Open access

Production of Biofuels from Cellulose of Woody Biomass

Written By

Pedram Fatehi

Submitted: 03 November 2011 Published: 29 August 2013

DOI: 10.5772/50740

From the Edited Volume

Cellulose - Biomass Conversion

Edited by Theo van de Ven and John Kadla

Chapter metrics overview

4,047 Chapter Downloads

View Full Metrics

1. Introduction

Today, fossil resources supply approximately 86% of energy and 96% of organic chemicals used in the world [1]. The continuous depletion of petroleum fuel is one of the prime concerns these days. Other concerns associated with the large-scale utilization of fossil fuels are availability, global warming and uneven geographic distribution [2,3]. Also, the global population is expected to increase by approximately 3 billion people by 2050, which substantially increases the need for fuels. One estimate indicated that the world energy consumption would increase by 35% over the next 20 years in order to meet the growing demand of industrialized countries and the rapid development of emerging economies [2]. As the fossil fuels are depleting in the coming years, new technologies should be developed in order to produce fuels from renewable biomass resources [4].

Currently, biomass accounts for 9.8% of the world’s primary energy use annually, among which 30% is used in modern forms (e.g. liquid biofuel and steam), and 70% is used traditionally (combustion for domestic heating with the energy density of 15-20 MJ/kg) [2,4,5]. Biofuel is a type of fuel whose energy is derived from biomass. It includes solid biofuels such as wood pellets, wood cube and wood puck; liquid biofuels such as ethanol and butanol; and gaseous biofuel such as hydrogen. The world liquid biofuel production would increase from 1.9 million of barrels per day in 2010 to 5.9 million barrel per day in 2030 [2]. In the United States, expectation for the production of biofuels is 136 billion litter mandated by 2022, of which 61 million liter are made from cellulosic materials [6].

As described earlier, biomass has directly been used for producing heat and electricity. The power generation engines were designed so that they directly used biomass as energy source during World War II. However, this direct utilization has several major problems: if used as biofuel, the uneven geographical distribution of biomass necessitate its transportation, the bulky nature of biomass (low heat value) leads to costly and complicated transportation systems; engines utilizing biomass usually possess a poor efficiency and high environmental impact (e.g. CO2 emission). However, according to the first law of thermodynamics, any biological or chemical conversion of biomass to biofuel will consume energy, thus a part of the energy stored in biomass will be lost during the conversion and the biofuel (product) will ultimately have a lower energy than will biomass (raw material). There are several advantages for the conversion of biomass to biofuel that outweighs this deficiency: there is a large demand for liquid fuel (e.g. ethanol) in the transportation sector; the biofuel production process is more environmentally friendly (i.e. less CO2 emission); the digested materials from biorefinery can be readily used as an excellent and sustainable fertilizer for cultivation and crops (a true recyclable process); the energy density (MJ/kg) of biofuel will be higher than that of biomass, and the common problem of combustion, e.g. fly ash disposal and super heater corrosion, can be eliminated via biofuel production [2,7,8].

One of the challenges of biofuel production is the difficulties in increasing the bulk density of the resource, while preserving its energy content [9]. The future biofuel should 1) have a high energy density on a mass; 2) be produced at yields near the stoichiometric maximum from a given biomass feed; 3) be compatible with existing fuel distribution infrastructures; 4) have a minimum impact on environment; and 5) not affect the global food supplies [9-11].

The first generation biofuels are presently produced from sugars, starches and vegetable oils, but these products have several issues: 1) their availability is limited by soil fertility and per-hectare yield; and 2) their contribution to savings of CO2 emissions and fossil energy consumption are limited by high energy input for their cultivations and conversions [12-14]. However, lignocellulosic biomass seems to be more attractive because 1) it is the most widespread renewable source available on earth (overall chemical energy stored in biomass by plants is approximately 6-7 times of total human energy consumption annually [15]); 2) it is locally available in many countries; and more importantly 3) it does not compete with food or food industries [12]. However, the conversion of woody biomass to fermentable sugars is more difficult than that of agro-based biomass because of the presence of more hemicelluloses (not easily fermentable) and lignin as well as more condensed and crystallized structure of cellulose in woody biomass [9].

Current technologies to produce biofuel from cellulosic materials involve gasification, pyrolysis/liquefaction and hydrolysis of biomass. This book chapter excludes 1) the studies on the gasification and pyrolysis/liquefaction of biomass (as an intact raw material) for biofuel production; and 2) the studies on the production of fuel additives, e.g. levulinic acid and furfural [16] and the studies on the production of biodiesel [17-22]. Instead, recent advancements and challenges associated with the production of ethanol, butanol, hydrogen and new furan-based biofuel from cellulosic biomass will be discussed.

In order to produce biofuel from cellulose of biomass, the lignocellulosic biomass should be first dissembled to facilitate the isolation of cellulose from other constituents, i.e. lignin and hemicelluloses. Subsequently, cellulose macromolecules should be depolymerized, as depolymerization significantly improves the chemical and biological conversions of cellulose to biofuel. Then, the depolymerized cellulose, i.e. glucose, should be converted to biofuel via biological treatments, and finally the biofuel should be purified. Additionally, biomass should be deoxygenated as the presence of oxygen reduces the heat content of molecules and creates high polarity, which impairs its blending with existing fossil fuels [12].

Advertisement

2. Pretreatment of biomass

The complex structure of lignocellulosic biomass makes its utilization in biofuel production difficult. Lignocellulosic raw materials are generally composed of 40-50% cellulose, 25-35% hemicelluloses and 15-20% lignin [8]. A pretreatment stage is necessary to dissociate the plant cell wall in order to improve the accessibility of chemicals and/or microorganisms to cellulose for possible conversions [23]. The pretreatment processes target the removal of lignin, which improves the digestibility of cellulose in the following hydrolysis process [24]. Table 1 lists various pretreatment processes of woody biomass conducted in the past in order to improve the performance of fermentation processes in producing biofuels.

Pretreatment typeExample
Physical pretreatmentBall milling, Irradiation
Physicochemical pretreatmentSteam explosion, hot water pretreatment
Chemical pretreatmentAcid, alkali, solvent
BiologicalFungi
EnzymeCellulase

Table 1.

Various pretreatment processes of lignocellulosic feedstock [14]

Physical pretreatment consists of mechanical disruption of lignocelluloses, which is an environmentally friendly process. This process increases the surface area of biomass and decreases the crystallinity of cellulose, but it does not cause an expensive mass loss [25]. Irradiation using gamma rays, electron rays and microwaves are other physical methods to beak the structure of lignocelluloses. Microwave irradiation has been applied in many fields including food drying chemical synthesis and extraction [14].

Physicochemical pretreatment is another approach to separate the lignocelluloses of woody materials. Hydrothermal treatment, such as hot water pretreatment and steam explosion, is a suitable method particularly prior to enzyme hydrolysis. Hot water pretreatment process is conducted under pressure at an elevated temperature of 230-240 °C for 15 min to maintain water in liquid form, which produces less inhibitory compounds (e.g. furfural) compared to steam explosion method [26,27]. However, the viscosity of the spent liquor produced in this method is rather high, which makes its handling process challenging. The steam explosion is practically applied in industry via steaming biomass at an elevated temperature, e.g. 170 °C [28-30]. To limit the production of inhibitors, process conditions should be precisely adjusted. Steam explosion has different subcategories, such as ammonia fiber explosion and acid-explosion, in which acid or ammonia is also added to the system during the steaming process.

The chemical pretreatment of cellulosic biomass includes oxidizing hydrolysis, acid or alkaline hydrolysis and solvent extraction. In this context, lime treatment (e.g. treatment of woody biomass at 180 °C with lime solutions) has been considered as an effective chemical pretreatment method, because of its low cost and wide use in agro- and wood-based processes [4,27,31-34]. Pretreatment with dilute acid at intermediate temperatures (e.g. 160 °C) is usually considered the most cost-effective method to loosen the cell wall matrix via degrading the hemicelluloses of biomass [27,35].

Biological pretreatment, such as fungal, is milder in its operational conditions than physical or chemical pretreatment. The oxidative biodegradation of lignin by white-rot fungi has been widely studied in the past [36]. The main advantages of biological pretreatment are low energy input, no chemical requirement, mild environmental conditions and environmentally friendly working manner [25]. However, the biological pretreatment processes usually need a long retention time and have a low yield. They are also sensitive to the process conditions such as temperature and pH. Enzymatic pretreatment of biomass has been comprehensively studied in the past and will be discussed in a separate section.

Advertisement

3. Hydrolysis

Generally, microorganisms have poor cellulose/biomass digestibility and a limited efficiency in producing biofuels. Hydrolysis has been commonly applied in industrial scales to improve the efficiency of microorganisms in producing biofuels prior to fermentation, which relies on the decomposition of polysaccharides to monosaccharides [37].

3.1. Enzyme hydrolysis

In this process, enzyme is used for decomposing polysaccharides into monosaccharides. Microorganisms that usually produce enzyme are fungal species, such as hypocerea jecornia, trichoderma reesei, and bacteria species, such as clostridium thermocellum, cellulomonas flavigena. Three main enzymes hydrolyzing cellulose to glucose are cellulase, 1-4- β-D-endoglucanase and 1-4- β-D-cellobiohydrolase [38]. Process parameters, e.g. pH, temperature, time, significantly affect the performance of enzyme hydrolysis. Furthermore, porosity and crystallinity and the lignin content of biomass seem to significantly influence the efficiency of this process [25].

Equation 1 describes the enzymatic hydrolysis model of cellulose to glucose [39]:

X=1-Ke+e0KeKae0t+1+e0bE1

where e0 is the initial enzyme concentration (g/l), X is conversion efficiency (<1), t is the reaction time (h), ka is the enzyme deactivation constant (g/l.h), b=kekak1 is the fitted constant (dimensionless), ke is the adsorption equilibrium constant (g/l), and k1 is the rate constant of sugar formation (1/h). When e0, X converges to a constant at constant time according to equation 2:

X=1-1KeKat+1bE2

and when t, X converges to a maximum conversion of 1.

Figure 1 shows the results of one study on the enzyme hydrolysis of pre-steamed corn Stover at 50 °C and pH 4.8, the conversion of celluloses to reducing sugar was 60% after 48h. Fitting the experimental data of Figure 1 into equation 1 resulted in ke, ka, k1, and b of 0.9975 (g/l), 0.9837 (l/g.h), 0.2843 (1/h), and 0.2897, respectively [39].

Figure 1.

The conversion of cellulose and the concentration of reducing sugars as a function of time in the enzymatic pretreatment. □, ∆ experimental points; solid and dash lines denote the model (Eq. 1) values [39].

Enzymatic hydrolysis has several advantages compared with acid hydrolysis such as a lower equipment cost and higher glucose yield without sugar degrading products or by-products [24].

3.2. Acid hydrolysis

This method has been extensively applied to convert oligomeric sugars to monomeric sugars in the past [40-43]. It has a short process time (i.e. less than 1 h) and produces a higher sugar yield (>85%), but operates at a relatively high temperature, e.g. >120 °C (compared to enzyme hydrolysis). However, acid hydrolysis may result in the production of undesirable by-products that should be eliminated prior to fermentation processes. These detoxification processes are generally costly and complicated [25,40-42].

Advertisement

4. Detoxification process

Natural occurring and process-induced compounds may retard the fermentation processes for producing biofuels, which complicates the fermentation process [42-46]. These inhibitors are phenols, acetic acid, furfural, metal ions and lignosulfonates, and can chemically or biologically be removed from hydrolysates prior to fermentation processes.

Boiling and overliming have been extensively used in different studies to reduce the concentration of the inhibitors [44-46]. Boiling was proposed to reduce the concentration of volatile components, e.g. furfural, and overliming was proposed to create some insoluble inorganic salts that adsorb inhibitors [47]. Alternatively, the inhibitors can be removed by employing the concept of adsorption and flocculation phenomena. In this regard, commercial adsorbents, e.g. activated carbons, fillers, e.g. calcium carbonate or lime, or ion exchange resins may be suitable choices. The adsorption/flocculation processes could effectively remove these inhibitors and research in improving the performance of this process is on-going [28,30,48-50].

Alternatively, the inhibitors can be removed from hydrolysates via biological treatments. For example, a mutant yeast, S. cerevisiae YGSCD 308.3, was applied to reduce the acetic acid content of the ammonia-based hardwood hydrolysate in one study [51]. This yeast grows on acetic acid, but not on xylose, glucose, mannose and fructose. This detoxification process resulted in an ethanol production (fermentation was conducted at a temperature of 30 °C, and 300 rpm for 24 h) with a 73% yield of that of the maximum theoretical value in a laboratory scale. However, the presence of acetic acid did not allow ethanol production in the control sample of this study [51].

Advertisement

5. Ethanol production

As of January 2008, 136 ethanol plants in the US produced 7.5 billion gallons of ethanol annually, and this number is expected to increase by 13.3 billion gallons/year via constructing additional 62 plants [4]. The current biofuel market is largely dominated by ethanol (90% of the world biofuel production) [2]. However, most of ethanol produced today comes from starch (as in maize grains) or sucrose (from sugarcane), i.e. first generation biofuel. Lignocellulosic biomass, on the other hand, represents more abundant feedstock for ethanol production, i.e. the second generation biofuel [4]. However, the cost of producing lignocellulosic ethanol could be almost double of that of corn-derived ethanol [2]. In this context, the US department of Energy (DOE) has committed over $ 1 billion dollars toward a realization of a 2012 goal of making lignocellulosic ethanol at a competitive cost of $1.33 per gallon [52].

The production of ethanol from biomass has been criticized in the past since a large amount of CO2 is produced (released) as the by-product of this fermentation process. However, one study showed that, if softwood (unspecified species and fermentation conditions) biomass were considered as raw material, and ethanol were produced from cellulose and hemicelluloses and FT-diesel from lignin, this integrated process could lead to 54% of mass conversion efficiency, 67% of carbon conservation efficiency, but more interestingly, 88% of energy conversion efficiency [12]. In other words, the majority of mass, but just 12% of energy, would be lost in ethanol production process. Consequently, ethanol production process would definitely improve the energy density of biomass, which is a critical factor of biofuel. This analysis depicted that it was extremely crucial to widen the assessments beyond the sole mass balance to obtain a complete understanding of biofuel production [12].

5.1. Process alternatives

The biological processes of ethanol production usually involve hydrolysis, which breaks cellulose to glucose, and fermentation to convert glucose to ethanol. Ethanol can be produced via separate hydrolysis and fermentation (SHF) or simultaneous saccharification and fermentation (SSF). The SHF facilitates the operation of hydrolysis and fermentation processes under separate optimized conditions. Previous studies on pre-steamed agro-based raw materials (corn Stover) revealed that the yield (ethanol produced per unit mass of dried feedstock g/g) of SHF was higher than that of SSF [53]. Also, SHF has a faster hydrolysis rate than does SSF under optimized conditions [11,39]. In SSF process, cellulose is hydrolyzed to glucose by cellulase, while yeast spontaneously ferments glucose to ethanol [54]. The SSF usually has a higher productivity (ethanol produced per unit mass of dried feedstock per unit time, g/g.h) than does SHF, since SSF has a shorter operating time [39]. Other advantages of SSF over SHF include less inhibition of enzymes and a longer time of enzymatic hydrolysis. Usually, fermentation temperature and the presence of ethanol are not at optimized conditions for cellulase activities in the SSF process, which leads to poor enzyme performance. To overcome this difficulty, several approaches were followed in the past, which are demonstrated as follows:

  1. Enzyme immobilization onto a solid support was proposed to improve the enzyme activity at non-optimal conditions (e.g. SSF). In one study, dissolved cellulase (500 µl, in 10 mM acetate buffer, pH 4.8 with 0.01 % (w/v) sodium azide) was immobilized on Aerosol OX-50 silica (40 nm average diameter particle, 1.4 m2 surface area), and the SSF was carried out using this coated silica [54]. The results showed that the ethanol yields of the SSF process containing immobilized enzyme were 2.3 and 2.1 times higher than that of the SSF process containing enzyme in solutions at pH 4.8 and 5.3, respectively. The higher ethanol yield of the immobilized enzyme process was likely due to higher glucose yields as a result of increased enzymatic stability at the non-optimal enzyme conditions required for the SSF [54]. However, this process results in a higher ethanol yield under the optimum conditions of fermentation rather than that of enzyme hydrolysis. This is because 1) the optimum conditions for the immobilized enzyme reaction may not coincide with those for the enzyme in solutions; 2) although the same mass of enzyme can be used in the solution and immobilized systems, it does not mean that the same amount of enzyme is available to participate in the hydrolysis reactions; and 3) enzyme immobilization results in a random orientation of adsorbed enzyme on the silica substrates, and this leads to some inactive enzyme due to buried or inaccessible active sites [54].

  2. Semi-simultaneous saccharification and fermentation (SSSF) is another approach to produce ethanol. In this method, the hydrolysis process is applied under optimized conditions, which breaks down cellulose to glucose, and subsequently fermentation is conducted on the product of hydrolysis without removing the hydrolystate. In other words, it is an operating mode between the SHF and SSF, thus it has the advantages of both SHF and SSF (a higher productivity and yield). In one study on microcrystalline cellulose (Avicel PH101), three alternative processes were carried out to produce ethanol: 24h hydrolysis+48h SSF (called SSSF), 72 h SSF, and 24 h hydrolysis+48h fermentation (SHF) [39]. The hydrolysis process was conducted under the conditions of 50 °C, pH 4.8 and 2 g Novozymes enzyme, while the fermentation was performed under the conditions of 36 °C, pH 4.8 and 300 rpm using S. cereviasiae. The results showed that the optimal ethanol productivity for SSSF, SSF and SHF were 0.222, 0.194, and 0.176 g/l.h, respectively. The corresponding maximum ethanol concentration was 16, 14 and 12.6 g/l with equivalent theoretical ethanol yields of 70.5%, 61.8%, and 56.1%, respectively [39].

5.2. Theory

The theoretical ethanol yield (Y) can be calculated using equation 3 [39]:

Y=0.9E0.511C0100%E3

where E is the final ethanol concentration (g/l) and C0 is the initial cellulose concentration (g/l). Also, the fermentation efficiency (ef), from sugar to ethanol, can be determined using equation 4:

ef=E0.511Gh100%E4

where Gh is the glucose concentration after hydrolysis (g/l). When glucan in cellulose is completely converted into glucose (C0=0.9 Gh). The theoretical yield of SSF yield is equal to fermentation efficiency [39].

5.3. Microorganism for ethanol production

Ethanol production via fermentation has been the subject of several research activities. In this regard, many yeasts and bacteria have been introduced or modified and their ethanol production efficiencies have been assessed.

5.3.1. Bacteria

Escherichia coli can consume hexoses and pentoses for ethanol production. However, E. coli is less rebust against several factors including pH, salt concentration and temperature. It also exhibits a low ethanol (<35 g/l) and butanol tolerance (< 20g/l) [51]. Z. mobilis has also been applied in ethanol production, has a tolerance of up to 120 g/l ethanol, and its ethanol production yield approaches 97% of the maximum theoretical value under optimized conditions. However, it can only ferment glucose, fructose and sucrose (hexoses), and it has a low tolerance to acetic acid [55,56]. Clostridia have also been applied in ethanol production under optimized temperature of 35-37 °C at a pH range of 6 and 9, which can ferment hexoses and pentoses [57]. Corynebacterium glutamicum has also been applied for ethanol production, but it cannot ferment pentoses, unlike E. coli. [58].

5.3.2. Yeast

Hexoses can effectively be converted to ethanol with a high yield (0.4-0.51 g ethanol/ g glucose) and a high productivity (up to 1 g/l.h) by Saccharomyces cerevisiae or re-combinant S. cerevisiae [53,55]. S. cerevisiae is the best known microorganism used for fermenting glucose to ethanol, but it cannot ferment pentoses. In this respect, the fermentation rate of xylose is 3-12 times lower than that of glucose by S. cerevisiae [45]. Kluyveromyces is another thermotolorant species with up to 98% of the maximum theoretical ethanol yield at a temperature above 40 °C, but it cannot accommodate pentose [59]. Hansenula polymorpha is another thermotolorant species with an optimized temperature of 37 °C, and its fermentation operation is possible up to a temperature of 48 °C. Kluyveromyces and H. polymorpha are suitable for SSF processes [60]. P. stipitis is another naturally fermenting pentose and hexose. It can be used for SSF set-up even though its optimized growth temperature is around 30 °C [53,55].

5.4. Ethanol recovery

The ethanol production is coincided with yeast cell growth in fermentation, hence the yeast is a by-product of the process. Pure yeast is a value-added product of the process and can inevitably decrease the net cost of the process [53]. Currently, centrifugation is applied to separate yeast cells from ethanol, which is an expensive process with a high energy demand. Yeast cell immobilization technologies using inert carries or chemicals have also been applied in ethanol industry for separating ethanol from cells [53]. Alternatively, yeast flocculation process was introduced in the 1980s and commercialized in 2005 in China and comprehensively used in brewing industry. It involves lectin-like proteins and selectively binds the mannose resides of cell wall of adjacent yeast cell. In this flocculation process, calcium ions are needed and the flocculation occurs spontaneously [61-63]. Upon formation, the flocs would either sediment (large yeast) or rise to the surface (ale yeasts). This flocculation process has the advantages of 1) allowing greater yeast cell biomass concentrations because of no inert carrier; 2) being a simpler and more economically competitive process and 3) fostering the yeast cell viability because the continuous renewal of cells resulting from breaking up of relatively large flocs. The yeast flocs can be purged from the fermenter maintaining the biomass concentration inside the fermenter at specified levels [64]. Different configurations of immobilization process are available for industry: airlift, single packed column, two-stage packed column and CO2 suspended bed [64].

5.5. SPORL and dilute acid ethanol production processes

Sulfite pretreatment to overcome recalcitrance of lignocellulose (SPORL) has been developed by the U.S. Forest Service, Forest Products Laboratory and the University of Wisconsin-Madison as a promising biorefinery process to produce ethanol from cellulosic materials [32,65,66]. The SPORL process has a short chemical pretreatment at a high temperature to remove recalcitrance of the substrate without a significant delignification, and a disk refiner to increase the surface area, which is necessary for the sugar dissolution in the latter stages for fermentation. The potential products of the SPORL process are ethanol, hemicellulosic sugars and lignosulfonate [67]. Figure 2 shows the process diagram of the SPORL process. As can be seen, the wood chips are pretreated with bisulfite and/or sulfuric acid at a temperature of 160-190 °C, a pH of 2-5, and liquid/wood ratio of 2-3 for 10-30 min. The bisulfite charge is 1-3% and 6-9% for hardwood and softwood species, respectively, and acid ranges between 1 and 2% [68]. The solid substrates of the chemical treatment are then fibrilized using a mechanical disk refiner. The chemical pretreatment has a direct effect on the refining stage of the SPORL process. Subsequently, the solid substrate is enzymatically hydrolyzed, fermented, and distilled to produce ethanol [67]. Hemicelluloses are also isolated from the spent liquor and fermented to ethanol, while lignosulfonate is a by-product of this process. Softwoods species have usually poor digestibility in enzymatic saccharification [65]. The SPORL process is particularly effective in improving the enzymatic saccharification efficiency of softwoods.

Figure 2.

Process flow diagram of the SPORL or dilute acid treatment [67].

Alternatively, ethanol can be produced by pretreating biomass with sulfuric acid, and subsequent mechanical pulping (refining) of the pretreated biomass. This process is called dilute acid process. However, a high temperature and a very low pH of this process impart serious equipment corrosion problems [70].

It was reported that the SPORL pretreatment was more efficient than the dilute acid pretreatment for recovering sugars from the solid substrate and spent liquor. Inhibitors were more in the spent liquor of the dilute acid pretreatment, but the concentration of lignin was higher in the spent liquor of the SPORL pretreatment. Due to the lower inhibitors presented in the spent liquor, the ethanol production would be facilitated with the SPORL system. The ethanol production from the substrate and spent liquor of pine species under various SPORL pretreatment conditions at 180 °C for 25 min and then fermenting via using S. cerevisiae (ATCC 200062) at 32 °C for 72 h at 100 rpm are listed in Table 2 [70].

Acid charge, %Bisulfate charge, %Initial pHSubstrate ethanol1Substrate ethanol efficiency2Hydrosate ethanol1Hydrolysate ethanol efficiency2Total ethanol yield
2.2141.8136.773.752.865.1189.5(49.2)
2.2181.9209.483.866.994.2276.3(71.7)
1.482.3193.276.436.370.4229.5(59.6)
084.2165.669.61.711.3166.7(43.3)

Table 2.

Ethanol production in the SPORL system of pine species under various conditions [70].

As can be seen, by increasing the bisulfite concentration from 4% to 8%, the ethanol production and efficiency were increased for both the substrate and spent liquor, resulting in 21% increase in the total ethanol production. It is also noticeable that the ethanol production from the substrate or spent liquor was reduced by reducing the acid charge (increasing pH) of the pretreatment.

Advertisement

6. Butanol production

Although ethanol has been considered as a promising biofuel, it has several drawbacks: the heat value of ethanol is 27 MJ/kg, while that of FT-diesel is 42.7 MJ/kg implying that ethanol contain a much lower energy density compared with diesel fuel used in automobile industry [2,12]; it has also a high water solubility that prevents it from being an ideal biofuel, and high concentration ethanol blends can cause corrosion of some metallic components in tanks and deterioration of rubbers and plastic used in car engines [2]. Consequently, the incentives for obtaining a better alternative biofuel are high.

The industrial synthesis of biobutanol was commenced during 1912-1914 by acetone-butanol- ethanol (ABE) fermentation of molasses and cereal grains using clostridium acetobutylicum [71]. However, butanol was primarily used as a solvent for the production of other chemicals prior to 2005. Butanol has similar energy density and polarity to those of gasoline [71]. It has an adequate blending ability with gasoline and compatibility to combustion engines [72]. It can be shipped through existing fuel pipelines, whereas ethanol must be transported via rail and truck [73,74]. It has octane-improving power and low volatility (six times less than the volatility of ethanol). These novel properties made butanol a promising biofuel [75]. The economics of butanol fermentation is favorable even with the present technology [76]. However, the capital cost of butanol fermentation is presently higher, but its production cost is less, than that of petrochemical process to produce butanol [77]. Despite steadily growing production, its market remains tight and its price high due to low investment and hefty demand in coming years [76].

6.1. Butanol fermentation process

To produce butanol via fermentation as the second generation biofuel, cellulose should be initially converted to glucose, as most of the butanol-fermenting microorganisms can digest glucose and not cellulose. Similar to ethanol production, the enzymatic hydrolysis of polysaccharides to monosaccharides and then fermentation to biosolvents were carried out in the past [78]. The drawbacks of this process is its energy intensity, which makes its commercialization costly [78,79].

The production of butanol from cellulose mainly relies on the application of clostridia species (C. acetobutylicum, C. beijerinckii, C. pasteurianum). Clostridia acetone-butanol-ethanol (ABE) fermentation from carbohydrates used to be the largest biotechnological process second to yeast ethanol fermentation and the largest process ever conducted under sterile conditions [72,76]. Clostridia species have several advantages: their thermophilic nature permits their utilization at a high temperature of 60 °C (facilitates sterilization process), they can digest pentoses, and more interestingly they can ferment cellulose, which implies that cellulose hydrolysis to glucose and glucose fermentation can be proceeded spontaneously. Butanol can only be generated if the cells enter sporulation process, but this process discourages cell growth. Thus, balancing cell growth and sporulation timing is a key factor influencing the carbon flow towards cell growth and butanol generation [74]. However, these species have some disadvantages including the low solvent resistance, comparative difficulty in genetic modification, and increased energetic demand for cellulase production under anaerobic environments [80].

In the ABE process, the coproduction of acetone, butanol and ethanol causes a poor selectivity with respect to butanol production. The theoretical mass and energy yields of ABE fermentation are 37% and 94%, respectively [81]. It was reported that the substrate costs account for 60% of the total production cost, and the butanol production will not be feasible if the fermentation yield is less than 25% [80]. The best results reported for the ABE process were 8.2 g/l acetone, 2.2 g/l ethanol and 17.6 g/l butanol [82].

6.2. Production improvement

The production of butanol faces some challenges: 1) selection of sustainable biomass 2) low production rate, 3) constrains executed on butanol inhibition and 4) high product recovery costs [72,83]. Different strategies can be performed to improve the production of butanol via fermentation as described in the following sections.

6.2.1. Eliminating inhibitors

The phenolic compounds of cellulosic materials inhibit the butanol fermentation process (similar that of ethanol fermentation process) [78]. In one study, the majority of inhibitory compounds were phenolic compounds, while furfural and hydroxymethyl furfural (HMF) did not affect the cell growth and ABE fermentation [84]. It was reported that, during the initial growth phase of cells, other by-products were also produced such as acetic acid and butyric acid, which inhibited the butanol production [84]. In one study, the presence of 1 g/l ferulic acid and vanillin acid reduced the cell growth by 70% and 56%, respectively [84]. The choice of detoxification method would depend on the compositions of hydrolysates and the species of fermenting microorganism [84]. Previously, lime treatment [85], evaporation, adsorption using ion exchange resins and activated carbon were used prior to fermentation as detoxification methods for butanol production [84]. However, the detoxification efficiency depends on the chemical structure of inhibitors [84]. Other inhibitors of the ABE process are substrate inhibition, salt concentration, presence of dead cells, low water activity, O2 diffusion, macromolecules accumulation and nutrient deficiency [86].

6.2.2. Removal of butanol

Clostridia species are known to be solventogenic in producing acetone, butanol, and ethanol, but are still subjected to negative inhibition by their own products [72,74]. The hurdles are being resolved using genetic engineering techniques, metabolic engineering strategies and integrated continuous fermentation processes [72]. In this context, one study showed that the presence of butanol at a concentration of higher than 7.4 g/l impaired the cell growth [87,88]. In this case, butanol may penetrate the cytoplasmic membrane and disrupt several physicochemical characteristics of the cells [72]. On the contrary, the Clostridium BOH3 species showed the advantage of high resistance against butanol (up to 16 g/l) [74].

In the literature, the simultaneous production and the removal of butanol was carried out in order to maintain a low butanol concentration in the fermentation medium (broth) using liquid-liquid extraction, adsorption, perstraction, reverse osmosis, pre-evaporation and gas stripping [89], among which gas stripping has received great attentions. Gas stripping offers several advantages including feasibility and simplicity of the process, reduction in butanol concentration without affecting culture, concentration of nutrients and reaction intermediates [86]. Furthermore, a membrane separation with super critical extraction may be a feasible method in butanol removal in future [72].

6.2.3. Modifying microorganism

One alternative to increase the butanol production was reported to be the genetic engineering for strain improvement with insertion of butanol producing gene of Clostridia in high butanol tolerant organisms. In this context, the genetic manipulating of Clostridia was reported to decrease its sensitivity to the presence of butanol, which eventually increased the butanol concentration up to 17.8 g/l in the fermentation medium [87]. Also, research on aerobic producing butanol using genetically engineered organics like, E. coli, are being attempted [72]. The strain improvement is effective in improving the yield, but has marginal influence on the economics.

6.3. Process configuration

The fermentation of glucose to butanol can be conducted in batch or continuous process. Generally, continuous butanol processes, (free cells, immobilized cells, and cell recycling) are more economical over batch processes. Other advantages are the reduction in sterilization and inoculation times and the superior productivity. In free cell continuous fermentation, cells are free to move within the fermentation broth due to agitation or air lifting. This maintains the microbial cells and nutrients in the suspension and helps in promoting mass transfer [72]. In immobilized cell fermentation, cells are stationary, which have the advantages of the long survival time of cells in solventogenic phase. In one study, immobilized cell fermentation using C. acetobutylicum produced 20% higher butanol yield than did free cell fermentation [90]. However, the scale up process using immobilized technology seems to have technical problems due to excessive cell growth in the packed beds and blockage [85]. Recycling technique using a membrane technology (e.g. filter) after the fermentation has also been attempted in the literature and the results showed 10 times higher cell concentration and 6 times higher butanol production compared with conventional continuous process using C. saccharoperbutylacetonicum [85]. On the contrary, batch process has a less capital cost and contamination problem (sterilization) in fermentation, and is more flexible.

Advertisement

7. Hydrogen production

Hydrogen has been considered as one of the major energy carriers in future due to its high energy content and its capability to overcome the air pollution and global warming [91]. Renewable carbohydrates (e.g. cellulose) are suitable raw materials for hydrogen production, i.e. second generation biofuel, because they are less expensive than other hydrogen carriers, e.g. hydrocarbons, biodiesel, methanol and ethanol [3,37]. Glucose is widely used substrate for hydrogen production [37]. Today, the production cost of hydrogen from renewable biomass is appealing ($ 60/dry ton or $ 3.6 per GJ) [15]. The most important energy application of hydrogen is transportation, especially for light-duty vehicles [3]. However, the large scale implementation of the hydrogen economy has some obstacles: sustainable hydrogen production, high density hydrogen storage, hydrogen distribution infrastructure, fuel cell cost and life time, and safety concerns [3]. Thus, new technologies and strategies should be developed to make the hydrogen production more economically attractive and industrially feasible [14].

7.1. Hydrogen production processes

Table 3 lists various pathways to produce hydrogen from glucose [3]. As can be seen, the theoretical and practical yields of hydrogen production via chemical catalytic reactions, i.e. gasification, pyrolysis and hydrolysis (accompanied by aqueous phase reforming (APR)), of glucose are in a similar range. The gasification is conducted at a high temperature of 1000 K in the presence of oxygen and water. Pyrolysis is carried out at a high temperature but in the absence of oxygen. The main advantage of APR over gasification is the lower extent of undesirable decomposition reaction [3]. The APR is carried out at a lower temperature (400-550 K) and a medium pressure (50-70 bar) [3]. The water medium of this process promotes the occurrence of the hydrogen production reaction. Although the chemical catalytic reaction seem to have considerably higher hydrogen production yields, their prerequisite high energy input and poor selectivity toward hydrogen production are barriers in their industrial implementations.

However, the biological conversion of cellulose to hydrogen is performed at much lower temperatures, which implies that the energy input of this process is much lower than that of chemical catalytic processes. In contrast, the theoretical and practical yields of hydrogen production via this method are rather low (see dark anaerobic fermentation in Table 3) [30,92,93]. In principle, up to 12 mol of hydrogen can be produced per mole of glucose and water via fermentation. However, natural microorganisms produce as much as 4 mol of hydrogen/ mol of glucose along with 1 mol of acetate [94]. The current yield of hydrogen production in cellulose fermentation ranges from 1 to 2 mol H2/mole of hexose sugar [95]. To increase the hydrogen yield, a bioelectrochemically-assisted microbial fuel reactor was reported to convert 2 mol of acetate to up to 8 mol of hydrogen with the aid of electricity [96]. In this respect, the overall production yield of hydrogen increased to 9 mol per mol of glucose [96]. However, this process also needs electricity input and has not been studied in large scales.

MethodTheoretical yield, %Practical yield, %Energy efficiency, %
Dark fermentation (DF)41-3.210-30
DF+ electricity-assisted microbial fuel cell12975
Ethanol fermentation/partial oxidation reforming10960
Gasification122-835-50
Pyrolysis122.5-830-50
Hydrolysis+ aqueous reforming126-830-50
Synthetic pathway biotransformation1212122

Table 3.

Methods for converting glucose to hydrogen as biofuel [3].

Synthetic pathway biotransformation is a new biocatalytic technology based on the application of enzymes to convert cellulose to hydrogen. This process is much simpler than biological treatments. The biotransformation of carbohydrates to hydrogen by cell-free synthetic pathway (enzymes) has numerous advantages: high production yield (12 H2/glucose unit), 100% selectivity, high energy conversion efficiency (122% based on combustion energy), high purity hydrogen generated, mild reaction conditions, low cost bioreactor and no toxicity hazards. In one study using 2 mM cellubiose concentration as the substrate, the overall yields of hydrogen and CO2 were 11.2 and 5.64 mol per mol of anhydroglucose unit corresponding to 93% and 94% of the theoretical yields, respectively [97]. Furthermore, these enzymatic reactions are reversible, thus the removal of products favors the unidirectional reaction towards the desired products. One study on the biotransformation of starch and water revealed that these reactions were spontaneous and endothermic (∆G0= -49.8 kJ/mol, ∆H0=+598 kJ/mol). Thus, these reactions are driven by entropy gain rather than enthalpy loss [3]. Such entropy-driven reactions can generate more output chemical energy in the form of hydrogen than input energy in the form of polysaccharides [97]. This is very interesting as most of the chemical conversions of cellulose/glucose to biofuel are exothermic. In other words, the output biofuel has less energy than do raw materials, i.e. cellulose/glucose. However, this method currently has a low hydrogen production rate [3]. Research in this area is on-going and no concrete conclusion has been made yet.

7.2. Biological production of hydrogen

The biological production of hydrogen has received great attentions due to its potential as an inexhaustible, low-cost and renewable source of clean energy [91]. The photosynthetic production of hydrogen basically uses CO2 and water via direct photolysis. The yield of this process is high, but there is a shortage of photobioreactors on large scales, which constrains its investigation in laboratory scales [98]. Compared to photosynthesis, anaerobic hydrogen production process is more feasible (less expensive), has higher rates and thus has been widely studied [99]. This process seems to be closer to be implemented in commercial scales.

Several breakthroughs have been made in understanding the fundamentals of hydrogen production including the isolation of microbial strains with a high hydrogen production capacity and the optimization of the microbial fermentation process [100,101]. The performance of anaerobic fermentation depends on a number of factors, e.g. temperature, pH, alkalinity oxidation-reduction potential, particle size of lignocellulosic materials, substrate content and inoculum source [91,102]. It was reported that the optimized pH for producing hydrogen in a dark fermentation is between 5.5 and 6.7 [102]. However, there are contradictory reports about the influence of ethanol on hydrogen production in dark fermentation processes in that some studies reported a possible competition between ethanol and hydrogen production [103-105], while others reported a high hydrogen production accompanied by a high ethanol production [106,107]. In another study, the addition of ethanol to the growth medium at the initiation of the fermentation process resulted in 54% H2 and 25% acetate increases, respectively, using C. Thermocellum bacteria [108].

In addition to hydrogen, organic acids are produced in dark fermentation. To achieve adequate overall energy efficiency, the energy stored in the organic acids produced in the dark fermentation processes should also be utilized. This can be conducted via combining the concepts of dark fermentation and anaerobic digestion [109].

On the contrary, dark fermentation has some drawbacks: 1) poor yield per substrate in the conversion of biomass to H2 by microbial fermentation; 2) sensitivity to end-product accumulation; and 3) high environmental and economic costs of biomass production to generate fermentable substrate [13,110-112].

7.3. Microorganisms to produce hydrogen

Thermophilic cellulosic bacteria promote their greater operating temperature to produce hydrogen, which facilitate biomass pretreatment, maximize enzymatic reaction rates, and favor the equilibrium point of H2 in direction of H2 production [13,14]. Clostridium thermocellum and caldicellulosiruptor saccharolyticus have been the main focus of hydrogen production research [95]. Caldicellulosiruptor saccharolyticus is a gram-positive and extremely thermophilic, has been reported to produce hydrogen from cellulose even at a high temperature of 70 °C, and is the most promising candidate for large scales hydrogen production [98,113]. Hydrogen has been reported to be produced from pentoses via using thermophile Caldicellulosiruptor saccharolytricus with a high yield of 334.7 ml H2/g of sugar accounting for 67% of maximum theoretical yield of 497.6 ml H2/g of sugar [37,114,115]. Furthermore, the optimum pH in the fermentation of cellulose to produce hydrogen via C. Thermocellum was between 7 and 7.2 [116].

In the literature, some mesophilic bacteria have been investigated for hydrogen production including Clostridium cellulolyticum, Clostidium ellulovorans, Clostridium phytofermentans and Clostridium termitidis. However, their low operating temperature (32-40 °C) introduces additional operating units to the hydrogen production process and increases the risks in process operations, which constrains the practical implementation of mesophilic bacteria in large industrial scales [3].

7.4. Fermentation culture

Although the pure culture usually provides a high production efficiency, it is difficult to apply in industrial scales. In fact, the preparation of pure cultures in indutrial scales is difficult and expensive, which will definitely affect the production cost of hydrogen. Instead, hydrogen can be produced with mixed cultures. The mixed culture has been claimed to be more effective in substrate conversion than pure culture [95].

However, mixed culture may encounter the drawbacks of the competition of substrates with non-hydrogen producing microbial population as well as the consumption of produced hydrogen by hydrogen consuming bacteria. One alternative to address this difficulty is to pretreat the mixed culture with base, heat and/or an anaerobic condition, which eliminates/inhibits the non-producing/consuming hydrogen bacteria [34,113,117]. Heat pretreatment is commonly used in anaerobic fermentation in order to improve the hydrogen production by activating spore-forming clostridium and inhibiting hydrogen consuming non-spore-forming bacteria. However, the heat pretreatment might inactivate hydrogen producing bacteria existing in natural feedstock [34,37]. The elucidation of anaerobic activated sludge microbial community utilizing monosaccharides will be an important foundation towards the industrialization of hydrogen production. In one study, a mixed microbial culture was obtained via the anaerobic activation of sludge in a continuous stirred-tank reactor (29 days of acclimatization) [118]. In this study, glucose had the highest specific hydrogen production rate (358 mL/g.g of mixed liquid volatile suspended solid) and conversion rate (82 mL/g glucose) among glucose, fructose, galactose and arabinose under the fermentation conditions of 35 °C and pH of 5 [34,118].

7.5. Improving production efficiency

7.5.1. Dark fermentation

The presence of hydrogen in fermentation reactors seems to affect the performance of dark fermentation process. It was reported that sparling the bioreactor with nitrogen would bring the concentration of hydrogen at a low level in the fermentation. Alternatively, using hollow fiber silicone rubber membrane effectively reduced biogas partial pressure in the fermentation resulting 10% improvement in the rate and 15% increase in the yield of hydrogen production [119]. To improve the hydrogen production, some strategies were carried out in the literature involving bioprocess engineering and bioreactor design that maintains a neutral pH during fermentation and ensures the rapid removal of hydrogen and CO2 from the aqueous phase [95].

7.5.2. Enzymatic treatment

Enzymes applied in the biotransformation of hydrogen are expensive. A combination of enzyme immobilization and thermo stable enzymes was reported to increase the life time of enzyme used for hydrogen production [120]. In one study, changing the process parameters (e.g. temperature, enzyme concentration, substrate concentration and metabolic channeling) in enzymatic reaction were reported to affect the hydrogen production rates [3]. A conservative estimation reported that the hydrogen production rates could increase to 23.6 H2/l/h using a high-cell density [121]. The over-expression of enzymes catalyzing reactions towards the desired product is another method to increase the hydrogen production. The heterologous gene expression of pyruvate decarboxylate and alcohol dehydrogenase from Zymomonas mobilis within C. cellulolyticum was shown to increase the hydrogen yield by 75%, while that of acetate and ethanol increased by 93% and 53%, respectively [122].

7.6. Process configuration

Similar to ethanol production, hydrogen production can be conducted in a one- or two-stage process. The simultaneous saccharification and fermentation process, i.e. one-stage process, is less expensive and more commercially feasible, but may be less efficient since the preferred conditions for cellulose degradation and dark fermentation could be significantly different [14]. An important parameter in this process is to choose a microorganism, such as thermococcus kodakaensis, clostridium thermolacticum and clostridium thermocellum, that has the capability to produce cellulolytic enzymes and hydrogen simultaneously [14,123]. In this process, the production rate and efficiency are limited by enzymatic saccharification. The two-stage hydrogen production process, on the other hand, is performed via cellulose hydrolysis in one stage and fermentation of the hexose in another stage [124]. This process might be more effective in terms of hydrogen yield, but it is more complicated.

Advertisement

8. Production of furan-based biofuel

Recently, a new second generation biofuel was introduced via the chemical conversion of cellulose. In this approach, hydroxymethyl furfural (HMF) is initially produced from cellulose, and HMF is subsequently converted to 2,5-dimethylfuran (DMF). DMF has an energy content of 31.5 MJ/l, which is similar to that of gasoline (35 MJ/l) and 40% greater than that of ethanol (23 MJ/l) [1]. DMF (bp 92-94 °C) is less volatile than ethanol (bp 78 °C), and is immiscible with water, which makes it an appealing liquid biofuel [1].

Figure 3 shows the process block diagram of DMF production from fructose. This process contains two main parts: 1) HMF production and its purification and 2) DMF production from HMF and its downstream purification.

Figure 3.

DMF production from fructose [6].

Two-solvent catalyst system of butanol/water (at 180 °C) was proposed for producing HMF from fructose [1]. In this process, the conversion of glucose/fructose takes place in the aqueous phase (with HMF yield of 83%), in which HCl acts as a catalyst to convert fructose to HMF. Also, NaCl is added to the system and enhances the transportation of HMF from aqueous phase to organic phase (butanol), which prevents HMF from further degradation in the aqueous phase. Afterwards, the HMF is purified via various distillation/separation units and butanol is recycled to the HMF production reactor [1].

In another study, methyl isobutyl ketone (MIBK)/water system (in the present of TiO2 and 250 °C) resulted in 35% HMF yield from glucose [125,126]. In this respect, ionic liquids, e.g.1-ethyl-3-methyl-imidazolium chloride ionic (in the presence of LiCl, CuCl or CrCl) showed a higher yield (>65%) at a temperature of 160 °C [127,128]. In another research, dimethyl sulfoxide (DMSO) was used as solvent to produce HMF from glucose with HMF yield of 53% [129]. However, these solvents are usually toxic and difficult to prepare, and their separation process from HMF is challenging [129]. These drawbacks will most likely limits their application in producing HMF in laboratory scales.

Alternatively, the hydrothermal conversion of cellulose to HMF (275-300 °C for less than 30 min) in homogenous systems in the presence of sulphuric acid resulted in a 20% HMF yield. However, this system under a neutral condition produced a lower yield (<16%), but a product (HMF) with a higher purity (50-60%) [130]. In another research, the yield of HMF from fructose was 65% in the presence of phosphoric acid, which was conducted at 228 °C for less than 5 min [1]. The most promising and industrially attractive method to produce HMF was the acetic acid/water mixture (10 g/l acid at 200 °C for 20 min), which resulted in the production of HMF from fructose with 60% yield [131]. This process resulted in 53% HMF yield under the same conditions, but without acetic acid [131]. Alternatively, the hot compressed water treatment of cellulose at a temperature of 523 K for 5 min (pressure 2.5 MPa under nitrogen) produced 15% HMF, and the addition of 10% (wt.) TiO2 to the system increased the yield to 28% [1]. This method is more feasible than other homogenous catalytic methods, as the catalyst in this system can be easily separated and thus recycled to the system [132]. Furthermore, the addition of acetone to water/TiO2 system induced a higher HMF yield (35%). This process was conducted under atmospheric condition, which is more industrially attractive, but the presence of acetone in the system brings recycling issues and uncertainties at large scale applications [1].

The produced HMF in Figure 3 will be fed into a PFTR reactor, in which H2 is added (in the presence of chromium II or copper-ruthenium-carbon catalysts) that is necessary for the conversion of HMF to DMF [6]. Finally, DMF is purified using several distillation/separation processes, and the organic solvent (butanol) is recycled to the system (Figure 3).

However, this process has several drawbacks: 1) it uses hydrogen which is a biofuel and currently expensive; 2) it uses butanol (another biofuel) as a solvent, which brings difficulties to its large scale implementations; 3) it uses NaCl to enhance the extraction of HMF from aqueous phase to organic butanol. NaCl introduces uncertainties in the downstream processes and its removal is cost intensive; 4) as described above, expensive chromium II or copper-ruthenium-carbon was used as a catalyst in the PFTR reactor [6]. It was reported that the production cost of DMF using this process is approximately 2 $/l, which is not presently economical. The process optimization seemed to lower the production cost to approximately 1 $/l, which is still high and not competitive with other appealing biofuels [6].

Alternatively, cellulose can be converted to 5-chloromethylfurfural (CMF) via heating cellulose in concentrated HCl at the presence of LiCl. Subsequently, the products should be extracted with 1,2-dichloroethane. This process yielded 71% CMF in one study [133]. The CMF was then converted to DMF and 5-ethoxymethylfurfural. This process yielded 84% isolated DMF, but the presence of LiCl is considered as one drawback of this system [133]. Alternatively, the CMF was converted to ethoxymethylfurfural (EMF) via stirring in ethanol solution. EMF has a boiling point of 235 °C, and energy density that is similar to energy density of gasoline and 40% higher than that of ethanol [133]. The EMF can also be used a biofuel, but this process seem to be complicated and faces with several technological challenges.

The most important parameters affecting the production cost of DMF via the aforementioned process are feedstock cost, production yield, by-product prices, catalyst cost, and total purchased equipment cost [6]. The productions of DMF and CMF are not feasible using current technologies. Although these chemicals were produced at laboratory scales, their commercialization is under serious questions as their production processes require various expensive solvents that are not easily recycled. The separation and purification of final products from solvents are also costly.

Advertisement

9. Conclusions

The pretreatment of woody materials is an important step for producing biofuels via fermentation. The physicochemical pretreatment is the most promising approach to dissemble cellulosic biomass. Enzymatic hydrolysis is very selective in terms of decomposing cellulose chains, but its process conditions are not very industrially attractive. However, acid hydrolysis is fast with a high yield, but it produces some by-products. These by-products are inhibitors of downstream fermentation processes and hence should be removed from the process prior to fermentation. Adsorption and evaporation have been the most successful approaches to eliminate these inhibitors in detoxification stages prior to fermentation. S. cerevisiae, C. acetobutylicum and C. thermocellum are the most promising microorganisms for ethanol, butanol and hydrogen productions, respectively. Presently, the major challenges in the production of these biofuels are the rather low production yield of biofuels and the sensitivity of microorganisms to the presence of inhibitors and biofuels in the fermentation media. The productions of ethanol and butanol from woody biomass are close to be commercialized, but hydrogen production is still facing with difficulties in increasing the production rate. Although 2,5-dimethylfuran (DMF) has appealing properties and can be potentially used as biofuel, its production with present technologies is not economical and more industrially attractive processes should be developed in order to have a commercialized DMF process.

Advertisement

Acknowledgement

The author would like to thank NSERC Canada for providing funds for this review research.

References

  1. 1. Binder JB, Raines RT.2009Simple Chemical Transformation of Lignocellulosic Biomass into Furrans for Fuels and Chemicals. Journal of American Chemical Society 2009; 13119791985
  2. 2. Serrano-RuizJ. C.WestR. M.DumesicJ.2010Catalytic Conversion of Renewable Biomass Resources to Fuels and Chemicals. Chemical and Biomolecular Engineering-Annual Review 2010; 179100
  3. 3. Zhang Y-HP.2010Renewable Carbohydrates Are a Potential High-Density Hydrogen Carrier. International Journal of Hydrogen Energy 2010; 351033410342
  4. 4. SivakumarG.ValiD. R.XuJ.BurnerD. M.Jr GeJ. O. L.WeathersX.P. J.2010Bioethanol and Biodiesel: Alternative Liquid Fuels for Future Generations. Engineering and Life Science 2010; 101818
  5. 5. Visser EM, Filho DO, Martins MA, Steward BL.2011Bioethanol Production Potential from Brazilian Biodiesel Co-products. Biomass and Bioenergy 2011; 35489494
  6. 6. Kazi FK, Patel AD, Serrano-Ruiz JC, Dumesic JA, Anex RP.2011Techno-Economic Analysis of Dimethylfuran (DMF) and Hydroxymethylfurfural (HMF) Production from Pure Fructose in Catalytic Processes. Chemical Engineering Journal 2011; 169329338
  7. 7. AlvandiS.AgrawalA.2008Experimental Study of Combustion of Hydrogen Syngas/Methane Fuel Mixtures in a Porous Burner. International Journal of Hydrogen Energy 2008; 33414071415
  8. 8. KaparajuP.SerranoM.ThomsenA. B.KongjanP.AngelidakiI.2009Bioethanol, Biohydrogen and Biogas Production from Wheat Straw in a Biorefinery Concept. Bioresource Technology 2009; 10025622568
  9. 9. FischerC. R.Klein-MarcuschamerD.StephanopoulosG.2008Selection and Optimization of Microbial Hosts for Biofuels Production. Metabolic Engineering 2008; 10295304
  10. 10. JohnsonJ. M. F.MDColemanGesch. R.JaradatA.MitchellR.ReicoskyD.WilhelmW. W.2007Biomass-bioenergy Crops in the United States: A Changing Paradigm. American Journal of Plant Science and Biotechnology 2007; 1128
  11. 11. WeberC.FarwickA.BenischF.BratD.DietzH.SubtilT.BolesE.2010Trends and Challenges in the Microbial Production of Lignocellulosic Bioalcohol Fuels. Applied Microbiology and Biotechnology 2010; 8713031315
  12. 12. CherubiniF.StrommanA. H.2010Production of Biofuels and Biochemicals from Lignocellulosic Biomass: Estimation of Maximum Theoretical Yield and Efficiencies Using Matrix Algebra. Energy and Fuels 2010; 2426572666
  13. 13. Jones PR.2008Improving Fermentative Biomass-Derived H2-Production by Engineering Microbial Metabolism. International Journal of Hydrogen Energy 2008; 3351225130
  14. 14. Cheng CL, Lo YC, Lee KS, Duu JL, Lin CY, Chang JS.2011Biohydrogen Production from Lignocellulosic Feedstock. Bioresource Technology 2011; 10285148523
  15. 15. Zhang Y-HP.2009A Sweet Out-Of-The-Box Solution to the Hydrogen Economy: Is the Sugar-Powered Car Science Fiction? Energy and Environmental Science 2009; 2272282
  16. 16. DautzenbergG.GerhardtM.KammB.2011Bio-based Fuels and Fuel Additives from Lignocellulose Feedstock via the Production of Levulinic Acid and Furfural. Holzforschung 2011; 65439451
  17. 17. KodaR.TakaoN.ShinjiH.SriappareddyT.KazunoriN.TsutomuN.ChiakiO.HidekiF.KondoA.2010Ethanolysis of Rapeseed Oil to Produce Biodiesel Fuel Catalyzed by Fusarium Heterosporum Lipase-Expressing Fungus Immobilized Whole-Cell Biocatalysts. Journal of Molecular Catalysis B: Enzymatic 2010: 66(1-2): 101-104.
  18. 18. AndréA.DiamantopoulouP.PhilippoussisA.SarrisD.KomaitisM.PapanikolaouS.2010Biotechnological Conversions of Bio-Diesel Derived Waste Glycerol into Added-Value Compounds by Higher Fungi: Production of Biomass, Single Cell Oil and Oxalic Acid. Industrial Crops and Products 2010; 312407416
  19. 19. VenkataSubhash. G.VenkataMohan. S.2011Biodiesel Production from Isolated Oleaginous Fungi Aspergillus sp. Using Corncob Waste Liquor as a Substrate. Bioresource Technology 2011; 1021992869290
  20. 20. MengX.YangJ.XuX.ZhangL.NieQ.XianM.2009Biodiesel Production from Oleaginous Microorganisms. Renewable Energy 2009; 34115
  21. 21. AraiS.NakashimaK.TaninoT.OginoC.KondoA.FukudaH.2010Production of Biodiesel Fuel from Soybean Oil Catalyzed by Fungus Whole-Cell Biocatalysts in Ionic Liquids. Enzyme and Microbial Technology 2010; 465155
  22. 22. XiaC.ZhangJ.ZhangW.HuB.2011A New Cultivation Method for Microbial Oil Production: Cell Pelletization and Lipid Accumulation by Mucor Circinelloides. Biotechnology for Biofuels 2011; 41524
  23. 23. FanY.ZhangY.ZhangS.HouH.RenB.2006Efficient Conversion of Wheat Straw Wastes into Biohydrogen Gas by Cow Dung Compost. Bioresource Technology 2006; 973500505
  24. 24. CaraC.MoyaM.BallesterosI.MJNegroGonzalez. A.RuizE.2007Influence of Solid Loading on Enzymatic Hydrolysis of Steam Exploded or Liquid Hot Water Pretreated Olive Tree Biomass. Process Biochemistry 2007; 4210031009
  25. 25. LinZ. X.HuangH.ZhangH. M.ZhangL.YanL. S.ChenJ. W.2010Ball Milling Pretreatment of Corn Stover for Enhancing the Efficiency of Enzymatic Hydrolysis. Applied Biochemistry and Biotechnology 2010; 16218721880
  26. 26. MJNegroManzanares. P.OlivaJ. M.BallesterosI.BallesterosM.2003Changes in Various Physical/Chemical Parameters of Pinus Pinaster Wood after Steam Explosion Pretreatment. Biomass and Bioenergy 2003; 25301308
  27. 27. Cheng JJ, Timilsina GR.2011Status and Barriers of Advanced Biofuel Technologies: A Review. Renewable Energy 2011; 3635413549
  28. 28. ShiH.FatehiP.XiaoH.NiY.2011A Combined Acidification/PEO Flocculation Process to Improve the Lignin Removal from the Pre-Hydrolysis Liquor of Kraft-Based Dissolving Pulp Production Process. Bioresource Technology 2011; 10251775182
  29. 29. LiuZ.FatehiP.MSJahanNi. Y.2011Separation of Lignocellulosic Materials by Combined Processes of Prehydrolysis and Ethanol Extraction. Bioresource Technology 2011; 10212641269
  30. 30. FatehiP.ShenJ.NiY.2011Treatment of Pre-Hydrolysis Liquor of Kraft-Based Dissolving Pulp Process with Surfactant and Calcium Oxide. Journal of Science and Technology for Forest Products and Processes 2011; 133137
  31. 31. ShenJ.FatehiP.SoleymaniP.NiY.2011A Process to Utilize the Lignocelluloses of Pre-Hydrolysis Liquor in the Lime Kiln of Kraft-Based Dissolving Pulp Production Process. Bioresource Technology 2011; 1021003510039
  32. 32. FatehiP.NiY.2011Integrate Forest Biorefinery-Sulfite Pulping. In: Zhu J, Zhang X, Pan X. (ed.) Sustainable Production of Fuels, Chemicals, and Fibers from Forest Biomass. American Chemical Society Symposium Series 2011; 1067: 409441
  33. 33. FatehiP.NiY.2011Integrated Forest Biorefinery- Prehydrolysis/Dissolving Pulping Process. In: Zhu J, Zhang X, Pan X. (ed.) Sustainable Production of Fuels, Chemicals, and Fibers from Forest Biomass. American Chemical Society Symposium Series 2011; 1067: 475506
  34. 34. ChairattanamanokornP.PenthamakeeratiP.ReungsangA.LoY. C.LuW. B.ChangJ. S.2009Production of Biohydrogen from Hydrolyzed Bagasse with Thermally Preheated Sludge. International Journal of Hydrogen Energy 2009; 3476127617
  35. 35. Gomez LD, Stelle-King CG, McQueen-Mason SJ.2008Sustainable Liquid Biofuels from Biomass: the Writing’s on the Walls. New Phytologist 2008; 178473485
  36. 36. TaniguchiM.SuzukiH.WatanabeD.SakaiK.HoshinoK.TanakT.2005Evaluation of Pretreatment with Pleurotus Ostreatus for Enzymatic Hydrolysis of Rice Straw. Journal of Bioscience and Bioenergy 2005; 1006637643
  37. 37. Lin CY, Hung WC.2008Enhancement of Fermentative Hydrogen/Ethanol Production from Cellulose Using Mixed Anaerobic Cultures. International Journal of Hydrogen Energy 2008; 3336603667
  38. 38. HarnpicharnchaiP.ChampredaV.SornlakeW.EuwilaichitrL. A.2009A Thermotolerant Beta-glucosidase Isolated from an Endophytic Fungi, Periconia sp., with a Possible Use for Biomass Conversion to Sugars. Protein Expression Purification Journal 2009; 676169
  39. 39. ShenJ.AgblevorF. A.2010Modeling Semi-Simultaneous Saccharification and Fermentation of Ethanol Production from Cellulose. Biomass and Bioenergy 2010; 3410981107
  40. 40. SaeedA.FatehiP.NiY.van HeiningenA.2011Impact of Furfural on the Sugar Analysis of Pre-Hydrolysis Liquor of Kraft-Based Dissolving Pulp Production Process Using the HPAEC Technique. BioResources 2011; 6217071718
  41. 41. LiuZ.FatehiP.NiY.2011A Proposed Process for Utilizing the Hemicelluloses of Pre-Hydrolysis Liquor in Papermaking. Bioresource Technology 2011; 10296139618
  42. 42. LiuX.FatehiP.NiY.2011Adsorption of Lignocellulosic Materials Dissolved in Hydrolysis Liquor of Kraft-Based Dissolving Pulp Production Process on Polymer-Modified Activated Carbons. Journal of Science and Technology for Forest Products and Processes 2011; 114654
  43. 43. LeschinskyM.ZuckerstatterG.WeberH. K.PattR.SixtaH.2008Effect of Autohydrolyis of Eucalyptus Globulus Wood on Lignin Structure. Part 2: Effect of Autohydrolysis Intensity. Holzforschung 2008; 62653658
  44. 44. HelleS.CameronD.LamJ.WhiteB.DuffS.2003Effect of Inhibitory Compounds Found in Biomass Hydrolysates on Growth and Xylose Fermentation by a Genetically Engineered Strain of S. Cerevisiae. Enzyme and Microbial Technology 2003; 33786792
  45. 45. HelleS. S.MurryA.LamJ.CameronD. R.DuffS. J. B.2004Xylose Fermentation by Genetically Modified Saccharomyces Cerevisiae 259ST in Spent Sulfite Liquor. Bioresource Technology 2004; 92163171
  46. 46. HelleS. S.LinT.DuffS. J. B.2008Optimization of Spent Sulfite Liquor Fermentation. Enzyme and Microbial Technology 2008; 42259264
  47. 47. Nigam JN. Ethanol2001Production from Hardwood Spent Sulfite Liquor Using an Adapted Strain of Pichia Stipitis. Journal of Industrial Microbiology and Biotechnology 2001; 26145150
  48. 48. SaeedA.FatehiP.NiY.2011Chitosan as a Flocculant for Pre-Hydrolysis Liquor of Kraft-Based Dissolving Pulp Production Process. Carbohydrate Polymer 2011; 8616301636
  49. 49. LiuX.FatehiP.NiY.2011Adsorption of Lignocellulosic Materials Dissolved in Pre-Hydrolysis Liquor of Kraft-Based Dissolving Pulp Process on Oxidized Activated Carbons. Industrial Engineering and Chemistry Research 2011; 501170611711
  50. 50. ShenJ.FatehiP.SoleimaniP.NiY.2012Lime Treatment of Pre-hydrolysis Liquor from the Kraft-based Dissolving Pulp Production Process. Industrial Engineering and Chemistry Research 2012; 51662667
  51. 51. SchneiderH.1996Selective Removal of Acetic Acid from Hardwood-Spent Sulfite Liquor Using a Mutant Yeast. Enzyme and Microbial Technology 1996; 199498
  52. 52. WaltzE.2008Cellulosic Ethanol Booms despite Unproven Business Models. Natural Biotechnology 2008; 2689
  53. 53. OhgrenK.BuraR.LesnickiG.SaddlerJ.ZacchiG.2007A Comparison between Simultaneous Saccharification and Fermentation and Separate Hydrolysis and Fermentation Using Steam-Pretreated Corn Stover. Process Biochemistry 2007; 425834839
  54. 54. Lupoi JS, Smith EA.2011Evaluation of Nanoparticle-Immobilized Cellulase for Improved Ethanol Yield in Simultaneous Saccharification and Fermentation Reactions. Biotechnology and Bioengineering 2011; 1081228352843
  55. 55. KnoshaugE. P.ZhangM.2009Butanol Tolerance in a Selection of Microorganisms. Applied Biochemistry and Biotechnology 2009; 1531320
  56. 56. RogersP. L. K.LeeJ.SkotnickiM. L.TribeD. E.1982Ethanol Production by Zymononas mobilis. Advanced Biochemical Engineering 1982; 223784
  57. 57. Warnick TA, Methe BA, Leschine SB.2002Clostridium Phytofermentans sp. nov., a Celluloytic Mesophile from Forest Soil. International Journal of System and Evolution Microbiology 2002; 5211551160
  58. 58. KawaguchiH.SasakiM.AAVertesInui. M.YukawaH.2009Identification and Functional Analysis of the Gene Cluster for L-arabinose Utilization in Corynebacterium Glutaicum. Applied Environmental Microbiology 2009; 751134193429
  59. 59. BanatI. M.NigamP.MarchantR.1992Isolation of Thermotolorant, Fermentative Yeasts Growing at 52 °C and Producing Ethanol at 45 °C and 50 °C. World Journal of Microbiology and Biotechnology 1992; 8259263
  60. 60. Ryabova OB, Chmil OM, Sibirny AA.2003Xylose and Cellobiose Fermentation to Ethanol by the Thermotolorant Methylotrophic Yeast Hansenula Polymorpha. FEMS Yeast Research 2003; 4157164
  61. 61. VerstrepenK. J.DerdelelinckxG.VerachtertH.DelvauxF. R.2003Yeast Flocculation: What Brewers Should Know. Applied Microbiology and Biotechnology 2003; 61197205
  62. 62. XuT. J.ZhaoQ.BaiF. W.2005Continuous Ethanol Production Using Self-Flocculating Yeast in a Cascade of Fermentors. Enzyme and Microbial Technology 2005; 37634640
  63. 63. BaiF. W.AndersonW. A.Moo-YoungM.2008Ethanol Fermentation Technologies from Sugar and Starch Feedstocks. Biotechnology Advances 2008; 2689105
  64. 64. GeX. M.ZhangL.BaiF. W.2006Impact of Temperature, pH, Divalent Cations, Sugars, and Ethanol on the Flocculating of SPSC01. Enzyme and Microbial Technology 2006; 39783787
  65. 65. ZhuJ. Y.PanX. J.WangG. S.GleisnerR.Sulfite2009Pretreatment (SPORL) for Robust Enzymatic Saccharification of Spruce and Red Pine. Bioresource Technology 2009; 10024112418
  66. 66. WangG. S.PanX. J.ZhuJ. Y.GleisnerR.RockwoodD.2009Sulfite Pretreatment to Overcome Recalcitrance of Lignocellulose (SPORL) for Robust Enzymatic Saccharification of Hardwoods. Biotechnology Progress 2009; 25410861093
  67. 67. ZhuW.ZhuJ. Y.GleisnerR.PanX. J.2010On Energy Consumption for Size-Reduction and Yields from Subsequent Enzymatic Saccharification of Pretreated Lodgepole Pine. Bioresource Technology 2010; 10127822792
  68. 68. Zhu JY, Pan XJ.2010Woody Biomass Pretreatment for Cellulosic Ethanol Production: Technology and Energy Consumption Evaluation. Bioresource Technology 2010; 10149925002
  69. 69. ShuaiL.YangQ.ZhuJ. Y.LuF. C.WeimerP. J.RalphJ.PanX. J.2010Comparative Study of SPORL and Dilute-Acid Pretreatments of Spruce for Cellulosic Ethanol Production. Bioresource Technology 2010; 10131063114
  70. 70. ZhuJ. Y.ZhuW.O’BryanP.BSDienTian. S.GleisnerR.PanX. J.2010Ethanol Production from SPORL-Pretreated Lodgepole Pine: Preliminary Evaluation of Mass Balance and Process Energy Efficiency. Applied Microbiology and Biotechnology 2010; 8613551365
  71. 71. DurreP.2007Biobutanol: An Attractive Biofuel. Biotechnology Journal 2007; 215251534
  72. 72. KumarM.GayenK.2011Developments in Biobutanol Production: New Insights. Applied Energy 2011; 8819992012
  73. 73. SchwarzW. H.GapesR.2006Butanol-Rediscovering a Renewable Fuel. BioWorld Europe 2006; 11619
  74. 74. BramonoS. E.LamY. S.OngS. L.HeJ.2011A Mesophilic Clostridium Species That Produces Butanol from Monosaccharides and Hydrogen from Polysaccharides. Bioresource Technology 2011; 10295589563
  75. 75. KeisS.ShaheenR.JonesT. D.2011Emended Description of Clostridium Acetobutylicum and Clostridium Beijerinckii and Descriptions of Clostridium Saccharoperbutylacetonicum sp. nov. and Clostridium Saccharobutylicum sp. nov. International Journal of System Evolution and Microbiology 2011; 5120952103
  76. 76. ZverlovV. V.BerezinaO.VelikodvorskayaG. A.SchwarzW. H.2006Bacterial Acetone and Butanol Production by Industrial Fermentation in the Soviet Union: Use of Hydrolyzed Agricultural Waste for Biorefinery. Applied Microbiology and Biotechnology 2006; 71587597
  77. 77. MarlattJ. A.DattaR.1986Acetone-Butanol Fermentation Process Development and Economic Evaluation. Biotechnology Progress 1986; 212328
  78. 78. EzejiT.QureshiN.BlaschekH. P.2007Butanol Production from Agriculture Residues: Impact of Degradation Products on Clostridium Beijerinckii Growth and Butanol Fermentation. Biotechnology and Bioengineering 2007; 97614601469
  79. 79. EzejiT.BlaschekH. P.2008Fermentation of Dried Distillers’s Grains and Solubles (DDGS) Hydrolysates to Solvents and Value-Added Products by Solventogenic Clostridia. Bioresource Technology 2008; 991252325242
  80. 80. DemainA. L.NewcombM.WuJ. H. D.2005Cellulose, Clostridia and Ethanol. Microbiology and Molecular Biology Review 2005; 69124154
  81. 81. QureshiN.BlaschekH. P.2001ABE Production from Corn: a Recent Economic Evaluation. Journal of Industrial Microbiology and Biotechnology 2001; 27292297
  82. 82. Nair RV, Green EM, Watson DE, Bennett GN, Papoutsakis ET.1999Regulation of the Sol Locus Genes for Butanol and Acetone Formation in Clostridium Acetobutylicum ATCC 824 by A Putative Transcriptional Repressor. Journal of Bacteriology 1999; 181319330
  83. 83. TashiroY.TakedaK.KobayashiG.SonomotoK.IshizakiA.YoshinoS.2004High Butanol Production by Clostridium Saccharoperbutylactonicum N1-4 in Fed-Batch Culture with pH-Stat Continuous Butyric Acid and Glucose Feeding Method. Journal of Bioscience and Bioengineering 2004; 984263268
  84. 84. ChoD. H.LeeY. J.UmY.SangB. I.KimY. H.2009Detoxification of Model Phenolic Compounds in Lignocellulosic Hydrolysates with Peroxidase for Butanol Production from Clostridium Beijerinckii. Applied Microbiology and Biotechnology 2009; 8310351043
  85. 85. QureshiN.LaiL. L.BlaschekH. P.2004Scale-up of a High Productivity Continuous Biofilm Reactor to Produce Butanol by Adsorbed Cells of Clostridium Beijerinckii. Food Bioproduct Processing 2004; 82164173
  86. 86. MaddoxI. S.QureshiN.ThomsonK. R.1999Production of Acetone-Butanol-Ethanol from Concentrated Substrates Using Clostridium Acetobutylicum in an Integrated Fermentation-Product Removal Process. Process Biochemistry 1999; 303209215
  87. 87. LeeS. Y.ParkJ. H.JangS. H.NielsenL. K.KimJ.JungK. S.2008Fermentative Butanol Production by Clostridia. Biotechnology and Bioengineering 2008; 1012209228
  88. 88. LiuS.QureshiN.How Microbes Tolerate Ethanol and Butanol. New Biotechnology 200926117121
  89. 89. ZhengY. N.LiL. Z.XianM.MaYangY. J.XuJ. M.HeX.D. Z.2009Problems with the Microbial Production of Butanol. Journal of Industrial Microbiology and Biotechnology 2009; 3611271138
  90. 90. Huang WC, Ramey DE, Yang ST.2004Continuous Production of Butanol by Clostridium Acetobutylicum Immobilized in a Fibrous Bed Bioreactor. Applied Biochemistry and Biotechnology 2004; 115887898
  91. 91. YuanX.ShiX.ZhangP.WeiY.GuoR.WangL.2011Anaerobic Biohydrogen Production from Wheat Straw stalk by Mixed Microflora: Kinetic Model and Particle Size Influence. Bioresource Technology 2011; 10290079012
  92. 92. Adams MWW, Stiefel EL.1998Biological Hydrogen Production: Not So Elementary. Science 1998; 28218421843
  93. 93. Hallenbeck PC, Benemann JR.2002Biological Hydrogen Production: Fundamentals and Limiting Processes. International Journal of Hydrogen Energy 2002; 2711851193
  94. 94. ThauerK.JungermannK.DeckerK.1977Energy Conservation in Chemotrophic Anaerobic Bacteria. Bacteriology Review 1977; 41100180
  95. 95. LevinB. L.CarereC. R.CicekN.SparlingR.2009Challenges for Biohydrogen Production via Direct Lignocellulose Fermentation. International Journal of Hydrogen Energy 2009; 3473907403
  96. 96. Logan BE, Regan JM.2006Microbial Fuel Cells-Challenges and Applications. Environmental Science and Technology 2006; 4051725180
  97. 97. YeX.WangY.HopkinsR. C.AdamsM. W. W.EvansB. R.MielenzJ. R.ZhangY. H.2009Spontaneous High-Yield Production of Hydrogen from Cellulosic Materials and Water Catalyzed by Enzyme Cocktails. Chemistry and Sustainability 2009; 2149152
  98. 98. HerbelZ.RakhelyG.BagiZ.IvanovaG.AcsN.EtelkaK.KovacsL.2010Exploitation of the Extremely Thermophilic Caldicellulosiruptor Saccharolyticus in Hydrogen and Biogas Production from Biomasses. Environmental Technology 2010; 31(8-9): 1017-1024.
  99. 99. FangH. H. P.ZhangT.LiuH.2003Biohydrogen Production from Starch in Wastewater under Thermophilic Conditions. Journal of Environment and Management 2003; 69149156
  100. 100. RenN. Q.WangB. Z.MaF.1995Hydrogen Bio-production of Carbohydrate by Anaerobic Activated Sludge Process. In: Proceeding of Water Environmental Federation and 68th Annual Conference Expo, Miami Beach, Florida, Oct 21-25 1995, 145-153.
  101. 101. MARachmanFurutani. Y.NakashimadaY.KakizonoT.NishioN.1997Enhanced Hydrogen Production in Altered Mixed Acid Fermentation of Glucose by Enterobacter Aerogenes. Journal of Fermentation Bioengineering 1997; 83358363
  102. 102. Li JZ, Ren NQ.1998The Operational Controlling Strategy about the Optimal Fermentation Type of Acidogenic Phase. China Environmental Science 1998; 185398402
  103. 103. LiC.FangH. H. P.2007Fermentative Hydrogen Production from Wastewater and Solid Wastes by Mixed Cultures. Critical Review in Environmental Science and Technology 2007; 37139
  104. 104. HawkesF. R.DinsdaleR.HawkesD. L.HussyL.2002Sustainable Fermentative Hydrogen Production: Challenges for Process Optimization. International Journal of Hydrogen Energy 2002; 2713351347
  105. 105. KhanalS. K.ChenW. H.LiL.SungS.2005Biological Hydrogen Production: Effect of pH and Intermediated Products. International Journal of Hydrogen Energy 2005; 2911231131
  106. 106. RenN.LiJ.LiB.WangY.LiuS.2006Biohydrogen Production from Molasses by Anaerobic Fermentation with a Pilot Scale Bioreactor System. International Journal of Hydrogen Energy 2006; 3121472157
  107. 107. KoshinenP. E. P.LayC. H.BeckS. R.TolvanenA. H.KaksonenA. H.OrlygssonJ.LinC. Y.PuhakkaJ. A.2008Bioprospecting Thermophilic Microorganisms from Icelandic Hot Springs for Sustainable Energy (H2 and/or Ethanol) Production. Energy and Fuels 2008; 22124140
  108. 108. RydzakT.LevinD. B.CicekN.SparlingR.2011Growth Phase Dependent Enzyme Profile of Pyruvate Catabolism and End-Product Formation in Clostridium Thermocellum ATCC 27405. Journal of Biotechnology 2011; 140169175
  109. 109. LjunggrenM.ZacchiG.2010Techno-Economic Analysis of a Two-step Biological Process Producing Hydrogen and Methane. Bioresource Technology 2010; 10177807788
  110. 110. HillJ.Nelsone.TilmanD.PolaskyS.TiffanyD.2006Environmental, Economic, and Energetic Costs and Benefits of Biodiesel and Ethanol Biofuels. Proceeding of National Academic Science USA 2006; 1031120611210
  111. 111. AngenentL. T.KarimK.Al-dahhanM. H.BAWrenn-EspinosaDomiguez.R.2004Production of Bioenergy and Biochemicals from Industrial and Agricultural Wastewater. Trends in Biotechnology 2004; 22477485
  112. 112. De VrijeT.MarsA. E.MABuddeLai. M. H.DijkemaC.de WarrdP.ClaassenP. A.2007Glycolytic Pathway and Hydrogen Yield Studies of the Extreme Thermophile Caldicellulosiruptor Saccharolyticus. Applied Microbiology and Biotechnology 2007; 7413581367
  113. 113. PanagiotopoulosI. A.BakkerR. R.de VrijeT.KoukiosE. G.ClaassenP. A. M.2010Pretreatment of Sweet Sorghum Bagasse for Hydrogen Production by Caldicellulosiruptor Saccharolyticus. International Journal of Hydrogen Energy 2010; 3577387747
  114. 114. KadarZ.de VrijeT.van NoordenG.BuddeM.SzengyelZ.ReczeyK.ClaassenP.2004Yield from Glucose, Xylose, and Paper Sludge Hydrolysate during Hydrogen Production by the Extreme Thermophile Caldaicellulosriruptur Saccharolyticus. Applied Biochemistry and Biotechnology 2004; 1141497508
  115. 115. LoY.ChenW.HungC.ChenS.ChangJ.2008Dark H2 Fermentation from Sucrose and Xylose Using H2Producing Indigenous Bacteria: Feasibility and Kinetic Studies. Water Research 2008; 42(4-5): 827-842.
  116. 116. IslamR.CicekN.SparlingR.LevinD. B.2009Influence of Initial Cellulose Concentration on the Carbon Flow Distribution during Batch Fermentation by Clostridium Thermocellum ATCC 27405. Applied Microbiology and Biotechnology 2009; 82141148
  117. 117. Modan SV, Bhaskar YV, Krishna PM, Rao NC, Babu VL, Sarna PN.2007Biohydrogen Production from Chemical Wastewater as Substrate by Selectively Enriched Anaerobic Mixed Consortia: Influence of Fermentation pH and Substrate Composition. International Journal of Hydrogen Energy 2007; 3222862295
  118. 118. LiJ.RenN.LiB.QinZ.HeJ.2008Anaerobic Biohydrogen Production from Monosaccharides by a Mixed Microbial Community Culture. Bioresource Technology 2008; 9965286537
  119. 119. Liang TM, Cheng SS, Wu KL.2002Behaviour Study on Hydrogen Fermentation Reactor Installed with Silicone Rubber Membrane. International Journal of Hydrogen Energy 2002; 2711571165
  120. 120. MyungS.WangY. R.Zhang-HY.P.2010Fructose-1,6 Bisphosphatase from a Hyper-Thermophilic Bacterium Thermotoga Maritime: Characterization, Metabolite Stability and its Implications. Process Biochemistry 2010; 451218821887
  121. 121. YoshidaA.NishimuraT.KawaguchiH.InuiM.YukawaH.2005Enhanced Hydrogen Production from Formic Acid by Formate Hydrogen Lyase-Overexpressing Escherichia Coli Strains. Applied Environmental Microbiology 2005; 7167626768
  122. 122. GuedonE.DesvauxM.PetitdemanageH.2002Improvement of Cellulolytic Properties of Clostridium Cellulolyticum by Metabolic Engineering. Applied Environmental Microbiology 2002; 685358
  123. 123. KanaiT.ImankaH.NakajimaA.UwamoriK.OmoriY.FukuiT.AtomiH.ImanakaT.2005Continuous Hydrogen Production by the Hyperthermophilic Archaeon, Thermococcus Kadakaraensis KOD1. Journal of Biotechnology 2005; 116271282
  124. 124. MosierN.wymanC.DaleB.ElanderR.LeeY. Y.HoltzappleM.LadischM.2005Features of Promising Technologies for Pretreatment of Lignocellulosic Biomass. Bioresource Technology 2005; 96673686
  125. 125. Mc NeffC. V.NowlanD. T.Mc NeffL. C.YanB.FedieR. L.2010Continuous Production of 5-hydroxymethylfurfural from Simple and Complex Carbohydrates. Applied Catalyst A: General 2010; 3846569
  126. 126. SuY.BrownH. M.HuangX.ZhouX. D.AmonetteJ. E.ZhangZ. C.2009Single-Step Conversion of Cellulose to 5-Hydroxymethylfurfural (HMF), A Versatile Platform Chemical. Applied Catalyst A: General 2009; 361117122
  127. 127. WangP.YuH.ZhanS.WangS.2011Catalytic Hydrolysis of Lignocellulosic Biomass into 5-Hydroxymethylfurfural in Ionic Liquid. Bioresource Technology 2011; 10241794183
  128. 128. TaoF.SongH.ChouL.2011Catalytic Conversion of Cellulose to Chemicals and Ionic Liquid. Carbohydrate Research 2011; 3465863
  129. 129. ChhedaJ. N.Roman-LeshkovY.DumesicJ. A.2007Production of 5-HydroxymethylFurfural and Furfural by Dehydration of Biomass-Derived Mono- and Poly-Saccharides. Green Chemistry 2007; 9342350
  130. 130. YinS.PanY.TanZ.2011Hydrothermal Conversion of Cellulose to 5-Hydroxymethyl Furfural. International Journal of Green Energy 2011; 8234247
  131. 131. LiY.LuX.YuanL.LiuX.2009Fructose Decomposition Kinetics in Organic acid-Enriched High Temperature Liquid Water. Biomass and Bioenergy 2009; 3311821187
  132. 132. ChareonlimkunA.ChampredaV.ShotiprulA.LaosiripojanaN.2010Reactions of C5 and C6-Sugars and Lignocellulose under Hot Compressed Water (HCW) in the Presence of Heterogeneous Acid Catalysts. Fuel 2010; 8928732880
  133. 133. MascalM.NikitinE. B.2008Direct, High-Yield Conversion of Cellulose into Biofuel. Angewandte Chemie (International Edition) 2008; 4779247926

Written By

Pedram Fatehi

Submitted: 03 November 2011 Published: 29 August 2013