1-s2.0-S0196890410003791-main

18
Production of bioethanol from lignocellulosic materials via the biochemical pathway: A review Mustafa Balat Sila Science & Energy Company, University Mah, Trabzon, Turkey article info Article history: Received 1 January 2010 Accepted 15 August 2010 Available online 6 September 2010 Keywords: Bioethanol Lignocellulosic material Pretreatment Enzymatic hydrolysis Fermentation abstract 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. Bioeth- anol can be produced from different kinds of raw materials. These raw materials are classified 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. Lignocellu- losic 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. Ó 2010 Elsevier Ltd. All rights reserved. 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) emis- sions and 19% of global carbon dioxide (CO 2 ) emissions [2]. CO 2 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 finite nature of fossil fuels, increasing concerns regarding environmental impact, especially related to greenhouse gas (GHG) emissions, and health and safety considerations are forc- ing the search for new energy sources and alternative ways to power the world’s motor vehicles. An alternative fuel must be technically feasible, economically competitive, environmentally acceptable, and readily available [6]. Numerous potential alterna- tive fuels have been proposed, including bioethanol, biodiesel, methanol, hydrogen, boron, natural gas, liquefied petroleum gas (LPG), Fischer–Tropsch 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 CO 2 -cycle in combustion, (3) biofuels have a considerable environ- mentally friendly potential, (4) there are many benefits the envi- ronment, economy and consumers in using biofuels, and (5) they are biodegradable and contribute to sustainability. The major ben- efits of biofuels are given in Table 1. The use of biofuels can con- tribute 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 coun- tries 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 po- tential for developing countries to benefit from the increased glo- bal 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 effi- cient 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 thermo- chemical processes such as pyrolysis [10,11], gasification [12,13], liquefaction [14], supercritical fluid 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 gasification, 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, Fischer–Tropsch liquids, and biohydrogen. The term biofuels can refer to fuels for direct combustion for electricity production, but 0196-8904/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.enconman.2010.08.013 Tel.: +90 462 871 3025; fax: +90 462 871 3110. E-mail address: [email protected] Energy Conversion and Management 52 (2011) 858–875 Contents lists available at ScienceDirect Energy Conversion and Management journal homepage: www.elsevier.com/locate/enconman

description

fuels

Transcript of 1-s2.0-S0196890410003791-main

Page 1: 1-s2.0-S0196890410003791-main

Energy Conversion and Management 52 (2011) 858–875

Contents lists available at ScienceDirect

Energy Conversion and Management

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

Production of bioethanol from lignocellulosic materials via the biochemicalpathway: 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

Article history:Received 1 January 2010Accepted 15 August 2010Available online 6 September 2010

Keywords:BioethanolLignocellulosic materialPretreatmentEnzymatic hydrolysisFermentation

0196-8904/$ - see front matter � 2010 Elsevier Ltd. Adoi:10.1016/j.enconman.2010.08.013

⇑ Tel.: +90 462 871 3025; fax: +90 462 871 3110.E-mail address: [email protected]

Bioethanol is by far the most widely used biofuel for transportation worldwide. Production of bioethanolfrom biomass is one way to reduce both consumption of crude oil and environmental pollution. Bioeth-anol can be produced from different kinds of raw materials. These raw materials are classified into threecategories of agricultural raw materials: simple sugars, starch and lignocellulose. The price of the rawmaterials is highly volatile, which can highly affect the production costs of the bioethanol. One majorproblem with bioethanol production is the availability of raw materials for the production. Lignocellu-losic biomass is the most promising feedstock considering its great availability and low cost, but thelarge-scale commercial production of fuel bioethanol from lignocellulosic materials has still not beenimplemented.

� 2010 Elsevier Ltd. All rights reserved.

1. Introduction

Today, the transportation sector worldwide is almost entirelydependent on petroleum-based fuels. It is responsible for 60% ofthe world oil consumption [1]. In addition, transportation sectoraccounts for more than 70% of global carbon monoxide (CO) emis-sions and 19% of global carbon dioxide (CO2) emissions [2]. CO2

emissions from a gallon of gasoline are about 8 kg [3]. Aroundthe world, there were about 806 million cars and light trucks onthe road in 2007 [4]. These numbers are projected to increase to1.3 billion by 2030 and to over 2 billion vehicles by 2050 [5]. Thisgrowth will affect the stability of ecosystems and global climate aswell as global oil reserves. The dramatic increase in the price ofpetroleum, the finite nature of fossil fuels, increasing concernsregarding environmental impact, especially related to greenhousegas (GHG) emissions, and health and safety considerations are forc-ing the search for new energy sources and alternative ways topower the world’s motor vehicles. An alternative fuel must betechnically feasible, economically competitive, environmentallyacceptable, and readily available [6]. Numerous potential alterna-tive fuels have been proposed, including bioethanol, biodiesel,methanol, hydrogen, boron, natural gas, liquefied petroleum gas(LPG), Fischer–Tropsch fuel, p-series, electricity, and solar fuels.

Biomass-based fuels, also known as biofuels offer manyadvantages over petroleum-based fuels [7]: (1) biofuels are easilyavailable from common biomass sources, (2) they are represent aCO2-cycle in combustion, (3) biofuels have a considerable environ-

ll rights reserved.

mentally friendly potential, (4) there are many benefits the envi-ronment, economy and consumers in using biofuels, and (5) theyare biodegradable and contribute to sustainability. The major ben-efits of biofuels are given in Table 1. The use of biofuels can con-tribute to the mitigation of GHG emissions, provide a clean andtherefore sustainable energy source, and increase the agriculturalincome for rural poor in developing countries. Developing coun-tries have a comparative advantage for biofuel production becauseof greater availability of land, favorable climatic conditions foragriculture and lower labour costs. However, there may be othersocio-economic and environmental implications affecting the po-tential for developing countries to benefit from the increased glo-bal demand for biofuel [8]. Large-scale production of biofuelsoffers an opportunity for certain developing countries to reducetheir dependence on oil imports. In developed countries there isa growing trend towards employing modern technologies and effi-cient bioenergy conversion using a range of biofuels, which arebecoming cost-wise competitive with fossil fuels [9].

Biofuels are made from bio-based materials through thermo-chemical processes such as pyrolysis [10,11], gasification [12,13],liquefaction [14], supercritical fluid extraction [15], supercriticalwater liquefaction [16] and biochemical [17]. Thermo-chemicalreforming of biomass concerns the processes of catalytic andnon-catalytic pyrolysis as well as the gasification, which aims atthe maximization of the production of energetically exploitableliquid and gaseous products.

Biofuels include bioethanol, biomethanol, vegetable oils,biodiesel, biogas, biosynthetic gas (bio-syngas), bio-oil, bio-char,Fischer–Tropsch liquids, and biohydrogen. The term biofuels canrefer to fuels for direct combustion for electricity production, but

Page 2: 1-s2.0-S0196890410003791-main

Table 1Major benefits of biofuels.

Economic impacts SustainabilityFuel diversityIncreased number of rural manufacturing jobsIncreased income taxesIncreased investments in plant and equipmentAgricultural developmentInternational competitivenessReducing the dependency on imported petroleum

Environmental impacts Greenhouse gas reductionsReducing of air pollutionBiodegradabilityHigher combustion efficiencyImproved land and water useCarbon sequestration

Energy security Domestic targetsSupply reliabilityReducing use of fossil fuelsReady availabilityDomestic distributionRenewability

Table 2Physical and chemical properties of ethanol, methanol and gasoline.

Property MethanolCH3OH

EthanolC2H5OH

GasolineC4-C12

Molecular weight (g/mol) 32 46 �114Specific gravity 0.789 (298 K) 0.788 (298 K) 0.739

(288.5 K)Vapor density rel. to air 1.10 1.59 3.0–4.0Liquid density (g cm�3

at 298 K)0.79 0.79 0.74

Boiling point (K) 338 351 300–518Melting point (K) 175 129 –Heat of evaporation (Btu/lb) 472 410 135

Heating value (kBTU gal�1)Lower 58 74 111Upper 65 85 122

Tank design pressure (psig) 15 15 15Viscosity (cp) 0.54 1.20 0.56Flash point (K) 284 287 228

Flammability/explosionlimits

(%) Lower (LFL) 6.7 3.3 1.3(%) Upper (UFL) 36 19 7.6

Auto ignitiontemperature (K)

733 636 523–733

M. Balat / Energy Conversion and Management 52 (2011) 858–875 859

is generally used for liquid fuels for transportation sector [18].Renewable liquid biofuels for transportation have recently at-tracted huge attention in different countries all over the world be-cause of its renewability, sustainability, common availability,regional development, rural manufacturing jobs, reduction ofGHG emissions, and its biodegradability [19].

Bioethanol is by far the most widely used biofuel for transpor-tation worldwide. Bioethanol and bioethanol/gasoline blends havea long history as alternative transportation fuels. It has been usedin Germany and France as early as 1894 by the then incipientindustry of internal combustion engines (ICEs) [20]. Brazil has uti-lized bioethanol as a fuel since 1925. By that time, the productionof bioethanol was 70 times bigger than the production and con-sumption of petrol [21]. The use of bioethanol for fuel was wide-spread in Europe and the United States until the early 1900s.Because it became more expensive to produce than petroleum-based fuel, especially after World War II, bioethanol’s potentialwas largely ignored until the oil crisis of the 1970s [22]. Sincethe 1980s, there has been an increased interest in the use of bio-ethanol as an alternative transportation fuel.

To ensure that ‘‘good” bioethanol is produced, with reference toGHG benefits, the following demands must be met [23]: (1) bioeth-anol plants should use biomass and not fossil fuels, (2) cultivationof annual feedstock crops should be avoided on land rich in carbon(above and below ground), such as peat soils used as permanentgrassland, (3) by-products should be utilized efficiently in orderto maximize their energy and GHG benefits, and (4) nitrous oxideemissions should be kept to a minimum by means of efficient fer-tilization strategies, and the commercial nitrogen fertilizer utilizedshould be produced in plants which have nitrous oxide gas clean-ing. Bioethanol is a fuel derived from renewable sources of feed-stock; typically plants such as wheat, sugar beet, corn, straw, andwood. Bioethanol is an alternative fuel that is produced almost en-tirely from food crops. It represents an important, renewable liquidfuel for motor vehicles. Producing bioethanol as a transportationfuel can help reduce CO2 buildup in two important ways: by dis-placing the use of fossil fuels, and by recycling the CO2 that is re-leased when it is combusted as fuel. An important advantage ofcrop-based bioethanol is its GHG benefits [24].

Solubility in H2O (%) Miscib.(100%)

Miscib.(100%)

Negl. (�0.01)

Azeotrope with H2O None 95% EtOH ImmisciblePeak flame temperature (K) 2143 2193 2303Minimum ignition energy

in air (mJ)0.14 0.23 –

2. Bioethanol as a transportation fuel

The alcohols are oxygenates fuels that the alcohol molecule hasone or more oxygen, which decreases to the combustion heat.

Practically, any of the organic molecules of the alcohol family canbe used as a fuel. The alcohols can be used for motor fuels aremethanol (CH3OH), bioethanol (C2H5OH), propanol (C3H7OH),butanol (C4H9OH). However, only methanol and bioethanol fuelsare technically and economically suitable for internal combustionengines (ICEs) [24].

Bioethanol is ethyl alcohol, grain alcohol, or chemically C2H5OHor EtOH. It has high octane number (108) [25], both permit the ris-ing of the compression ratio and gives lower emission [26]. Octanenumber is a measure of the gasoline quality for prevention of earlyignition, which leads to cylinder knocking. The fuels with higheroctane numbers are preferred in spark-ignition ICEs. An oxygenatefuel such as bioethanol is provides a reasonable antiknock value[3]. Disadvantages of bioethanol include its lower energy densitythan gasoline (but about 35% higher than that of methanol), its cor-rosiveness, low flame luminosity, lower vapor pressure (makingcold starts difficult), miscibility with water, and toxicity to ecosys-tems [27], increase in exhaust emissions of acetaldehyde, and in-crease in vapor pressure (and evaporative emissions) whenblending with gasoline. Physical and chemical properties of bioeth-anol, methanol and gasoline are given in Table 2 [28].

Bioethanol has been used as a modern biofuel, applied directlyas a gasoline improver or gasoline subsistent, or in the form ofETBE (ethyl tertiary butyl ether) for currently added syntheti-cally-produced octane enhancers and in bioethanol–diesel blendswith particular purpose to reduce the emissions of exhaust gasses[29]. Bioethanol is most commonly blended with gasoline in con-centrations of 10% bioethanol to 90% gasoline, known as E10 andnicknamed ‘‘gasohol”. Bioethanol can be used as a 5% blend withpetrol under the European Union (EU) quality standard EN 228.This blend requires no engine modification and is covered by vehi-cle warranties. With engine modification, bioethanol can be usedat higher levels, for example, E85 [30]. Some countries has exer-cised biofuel program both form bioethanol–gasoline blendprogram such as the United States (E10 and for flexible-fuelvehicle-FFV E85), Canada (E10 and for FFV E85), Sweden (E5 andfor FFV E85), India (E5), Australia (E10), Thailand (E10), China

Page 3: 1-s2.0-S0196890410003791-main

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

Table 3Bioethanol yields from different energy crops.

Country Energy crop Bioethanol yield (l/ha)

Brazil Sugarcane, 100% 6641USA Corn, 98% 3770

Sweet sorghum, 2% 1365

China Corn, 70% 2011Wheat, 30% 1730

EU-27 Wheat, 48% 1702Sugar beet, 29% 5145

Canada Corn, 70% 3460Wheat, 30% 1075

860 M. Balat / Energy Conversion and Management 52 (2011) 858–875

(E10), Columbia (E10), Peru (E10), Paraguay (E7), Brazil (E20, E25and FFV any blend).

The reduced CO2 emissions mean that bioethanol is good for theenvironment. Using bioethanol-blended fuel for automobiles cansignificantly reduce petroleum use and exhaust GHG emission[24]. On a life-cycle basis, not all biofuels are equal in terms ofenvironmental benefits. Fig. 1 demonstrates the lower GHG emis-sions resulting from the use of biofuels compared to gasoline ona life-cycle basis. As Fig. 1 illustrates, corn-based bioethanol offersrather limited benefits, as it reduces GHG emissions by only 18%compared to gasoline. In contrast, sugarcane and cellulosic bioeth-anol result in almost 90% lower emissions [31].

The net energy balance of biomass to bioethanol conversion isthe key parameter that explains the interest in using bioethanolfuel instead of fossil gasoline. From a life-cycle assessment (LCA)viewpoint, the ratio of the energy content of bioethanol to thenet non-renewable primary energy (allocated to bioethanol) con-sumed in the whole production process from biomass productionto 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 consumedin making it. On the other hand, sugarcane and cellulosic bioetha-nol yield renewable energy nine times worth the fossil energy usedto produce them [31].

3. An overview of bioethanol feedstocks

Bioethanol can be produced from different kinds of raw materi-als. The raw materials are classified into three categories of agricul-tural 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). Cur-rently, a focus is on bioethanol production from crops, such as corn,wheat, sugar cane, as well as on highly abundant agriculturalwastes.

One major problem with bioethanol production is the availabil-ity of raw materials for the production. The availability of feedstockfor bioethanol can vary considerably from season to season and de-pends on geographic locations. Locally available agricultural bio-mass will be used for the bioethanol production [33]. For a givenproduction line, the comparison of the feedstocks includes severalissues [34]: (1) chemical composition of the biomass, (2) cultiva-tion practices, (3) availability of land and land use practices, (4)use of resources, (5) energy balance, (6) emission of greenhousegases, acidifying gases and ozone depletion gases, (7) absorptionof minerals to water and soil, (8) injection of pesticides, (9) soilerosion, (10) contribution to biodiversity and landscape valuelosses, (11) farm-gate price of the biomass, (12) logistic cost (trans-

port and storage of the biomass), (13) direct economic value of thefeedstocks taking into account the co-products, (14) creation ormaintain of employment, and (15) water requirements and wateravailability.

Brazil utilizes sugarcane for bioethanol production while theUnited States and Europe mainly use starch from corn, and fromwheat and barley, respectively. Sugarcane as a biofuel crop hasmuch expanded in the last decade, yielding anhydrous bioethanol(gasoline additive) and hydrated bioethanol by fermentation anddistillation of sugarcane juice and molasses [35]. Brazil’s sugarcaneyield averages about 82.4 tons/ha [36]. The yield of bioethanol perhectare, currently at around 6650 l/ha (Table 3) [37]. Brazil is thelargest single producer of sugarcane with about 31% of global pro-duction [35]. It has nearly 9 million hectares of sugarcane undercultivation. Sugar beet crops are grown in most of the EU-25 coun-tries, and yield substantially more bioethanol per hectare thanwheat.

The United States is predominantly a producer of bioethanol de-rived from corn, and production is concentrated in Midwesternstates with abundant corn supplies [38]. Feedstock availability isnot expected to be a constraint for bioethanol production overthe next decade. Corn is expected to remain the predominant feed-stock in the United States, although its share likely will declinemodestly by 2015. Corn-based bioethanol production in most ofthe countries assessed is limited, especially compared to the Uni-ted States. Only Canada reported explicit plans for significant fu-ture development of corn-based bioethanol, although China hasused corn as a feedstock in the past and Argentina is looking atthe 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 canhighly affect the production costs of the bioethanol [40]. Lignocel-lulosic materials serve as a cheap and abundant feedstock, which isrequired to produce fuel bioethanol from renewable resources atreasonable costs. In 2007 the US Department of Energy providedmore than US$1 billion toward lingocellulosic bioethanol projects,with the goal of making the fuel cost competitive at US$1.33 pergallon by 2012 [41]. The level of support provided by the EU isfar less, but is still significant (approximately US$68 million in2006) [41].

Lignocellulosic materials can be classified in four groups basedon type of resource: (1) forest residues, (2) municipal solid waste,(3) waste paper, and (4) crop residue resources. Literature reportsseveral papers on utilization of various lignocellulosic waste mate-rials such as rice straw [42], corn stover [43], switchgrass [44],palm bagasse [45], etc.

Lignocellulosic materials could produce up to 442 billion litersper year of bioethanol [46]. Rice straw is one of the abundant

Page 4: 1-s2.0-S0196890410003791-main

M. Balat / Energy Conversion and Management 52 (2011) 858–875 861

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

4.2. Chemical structure and basic components of lignocellulosicmaterials

Chemical composition of lignocellulosic materials is a key factoraffecting efficiency of biofuel production during conversion pro-cesses. The structural and chemical composition of lignocellulosicmaterials is highly variable because of genetic and environmentalinfluences and their interactions [48]. A typical chemical composi-tion of lignocellulosic materials is 48 wt.% C, 6 wt.% H, and 45 wt.%O, the inorganic matter being a minor component [49]. The proxi-mate analysis of rice straw and wheat straw shows components asfollow: volatile matter (65.47%, 75.27%), fixed carbon (15.86%,17.71%) and ash (18.67%, 7.02%), respectively [50].

Lignocelluloses consist mainly of cellulose, hemicellulose andlignin; these components build up about 90% of dry matter inlignocelluloses, with the rest consisting of e.g. extractive and ash[51]. The basic structure of all woody biomass consists of three ba-sic polymers: cellulose (C6H10O5)x, hemicelluloses such as xylan(C5H8O4)m, and lignin [C9H10O3(OCH3)0.9–1.7]n in trunk, foliage,and bark. The proportion of these wood constituents varies be-tween species, and there are distinct differences between hard-woods and softwoods. Cellulose + hemicellulose contents aremore in hardwoods (78.8%) than softwoods (70.3%), but lignin ismore in softwoods (29.2%) than hardwoods (21.7%) [52]. The struc-tural composition of various types of lignocellulosic-biomassmaterials are given in Table 4 [53].

Cellulose and hemicellulose, which typically make up two-thirds of cell wall dry matter, are polysaccharides that can behydrolyzed to sugars and then fermented to bioethanol. Processperformance, i.e. Bioethanol yield from biomass, is directly relatedto cellulose, hemicellulose, and individual sugar concentration inthe feedstock [54]. The lignin cannot be used for bioethanolproduction.

Cellulose, the major component of plant biomass (30–60% of totalfeedstock dry matter), is a linear polymer of glucose; the orientationof the linkages and additional hydrogen bonding make the polymerrigid and difficult to break. In hydrolysis the polysaccharide isbroken down to free sugar molecules by the addition of water [55].

Table 4Composition of various types of lignocellulosic-biomass materials (% dry weight).

Material Cellulose Hemicelluloses Lignin Ash Extractives

Algae (green) 20–40 20–50 – – –Cotton, flax,

etc.80–95 5–20 – – –

Grasses 25–40 25–50 10–30 – –Hardwoods 45 ± 2 30 ± 5 20 ± 4 0.6 ± 0.2 5 ± 3Hardwood

barks22–40 20–38 30–55 0.8 ± 0.2 6 ± 2

Softwoods 42 ± 2 27 ± 2 28 ± 3 0.5 ± 0.1 3 ± 2Softwood

barks18–38 15–33 30–60 0.8 ± 0.2 4 ± 2

Cornstalks 39–47 26–31 3–5 12–16 1–3Wheat straw 37–41 27–32 13–15 11–14 7 ± 2Newspapers 40–55 25–40 18–30 – –Chemical

pulps60–80 20–30 2–10 – –

This process is also known as saccharification. The product, glucose,is a six-carbon sugar.

Hemicellulose (20–40% of total feedstock dry matter) is a short,highly branched polymer of five-carbon (pentoses) and six-carbon(hexoses) sugars. Specifically, hemicellulose contains xylose andarabinose (five-carbon sugars) and galactose, glucose, and man-nose (six-carbon sugars). Hemicellulose is more readily hydrolyzedcompared to cellulose because of its branched, amorphous nature[48]. The dominant sugars in hemicelluloses are mannose in soft-woods and xylose in hardwoods and agriculture residues [56].

Lignin (15–25% of total feedstock dry matter) is an aromaticpolymer synthesised from phenylpropanoid precursors. The basicchemical phenylpropane units of lignin (primarily syringyl, guaia-cyl and p-hydroxy phenol) are bonded together by a set of linkagesto form a very complex matrix [57]. This matrix comprises a vari-ety 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 first and second cat-egory, respectively. Softwoods generally contain more lignin thanhardwoods [59]. Lignin contents on a dry basis in both softwoodsand hardwoods generally range from 20% to 40% by weight andfrom 10% to 40% by weight in various herbaceous species, suchas bagasse, corncobs, peanut shells, rice hulls and straws [60]. Lig-nin is one of the drawbacks of using lignocellulosic-biomass mate-rials in fermentation, as it makes lignocellulose resistant tochemical and biological degradation [56].

5. Bioethanol from lignocellulosic materials via the biochemicalpathway

Biochemical conversion of lignocellulosic materials throughsaccharification and fermentation is a major pathway for bioetha-nol production from biomass. Bioconversion of lignocellosics tobioethanol is difficult due to: (1) the resistant nature of biomassto breakdown; (2) the variety of sugars which are released whenthe hemicellulose and cellulose polymers are broken and the needto find or genetically engineer organisms to efficiently fermentthese sugars; (3) costs for collection and storage of low densitylignocellosic materials. Generic block diagram of bioethanol pro-duction from lignocellulose materials is given in Fig. 2 [61]. The ba-sic process steps in producing bioethanol from lignocellulosicmaterials are: pretreatment, hydrolysis, fermentation and productseparation/distillation.

5.1. Pretreatment of lignocellulosic materials

The recalcitrance of lignocellulose is one of the major barriers tothe economical production of bioethanol. The technical approachto overcome recalcitrance has been pretreatment of biomass feed-stock to remove the barriers and make cellulose more accessible tohydrolytic enzymes for conversion to glucose [62]. The goals ofpretreatment on lignocellulosic material are depicted in Fig. 3[63]. If the pretreatment is not efficient enough the resultant resi-due is not easily hydrolyzable by cellulase enzyme and if it is moresevere, result is the production of toxic compounds which inhibitthe microbial metabolism [64].

Pretreatment has been viewed as one of the most expensiveprocessing steps within the conversion of biomass to fermentablesugar [65]. There is huge scope in lowering the cost of pretreat-ment process through extensive R&D approaches. Pretreatmentof cellulosic biomass in cost effective manner is a major challengeof cellulose to bioethanol technology research and development[66]. Taherzadeh and Karimi [56] has summarized the prerequi-sites for an ideal lignocellulose pretreatment; it should: (1) pro-duction of reactive cellulosic fiber for enzymatic attack, (2)

Page 5: 1-s2.0-S0196890410003791-main

Fig. 2. Generic block diagram of bioethanol production from lignocellulosebiomass. Possibilities for reaction–reaction integration are shown inside the shadedboxes: SSF – simultaneous saccharification and fermentation; SSFC – simultaneoussaccharification 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].

862 M. Balat / Energy Conversion and Management 52 (2011) 858–875

avoiding destruction of hemicelluloses and cellulose, (3) avoidingformation of possible inhibitors for hydrolytic enzymes and fer-menting microorganisms, (4) minimizing the energy demand, (e)reducing the cost of size reduction for feedstocks, (5) reducingthe cost of material for construction of pretreatment reactors, (6)producing less residues, and (7) consumption of little or no chem-ical and using a cheap chemical. Pretreatment is crucial for ensur-ing 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 explo-sion/autohydrolysis, hydrothermolysis, and wet oxidation), chem-ical (alkali, dilute acid, oxidizing agents, and organic solvents),and biological processes have been used for pretreatment of ligno-cellulosic materials. However, not all of these methods have yetdeveloped enough to be feasible technically or economically forlarge-scale processes. For example, milling could be applied to cre-ate a better steam explosion by reducing the chip size [67]. Advan-tages and disadvantages of various pretreatment processes forlignocellulosic materials are summarized in Table 5 [68]. Many

researchers have investigated the effect of different pretreatmentmethods upon various lignocellulosic materials such as corn stover[69], wheat straw [70], switchgrass [71], rice straw [72], and sug-arcane bagasse [73].

5.1.1. Physical pretreatment5.1.1.1. Mechanical comminution. Lignocellulosic materials can becomminuted by a combination of chipping, grinding, and millingto reduce cellulose crystallinity. The size of the materials is usually10–30 mm after chipping and 0.2–2 mm after milling or grinding[68,74,75]. Vibratory ball milling was found to be more effectivethan ordinary ball milling in reducing cellulose crystallinity ofspruce and aspen chips and in improving their digestibility[68,76]. Power requirements of mechanical comminution dependon the final 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 mechanicalcomminution are regarded as high for hardwood, which consumes130 kW h/ton to reduce the particle size to 1.6 mm. To reduce thesize of corn stover with mechanical comminution to 1.6 mm re-quires far less energy, consuming only 14 kW h/ton [79]. Thesemechanical pretreatment techniques are time-consuming, energyintensive, or expensive to process. The compression milling isapparently the only comminution process that has been testedusing a production-scale apparatus [80].

5.1.1.2. Pyrolysis. Pyrolysis has also been used for pretreating ligno-cellulosic materials, since biomass can be used as substrate for afast pyrolysis for thermal conversion of cellulose and hemicelluloseinto fermentable sugars with good yields [81]. When the groundcellulosic materials are treated at temperatures greater than573 K, cellulose rapidly decomposes to produce gaseous productsand residual char [74,82]. Pyrolysis pretreatment prior to enzy-matic hydrolysis of three waste cellulosic materials (office paper,newspaper and cardboard) was examined by Leustean [75].Table 6 shows after treatment and enzymatic hydrolysis the reduc-ing sugar concentration. The pyrolysis pretreatment of groundmaterial is improved the conversion of cellulose to glucose yieldfrom enzymatic hydrolysis [75].

5.1.2. Physico-chemical pretreatment5.1.2.1. Steam explosion (autohydrolysis). Steam explosion is themost commonly used method for the pretreatment of lignocellu-losic materials [78]. To summarize the effects of steam explosiontreatment on lignocellulosics reported in the literature [83]: (1)steam explosion treatment increases crystallinity of cellulose bypromoting crystallization of the amorphous portions; (2) hemicel-lulose is easily hydrolyzed by steam explosion treatment; (3) thereis evidence that steam explosion promotes delignification. In thismethod, chipped biomass is treated with high-pressure saturatedsteam and then the pressure is swiftly reduced, which makes thematerials undergo an explosive decomposition.

Steam explosion, compared to other pretreatment methods, of-fers potential for lower capital investment, significantly lowerenvironmental impact, more potential for energy efficiency, lesshazardous process chemicals and conditions and complete sugarrecovery [81]. The conventional mechanical methods require 70%more energy than steam explosion to achieve the same size reduc-tion [68,74,84]. Steam explosion is considered the most cost effec-tive option for hardwood and agriculture residues, but is lesseffective for softwood [84]. Most important factors affecting effec-tiveness of steam explosion are particle size, temperature and res-idence time and the combined effect of both temperature and timeis described by severity factor (Ro), which is optimal for maximumsugar yield between 3.0 and 4.5 [81]. Steam explosion is initiatedat a temperature of 433–533 K with a corresponding pressure

Page 6: 1-s2.0-S0196890410003791-main

Table 5Advantages and disadvantages of various pretreatment processes for lignocellulosic materials [68].

Pretreatment process Advantages Limitations and disadvantages

Mechanical comminution Reduces cellulose crystallinity Power consumption usually higher than inherent biomass energySteam explosion Causes hemicellulose degradation

and lignin transformation; cost-effectiveDestruction of a portion of the xylan fraction;incomplete disruption of the lignin-carbohydrate matrix;generation of compounds inhibitory to microorganisms

AFEX Increases accessible surface area,removes lignin and hemicellulose to an extent;does not produce inhibitors for down-stream processes

Not efficient for biomass with high lignin content

CO2 explosion Increases accessible surface area; cost-effective;does not cause formation of inhibitory compounds

Does not modify lignin or hemicelluloses

Ozonolysis Reduces lignin content; does not produce toxic residues Large amount of ozone required; expensiveAcid hydrolysis Hydrolyzes hemicellulose to xylose

and other sugars; alters lignin structureHigh cost; equipment corrosion; formation of toxic substances

Alkaline hydrolysis Removes hemicelluloses and lignin;increases accessible surface area

Long residence times required; irrecoverablesalts formed and incorporated into biomass

Organosolv Hydrolyzes lignin and hemicelluloses Solvents need to be drained from the reactor,evaporated, condensed, and recycled; high cost

Pyrolysis Produces gas and liquid products High temperature; ash productionPulsed electrical field Ambient conditions; disrupts plant cells; Process needs more researchBiological Simple equipment degrades lignin

and hemicelluloses; low energy requirementsRate of hydrolysis is very low

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

M. Balat / Energy Conversion and Management 52 (2011) 858–875 863

0.69–4.83 MPa for several seconds to a several minutes before thematerial is exposed to atmospheric pressure for cooling [85].

Uncatalyzed steam explosion refers to a pretreatment tech-nique in which lignocellulosic biomass is rapidly heated by high-pressure steam without addition of any chemicals. The biomass/steam mixture is held for a period of time to promote hemicellu-lose hydrolysis, and terminated by an explosive decompression[86]. Negro et al. [87] studied steam explosion and liquid hot watermethods for pretreatment of poplar (Populus nigra) biomass. Thebest results were obtained in steam explosion pretreatment at483 K and 4 min, taking into account cellulose recovery above95%, enzymatic hydrolysis yield of about 60%, and 41% xyloserecovery in the liquid fraction. Addition of H2SO4 (or SO2) or CO2

[typically 0.3–3% (w/w)] in steam explosion can decrease timeand temperature, effectively improve hydrolysis, decrease the pro-duction of inhibitory compounds, and lead to complete removal ofhemicellulose [68]. H2SO4 is a strong catalyst that highly improvesthe hemicellulose removal but also easily yields inhibitory sub-

Table 6Pyrolysed cellulosic materials [75].

Office paper Newspaper Cardboard

Reducing sugar (mg/ml) 0.14 0.11 0.11

stances [88]. Ballesteros et al. [89] applied acid-catalyzed steamexplosion pretreatment of wheat straw for bioethanol productionby varying the temperature (433–473 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 conditionsto obtain high conversion yield to bioethanol (approx 80% of theo-retical) of cellulose-rich residue after steam explosion are 463 Kand 10 min or 473 K and 5 min, in acid-impregnated straw. Usinga H2SO4-catalyzed steam explosion process for pretreatment ofSalix chips, at 473 K for either 4 or 8 min using 0.5% sulfuric acid,resulted in glucose recovery about 92% and 86% xylose recoveryafter enzymatic hydrolysis [90]. SO2 appears more appealing thanH2SO4 in steam explosion since the former requires milder andmuch less expensive reactor material, generates less gypsum,yields more xylose, and produces more digestible substrate withhigh fermentability [76]. The treatment can be carried out by 1–4% SO2 (w/w substrate) at elevated temperatures, e.g. 433–503 K,for a period of e.g. 10 min [56]. The main drawback of SO2 is itshigh toxicity, which may pose safety and health risks. However,SO2 is used in various industrial processes using established tech-niques [91].

Two-step pretreatment has been suggested in several studies asa means of increasing the sugar recovery [92,93]. In the first step,steam is performed using low temperature to solubilize hemicellu-

Page 7: 1-s2.0-S0196890410003791-main

864 M. Balat / Energy Conversion and Management 52 (2011) 858–875

losic fraction, and cellulose fraction is subjected to a second steamexplosion pretreatment step at a temperature higher than 483 K. Itoffers some additional advantages (higher bioethanol yields, betteruse of raw material and lower enzyme dosages during steamexplosion) [81].

5.1.2.2. Ammonia fiber explosion. Ammonia fiber explosion (AFEX) isone of the alkaline physico-chemical pretreatment processes. Inthis process, the material is subjected to liquid ammonia at hightemperature and pressure, and a subsequent fast decompression,similar to the steam explosion, which causes a fast saccharificationof the lignocellulosic material [94]. In a typical AFEX process, thedosage of liquid ammonia is 1–2 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 ammonialoading, temperature, water loading, blowdown pressure, time,and number of treatments [56]. This system does not directly lib-erate any sugars, but allows the polymers (hemicellulose and cel-lulose) to be attacked enzymatically and reduced to sugars [95].AFEX pretreatment yields optimal hydrolysis rates for pretreatedlignocellulosics with close to theoretical yields at low enzymeloadings (<5 FPU/g of biomass or 20 FPU/g cellulose) [86].

AFEX pretreatment has been demonstrated to markedly im-prove the saccharification rates of numerous herbaceous cropsand grasses [78]. It has been applied to various lignocellulosicmaterials, including rice straw, municipal solid waste, newspaper,sugar beet pulp, sugarcane bagasse, corn stover, switchgrass,miscanthus, apsen chips, etc. [76]. The optimal conditions for pre-treatment of switchgrass with AFEX were reported by Alizadeh etal. [96]. The optimal pretreatment conditions were found to benear 373 K reactor temperature, and ammonia loading of 1 kg ofammonia per kilogram of dry matter with 80% moisture content(dry weight basis) at 5 min residence time. These conditionsyielded a sixfold improvement in hydrolysis efficiency. Teymouriet al. [97] evaluated the optimum process conditions for the pre-treatment of corn stover. The optimal pretreatment conditionswere found to be a temperature of 373 K, an ammonia loading of1 kg of ammonia per kilogram of dry matter, a moisture contentof 60% (dry weight basis), and a residence time of 5 min. Approxi-mately 98% of the theoretical glucose yield was obtained duringenzymatic hydrolysis of the optimal treated corn stover using 60filter paper units (FPU) of cellulase enzyme/g of glucan (equal to22 FPU/g of dry corn stover). The bioethanol yield from this samplewas increased up to 2.2 times over that of untreated sample. Thecomposition of the materials after AFEX pretreatment was essen-tially the same as the original materials [74]. Holtzapple et al.[98] obtained over 90% of hydrolysis of cellulose and hemicelluloseafter AFEX pretreatment of Bermuda grass (approximately 5% lig-nin) and bagases (15% lignin). AFEX works only moderately andis not attractive for the biomass with high lignin content. Lignincontent in grasses (15–20%) is relatively low when compared withhardwoods and softwood (20–35%) [98]. This could be one of themajor reasons that grasses can be more easily digested comparedto AFEX treated hardwoods following AFEX pretreatment [99].Ammonia must be recycled after the pretreatment to reduce thecost and protect the environment [56]. A possible approach is to re-cover the ammonia after the pretreatment by evaporation [100].

5.1.2.3. Liquid hot-water pretreatment. Cooking of lignocellulosicmaterials in liquid hot water (LHW) is one of the hydrothermalpretreatment methods applied for pretreatment of lignocellulosicmaterials since several decades ago in e.g. pulp industries [56].LHW subjects biomass to hot water in liquid state at high pressureduring a fixed period and it presents elevated recovery rates forpentoses and generates low amount of inhibitors [81]. This pre-treatment process usually has involved temperatures of 473–

503 K for up to 15 min. Around 40–60% of the total mass is dis-solved in this process, with 4–22% of the cellulose, 35–60% of thelignin and all of the hemicellulose being removed [101]. If the pHis maintained between 4 and 7, the degradation of monosaccharidesugars can be minimized [102].

5.1.3. Chemical pretreatment5.1.3.1. Ozonolysis. Ozonolysis involves using ozone gas to breakdown the lignin and hemicellulose and increase the biodegradabil-ity of the cellulose. The pretreatment is usually carried out at roomtemperature and is effective at lignin removal without the forma-tion of toxic by-products [103]. Ozonation has been widely used toreduce the lignin content of both agricultural and forestry wastes[104]. Ozonolysis has been shown to break down 49% of lignin incorn stalks and 55–59% of lignin in autohydrolyzed (hemicellulosefree) corn stalks [105]. In a study [106], wheat and rye straws werepretreated with ozone to increase the enzymatic hydrolysis extentof potentially fermentable sugars. Enzymatic hydrolysis yields ofup to 88.6% and 57% were obtained compared to 29% and 16% innon-ozonated wheat and rye straw respectively. A drawback ofozonolysis is that a large amount of ozone is required, which canmake the process expensive [68].

5.1.3.2. Alkaline pretreatment. Alkali pretreatment refers to theapplication of alkaline solutions to remove lignin and various uron-ic acid substitutions on hemicellulose that lower the accessibilityof enzyme to the hemicellulose and cellulose [107,108]. These pro-cesses utilize lower temperatures and pressures compared to otherpretreatment technologies. Alkali pretreatment may be carried outat ambient conditions, but pretreatment time is measured in termsof hours or days rather than minutes or seconds [86]. Regardlessthe advantages, these methods present difficulties from the pointof view of the process economy for obtaining fuels [94]. Sodium,potassium, calcium and ammonium hydroxide are appropriatechemicals for pretreatment. Of these four, NaOH has been studiedthe most [68]. Dilute NaOH treatment of lignocellulosic biomasscauses swelling, leading to an increase in the internal surface area,a decrease in crystallinity, separation of structural linkages be-tween lignin and carbohydrates, and disruption of the lignin struc-ture [109]. Millet et al. [110] reported that the digestibility ofNaOH-treated hardwood to increase from 14% to 55% with the de-crease of lignin content from 24%–55% to 20%. However, no effectof dilute NaOH pretreatment was found for softwoods with lignincontent greater than 26%. Silverstein et al. [107] reported >65% lig-nin reduction in cotton stalk treated with 2% NaOH for 90 min at394 K/15 psi. Cellulases produced by Bacillus subtilis for the sac-charification of wheat straw, rice straw and bagasse were usedby Akhtar et al. [111]. Pretreatment of these substrates with 2%NaOH was found to be more effective for increasing the saccharifi-cation. The saccharification rates of 33.0%, 25.5% and 35.5% wereobtained with 2% NaOH pretreated wheat straw, rice straw and ba-gasse, respectively. In a study [112], a combination of NaOH treat-ment and homogenization was used as a pretreatment to enhancethe enzymatic hydrolysis of corn stover. The highest glucose yield(6.25 g/l) was obtained when the corn stover was pretreated by acombination of 1.0 N NaOH treatment and homogenization (Fig. 5).

Lime (Ca(OH)2) as compared to NaOH and KOH has lower costand less significant safety requirements. It can be recovered fromhydrolyzate by reaction with CO2, so that formed carbonate canthen be reconverted to lime [113]. To make lime as efficient asother alkalis in enhancing the digestibility of lignocellulose, appro-priate pretreatment conditions need to be employed [114]. Lime,water, and an oxidizing agent (air or O2) are mixed with the bio-mass at temperatures ranging from 313 to 423 K for a period rang-ing from hours to weeks [115]. Two types of lime treatment havebeen explored: (1) short-term and (2) long-term. Short-term lime

Page 8: 1-s2.0-S0196890410003791-main

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

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

M. Balat / Energy Conversion and Management 52 (2011) 858–875 865

pretreatment involves boiling the biomass with a lime loading of0.1 g Ca(OH)2/g dry biomass at temperatures of 358–408 for 1–3 h [116]. Long-term pretreatment involves using the same limeloading at lower temperatures (313–328 K) for 4–6 weeks in thepresence 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-atmoxygen) [120]. Saha and Cotta [118] obtained maximum total sugaryield (451 ± 3mg g�1 straw; glucose, 252 ± 6 mg; xylose, 173 ±3 mg; arabinose, 27 ± 2 mg; 65% conversion) by lime pretreatment(100mg g�1 straw, 394 K, 1 h). The authors also investigated theeffects of pH (3.5–6.5) and temperature (298–343 K) on the enzy-matic hydrolysis of lime pretreated wheat straw (8.6%, w/v) using acombination of three enzymes (cellulose, b-glucosidase, and hemi-cellulase), each enzyme at a dose level of 0.05 ml g�1 substrate.Fig. 6 shows the effect of pH and temperature on the enzymatichydrolysis of lime pretreated wheat straw.

Alkaline peroxide is one of the effective pretreatment methodsthat can improve the enzymatic hydrolysis by delignification of lig-nocellulosic materials. In this method, the lignocelluloses are soakedin pH-adjusted water (e.g. to pH 11–12 using NaOH) containingH2O2 at room temperatures for a period of time (e.g. 6–24 h) [56].

5.1.3.3. Acid pretreatment. Acid pretreatments normally aim forhigh yields of sugars from lignocellulosic materials. Acid pretreat-ment involves the use of sulfuric, nitric, or hydrochloric acids to re-move hemicellulose components and expose cellulose forenzymatic digestion [107]. The acid pretreatment can operateeither under a high temperature and low acid concentration (diluteacid pretreatment) or under a low temperature and high acid con-centration (concentrated acid pretreatment) [56]. Dilute acidhydrolysis has been successfully developed for pretreatment of lig-nocellulosic materials. The dilute acid pretreatment works fairlywell on agricultural feedstocks, such as corn stover and rice/wheatstraw [62]. While dilute acid pretreatments are known to improveenzymatic hydrolysis, their cost is relatively high compared tophysico-chemical pretreatments [44]. This pretreatment methodgives high reaction rates and significantly improves cellulosehydrolysis [47]. There are primarily two types of dilute acid pre-treatment processes: low solids loading (5–10% [w/w]), high tem-perature (T > 433 K), continuous-flow processes and high solids

loading (10–40% [w/w]), lower temperature (T < 433 K), batch pro-cesses [121]. In general, higher pretreatment temperatures andshorter reactor residence times result in higher soluble xyloserecovery yields and enzymatic cellulose digestibility. Higher-tem-perature dilute acid pretreatment has been shown to increase cel-lulose digestibility of pretreated residues [122]. Depending on thesubstrate and the conditions used, between 80% and 95% of thehemicellulosic sugars can be recovered by dilute acid pretreatmentfrom the lignocellulosic material [47,123,124]. Silverstein et al.[107] reported 95% xylan reduction in cotton stalk treated with2% H2SO4 for 90 min at 394 K/15 psi.

In recent years, treatment of lignocellulosic biomass with dilutesulfuric acid has been primarily used as a means of hemicellulosehydrolysis and pretreatment for enzymatic hydrolysis of cellulose[125]. Sulfuric acid at concentrations usually below 4 wt.%, hasbeen of the most interest in such studies as it is inexpensive andeffective [68]. Dilute sulfuric acid pretreatment (0.2–2.0% sulfuricacid, 394–493 K) of lignocellulose serves three important functionsin the conversion process [121]: (1) hydrolysis of the hemicellulosecomponents to produce a syrup of monomeric sugars, (2) exposureof cellulose for enzymatic digestion by removal of hemicelluloseand part of the lignin, and (3) solubilization of heavy metals whichmay be contaminating the feedstock. In spite of these benefits, di-lute sulfuric acid has some important disadvantages [76]: (1) cor-rosion that mandates expensive materials of construction, (2)acidic prehydrolyzates must be neutralized before the sugars pro-ceed to fermentation, (3) gypsum has problematic reverse solubil-ity characteristics when neutralized with inexpensive calciumhydroxide, (4) formation of degradation products and release ofnatural biomass fermentation inhibitors are other characteristicsof acid pretreatment, (5) disposal of neutralization salts is needed,and (6) biomass particle size reduction is necessary.

Page 9: 1-s2.0-S0196890410003791-main

866 M. Balat / Energy Conversion and Management 52 (2011) 858–875

5.1.4. Biological pretreatmentBiological pretreatment involves microorganisms such as

brown-, white- and soft-rot fungi that are used to degrade ligninand solubilize hemicellulose. White-rot fungi are the most effectivebiological pretreatment of lignocellulosic materials [56,68,74]. Leeet al. [126] evaluated biological pretreatment of Japanese red pine(Pinus densiflora) using three white-rot fungi (Ceriporia lacerata,Stereum hirsutum, and Polyporus brumalis). pretreatment withS. hirsutum resulted in selective degradation of the lignin rather thanthe holocellulose component. The advantages of biological pretreat-ment include low energy requirement and mild environmental con-ditions. However, the rate of hydrolysis in most biologicalpretreatment process is very low [74].

5.2. Hydrolysis techniques

The carbohydrate polymers in lignocellulosic materials need tobe converted to simple sugars before fermentation, through a pro-cess called hydrolysis [127]. Various methods for the hydrolysis oflignocellulosic materials have recently been described. The mostcommonly applied methods can be classified in two groups: chem-ical hydrolysis (dilute and concentrated acid hydrolysis) and enzy-matic hydrolysis. There are some other hydrolysis methods inwhich no chemicals or enzymes are applied. For instance, lignocellu-lose may be hydrolyzed by gamma-ray or electron-beam irradiation,or microwave irradiation. However, those processes are commer-cially unimportant.

Several products can result from hydrolysis of lignocellulosicmaterial (Fig. 7) [59]. When hemicelluloses are hydrolyzed to xy-lose, mannose, acetic acid, galactose, and glucose are liberated.The main application of xylose is its bioconversion to xylitol, afunctional sweetener with important technological properties[128]. Hemicellulosic hydrolysis can be generalized as

Hemicelluloses ! Xylan ! Xylose ! Furfural ð1Þ

Acetyl groups ! Acetic acid ð2Þ

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

Cellulose ! Glucan ! Glucose

! Decomposition products ð3Þ

Fig. 7. Main degradation products occurring during hydrolysis of lignocellulosicmaterial.

5.2.1. Chemical hydrolysisChemical hydrolysis involves exposure of lignocellulosic mate-

rials to a chemical for a period of time at a specific temperature,and results in sugar monomers from cellulose and hemicellulosepolymers [127]. In the chemical hydrolysis, the pretreatment andthe hydrolysis may be carried out in a single step. Acids are pre-dominantly applied in chemical hydrolysis [127]. There are two ba-sic types of acid hydrolysis processes: dilute acid and concentratedacid, each with variations.

5.2.1.1. Dilute acid hydrolysis. Dilute acid hydrolysis is the oldesttechnology for converting cellulose biomass to bioethanol. Thisprocess is conducted under high temperature and pressure, andhas a reaction time in the range of seconds or minutes, which facil-itates continuous processing. Dilute acid process involves a solu-tion of about 1% H2SO4 concentration in a continuous flowreactor at a high temperature (about 488 K). Most dilute acid pro-cesses are limited to a sugar recovery efficiency of around 50%[132]. The combination of acid and high temperature and pressuredictate special reactor materials, which can make the reactorexpensive. The first reaction converts the cellulosic materials to su-gar and the second reaction converts the sugars to other chemicals.Unfortunately, the conditions that cause the first reaction to occuralso are the right conditions for the second to occur [53].

Dilute acid hydrolysis occurs in two-stages to take advantage ofthe differences between hemicellulose and cellulose. The first-stage is performed at low temperature to maximize the yield fromthe hemicellulose, and the second, higher temperature stage isoptimized for hydrolysis of the cellulose portion of the feedstock[133]. The first-stage is conducted under mild process conditions(e.g. 0.7% H2SO4, 463 K) to recover the five-carbon sugars, whilein the second stage only the remaining solids with the more resis-tant cellulose undergo harsher conditions (488 K, but a milder 0.4%H2SO4) to recover the six-carbon sugars [55]. Modern experimentalyields for this two-stage acid process (3 min per stage) are 89% formannose, 82% for galactose, but only 50% for glucose with a glu-cose to bioethanol conversion that is 90% of the theoretical maxi-mum [102]. Schematic flowsheet for dilute acid hydrolysis isgiven in Fig. 8. The primary challenge for dilute acid hydrolysisprocesses is how to raise glucose yields higher than 70% in an eco-nomically viable industrial process while maintaining a high cellu-lose hydrolysis rate and minimizing glucose decomposition. Forrapid continuous processes, in order to allow adequate acid pene-tration, feedstocks must also be reduced in size so that the maxi-mum 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 hemicellulo-sic sugars while the subsequent stage requires the biomass to bedried followed by the addition of concentrated sulfuric acid (70–90%) [102]. The acid concentration used in concentrated acid-hydrolysis process is in the range of 10–30% [134]. Reaction timesare typically much longer than for dilute acid process. This processprovides a complete and rapid conversion of cellulose to glucoseand hemicelluloses to five-carbon sugars with little degradation.The critical factors needed to make this process economically via-ble are to optimize sugar recovery and cost effectively recovers theacid for recycling [135].

In comparison to dilute acid hydrolysis, concentrated acidhydrolysis leads to little sugar degradation and gives sugar yieldsapproaching 100% [136]. Table 7 shows the yields of bioethanolby concentrated sulfuric acid hydrolysis from cornstalks. The con-centrated acid process offers more potential for cost reductionsthan the dilute acid process. However, environment and corrosionproblems and the high cost of acid consumption and recovery pres-ent major barriers to economic success [136].

Page 10: 1-s2.0-S0196890410003791-main

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

M. Balat / Energy Conversion and Management 52 (2011) 858–875 867

5.2.2. Enzymatic hydrolysisAcid hydrolysis has a major disadvantage where the sugars are

converted to degradation products like tars. This degradation canbe prevented by using enzymes favoring 100% selective conversionof cellulose to glucose. When hydrolysis is catalyzed by such en-zymes, the process is known as enzymatic hydrolysis [137]. Enzy-matic hydrolysis of natural lignocellulosic materials is a very slowprocess because cellulose hydrolysis is hindered by structuralparameters of the substrate, such as lignin and hemicellulose con-tent, surface area, and cellulose crystallinity [138]. Utility cost ofenzymatic hydrolysis is low compared to acid or alkaline hydroly-sis because enzyme hydrolysis is usually conducted at mild condi-tions (pH 4.8) and temperature (318–323 K) and does not have acorrosion problem [74]. The enzymatic hydrolysis has currentlyhigh yields (75–85%) and improvements are still projected (85–95%), as the research field is only a decade young [55]. Comparisonof process conditions and performance of three cellulose hydrolysisprocesses is given in Table 8. Enzymatic hydrolysis is an environ-mentally friendly alternative that involves using carbohydratedegrading enzymes (cellulases and hemicellulases) to hydrolyzelignocelluloses into fermentable sugars [44].

5.2.2.1. Enzymatic hydrolysis of cellulose. Cellulose is typicallyhydrolyzed by an enzyme called cellulase. These enzymes are pro-duced by several microorganisms, commonly by bacteria and fun-gi. These microorganisms can be aerobic or anaerobic, mesophilicor thermophilic. Bacteria belonging to Clostridium, Cellulomonas,

Table 7Yields of bioethanol by concentrated sulfuric acid hydrolysis from cornstalks [53].

Amount of cornstalk (kg) 1000Cellulose content (kg) 430Cellulose conversion and recovery

efficiency (% dry weight)76

Bioethanol stoichiometric yield (% dry weight) 51Glucose fermentation efficiency (% dry weight) 75Bioethanol yield from glucose (kg) 130Amount of cornstalk (kg) 1000Hemicelluloses content (kg) 290Hemicelluloses conversion and recovery

efficiency (% dry weight)90

Bioethanol stoichiometric yield (% dry weight) 51Xylose fermentation efficiency (% dry weight) 50Bioethanol yield from xylose (kg) 66Total bioethanol yield from 1000 kg of cornstalks 196 kg

(225.7 L = 59 gallons)

Bacillus, Thermomonospora, Ruminococcus, Bacteriodes, Erwinia,Acetovibrio, Microbispora, and Streptomyces can produce cellulaseseffectively [74]. Fungi such as Sclerotium rolfsii, P. chrysosporiumand species of Trichoderma, Aspergillus, Schizophyllum and Penicili-um are used to produce cellulases [139]. Mutant strains of Tricho-derma sp. (T. viride, T. reesei, T. longibrachiatum) have long beenconsidered to be the most productive and powerful destroyers ofcrystalline cellulose [140]. Commercial products of various T. reeseiisolates have been available for a long time in cereal foods applica-tions, the brewing industry, fruit and vegetable processing andhave also been widely evaluated and applied in relation to bioeth-anol production processes. T. reesei secretes high amounts of en-zymes, up to 100 g l�1 [141].

Cellulase is a group of enzymes that synergistically hydrolyzescellulose (Fig. 9) [142]. The widely accepted mechanism for enzy-matic cellulose hydrolysis involves synergistic actions by endoglu-canses (EG, endo-1,4-b-D-glucanases, or EC 3.2.1.3.), exoglucanasesor cellobiohydrolases (CBH, 1,4-b-D-glucan cellobiohydrolases, orEC 3.2.1.91.), and b-glucosidases (BGL, cellobiases or EC 3.2.1.21).EG hydrolyze accessible intramolecular b-1,4-glucosidic bonds ofcellulose chains randomly to produce new chain ends; CBH proc-essively cleave cellulose chains at the ends to release soluble cello-biose or glucose; and BGL hydrolyze cellobiose to glucose in orderto eliminate cellobiose inhibition [143]. BGL complete the hydroly-sis process by catalyzing the hydrolysis of cellobiose to glucose.The supplementation of b-glucosidase in hydrolysis is requireddue to its insufficient amount from T. reesei, to prevent cellulasesinhibition resulted from cellobiose accumulation [114]. During cel-lulose hydrolysis, the solid substrate characteristics vary, includ-ing: (1) changes in the cellulose chain end number resultingfrom generation by EG and consumption by CBH and (2) changesin cellulose accessibility resulting from substrate consumptionand cellulose fragmentation [143].

There are different factors that affect the enzymatic hydrolysisof cellulose, namely, substrates, cellulase activity, reaction condi-tions (temperature, pH as well as other parameters), and a strongproduct inhibition. To improve the yield and rate of enzymatichydrolysis, research has been focused on optimizing the hydrolysisprocess and enhancing the cellulase activity [74]. The rate of enzy-matic hydrolysis of cellulose is dependent upon several structuralfeatures of the cellulose. The cellulose features known to affectthe rate of hydrolysis include: (1) molecular structure of cellulose,(2) crystallinity of cellulose, (3) surface area of cellulose fiber, (4)degree of swelling of cellulose fiber, (5) degree of polymerization,and (6) associated lignin or other materials [144]. A low substrate

Page 11: 1-s2.0-S0196890410003791-main

Table 8Comparison of process conditions and performance of three hydrolysis processes [55].

Consumables Temperature (K) Time Glucose yield (%) Available

Dilute acid <1% H2SO4 488 3 min 50–70 NowConcentrated acid 30–70% H2SO4 313 2–6 h 90 NowEnzymatic Cellulase 323 1.5 days 75 ? 95 Now ? 2020

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

Fig. 10. Effects of T. reesei ZU-02 cellulase dosage (presented as filter paperactivities per gram of substrate, FPU g�1 substrate) on the enzymatic hydrolysis ofdilute acid-treated corncob [146].

868 M. Balat / Energy Conversion and Management 52 (2011) 858–875

concentration gives low yield and rate, and a high cellulase dosagemay increase the costs disproportional [55]. A cellulase dosage of 10FPU (filter paper units) per gram of biomass is often used in labora-tory studies because it provides a hydrolysis profile with high levelsof glucose yield in a reasonable time (48–72 h) at a reasonable en-zyme cost [145]. Chen et al. [146] investigated effects of cellulasedosage on the enzymatic hydrolysis of dilute acid-treated corncob.Hydrolysis experiments were performed with 100 g l�1 substrateand different dosages of T. reesei ZU-02 cellulase (FPU g�1 substrate)at pH 4.8 and 323 K. Results of this study are shown in Fig. 10. Asshown in Fig. 10, reducing sugar concentration and hydrolysis yieldhad a similar variation trend, that is, both increased sharply with cel-lulase dosage varying from 10 to 20 FPU g�1 substrate, and basicallyleveled off from 20 to 30 FPU g�1 substrate.

One limitation with using cellulase is that there is a reduction inrates due to end product (cellobiose and glucose) inhibition. Simul-taneous saccharification and fermentation (SSF) overcomes thisproblem by hydrolyzing cellulose and fermenting the hydrolysisproduct at the same time [147].

5.2.2.2. Enzymatic hydrolysis of hemicelluloses. There is a great inter-est in the enzymatic hydrolysis of xylan because of possible appli-cations in ruminal digestion, waste treatment, fuel and chemicalproduction, and paper manufacture [148]. Unlike cellulose, xylansare chemically quite complex, and their degradation requires mul-tiple enzymes. Xylan-degrading enzymes are produced by a widevariety of fungi and bacteria such as Trichodrema spp. [149,150],Penicillium spp. [151,152], Talaromyces spp. [151,153] Aspergillusspp. [154], and Bacillus spp. [155].

Enzymatic hydrolysis of xylan involves a multi-enzyme system,including endoxylanase, exoxylanase, ß-xylosidase, a-arabinofura-nosidase, a-glucoronisidase, acetyl xylan esterase, and ferulic acidesterase [156]. Table 9 presents most important enzyme activitiesrequired for hydrolysis of xylooligosaccharides obtained fromhardwoods and herbaceous type materials [113]. The endoxylan-ase attacks the main chains of xylans and b-xylosidase hydrolyzesxylooligosaccharides to xylose. The a-arabinofuranosidase and a-glucuronidase remove the arabinose and 4-0-methyl glucuronicacid substituents, respectively, from the xylan backbone [157].Hemicellulolytic esterases include acetyl esterases which hydro-

lyze the acetyl substitutions on xylose moieties, and feruloyl ester-ases which hydrolyze the ester bond between the arabinosesubstitutions and ferulic acid. Feruloyl esterases aid the releaseof hemicellulose from lignin and renders the free polysaccharideproduct 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 ismuch greater than for cellulose hydrolysis, accessibility to the sub-strate is easier since xylan does not form tight crystalline struc-tures [44].

5.3. Fermentation

The supernatant from enzymatic hydrolysis of lignocellulosescan contain both six-carbon (hexoses) and five-carbon (pentoses)sugars (if both cellulose and hemicellulose are hydrolyzed).Depending on the lignocellulose source, the hydrolysate typicallyconsists of glucose, xylose, arabinose, galactose, mannose, fucose,

Page 12: 1-s2.0-S0196890410003791-main

Table 9Relevant enzymatic activities for enzymatic posthydrolysis of xylooligosaccharides [113].

Enzyme EC Hydrolyzed linkage Substrate Main product

Endoxylanase 3.2.1.8 Internal b-1,4 Main chain OligomersExoxylanase n.c. Terminal b-1,4 (reducing end) Main chain Xylose, xylobioseb-Xylosidase 3.2.1.37 Terminal b-1,4 (non-reducing end) Oligomers XyloseArabinosidase 3.2.1.55 Side groups ArabinoseGlucoronisidase 3.2.1.139 Side groups Methylglucuronic acidsAcetyl xylan esterases 3.1.1.72 Ester bond Side groups Acetic acidFeruloyl esterases 3.1.1.73 Ester bond Side groups Ferulic acid

n.c.: not yet classified.

Table 10Important traits for bioethanol fermentation process [160].

Trait Requirement

Bioethanol yield >90% of theoreticalBioethanol tolerance >40 g l�1

Bioethanol productivity >1 g l�1 h�1

Robust grower and simple growthrequirements

Inexpensive mediumformulation

Able to grow in undiluted hydrolysates Resistance to inhibitorsCulture growth conditions retard

contaminantsAcidic pH or highertemperatures

Table 11Comparison between modified Z. mobilis and E. coli [94].

Z. mobilis E. coli

Bioethanol (g l�1) 62 27Bioethanol yielda (%) 97 90Bioethanol productivity (g l�1 h�1) 1.29 0.92

a Estimation from the theoretical yields.

M. Balat / Energy Conversion and Management 52 (2011) 858–875 869

and rhamnose [44]. One ton of glucan, galactan, or mannan yields1.11 tons of six-carbon sugars and could be fermented theoreti-cally into 172.0 gallons of bioethanol [159]. One ton of arabinanor xylan yields 1.14 tons of five-carbon sugars and could be fer-mented theoretically into 176.0 gallons of bioethanol [159]. Micro-organisms can be used to ferment all lignocellulose-derived sugarsto bioethanol.

5.3.1. Microorganisms related to bioethanol fermentationMicroorganisms for bioethanol fermentation can best be de-

scribed in terms of their performance parameters and other require-ments such as compatibility with existing products, processes andequipment. The performance parameters of fermentation are: tem-perature range, pH range, alcohol tolerance, growth rate, productiv-ity, osmotic tolerance, specificity, yield, genetic stability, andinhibitor tolerance. The characteristics required for an industriallysuitable microorganism are summarized in Table 10 [160].

Traditionally, Saccharomyces cerevisiae and Zymomonas mobilishave been used for bioethanol fermentation. They are capable ofefficiently fermenting glucose into bioethanol, but are unable toferment xylose [44]. Natural xylose-fermenting yeasts, such asPichia stipitis, Candida shehatae, and Candida parapsilosis, canmetabolize xylose via the action of xylose reductase (XR) to con-vert xylose to xylitol, and of xylitol dehydrogenase (XDH) to con-vert xylitol to xylulose. Therefore, bioethanol fermentation fromxylose can be successfully performed by recombinant S. cerevisiaecarrying 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 metab-olized through the pentose phosphate pathway (PPP) [162].

The most frequently used microorganism for fermenting bio-ethanol in industrial processes is S. cerevisiae, which has provedto be very robust and well suited to the fermentation of lignocellu-losic hydrolysates [91]. S. cerevisiae can easily ferment hexoses, buthardly xylose in lignocellulose hydrolysates, because S. cerevisiaelacks enzymes that convert xylose to xylulose [163]. However, thisyeast can ferment xylulose [164]. For xylose-using S. cerevisiae,high bioethanol yields from xylose also require metabolic engi-neering strategies to enhance the xylose flux [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. mobi-lis, a Gram-negative bacterium, is well recognized for its ability toefficiently produce bioethanol at high rates from glucose, fructose,and sucrose. When Z. mobilis and S. cerevisiae were compared fortheir efficiency to produce bioethanol from glucose and starchhydrolysate, higher yield was observed for Z. Mobilis [166]. Com-parative performance trials on glucose have shown that Z. mobiliscan achieve 5% higher bioethanol yields and up to 5-fold higherbioethanol volumetric productivity compared to traditional S. cere-visiae yeast [167]. It has a theoretical yield of 97% [94]. Z. mobilisefficiently produces bioethanol from the hexose sugars glucoseand fructose but not from pentose sugars, although a xylose fer-menting Z. mobilis was generated by introducing a xylose-metabo-lizing pathway from E. coli [165]. Modified Z. mobilis has theadvantages of requiring a minimum of nutrients, growing at lowpH and high temperatures, and it is considered generally recog-nized as safe (GRAS). A comparison of modified Z. mobilis andE. coli showing their respective advantages is shown in Table 11[94]. E. coli and K. oxytoca naturally metabolize arabinose, such thatthe ethanologenic strains ferment all lignocellulose-derived sugars[165]. The construction of E. coli strains to selectively produce bio-ethanol was one of the first successful applications of metabolicengineering [168]. E. coli, as a biocatalyst for bioethanol produc-tion, has ability to ferment a wide spectrum of sugars, no require-ments for complex growth factors, and prior industrial use (e.g. forproduction of recombinant protein). The major disadvantagesassociated with using E. coli cultures are a narrow and neutralpH growth range (6.0–8.0), less hardy cultures compared to yeast,and public perceptions regarding the danger of E. coli strains. Thelack of data on the use of residual E. coli cell mass as an ingredientin animal feed is also an obstacle to its application [160,169]. K.oxytoca is an enteric bacterium found growing in paper and pulpstreams as well as around other sources of wood. The microorgan-ism is capable of growing at a pH at least as low as 5.0 and temper-atures as warm as 308 K. It can grow on a wide variety of sugarsincluding hexoses and pentoses, as well as on cellobiose and cello-triose [160].

Thermophilic anaerobic bacteria have also been extensivelyexamined for their potential as bioethanol producers. These bacte-ria include Thermoanaerobacter ethanolicus [170], Clostridium ther-mohydrosulfuricum [171], Thermoanaerobacter mathranii [172],Thermoanaerobium brockii [173], Clostridium thermosaccharolyticum[174], etc. Thermophilic anaerobic bacteria have a distinct advan-tage over conventional yeasts for bioethanol production in their

Page 13: 1-s2.0-S0196890410003791-main

870 M. Balat / Energy Conversion and Management 52 (2011) 858–875

ability to use a variety of inexpensive biomass feedstocks and theirability to withstand temperature extremes [175]. The low bioetha-nol tolerance of thermophilic anaerobic bacteria (<2%, v/v) is a ma-jor obstacle for their industrial exploitation for bioethanolproduction [176].

5.3.2. Fermentation techniquesFermentation can be performed as a batch, fed-batch or contin-

uous process. The choice of most suitable process will depend uponthe kinetic properties of microorganisms and type of lignocellu-losic hydrolysate in addition to process economics aspects [66].Batch culture can be considered as a closed culture system whichcontains an initial, limited amount of nutrient, which is inoculatedwith microorganisms to allow the fermentation [177]. It is verysimple method, during the fermentation nothing is added afterinoculation except possibly acid or alkali for pH control or air foraerobic fermentations.

Fed-batch reactors are widely used in industrial applicationsbecause they combine the advantages from both batch and contin-uous processes [178]. The major advantage of fed-batch, compar-ing to batch, is the ability to increase maximum viable cellconcentration, prolong culture lifetime, and allow product accu-mulation to a higher concentration [179]. This process allows forthe maintenance of critical process variables (e.g. temperature,pH, and dissolved oxygen) at specific levels through feedback con-trol [180].

In the continuous process, feed, which contains substrate, cul-ture medium and other required nutrients, is pumped continu-ously into an agitated vessel where the microorganisms areactive. The product, which is taken from the top of the bioreactor,contains bioethanol, cells, and residual sugar [181]. One of the firstapproaches taken in improving the yeast bioethanol fermentationprocess involved operating the fermenters in a continuous moderather than the more conventional batch mode and therebyincreasing the productivity about threefold from about 2 to 6 gEtOH/l/h [182]. Operating continuously at higher cell densitiesusing cell recycle reactors was another effective means of greatlyincreasing the productivity. A single-stage continuous stirred-tankreactor (CSTR) with cell recycle operating at high biomass loading(50–80 g yeast/l) has a bioethanol productivity of 30–40 g EtOH/l/h[182].

5.3.3. Hydrolysis and fermentation strategies5.3.3.1. Separate hydrolysis and fermentation (SHF). Enzymatichydrolysis performed separately from fermentation step is knownas separate hydrolysis and fermentation (SHF). In the SHF configu-ration the joint liquid flow from both hydrolysis reactors first en-ters the glucose fermentation reactor. The mixture is thendistilled to remove the bioethanol leaving the unconverted xylosebehind. In a second reactor, xylose is fermented to bioethanol, andthe bioethanol is again distilled [55]. The advantage of SHF is theability to carry out each step under optimal conditions, i.e. enzy-matic hydrolysis at 318–323 K and fermentation at about 303 K[183]. The disadvantage of this method is the inhibition of cellulaseand b-glucosidase enzymes by glucose released during hydrolysis,which calls for lower solids loadings and higher enzyme loadingsto achieve reasonable yields [121].

5.3.3.2. Simultaneous saccharification and fermentation (SSF). Simul-taneous saccharification and fermentation (SSF) is a promising pro-cess option for production of bioethanol from lignocellulosicmaterials [184]. This process is often effective when combinedwith dilute acid or high temperature hot-water pretreatment. InSSF, cellulases and xylanases convert the carbohydrate polymersto fermentable sugars. These enzymes are notoriously susceptibleto feedback inhibition by the products – glucose, xylose, cellobiose,

and other oligosaccharides [123]. This process has an enhancedrate of hydrolysis, needs lower enzyme loading, results in higherbioethanol yields, and reduces the risk of contamination. Presently,an SSF process for e.g. wheat straw hydrolyzate can be expected togive final bioethanol concentrations close to 40 g l�1 with a yieldbased on total hexoses and pentoses higher than 70% [185].

SSF requires compatible fermentation and saccharification con-ditions, with a similar pH, temperature and optimum substrateconcentration [186]. In many cases, the low pH, e.g. lower than5, and high temperature, e.g. >313 K, may be favorable for enzy-matic hydrolysis, whereas the low pH can surely inhabit the lacticacid production and the high temperature may affect adversely thefungal cell growth [187]. Trichoderma reesei cellulases, which con-stitute the most active preparations, have optimal activity at pH4.5 and 328 K. For Saccharomyces cultures SSF are typically con-trolled at pH 4.5 and 310 K [160]. A typical fermentation will take5–7 days, depending on the accessibility of the cellulose and initialsolids loading of the fermentation. The long residence time maymake contamination control difficult in a continuous process, butmay be manageable in a batch process [188].

Major advantages of SSF as described by Sun and Cheng [74], in-clude: (1) increase of hydrolysis rate by conversion of sugars thatinhibit the cellulase activity; (2) lower enzyme requirement; (3)higher product yields; (4) lower requirements for sterile conditionssince glucose is removed immediately and bioethanol is produced;(5) shorter process time; and (6) less reactor volume. The majoradvantage of SSF is that the immediate consumption of sugars bythe microorganism produces low sugar concentrations in the fer-mentor, which significantly reduces enzyme inhibition comparedto SHF [188]. The main disadvantage of SSF lies in different temper-ature optima for saccharification (323 K) and fermentation (308 K)[189].

The most important process improvement made for the enzy-matic hydrolysis of biomass is the introduction of SSF, which hasbeen improved to include combines the cellulase enzymes and fer-menting microbes in one vessel to improve the bioethanol produc-tion economics. The technology has been improved to include theco-fermentation of multiple sugar substrates, i.e., simultaneoussaccharification of both cellulose (to glucose) and hemicellulose(to xylose), and co-fermentation of both glucose and xylose bygenetically engineered microbes in the same broth [136].

5.3.3.3. Direct microbial conversion (DMC). Direct microbial conver-sion (DCM) combines cellulase production, cellulose hydrolysisand glucose fermentation into a single step. This process is attrac-tive in that it reduces the number of reactors, simplifies operation,and reduces the cost of chemicals [121]. DCM seems the logicalendpoint in the evolution of bioethanol production from lignocel-lulosic materials. Application of DCM entails no operating cost orcapital investment for dedicated enzyme production (or purchase),reduced diversion of substrate for enzyme production, and com-patible enzyme and fermentation systems [55]. The disadvantagesare 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 approachcommercial viability, advancements in product recovery technolo-gies will be required. For cases in which fermentation products aremore volatile than water, recovery by distillation is often the tech-nology of choice. Distillation technologies that will allow the eco-nomic recovery of dilute volatile products from streamscontaining a variety of impurities have been developed and

Page 14: 1-s2.0-S0196890410003791-main

Table 12Estimates of the costs of bioethanol production from different feedstock (US$/l) [201].

Feedstock Production costrangea in 2005

Projected productioncost rangea in 2030

Sugarcane 0.20–0.50 0.20–0.35Corn 0.60–0.80 0.35–0.55Sugar beet 0.62–0.82 0.40–0.60Wheat 0.70–0.95 0.45–0.65Lignocellulose 0.80–1.10 0.25–0.65

a Excluding any subsidies to bioethanol production.

M. Balat / Energy Conversion and Management 52 (2011) 858–875 871

commercially demonstrated [190]. A distillation system separatesthe bioethanol from water in the liquid mixture.

The first step is to recover the bioethanol in a distillation or beercolumn, where most of the water remains with the solids part. Theproduct (37% bioethanol) is then concentrated in a rectifying col-umn to a concentration just below the azeotrope (95%) [55]. Theremaining bottoms product is fed to the stripping column to re-move additional water, with the bioethanol distillate from strip-ping being recombined with the feed to the rectifier [191]. Therecovery of bioethanol in the distillation columns in the plant isfixed to be 99.6% to reduce bioethanol losses [192].

After the first effect, solids are separated using a centrifuge anddried in a rotary dryer. A portion (25%) of the centrifuge effluent isrecycled to fermentation and the rest is sent to the second andthird evaporator effects. Most of the evaporator condensate is re-turned to the process as fairly clean condensate (a small portion,10%, is split off to waste water treatment to prevent build up oflow-boiling compounds) and the concentrated syrup contains15–20% by weight total solids [193].

6. Bioethanol economy

Considering that up to now the cost of bioethanol was consider-ably higher than the cost of fossil gasoline supply, national govern-ments had to enact special policies in order to encourageproduction and use of bioethanol in the transportation sector. Ingeneral, the following three main approaches can be distinguishedin the implementation of biofuels supporting policies and regula-tion: (1) taxation-based policies, (2) agriculture-based policies/subsidies, and (3) fuel mandates [34]. At present, the developmentand promotion of biofuels are mainly driven by the agriculturalsector and green lobbies rather than the energy sector. In fact, mostbiofuel 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 pro-gression of more efficient and cheaper technology, biofuels couldbe a cost-effective alternative in the near future in many countries[194].

The cost for bioethanol production can vary substantiallydepending on several factors, e.g. feedstock costs and by-productsrevenues, cost of process energy, investment costs (related to thetype of feedstock), plant location and transportation cost andfinancing costs [195]. Brazilian bioethanol is far more competitivethan that produced in the United States from corn or in Europefrom sugar beet, because of shorter processing times, lower labourcosts, lower transport costs and input costs [196]. Bioethanol pro-duction from sugarcane is very economical in Brazil because of twoprimary reasons. Brazil dropped support of sugar prices to supportthe bioethanol industry with government established mandates forthe blending of bioethanol with gasoline. This drastically loweredthe cost of the feedstock, sugarcane, and created a demand forand supported the price of bioethanol. In addition, Brazil’s vastland area of cultivatable acreage means that land devoted to sugar-cane production for bioethanol is not in competition with land de-voted for food production [197]. Bioethanol from sugarcane inBrazil costs US$0.23–0.29/l [198], while in the EU and the UnitedStates sugar beet and corn-derived bioethanol cost US$0.29/l[199] and US$0.53/l [200], respectively. Other efficient sugar pro-ducing countries such as Pakistan, Swaziland and Zimbabwe haveproduction costs similar to Brazil’s [194].

The cost of raw material, which varies considerably betweendifferent studies (US$22–US$61 per metric ton dry matter), andthe capital costs, which makes the total cost dependent on plantcapacity, contribute most to the total production cost [165]. Thecost and availability of feedstock was crucial because in most bio-

fuels, feedstock represents 60–75% of the total bioethanol produc-tion cost [3]. Estimates of the costs of bioethanol production fromdifferent feedstock are shown in Table 12. The cost figures can becompared 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 foodcrops such as corn grain and sugarcane, but the limited supply ofthese crops can lead to competition between their use in bioetha-nol production and food provision [202]. Using food crops to pro-duce 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. Growingcrops for fuel squanders land, water, and energy resources vitalfor the production of food for people [203]. In 2007, when US retailfood prices rose 4% above 2006 levels and twice as fast as overallcore inflation (2.3%), consumers took notice [204]. The bioetha-nol-driven surge in corn demand has fueled a sharp rise in cornprices. For example, the futures contract for March 2007 corn onthe Chicago Board of Trade, rose from US$2.50 per bushel in Sep-tember 2006 to a contract high of over US$4.16 per bushel in Jan-uary 2007 (a rise of 66%). This sharp rise in corn prices owes itsorigins largely to increasing corn demand spurred by the rapidexpansion of corn-based bioethanol production capacity in theUnited States since mid-2006 [205]. Higher corn prices were, inpart, driven by demand to make bioethanol and these higher priceseffectively bid acres away from other crops that provided lower re-turns, such as soybeans, wheat, and hay [204]. Using corn for bio-ethanol 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 consid-ering its great availability and low cost, but the large-scale com-mercial production of fuel bioethanol from lignocellulosicmaterials has still not been implemented. Today the productioncost of bioethanol from lignocellulose is still too high, which isthe major reason why bioethanol has not made its breakthroughyet. Pretreatment has been viewed as one of the most expensiveprocessing steps in cellulosic biomass to fermentable sugars con-version with costs as high as US$0.08/l bioethanol produced [86].Enzyme pricing is assumed such that the total contribution of en-zymes to production costs is about US$0.04/l of bioethanol withsome variation depending upon actual bioethanol yields resultingfrom the particular pretreatment approach [206]. Significantgrowth of the bioethanol industry will depend on the developmentof new processes that convert cellulosic biomass from non-foodcrops and waste materials into bioethanol [207].

7. Conclusion

Recently, there has been growing interest in biofuels due to therising energy costs and environmental problems. Bioethanol is byfar the most widely used biofuel for transportation worldwide. Itwill continue to be developed as a transport fuel produced in trop-ical latitudes and traded internationally, for use primarily as a gas-oline additive.

Page 15: 1-s2.0-S0196890410003791-main

872 M. Balat / Energy Conversion and Management 52 (2011) 858–875

Bioethanol production generally utilizes derivatives from foodcrops such as corn grain and sugarcane, but the limited supply ofthese crops can lead to competition between their use in bioetha-nol production and food provision. The price of the raw materials isalso highly volatile, which can highly affect the production costs ofthe bioethanol. Lignocellulosic materials serve as a cheap andabundant feedstock, which is required to produce fuel bioethanolfrom renewable resources at reasonable costs.

Lignocellulosic biomass can be converted to bioethanol byhydrolysis and subsequent fermentation. Lignocellulose is oftenhydrolyzed by acid treatment; the hydrolysate obtained is thenused 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, microorganismsshould be required to efficiently ferment these sugars for the suc-cessful industrial production of bioethanol.

Acknowledgement

The author would like to thank Professor Ayhan Demirbas� forhis very considerable help and encouragement throughout thecourse of this work.

References

[1] International Energy Agency (IEA). Key world energy statistics2008. 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 bio-ethanol fuel. Appl Energy 2009;86:2273–82.

[4] Plunkett JW. Plunkett’s automobile industry Almanac 2008: automobile, truckand specialty vehicle industry market research, statistics, trends & leadingcompanies. 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). Mobility2030: meeting the challenges to sustainability. The sustainable mobilityproject, Geneva, Switzerland; 2004. <www.wbcsd.ch/web/publications/mobility> [accessed October 2009].

[6] Meher LC, Sagar DV, Naik SN. Technical aspects of biodiesel production bytransesterification—a review. Renew Sustain Energy Rev 2006;10:248–68.

[7] Demirbas A. Biofuels sources, biofuel policy, biofuel economy and globalbiofuel projections. Energy Convers Manage 2008;49:2106–16.

[8] Forum for Agricultural Research in Africa (FARA). Bioenergy value chainresearch and development Stakes and Opportunities. Burkina Faso, FARADiscussion 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 developingcountries. Energy Convers Manage 2007;48:2386–98.

[10] Ertas M, Alma MH. Slow pyrolysis of chinaberry (Melia azedarach L.) seeds:Part I. The influence of pyrolysis parameters on the product yields. EnergyEduc Sci Technol Part A 2011;26:143–54.

[11] Balat M, Balat M, Kırtay E, Balat H. Main routes for the thermo-conversion ofbiomass into fuels and chemicals. Part 1: pyrolysis systems. Energy ConversManage 2009;50:3147–57.

[12] De Kam MJ, Morey RV, Tiffany DG. Biomass integrated gasification combinedcycle for heat and power at ethanol plants. Energy Convers Manage2009;50:1682–90.

[13] Balat M, Balat M, Kırtay E, Balat H. Main routes for the thermo-conversion ofbiomass into fuels and chemicals. Part 2: gasification systems. EnergyConvers Manage 2009;50:3158–68.

[14] Liu Z, Zhang FS. Effects of various solvents on the liquefaction of biomass toproduce fuels and chemical feedstocks. Energy Convers Manage2008;49:3498–504.

[15] Sener U, Genel Y, Saka C, Kilicel F, Kucuk MM. Supercritical fluid extraction ofcotton stalks. Energy Source A 2010;32:20–5.

[16] Demirbas A. Diesel-like fuel from tallow by pyrolysis and supercritical waterliquefaction. Energy Source A 2009;31:824–30.

[17] Mohanty SK, Behera S, Swain MR, Ray RC. Bioethanol production from mahula(Madhuca latifolia L.) flowers by solid-state fermentation. Appl Energy2009;86:640–4.

[18] Balat H. Prospects of biofuels for a sustainable energy future: A criticalassessment. Energy Educ Sci Technol Part A 2010;24:85–111.

[19] Demirbas A. Biofuels securing the planet’s future energy needs. EnergyConvers Manage 2009;50:2239–49.

[20] Demirbas A, Karslioglu S. Biodiesel production facilities from vegetable oilsand animal fats. Energy Source A 2007;29:133–41.

[21] Lang X, Macdonald DG, Hill GA. Recycle bioreactor for bioethanol productionfrom wheat starch II. Fermentation and economics. Energy Source2001;23:427–36.

[22] Demirbas MF, Balat M, Balat H. Potential contribution of biomass to thesustainable energy development. Energy Convers Manage 2009;50:1746–60.

[23] Börjesson P. Good or bad bioethanol from a greenhouse gas perspective–What determines this? Appl Energy 2009;86:589–94.

[24] Demirbas A. Producing and using bioethanol as an automotive fuel. EnergySource B 2007;2:391–401.

[25] Balat M. Bioethanol as a vehicular fuel: a critical review. Energy Source A2009;31:1242–55.

[26] Çelik MB. Experimental determination of suitable ethanol–gasoline blendrate at high compression ratio for gasoline engine. Appl Therm Eng2008;28:396–404.

[27] Demirbas A. Fuel alternatives to gasoline. Energy Source B 2007;2:311–20.[28] Stokes H. Alcohol fuels (ethanol and methanol): safety. In: Presentation at

ETHOS conference 2005, Seattle, Washington, January 29–30; 2005.[29] Pejin D, Mojovic LC, Vucurovic V, Pejin J, Dencic S, Rakin M. Fermentation of

wheat and triticale hydrolysates: a comparative study. Fuel 2009;88:1625–8.[30] Balat M. Global status of biomass energy use. Energy Source A 2009;31:

1160–73.[31] Philippidis G. The potential of biofuels in the Americas. Energy cooperation

and security in the hemisphere task force, center for hemispheric policy – TheUniversity 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 Res2005;64:809–21.

[33] Mojovic L, Pejin D, Grujic O, Pejin J, Rakin M, Vukasinovic M, et al. Progress inthe production of bioethanol on starch-based feedstock. Chem Ind Chem EngQ 2009;15:211–26.

[34] Gnansounou E, Bedniaguine D, Dauriat A. Promoting bioethanol productionthrough clean development mechanism: findings and lessons learnt fromASIATIC project. In: Proceedings of the 7th IAEE European energy conference,Bergen, Norway, August 28–30; 2005.

[35] Hartemink AE. Sugarcane for bioethanol: soil and environmental issues. AdvAgron 2008;99:125–82.

[36] Prabhakar SVRK, Elders M. CO2 Reduction potential of biofuels in Asia: issuesand policy implications. In: International conference on energy security andclimate change: issues, strategies, and options (ESCC 2008), Bangkok,Thailand, August 6–8; 2009.

[37] Trostle R. Global agricultural supply and demand: factors contributing to therecent 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 20–22;2006.

[39] Kline KL, Oladosu GA, Wolfe AK, Perlack RD, Dale VH, McMahon M. Biofuelfeedstock assessment for selected countries. Oak Ridge national laboratorytechnical 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 gasolinesubstitution in Thailand and its competitiveness. Thammasat Int J Sci Technol2007;12:69–80.

[41] Slade R, Bauen A, Shah N. The greenhouse gas emissions performance ofcellulosic ethanol supply chains in Europe. Biotechnol Biofuels 2009;2:1–19.

[42] Yao R, Qi B, Deng S, Liu N, Peng S, Cui Q. Use of surfactants in enzymatichydrolysis of rice straw and lactic acid production from rice straw bysimultaneous saccharification and fermentation. BioResources 2007;2:389–98.

[43] Wang L, Hanna MA, Weller CL, Jones DD. Technical and economical analysesof combined heat and power generation from distillers grains and corn stoverin ethanol plants. Energy Convers Manage 2009;50:1704–13.

[44] Keshwani DR, Cheng JJ. Switchgrass for bioethanol and other value-addedapplications: a review. Bioresour Technol 2009;100:1515–23.

[45] Carvalho LG, Gomes AM, Aranda DAG, Pereira N. Ethanol from lignocellulosicresidues of palm oil industry. In: 11th International conference on advancedmaterials, Rio de Janeiro, Brazil, September 20–25; 2009.

[46] Bohlmann GM. Process economic considerations for production of ethanolfrom biomass feedstocks. Ind Biotechnol 2006;2:14–20.

[47] Karimi K, Emtiazi G, Taherzadeh MJ. Ethanol production from dilute-acidpretreated rice straw by simultaneous saccharification and fermentation withMucor indicus, Rhizopus oryzae, and Saccharomyces cerevisiae. EnzymeMicrobiol Technol 2006;40:138–44.

[48] Lee D, Owens VN, Boe A, Jeranyama P. Composition of herbaceous biomassfeedstocks. South Dakota State University Publication, SGINC1-07, Brookings,SD; June 2007.

[49] Molina-Sabio M, Rodríguez-Reinoso F. Role of chemical activation in thedevelopment of carbon porosity. Colloid Surface Physicochem Eng Aspect2004;241:15–25.

[50] Jenkins BM, Baxter LL, Miles Jr TR, Miles TR. Combustion properties ofbiomass. Fuel Process Technol 1998;54:17–46.

[51] Dehkhoda A. Concentrating lignocellulosic hydrolysate by evaporation and itsfermentation by repeated fedbatch using flocculating Saccharomycescerevisiae. Master thesis, Industrial Biotechnology Boras University andSEKAB E-Technology, Sweden; 2008.

Page 16: 1-s2.0-S0196890410003791-main

M. Balat / Energy Conversion and Management 52 (2011) 858–875 873

[52] Balat M. Gasification of biomass to produce gaseous products. Energy SourceA 2009;31:516–26.

[53] Demirbas A. Bioethanol from cellulosic materials: a renewable motor fuelfrom biomass. Energy Source 2005;27:327–37.

[54] Karunanithy C, Muthukumarappan K, Julson JL. Enzymatic hydrolysis of cornstover pretreated in high shear bioreactor. In: ASABE Annual internationalmeeting, Paper No. 084114, Rhode Island, June 29–July 2; 2008.

[55] Hamelinck CN, Van Hooijdonk G, Faaij APC. Ethanol from lignocellulosicbiomass: techno-economic performance in short-, middle- and long-term.Biomass Bioenergy 2005;28:384–410.

[56] Taherzadeh MJ, Karimi K. Pretreatment of lignocellulosic wastes to improveethanol and biogas production: a review. Int J Mol Sci 2008;9:1621–51.

[57] Demirbas A. Heavy metal adsorption onto agro-based waste materials: areview. J Hazard Mater 2008;157:220–9.

[58] Feldman D, Banu D, Natansohn A, Wang J. Structure–properties relations ofthermally cured epoxy-lignin polyblends. J Appl Polym Sci 1991;42:1537–50.

[59] Demirbas A. Products from lignocellulosic materials via degradationprocesses. Energy Source A 2008;30:27–37.

[60] Yaman S. Pyrolysis of biomass to produce fuels and chemical feedstocks.Energy Convers Manage 2004;45:651–71.

[61] Cardona Alzate CA, Sanchez Toro OJ. Energy consumption analysis ofintegrated flow sheets for production of fuel ethanol from ligno-cellulosicbiomass. Energy 2006;31:2447–59.

[62] Zhu JY, Wang GS, Pan XJ, Gleisner R. The status of and key barriers inlignocellulosic ethanol production: a technological perspective. In:International conference on biomass energy technologies, Guangzhou,China, December 3–5; 2008.

[63] Hsu TA, Ladisch MR, Tsao GT. Alcohol from cellulose. Chem Technol1980;10:315–9.

[64] Kodali B, Pogaku R. Pretreatment studies of rice bran for the effectiveproduction of cellulose. Electron J Environ Agric Food Chem 2006;5:1253–64.

[65] Zhang Y, Pan Z, Zhang R. Overview of biomass pretreatment for cellulosicethanol production. Int J Agric Biol Eng 2009;2:51–68.

[66] Chandel AK, Es C, Rudravaram R, Narasu ML, Rao LV, Ravindra P. Economicsand environmental impact of bioethanol production technologies: anappraisal. Biotechnol Mol Biol Rev 2007;2:14–32.

[67] Taherzadeh MJ, Karimi K. Enzymatic-based hydrolysis processes for ethanolfrom lignocellulosic materials: a review. BioResources 2007;2:707–38.

[68] Kumar P, Barrett DM, Delwiche MJ, Stroeve P. Methods for pretreatment oflignocellulosic biomass for efficient hydrolysis and biofuel production. IndEng Chem Res 2009;48:3713–29.

[69] Chen M, Zhao J, Xia L. Comparison of four different chemical pretreatments ofcorn stover for enhancing enzymatic digestibility. Biomass Bioenergy2009;33:1381–5.

[70] Sun F, Chen H. Enhanced enzymatic hydrolysis of wheat straw by aqueousglycerol pretreatment. Bioresour Technol 2008;99:6156–61.

[71] Esteghlalian A, Hashimoto AG, Fenske JJ, Penner MH. Modeling andoptimization of the dilute-sulfuric-acid pretreatment of corn stover, poplarand switchgrass. Bioresour Technol 1997;59:129–36.

[72] Zhang Q, Cai W. Enzymatic hydrolysis of alkali-pretreated rice straw byTrichoderma reesei ZM4-F3. Biomass Bioenergy 2008;32:1130–5.

[73] Rabelo SC, Filho RM, Costa AC. Lime pretreatment of sugarcane bagasse forbioethanol production. Appl Biochem Biotechnol 2009;153:139–53.

[74] Sun Y, Cheng J. Hydrolysis of lignocellulosic materials for ethanol production:a review. Bioresour Technol 2002;83:1–11.

[75] Leustean I. Bioethanol from lignocellulosic materials. J AgroalimentaryProcess Technol 2009;15:94–101.

[76] Zheng Y, Pan Z, Zhang R. Overview of biomass pretreatment for cellulosicethanol production. Int J Agric Biol Eng 2009;2:51–68.

[77] Cadoche L, Lopez GD. Assessment of size reduction as a preliminary step inthe production of ethanol from lignocellulosic wastes. Biol Wastes1989;30:153–7.

[78] McMillan JD. Pretreatment of lignocellulosic Biomass. In: Himmel ME, BakerJO, Overend RP, editors. Enzymatic conversion of biomass for fuelsproduction, ACS Symposium Series 566. American Chemical Society,Washington, DC; 1994. p. 292–324.

[79] Ruffell J. Pretreatment and hydrolysis of recovered fibre for ethanolproduction. 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-scalebioethanol production from lignocellulose: a review. J Sci Ind Res2008;67:874–84.

[82] Pasha C, Rao LV. Thermotolerant yeasts for bioethanol production usinglignocellulosic substrates. In: Satyanarayana T, Kunze G, editors. Yeastbiotechnology: diversity and applications. Netherlands: Springer; 2009. p.551–88.

[83] Jeoh T. Steam explosion pretreatment of cotton gin waste for fuel ethanolproduction. Master’s thesis (Agblevor FA, Chen JS, Helm RF advs.), VirginiaTech 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;50:1–39.

[85] Heerah AS, Mudhoo A, Mohee R, Sharma SK. Steam pre-treatment oflignocellulosic wastes for biomethanogenesis: a preliminary study. RasayanJ Chem 2008;1:503–14.

[86] Mosier N, Wyman C, Dale B, Elander R, Holtzapple YYLM, Ladisch M. Featuresof promising technologies for pretreatment of lignocellulosic biomass.Bioresour Technol 2005;96:673–86.

[87] Negro MJ, Manzanares P, Ballesteros I, Oliva JM, Cabanas A, Ballesteros M.Hydrothermal pretreatment conditions to enhance ethanol production frompoplar biomass. Appl Biochem Biotechnol 2003;105:87–100.

[88] Bertilsson M. Simultaneous saccharification and fermentation of spruce – acomparison 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. ApplBiochem Biotechnol 2006;130:496–508.

[90] Sassner P, Martensson CG, Galbe M, Zacchi G. Steam pretreatment of H2SO4-impregnated Salix for the production of bioethanol. Bioresour Technol2008;99:137–45.

[91] Galbe M, Zacchi G. A review of the production of ethanol from softwood. ApplMicrobiol Biotechnol 2002;59:618–28.

[92] Söderström J, Pilcher L, Galbe M, Zacchi G. Two-step steam pretreatment ofsoftwood by dilute H2SO4 impregnation for ethanol production. BiomassBioenergy 2003;24:475–86.

[93] Söderström J, Galbe M, Zacchi G. Effect of washing on yield in one- and two-step steam pretreatment of softwood for production of ethanol. BiotechnolProg 2008;20:744–9.

[94] Abril D, Abril A. Ethanol from lignocellulosic biomass. Cien Inv Agr2009;36:177–90.

[95] Dale MC, Moelhman M. Enzymatic simultaneous saccharification andfermentation (SSF) of biomass to ethanol in a pilot 130 liter multistagecontinuous reactor separator. In: Ninth biennial bioenergy conference,Buffalo, New York, October 15–19; 2000.

[96] Alizadeh H, Teymouri F, Gilbert TI, Dale BE. Pretreatment of switchgrass byammonia fiber explosion (AFEX). Appl Biochem Biotechnol 2005;121:1133–41.

[97] Teymouri F, Laureano-Pérez L, Alizadeh H, Dale BE. Ammonia fiber explosiontreatment of corn stover. Appl Biochem Biotechnol 2004;115:951–63.

[98] Holtzapple MT, Jun J-H, Ashok G, Patibandla SL, Dale BE. The ammonia freezeexplosion (AFEX) process: a practical lignocellulose pretreatment. ApplBiochem Biotechnol 1991;28(/29):59–74.

[99] Balan V, Sousa LC, Chundawat SPS, Marshall D, Sharma LN, Chambliss CK,et al. Enzymatic diagestibility and pretreatment degradation products ofAFEX-treated hardwoods (Populus nigra). Biotechnol Prog 2009;25:365–75.

[100] Lee YJ. Oxidation of sugarcane bagasse using a combination of hypochlorite andperoxide. Master’s thesis (Day DF, Godber JS, Janes ME, advs), Department ofFood Science, Graduate Faculty of the Louisiana State University andAgricultural and Mechanical College, October 28; 2005.

[101] Hu G, Heitmann JA, Rojas OJ. Feedstock pretreatment strategies for producingethanol from wood, bark and forest residues. BioResources 2008;3:270–94.

[102] Hayes DJ. An examination of biorefining processes, catalysts and challenges.Catal Today 2009;145:138–51.

[103] Vidal PF, Molinier J. Ozonolysis of lignin – improvement of in vitrodigestibility of poplar sawdust. Biomass 1988;16:1–17.

[104] Neeley WC. Factor affecting the pretreatment of bio-mass with gaseousozone. Biotechnol Bioeng 1984;26:59–65.

[105] Williams KC. Subcritical water and chemical pretreatments of cotton stalk forthe production of ethanol. Master’s thesis, Biological and AgriculturalEngineering, North Carolina State University, Raleigh, NC, USA; 2006.

[106] García-Cubero MT, Gonzalez-Benito G, Indacoechea I, Coca M, Bolado S. Effectof ozonolysis pretreatment on enzymatic digestibility of wheat and ryestraw. Bioresour Technol 2009;100:1608–13.

[107] Silverstein RA, Chen Y, Sharma-Shivappa RR, Boyette MD, Osborne J. Acomparison of chemical pretreatment methods for improvingsaccharification of cotton stalks. Bioresour Technol 2008;98:3000–11.

[108] Han M, Moon SK, Kim Y, Kim Y, Chung B, Choi GW. Bioethanol productionfrom ammonia percolated wheat straw. Biotechnol Bioprocess Eng2009;14:606–11.

[109] Fang LT, Gharpuray MM, Lee YH. Cellulose hydrolysis biotechnologymonographs. Berlin, Germany: Springer; 1987. p. 55.

[110] Millet MA, Baker AJ, Scatter LD. Physical and chemical pretreatment forenhancing cellulose. Appl Microbiol Biotechnol 1976;29:462–8.

[111] Akhtar MS, Saleem M, Akhtar MW. Saccharification of lignocellulosicmaterials by the cellulases of Bacillus subtilis. Int J Agr Biol 2001;3:199–202.

[112] Li Y, Ruan R, Chen PL, Liu Z, Pan X, Lin X, et al. Enzymatic hydrolysis of cornstover pretreated by combined dilute alkaline treatment andhomogenization. Trans ASAE 2004;47:821–5.

[113] Carvalheiro F, Duarte LC, Girio FM. Hemicellulose biorefineries: a review onbiomass pretreatments. J Sci Ind Res 2008;67:849–64.

[114] Wang Z, Cheng JJ. Lime pretreatment of coastal bermudagrass for bioethanolproduction. In: ASABE meeting, Reno, Nevada, June 21–24; 2009.

[115] Balat M, Balat H, Oz C. Progress in bioethanol processing. Progr EnergyCombust Sci 2008;34:551–73.

[116] Ramirez RS. Long-term lime pretreatment of poplar wood. Master’s thesis,Texas A&M University; December 2005.

[117] Chang VS, Burr B, Holtzapple MT. Lime pretreatment of switchgrass. ApplBiochem Biotechnol 1997;63–65:3–19.

[118] Saha BC, Cotta MA. Enzymatic hydrolysis and fermentation of lime pretreatedwheat straw to ethanol. J Chem Tech Biotechnol 2007;82:913–9.

Page 17: 1-s2.0-S0196890410003791-main

874 M. Balat / Energy Conversion and Management 52 (2011) 858–875

[119] Karr WE, Holtzapple MT. Using lime pretreatment to facilitate the enzymatichydrolysis of corn stover. Biomass Bioenergy 2000;18:189–99.

[120] Chang VS, Nagwani M, Kim CH, Holtzapple MT. Oxidative lime pretreatmentof high-lignin biomass. Appl Biochem Biotechnol 2001;94:1–28.

[121] Silverstein RA. A comparison of chemical pretreatment methods forconverting cotton stalks to ethanol. Master’s 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 andmoisture on dilute-acid steam explosion pretreatment of corn stover andcellulase enzyme digestibility. Appl Biochem Biotechnol 2003;105:165–78.

[123] Jeffries TW, Jin YS. Ethanol and thermotolerance in the bioconversion ofxylose by yeasts. Adv Appl Microbiol 2000;47:221–68.

[124] Torget R, Hatzis C, Hayward TK, Hsu TA, Philippidis GP. Optimization ofreverse-flow, 2-temperature, dilute-acid pretreatment to enhance biomassconversion to ethanol. Appl Biochem Biotechnol 1996;58:85–101.

[125] Lee YY, Iyer P, Torget RW. Dilute-acid hydrolysis of lignocellulosic biomass.Adv Biochem Eng Biotechnol 1999;65:93–115.

[126] Lee JW, Gwak KS, Park JY, Park MJ, Choi DH, Kwon M, et al. Biologicalpretreatment of softwood Pinus densiflora by three white rot fungi. JMicrobiol 2007;45:485–91.

[127] Taherzadeh MJ, Karimi K. Acid-based hydrolysis processes for ethanol fromlignocellulosic materials: a review. BioResources 2007;2:472–99.

[128] Gamez S, Ramirez JA, Garrote G, Vazquez M. Manufacture of fermentablesugar solutions from sugar cane bagasse hydrolyzed with phosphoric acid atatmospheric pressure. J Agric Food Chem 2004;52:4172–7.

[129] Gullu DE. Effect of catalyst on yield of liquid products from biomass viapyrolysis. Energy Source 2003;25:753–65.

[130] Dunlop AP. Furfural formation and behaviour. Ind Eng Chem 1948;40:204–9.[131] Ulbricht RJ, Sharon J, Thomas J. A review of 5-hydroxymethylfurfura HMF in

parental solutions. Fundam Appl Toxicol 1984;4:843–53.[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 changeinitiative, 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 criticalstep for production of ethanol from mixed wood chips. Biomass Bioenergy2002;22:401–4.

[135] Demirbas A. The importance of bioethanol and biodiesel from biomass.Energy Source B 2008;3:177–85.

[136] Yu Y, Lou X, Wu H. Some recent advances in hydrolysis of biomass in hot-compressed water and its comparisons with other hydrolysis methods.Energy Fuels 2008;22:46–60.

[137] Pike PW, Sengupta D, Hertwig TA. Integrating biomass feedstocks intochemical production complexes using new and existing processes. MineralsProcessing 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 ofcellulosic substrates. Holzforschung 2006;60:398–401.

[139] Mani S, Tabil LG, Opoku A. Ethanol from agricultural crop residues – anoverview. In: ASAE/CSAE north-central intersectional meeting, Saskatoon,Saskatchewan, Canada, September 27–28; 2002.

[140] Zhou J, Wang YH, Chu J, Zhuang YP, Zhang SL, Yin P. Identification andpurification of the main components of cellulases from a mutant strain ofTrichoderma viride T 100-14. Bioresour Technol 2008;99:6826–33.

[141] Viksø-Nielsen A. Recent development in enzymes for biomass hydrolysis. In:Maj M, Kirsch N, editors. First European workshop on biotechnology forlignocellulose biorefineries. Forest & Landscape Denmark, University ofCopenhagen, Copenhagen, March 27–28; 2008.

[142] Kim SH. Lime pretreatment and enzymatic hydrolysis of corn stover. Doctoraldissertation, Texas A&M University; May 2004.

[143] Zhang YHP, Himmel ME, Mielenz JR. Outlook for cellulase improvement:screening and selection strategies. Biotechnol Adv 2006;24:452–81.

[144] Detroy RW, Julian GS. Biomass conversion: fermentation chemicals and fuels.Crit Rev Microbiol 1982;10:203–28.

[145] Gregg DJ, Saddler JN. Factors affecting cellulose hydrolysis and the potentialof enzyme recycle to enhance the efficiency of an integrated wood to ethanolprocess. Biotechnol Bioeng 1996;51:375–83.

[146] Chen M, Xia L, Xue P. Enzymatic hydrolysis of corncob and ethanol productionfrom cellulosic hydrolysate. Int Biodeterior Biodegrad 2007;59:85–9.

[147] Rezaei F, Richard TL, Logan BE. Enzymatic hydrolysis of cellulose coupledwith electricity generation in a microbial fuel cell. Biotechnol Bioeng2008;101:1163–9.

[148] Silva CHC, Puls J, De Sousa MV, Filho EXF. Purification and characterization ofa low molecular weight xylanase from solid-state cultures of Aspergillusfumigatus Fresenius. Rev Microbiol 1999;30:114–9.

[149] Wong KKY, Saddler JN. Trichoderma xylanases, their properties andapplications. In: Visser J, Beldman G, Kusters-van Someren M, Voragen AGJ,editors. Xylans and xylanases. Amsterdam: Elsevier; 1992. p. 171–86.

[150] Haltrich D, Nidetzky B, Kulbe KD, Steiner W, Zupancic S. Production of fungalxylanases. Bioresour Technol 1996;58:137–61.

[151] Filho EXF, Tuohy MG, Puls J, Coughlan MP. The xylan-degrading enzyme-systems of Penicillium capsulatum and Talaromyces emersonii. Biochem SocTrans 1991;19:S25.

[152] Jorgensen H, Eriksson T, Borjesson J, Tjerneld F, Olsson L. Purification andcharacterization of five cellulases and one xylanase from Penicilliumbrasilianum IBT 20888. Enzyme Microb Technol 2003;32:851–61.

[153] Tuohy MG, Puls J, Claeyssens M, Vrsanska M, Coughlan MP. The xylan-degrading enzyme system of Talaromyces emersonii: novel enzymes withactivity against aryl beta-D-xylosides and unsubstituted xylans. Biochem J1993;290:515–23.

[154] Dos Reis S, Costa MAF, Peralta RM. Xylanase production by a wild strain ofAspergillus nidulans. Acta Sci: Biol Sci 2003;1:221–5.

[155] Virupakshi S, Babu KG, Gaikwad SR, Naik GR. Production of a xylanolyticenzyme by a thermoalkaliphilic Bacillus sp. JB-99 in solid state fermentation.Process Biochem 2005;40:431–5.

[156] Saha BC. Lignocellulose biodegradation and applications in biotechnology.ACS Symp Ser 2004;889:2–34.

[157] Saha BC. Hemicellulose bioconversion. J Ind Microbiol Biotechnol2003;30:279–91.

[158] Howard RL, Abotsi E, Jansen van Rensburg EL, Howard S. Lignocellulosebiotechnology: issues of bioconversion and enzyme production. Afr JBiotechnol 2003;2:602–19.

[159] Szulczyk KR, McCarl BA, Cornforth G. Market penetration of ethanol. RenewSustain Energy Rev 2010;14:394–403.

[160] Dien BS, Cotta MA, Jeffries TW. Bacteria engineered for fuel ethanolproduction: current status. Appl Microbiol Biotechnol 2003;63:258–66.

[161] Katahira S, Mizuike A, Fukuda H, Kondo A. Ethanol fermentation fromlignocellulosic hydrolysate by a recombinant xylose- andcellooligosaccharide-assimilating yeast strain. Appl Microbiol Biotechnol2006;72:1136–43.

[162] Zaldivar J, Nielsen J, Olsson L. Fuel ethanol production from lignocellulose: achallenge for metabolic engineering and process integration. Appl MicrobiolBiotechnol 2001;56:17–34.

[163] Tian S, Zang J, Pan Y, Liu J, Yuan Z, Yan Y, et al. Construction of a recombinantyeast strain converting xylose and glucose to ethanol. Front Biol China2008;3:165–9.

[164] Shi XQ, Jeffries TW. Anaerobic growth and improved fermentation of Pichiastipitis bearing a URA1 gene from Saccharomyces cerevisiae. Appl MicrobiolBiotechnol 1998;50:339–45.

[165] Hahn-Hagerdal B, Galbe M, Gorwa-Grauslund MF, Liden G, Zacchi G. Bio-ethanol – the fuel of tomorrow from the residues of today. Trends Biotechnol2006;24:549–56.

[166] Gunasekaran P, Raj KC. Ethanol fermentation technology – Zymomonasmobilis. Curr Sci 1999;77:56–68.

[167] Saez-Miranda JS, Saliceti-Piazza L, McMillan JD. Measurement and analysis ofintracellular ATP levels in metabolically engineered Zymomonas mobilisfermenting glucose and xylose mixtures. Biotechnol Prog 2008;22:359–68.

[168] Millichip RJ, Doelle HW. Large-scale ethanol production from Milo (Sorghum)using Zymomonas mobilis. Process Biochem 1989;24:141–5.

[169] Lin Y, Tanaka S. Ethanol fermentation from biomass resources: current stateand prospects. Appl Microbiol Biotechnol 2006;69:627–42.

[170] Avci A, Donmez S. Effect of zinc on ethanol production by twoThermoanaerobacter strains. Process Biochem 2006;41:984–9.

[171] Cook GM, Morgan HW. Hyperbolic growth of Thermoanaerobacterthermohydrosulfuricus (Clostridium thermohydrosulfuricum) increases ethanolproduction in pH-controlled batch culture. Appl Microbiol Biotechnol1994;41:84–9.

[172] Larsen L, Nielsen P, Ahring BK. Thermoanaerobacter mathranii sp. Nov., anethanol-producing, extremely thermophilic anaerobic bacterium from a hotspring in Iceland. Arch Microbiol 1997;168:114–9.

[173] Lamed R, Zeikus JG. Glucose fermentation pathway of Thermoanaerobiumbrockii. J Bacteriol 1980;14:1251–7.

[174] Baskaran S, Ahn HJ, Lynd LR. Investigation of the ethanol tolerance ofClostridium thermosaccharolyticum in continuous culture. Biotechnol Prog1995;11:276–81.

[175] Knutson BL, Strobel HJ, Nokes SE, Dawson KA, Berberich JA, Jones CR. Effect ofpressurized solvents on ethanol production by the thermophilic bacteriumClostridium thermocellum. J Supercritical Fluids 1999;16:149–56.

[176] Georgieva TI, Skiadas IV, Ahring BK. Effect of temperature on ethanoltolerance of a thermophilic anaerobic ethanol producer ThermoanaerobacterA10: modeling and simulation. Biotechnol Bioeng 2007;98:1161–70.

[177] Abtahi Z. Ethanol and glucose tolerance of M. indicus in aerobic and anaerobicconditions. 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 EnvironmentalEngineering, University of Oulu, Report A No. 21, Finland; June 2003.

[179] Frison A, Memmert K. Fed-batch process development for monoclonalantibody production with cellferm-pro. Gen Eng News 2002;22:66–7.

[180] Gunther JC, Seborg DE, Baclaski J. Fault detection and diagnosis in industrialfed-batch fermentation. In: 2006 American control conference, Minneapolis,MN, June 14–16; 2006.

[181] Caylak B, Sukan FV. Comparison of different production processes forbioethanol. Turk J Chem 1998;22:351–9.

[182] Lawford HG. A new approach to improving the performance of Zymomonas incontinuous ethanol fermentations. Appl Biochem Biotechnol 1988;17:203–19.

[183] Tengborg C, Galbe M, Zacchi G. Influence of enzyme loading and physicalparameters on the enzymatic hydrolysis of steam-pretreated softwood.Biotechnol Prog 2008;17:110–7.

Page 18: 1-s2.0-S0196890410003791-main

M. Balat / Energy Conversion and Management 52 (2011) 858–875 875

[184] Bertilsson M, Olofsson K, Liden G. Prefermentation improves xyloseutilization in simultaneous saccharification and co-fermentation ofpretreated spruce. Biotechnol Biofuels 2009;2:8.

[185] Olofsson K, Bertilsson M, Liden G. A short review on SSF – an interestingprocess option for ethanol production from lignocellulosic feedstocks.Biotechnol Biofuels 2008;1:7.

[186] Ballesteros M, Oliva JM, Negro MJ, Manzanares P, Ballesteros I. Ethanol fromlignocellulosic materials by a simultaneous saccharification and fermentationprocess (SFS) with Kluyveromyces marxianus CECT 10875. Process Biochem2004;39:1843–8.

[187] Huang LP, Jin B, Lant P, Zhou J. Simultaneous saccharification andfermentation of potato starch wastewater to lactic acid by Rhizopus oryzaeand Rhizopus arrhizus. Biochem Eng J 2005;23:265–76.

[188] Schell DJ, Ruth MF, Tucker MP. Modeling the enzymatic hydrolysis of dilute-acid pretreated douglas Fir. In: 20th Symposium on biotechnology for fuelsand chemicals, Gatlinburg, TN, May 3–7; 1998.

[189] Krishna SH, Reddy TJ, Chowdary GV. Simultaneous saccharification andfermentation of lignocellulosic wastes to ethanol using a thermotolerantyeast. Bioresour Technol 2001;77:193–6.

[190] Madson PW, Lococo DB. Recovery of volatile products from dilute high-fouling process streams. Appl Biochem Biotechnol 2000;84–86:1049–61.

[191] Kwiatkowski JR, McAloon AJ, Taylor F, Johnston DB. Modeling the process andcosts of fuel ethanol production by the corn dry-grind process. Ind Crop Prod2006;23:288–96.

[192] Karuppiah R, Peschel A, Grossmann IE, Martín M, Martinson W, Zullo L.Energy optimization for the design of corn-based ethanol plants. AIChE J2008;54:1499–525.

[193] McAloon A, Taylor F, Yee W, Ibsen K, Wooley R. Determining the cost ofproducing ethanol from corn starch and lignocellulosic feedstocks. NationalRenewable Energy Laboratory, Technical Report, NREL/TP-580-28893,Golden, Colorado; October 2000.

[194] De Fraiture C, Giordano M, Liao Y. Biofuels and implications for agriculturalwater use: blue impacts of green energy. Water Policy 2008;10:67–81.

[195] Elam N. Alternative fuels (ethanol) in Sweden. Investigation and evaluationfor IEA Bioenergy, Task 27, Atrax Energi AB, Göteborg, Sweden; 2000.

[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 December2009].

[197] Shapouri H, Salassi M, Nelson J. The economic feasibility of ethanolproduction 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 developingcountries. 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 prospectsgroup. 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 ofthe genes required for tolerance to vanillin, which is a potential inhibitor ofbioethanol fermentation, in Saccharomyces cerevisiae. Biotechnol Biofuels2008;1:1–6.

[203] Pimentel D, Marklein A, Toth MA, Karpoff M, Paul GS, McCormack R, et al.Food versus biofuels: environmental and economic costs. Human Ecol2009;37:1–12.

[204] Tiffany DG. Economic and environmental impacts of US corn ethanolproduction and use, vol. 5. Federal Reserve Bank of St. Louis RegionalEconomic Development; 2009. p. 42–58.

[205] Yacobucci BD, Schnepf R. Ethanol and biofuels: agriculture, infrastructure,and market constraints related to expanded production. CRS report forcongress, Order Code RL33928, March 16; 2007.

[206] Eggeman T, Elander RT. Process and economic analysis of pretreatmenttechnologies. Bioresour Technol 2005;96:2019–25.

[207] Woodson M, Jablonowski CJ. An economic assessment of traditional andcellulosic ethanol technologies. Energy Source B 2008;3:372–83.