You are on page 1of 18

Energy Conversion and Management 52 (2011) 858875

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Production of bioethanol from lignocellulosic materials via the biochemical pathway: A review
Mustafa Balat
Sila Science & Energy Company, University Mah, Trabzon, Turkey

a r t i c l e

i n f o

a b s t r a c t
Bioethanol is by far the most widely used biofuel for transportation worldwide. Production of bioethanol from biomass is one way to reduce both consumption of crude oil and environmental pollution. Bioethanol can be produced from different kinds of raw materials. These raw materials are classied into three categories of agricultural raw materials: simple sugars, starch and lignocellulose. The price of the raw materials is highly volatile, which can highly affect the production costs of the bioethanol. One major problem with bioethanol production is the availability of raw materials for the production. Lignocellulosic biomass is the most promising feedstock considering its great availability and low cost, but the large-scale commercial production of fuel bioethanol from lignocellulosic materials has still not been implemented. 2010 Elsevier Ltd. All rights reserved.

Article history: Received 1 January 2010 Accepted 15 August 2010 Available online 6 September 2010 Keywords: Bioethanol Lignocellulosic material Pretreatment Enzymatic hydrolysis Fermentation

1. Introduction Today, the transportation sector worldwide is almost entirely dependent on petroleum-based fuels. It is responsible for 60% of the world oil consumption [1]. In addition, transportation sector accounts for more than 70% of global carbon monoxide (CO) emissions and 19% of global carbon dioxide (CO2) emissions [2]. CO2 emissions from a gallon of gasoline are about 8 kg [3]. Around the world, there were about 806 million cars and light trucks on the road in 2007 [4]. These numbers are projected to increase to 1.3 billion by 2030 and to over 2 billion vehicles by 2050 [5]. This growth will affect the stability of ecosystems and global climate as well as global oil reserves. The dramatic increase in the price of petroleum, the nite nature of fossil fuels, increasing concerns regarding environmental impact, especially related to greenhouse gas (GHG) emissions, and health and safety considerations are forcing the search for new energy sources and alternative ways to power the worlds motor vehicles. An alternative fuel must be technically feasible, economically competitive, environmentally acceptable, and readily available [6]. Numerous potential alternative fuels have been proposed, including bioethanol, biodiesel, methanol, hydrogen, boron, natural gas, liqueed petroleum gas (LPG), FischerTropsch fuel, p-series, electricity, and solar fuels. Biomass-based fuels, also known as biofuels offer many advantages over petroleum-based fuels [7]: (1) biofuels are easily available from common biomass sources, (2) they are represent a CO2-cycle in combustion, (3) biofuels have a considerable environ Tel.: +90 462 871 3025; fax: +90 462 871 3110.
E-mail address: mustafabalat@yahoo.com 0196-8904/$ - see front matter 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.enconman.2010.08.013

mentally friendly potential, (4) there are many benets the environment, economy and consumers in using biofuels, and (5) they are biodegradable and contribute to sustainability. The major benets of biofuels are given in Table 1. The use of biofuels can contribute to the mitigation of GHG emissions, provide a clean and therefore sustainable energy source, and increase the agricultural income for rural poor in developing countries. Developing countries have a comparative advantage for biofuel production because of greater availability of land, favorable climatic conditions for agriculture and lower labour costs. However, there may be other socio-economic and environmental implications affecting the potential for developing countries to benet from the increased global demand for biofuel [8]. Large-scale production of biofuels offers an opportunity for certain developing countries to reduce their dependence on oil imports. In developed countries there is a growing trend towards employing modern technologies and efcient bioenergy conversion using a range of biofuels, which are becoming cost-wise competitive with fossil fuels [9]. Biofuels are made from bio-based materials through thermochemical processes such as pyrolysis [10,11], gasication [12,13], liquefaction [14], supercritical uid extraction [15], supercritical water liquefaction [16] and biochemical [17]. Thermo-chemical reforming of biomass concerns the processes of catalytic and non-catalytic pyrolysis as well as the gasication, which aims at the maximization of the production of energetically exploitable liquid and gaseous products. Biofuels include bioethanol, biomethanol, vegetable oils, biodiesel, biogas, biosynthetic gas (bio-syngas), bio-oil, bio-char, FischerTropsch liquids, and biohydrogen. The term biofuels can refer to fuels for direct combustion for electricity production, but

M. Balat / Energy Conversion and Management 52 (2011) 858875 Table 1 Major benets of biofuels. Economic impacts Sustainability Fuel diversity Increased number of rural manufacturing jobs Increased income taxes Increased investments in plant and equipment Agricultural development International competitiveness Reducing the dependency on imported petroleum Greenhouse gas reductions Reducing of air pollution Biodegradability Higher combustion efciency Improved land and water use Carbon sequestration Domestic targets Supply reliability Reducing use of fossil fuels Ready availability Domestic distribution Renewability

859

Environmental impacts

Energy security

is generally used for liquid fuels for transportation sector [18]. Renewable liquid biofuels for transportation have recently attracted huge attention in different countries all over the world because of its renewability, sustainability, common availability, regional development, rural manufacturing jobs, reduction of GHG emissions, and its biodegradability [19]. Bioethanol is by far the most widely used biofuel for transportation worldwide. Bioethanol and bioethanol/gasoline blends have a long history as alternative transportation fuels. It has been used in Germany and France as early as 1894 by the then incipient industry of internal combustion engines (ICEs) [20]. Brazil has utilized bioethanol as a fuel since 1925. By that time, the production of bioethanol was 70 times bigger than the production and consumption of petrol [21]. The use of bioethanol for fuel was widespread in Europe and the United States until the early 1900s. Because it became more expensive to produce than petroleumbased fuel, especially after World War II, bioethanols potential was largely ignored until the oil crisis of the 1970s [22]. Since the 1980s, there has been an increased interest in the use of bioethanol as an alternative transportation fuel. To ensure that good bioethanol is produced, with reference to GHG benets, the following demands must be met [23]: (1) bioethanol plants should use biomass and not fossil fuels, (2) cultivation of annual feedstock crops should be avoided on land rich in carbon (above and below ground), such as peat soils used as permanent grassland, (3) by-products should be utilized efciently in order to maximize their energy and GHG benets, and (4) nitrous oxide emissions should be kept to a minimum by means of efcient fertilization strategies, and the commercial nitrogen fertilizer utilized should be produced in plants which have nitrous oxide gas cleaning. Bioethanol is a fuel derived from renewable sources of feedstock; typically plants such as wheat, sugar beet, corn, straw, and wood. Bioethanol is an alternative fuel that is produced almost entirely from food crops. It represents an important, renewable liquid fuel for motor vehicles. Producing bioethanol as a transportation fuel can help reduce CO2 buildup in two important ways: by displacing the use of fossil fuels, and by recycling the CO2 that is released when it is combusted as fuel. An important advantage of crop-based bioethanol is its GHG benets [24]. 2. Bioethanol as a transportation fuel The alcohols are oxygenates fuels that the alcohol molecule has one or more oxygen, which decreases to the combustion heat.

Practically, any of the organic molecules of the alcohol family can be used as a fuel. The alcohols can be used for motor fuels are methanol (CH3OH), bioethanol (C2H5OH), propanol (C3H7OH), butanol (C4H9OH). However, only methanol and bioethanol fuels are technically and economically suitable for internal combustion engines (ICEs) [24]. Bioethanol is ethyl alcohol, grain alcohol, or chemically C2H5OH or EtOH. It has high octane number (108) [25], both permit the rising of the compression ratio and gives lower emission [26]. Octane number is a measure of the gasoline quality for prevention of early ignition, which leads to cylinder knocking. The fuels with higher octane numbers are preferred in spark-ignition ICEs. An oxygenate fuel such as bioethanol is provides a reasonable antiknock value [3]. Disadvantages of bioethanol include its lower energy density than gasoline (but about 35% higher than that of methanol), its corrosiveness, low ame luminosity, lower vapor pressure (making cold starts difcult), miscibility with water, and toxicity to ecosystems [27], increase in exhaust emissions of acetaldehyde, and increase in vapor pressure (and evaporative emissions) when blending with gasoline. Physical and chemical properties of bioethanol, methanol and gasoline are given in Table 2 [28]. Bioethanol has been used as a modern biofuel, applied directly as a gasoline improver or gasoline subsistent, or in the form of ETBE (ethyl tertiary butyl ether) for currently added synthetically-produced octane enhancers and in bioethanoldiesel blends with particular purpose to reduce the emissions of exhaust gasses [29]. Bioethanol is most commonly blended with gasoline in concentrations of 10% bioethanol to 90% gasoline, known as E10 and nicknamed gasohol. Bioethanol can be used as a 5% blend with petrol under the European Union (EU) quality standard EN 228. This blend requires no engine modication and is covered by vehicle warranties. With engine modication, bioethanol can be used at higher levels, for example, E85 [30]. Some countries has exercised biofuel program both form bioethanolgasoline blend program such as the United States (E10 and for exible-fuel vehicle-FFV E85), Canada (E10 and for FFV E85), Sweden (E5 and for FFV E85), India (E5), Australia (E10), Thailand (E10), China
Table 2 Physical and chemical properties of ethanol, methanol and gasoline. Property Molecular weight (g/mol) Specic gravity Vapor density rel. to air Liquid density (g cm3 at 298 K) Boiling point (K) Melting point (K) Heat of evaporation (Btu/lb) Heating value (kBTU gal1) Lower Upper Tank design pressure (psig) Viscosity (cp) Flash point (K) Flammability/explosion limits (%) Lower (LFL) (%) Upper (UFL) Auto ignition temperature (K) Solubility in H2O (%) Azeotrope with H2O Peak ame temperature (K) Minimum ignition energy in air (mJ) Methanol CH3OH 32 0.789 (298 K) 1.10 0.79 338 175 472 58 65 15 0.54 284 Ethanol C2H5OH 46 0.788 (298 K) 1.59 0.79 351 129 410 74 85 15 1.20 287 Gasoline C4-C12 114 0.739 (288.5 K) 3.04.0 0.74 300518 135 111 122 15 0.56 228

6.7 36 733 Miscib. (100%) None 2143 0.14

3.3 19 636 Miscib. (100%) 95% EtOH 2193 0.23

1.3 7.6 523733 Negl. (0.01) Immiscible 2303

860

M. Balat / Energy Conversion and Management 52 (2011) 858875 Table 3 Bioethanol yields from different energy crops. Country Brazil USA China EU-27 Canada Energy crop Sugarcane, 100% Corn, 98% Sweet sorghum, 2% Corn, 70% Wheat, 30% Wheat, 48% Sugar beet, 29% Corn, 70% Wheat, 30% Bioethanol yield (l/ha) 6641 3770 1365 2011 1730 1702 5145 3460 1075

Fig. 1. Reduction in GHG emissions, compared to gasoline, by bioethanol produced from a variety of feedstocks (on a life-cycle basis) [31].

(E10), Columbia (E10), Peru (E10), Paraguay (E7), Brazil (E20, E25 and FFV any blend). The reduced CO2 emissions mean that bioethanol is good for the environment. Using bioethanol-blended fuel for automobiles can signicantly reduce petroleum use and exhaust GHG emission [24]. On a life-cycle basis, not all biofuels are equal in terms of environmental benets. Fig. 1 demonstrates the lower GHG emissions resulting from the use of biofuels compared to gasoline on a life-cycle basis. As Fig. 1 illustrates, corn-based bioethanol offers rather limited benets, as it reduces GHG emissions by only 18% compared to gasoline. In contrast, sugarcane and cellulosic bioethanol result in almost 90% lower emissions [31]. The net energy balance of biomass to bioethanol conversion is the key parameter that explains the interest in using bioethanol fuel instead of fossil gasoline. From a life-cycle assessment (LCA) viewpoint, the ratio of the energy content of bioethanol to the net non-renewable primary energy (allocated to bioethanol) consumed in the whole production process from biomass production to its conversion into bioethanol. As the approach is LCA oriented, the energy input must be estimated in terms of primary energy [32]. Studies have shown that corn-based bioethanol yields 20 30% more energy, typically fossil fuel energy, than is consumed in making it. On the other hand, sugarcane and cellulosic bioethanol yield renewable energy nine times worth the fossil energy used to produce them [31]. 3. An overview of bioethanol feedstocks Bioethanol can be produced from different kinds of raw materials. The raw materials are classied into three categories of agricultural raw materials: sucrose-containing feedstocks (e.g. sugar cane, sugar beet, sweet sorghum and fruits), starch materials (e.g. corn, milo, wheat, rice, potatoes, cassava, sweet potatoes and barley) and lignocellulosic materials (e.g. wood, straw and grasses). Currently, a focus is on bioethanol production from crops, such as corn, wheat, sugar cane, as well as on highly abundant agricultural wastes. One major problem with bioethanol production is the availability of raw materials for the production. The availability of feedstock for bioethanol can vary considerably from season to season and depends on geographic locations. Locally available agricultural biomass will be used for the bioethanol production [33]. For a given production line, the comparison of the feedstocks includes several issues [34]: (1) chemical composition of the biomass, (2) cultivation practices, (3) availability of land and land use practices, (4) use of resources, (5) energy balance, (6) emission of greenhouse gases, acidifying gases and ozone depletion gases, (7) absorption of minerals to water and soil, (8) injection of pesticides, (9) soil erosion, (10) contribution to biodiversity and landscape value losses, (11) farm-gate price of the biomass, (12) logistic cost (trans-

port and storage of the biomass), (13) direct economic value of the feedstocks taking into account the co-products, (14) creation or maintain of employment, and (15) water requirements and water availability. Brazil utilizes sugarcane for bioethanol production while the United States and Europe mainly use starch from corn, and from wheat and barley, respectively. Sugarcane as a biofuel crop has much expanded in the last decade, yielding anhydrous bioethanol (gasoline additive) and hydrated bioethanol by fermentation and distillation of sugarcane juice and molasses [35]. Brazils sugarcane yield averages about 82.4 tons/ha [36]. The yield of bioethanol per hectare, currently at around 6650 l/ha (Table 3) [37]. Brazil is the largest single producer of sugarcane with about 31% of global production [35]. It has nearly 9 million hectares of sugarcane under cultivation. Sugar beet crops are grown in most of the EU-25 countries, and yield substantially more bioethanol per hectare than wheat. The United States is predominantly a producer of bioethanol derived from corn, and production is concentrated in Midwestern states with abundant corn supplies [38]. Feedstock availability is not expected to be a constraint for bioethanol production over the next decade. Corn is expected to remain the predominant feedstock in the United States, although its share likely will decline modestly by 2015. Corn-based bioethanol production in most of the countries assessed is limited, especially compared to the United States. Only Canada reported explicit plans for signicant future development of corn-based bioethanol, although China has used corn as a feedstock in the past and Argentina is looking at the possibility of corn as biofuel feedstock in the future [39]. 4. Lignocellulosic-biomass materials 4.1. Availability of lignocellulosic material The price of the raw materials is also highly volatile, which can highly affect the production costs of the bioethanol [40]. Lignocellulosic materials serve as a cheap and abundant feedstock, which is required to produce fuel bioethanol from renewable resources at reasonable costs. In 2007 the US Department of Energy provided more than US$1 billion toward lingocellulosic bioethanol projects, with the goal of making the fuel cost competitive at US$1.33 per gallon by 2012 [41]. The level of support provided by the EU is far less, but is still signicant (approximately US$68 million in 2006) [41]. Lignocellulosic materials can be classied in four groups based on type of resource: (1) forest residues, (2) municipal solid waste, (3) waste paper, and (4) crop residue resources. Literature reports several papers on utilization of various lignocellulosic waste materials such as rice straw [42], corn stover [43], switchgrass [44], palm bagasse [45], etc. Lignocellulosic materials could produce up to 442 billion liters per year of bioethanol [46]. Rice straw is one of the abundant

M. Balat / Energy Conversion and Management 52 (2011) 858875

861

lignocellulosic waste materials in the world. It is annually produced about 731 million tons which is distributed in Africa (20.9 million tons), Asia (667.6 million tons), Europe (3.9 million tons), America (37.2 million tons) and Oceania (1.7 million tons). This amount of rice straw can potentially produce 205 billion liters bioethanol per year, which is the largest amount from a single biomass feedstock [47].

4.2. Chemical structure and basic components of lignocellulosic materials Chemical composition of lignocellulosic materials is a key factor affecting efciency of biofuel production during conversion processes. The structural and chemical composition of lignocellulosic materials is highly variable because of genetic and environmental inuences and their interactions [48]. A typical chemical composition of lignocellulosic materials is 48 wt.% C, 6 wt.% H, and 45 wt.% O, the inorganic matter being a minor component [49]. The proximate analysis of rice straw and wheat straw shows components as follow: volatile matter (65.47%, 75.27%), xed carbon (15.86%, 17.71%) and ash (18.67%, 7.02%), respectively [50]. Lignocelluloses consist mainly of cellulose, hemicellulose and lignin; these components build up about 90% of dry matter in lignocelluloses, with the rest consisting of e.g. extractive and ash [51]. The basic structure of all woody biomass consists of three basic polymers: cellulose (C6H10O5)x, hemicelluloses such as xylan (C5H8O4)m, and lignin [C9H10O3(OCH3)0.91.7]n in trunk, foliage, and bark. The proportion of these wood constituents varies between species, and there are distinct differences between hardwoods and softwoods. Cellulose + hemicellulose contents are more in hardwoods (78.8%) than softwoods (70.3%), but lignin is more in softwoods (29.2%) than hardwoods (21.7%) [52]. The structural composition of various types of lignocellulosic-biomass materials are given in Table 4 [53]. Cellulose and hemicellulose, which typically make up twothirds of cell wall dry matter, are polysaccharides that can be hydrolyzed to sugars and then fermented to bioethanol. Process performance, i.e. Bioethanol yield from biomass, is directly related to cellulose, hemicellulose, and individual sugar concentration in the feedstock [54]. The lignin cannot be used for bioethanol production. Cellulose, the major component of plant biomass (3060% of total feedstock dry matter), is a linear polymer of glucose; the orientation of the linkages and additional hydrogen bonding make the polymer rigid and difcult to break. In hydrolysis the polysaccharide is broken down to free sugar molecules by the addition of water [55].

This process is also known as saccharication. The product, glucose, is a six-carbon sugar. Hemicellulose (2040% of total feedstock dry matter) is a short, highly branched polymer of ve-carbon (pentoses) and six-carbon (hexoses) sugars. Specically, hemicellulose contains xylose and arabinose (ve-carbon sugars) and galactose, glucose, and mannose (six-carbon sugars). Hemicellulose is more readily hydrolyzed compared to cellulose because of its branched, amorphous nature [48]. The dominant sugars in hemicelluloses are mannose in softwoods and xylose in hardwoods and agriculture residues [56]. Lignin (1525% of total feedstock dry matter) is an aromatic polymer synthesised from phenylpropanoid precursors. The basic chemical phenylpropane units of lignin (primarily syringyl, guaiacyl and p-hydroxy phenol) are bonded together by a set of linkages to form a very complex matrix [57]. This matrix comprises a variety of functional groups, such as hydroxyl, methoxyl and carbonyl, which impart a high polarity to the lignin macromolecule [58]. Softwood and hardwood lignins belong to the rst and second category, respectively. Softwoods generally contain more lignin than hardwoods [59]. Lignin contents on a dry basis in both softwoods and hardwoods generally range from 20% to 40% by weight and from 10% to 40% by weight in various herbaceous species, such as bagasse, corncobs, peanut shells, rice hulls and straws [60]. Lignin is one of the drawbacks of using lignocellulosic-biomass materials in fermentation, as it makes lignocellulose resistant to chemical and biological degradation [56]. 5. Bioethanol from lignocellulosic materials via the biochemical pathway Biochemical conversion of lignocellulosic materials through saccharication and fermentation is a major pathway for bioethanol production from biomass. Bioconversion of lignocellosics to bioethanol is difcult due to: (1) the resistant nature of biomass to breakdown; (2) the variety of sugars which are released when the hemicellulose and cellulose polymers are broken and the need to nd or genetically engineer organisms to efciently ferment these sugars; (3) costs for collection and storage of low density lignocellosic materials. Generic block diagram of bioethanol production from lignocellulose materials is given in Fig. 2 [61]. The basic process steps in producing bioethanol from lignocellulosic materials are: pretreatment, hydrolysis, fermentation and product separation/distillation. 5.1. Pretreatment of lignocellulosic materials The recalcitrance of lignocellulose is one of the major barriers to the economical production of bioethanol. The technical approach to overcome recalcitrance has been pretreatment of biomass feedstock to remove the barriers and make cellulose more accessible to hydrolytic enzymes for conversion to glucose [62]. The goals of pretreatment on lignocellulosic material are depicted in Fig. 3 [63]. If the pretreatment is not efcient enough the resultant residue is not easily hydrolyzable by cellulase enzyme and if it is more severe, result is the production of toxic compounds which inhibit the microbial metabolism [64]. Pretreatment has been viewed as one of the most expensive processing steps within the conversion of biomass to fermentable sugar [65]. There is huge scope in lowering the cost of pretreatment process through extensive R&D approaches. Pretreatment of cellulosic biomass in cost effective manner is a major challenge of cellulose to bioethanol technology research and development [66]. Taherzadeh and Karimi [56] has summarized the prerequisites for an ideal lignocellulose pretreatment; it should: (1) production of reactive cellulosic ber for enzymatic attack, (2)

Table 4 Composition of various types of lignocellulosic-biomass materials (% dry weight). Material Algae (green) Cotton, ax, etc. Grasses Hardwoods Hardwood barks Softwoods Softwood barks Cornstalks Wheat straw Newspapers Chemical pulps Cellulose 2040 8095 2540 45 2 2240 42 2 1838 3947 3741 4055 6080 Hemicelluloses 2050 520 2550 30 5 2038 27 2 1533 2631 2732 2540 2030 Lignin 1030 20 4 3055 28 3 3060 35 1315 1830 210 Ash 0.6 0.2 0.8 0.2 0.5 0.1 0.8 0.2 1216 1114 Extractives 53 62 32 42 13 72

862

M. Balat / Energy Conversion and Management 52 (2011) 858875

researchers have investigated the effect of different pretreatment methods upon various lignocellulosic materials such as corn stover [69], wheat straw [70], switchgrass [71], rice straw [72], and sugarcane bagasse [73]. 5.1.1. Physical pretreatment 5.1.1.1. Mechanical comminution. Lignocellulosic materials can be comminuted by a combination of chipping, grinding, and milling to reduce cellulose crystallinity. The size of the materials is usually 1030 mm after chipping and 0.22 mm after milling or grinding [68,74,75]. Vibratory ball milling was found to be more effective than ordinary ball milling in reducing cellulose crystallinity of spruce and aspen chips and in improving their digestibility [68,76]. Power requirements of mechanical comminution depend on the nal particle size and the biomass characteristics [77]. Power requirements increase rapidly with decreasing particle size, as shown in Fig. 4 [78]. The energy requirements of mechanical comminution are regarded as high for hardwood, which consumes 130 kW h/ton to reduce the particle size to 1.6 mm. To reduce the size of corn stover with mechanical comminution to 1.6 mm requires far less energy, consuming only 14 kW h/ton [79]. These mechanical pretreatment techniques are time-consuming, energy intensive, or expensive to process. The compression milling is apparently the only comminution process that has been tested using a production-scale apparatus [80]. 5.1.1.2. Pyrolysis. Pyrolysis has also been used for pretreating lignocellulosic materials, since biomass can be used as substrate for a fast pyrolysis for thermal conversion of cellulose and hemicellulose into fermentable sugars with good yields [81]. When the ground cellulosic materials are treated at temperatures greater than 573 K, cellulose rapidly decomposes to produce gaseous products and residual char [74,82]. Pyrolysis pretreatment prior to enzymatic hydrolysis of three waste cellulosic materials (ofce paper, newspaper and cardboard) was examined by Leustean [75]. Table 6 shows after treatment and enzymatic hydrolysis the reducing sugar concentration. The pyrolysis pretreatment of ground material is improved the conversion of cellulose to glucose yield from enzymatic hydrolysis [75]. 5.1.2. Physico-chemical pretreatment 5.1.2.1. Steam explosion (autohydrolysis). Steam explosion is the most commonly used method for the pretreatment of lignocellulosic materials [78]. To summarize the effects of steam explosion treatment on lignocellulosics reported in the literature [83]: (1) steam explosion treatment increases crystallinity of cellulose by promoting crystallization of the amorphous portions; (2) hemicellulose is easily hydrolyzed by steam explosion treatment; (3) there is evidence that steam explosion promotes delignication. In this method, chipped biomass is treated with high-pressure saturated steam and then the pressure is swiftly reduced, which makes the materials undergo an explosive decomposition. Steam explosion, compared to other pretreatment methods, offers potential for lower capital investment, signicantly lower environmental impact, more potential for energy efciency, less hazardous process chemicals and conditions and complete sugar recovery [81]. The conventional mechanical methods require 70% more energy than steam explosion to achieve the same size reduction [68,74,84]. Steam explosion is considered the most cost effective option for hardwood and agriculture residues, but is less effective for softwood [84]. Most important factors affecting effectiveness of steam explosion are particle size, temperature and residence time and the combined effect of both temperature and time is described by severity factor (Ro), which is optimal for maximum sugar yield between 3.0 and 4.5 [81]. Steam explosion is initiated at a temperature of 433533 K with a corresponding pressure

Fig. 2. Generic block diagram of bioethanol production from lignocellulose biomass. Possibilities for reactionreaction integration are shown inside the shaded boxes: SSF simultaneous saccharication and fermentation; SSFC simultaneous saccharication and co-fermentation. Main stream components are: C cellulose; H hemicellulose; L lignin; G glucose; P pentose; I inhibitors; EtOH ethanol [61].

Fig. 3. Schematic of goals of pretreatment on lignocellulosic material [63].

avoiding destruction of hemicelluloses and cellulose, (3) avoiding formation of possible inhibitors for hydrolytic enzymes and fermenting microorganisms, (4) minimizing the energy demand, (e) reducing the cost of size reduction for feedstocks, (5) reducing the cost of material for construction of pretreatment reactors, (6) producing less residues, and (7) consumption of little or no chemical and using a cheap chemical. Pretreatment is crucial for ensuring good ultimate yields of sugars from both polysaccharides. Hydrolysis without preceding pretreatment yields typically <20%, whereas yields after pretreatment often exceed 90% [55]. Physical (milling and grinding), physico-chemical (steam explosion/autohydrolysis, hydrothermolysis, and wet oxidation), chemical (alkali, dilute acid, oxidizing agents, and organic solvents), and biological processes have been used for pretreatment of lignocellulosic materials. However, not all of these methods have yet developed enough to be feasible technically or economically for large-scale processes. For example, milling could be applied to create a better steam explosion by reducing the chip size [67]. Advantages and disadvantages of various pretreatment processes for lignocellulosic materials are summarized in Table 5 [68]. Many

M. Balat / Energy Conversion and Management 52 (2011) 858875 Table 5 Advantages and disadvantages of various pretreatment processes for lignocellulosic materials [68]. Pretreatment process Mechanical comminution Steam explosion Advantages Reduces cellulose crystallinity Causes hemicellulose degradation and lignin transformation; cost-effective Increases accessible surface area, removes lignin and hemicellulose to an extent; does not produce inhibitors for down-stream processes Increases accessible surface area; cost-effective; does not cause formation of inhibitory compounds Reduces lignin content; does not produce toxic residues Hydrolyzes hemicellulose to xylose and other sugars; alters lignin structure Removes hemicelluloses and lignin; increases accessible surface area Hydrolyzes lignin and hemicelluloses Produces gas and liquid products Ambient conditions; disrupts plant cells; Simple equipment degrades lignin and hemicelluloses; low energy requirements Limitations and disadvantages

863

AFEX

Power consumption usually higher than inherent biomass energy Destruction of a portion of the xylan fraction; incomplete disruption of the lignin-carbohydrate matrix; generation of compounds inhibitory to microorganisms Not efcient for biomass with high lignin content

CO2 explosion Ozonolysis Acid hydrolysis Alkaline hydrolysis Organosolv Pyrolysis Pulsed electrical eld Biological

Does not modify lignin or hemicelluloses Large amount of ozone required; expensive High cost; equipment corrosion; formation of toxic substances Long residence times required; irrecoverable salts formed and incorporated into biomass Solvents need to be drained from the reactor, evaporated, condensed, and recycled; high cost High temperature; ash production Process needs more research Rate of hydrolysis is very low

Fig. 4. Energy requirements for ball milling municipal solid waste [78].

0.694.83 MPa for several seconds to a several minutes before the material is exposed to atmospheric pressure for cooling [85]. Uncatalyzed steam explosion refers to a pretreatment technique in which lignocellulosic biomass is rapidly heated by highpressure steam without addition of any chemicals. The biomass/ steam mixture is held for a period of time to promote hemicellulose hydrolysis, and terminated by an explosive decompression [86]. Negro et al. [87] studied steam explosion and liquid hot water methods for pretreatment of poplar (Populus nigra) biomass. The best results were obtained in steam explosion pretreatment at 483 K and 4 min, taking into account cellulose recovery above 95%, enzymatic hydrolysis yield of about 60%, and 41% xylose recovery in the liquid fraction. Addition of H2SO4 (or SO2) or CO2 [typically 0.33% (w/w)] in steam explosion can decrease time and temperature, effectively improve hydrolysis, decrease the production of inhibitory compounds, and lead to complete removal of hemicellulose [68]. H2SO4 is a strong catalyst that highly improves the hemicellulose removal but also easily yields inhibitory subTable 6 Pyrolysed cellulosic materials [75]. Ofce paper Reducing sugar (mg/ml) 0.14 Newspaper 0.11 Cardboard 0.11

stances [88]. Ballesteros et al. [89] applied acid-catalyzed steam explosion pretreatment of wheat straw for bioethanol production by varying the temperature (433473 K), the residence time (5, 10 or 20 min) and the acid concentration [H2SO4 0.9% (w/w)]. According to results of this study, the best pretreatment conditions to obtain high conversion yield to bioethanol (approx 80% of theoretical) of cellulose-rich residue after steam explosion are 463 K and 10 min or 473 K and 5 min, in acid-impregnated straw. Using a H2SO4-catalyzed steam explosion process for pretreatment of Salix chips, at 473 K for either 4 or 8 min using 0.5% sulfuric acid, resulted in glucose recovery about 92% and 86% xylose recovery after enzymatic hydrolysis [90]. SO2 appears more appealing than H2SO4 in steam explosion since the former requires milder and much less expensive reactor material, generates less gypsum, yields more xylose, and produces more digestible substrate with high fermentability [76]. The treatment can be carried out by 1 4% SO2 (w/w substrate) at elevated temperatures, e.g. 433503 K, for a period of e.g. 10 min [56]. The main drawback of SO2 is its high toxicity, which may pose safety and health risks. However, SO2 is used in various industrial processes using established techniques [91]. Two-step pretreatment has been suggested in several studies as a means of increasing the sugar recovery [92,93]. In the rst step, steam is performed using low temperature to solubilize hemicellu-

864

M. Balat / Energy Conversion and Management 52 (2011) 858875

losic fraction, and cellulose fraction is subjected to a second steam explosion pretreatment step at a temperature higher than 483 K. It offers some additional advantages (higher bioethanol yields, better use of raw material and lower enzyme dosages during steam explosion) [81]. 5.1.2.2. Ammonia ber explosion. Ammonia ber explosion (AFEX) is one of the alkaline physico-chemical pretreatment processes. In this process, the material is subjected to liquid ammonia at high temperature and pressure, and a subsequent fast decompression, similar to the steam explosion, which causes a fast saccharication of the lignocellulosic material [94]. In a typical AFEX process, the dosage of liquid ammonia is 12 kg ammonia/kg dry biomass, the temperature is 363 K, and the residence time is 30 min [68,74]. The effective parameters in the AFEX process are ammonia loading, temperature, water loading, blowdown pressure, time, and number of treatments [56]. This system does not directly liberate any sugars, but allows the polymers (hemicellulose and cellulose) to be attacked enzymatically and reduced to sugars [95]. AFEX pretreatment yields optimal hydrolysis rates for pretreated lignocellulosics with close to theoretical yields at low enzyme loadings (<5 FPU/g of biomass or 20 FPU/g cellulose) [86]. AFEX pretreatment has been demonstrated to markedly improve the saccharication rates of numerous herbaceous crops and grasses [78]. It has been applied to various lignocellulosic materials, including rice straw, municipal solid waste, newspaper, sugar beet pulp, sugarcane bagasse, corn stover, switchgrass, miscanthus, apsen chips, etc. [76]. The optimal conditions for pretreatment of switchgrass with AFEX were reported by Alizadeh et al. [96]. The optimal pretreatment conditions were found to be near 373 K reactor temperature, and ammonia loading of 1 kg of ammonia per kilogram of dry matter with 80% moisture content (dry weight basis) at 5 min residence time. These conditions yielded a sixfold improvement in hydrolysis efciency. Teymouri et al. [97] evaluated the optimum process conditions for the pretreatment of corn stover. The optimal pretreatment conditions were found to be a temperature of 373 K, an ammonia loading of 1 kg of ammonia per kilogram of dry matter, a moisture content of 60% (dry weight basis), and a residence time of 5 min. Approximately 98% of the theoretical glucose yield was obtained during enzymatic hydrolysis of the optimal treated corn stover using 60 lter paper units (FPU) of cellulase enzyme/g of glucan (equal to 22 FPU/g of dry corn stover). The bioethanol yield from this sample was increased up to 2.2 times over that of untreated sample. The composition of the materials after AFEX pretreatment was essentially the same as the original materials [74]. Holtzapple et al. [98] obtained over 90% of hydrolysis of cellulose and hemicellulose after AFEX pretreatment of Bermuda grass (approximately 5% lignin) and bagases (15% lignin). AFEX works only moderately and is not attractive for the biomass with high lignin content. Lignin content in grasses (1520%) is relatively low when compared with hardwoods and softwood (2035%) [98]. This could be one of the major reasons that grasses can be more easily digested compared to AFEX treated hardwoods following AFEX pretreatment [99]. Ammonia must be recycled after the pretreatment to reduce the cost and protect the environment [56]. A possible approach is to recover the ammonia after the pretreatment by evaporation [100]. 5.1.2.3. Liquid hot-water pretreatment. Cooking of lignocellulosic materials in liquid hot water (LHW) is one of the hydrothermal pretreatment methods applied for pretreatment of lignocellulosic materials since several decades ago in e.g. pulp industries [56]. LHW subjects biomass to hot water in liquid state at high pressure during a xed period and it presents elevated recovery rates for pentoses and generates low amount of inhibitors [81]. This pretreatment process usually has involved temperatures of 473

503 K for up to 15 min. Around 4060% of the total mass is dissolved in this process, with 422% of the cellulose, 3560% of the lignin and all of the hemicellulose being removed [101]. If the pH is maintained between 4 and 7, the degradation of monosaccharide sugars can be minimized [102]. 5.1.3. Chemical pretreatment 5.1.3.1. Ozonolysis. Ozonolysis involves using ozone gas to break down the lignin and hemicellulose and increase the biodegradability of the cellulose. The pretreatment is usually carried out at room temperature and is effective at lignin removal without the formation of toxic by-products [103]. Ozonation has been widely used to reduce the lignin content of both agricultural and forestry wastes [104]. Ozonolysis has been shown to break down 49% of lignin in corn stalks and 5559% of lignin in autohydrolyzed (hemicellulose free) corn stalks [105]. In a study [106], wheat and rye straws were pretreated with ozone to increase the enzymatic hydrolysis extent of potentially fermentable sugars. Enzymatic hydrolysis yields of up to 88.6% and 57% were obtained compared to 29% and 16% in non-ozonated wheat and rye straw respectively. A drawback of ozonolysis is that a large amount of ozone is required, which can make the process expensive [68]. 5.1.3.2. Alkaline pretreatment. Alkali pretreatment refers to the application of alkaline solutions to remove lignin and various uronic acid substitutions on hemicellulose that lower the accessibility of enzyme to the hemicellulose and cellulose [107,108]. These processes utilize lower temperatures and pressures compared to other pretreatment technologies. Alkali pretreatment may be carried out at ambient conditions, but pretreatment time is measured in terms of hours or days rather than minutes or seconds [86]. Regardless the advantages, these methods present difculties from the point of view of the process economy for obtaining fuels [94]. Sodium, potassium, calcium and ammonium hydroxide are appropriate chemicals for pretreatment. Of these four, NaOH has been studied the most [68]. Dilute NaOH treatment of lignocellulosic biomass causes swelling, leading to an increase in the internal surface area, a decrease in crystallinity, separation of structural linkages between lignin and carbohydrates, and disruption of the lignin structure [109]. Millet et al. [110] reported that the digestibility of NaOH-treated hardwood to increase from 14% to 55% with the decrease of lignin content from 24%55% to 20%. However, no effect of dilute NaOH pretreatment was found for softwoods with lignin content greater than 26%. Silverstein et al. [107] reported >65% lignin reduction in cotton stalk treated with 2% NaOH for 90 min at 394 K/15 psi. Cellulases produced by Bacillus subtilis for the saccharication of wheat straw, rice straw and bagasse were used by Akhtar et al. [111]. Pretreatment of these substrates with 2% NaOH was found to be more effective for increasing the saccharication. The saccharication rates of 33.0%, 25.5% and 35.5% were obtained with 2% NaOH pretreated wheat straw, rice straw and bagasse, respectively. In a study [112], a combination of NaOH treatment and homogenization was used as a pretreatment to enhance the enzymatic hydrolysis of corn stover. The highest glucose yield (6.25 g/l) was obtained when the corn stover was pretreated by a combination of 1.0 N NaOH treatment and homogenization (Fig. 5). Lime (Ca(OH)2) as compared to NaOH and KOH has lower cost and less signicant safety requirements. It can be recovered from hydrolyzate by reaction with CO2, so that formed carbonate can then be reconverted to lime [113]. To make lime as efcient as other alkalis in enhancing the digestibility of lignocellulose, appropriate pretreatment conditions need to be employed [114]. Lime, water, and an oxidizing agent (air or O2) are mixed with the biomass at temperatures ranging from 313 to 423 K for a period ranging from hours to weeks [115]. Two types of lime treatment have been explored: (1) short-term and (2) long-term. Short-term lime

M. Balat / Energy Conversion and Management 52 (2011) 858875

865

Fig. 5. Effect of NaOH concentration on enzymatic hydrolysis of corn stover pretreated by combined NaOH treatment and homogenization (raw material = 2 mm corn stover, pretreatment conditions = NaOH treatment + homogenization, hydrolysis conditions = 20 GCU cellulose/g substrate at 323 K, PH 4.8) [112].

pretreatment involves boiling the biomass with a lime loading of 0.1 g Ca(OH)2/g dry biomass at temperatures of 358408 for 1 3 h [116]. Long-term pretreatment involves using the same lime loading at lower temperatures (313328 K) for 46 weeks in the presence of air [116]. Lime has been used to pretreat switchgrass (373 K for 2 h) [117], wheat straw (394 K for 1 h) [118], corn stover (373 K for 13 h) [119], and poplar wood (423 K for 6 h with 14-atm oxygen) [120]. Saha and Cotta [118] obtained maximum total sugar yield (451 3mg g1 straw; glucose, 252 6 mg; xylose, 173 3 mg; arabinose, 27 2 mg; 65% conversion) by lime pretreatment (100mg g1 straw, 394 K, 1 h). The authors also investigated the effects of pH (3.56.5) and temperature (298343 K) on the enzymatic hydrolysis of lime pretreated wheat straw (8.6%, w/v) using a combination of three enzymes (cellulose, b-glucosidase, and hemicellulase), each enzyme at a dose level of 0.05 ml g1 substrate. Fig. 6 shows the effect of pH and temperature on the enzymatic hydrolysis of lime pretreated wheat straw. Alkaline peroxide is one of the effective pretreatment methods that can improve the enzymatic hydrolysis by delignication of lignocellulosic materials. In this method, the lignocelluloses are soaked in pH-adjusted water (e.g. to pH 1112 using NaOH) containing H2O2 at room temperatures for a period of time (e.g. 624 h) [56]. 5.1.3.3. Acid pretreatment. Acid pretreatments normally aim for high yields of sugars from lignocellulosic materials. Acid pretreatment involves the use of sulfuric, nitric, or hydrochloric acids to remove hemicellulose components and expose cellulose for enzymatic digestion [107]. The acid pretreatment can operate either under a high temperature and low acid concentration (dilute acid pretreatment) or under a low temperature and high acid concentration (concentrated acid pretreatment) [56]. Dilute acid hydrolysis has been successfully developed for pretreatment of lignocellulosic materials. The dilute acid pretreatment works fairly well on agricultural feedstocks, such as corn stover and rice/wheat straw [62]. While dilute acid pretreatments are known to improve enzymatic hydrolysis, their cost is relatively high compared to physico-chemical pretreatments [44]. This pretreatment method gives high reaction rates and signicantly improves cellulose hydrolysis [47]. There are primarily two types of dilute acid pretreatment processes: low solids loading (510% [w/w]), high temperature (T > 433 K), continuous-ow processes and high solids

Fig. 6. (A) Effect of pH on the enzymatic hydrolysis of lime pretreated wheat straw at 318 K and (B) effect of temperature on the enzymatic hydrolysis of lime pretreated wheat straw at pH 5.0. Yield of total sugars at: (a) 6, (b) 24, and (c) 72 h [118].

loading (1040% [w/w]), lower temperature (T < 433 K), batch processes [121]. In general, higher pretreatment temperatures and shorter reactor residence times result in higher soluble xylose recovery yields and enzymatic cellulose digestibility. Higher-temperature dilute acid pretreatment has been shown to increase cellulose digestibility of pretreated residues [122]. Depending on the substrate and the conditions used, between 80% and 95% of the hemicellulosic sugars can be recovered by dilute acid pretreatment from the lignocellulosic material [47,123,124]. Silverstein et al. [107] reported 95% xylan reduction in cotton stalk treated with 2% H2SO4 for 90 min at 394 K/15 psi. In recent years, treatment of lignocellulosic biomass with dilute sulfuric acid has been primarily used as a means of hemicellulose hydrolysis and pretreatment for enzymatic hydrolysis of cellulose [125]. Sulfuric acid at concentrations usually below 4 wt.%, has been of the most interest in such studies as it is inexpensive and effective [68]. Dilute sulfuric acid pretreatment (0.22.0% sulfuric acid, 394493 K) of lignocellulose serves three important functions in the conversion process [121]: (1) hydrolysis of the hemicellulose components to produce a syrup of monomeric sugars, (2) exposure of cellulose for enzymatic digestion by removal of hemicellulose and part of the lignin, and (3) solubilization of heavy metals which may be contaminating the feedstock. In spite of these benets, dilute sulfuric acid has some important disadvantages [76]: (1) corrosion that mandates expensive materials of construction, (2) acidic prehydrolyzates must be neutralized before the sugars proceed to fermentation, (3) gypsum has problematic reverse solubility characteristics when neutralized with inexpensive calcium hydroxide, (4) formation of degradation products and release of natural biomass fermentation inhibitors are other characteristics of acid pretreatment, (5) disposal of neutralization salts is needed, and (6) biomass particle size reduction is necessary.

866

M. Balat / Energy Conversion and Management 52 (2011) 858875

5.1.4. Biological pretreatment Biological pretreatment involves microorganisms such as brown-, white- and soft-rot fungi that are used to degrade lignin and solubilize hemicellulose. White-rot fungi are the most effective biological pretreatment of lignocellulosic materials [56,68,74]. Lee et al. [126] evaluated biological pretreatment of Japanese red pine (Pinus densiora) using three white-rot fungi (Ceriporia lacerata, Stereum hirsutum, and Polyporus brumalis). pretreatment with S. hirsutum resulted in selective degradation of the lignin rather than the holocellulose component. The advantages of biological pretreatment include low energy requirement and mild environmental conditions. However, the rate of hydrolysis in most biological pretreatment process is very low [74]. 5.2. Hydrolysis techniques The carbohydrate polymers in lignocellulosic materials need to be converted to simple sugars before fermentation, through a process called hydrolysis [127]. Various methods for the hydrolysis of lignocellulosic materials have recently been described. The most commonly applied methods can be classied in two groups: chemical hydrolysis (dilute and concentrated acid hydrolysis) and enzymatic hydrolysis. There are some other hydrolysis methods in which no chemicals or enzymes are applied. For instance, lignocellulose may be hydrolyzed by gamma-ray or electron-beam irradiation, or microwave irradiation. However, those processes are commercially unimportant. Several products can result from hydrolysis of lignocellulosic material (Fig. 7) [59]. When hemicelluloses are hydrolyzed to xylose, mannose, acetic acid, galactose, and glucose are liberated. The main application of xylose is its bioconversion to xylitol, a functional sweetener with important technological properties [128]. Hemicellulosic hydrolysis can be generalized as

5.2.1. Chemical hydrolysis Chemical hydrolysis involves exposure of lignocellulosic materials to a chemical for a period of time at a specic temperature, and results in sugar monomers from cellulose and hemicellulose polymers [127]. In the chemical hydrolysis, the pretreatment and the hydrolysis may be carried out in a single step. Acids are predominantly applied in chemical hydrolysis [127]. There are two basic types of acid hydrolysis processes: dilute acid and concentrated acid, each with variations. 5.2.1.1. Dilute acid hydrolysis. Dilute acid hydrolysis is the oldest technology for converting cellulose biomass to bioethanol. This process is conducted under high temperature and pressure, and has a reaction time in the range of seconds or minutes, which facilitates continuous processing. Dilute acid process involves a solution of about 1% H2SO4 concentration in a continuous ow reactor at a high temperature (about 488 K). Most dilute acid processes are limited to a sugar recovery efciency of around 50% [132]. The combination of acid and high temperature and pressure dictate special reactor materials, which can make the reactor expensive. The rst reaction converts the cellulosic materials to sugar and the second reaction converts the sugars to other chemicals. Unfortunately, the conditions that cause the rst reaction to occur also are the right conditions for the second to occur [53]. Dilute acid hydrolysis occurs in two-stages to take advantage of the differences between hemicellulose and cellulose. The rststage is performed at low temperature to maximize the yield from the hemicellulose, and the second, higher temperature stage is optimized for hydrolysis of the cellulose portion of the feedstock [133]. The rst-stage is conducted under mild process conditions (e.g. 0.7% H2SO4, 463 K) to recover the ve-carbon sugars, while in the second stage only the remaining solids with the more resistant cellulose undergo harsher conditions (488 K, but a milder 0.4% H2SO4) to recover the six-carbon sugars [55]. Modern experimental yields for this two-stage acid process (3 min per stage) are 89% for mannose, 82% for galactose, but only 50% for glucose with a glucose to bioethanol conversion that is 90% of the theoretical maximum [102]. Schematic owsheet for dilute acid hydrolysis is given in Fig. 8. The primary challenge for dilute acid hydrolysis processes is how to raise glucose yields higher than 70% in an economically viable industrial process while maintaining a high cellulose hydrolysis rate and minimizing glucose decomposition. For rapid continuous processes, in order to allow adequate acid penetration, feedstocks must also be reduced in size so that the maximum particle dimension is in the range of a few millimeters [132]. 5.2.1.2. Concentrated acid hydrolysis. This process involves an acid (dilute or concentrated) pretreatment to liberate the hemicellulosic sugars while the subsequent stage requires the biomass to be dried followed by the addition of concentrated sulfuric acid (70 90%) [102]. The acid concentration used in concentrated acidhydrolysis process is in the range of 1030% [134]. Reaction times are typically much longer than for dilute acid process. This process provides a complete and rapid conversion of cellulose to glucose and hemicelluloses to ve-carbon sugars with little degradation. The critical factors needed to make this process economically viable are to optimize sugar recovery and cost effectively recovers the acid for recycling [135]. In comparison to dilute acid hydrolysis, concentrated acid hydrolysis leads to little sugar degradation and gives sugar yields approaching 100% [136]. Table 7 shows the yields of bioethanol by concentrated sulfuric acid hydrolysis from cornstalks. The concentrated acid process offers more potential for cost reductions than the dilute acid process. However, environment and corrosion problems and the high cost of acid consumption and recovery present major barriers to economic success [136].

Hemicelluloses ! Xylan ! Xylose ! Furfural Acetyl groups ! Acetic acid

1 2

Degradation of xylan yields eight main products: water, methanol, formic, acetic, and propionic acids, hydroxy-1-propanone, hydroxy-1-butanone and 2-furfuraldeyde [129].Under high temperature and pressure xylose is further degraded to furfural [130]. Similarly, 5-hydroxymethyl furfural (HMF) is formed from hexose degradation [131]. Cellulose is hydrolyzed to glucose. The following reaction is proposed for hydrolysis of cellulose:

Cellulose ! Glucan ! Glucose ! Decomposition products 3

Fig. 7. Main degradation products occurring during hydrolysis of lignocellulosic material.

M. Balat / Energy Conversion and Management 52 (2011) 858875

867

Fig. 8. Dilute acid hydrolysis (rst-stage and two-stages) and separate fermentation of pentose and hexose sugars [66].

5.2.2. Enzymatic hydrolysis Acid hydrolysis has a major disadvantage where the sugars are converted to degradation products like tars. This degradation can be prevented by using enzymes favoring 100% selective conversion of cellulose to glucose. When hydrolysis is catalyzed by such enzymes, the process is known as enzymatic hydrolysis [137]. Enzymatic hydrolysis of natural lignocellulosic materials is a very slow process because cellulose hydrolysis is hindered by structural parameters of the substrate, such as lignin and hemicellulose content, surface area, and cellulose crystallinity [138]. Utility cost of enzymatic hydrolysis is low compared to acid or alkaline hydrolysis because enzyme hydrolysis is usually conducted at mild conditions (pH 4.8) and temperature (318323 K) and does not have a corrosion problem [74]. The enzymatic hydrolysis has currently high yields (7585%) and improvements are still projected (85 95%), as the research eld is only a decade young [55]. Comparison of process conditions and performance of three cellulose hydrolysis processes is given in Table 8. Enzymatic hydrolysis is an environmentally friendly alternative that involves using carbohydrate degrading enzymes (cellulases and hemicellulases) to hydrolyze lignocelluloses into fermentable sugars [44]. 5.2.2.1. Enzymatic hydrolysis of cellulose. Cellulose is typically hydrolyzed by an enzyme called cellulase. These enzymes are produced by several microorganisms, commonly by bacteria and fungi. These microorganisms can be aerobic or anaerobic, mesophilic or thermophilic. Bacteria belonging to Clostridium, Cellulomonas,

Table 7 Yields of bioethanol by concentrated sulfuric acid hydrolysis from cornstalks [53]. Amount of cornstalk (kg) Cellulose content (kg) Cellulose conversion and recovery efciency (% dry weight) Bioethanol stoichiometric yield (% dry weight) Glucose fermentation efciency (% dry weight) Bioethanol yield from glucose (kg) Amount of cornstalk (kg) Hemicelluloses content (kg) Hemicelluloses conversion and recovery efciency (% dry weight) Bioethanol stoichiometric yield (% dry weight) Xylose fermentation efciency (% dry weight) Bioethanol yield from xylose (kg) Total bioethanol yield from 1000 kg of cornstalks 1000 430 76 51 75 130 1000 290 90 51 50 66 196 kg (225.7 L = 59 gallons)

Bacillus, Thermomonospora, Ruminococcus, Bacteriodes, Erwinia, Acetovibrio, Microbispora, and Streptomyces can produce cellulases effectively [74]. Fungi such as Sclerotium rolfsii, P. chrysosporium and species of Trichoderma, Aspergillus, Schizophyllum and Penicilium are used to produce cellulases [139]. Mutant strains of Trichoderma sp. (T. viride, T. reesei, T. longibrachiatum) have long been considered to be the most productive and powerful destroyers of crystalline cellulose [140]. Commercial products of various T. reesei isolates have been available for a long time in cereal foods applications, the brewing industry, fruit and vegetable processing and have also been widely evaluated and applied in relation to bioethanol production processes. T. reesei secretes high amounts of enzymes, up to 100 g l1 [141]. Cellulase is a group of enzymes that synergistically hydrolyzes cellulose (Fig. 9) [142]. The widely accepted mechanism for enzymatic cellulose hydrolysis involves synergistic actions by endoglucanses (EG, endo-1,4-b-D-glucanases, or EC 3.2.1.3.), exoglucanases or cellobiohydrolases (CBH, 1,4-b-D-glucan cellobiohydrolases, or EC 3.2.1.91.), and b-glucosidases (BGL, cellobiases or EC 3.2.1.21). EG hydrolyze accessible intramolecular b-1,4-glucosidic bonds of cellulose chains randomly to produce new chain ends; CBH processively cleave cellulose chains at the ends to release soluble cellobiose or glucose; and BGL hydrolyze cellobiose to glucose in order to eliminate cellobiose inhibition [143]. BGL complete the hydrolysis process by catalyzing the hydrolysis of cellobiose to glucose. The supplementation of b-glucosidase in hydrolysis is required due to its insufcient amount from T. reesei, to prevent cellulases inhibition resulted from cellobiose accumulation [114]. During cellulose hydrolysis, the solid substrate characteristics vary, including: (1) changes in the cellulose chain end number resulting from generation by EG and consumption by CBH and (2) changes in cellulose accessibility resulting from substrate consumption and cellulose fragmentation [143]. There are different factors that affect the enzymatic hydrolysis of cellulose, namely, substrates, cellulase activity, reaction conditions (temperature, pH as well as other parameters), and a strong product inhibition. To improve the yield and rate of enzymatic hydrolysis, research has been focused on optimizing the hydrolysis process and enhancing the cellulase activity [74]. The rate of enzymatic hydrolysis of cellulose is dependent upon several structural features of the cellulose. The cellulose features known to affect the rate of hydrolysis include: (1) molecular structure of cellulose, (2) crystallinity of cellulose, (3) surface area of cellulose ber, (4) degree of swelling of cellulose ber, (5) degree of polymerization, and (6) associated lignin or other materials [144]. A low substrate

868

M. Balat / Energy Conversion and Management 52 (2011) 858875

Table 8 Comparison of process conditions and performance of three hydrolysis processes [55]. Consumables Dilute acid Concentrated acid Enzymatic <1% H2SO4 3070% H2SO4 Cellulase Temperature (K) 488 313 323 Time 3 min 26 h 1.5 days Glucose yield (%) 5070 90 75 ? 95 Available Now Now Now ? 2020

Fig. 9. Mode of action of cellulolytic enzymes [142].

concentration gives low yield and rate, and a high cellulase dosage may increase the costs disproportional [55]. A cellulase dosage of 10 FPU (lter paper units) per gram of biomass is often used in laboratory studies because it provides a hydrolysis prole with high levels of glucose yield in a reasonable time (4872 h) at a reasonable enzyme cost [145]. Chen et al. [146] investigated effects of cellulase dosage on the enzymatic hydrolysis of dilute acid-treated corncob. Hydrolysis experiments were performed with 100 g l1 substrate and different dosages of T. reesei ZU-02 cellulase (FPU g1 substrate) at pH 4.8 and 323 K. Results of this study are shown in Fig. 10. As shown in Fig. 10, reducing sugar concentration and hydrolysis yield had a similar variation trend, that is, both increased sharply with cellulase dosage varying from 10 to 20 FPU g1 substrate, and basically leveled off from 20 to 30 FPU g1 substrate. One limitation with using cellulase is that there is a reduction in rates due to end product (cellobiose and glucose) inhibition. Simultaneous saccharication and fermentation (SSF) overcomes this problem by hydrolyzing cellulose and fermenting the hydrolysis product at the same time [147]. 5.2.2.2. Enzymatic hydrolysis of hemicelluloses. There is a great interest in the enzymatic hydrolysis of xylan because of possible applications in ruminal digestion, waste treatment, fuel and chemical production, and paper manufacture [148]. Unlike cellulose, xylans are chemically quite complex, and their degradation requires multiple enzymes. Xylan-degrading enzymes are produced by a wide variety of fungi and bacteria such as Trichodrema spp. [149,150], Penicillium spp. [151,152], Talaromyces spp. [151,153] Aspergillus spp. [154], and Bacillus spp. [155]. Enzymatic hydrolysis of xylan involves a multi-enzyme system, including endoxylanase, exoxylanase, -xylosidase, a-arabinofuranosidase, a-glucoronisidase, acetyl xylan esterase, and ferulic acid esterase [156]. Table 9 presents most important enzyme activities required for hydrolysis of xylooligosaccharides obtained from hardwoods and herbaceous type materials [113]. The endoxylanase attacks the main chains of xylans and b-xylosidase hydrolyzes xylooligosaccharides to xylose. The a-arabinofuranosidase and aglucuronidase remove the arabinose and 4-0-methyl glucuronic acid substituents, respectively, from the xylan backbone [157]. Hemicellulolytic esterases include acetyl esterases which hydro-

Fig. 10. Effects of T. reesei ZU-02 cellulase dosage (presented as lter paper activities per gram of substrate, FPU g1 substrate) on the enzymatic hydrolysis of dilute acid-treated corncob [146].

lyze the acetyl substitutions on xylose moieties, and feruloyl esterases which hydrolyze the ester bond between the arabinose substitutions and ferulic acid. Feruloyl esterases aid the release of hemicellulose from lignin and renders the free polysaccharide product more amenable to degradation by the other hemicellulases [158]. As in cellulase systems, xylan-degrading systems also exhibit. While the number of enzymes required for xylan hydrolysis is much greater than for cellulose hydrolysis, accessibility to the substrate is easier since xylan does not form tight crystalline structures [44]. 5.3. Fermentation The supernatant from enzymatic hydrolysis of lignocelluloses can contain both six-carbon (hexoses) and ve-carbon (pentoses) sugars (if both cellulose and hemicellulose are hydrolyzed). Depending on the lignocellulose source, the hydrolysate typically consists of glucose, xylose, arabinose, galactose, mannose, fucose,

M. Balat / Energy Conversion and Management 52 (2011) 858875 Table 9 Relevant enzymatic activities for enzymatic posthydrolysis of xylooligosaccharides [113]. Enzyme Endoxylanase Exoxylanase b-Xylosidase Arabinosidase Glucoronisidase Acetyl xylan esterases Feruloyl esterases n.c.: not yet classied. EC 3.2.1.8 n.c. 3.2.1.37 3.2.1.55 3.2.1.139 3.1.1.72 3.1.1.73 Hydrolyzed linkage Internal b-1,4 Terminal b-1,4 (reducing end) Terminal b-1,4 (non-reducing end) Substrate Main chain Main chain Oligomers Side groups Side groups Side groups Side groups Main product Oligomers Xylose, xylobiose Xylose Arabinose Methylglucuronic acids Acetic acid Ferulic acid

869

Ester bond Ester bond

Table 10 Important traits for bioethanol fermentation process [160]. Trait Bioethanol yield Bioethanol tolerance Bioethanol productivity Robust grower and simple growth requirements Able to grow in undiluted hydrolysates Culture growth conditions retard contaminants Requirement >90% of theoretical >40 g l1 >1 g l1 h1 Inexpensive medium formulation Resistance to inhibitors Acidic pH or higher temperatures

Table 11 Comparison between modied Z. mobilis and E. coli [94]. Z. mobilis Bioethanol (g l1) Bioethanol yielda (%) Bioethanol productivity (g l1 h1)
a

E. coli 27 90 0.92

62 97 1.29

Estimation from the theoretical yields.

and rhamnose [44]. One ton of glucan, galactan, or mannan yields 1.11 tons of six-carbon sugars and could be fermented theoretically into 172.0 gallons of bioethanol [159]. One ton of arabinan or xylan yields 1.14 tons of ve-carbon sugars and could be fermented theoretically into 176.0 gallons of bioethanol [159]. Microorganisms can be used to ferment all lignocellulose-derived sugars to bioethanol. 5.3.1. Microorganisms related to bioethanol fermentation Microorganisms for bioethanol fermentation can best be described in terms of their performance parameters and other requirements such as compatibility with existing products, processes and equipment. The performance parameters of fermentation are: temperature range, pH range, alcohol tolerance, growth rate, productivity, osmotic tolerance, specicity, yield, genetic stability, and inhibitor tolerance. The characteristics required for an industrially suitable microorganism are summarized in Table 10 [160]. Traditionally, Saccharomyces cerevisiae and Zymomonas mobilis have been used for bioethanol fermentation. They are capable of efciently fermenting glucose into bioethanol, but are unable to ferment xylose [44]. Natural xylose-fermenting yeasts, such as Pichia stipitis, Candida shehatae, and Candida parapsilosis, can metabolize xylose via the action of xylose reductase (XR) to convert xylose to xylitol, and of xylitol dehydrogenase (XDH) to convert xylitol to xylulose. Therefore, bioethanol fermentation from xylose can be successfully performed by recombinant S. cerevisiae carrying heterologous XR and XDH from P. stipitis, and xylulokinase (XK) from S. cerevisiae [161]. In bacteria, a xylose isomerase (XI) converts xylose to xylulose, which after phosphorylation, is metabolized through the pentose phosphate pathway (PPP) [162]. The most frequently used microorganism for fermenting bioethanol in industrial processes is S. cerevisiae, which has proved to be very robust and well suited to the fermentation of lignocellulosic hydrolysates [91]. S. cerevisiae can easily ferment hexoses, but hardly xylose in lignocellulose hydrolysates, because S. cerevisiae lacks enzymes that convert xylose to xylulose [163]. However, this yeast can ferment xylulose [164]. For xylose-using S. cerevisiae, high bioethanol yields from xylose also require metabolic engineering strategies to enhance the xylose ux [165]. Bacteria, such as Z. mobilis, Escherichia coli and Klebsiella oxytoca, have attracted particular interest, given their rapid fermentation,

which can be minutes compared to hours for yeasts [102]. Z. mobilis, a Gram-negative bacterium, is well recognized for its ability to efciently produce bioethanol at high rates from glucose, fructose, and sucrose. When Z. mobilis and S. cerevisiae were compared for their efciency to produce bioethanol from glucose and starch hydrolysate, higher yield was observed for Z. Mobilis [166]. Comparative performance trials on glucose have shown that Z. mobilis can achieve 5% higher bioethanol yields and up to 5-fold higher bioethanol volumetric productivity compared to traditional S. cerevisiae yeast [167]. It has a theoretical yield of 97% [94]. Z. mobilis efciently produces bioethanol from the hexose sugars glucose and fructose but not from pentose sugars, although a xylose fermenting Z. mobilis was generated by introducing a xylose-metabolizing pathway from E. coli [165]. Modied Z. mobilis has the advantages of requiring a minimum of nutrients, growing at low pH and high temperatures, and it is considered generally recognized as safe (GRAS). A comparison of modied Z. mobilis and E. coli showing their respective advantages is shown in Table 11 [94]. E. coli and K. oxytoca naturally metabolize arabinose, such that the ethanologenic strains ferment all lignocellulose-derived sugars [165]. The construction of E. coli strains to selectively produce bioethanol was one of the rst successful applications of metabolic engineering [168]. E. coli, as a biocatalyst for bioethanol production, has ability to ferment a wide spectrum of sugars, no requirements for complex growth factors, and prior industrial use (e.g. for production of recombinant protein). The major disadvantages associated with using E. coli cultures are a narrow and neutral pH growth range (6.08.0), less hardy cultures compared to yeast, and public perceptions regarding the danger of E. coli strains. The lack of data on the use of residual E. coli cell mass as an ingredient in animal feed is also an obstacle to its application [160,169]. K. oxytoca is an enteric bacterium found growing in paper and pulp streams as well as around other sources of wood. The microorganism is capable of growing at a pH at least as low as 5.0 and temperatures as warm as 308 K. It can grow on a wide variety of sugars including hexoses and pentoses, as well as on cellobiose and cellotriose [160]. Thermophilic anaerobic bacteria have also been extensively examined for their potential as bioethanol producers. These bacteria include Thermoanaerobacter ethanolicus [170], Clostridium thermohydrosulfuricum [171], Thermoanaerobacter mathranii [172], Thermoanaerobium brockii [173], Clostridium thermosaccharolyticum [174], etc. Thermophilic anaerobic bacteria have a distinct advantage over conventional yeasts for bioethanol production in their

870

M. Balat / Energy Conversion and Management 52 (2011) 858875

ability to use a variety of inexpensive biomass feedstocks and their ability to withstand temperature extremes [175]. The low bioethanol tolerance of thermophilic anaerobic bacteria (<2%, v/v) is a major obstacle for their industrial exploitation for bioethanol production [176]. 5.3.2. Fermentation techniques Fermentation can be performed as a batch, fed-batch or continuous process. The choice of most suitable process will depend upon the kinetic properties of microorganisms and type of lignocellulosic hydrolysate in addition to process economics aspects [66]. Batch culture can be considered as a closed culture system which contains an initial, limited amount of nutrient, which is inoculated with microorganisms to allow the fermentation [177]. It is very simple method, during the fermentation nothing is added after inoculation except possibly acid or alkali for pH control or air for aerobic fermentations. Fed-batch reactors are widely used in industrial applications because they combine the advantages from both batch and continuous processes [178]. The major advantage of fed-batch, comparing to batch, is the ability to increase maximum viable cell concentration, prolong culture lifetime, and allow product accumulation to a higher concentration [179]. This process allows for the maintenance of critical process variables (e.g. temperature, pH, and dissolved oxygen) at specic levels through feedback control [180]. In the continuous process, feed, which contains substrate, culture medium and other required nutrients, is pumped continuously into an agitated vessel where the microorganisms are active. The product, which is taken from the top of the bioreactor, contains bioethanol, cells, and residual sugar [181]. One of the rst approaches taken in improving the yeast bioethanol fermentation process involved operating the fermenters in a continuous mode rather than the more conventional batch mode and thereby increasing the productivity about threefold from about 2 to 6 g EtOH/l/h [182]. Operating continuously at higher cell densities using cell recycle reactors was another effective means of greatly increasing the productivity. A single-stage continuous stirred-tank reactor (CSTR) with cell recycle operating at high biomass loading (5080 g yeast/l) has a bioethanol productivity of 3040 g EtOH/l/h [182]. 5.3.3. Hydrolysis and fermentation strategies 5.3.3.1. Separate hydrolysis and fermentation (SHF). Enzymatic hydrolysis performed separately from fermentation step is known as separate hydrolysis and fermentation (SHF). In the SHF conguration the joint liquid ow from both hydrolysis reactors rst enters the glucose fermentation reactor. The mixture is then distilled to remove the bioethanol leaving the unconverted xylose behind. In a second reactor, xylose is fermented to bioethanol, and the bioethanol is again distilled [55]. The advantage of SHF is the ability to carry out each step under optimal conditions, i.e. enzymatic hydrolysis at 318323 K and fermentation at about 303 K [183]. The disadvantage of this method is the inhibition of cellulase and b-glucosidase enzymes by glucose released during hydrolysis, which calls for lower solids loadings and higher enzyme loadings to achieve reasonable yields [121]. 5.3.3.2. Simultaneous saccharication and fermentation (SSF). Simultaneous saccharication and fermentation (SSF) is a promising process option for production of bioethanol from lignocellulosic materials [184]. This process is often effective when combined with dilute acid or high temperature hot-water pretreatment. In SSF, cellulases and xylanases convert the carbohydrate polymers to fermentable sugars. These enzymes are notoriously susceptible to feedback inhibition by the products glucose, xylose, cellobiose,

and other oligosaccharides [123]. This process has an enhanced rate of hydrolysis, needs lower enzyme loading, results in higher bioethanol yields, and reduces the risk of contamination. Presently, an SSF process for e.g. wheat straw hydrolyzate can be expected to give nal bioethanol concentrations close to 40 g l1 with a yield based on total hexoses and pentoses higher than 70% [185]. SSF requires compatible fermentation and saccharication conditions, with a similar pH, temperature and optimum substrate concentration [186]. In many cases, the low pH, e.g. lower than 5, and high temperature, e.g. >313 K, may be favorable for enzymatic hydrolysis, whereas the low pH can surely inhabit the lactic acid production and the high temperature may affect adversely the fungal cell growth [187]. Trichoderma reesei cellulases, which constitute the most active preparations, have optimal activity at pH 4.5 and 328 K. For Saccharomyces cultures SSF are typically controlled at pH 4.5 and 310 K [160]. A typical fermentation will take 57 days, depending on the accessibility of the cellulose and initial solids loading of the fermentation. The long residence time may make contamination control difcult in a continuous process, but may be manageable in a batch process [188]. Major advantages of SSF as described by Sun and Cheng [74], include: (1) increase of hydrolysis rate by conversion of sugars that inhibit the cellulase activity; (2) lower enzyme requirement; (3) higher product yields; (4) lower requirements for sterile conditions since glucose is removed immediately and bioethanol is produced; (5) shorter process time; and (6) less reactor volume. The major advantage of SSF is that the immediate consumption of sugars by the microorganism produces low sugar concentrations in the fermentor, which signicantly reduces enzyme inhibition compared to SHF [188]. The main disadvantage of SSF lies in different temperature optima for saccharication (323 K) and fermentation (308 K) [189]. The most important process improvement made for the enzymatic hydrolysis of biomass is the introduction of SSF, which has been improved to include combines the cellulase enzymes and fermenting microbes in one vessel to improve the bioethanol production economics. The technology has been improved to include the co-fermentation of multiple sugar substrates, i.e., simultaneous saccharication of both cellulose (to glucose) and hemicellulose (to xylose), and co-fermentation of both glucose and xylose by genetically engineered microbes in the same broth [136]. 5.3.3.3. Direct microbial conversion (DMC). Direct microbial conversion (DCM) combines cellulase production, cellulose hydrolysis and glucose fermentation into a single step. This process is attractive in that it reduces the number of reactors, simplies operation, and reduces the cost of chemicals [121]. DCM seems the logical endpoint in the evolution of bioethanol production from lignocellulosic materials. Application of DCM entails no operating cost or capital investment for dedicated enzyme production (or purchase), reduced diversion of substrate for enzyme production, and compatible enzyme and fermentation systems [55]. The disadvantages are low bioethanol yields, caused by byproduct formation (acetate, lactate), low tolerance of the microorganism to bioethanol (3.5% w/ v), and limited growth in hydrolysate syrups [162]. 5.4. Product and solids recovery As biomass hydrolysis, and fermentation technologies approach commercial viability, advancements in product recovery technologies will be required. For cases in which fermentation products are more volatile than water, recovery by distillation is often the technology of choice. Distillation technologies that will allow the economic recovery of dilute volatile products from streams containing a variety of impurities have been developed and

M. Balat / Energy Conversion and Management 52 (2011) 858875

871

commercially demonstrated [190]. A distillation system separates the bioethanol from water in the liquid mixture. The rst step is to recover the bioethanol in a distillation or beer column, where most of the water remains with the solids part. The product (37% bioethanol) is then concentrated in a rectifying column to a concentration just below the azeotrope (95%) [55]. The remaining bottoms product is fed to the stripping column to remove additional water, with the bioethanol distillate from stripping being recombined with the feed to the rectier [191]. The recovery of bioethanol in the distillation columns in the plant is xed to be 99.6% to reduce bioethanol losses [192]. After the rst effect, solids are separated using a centrifuge and dried in a rotary dryer. A portion (25%) of the centrifuge efuent is recycled to fermentation and the rest is sent to the second and third evaporator effects. Most of the evaporator condensate is returned to the process as fairly clean condensate (a small portion, 10%, is split off to waste water treatment to prevent build up of low-boiling compounds) and the concentrated syrup contains 1520% by weight total solids [193].

Table 12 Estimates of the costs of bioethanol production from different feedstock (US$/l) [201]. Feedstock Sugarcane Corn Sugar beet Wheat Lignocellulose
a

Production cost rangea in 2005 0.200.50 0.600.80 0.620.82 0.700.95 0.801.10

Projected production cost rangea in 2030 0.200.35 0.350.55 0.400.60 0.450.65 0.250.65

Excluding any subsidies to bioethanol production.

6. Bioethanol economy Considering that up to now the cost of bioethanol was considerably higher than the cost of fossil gasoline supply, national governments had to enact special policies in order to encourage production and use of bioethanol in the transportation sector. In general, the following three main approaches can be distinguished in the implementation of biofuels supporting policies and regulation: (1) taxation-based policies, (2) agriculture-based policies/ subsidies, and (3) fuel mandates [34]. At present, the development and promotion of biofuels are mainly driven by the agricultural sector and green lobbies rather than the energy sector. In fact, most biofuel programs depend on subsidies and government programs, which can lead to market distortion and is costly for governments. Nevertheless, at sustained high oil prices and with a steady progression of more efcient and cheaper technology, biofuels could be a cost-effective alternative in the near future in many countries [194]. The cost for bioethanol production can vary substantially depending on several factors, e.g. feedstock costs and by-products revenues, cost of process energy, investment costs (related to the type of feedstock), plant location and transportation cost and nancing costs [195]. Brazilian bioethanol is far more competitive than that produced in the United States from corn or in Europe from sugar beet, because of shorter processing times, lower labour costs, lower transport costs and input costs [196]. Bioethanol production from sugarcane is very economical in Brazil because of two primary reasons. Brazil dropped support of sugar prices to support the bioethanol industry with government established mandates for the blending of bioethanol with gasoline. This drastically lowered the cost of the feedstock, sugarcane, and created a demand for and supported the price of bioethanol. In addition, Brazils vast land area of cultivatable acreage means that land devoted to sugarcane production for bioethanol is not in competition with land devoted for food production [197]. Bioethanol from sugarcane in Brazil costs US$0.230.29/l [198], while in the EU and the United States sugar beet and corn-derived bioethanol cost US$0.29/l [199] and US$0.53/l [200], respectively. Other efcient sugar producing countries such as Pakistan, Swaziland and Zimbabwe have production costs similar to Brazils [194]. The cost of raw material, which varies considerably between different studies (US$22US$61 per metric ton dry matter), and the capital costs, which makes the total cost dependent on plant capacity, contribute most to the total production cost [165]. The cost and availability of feedstock was crucial because in most bio-

fuels, feedstock represents 6075% of the total bioethanol production cost [3]. Estimates of the costs of bioethanol production from different feedstock are shown in Table 12. The cost gures can be compared with the cost of producing gasoline of around US$0.70/ l at oil prices of US$100 per barrel [201]. Bioethanol production generally utilizes derivatives from food crops such as corn grain and sugarcane, but the limited supply of these crops can lead to competition between their use in bioethanol production and food provision [202]. Using food crops to produce bioethanol raises major nutritional and ethical concerns. Nearly 60% of humans in the world are currently malnourished, so the need for grains and other basic foods is critical. Growing crops for fuel squanders land, water, and energy resources vital for the production of food for people [203]. In 2007, when US retail food prices rose 4% above 2006 levels and twice as fast as overall core ination (2.3%), consumers took notice [204]. The bioethanol-driven surge in corn demand has fueled a sharp rise in corn prices. For example, the futures contract for March 2007 corn on the Chicago Board of Trade, rose from US$2.50 per bushel in September 2006 to a contract high of over US$4.16 per bushel in January 2007 (a rise of 66%). This sharp rise in corn prices owes its origins largely to increasing corn demand spurred by the rapid expansion of corn-based bioethanol production capacity in the United States since mid-2006 [205]. Higher corn prices were, in part, driven by demand to make bioethanol and these higher prices effectively bid acres away from other crops that provided lower returns, such as soybeans, wheat, and hay [204]. Using corn for bioethanol increases the price of US beef, chicken, pork, eggs, breads, cereals, and milk from 10% to 30% [203]. Lignocellulosic biomass is the most promising feedstock considering its great availability and low cost, but the large-scale commercial production of fuel bioethanol from lignocellulosic materials has still not been implemented. Today the production cost of bioethanol from lignocellulose is still too high, which is the major reason why bioethanol has not made its breakthrough yet. Pretreatment has been viewed as one of the most expensive processing steps in cellulosic biomass to fermentable sugars conversion with costs as high as US$0.08/l bioethanol produced [86]. Enzyme pricing is assumed such that the total contribution of enzymes to production costs is about US$0.04/l of bioethanol with some variation depending upon actual bioethanol yields resulting from the particular pretreatment approach [206]. Signicant growth of the bioethanol industry will depend on the development of new processes that convert cellulosic biomass from non-food crops and waste materials into bioethanol [207]. 7. Conclusion Recently, there has been growing interest in biofuels due to the rising energy costs and environmental problems. Bioethanol is by far the most widely used biofuel for transportation worldwide. It will continue to be developed as a transport fuel produced in tropical latitudes and traded internationally, for use primarily as a gasoline additive.

872

M. Balat / Energy Conversion and Management 52 (2011) 858875 [20] Demirbas A, Karslioglu S. Biodiesel production facilities from vegetable oils and animal fats. Energy Source A 2007;29:13341. [21] Lang X, Macdonald DG, Hill GA. Recycle bioreactor for bioethanol production from wheat starch II. Fermentation and economics. Energy Source 2001;23:42736. [22] Demirbas MF, Balat M, Balat H. Potential contribution of biomass to the sustainable energy development. Energy Convers Manage 2009;50:174660. [23] Brjesson P. Good or bad bioethanol from a greenhouse gas perspective What determines this? Appl Energy 2009;86:58994. [24] Demirbas A. Producing and using bioethanol as an automotive fuel. Energy Source B 2007;2:391401. [25] Balat M. Bioethanol as a vehicular fuel: a critical review. Energy Source A 2009;31:124255. [26] elik MB. Experimental determination of suitable ethanolgasoline blend rate at high compression ratio for gasoline engine. Appl Therm Eng 2008;28:396404. [27] Demirbas A. Fuel alternatives to gasoline. Energy Source B 2007;2:31120. [28] Stokes H. Alcohol fuels (ethanol and methanol): safety. In: Presentation at ETHOS conference 2005, Seattle, Washington, January 2930; 2005. urovic ic LC, Vuc V, Pejin J, Denc S, Rakin M. Fermentation of [29] Pejin D, Mojovic wheat and triticale hydrolysates: a comparative study. Fuel 2009;88:16258. [30] Balat M. Global status of biomass energy use. Energy Source A 2009;31: 116073. [31] Philippidis G. The potential of biofuels in the Americas. Energy cooperation and security in the hemisphere task force, center for hemispheric policy The University of Miami, July 24; 2008. <www6.miami.edu/hemispheric-policy/ Philippidis> [accessed October 2009]. [32] Gnansounou E, Dauriat A. Ethanol fuel from biomass: a review. J Sci Ind Res 2005;64:80921. [33] Mojovic L, Pejin D, Grujic O, Pejin J, Rakin M, Vukasinovic M, et al. Progress in the production of bioethanol on starch-based feedstock. Chem Ind Chem Eng Q 2009;15:21126. [34] Gnansounou E, Bedniaguine D, Dauriat A. Promoting bioethanol production through clean development mechanism: ndings and lessons learnt from ASIATIC project. In: Proceedings of the 7th IAEE European energy conference, Bergen, Norway, August 2830; 2005. [35] Hartemink AE. Sugarcane for bioethanol: soil and environmental issues. Adv Agron 2008;99:12582. [36] Prabhakar SVRK, Elders M. CO2 Reduction potential of biofuels in Asia: issues and policy implications. In: International conference on energy security and climate change: issues, strategies, and options (ESCC 2008), Bangkok, Thailand, August 68; 2009. [37] Trostle R. Global agricultural supply and demand: factors contributing to the recent increase in food commodity prices. USDA economic research service, report WRS-0801, Washington, DC; July 2008. [38] Asher A. Opportunities in biofuels creating competitive biofuels markets. In: Biofuels Australasia 2006 conference, Sydney, Australia, November 2022; 2006. [39] Kline KL, Oladosu GA, Wolfe AK, Perlack RD, Dale VH, McMahon M. Biofuel feedstock assessment for selected countries. Oak Ridge national laboratory technical report, no. ORNL/TM-2007/224, Tennessee; 2008. <www.osti.gov/ bridge/servlets/purl/924080-y8ATDg/924080> [accessed October 2009]. [40] Yoosin S, Sorapipatana C. A study of ethanol production cost for gasoline substitution in Thailand and its competitiveness. Thammasat Int J Sci Technol 2007;12:6980. [41] Slade R, Bauen A, Shah N. The greenhouse gas emissions performance of cellulosic ethanol supply chains in Europe. Biotechnol Biofuels 2009;2:119. [42] Yao R, Qi B, Deng S, Liu N, Peng S, Cui Q. Use of surfactants in enzymatic hydrolysis of rice straw and lactic acid production from rice straw by simultaneous saccharication and fermentation. BioResources 2007;2:38998. [43] Wang L, Hanna MA, Weller CL, Jones DD. Technical and economical analyses of combined heat and power generation from distillers grains and corn stover in ethanol plants. Energy Convers Manage 2009;50:170413. [44] Keshwani DR, Cheng JJ. Switchgrass for bioethanol and other value-added applications: a review. Bioresour Technol 2009;100:151523. [45] Carvalho LG, Gomes AM, Aranda DAG, Pereira N. Ethanol from lignocellulosic residues of palm oil industry. In: 11th International conference on advanced materials, Rio de Janeiro, Brazil, September 2025; 2009. [46] Bohlmann GM. Process economic considerations for production of ethanol from biomass feedstocks. Ind Biotechnol 2006;2:1420. [47] Karimi K, Emtiazi G, Taherzadeh MJ. Ethanol production from dilute-acid pretreated rice straw by simultaneous saccharication and fermentation with Mucor indicus, Rhizopus oryzae, and Saccharomyces cerevisiae. Enzyme Microbiol Technol 2006;40:13844. [48] Lee D, Owens VN, Boe A, Jeranyama P. Composition of herbaceous biomass feedstocks. South Dakota State University Publication, SGINC1-07, Brookings, SD; June 2007. [49] Molina-Sabio M, Rodrguez-Reinoso F. Role of chemical activation in the development of carbon porosity. Colloid Surface Physicochem Eng Aspect 2004;241:1525. [50] Jenkins BM, Baxter LL, Miles Jr TR, Miles TR. Combustion properties of biomass. Fuel Process Technol 1998;54:1746. [51] Dehkhoda A. Concentrating lignocellulosic hydrolysate by evaporation and its fermentation by repeated fedbatch using occulating Saccharomyces cerevisiae. Master thesis, Industrial Biotechnology Boras University and SEKAB E-Technology, Sweden; 2008.

Bioethanol production generally utilizes derivatives from food crops such as corn grain and sugarcane, but the limited supply of these crops can lead to competition between their use in bioethanol production and food provision. The price of the raw materials is also highly volatile, which can highly affect the production costs of the bioethanol. Lignocellulosic materials serve as a cheap and abundant feedstock, which is required to produce fuel bioethanol from renewable resources at reasonable costs. Lignocellulosic biomass can be converted to bioethanol by hydrolysis and subsequent fermentation. Lignocellulose is often hydrolyzed by acid treatment; the hydrolysate obtained is then used for bioethanol fermentation by microorganisms such as yeast. Because such lignocellulose hydrolysate contains not only glucose, but also various monosaccharides (e.g. xylose, mannose, fructose, galactose, and arabinose) and oligosaccharides, microorganisms should be required to efciently ferment these sugars for the successful industrial production of bioethanol. Acknowledgement The author would like to thank Professor Ayhan Demirbas for his very considerable help and encouragement throughout the course of this work. References
[1] International Energy Agency (IEA). Key world energy statistics 2008. Paris: OECD/IEA; 2008. [2] Goldemberg J. Environmental and ecological dimensions of biofuels. In: Proceedings of the conference on the ecological dimensions of biofuels, Washington, DC, March 10; 2008. [3] Balat M, Balat H. Recent trends in global production and utilization of bioethanol fuel. Appl Energy 2009;86:227382. [4] Plunkett JW. Plunketts automobile industry Almanac 2008: automobile, truck and specialty vehicle industry market research, statistics, trends & leading companies. Plunkett Research Ltd., Houston, Texas; 2008. <www.amazon.com/ Plunketts-Automobile-Industry-Almanac-2008> [accessed October 2009]. [5] World Business Council for Sustainable Development (WBCSD). Mobility 2030: meeting the challenges to sustainability. The sustainable mobility project, Geneva, Switzerland; 2004. <www.wbcsd.ch/web/publications/ mobility> [accessed October 2009]. [6] Meher LC, Sagar DV, Naik SN. Technical aspects of biodiesel production by transestericationa review. Renew Sustain Energy Rev 2006;10:24868. [7] Demirbas A. Biofuels sources, biofuel policy, biofuel economy and global biofuel projections. Energy Convers Manage 2008;49:210616. [8] Forum for Agricultural Research in Africa (FARA). Bioenergy value chain research and development Stakes and Opportunities. Burkina Faso, FARA Discussion Paper; April 2008. <www.fara-africa.org/media/.../Bioenergy_ Discussion_Paper_April_2008> [accessed November 2009]. [9] Demirbas AH, Demirbas I. Importance of rural bioenergy for developing countries. Energy Convers Manage 2007;48:238698. [10] 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. [11] Balat M, Balat M, Krtay E, Balat H. Main routes for the thermo-conversion of biomass into fuels and chemicals. Part 1: pyrolysis systems. Energy Convers Manage 2009;50:314757. [12] De Kam MJ, Morey RV, Tiffany DG. Biomass integrated gasication combined cycle for heat and power at ethanol plants. Energy Convers Manage 2009;50:168290. [13] Balat M, Balat M, Krtay E, Balat H. Main routes for the thermo-conversion of biomass into fuels and chemicals. Part 2: gasication systems. Energy Convers Manage 2009;50:315868. [14] Liu Z, Zhang FS. Effects of various solvents on the liquefaction of biomass to produce fuels and chemical feedstocks. Energy Convers Manage 2008;49:3498504. [15] Sener U, Genel Y, Saka C, Kilicel F, Kucuk MM. Supercritical uid extraction of cotton stalks. Energy Source A 2010;32:205. [16] Demirbas A. Diesel-like fuel from tallow by pyrolysis and supercritical water liquefaction. Energy Source A 2009;31:82430. [17] Mohanty SK, Behera S, Swain MR, Ray RC. Bioethanol production from mahula (Madhuca latifolia L.) owers by solid-state fermentation. Appl Energy 2009;86:6404. [18] Balat H. Prospects of biofuels for a sustainable energy future: A critical assessment. Energy Educ Sci Technol Part A 2010;24:85111. [19] Demirbas A. Biofuels securing the planets future energy needs. Energy Convers Manage 2009;50:223949.

M. Balat / Energy Conversion and Management 52 (2011) 858875 [52] Balat M. Gasication of biomass to produce gaseous products. Energy Source A 2009;31:51626. [53] Demirbas A. Bioethanol from cellulosic materials: a renewable motor fuel from biomass. Energy Source 2005;27:32737. [54] Karunanithy C, Muthukumarappan K, Julson JL. Enzymatic hydrolysis of corn stover pretreated in high shear bioreactor. In: ASABE Annual international meeting, Paper No. 084114, Rhode Island, June 29July 2; 2008. [55] Hamelinck CN, Van Hooijdonk G, Faaij APC. Ethanol from lignocellulosic biomass: techno-economic performance in short-, middle- and long-term. Biomass Bioenergy 2005;28:384410. [56] Taherzadeh MJ, Karimi K. Pretreatment of lignocellulosic wastes to improve ethanol and biogas production: a review. Int J Mol Sci 2008;9:162151. [57] Demirbas A. Heavy metal adsorption onto agro-based waste materials: a review. J Hazard Mater 2008;157:2209. [58] Feldman D, Banu D, Natansohn A, Wang J. Structureproperties relations of thermally cured epoxy-lignin polyblends. J Appl Polym Sci 1991;42:153750. [59] Demirbas A. Products from lignocellulosic materials via degradation processes. Energy Source A 2008;30:2737. [60] Yaman S. Pyrolysis of biomass to produce fuels and chemical feedstocks. Energy Convers Manage 2004;45:65171. [61] Cardona Alzate CA, Sanchez Toro OJ. Energy consumption analysis of integrated ow sheets for production of fuel ethanol from ligno-cellulosic biomass. Energy 2006;31:244759. [62] Zhu JY, Wang GS, Pan XJ, Gleisner R. The status of and key barriers in lignocellulosic ethanol production: a technological perspective. In: International conference on biomass energy technologies, Guangzhou, China, December 35; 2008. [63] Hsu TA, Ladisch MR, Tsao GT. Alcohol from cellulose. Chem Technol 1980;10:3159. [64] Kodali B, Pogaku R. Pretreatment studies of rice bran for the effective production of cellulose. Electron J Environ Agric Food Chem 2006;5:125364. [65] Zhang Y, Pan Z, Zhang R. Overview of biomass pretreatment for cellulosic ethanol production. Int J Agric Biol Eng 2009;2:5168. [66] Chandel AK, Es C, Rudravaram R, Narasu ML, Rao LV, Ravindra P. Economics and environmental impact of bioethanol production technologies: an appraisal. Biotechnol Mol Biol Rev 2007;2:1432. [67] Taherzadeh MJ, Karimi K. Enzymatic-based hydrolysis processes for ethanol from lignocellulosic materials: a review. BioResources 2007;2:70738. [68] Kumar P, Barrett DM, Delwiche MJ, Stroeve P. Methods for pretreatment of lignocellulosic biomass for efcient hydrolysis and biofuel production. Ind Eng Chem Res 2009;48:371329. [69] Chen M, Zhao J, Xia L. Comparison of four different chemical pretreatments of corn stover for enhancing enzymatic digestibility. Biomass Bioenergy 2009;33:13815. [70] Sun F, Chen H. Enhanced enzymatic hydrolysis of wheat straw by aqueous glycerol pretreatment. Bioresour Technol 2008;99:615661. [71] Esteghlalian A, Hashimoto AG, Fenske JJ, Penner MH. Modeling and optimization of the dilute-sulfuric-acid pretreatment of corn stover, poplar and switchgrass. Bioresour Technol 1997;59:12936. [72] Zhang Q, Cai W. Enzymatic hydrolysis of alkali-pretreated rice straw by Trichoderma reesei ZM4-F3. Biomass Bioenergy 2008;32:11305. [73] Rabelo SC, Filho RM, Costa AC. Lime pretreatment of sugarcane bagasse for bioethanol production. Appl Biochem Biotechnol 2009;153:13953. [74] Sun Y, Cheng J. Hydrolysis of lignocellulosic materials for ethanol production: a review. Bioresour Technol 2002;83:111. [75] Leustean I. Bioethanol from lignocellulosic materials. J Agroalimentary Process Technol 2009;15:94101. [76] Zheng Y, Pan Z, Zhang R. Overview of biomass pretreatment for cellulosic ethanol production. Int J Agric Biol Eng 2009;2:5168. [77] Cadoche L, Lopez GD. Assessment of size reduction as a preliminary step in the production of ethanol from lignocellulosic wastes. Biol Wastes 1989;30:1537. [78] McMillan JD. Pretreatment of lignocellulosic Biomass. In: Himmel ME, Baker JO, Overend RP, editors. Enzymatic conversion of biomass for fuels production, ACS Symposium Series 566. American Chemical Society, Washington, DC; 1994. p. 292324. [79] Ruffell J. Pretreatment and hydrolysis of recovered bre for ethanol production. Master of Applied Science, The University of British Columbia; August 2008. [80] Wyman C. Handbook on bioethanol: production and utilization. Washington, DC: Taylor and Francis; 1996. [81] Tomas-Pejo E, Olive JM, Ballesteros M. Realistic approach for full-scale bioethanol production from lignocellulose: a review. J Sci Ind Res 2008;67:87484. [82] Pasha C, Rao LV. Thermotolerant yeasts for bioethanol production using lignocellulosic substrates. In: Satyanarayana T, Kunze G, editors. Yeast biotechnology: diversity and applications. Netherlands: Springer; 2009. p. 55188. [83] Jeoh T. Steam explosion pretreatment of cotton gin waste for fuel ethanol production. Masters thesis (Agblevor FA, Chen JS, Helm RF advs.), Virginia Tech University, VA; 1998. [84] Prasad S, Singh A, Joshi HC. Ethanol as an alternative fuel from agricultural, industrial and urban residues. Resour, Conserv Recycl 2007;50:139. [85] Heerah AS, Mudhoo A, Mohee R, Sharma SK. Steam pre-treatment of lignocellulosic wastes for biomethanogenesis: a preliminary study. Rasayan J Chem 2008;1:50314.

873

[86] Mosier N, Wyman C, Dale B, Elander R, Holtzapple YYLM, Ladisch M. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour Technol 2005;96:67386. [87] Negro MJ, Manzanares P, Ballesteros I, Oliva JM, Cabanas A, Ballesteros M. Hydrothermal pretreatment conditions to enhance ethanol production from poplar biomass. Appl Biochem Biotechnol 2003;105:87100. [88] Bertilsson M. Simultaneous saccharication and fermentation of spruce a comparison of pretreatment conditions and different enzyme preparations. Master thesis, Department of Chemical Engineering, Lund University, Lund, Sweden; February 2007. [89] Ballesteros I, Negro MJ, Oliva JM, Cabanas A, Manzanares P, Ballesteros M. Ethanol production from steam-explosion pretreated wheat straw. Appl Biochem Biotechnol 2006;130:496508. [90] Sassner P, Martensson CG, Galbe M, Zacchi G. Steam pretreatment of H2SO4impregnated Salix for the production of bioethanol. Bioresour Technol 2008;99:13745. [91] Galbe M, Zacchi G. A review of the production of ethanol from softwood. Appl Microbiol Biotechnol 2002;59:61828. [92] Sderstrm J, Pilcher L, Galbe M, Zacchi G. Two-step steam pretreatment of softwood by dilute H2SO4 impregnation for ethanol production. Biomass Bioenergy 2003;24:47586. [93] Sderstrm J, Galbe M, Zacchi G. Effect of washing on yield in one- and twostep steam pretreatment of softwood for production of ethanol. Biotechnol Prog 2008;20:7449. [94] Abril D, Abril A. Ethanol from lignocellulosic biomass. Cien Inv Agr 2009;36:17790. [95] Dale MC, Moelhman M. Enzymatic simultaneous saccharication and fermentation (SSF) of biomass to ethanol in a pilot 130 liter multistage continuous reactor separator. In: Ninth biennial bioenergy conference, Buffalo, New York, October 1519; 2000. [96] Alizadeh H, Teymouri F, Gilbert TI, Dale BE. Pretreatment of switchgrass by ammonia ber explosion (AFEX). Appl Biochem Biotechnol 2005;121:113341. [97] Teymouri F, Laureano-Prez L, Alizadeh H, Dale BE. Ammonia ber explosion treatment of corn stover. Appl Biochem Biotechnol 2004;115:95163. [98] Holtzapple MT, Jun J-H, Ashok G, Patibandla SL, Dale BE. The ammonia freeze explosion (AFEX) process: a practical lignocellulose pretreatment. Appl Biochem Biotechnol 1991;28(/29):5974. [99] Balan V, Sousa LC, Chundawat SPS, Marshall D, Sharma LN, Chambliss CK, et al. Enzymatic diagestibility and pretreatment degradation products of AFEX-treated hardwoods (Populus nigra). Biotechnol Prog 2009;25: 36575. [100] Lee YJ. Oxidation of sugarcane bagasse using a combination of hypochlorite and peroxide. Masters thesis (Day DF, Godber JS, Janes ME, advs), Department of Food Science, Graduate Faculty of the Louisiana State University and Agricultural and Mechanical College, October 28; 2005. [101] Hu G, Heitmann JA, Rojas OJ. Feedstock pretreatment strategies for producing ethanol from wood, bark and forest residues. BioResources 2008;3:27094. [102] Hayes DJ. An examination of biorening processes, catalysts and challenges. Catal Today 2009;145:13851. [103] Vidal PF, Molinier J. Ozonolysis of lignin improvement of in vitro digestibility of poplar sawdust. Biomass 1988;16:117. [104] Neeley WC. Factor affecting the pretreatment of bio-mass with gaseous ozone. Biotechnol Bioeng 1984;26:5965. [105] Williams KC. Subcritical water and chemical pretreatments of cotton stalk for the production of ethanol. Masters thesis, Biological and Agricultural Engineering, North Carolina State University, Raleigh, NC, USA; 2006. [106] Garca-Cubero MT, Gonzalez-Benito G, Indacoechea I, Coca M, Bolado S. Effect of ozonolysis pretreatment on enzymatic digestibility of wheat and rye straw. Bioresour Technol 2009;100:160813. [107] Silverstein RA, Chen Y, Sharma-Shivappa RR, Boyette MD, Osborne J. A comparison of chemical pretreatment methods for improving saccharication of cotton stalks. Bioresour Technol 2008;98:300011. [108] Han M, Moon SK, Kim Y, Kim Y, Chung B, Choi GW. Bioethanol production from ammonia percolated wheat straw. Biotechnol Bioprocess Eng 2009;14:60611. [109] Fang LT, Gharpuray MM, Lee YH. Cellulose hydrolysis biotechnology monographs. Berlin, Germany: Springer; 1987. p. 55. [110] Millet MA, Baker AJ, Scatter LD. Physical and chemical pretreatment for enhancing cellulose. Appl Microbiol Biotechnol 1976;29:4628. [111] Akhtar MS, Saleem M, Akhtar MW. Saccharication of lignocellulosic materials by the cellulases of Bacillus subtilis. Int J Agr Biol 2001;3:199202. [112] Li Y, Ruan R, Chen PL, Liu Z, Pan X, Lin X, et al. Enzymatic hydrolysis of corn stover pretreated by combined dilute alkaline treatment and homogenization. Trans ASAE 2004;47:8215. [113] Carvalheiro F, Duarte LC, Girio FM. Hemicellulose bioreneries: a review on biomass pretreatments. J Sci Ind Res 2008;67:84964. [114] Wang Z, Cheng JJ. Lime pretreatment of coastal bermudagrass for bioethanol production. In: ASABE meeting, Reno, Nevada, June 2124; 2009. [115] Balat M, Balat H, Oz C. Progress in bioethanol processing. Progr Energy Combust Sci 2008;34:55173. [116] Ramirez RS. Long-term lime pretreatment of poplar wood. Masters thesis, Texas A&M University; December 2005. [117] Chang VS, Burr B, Holtzapple MT. Lime pretreatment of switchgrass. Appl Biochem Biotechnol 1997;6365:319. [118] Saha BC, Cotta MA. Enzymatic hydrolysis and fermentation of lime pretreated wheat straw to ethanol. J Chem Tech Biotechnol 2007;82:9139.

874

M. Balat / Energy Conversion and Management 52 (2011) 858875 [152] Jorgensen H, Eriksson T, Borjesson J, Tjerneld F, Olsson L. Purication and characterization of ve cellulases and one xylanase from Penicillium brasilianum IBT 20888. Enzyme Microb Technol 2003;32:85161. [153] Tuohy MG, Puls J, Claeyssens M, Vrsanska M, Coughlan MP. The xylandegrading enzyme system of Talaromyces emersonii: novel enzymes with activity against aryl beta-D-xylosides and unsubstituted xylans. Biochem J 1993;290:51523. [154] Dos Reis S, Costa MAF, Peralta RM. Xylanase production by a wild strain of Aspergillus nidulans. Acta Sci: Biol Sci 2003;1:2215. [155] Virupakshi S, Babu KG, Gaikwad SR, Naik GR. Production of a xylanolytic enzyme by a thermoalkaliphilic Bacillus sp. JB-99 in solid state fermentation. Process Biochem 2005;40:4315. [156] Saha BC. Lignocellulose biodegradation and applications in biotechnology. ACS Symp Ser 2004;889:234. [157] Saha BC. Hemicellulose bioconversion. J Ind Microbiol Biotechnol 2003;30:27991. [158] Howard RL, Abotsi E, Jansen van Rensburg EL, Howard S. Lignocellulose biotechnology: issues of bioconversion and enzyme production. Afr J Biotechnol 2003;2:60219. [159] Szulczyk KR, McCarl BA, Cornforth G. Market penetration of ethanol. Renew Sustain Energy Rev 2010;14:394403. [160] Dien BS, Cotta MA, Jeffries TW. Bacteria engineered for fuel ethanol production: current status. Appl Microbiol Biotechnol 2003;63:25866. [161] Katahira S, Mizuike A, Fukuda H, Kondo A. Ethanol fermentation from lignocellulosic hydrolysate by a recombinant xyloseand cellooligosaccharide-assimilating yeast strain. Appl Microbiol Biotechnol 2006;72:113643. [162] Zaldivar J, Nielsen J, Olsson L. Fuel ethanol production from lignocellulose: a challenge for metabolic engineering and process integration. Appl Microbiol Biotechnol 2001;56:1734. [163] Tian S, Zang J, Pan Y, Liu J, Yuan Z, Yan Y, et al. Construction of a recombinant yeast strain converting xylose and glucose to ethanol. Front Biol China 2008;3:1659. [164] Shi XQ, Jeffries TW. Anaerobic growth and improved fermentation of Pichia stipitis bearing a URA1 gene from Saccharomyces cerevisiae. Appl Microbiol Biotechnol 1998;50:33945. [165] Hahn-Hagerdal B, Galbe M, Gorwa-Grauslund MF, Liden G, Zacchi G. Bioethanol the fuel of tomorrow from the residues of today. Trends Biotechnol 2006;24:54956. [166] Gunasekaran P, Raj KC. Ethanol fermentation technology Zymomonas mobilis. Curr Sci 1999;77:5668. [167] Saez-Miranda JS, Saliceti-Piazza L, McMillan JD. Measurement and analysis of intracellular ATP levels in metabolically engineered Zymomonas mobilis fermenting glucose and xylose mixtures. Biotechnol Prog 2008;22:35968. [168] Millichip RJ, Doelle HW. Large-scale ethanol production from Milo (Sorghum) using Zymomonas mobilis. Process Biochem 1989;24:1415. [169] Lin Y, Tanaka S. Ethanol fermentation from biomass resources: current state and prospects. Appl Microbiol Biotechnol 2006;69:62742. [170] Avci A, Donmez S. Effect of zinc on ethanol production by two Thermoanaerobacter strains. Process Biochem 2006;41:9849. [171] Cook GM, Morgan HW. Hyperbolic growth of Thermoanaerobacter thermohydrosulfuricus (Clostridium thermohydrosulfuricum) increases ethanol production in pH-controlled batch culture. Appl Microbiol Biotechnol 1994;41:849. [172] Larsen L, Nielsen P, Ahring BK. Thermoanaerobacter mathranii sp. Nov., an ethanol-producing, extremely thermophilic anaerobic bacterium from a hot spring in Iceland. Arch Microbiol 1997;168:1149. [173] Lamed R, Zeikus JG. Glucose fermentation pathway of Thermoanaerobium brockii. J Bacteriol 1980;14:12517. [174] Baskaran S, Ahn HJ, Lynd LR. Investigation of the ethanol tolerance of Clostridium thermosaccharolyticum in continuous culture. Biotechnol Prog 1995;11:27681. [175] Knutson BL, Strobel HJ, Nokes SE, Dawson KA, Berberich JA, Jones CR. Effect of pressurized solvents on ethanol production by the thermophilic bacterium Clostridium thermocellum. J Supercritical Fluids 1999;16:14956. [176] Georgieva TI, Skiadas IV, Ahring BK. Effect of temperature on ethanol tolerance of a thermophilic anaerobic ethanol producer Thermoanaerobacter A10: modeling and simulation. Biotechnol Bioeng 2007;98:116170. [177] Abtahi Z. Ethanol and glucose tolerance of M. indicus in aerobic and anaerobic conditions. Master thesis, University College of Boras School of Engineering, Boras, Sweden; 2008. [178] Saarela U, Leiviska K, Juuso E. Modelling of a fed-batch fermentation process. Control Engineering Laboratory, Department of Process and Environmental Engineering, University of Oulu, Report A No. 21, Finland; June 2003. [179] Frison A, Memmert K. Fed-batch process development for monoclonal antibody production with cellferm-pro. Gen Eng News 2002;22:667. [180] Gunther JC, Seborg DE, Baclaski J. Fault detection and diagnosis in industrial fed-batch fermentation. In: 2006 American control conference, Minneapolis, MN, June 1416; 2006. [181] Caylak B, Sukan FV. Comparison of different production processes for bioethanol. Turk J Chem 1998;22:3519. [182] Lawford HG. A new approach to improving the performance of Zymomonas in continuous ethanol fermentations. Appl Biochem Biotechnol 1988;17:20319. [183] Tengborg C, Galbe M, Zacchi G. Inuence of enzyme loading and physical parameters on the enzymatic hydrolysis of steam-pretreated softwood. Biotechnol Prog 2008;17:1107.

[119] Karr WE, Holtzapple MT. Using lime pretreatment to facilitate the enzymatic hydrolysis of corn stover. Biomass Bioenergy 2000;18:18999. [120] Chang VS, Nagwani M, Kim CH, Holtzapple MT. Oxidative lime pretreatment of high-lignin biomass. Appl Biochem Biotechnol 2001;94:128. [121] Silverstein RA. A comparison of chemical pretreatment methods for converting cotton stalks to ethanol. Masters thesis (Sharma R, adv.), Biological and Agricultural Engineering, North Carolina State University, December 16; 2004. [122] Tucker MP, Kim KH, Newman MM, Nguyen QA. Effects of temperature and moisture on dilute-acid steam explosion pretreatment of corn stover and cellulase enzyme digestibility. Appl Biochem Biotechnol 2003;105:16578. [123] Jeffries TW, Jin YS. Ethanol and thermotolerance in the bioconversion of xylose by yeasts. Adv Appl Microbiol 2000;47:22168. [124] Torget R, Hatzis C, Hayward TK, Hsu TA, Philippidis GP. Optimization of reverse-ow, 2-temperature, dilute-acid pretreatment to enhance biomass conversion to ethanol. Appl Biochem Biotechnol 1996;58:85101. [125] Lee YY, Iyer P, Torget RW. Dilute-acid hydrolysis of lignocellulosic biomass. Adv Biochem Eng Biotechnol 1999;65:93115. [126] Lee JW, Gwak KS, Park JY, Park MJ, Choi DH, Kwon M, et al. Biological pretreatment of softwood Pinus densiora by three white rot fungi. J Microbiol 2007;45:48591. [127] Taherzadeh MJ, Karimi K. Acid-based hydrolysis processes for ethanol from lignocellulosic materials: a review. BioResources 2007;2:47299. [128] Gamez S, Ramirez JA, Garrote G, Vazquez M. Manufacture of fermentable sugar solutions from sugar cane bagasse hydrolyzed with phosphoric acid at atmospheric pressure. J Agric Food Chem 2004;52:41727. [129] Gullu DE. Effect of catalyst on yield of liquid products from biomass via pyrolysis. Energy Source 2003;25:75365. [130] Dunlop AP. Furfural formation and behaviour. Ind Eng Chem 1948;40:2049. [131] Ulbricht RJ, Sharon J, Thomas J. A review of 5-hydroxymethylfurfura HMF in parental solutions. Fundam Appl Toxicol 1984;4:84353. [132] Badger PC. Ethanol from cellulose: a general review. In: Janick J, Whipkey A, editors. Trends in new crops and new uses. Alexandria, VA: ASHS Press; 2002. [133] Farooqi R, Sam AG. Ethanol as a transportation fuel. Centre for applied business research in energy and the environment (CABREE) climate change initiative, University of Alberta, Canada, April 29; 2004. <www.business. ualberta.ca/cabree> [accessed December 2009]. [134] Iranmahboob J, Nadim F, Monemi S. Optimizing acid-hydrolysis: a critical step for production of ethanol from mixed wood chips. Biomass Bioenergy 2002;22:4014. [135] Demirbas A. The importance of bioethanol and biodiesel from biomass. Energy Source B 2008;3:17785. [136] Yu Y, Lou X, Wu H. Some recent advances in hydrolysis of biomass in hotcompressed water and its comparisons with other hydrolysis methods. Energy Fuels 2008;22:4660. [137] Pike PW, Sengupta D, Hertwig TA. Integrating biomass feedstocks into chemical production complexes using new and existing processes. Minerals Processing Research Institute, Louisiana State University, Baton Rouge, LA, November 3; 2008. [138] Pan X, Gilkes N, Saddler JN. Effect of acetyl groups on enzymatic hydrolysis of cellulosic substrates. Holzforschung 2006;60:398401. [139] Mani S, Tabil LG, Opoku A. Ethanol from agricultural crop residues an overview. In: ASAE/CSAE north-central intersectional meeting, Saskatoon, Saskatchewan, Canada, September 2728; 2002. [140] Zhou J, Wang YH, Chu J, Zhuang YP, Zhang SL, Yin P. Identication and purication of the main components of cellulases from a mutant strain of Trichoderma viride T 100-14. Bioresour Technol 2008;99:682633. [141] Viks-Nielsen A. Recent development in enzymes for biomass hydrolysis. In: Maj M, Kirsch N, editors. First European workshop on biotechnology for lignocellulose bioreneries. Forest & Landscape Denmark, University of Copenhagen, Copenhagen, March 2728; 2008. [142] Kim SH. Lime pretreatment and enzymatic hydrolysis of corn stover. Doctoral dissertation, Texas A&M University; May 2004. [143] Zhang YHP, Himmel ME, Mielenz JR. Outlook for cellulase improvement: screening and selection strategies. Biotechnol Adv 2006;24:45281. [144] Detroy RW, Julian GS. Biomass conversion: fermentation chemicals and fuels. Crit Rev Microbiol 1982;10:20328. [145] Gregg DJ, Saddler JN. Factors affecting cellulose hydrolysis and the potential of enzyme recycle to enhance the efciency of an integrated wood to ethanol process. Biotechnol Bioeng 1996;51:37583. [146] Chen M, Xia L, Xue P. Enzymatic hydrolysis of corncob and ethanol production from cellulosic hydrolysate. Int Biodeterior Biodegrad 2007;59:859. [147] Rezaei F, Richard TL, Logan BE. Enzymatic hydrolysis of cellulose coupled with electricity generation in a microbial fuel cell. Biotechnol Bioeng 2008;101:11639. [148] Silva CHC, Puls J, De Sousa MV, Filho EXF. Purication and characterization of a low molecular weight xylanase from solid-state cultures of Aspergillus fumigatus Fresenius. Rev Microbiol 1999;30:1149. [149] Wong KKY, Saddler JN. Trichoderma xylanases, their properties and applications. In: Visser J, Beldman G, Kusters-van Someren M, Voragen AGJ, editors. Xylans and xylanases. Amsterdam: Elsevier; 1992. p. 17186. [150] Haltrich D, Nidetzky B, Kulbe KD, Steiner W, Zupancic S. Production of fungal xylanases. Bioresour Technol 1996;58:13761. [151] Filho EXF, Tuohy MG, Puls J, Coughlan MP. The xylan-degrading enzymesystems of Penicillium capsulatum and Talaromyces emersonii. Biochem Soc Trans 1991;19:S25.

M. Balat / Energy Conversion and Management 52 (2011) 858875 [184] Bertilsson M, Olofsson K, Liden G. Prefermentation improves xylose utilization in simultaneous saccharication and co-fermentation of pretreated spruce. Biotechnol Biofuels 2009;2:8. [185] Olofsson K, Bertilsson M, Liden G. A short review on SSF an interesting process option for ethanol production from lignocellulosic feedstocks. Biotechnol Biofuels 2008;1:7. [186] Ballesteros M, Oliva JM, Negro MJ, Manzanares P, Ballesteros I. Ethanol from lignocellulosic materials by a simultaneous saccharication and fermentation process (SFS) with Kluyveromyces marxianus CECT 10875. Process Biochem 2004;39:18438. [187] Huang LP, Jin B, Lant P, Zhou J. Simultaneous saccharication and fermentation of potato starch wastewater to lactic acid by Rhizopus oryzae and Rhizopus arrhizus. Biochem Eng J 2005;23:26576. [188] Schell DJ, Ruth MF, Tucker MP. Modeling the enzymatic hydrolysis of diluteacid pretreated douglas Fir. In: 20th Symposium on biotechnology for fuels and chemicals, Gatlinburg, TN, May 37; 1998. [189] Krishna SH, Reddy TJ, Chowdary GV. Simultaneous saccharication and fermentation of lignocellulosic wastes to ethanol using a thermotolerant yeast. Bioresour Technol 2001;77:1936. [190] Madson PW, Lococo DB. Recovery of volatile products from dilute highfouling process streams. Appl Biochem Biotechnol 2000;8486: 104961. [191] Kwiatkowski JR, McAloon AJ, Taylor F, Johnston DB. Modeling the process and costs of fuel ethanol production by the corn dry-grind process. Ind Crop Prod 2006;23:28896. [192] Karuppiah R, Peschel A, Grossmann IE, Martn M, Martinson W, Zullo L. Energy optimization for the design of corn-based ethanol plants. AIChE J 2008;54:1499525. [193] McAloon A, Taylor F, Yee W, Ibsen K, Wooley R. Determining the cost of producing ethanol from corn starch and lignocellulosic feedstocks. National Renewable Energy Laboratory, Technical Report, NREL/TP-580-28893, Golden, Colorado; October 2000. [194] De Fraiture C, Giordano M, Liao Y. Biofuels and implications for agricultural water use: blue impacts of green energy. Water Policy 2008;10:6781. [195] Elam N. Alternative fuels (ethanol) in Sweden. Investigation and evaluation for IEA Bioenergy, Task 27, Atrax Energi AB, Gteborg, Sweden; 2000.

875

[196] Mathews J. A biofuels manifesto: why biofuels industry creation should be priority number one for the World Bank and for developing countries. Management Macquarie University; Sydney; 2006. <www.scidev.net/.../ a-biofuels-manifesto-why-biofuels-industry-creatio> [accessed December 2009]. [197] Shapouri H, Salassi M, Nelson J. The economic feasibility of ethanol production from sugar in the United States. US Department of Agriculture (USDA); July 2006. <www.usda.gov/oce> [accessed December 2009]. [198] Kojima M, Johnson T. Potential for biofuels for transport in developing countries. Energy sector management assistance programme, Joint UNDP/ World Bank, Washington, DC; October 2005. [199] Mitchell D. A note on rising food prices. World bank development prospects group. World Bank, Washington, DC, April 8; 2008. [200] Christensen K, Smith A. The case for hemp as a biofuel. Vote Hemp Inc. Report, Brattleboro, VT; 2008. [201] Dawson B, Spannagle M. The complete guide to climate change. New York (NY): Matt. Taylor & Francis Routledge; 2009. [202] Endo A, Nakamura T, Ando A, Tokuyasu K, Shima J. Genome-wide screening of the genes required for tolerance to vanillin, which is a potential inhibitor of bioethanol fermentation, in Saccharomyces cerevisiae. Biotechnol Biofuels 2008;1:16. [203] Pimentel D, Marklein A, Toth MA, Karpoff M, Paul GS, McCormack R, et al. Food versus biofuels: environmental and economic costs. Human Ecol 2009;37:112. [204] Tiffany DG. Economic and environmental impacts of US corn ethanol production and use, vol. 5. Federal Reserve Bank of St. Louis Regional Economic Development; 2009. p. 4258. [205] Yacobucci BD, Schnepf R. Ethanol and biofuels: agriculture, infrastructure, and market constraints related to expanded production. CRS report for congress, Order Code RL33928, March 16; 2007. [206] Eggeman T, Elander RT. Process and economic analysis of pretreatment technologies. Bioresour Technol 2005;96:201925. [207] Woodson M, Jablonowski CJ. An economic assessment of traditional and cellulosic ethanol technologies. Energy Source B 2008;3:37283.

You might also like