effect of various alkali salts on the yield of biocrude from the hydrothermal liquefaction of pine wood. Based on the conversion and yield of liquid products, the catalytic activity ranks as: K2CO3 > KOH > Na2CO3 > NaOH. In noncatalytic liquefaction, the solid residue was about 42 wt.%, whereas 0.94 M K2CO3 gave a more complete conversion with residue of only 4.0 wt.%. In noncatalytic experiments, furan derivatives were observed, whereas the catalytic runs produced mainly phenolic compounds. Additional catalysts, including various organic and inorganic acids (H2SO4, HCl, acetic acid), metal ions (Zn2+, Ni2+, Co2+, and Cr3+), salt and bases [CaCO3, Ca(OH)2, HCOONa, and HCOOK], and CO2 have been tested with varying degrees of success [123]. For example, Minowa et al. [124,125] compared sodium carbonate with reduced-nickel catalyst in the 475625 K temperature range, to conclude the role of alkali catalyst in inhibiting the formation of char from biocrude (stabilization of biocrude), resulting in more biocrude production. On the other hand, the nickel catalyst catalyzed the steam reforming reaction of aqueous products as intermediates and the methanation reaction [124,125]. Reducing gases, such as CO and H2, have been investigated in the early liquefaction experiments, in hopes of further reducing the oxygen content of biocrude [126,127]. But recently, experiments have shown that the effect is very minor [128]. On the other hand, intentional addition of CO2 to the reactor causes a noticeable increase in the oxygen content of biocrude. 6.2. Properties of biocrude Biocrude is solid at room temperature and is typically liquid at temperatures above 353 K. A large portion is chloroform (or acetone)-soluble, with a small amount of insoluble material of high molecular weight. Oxygen content is typically 1020 wt.% and is furanic or aromatic in nature. The heating value is in the range 3036 MJ/kg. In hydrothermal liquefaction, a small content of organics and heating value remains in the aqueous phase which should be used or recycled for full recovery. Biocrude contains a mixture of a large number of molecules with molecular weight as large as 2000. These components can be classied into organic acids, alcohols, aldehydes, ketones, aromatic hydrocarbons, aliphatic hydrocarbons, furans, and phenols. Some of the prominent compounds in the biocrude from corn stover are phenol, guaiacol, 4-ethyl-phenol, 2-methoxy-4-methyl-phenol, 4-ethylguaiacol, 2,6-dimethoxyphenol, 1,2,4-trimethoxybenzene, 5-tert-butylpyrogallol, 1,10-propylidenebis-benzene, 1-(4-hydroxy-3,5-dimethoxyphenyl)-ethanone, acetic acid, 1-hydroxy-2-propanone, furfural, 3-methyl-2-cyclopenten-1-one, 2,5-hexanedione, and desaspidinol [129]. Typically, the biocrude distillate contains 1923% aromatic hydrocarbons, 1135% phenols, 913% naphthenes, 511% aliphatic hydrocarbons, 714% aldehydes and ketons, and 310% alcohols, on mass basis [130]. Viscosity of biocrude varies greatly with the liquefaction conditions. For biocrude from the batch liquefaction with 5:1 water-tobiomass ratio and use of 10 wt.% (of biomass) sodium carbonate catalyst, the viscosity at 298 K is 313 106 cP, which increased to 323 106 cP upon storage for a month. The high and increasing viscosity indicates a poor ow characteristic and stability. The increase in the viscosity can be attributed to the continuing

26

A. Demirbas / Applied Energy 88 (2011) 1728

polymerization and oxidative coupling reactions in the biocrude upon storage. Although stability of biocrude is typically better than that of bio-oil, the viscosity of biocrude is much higher [131]. 6.3. Renement and upgrading of biocrude The two key properties of biocrude that differ from the petroleum crude are high oxygen content and molecular weight (or viscosity). For compatibility with the petroleum liquids, biocrude needs to be upgraded by hydrogen treatment. The addition of hydrogen to the biocrude molecules should be performed not to saturate the aromatic rings, but to remove oxygen and hydrocrack molecules. Hence, the hydrodeoxygenation (HDO) process is used, which is somewhat different than the hydro-treating used in the petroleum industry to remove nitrogen and sulfur. For HDO, hydrogen and biocrude are fed to a catalytic reactor maintained at about at 525675 K and 1018 MPa. Various catalysts have been successfully tested, including CoMo, NiMo, NiW, Ni, Co, Pd, and CuCrO, on a variety of supports, such as Al2O3, c-Al2O3, SiO2Al2O3, and Y-zeolite/Al2O3. In a catalytic HDO of biocrude, oxygen content was reduced from 10 to <0.1 wt.% with hydrogen consumption of 4 wt.% on intake and C1C4 gas production of only 2.4 wt.% [132]. The product included liquids similar to diesel, jet fuel, and a small portion of naphtha, with a cetane number of >50. Out of various catalyst choices, a promising candidate is the sulded form of CoMo catalyst due to its high activity. However, the product may become contaminated with sulfur, and the catalyst may get deactivated by coke deposition, and, mostly notably, may get poisoned by trace amounts of water present [133,134]. Recent development in the reductive upgrading of biocrude offers an attractive alternative route, which allows upgrading in the aqueous medium [135]. In fact, the separation from aqueous media would also be facile due to the nonmiscibility of hydrocarbons upon cooling to ambient temperature. In aqueous HDO [136], biocrude is added to acidic aqueous medium along with Pd/C catalyst, and hydrogen is fed to the reactor. Typically, the reaction is carried out at 475525 K for 30 min at 5 MPa. The oxygen is removed as H2O, CH3OH, and other oxygenates. A high conversion (about 100%) and high selectivity toward hydrocarbons are achieved. The hydrothermal liquefaction requires the use of high pressures, which necessitates the design of special reactors and separators, requiring a high capital investment. Despite several key benets, the widespread commercial use of hydrothermal technology has been hindered due to the high capital cost and a few critical issues that need to be resolved [42]. 7. Conclusion It is well known that transport is almost totally dependent on fossil particularly petroleum based fuels such as gasoline, diesel fuel, liqueed petroleum gas, and compressed natural gas. Of the special concerns are the liquid fuels for use in the automobiles. Hence, there is a widespread recent interest in learning more about obtaining liquid fuels from non-fossil sources. The combination of rising oil prices, issues of security, climate instability and pollution, and deepening poverty in rural and agricultural areas, is propelling governments to enact powerful incentives for the use of these fuels, which is in turn sparking investment. In fact, the world is on the verge of an unprecedented increase in the production and use of biofuels for transport. There are two global liquid transportation biofuels that might replace gasoline and diesel fuel; these are bioethanol and biodiesel, respectively. Transport is one of the main energy consuming sec-

tors. It is assumed that biodiesel is used as a petroleum diesel replacement and that bioethanol is used as a gasoline replacement. Biofuels production costs can vary widely by feedstock, conversion process, scale of production and region. The major economic factor to consider for input costs of biodiesel production is the feedstock, which is about 7580% of the total operating cost. Other important costs are labor, methanol and catalyst, which must be added to the feedstock. On an energy basis, ethanol is currently more expensive to produce than gasoline in all regions considered. Only ethanol produced in Brazil comes close to competing with gasoline. Ethanol produced from corn in the US is considerably more expensive than from sugar cane in Brazil, and ethanol from grain and sugar beet in Europe is even more expensive. The simultaneous production of biomethanol (from sugar juice) in parallel to the production of bioethanol, appears economically attractive in locations where hydroelectricity is available at very low cost. Production costs for ethanol are much lower in countries with a warm climate, with Brazil probably the lowest-cost producer in the world. Production costs in Brazil, using sugar cane as the feedstock, have recently been recorded at less than half the costs in Europe. The generally larger US conversion plants produce biofuels, particularly ethanol, at lower cost than plants in Europe. The production liquid fuels for the transportation sector is of importance for continued vitality of our industrialized society. Research is being conducted worldwide to develop new technologies for the generation of liquid fuels from renewable resources. Currently, biodiesel is produced by transesterication of vegetable oils and animal fats, and ethanol by fermentation of glucose. Processes for the efcient pyrolysis and liquefaction of biomass to produce bio-oil and bicrude are being developed. Bio-oils are liquid fuels made from biomass materials, such as agricultural crops, municipal wastes and agricultural and forestry by-products via thermochemical processes. The bio-oil from wood is typically a liquid, almost black through dark red brown. The density of the liquid is about 1200 kg/m3, which is higher than that of fuel oil and signicantly higher than that of the original biomass. The bio-oils have water contents of typically 1433 wt.%, which cannot be removed by conventional methods like distillation. Phase separation may occur above certain water contents. The higher heating value (HHV) is below 27 MJ/kg (compared to 43 46 MJ/kg for conventional fuel oils). Compared with petroleum, the biocrude has its own characteristics, such as higher oxygen content, consisting of various types of molecules of wide-ranging molecular weight, and being in solid state at ambient temperature. Biocrude has an oxygen content of 1020 wt.% and heating value of about 35 MJ/kg, which can be further improved by HDO to produce liquids similar to diesel and jet fuel. References
[1] Demirbas A. Political, economic and environmental impacts of biofuels: a review. Appl Energy 2009;86:S10817. [2] Demirbas AH. Biofuels for future transportation necessity. Energy Educ Sci Technol Part A 2010;26:1323. [3] Kecebas A, Alkan MA. Educational and consciousness-raising movements for renewable energy in Turkey. Energy Educ Sci Technol Part B 2009;1:15770. [4] Demirbas A. Energy concept and energy education. Energy Educ Sci Technol Part B 2009;1:85101. [5] Kurnaz MA, Calik M. A thematic review of energy teaching studies: focuses, needs, methods, general knowledge claims and implications. Energy Educ Sci Technol Part B 2009;1:126. [6] Demirbas A. Concept of energy conversion in engineering education. Energy Educ Sci Technol Part B 2009;1:18397. [7] Hertel TW, Tyner WE, Birur DP. The global impacts of biofuel mandates. Energy J 2010;31:75100. [8] Cardak O. The determination of the knowledge level of science students on energy ow through a word association test. Energy Educ Sci Technol Part B 2009;1:13955.

A. Demirbas / Applied Energy 88 (2011) 1728 [9] Demirbas A. Social, economic, environmental and policy aspects of biofuels. Energy Educ Sci Technol Part B 2009;1:75109. [10] Tatli ZH. Computer based education: online learning and teaching facilities. Energy Educ Sci Technol Part B 2009;1:17281. [11] Kan A. General characteristics of waste management: a review. Energy Educ Sci Technol Part A 2009;23:5569. [12] Demirbas A. Bioreneries: current activities and future developments. Energy Convers Manage 2009;50:2782801. [13] Demirbas A. Biofuels securing the planets future energy needs. Energy Convers Manage 2009;50:223949. [14] Demirbas AH. Inexpensive oil and fats feedstocks for production of biodiesel. Energy Educ Sci Technol Part A 2009;23:113. [15] Demirbas A. Biofuels sources, biofuel policy, biofuel economy and global biofuel projections. Energy Convers Manage 2008;49:210616. [16] Caglar A. Valorization of tea wastes by pyrolysis. Energy Educ Sci Technol Part A 2009;23:13544. [17] Demirbas A. Biodiesel from waste cooking oil via base-catalytic and supercritical methanol transesterication. Energy Convers Manage 2009;50:9237. [18] Balat M. Progress in biogas production processes. Energy Educ Sci Technol 2008;22:1536. [19] Saidur R. Energy, economics and environmental analysis for chillers in ofce buildings. Energy Educ Sci Technol Part A 2010;25:116. [20] Demirbas A. Progress and recent trends in biodiesel fuels. Energy Convers Manage 2009;50:1434. [21] Keskin A, Emiroglu AO. Catalytic reduction techniques for post-combustion diesel engine exhaust emissions. Energy Educ Sci Technol Part A 2010;25:87103. [22] Demirbas MF. Bioreneries for biofuel upgrading: a critical review. Appl Energy 2009;86:S15161. [23] Balat M, Balat H. Recent trends in global production and utilization of bioethanol fuel. Appl Energy 2009;86:227382. [24] Cerci Y, Ekin AB. Performance evaluation of a 250 L household refrigerator. Energy Educ Sci Technol Part A 2010;26:2536. [25] Mahlia TMI, Saidur R, Husnawan M, Masjuki HH, Kalam MA. An approach to estimate the life-cycle cost of energy efciency improvement of room air conditioners. Energy Educ Sci Technol Part A 2010;26:111. [26] TRS (The Royal Society). Sustainable biofuels: prospects and challenges. Policy document 01/08, London, January 14; 2008. [27] Balat H. Prospects of biofuels for a sustainable energy future: a critical assessment. Energy Educ Sci Technol Part A 2010;24:85111. [28] Demirbas B. Biomass business and operating. Energy Educ Sci Technol Part A 2010;26:3747. [29] Yenikaya C, Yaman H, Atar N, Erdogan Y, Colak F. Biomass resources and decolorization of acidic dyes from aqueous solutions by biomass biosorption. Energy Educ Sci Technol Part A 2009;24:113. [30] Demirbas MF. Bio-oils from corn stover via supercritical water liquefaction. Energy Educ Sci Technol Part A 2009;23:97104. [31] Ozkurt I. Qualifying of safower and algae for energy. Energy Educ Sci Thecnol Part A 2009;23:14551. [32] Demirbas MF. Microalgae as a feedstock for biodiesel. Energy Educ Sci Technol Part A 2010;25:3143. [33] Demirbas A. Biodegradability of biodiesel and petrodiesel fuels. Energy Sour Part A 2009;3:16974. [34] Demirbas A. Production of biodiesel from algae oils. Energy Sour Part A 2009;31:1638. [35] Sigar CP, Soni SL, Mathur J, Sharma D. Performance and emission characteristics of vegetable oil as diesel fuel extender. Energy Sour Part A 2009;31:13948. [36] Bozkurt H. Solar energy: solar semiconductor devices and photoelectrochemical cells. Future Energy Sour 2010;2:87147. [37] Ma F, Hanna MA. Biodiesel production: a review. Biores Technol 1999;70:115. [38] Zhang Y, Dub MA, McLean DD, Kates M. Biodiesel production from waste cooking oil: 2. Economic assessment and sensitivity analysis. Biores Technol 2003;90:22940. [39] Ertas M, Alma MH. Slow pyrolysis of chinaberry (Melia azedarach L.) seeds: characterization of bio-oils and bio-chars. Future Energy Sour 2010;2:14970. [40] Berl E. Production of oil from plant material. Science 1994;99:30912. [41] Aitani AM. Oil rening and products. In: Cleveland CJ, editor. Encyclopedia of energy. New York (NY): Elsevier; 2004. p. 15729. [42] Peterson AA, Vogel F, Lachance RP, Froling M, Antal Jr MJ, Tester JW. Thermochemical biofuel production in hydrothermal media: a review of suband supercritical water technologies. Energy Environ Sci 2008;1:3265. [43] Lede J. Solar thermochemical conversion of biomass. Solar Energy 1998;65:313. [44] Ucar A. The environmental impact of optimum insulation thickness for external walls and at roofs of building in Turkeys different degree-day regions. Energy Educ Sci Technol Part A 2009;24:4969. [45] Sevim C. Rapid climate change problem and wind energy investments for Turkey. Energy Educ Sci Technol Part A 2010;25:5967. [46] Akpinar EK, Akpinar S. Modelling of weather characteristics and wind power density in Elazig-Turkey. Energy Educ Sci Technol Part A 2010;25:4557. [47] Rajamohan P, Rajasekhar RVJ, Shanmugan S, Ramanathan K. Energy and economic evaluation of xed focus type solar parabolic concentrator for

27

[48] [49]

[50]

[51]

[52]

[53]

[54] [55] [56] [57] [58]

[59] [60]

[61]

[62]

[63] [64] [65] [66] [67] [68] [69]

[70]

[71]

[72]

[73]

[74] [75]

[76] [77] [78]

[79]

[80]

community cooking applications. Energy Educ Sci Technol Part A 2010;26:4959. Saidur R, Lai YK. Parasitic energy savings in engines using nanolubricants. Energy Educ Sci Technol Part A 2010;26:6174. Ralegaonkar RV, Gupta R. Application of passive solar architecture for intelligent building construction: a review. Energy Educ Sci Technol Part A 2010;26:7585. Bugutekin A, Yilmaz M, Kentli A, Isikan MO. A mathematical model for condensation of bubbles injected through an orice into subcooled water. Energy Educ Sci Technol Part A 2010;24:15171. Kecebas A, Kayfeci M. Effect on optimum insulation thickness, cost and saving of storage design temperature in cold storage in Turkey. Energy Educ Sci Technol Part A 2010;25:11727. Cerci Y. Experimental investigation of capacitor effects on performance parameters planning for household refrigerator and energy systems. Energy Educ Sci Technol Part A 2009;24:1524. Bolukbasi A, Comakli K, Sahin S. Domestic energy savings: investigation of optimum insulation thicknesses for the external wall of rural houses in Turkey. Energy Educ Sci Technol Part A 2009;24:2537. Darici B, Ocal FM. The structure of European nancial system and nancial integration. Energy Educ Sci Technol Part B 2010;2:13345. Alodali MFB. E-multicipality application in Turkey and Sakarya metropolitan city. Energy Educ Sci Technol Part B 2010;2:21124. Yan J, Lin T. Biofuels in Asia. Appl Energy 2009;86:S1S10. Matsumoto N, Sano D, Elder M. Biofuel initiatives in Japan: strategies, policies, and future potential. Appl Energy 2009;86:S6976. UN (United Nations). The emerging biofuels market: regulatory, trade and development implications. United Nations conference on trade and development, New York and Geneva; 2006. Hammond GP, Kallu S, McManus MC. Development of biofuels for the UK automotive market. Appl Energy 2008;85:50615. Tian Y, Zhao L, Meng H, Sun L, Yan J. Estimation of un-used land potential for biofuels development in (the) Peoples Republic of China. Appl Energy 2009;86:S7785. Yang J, Huang J, Qiu H, Rozelle S, Sombilla MA. Biofuels and the Greater Mekong Subregion: assessing the impact on prices, production and trade. Appl Energy 2009;86:S3746. Malik US, Ahmed M, Sombilla MA, Cueno SL. Biofuels production for smallholder producers in the Greater Mekong Sub-region. Appl Energy 2009;86:S5868. Zhou A, Thomson E. The development of biofuels in Asia. Appl Energy 2009;86:S1120. Phalan B. The social and environmental impacts of biofuels in Asia: an overview. Appl Energy 2009;86:S219. Prabhakar SVRK, Elder M. Biofuels and resource use efciency in developing Asia: back to basics. Appl Energy 2009;86:S306. Balat H. Social, political and economic impacts of biodiesel. Social Polit Econ Cult Res 2009;1:89125. Leung Dennis YC, Wu X, Leung MKH. A review on biodiesel production using catalyzed transesterication. Appl Energy 2010;87:108395. Huang G, Chen F, Wei D, Zhang X, Chen G. Biodiesel production by microalgal biotechnology. Appl Energy 2010;87:3846. Li Q, Yan Y. Production of biodiesel catalyzed by immobilized Pseudomonas cepacia lipase from Sapium sebiferum oil in micro-aqueous phase. Appl Energy 2010;87:314854. Thamsiriroj T, Murphy JD. Is it better to import palm oil from Thailand to produce biodiesel in Ireland than to produce biodiesel from indigenous Irish rape seed? Appl Energy 2009;86:595604. Qi DH, Chen H, Geng LM, Bian YZH, Ren XCH. Performance and combustion characteristics of biodieseldieselmethanol blend fuelled engine. Appl Energy 2010;87:167986. Balo F. Energy and economic analyses of insulated exterior walls for four different city in Turkey. Energy Educ Sci Technol Part A 2011;26:17588. Gao C, Zhai Y, Ding Y, Wu Q. Application of sweet sorghum for biodiesel production by heterotrophic microalga Chlorella protothecoides. Appl Energy 2010;87:75661. Leduc S, Natarajan K, Dotzauer E, McCallum I, Obersteiner M. Optimizing biodiesel production in India. Appl Energy 2009;86:S12531. Yee KF, Tan KT, Abdullah AZ, Lee KT. Life cycle assessment of palm biodiesel: revealing facts and benets for sustainability. Appl Energy 2009;86:S189196. Leung DYC, Wu X, Leung MKH. A review on biodiesel production using catalyzed transesterication. Appl Energy 2010;87:108395. Balat M, Balat H. Progress in biodiesel processing. Appl Energy 2010;87:03062619. Ganapathy T, Murugesan K, Gakkhar RP. Performance optimization of Jatropha biodiesel engine model using Taguchi approach. Appl Energy 2010;86:247686. Shu Q, Gao J, Nawaz Z, Liao Y, Wang D, Wang J. Synthesis of biodiesel from waste vegetable oil with large amounts of free fatty acids using a carbon-based solid acid catalyst. Appl Energy 2010;87: 258996. Wen Z, Yu X, Tu S-T, Yan J, Dahlquist E. Synthesis of biodiesel from vegetable oil with methanol catalyzed by Li-doped magnesium oxide catalysts. Appl Energy 2010;87:7438.

28

A. Demirbas / Applied Energy 88 (2011) 1728 [109] Kaparaju P, Serrano M, Thomsen AB, Kongjan P, Angelidaki I. Bioethanol, biohydrogen and biogas production from wheat straw in a biorenery concept. Biores Technol 2009;100:25628. [110] Kubiatko M, Vaculov I. Project-based learning: characteristic and the experiences with application in the science subjects. Energy Educ Sci Technol Part B 2011;3:6574. [111] Phillip ES. Organic chemical reactions in supercritical water. Chem Rev 1999;99:60321. [112] Yamaguchi T. Structure of subcritical and supercritical hydrogen-bonded liquids and solutions. J Mol Liq 1998;78:4350. [113] Ragauskas AJ, Williams CK, Davison BH, Britovsek G, Cairney J, Eckert CA, et al. The path forward for biofuels and biomaterials. Science 2006;311:4849. [114] Yesodharan S. Supercritical water oxidation: an environmentally safe method for the disposal of organic wastes. Curr Sci 2002;82:111222. [115] Molten PM, Demmitt TF, Donovan JM, Miller RK. Mechanism of conversion of cellulose wastes to liquid in alkaline solution. In: Klass DL, editor. Energy from biomass and wastes III. Chicago (IL): Institute of Gas Technology; 1983. [116] Appell HR. In: Anderson L, Tilman DA, editors. New York: Academic Press; 1983. [117] Appell HR, Fu YC, Friedman S, Yavorsky PM, Wender I. Converting organic wastes to oil. US Burea of Mines report of investigation. No. 7560; 1971. [118] Chornet E, Overend RP. Biomass liquefaction: an overview. In: Overend RP, Milne TA, Mudge LK, editors. Fundamentals of thermochemical biomass conversion. New York (NY): Elsevier Applied Science; 1983. [119] Saidur R, Lai YK. Nanotechnology in vehicles weight reduction and associated energy savings. Energy Educ Sci Technol Part A 2011;26:87101. [120] Kumar S, Gupta RB. Biocrude production from switchgrass using subcritical water. Energy Fuel 2009;23:51519. [121] Demirbas A. Mechanisms of liquefaction and pyrolysis reactions of biomass. Energy Convers Manage 2000;41:63346. [122] Karagoz S, Bhaskar T, Muto A, Sakata Y, Oshiki T, Kishimoto T. Low temperature catalytic hydrothermal treatment of wood biomass: analysis of liquid products. Chem Eng J 2005;108:12737. [123] Miyazawa T, Funazukuri T. Polysaccharide hydrolysis accelerated by adding carbon dioxide under hydrothermal conditions. Biotechnol Prog 2005;21:17825. [124] Minowa T, Zhen F, Ogi T. Cellulose decomposition in hot-compressed water with alkali or nickel catalyst. J Supercrit Fluid 1998;13:2539. [125] Minowa T, Zhen F, Ogi T. Liquefaction of cellulose in hot-compressed water using sodium carbonate: production distribution at different reaction temperature. J Chem Eng Jpn 1999;30:18690. [126] Appell HR. The production of oil from wood waste. In: Anderson L, Tilman DA, editors. Fuels from waste. New York (NY): Academic Press; 1977. [127] Davis H, Figueroa C, Schaleger L. Hydrogen or carbon monoxide in the liquefaction of biomass. Adv Hydrogen Energy 1982;3:84962. [128] He BJ, Zhang Y, Yin Y, Funk TL, Riskowski GL. Effects of alternative process gases on the thermochemical conversion process of swine manure. Trans ASAE 2001;44:187380. [129] Zhang B, Keitz M, Valentas K. Thermal effects on hydrothermal biomass liquefaction. Appl Biochem Biotechnol 2008;147:14350. [130] Adjaye JD, Sharma RK, Bakhshi NN. Characterization and stability analysis of wood-derived bio-oil. Fuel Process Technol 1992;31:24156. [131] Elliott DC, Schiefelbein GF. Liquid hydrocarbon fuels from biomass. Am Chem Soc Div Fuel Chem Preprints 1989;34:11606. [132] Goudriaan F, Peferoen DGR. Liquid fuels from biomass via a hydrothermal process. Chem Eng Sci 1990;45:272934. [133] Laurent E, Delmon B. Inuence of water in the deactivation of a sulded NiMo/c-A12O3 catalyst during hydrodexygenation. J Catal 1994;146:28491. [134] Elliott DC. Historical developments in hydroprocessing bio-oils. Energy Fuel 2007;21:1792815. [135] Diaz E, Mohedano AF, Calvo L, Gilarranz MA, Casas JA, Rodriguez JJ. Hydrogenation of phenol in aqueous phase with palladium on activated carbon catalysts. Chem Eng J 2007;131:6571. [136] Zhao C, Kou Y, Lemonidou AA, Li XB, Lercher JA. Highly selective catalytic conversion of phenolic bio-oil to alkanes. Angew Chem Int Ed Engl 2009;48:398790. S3987/16.

[81] Tan RR, Foo DCY, Aviso KB, Ng DKS. The use of graphical pinch analysis for visualizing water footprint constraints in biofuel production. Appl Energy 2009;86:6059. [82] Divakara BN, Upadhyaya HD, Wani SP, Gowda CLL. Biology and genetic improvement of Jatropha curcas L.: a review. Appl Energy 2010;87:73242. [83] Hazar H, Aydin H. Performance and emission evaluation of a CI engine fueled with preheated raw rapeseed oil (RRO)-diesel blends. Appl Energy 2010;87:78690. [84] Qiu H, Huang J, Yang J, Rozelle S, Zhang Y, Zhang Y, et al. Bioethanol development in China and the potential impacts on its agricultural economy. Appl Energy 2010;87:7683. [85] Yu S, Tao J. Economic, energy and environmental evaluations of biomassbased fuel ethanol projects based on life cycle assessment and simulation. Appl Energy 2009;86:S178188. [86] Taleghani M, Ansari HR, Jennings P. Renewable energy education in sustainable architecture: lessons from developed and developing countries. Energy Educ Sci Technol Part B 2010;2:11131. [87] Demirbas A. Biofuels: securing the planets future energy need. London: Springer; 2009. [88] Demirbas A. Bioreneries for biomass upgrading facilities. London: Springer; 2009. [89] Sahin Y. Environmental impacts of biofuels. Energy Educ Sci Technol Part A 2011;26:12942. [90] Ertas M, Alma MH. Slow pyrolysis of chinaberry (Melia azedarach L.) seeds: part I. The inuence of pyrolysis parameters on the product yields. Energy Educ Sci Technol Part A 2011;26:14354. [91] Yumurtaci M, Kecebas A. Renewable energy and its university level education in Turkey. Energy Educ Sci Technol Part B 2011;3:14352. [92] Demirbas MF, Balat M. Recent advances on the production and utilization trends of bio-fuels: a global perspective. Energy Convers Manage 2006;47:237181. [93] Agranet FO. Lichts world ethanol & biofuels report; 2009. <http://www. agra-net.com/portal2/home.jsp?template=productpage&pubid=ag072>. [94] Demirbas A. Progress and recent trends in biofuels. Prog Energy Combust Sci 2007;33:118. [95] Altun S. Fuel properties of biodiesels produced from different feedstocks. Energy Educ Sci Technol Part A 2011;26:16574. [96] Babu BV, Chaurasia AS. Modeling for pyrolysis of solid particle: kinetics and heat transfer effects. Energy Convers Manage 2003;44:225175. [97] Kumar S, Gupta RB. Hydrolysis of microcrystalline cellulose in sub- and supercritical water in a continuous ow reactor. Ind Eng Chem Res 2008;47:93219. [98] Bridgwater AV, Meier D, Radlein D. An overview of fast pyrolysis of biomass. Org Geochem 1999;30:147993. [99] Elliot DC, Beckman D, Bridgewater AV, Diebold JP, Gevert SB, Solantausta Y. Developments in direct thermochemical liquefaction of biomass: 19831990. Energy Fuels 1991;5:399410. [100] Tran DQ, Charanjit R. A kinetic model for pyrolysis of Douglas r bark. Fuel 1978;57:2938. [101] Bridgwater AV. Renewable fuels and chemicals by thermal processing of biomass. Chem Eng J 2003;91:87102. [102] Mohan D, Pittman CU, Steele PH. Pyrolysis of wood/biomass for bio-oil: a critical review. Energy Fuels 2006;20:84889. [103] Goyal HB, Seal D, Saxena RC. Biofuels from thermochemical conversion of renewable resources: a review. Renew Sustain Energy Rev 2008;12:50417. [104] Demirbas A. Fuel conversional aspects of palm oil and sunower oil. Energy Sour 2001;25:45766. [105] Demirbas A. The inuence of temperature on the yields of compounds existing in bio-oils obtained from biomass samples via pyrolysis. Fuel Proc Technol 2007;88:5917. [106] Arpa O, Yumrutas R, Demirbas A. Production of diesel-like fuel from waste engine oil by pyrolitic distillation. Appl Energy 2010;87:1227. [107] Zhang SP, Yan YJ, Ren JW, Li TC. Study of hydrodeoxygenation of bio-oil from the fast pyrolysis of biomass. Energy Sour 2003;25:5765. [108] de la Puente G, Gil A, Pis JJ, Grange P. Effects of support surface chemistry in hydrodeoxygenation reactions over CoMo/activated carbon sulded catalysts. Langmuir 1999;15:58006.

S-ar putea să vă placă și