Anda di halaman 1dari 16

Review

The new forestry biofuels sector


Yunqiao Pu, Dongcheng Zhang, Preet M. Singh and Arthur J. Ragauskas, Georgia Institute of Technology, USA Received October 9, 2007; revised version received November 12, 2007; accepted November 13, 2007 Published online December 19, 2007 in Wiley InterScience (www.interscience.wiley.com); DOI: 10.1002/bbb.48; Biofuels, Bioprod. Bioref. 2:5873 (2008) Abstract: Societys increasing demand for transportation fuels has assured a viable future for the development of renewable fuels. Although rst-generation biofuels are dependent on starches, sugars and vegetable oils, the need to generate higher volumes of biofuels at lower cost has shifted the research focus to cellulosic ethanol. The utilization of lignocellulosics for the sustainable manufacturing of biofuels is critically dependent on the chemical constituents of the starting biomass and the desired fuel properties. This review examines the major chemical constituents of biomass and the recent advances in their conversion to biofuels, with a special emphasis on the forest residues and woody-energy crops to bioethanol. 2007 Society of Chemical Industry and John Wiley & Sons, Ltd Keywords: cellulose; hemicellulose; lignin; wood; biorenery; biofuels; pretreatment; saccharication; fermentation; pyrolysis; gasication

Introduction
he events of the last few years have brought into sharp focus the need to develop sustainable green technologies for many of our most basic manufacturing and energy needs. Since the beginning of the new millennium, we have witnessed an ever-increasing merger of technical, economical and societal demands for sustainable technologies. Indeed, hardly a day goes by in which the issues of energy security, climate change, cradle-to-cradle product development are not discussed in public and professional forums.13 Accompanying these interests, science and engineering have made tremendous strides to begin to answer these challenges. Indeed, it is the intersection of science, business and public policy that has launched a new green, industrial revolution that promises to dramatically alter our world.4 At the cornerstone of this green industrial revolution is the integrated biorefinery.5 This is a biomass processing facility that integrates our ability to tailor biomass productivity and

processability with conversion processes, with the equipment to produce a range of fuels, power, and chemicals from biomass.6 It fully utilizes all components of biomass to make a range of foods, fuels, chemicals, feeds, materials, heat and power in proportions that maximizes sustainable, economic development. As such, this vision seeks to develop a new carbohydrate-lignin economy that will initially supplement todays petroleum economy and, as these non-renewable resources are consumed, will become the primary resource for fuels, chemicals and materials. Todays bioethanol and biodiesel plants represent the first-generation biorefineries that utilize readily processable bioresources such as sucrose, starches and plant oils.79 As has been highlighted in several reviews, societys ability to displace substantial amounts of nonrenewable petroleum reserves with renewable resources rests on its ability to secure large amounts of low-cost biomass. For example, Perlack et al., identified 1.3 billion dry tons of biomass potential/year in the USA which could be directed to biofuels production; enough to address approximately

Correspondence to: Arthur J. Ragauskas, School of Chemistry and Biochemistry, Georgia Institute of Technology, 500 10th Street NW, Atlanta, GA 30332. E-mail: arthur.ragauskas@chemistry.gatech.edu

58

2007 Society of Chemical Industry and John Wiley & Sons, Ltd

Review: New forestry biofuels sector

Y Pu et al.

one-third of current demand for transportation fuels in the United States.10 A subsequent workshop titled Breaking the biological barriers to cellulosic ethanol supported this hypothesis, and with a more aggressive research program on improving energy crops, the biomass replacement potential could be even greater.11 An analysis of the US bioresource basin suggests that approximately 30% of this biomass would originate from forest resources including: wood from forestlands, wood-related mill residues, and terrestrial urban wood residues. The exact distribution of these resources is clearly sensitive to geographical locations and the next generation of biorefineries will need to be engineered to utilize local bioresources. A unique web resource that summarizes the theoretical potential ethanol yield from biomass, including woody plants, is the US Department of Energy, Energy efficiency and renewable energy biomass program website.12 The theoretical ethanol yields for forest thinnings, hardwood sawdust and mixed paper were predicted at 81.5, 100.8 and 116.2 gallons/ton of dry feedstock, respectively.

these linear cellulose chains, which stiffens the chains and promotes aggregation into a crystalline structure. These properties give cellulose a multitude of crystalline fiber structures and morphologies. The degree of crystallinity of select cellulose samples are presented in Table 1.14 The ultrastructure of native cellulose (cellulose I) has been shown to possess an additional complexity in the form of two crystal phases: I and I.15 The relative amounts of I and I have been found to vary between samples from different origins. The I-rich specimens have been found in the cell wall of some algae and in bacterial cellulose, whereas I-rich specimens have been found in cotton, wood, and ramie fibers.16 The crystal and molecular structure of cellulose I has been examined recently by Nishiyama et al., using atomic-resolution synchrotron and neutron diff raction data recorded from cellulose isolated from alga and tunicin.17 Most native samples of cellulose also have varying degrees of amorphous cellulose, which is more reactive to chemical and enzymatic attack. Hemicellulose After cellulose, the next major polysaccharide resource is plant hemicelluloses. Unlike cellulose, hemicelluloses have lower DP values (i.e., typically 50300), frequently have side chain groups and are essentially amorphous. The main hemicelluloses of soft wood (SW) are galactoglucomannans (Fig. 2) and arabinoglucuronoxylan (Fig. 3), while in hardwood (HW) it is glucuronoxylan (Fig. 4). Table 2 summarizes

Woody biomass resources


The conversion of these bioresources to value-added materials and chemicals rests primarily on our abilities to manipulate the chemistry and biochemistry of cellulose, hemicellulose and lignin. These three biopolymers are the main global plant resources and further research is needed to efficiently convert these bio building blocks to biofuels, biochemicals, biomaterials and biopower. Cellulose Of the three bioresources, cellulose is chemically the simplest structure as it is a linear polymer of (14) glucopyranosyl (Fig. 1) with a degree of polymerization (DP) varying from ~10,000 in cotton to less than 500 in several industrially processed materials.13 The cellulose chain has a strong tendency to form intra- and inter-molecular hydrogen bonds by the hydroxyl groups on

Table 1. X-Ray crystallinity of some cellulose materials. Sample


Cotton linters Sulte dissolving pulp Prehydrolyzed sulfate pulp Viscose rayon Regenerated cellulose lm

X-ray crystallinity (%)


5663 5056 4045 2740 4045

HO O HOH2C

OH O O

HOH2C HO

O O

HO

OH O O

HOH2C HO

O O OH

OH

HOH2C

Figure 1. The structure of cellulose.

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

59

Y Pu et al.

Review: New forestry biofuels sector

HOH2C O HO

O O OH

RO OR O O OH O

HOH2C RO

O OR

RO O HOH2C OR O

R: H or Ac HO

HO

O CH2OH

Figure 2. Principal structure of galactoglucomannans in softwood.

O HO

OH O

OH O O O OH

OH O

OH O O

O
OH

O HO H3CO HOOC OH O HOH2C

OH

Figure 3. Principal structure of arabinoglucuronoxylan in softwood.


OR O O
RO

O RO

OR O

O O OR R: H or Ac

RO

OR O O

O HO H3CO HOOC OH O

Figure 4. Principal structure of glucuronoxylan in hardwood.

Table 2. The major hemicellulose components in softwood and hardwood.19,20 Composition Wood Hemicellulose type
Galacto-glucomannan SW Arabino-glucuronoxylan 710

Amount (% on wood)
1015

Units21
-D-Manp -D-Glcp -D-Galp Acetyl -D-Xylp 4-O-Me--D-GlcpA -L-Araf -D-Xylp 4-O-Me--D-GlcpA Acetyl -D-Manp -D-Glcp

Molar ratios
4 1 0.1 1 10 2 1.3 10 1 7 12 1

Linkage
14 14 16 14 12 13 14 12 14 14

~DP
100

100

Glucuronoxylan HW Glucomannan

1530

200

25

200

the main structural features of hemicelluloses appearing in common soft wood and hardwood resources.18 In addition, most sugar components can take part in the formation of lignin-carbohydrate complexes (LCC) by

covalent linkages between lignin and carbohydrates. 22,23 The most frequently suggested LCC-linkages in native wood are benzyl ester, benzyl ether, and glycosidic linkages. 24

60

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

Review: New forestry biofuels sector

Y Pu et al.

Lignin This biopolymer is an amorphous, cross-linked, and three dimensional phenolic polymer.25 The biosynthesis of lignin stems from the polymerization of three types of phenylpropane units as monolignols: coniferyl, sinapyl, and p-coumaryl alcohol.26,27 Figure 5 depicts these three structures. Soft wood lignin is composed mainly of coniferyl alcohol units, while hardwood lignin is composed mainly of coniferyl and sinapyl alcohol units. The polymerization process is initiated by an enzymecatalyzed oxidation of the monolignol phenolic hydroxyl groups to yield free radicals. A monolignol free radical can then couple with another monolignol free radical to generate a dilignol. Subsequent nucleophilic attack by water, alcohols, or phenolic hydroxyl groups on the benzyl carbon of the quinone methide intermediate restores the aromaticity of the benzene ring. The generated dilignols then undergo further polymerization to form protolignin. Although the exact structure of protolignin is unknown, improvements in methods for identifying lignin-degradation products and advancements in spectroscopic methods have

enabled scientists to elucidate the predominant structural features of lignin. Figure 6 depicts some of the common linkages found in soft wood lignin.28,29 The typical abundance of these types of linkages and functional groups in soft woods are shown in Tables 3 and 4.28 Lignin is much less hydrophilic than either cellulose or hemicelluloses and it has a general effect of inhibiting water adsorption and fiber swelling.
C C O C C C C O

C C O C O

O -O-4

O -O-4 O dibenzodioxocin

C C C

C C C

C C C

O -5 C C C

O 5-5

HO


1 6 5 2 3 4

Coniferyl alcohol/guaiacyl: R1 = OMe, R2 = H Sinapyl alcohol/ syringyl: R1 = R2 = OMe p-Coumaryl alcohol: R 1 = R2 = H

C C C

C C C

C C C

C C C

O O O - O O -1

R2

R1

4-O-5

OH

Figure 5. Three building blocks of lignin.

Figure 6. Common linkages between phenylpropane units in softwood lignin.28

Table 3. Proportions of different types of linkages connecting the phenylpropane units in softwood lignin. Linkage typea
-O-4 -5 5-5 5-5/-O-4 4-O-5 -1 -

Dimer structure
Phenylpropane -aryl ether Phenylcoumaran Biphenyl Dibenzodioxicin Diaryl ether 1,2-Diaryl propane --linked structures

Percentage
50 912 1525 1015 4 7 2

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

61

Y Pu et al.

Review: New forestry biofuels sector

Table 4. Functional groups in spruce lignin. Milled wood lignin per 100 C9 units
0.8 39 23.6 11.2 4.9

of the substrate. The effect of virgin lignin, redeposited lignin after pretreatment33 and LCC on the bioavailability of other cell-wall components is thought to play a large role in the physical restriction mechanism. Other factors have been proposed, including non-specific association between lignin and deconstruction enzymes (i.e., cellulase, xylanase, etc.).3436 Interestingly, recent studies at the National Renewable Energy Laboratory (NREL) have suggested that low levels of lignin may actually enhance cellulose hydrolysis.37 This effect has been attributed to a physical separation of microcellulose fibrils enhancing cellulase access/activity. Recent studies by Pu et al.,38 Hayashi et al.,39 and others have demonstrated that depolymerization of fibrous cellulose by cellulase exhibits selectivity toward the more reactive amorphous, paracrystalline and I forms of cellulose leaving behind a more recalcitrant crystalline form of cellulose. In contrast, other reports have suggested that the cellulose crystallinity index after hydrolysis does not change.40 Clearly, selectivity of cellulase hydrolysis and its impact on residual crystalline structure needs further investigation since it is well known that fungal cellulase hydrolysis of amorphous cellulose is 330 times faster than crystalline cellulose.41,42 The role of acetylated hemicelluloses for both soft woods and hardwoods has also been suggested to impact enzymatic deconstruction of polysaccharides.43 The efficient, cost-effective depolymerization of these polysaccharides to monosaccharides remains a key challenge in the utilization of these bioresources for fermentation to ethanol.44,45 To date, effective utilization of these bioresources is predicated on a pretreatment that reduces biomass recalcitrance.

Functional group
Carbonyl Olenic + substituted aromatic C Aliphatic CHx-OR Methoxyl Aliphatic CHx

Conversion of biomass to biofuels


As highlighted by Petruss recent renewable fuels article, the utilization of these primary bioresources for biofuels production centers about deoxygenation chemistry.30 Table 5 summarizes that cellulose, hemicelluloses and lignin are all too-rich in oxygen-function groups in comparison to typically hydrocarbon-based gasoline and diesel fuels. To date, the most successful technological route for the conversion of plant biomass to biofuels is the fermentation route, with bioethanol derived from starch and sucrose now becoming a common 510% fuel supplement. Although all biorefineries are regional and no one technology will address all needs, for many geographical locations to attain higher levels of renewable fuels this will require the utilization of lignocellulosics. This change in bioresource is based on the greater availability of lignocellulosic biomass, potential lower cost and the avoidance of the food or fuel arguments.31 Currently, the overall cost of converting lignocellulosic material to ethanol is higher than well-established commercial starch to bioethanol technologies. This is primarily due to the recalcitrance of lignocellulosics.32 Lignin is the most recalcitrant component of the plant cell wall. In general, the higher the proportion of lignin, the lower the bioavailability

Current pretreatment technologies


The objective of pretreating lignocellulosics is to alter the structure of biomass in order to make the cellulose and hemicelluloses more accessible and amenable to hydrolytic enzymes that can generate fermentable sugars. Effective pretreatment technologies need to address several important criteria, including: minimization of hemicelluloses degradation products, limiting the formation of by-products that inhibit ethanol fermentation, lignin alterations,46 minimal energy, capital and operating costs. Some of the most studied lignocellulosic/wood pretreatments are summarized below:

Table 5. General chemical composition of bioresources and petroleum.


Cellulose/starch Hemicellulose SW Lignin Gasoline Diesel [C6(H2O)5]n [C5(H2O)4]n/[C6(H2O)5]n [C10H12O4]n ~C6H14C12H26 ~C10H22 to C15H32

62

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

Review: New forestry biofuels sector

Y Pu et al.

Uncatalyzed steam explosion involves rapidly heating biomass with steam at elevated temperatures (~190240 oC) with residence times of ~38 min, followed by explosive decompression. This treatment promotes hemicellulose hydrolysis and opens up the plant cell structure, although enhanced digestability of cellulose is only weakly correlated with the physical effects.47,48 This autohydrolysis procedure has been shown to be effective with agricultural residues and hardwoods, but not as beneficial for soft woods.49,50 To improve the efficiency of this process, several additive technologies have been examined including pre-impregnation with SO251 and post-alkaline hydrogen peroxide treatment.52 Hot water autocatalyzed pretreatments at 200230 oC for up to 15 minutes can result in extensive hemicellulose hydrolysis but, the high lignin content biomass reduces subsequent cellulase hydrolysis.53 Depending on the conditions employed, 3046% of the lignin of corn stovers could be removed. The production of possible inhibitors such as furfural and hydroxymethyl furfural was reported to account for less than 3% of the original carbohydrates54 as xylan release frequently results in oligomers. Subsequent enzymatic hydrolysis of cellulose has been reported to yield glucose in 2595% yields, with the latter only being accomplished with physical milling. The application of this pretreatment to hardwoods has been reported, resulting in 90% conversion of glucose to ethanol after simultaneous saccharification and fermentation (SSF).55 Allen et al., published a comparison study of a hot-water treatment versus a dilute-acid pretreatment and both yielded comparable conversion to ethanol under optimized conditions, although the severity of the former pretreatment had to be much higher.56 Dilute acid pretreatment has been extensively studied and typically employs 0.42% H2SO4 (note: nitric, sulfur dioxide, and phosphoric acid have also been studied) at temperatures of 160220 oC to remove hemicelluloses and enhance cellulase digestion of cellulose.19 The acidic conditions used have been shown to enhance total sugar release after enzymatic hydrolysis to ~93% for corn stovers and ~82% for soft wood.57 The pretreatment conditions impact not only the plant polysaccharides but also lignin.58 For soft woods, a two-stage acidic pretreatment has been used to tailor the reactivity of cellulose and hemicellulose. This tailored

approach has been reported to increase sugar yields by 10% and reduces cellulase requirements by about 50%.59 The use of SO2 on spruce woodchips is of exceptional interest as it yields a more reactive material with less inhibitory compounds than dilute acid and this is reflected in higher ethanol yields after saccharification and fermentation.60 Recently, diethyloxylate has been reported as a potential acidic pretreatment reagent for wood and other treatments are also being developed.61 Aqueous lime or NaOH pretreatment has been shown to be effective for wheat straw and sugar bagasse with lower temperatures than acid treatments; however, the treatment times are hours long. For example, Chang et al., used lime with wheat straw at 85 oC, for 3 h.62 The use of an alkaline treatment incurs additional capital cost, as the recovery of salts requires a lime kiln to regenerate the base. The efficiency of alkaline treatments to convert recalcitrant biomass for subsequent cellulase treatments has focused on the application of a supplement oxidant, such as oxygen or hydrogen peroxide. It has been reported that this protocol dissolves the hemicelluloses, degrades lignin, and yields a cellulose fraction that is very accessible to enzymes for hydrolysis and fermentation to ethanol. An improved version of this pretreatment is the utilization of oxygen under alkaline conditions. An oxidative lime treatment63 and other wet-oxidations64 have been shown to improve the effectiveness of this pretreatment technology especially for wood-related bioresources. Ammonia pretreatment involves pretreating biomass with an aqueous solution (515%) at temperatures of 160180 oC. The ammonia reacts with lignin causing depolymerization and cleavage of select lignin-carbohydrate bonds. Agricultural residuals and herbaceous plants treated in this manner exhibit an excellent response to cellulase.65,66 Unfortunately, hardwoods and soft woods are not efficiently treated by this technology, with conversion yields of glucose to ethanol being reported to be less than 85%.67 In all cases, ammonia recovery is an additional cost and important consideration. Organosolv pretreatment of biomass resides on the use of an organic solvent system (i.e., ethanol/water,68,69 acetone/water,70 methyl isobutyl ketone/ethanol/water71) with enhanced solubilizing properties, due to the organic component. Usually, the resultant cellulosic fraction is highly

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

63

Y Pu et al.

Review: New forestry biofuels sector

susceptible to enzymatic hydrolysis, generating very high yields of glucose that can be readily converted to ethanol. Pan et al., have shown that ~88% of the cellulose could be recovered after the organosolv pretreatment of hybrid poplar and 85% was converted to glucose upon subsequent enzyme hydrolysis.72 Even better results were reported for infected lodgepole pine.73 Several authors have also indicated that the wood-based biofinery of the future will garner additional revenues from the extracted lignin and hemicellulose streams.7476 In all these and other pretreatment technologies, differences in cell-wall structure and chemistry impact how hardwoods and soft woods respond to chemical pretreatments. Several authors have indicted that the recalcitrance of soft wood resources is greater than hardwoods which is exhibited in reduced digestability by cellulase.77,78 The exact chemical constituents and ultrastructures that contribute to this effect are not well understood as recently highlighted by Mosier et al: Greater fundamental understanding of the chemical and physical mechanism that occur during pretreatment along with an improved understanding of the relationship between the chemical composition and physico-chemical structure of lignocellulosics on the enzymatic digestion of cellulose and hemicellulose is required for the generation of effective pretreatments.79 Future fundamental research into these issues promises to have a far-reaching beneficial effect in accelerating the development of low-cost biofuels.

Plant genetics, recalcitrance and future pretreatments


The need to improve the effectiveness of pretreatment technologies for wood is driven primarily by the fact that it remains the most costly step in the overall conversion of wood to biofuels. Wyman has perhaps best summarized the state-of-the-art pretreatment capabilities: The only step more expensive than pretreatment is no pretreatment, because of its impact on virtually all other operations.80 In light of the well-known dependency of biomass recalcitrance on the plant resource, it is natural to consider the opportunity of reducing the recalcitrance of wood and other biomass via the genetic engineering of the biomass. Indeed, the forest products industry has extensively championed the use of plant genetics to tailor the composition, structure and

reactivity of soft wood and hardwood biopolymers, especially lignin. For example studies by Chiang et al., have inserted antisense 4CL and sense coniferaldehyde 5-hydroxylase genes into aspen to yield trees with each or both of these transgenes. Introduction of the former gene reduced lignin concentrations by 55% and the latter gave up to a three-fold increase in syringyl: guiacyl lignin.81 Huntley et al., reported that increased syringyl-lignin in transgenic poplars, by overexpressing F5H, increased chemical pulp ability by 60%.82 Likewise, Pilate et al., demonstrated that transgenic poplar with low CAD activity exhibited improved kraft pulping properties.83 These results highlight the potential to alter specific biopolymer constituents in woody plants which confer benefits in subsequent chemical operations such as kraft and soda pulping. It is reasonable to anticipate that as our knowledge of the beneficial physical-chemical impacts of pretreatments on the plant cell wall is developed, it will be possible to genetically engineer low-recalcitrance wood. For example, reduced lignin content, modifications in cellulose crystallinity, differing hemicellulose structures and reduced lignin-carbohydrate complexes have all been shown to decrease plant recalcitrance and it should be possible to engineer these same properties into woody plants and other bioresources.84,85 A recent report by Davison et al.,86 has demonstrated this approach, since changes in lignin content and syringylguaiacyl ratios of a second-generation Populus significantly benefited xylose release upon dilute sulfuric acid hydrolysis. Chen and Dixon have also reported comparable results for the acid hydrolysis of a series of alfalfa lines containing antisensing constructs for downregulating lignin.87 In brief, the lines with reduced lignin content released greater amounts of carbohydrates during acid pretreatment and in subsequent enzymatic hydrolysis. These results indicate that genetic control of lignin content and composition influences the hydrolyzability of the biomass and sets the stage for further developments. Given recent advances in plant genomics, it is anticipated that engineered changes in plant cell structure will yield lowrecalcitrant, highly productive agro-energy crops in the near future. This will have a dramatic impact on pretreatment technologies reducing the severity, capital and operating costs of this key stage in the conversion of biomass to biofuels.

64

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

Review: New forestry biofuels sector

Y Pu et al.

Cellulosic saccharication-fermentation technologies


Current: Following pretreatment, woody biomass can be converted into simple sugars by enzymatic deconstruction via a cellulase treatment. This remains the second most expensive component in the bioconversion of wood to bioethanol, despite the fact that research studies over the past decade have decreased cellulase costs by greater than a ten-fold basis.88 Numerous publications and reviews have highlighted the use of (i) separate hydrolysis and fermentation (SHF) and (ii) simultaneous saccharification and fermentation (SSF) to convert pretreated wood to ethanol.89,90 Ewanick et al., have reported that employing SSF on a SO2-steam exploded lodgepole pine provided 6177% yield of the theoretical maximum ethanol yield depending upon the severity of the pretreatment.91 A two-stage acid treatment of spruce using SO2 or H2SO4 was reported to give ethanol yields of 81 and 70% respectively (i.e., 94 and 79 gallons/dry ton).60,92 Comparable SSF ethanol yields have also been reported for an organosolv pretreated mixed soft wood furnish.93 The conversion of hardwoods to ethanol has also been extensively studied. SSF treatment of steam-exploded poplar and eucalyptus has been reported to provide 71 and 62% of the maximum theoretical yield of ethanol from glucose.94 Higher SSF ethanol yields of 90% from cellulose have been reported for an acidic hot-water treatment of yellow poplar, provided that the solids were washed with hot water to remove solubilized lignin.95 The role of inhibitors formed during steam explosion of poplar on SSF has been extensively studied. Undetoxified pretreated wood was reported to yield no ethanol even with high loadings of Saccharomyces cerevisiae, whereas water-rinsed biomass provided an 82% yield of ethanol.96 Bari et al., have also reported bioethanol yields of 85% from steam-exploded97 aspen chips and more recently demonstrated that SO2 impregnation can enhance this pretreatment technology.98 Although most authors have tailored their SSF stage to their exact bioresource, Berlin et al., have reported that cellulase treatments performed optimally on hardwood also exhibit superior performance on soft wood substrates.99 A promising approach to reducing cellulase cost is to capture and reuse the enzymes. For example, Tu et al., have shown that 51% of the applied cellulases could be recovered

by re-adsorption onto fresh lignocellulosic materials.100 In addition, it is well known that the lignin fraction in pretreated lignocellulosics is involved in unproductive binding to cellulosic enzymes that reduces the performance of the enzymes. The development of additives including proteins101 and surfactants102 that disrupt this association has been shown to enhance the efficiency of deconstruction enzymes. Future: A process challenge in the conversion of wood to biofuels is the efficient conversion of all wood sugars (i.e., C5 and C6) to ethanol, especially for hardwoods which have greater amounts of pentoses. Microorganisms that are able to ferment sugars to ethanol can be either yeasts103,104 or bacteria.105 Over the past decades, new methods in molecular biology, protein chemistry and genetic engineering have led to an increasing number of new strains, exhibiting improved characteristics to ferment the full spectrum of sugars available in hydrolyzates.106,107 One promising strategy has been to take a natural hexose ethanologen and add the pathways to convert other sugars. This strategy has been effective in adding pentose conversion to Saccharomyces cerevisiae, and to Zymomonas mobilis.108,109 These enhancements promise to further enhance the overall fermentation of mixed solutions of hexoses and pentoses to ethanol.110112 Although research studies over the past decade have decreased cellulase cost by greater than a ten-fold basis,113 they still remain a significant cost for SSF and SHF. An alternative approach to minimize the cost of cellulose deconstruction and conversion to ethanol is consolidated bioprocessing (CBP). CBP involves (i) bioproduction of cellulolytic enzymes from thermophilic anaerobic microbes, (ii) hydrolysis of plant polysaccharides to simple sugars and (iii) their subsequent fermentation to ethanol all in one stage.114 This bioprocess is projected to reduce the cost of bioethanol by a factor of four over SSF and these reduced costs and simplicity of operation have heightened research in this field. To date, the penultimate CBP system has not been developed but the basic pathways that need to be developed have been reported and research is ongoing.115,116

Non-biological pathways of converting biomass to biofuels


Although the biological route for converting biomass to biofuels is one of the most developed and promising

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

65

Y Pu et al.

Review: New forestry biofuels sector

technologies, several other competing processes are also being explored and developed. As an alternative approach, the production of a bio-oil from biomass by pyrolysis certainly is of the most direct methods of liquidifying natures key bioresources. Typically, this process can be accomplished with a conventional slow pyrolysis reaction involving a reactor temperature of ~500 oC and a vapor residence time of 530 min, or fast pyrolysis conditions involving a temperature range of 425500 oC with a very short vapor resident time <2 s.117 Both of these conditions employ an oxygen-depleted or oxygen-free atmosphere which leads to the fragmentation of lignin and plant polysaccharides. The resulting bio-oil has been shown to contain +200 constituents including substituted phenols, catechols, pyrones, oligomers of lignin, hydroxyaldehydes/ketones, sugars, carboxylic acids and water.118,119 Although the application of bio-oils as a transportation fuel has been reported, the complex/variable nature of these oils, their instability, corrosivity, and relatively high water content have limited their commercial applications.120122 In an attempt to address some of these limitations, several investigators have examined the application of catalysts on pyrolysis products. For example, Williams et al., reported the use of a ZSM-5 zeolite catalyst in the downstream section of a pyrolysis reactor and demonstrated that the pyrolysis oils were markedly less oxygenated due to dehydration, decarboxylation and decarbonylation reactions.123 A comparable effect was reported when fast pyrolysis vapors were passed through mesoporous catalysts.124,125 The use of zeolites for upgrading pyrolysis bio-oils has been reported to be complicated by the deactivation of the catalyst by reversible coke formation and irreversible dealumination.126 Clearly, the search for improved pyrolysis oils by catalytic upgrading provides much promise but needs to address catalytic activity and stability under the conditions employed. An alternative approach to generate fuels from plantderived sugars is based on a series of controlled catalytic dehydration-hydrogenation reactions which has the ability to convert C6/C5 sugars to their corresponding hydrocarbons.127 Of special significance to this aqueous-phase reforming approach is that the hydrogen needed for hydrogenation is produced in-situ from the polyol over a catalyst. The incorporation of additional condensation reactions, such

as Aldo-condensations into this refining pathway provides for higher molecular weight hydrocarbons (i.e., C9C15) that could be considered as viable components for gasoline or diesel, depending on the reaction conditions employed.128 These recent developments highlight the potential that catalytic refining of biomass to a biogasoline/diesel resource could provide an alternative route to renewable fuels.129 The final route for converting biomass to biofuels is via gasification converting plant biomass to a stream of H2 and CO (syngas) which can be used for the production of Fischer Tropsch hydrocarbons. Production of syngas from biomass is usually accomplished in a gasifier at temperatures between 800 and 1000 oC under 2030 bar.130 As an alternative to thermal treatment, a pyrolysis plasma of wood has been recently developed to yield a syngas stream.131 Although gasification technologies for biomass are reasonably well developed, the technical challenges reside in the Fisher Tropsch catalyzed reactions which are prone to poisoning from alkali metals, halides, nitrogen and sulfur gases sufficiently abundant in most biomass resources.132 In addition, the propensity to form tars during biomass gasification provides additional difficulties and leads to inefficiencies in the overall process. The most common approach to reduce tar formation is either a thermal or catalytic treatment.133135 The remaining contaminants in a biomass syngas need to be removed using scrubbers, sorbents, oxidative treatments or catalytic decomposition132 and each of these technologies adds further cost to the overall process. Once these impurities are removed, Fischer-Tropsch chemistry can efficiently convert syngas to aliphatic straight-chain hydrocarbons, primary alcohols, branched and unsaturated hydrocarbons. The large hydrocarbons can then be hydrocracked to form a high-quality diesel fuel.

Integrated pulp-biofuel bioreneries


An interesting alternative to the development of a virgin wood-based biofuel refi nery is to incorporate these manufacturing concepts into an existing pulp mill, as depicted in Fig. 7. This approach has several attractive features including access to established wood resources, wood-handling equipment, transportation systems, environmental permitting issues and access to a skilled workforce knowledgeable in the

66

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

Review: New forestry biofuels sector

Y Pu et al.

operations.144 Paper-mill sludge typically has a negative value as it needs to be properly landfi lled and hence is an attractive resource for SSF conversion to bioethanol.145,146 Furthermore, it has been well documented that this low-cost bioresource does not need a pretreatment prior to SSF but the presence of minerals, contaminants and difficulties in mixing paper-mill sludge provide additional complications to the overall process. Nonetheless, a recent study by Fan and Lynd suggested that a viable SSF process could be developed yielding a +15% internal rate of fi nancial return which provides a viable treatment option for the ~5 million tons of paper mill sludge generated annually in the USA.147
Figure 7. The scheme for integrated wood-based biorenery.

Tomorrows forest biorenery


processing of wood. The development of the next generation pulp-biofuel mills is being actively investigated on several fronts.18 For kraft pulp mills, two high-priority opportunities center about the next generation of chemical recovery operations and the transition from the conventional Thompson recovery furnace technologies to gasification of black liquor which could yield a syngas process stream.136138 Alternatively, ongoing studies have highlighted the potential to extract select hemicelluloses from woodchips prior to kraft pulping. It is well known that kraft pulping conditions extract select hemicelluloses which do not contribute to final pulp properties as summarized in Table 6. These extractable hemicelluloses could provide a valuable, high-volume resource of sugars for bioethanol production generating ~2040 million gallons ethanol/year/mill.140 Thorp has reported that the potential annual production of ethanol from pre-extraction of hemicellulose could approach 2 billion gallons of ethanol/year.141 Recent studies suggest that a pre-extraction of northern hardwood benefits kraft pulping whereas a soft wood furnish suffers from yield losses which needs to be addressed.142,143 An alternative resource for bioethanol production is to utilize waste cellulosic streams from paper recycling The practical application of the science and engineering associated with converting wood to biofuels is a rapidly moving target that will require constant updating. Nonetheless, near the end of 2007 several notable industrial developments have been announced. In the USA, the Department of Energy recently announced an investment of up to $385 million for six biorefinery projects with an industry cost share of more than $1.2 billion.148 When fully operational, these biorefineries are expected to produce more than 130 million gallons of cellulosic ethanol/year. Three of these plants have announced the utilization of wood as a bioresource converting it to bioethanol via thermochemical or biological routes, including: (1) ALICO, Inc., Florida will produce 13.9 million gallons of ethanol/year with a proposed 770 tons/day feedstock from yard, wood, vegetative wastes and eventually energy cane. (2) BlueFire Ethanol, Inc., California will site a biorefi nery on an existing landfi ll and produce about 19 million gallons of ethanol/year. The proposed plant will consume 700 tons/day feedstock of sorted green waste and wood waste from landfi lls.

Table 6. Changes in carbohydrate distribution before and after kraft pulping loblolly pine.139 Source
Wood Kraft Pulp

Glucose
67.9 84.9

Galactose
3.5 0.3

Mannose
17.7 7.1

Arabinose
2.1 0.5

Xylanose
8.8 7.1

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

67

Y Pu et al.

Review: New forestry biofuels sector

(3) Range Fuels, Georgia has begun constructing a plant that will produce about 40 million gallons of ethanol/year and 9 million gallons/year of methanol. The plant will use 1,200 tons/day of wood residues and wood-based energy crops. The application of thermal and biological technologies to convert wood to ethanol clearly suggests that both technology platforms are viable, each with their own unique strengths and concerns. It is anticipated that the research developments described earlier in the review will favorably impact these commercial developments. Meanwhile, the Tembec Temiscaming sulfite mill in Quebec is a modern example of how a pulp mill can grow into a biorefinery.149 Along with pulp production, the mill ferments spent sulfite cooking liquors with Saccharomyces cerevisiae to produce 4 million gallons/ year of food-grade ethanol. It has installed an anaerobic biogas unit that displaces ~80% of the natural gas required for high-yield pulp flash drying and produces lignosulfonates for commercial markets. Although the same approach is much more technically challenging for a kraft pulp mill, its potential has been noted by several industry leaders.150 Another example is the Flambeau River paper mill in Wisconsin that has announced a partnership with American Process Inc., for a cellulosic ethanol biorefinery. The biorefinery project will be designed to produce 20 million gallons of cellulosic ethanol/year from the mills spent pulping liquor. 151 Xethanol Corporation has reported that it has acquired a former medium-density fiberboard factory which it plans to re-open as a pilot plant to demonstrate the technical and economic viability of using wood chips for the production of cellulosic feedstock.152 Additional announcements of research consortiums and pilot plant developments targeted at utilizing waste streams from virgin and recycled pulp mills along with wood residues occur virtually on a monthly basis on the international scene.153 Recent improvements in biorefinery processing technology, energy costs and favorable government policy will only accelerate these business developments in the forest products industry.154

biorefinery to the end-user also needs attention. Pipelines are, by far, the most cost-effective means of transporting large quantities of fuel over long distances, whereas tankers are used to transport fuels, including ethanol, over short distances such as from small biorefineries to storage and distribution centers. Use of existing pipeline infrastructure, presently used to transport gasoline products, as well as new, dedicated pipelines may be considered for ethanol transportation. Existing gasoline pipelines are made out of carbon steel. Corrosion and stress-corrosion cracking of carbon steel structures, especially pipeline steel, are other concerns for ethanol storage and transportation. A 2003 survey of industry, reported by the American Petroleum Institute (API Technical Report 939-D),155 indicates that carbon steel may undergo stress corrosion cracking (SCC) in certain ethanol environments. This is not a widespread concern as the cracks have only been observed primarily in user terminals exposed to ethanol products, but not in ethanol producer tanks, in rail/tank car/shipping transportation, or in end-user systems (e.g., gas tanks). Preliminary studies have shown that certain minor constituents may affect SCC behavior of carbon steel. For example, the presence of oxygen and the aging of fuelgrade ethanol have been reported to increase SCC susceptibility of carbon steel.156,157 However, effects of ethanol source or aging, and resulting differences in the minor constituents, on corrosion and stress corrosion cracking of carbon steels are not very clear. Further work is needed to understand corrosion mechanisms and to identify fuel-grade ethanol environments that may cause SCC. This will help us find mitigation strategies where some chemical additives may be used or alternative materials may be selected for transportation and storage structures to reliably distribute ethanol.

Summary and conclusions


In closing, after an extended period of low energy costs and differing research priorities, a near global emphasis on renewable biofuels technologies has evolved in the new millennium. Although differing social, environmental and economic issues have elevated these needs, there is no denying this new challenge. Furthermore, advances in plant genomics, biotechnology, nanotechnology, catalysis, material science, life-cycle analysis and computational modeling suggest that advances in the field of renewable

Transportation of bioethanol
With anticipated widespread usage of bioethanol, an efficient and reliable transportation and distribution system from

68

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

Review: New forestry biofuels sector

Y Pu et al.

biofuels will progress at rates unattainable in past decades. As these technological advances leave the laboratory and impact commercial practices, it will bring to the forefront Morris vision of the new carbohydrate-economy in which major industrial sectors are dependent on the sustainable utilization of biomass, in harmony with global agricultural production.158 Acknowledgements The authors acknowledge the support of key sponsors including NSF Performance for Innovation Program (Award # EEC0525746) and National Research Initiative of the USDA Cooperative State Research, Education and Extension Service, grant number 2003-35504-13620. References
1. Holdren JP, Energy and sustainability. Science 315:737 (2007). 2. Chow J, Kopp RJ and Portney PR, Energy resources and global development. Science 302:15281531 (2003). 3. Hoffert MI, Caldeira K, Benford, G, Criswell DR, Green C, Herzog H, Jain AK, Kheshgi HS, Lackner KS, Lewis JS, Lightfoot HD, Manheimer W, Mankins JC, Mauel ME, Perkins LJ, Schlesinger ME, Volk T and Wigley TML, Advanced technology paths to global climate stability: energy for a greenhouse planet. Science 298:981987 (2002). 4. Grassian VH, Meyer G, Abruna H, Coates GW, Achenie LE, Allison T, Brunschwig B, Ferry J, Garcia-Garibay M, Gardea-Torresday J, Grey CP, Hutchison J, Li CJ, Liotta C, Ragauskas A, Minteer S, Mueller K, Roberts J, Sadik O, Schmehl R, Schneider W, Selloni A, Stair P, Stewart J, Thorn D, Tyson J, Voelker B, White JM and Wood-Black F, Chemistry for a sustainable future. Environ Sci Technol 41:48404846 (2007). 5. Clark JH, Budarin V, Deswarte FEI, Hardy JJE, Kerton FM, Hunt AJ, Luque R, Macquarrie DJ, Milkowski K, Rodriguez A, Samuel O, Tavener SJ, White RJ and Wilson AJ, Green chemistry and the biorenery: A partnership for a sustainable future. Green Chem 8:853860 (2006). 6. Ragauskas AJ, Williams CK, Davison BH, Britovsek G, Cairney J, Eckert CA, Frederick WJ Jr, Hallett JP, Leak DJ, Liotta CL, Mielenz JR, Murphy R, Templer R and Tschaplinski T, The path forward for biofuels and biomaterials. Science 311:484489 (2006). 7. Bothast RJ and Schlicher MA, Biotechnological processes for conversion of corn into ethanol. Appl Microbiol Biotechnol 67:1925 (2005). 8. Gray KA, Cellulosic ethanol state of the technology. Int Sugar J 109:150151 (2007). 9. Marchetti JM, Miguel VU and Errazu AF, Possible methods for biodiesel production. Renew Sust Energ Rev 11:13001311 (2007). 10. Perlack R, Wright LL, Turhollow AF, Graham RL, Stokes BJ and Erbach DC, Biomass as feedstock for bioenergy and bioproducts industry: The technical feasibility of a billion-ton annual supply, 2005, http:// feedstockreview.ornl.gov/pdf/billion_ton_vision.pdf 11. Breaking the biological barriers to cellulosic ethanol: a joint research agenda A research roadmap resulting from the biomass to biofuels

workshop Rockville, Maryland, 2006, http://genomicsgtl.energy.gov/ biofuels/2005workshop/b2blowres63006.pdf 12. http://www1.eere.energy.gov/biomass/printable_versions/ethanol_yield_ calculator.html 13. Klemm D, Heublein B, Fink HP and Bohn A, Cellulose: fascinating biopolymer and sustainable raw material. Angew Chem Int Ed 44:33583393 (2005). 14. Klemn D, Philipp B, Heinze T, Heinze U and Wagenknecht W, Comprehensive Cellulose Chemistry, Vol. 1. Fundamentals and analytical methods, Wiley-VCH, Weinheim (1998). 15. Kadla JF and Gilbert RD, Cellulose structure: a review. Cell Chem Technol 34:197216 (2001). 16. Atalla RH and VanderHart DL, Native cellulose: a composite of two distinct crystalline forms. Science 223:283285 (1984). 17. (i) Nishiyama Y, Sugiyama J, Chanzy H and Langan P, Crystal structure and hydrogen bonding system in cellulose I from synchrotron X-ray and neutron brious diffraction. J Am Chem Soc 125:1430014306 (2003) (ii) Nishiyama Y, Sugiyama J, Chanzy H and Langan P, Crystal structure and hydrogen bonding system in cellulose I from synchrotron X-ray and neutron brious diffraction. J Am Chem Soc 124:90749082 (2002). 18. Ragauskas AJ, Nagy M, Kim DH, Eckert CA, Hallett JP and Liotta CL, From wood to fuels: Integrating biofuels and pulp production. Ind Biotechnol 2:5565 (2006). 19. Willfoer S, Sundberg A, Pranovich A and Holmbom B, Polysaccharides in some industrially important hardwood species. Wood Sci Technol 39:601617 (2005). 20. Willfoer S, Sundberg A, Hemming J and Holmbom B, Polysaccharides in some industrially important softwood species. Wood Sci Technol 39:245257 (2005). 21. See http://www.chem.qmul.ac.uk/iupac/ 22. Barakat A, Winter H, Rondeau-Mouro C, Saake B, Chabbert B and Cathala B, Studies of xylan interactions and cross-linking to synthetic lignins formed by bulk and end-wise polymerization: a model study of lignin carbohydrate complex formation. Planta 226: 267281 (2007). 23. Bunzel M, Ralph J, Lu F, Hateld RD and Steinhart H, Lignins and ferulate-coniferyl alcohol cross-coupling products in cereal grains. J Agric Food Chem 52:64966502 (2004). 24. Lawoko M, Henriksson G and Gellerstedt G, Characterization of lignincarbohydrate complexes (LCCs) of spruce wood (Picea abies L) isolated with two methods. Holzforschung 60:156161 (2006). 25. Davin LB and Lewis NG, Lignin primary structures and dirigent sites. Curr Opin Biotechnol 16:407415 (2005). 26. Boerjan W, Ralph J and Baucher M, Lignin biosynthesis. Annu Rev Plant Biol 54:519546 (2003). 27. Pu Y, Jiang N and Ragauskas AJ, Ionic liquid as a green solvent for lignin. J Wood Chem Technol 27: 2333 (2007). 28. Chakar FS and Ragauskas AJ, Review of current and future softwood kraft lignin process chemistry. Ind Crop Prod 20: 131141 (2004). 29. Pu Y, Anderson S, Lucia L and Ragauskas AJ, Investigation of the photooxidative chemistry of acetylated softwood lignin. J Photochem Photobio A 163: 215221 (2004). 30. Petrus L and Noordermeer MA, Biomass to biofuels a chemical perspective. Green Chem 8:861867 (2006).

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

69

Y Pu et al.

Review: New forestry biofuels sector

31. Dale BE, Cellulosic ethanol and sustainability: There is no food vs fuel conict, 233rd ACS National Meeting, Chicago, IL (2007). 32. Lynd LR, Wyman CE and Gerngross TU, Biocommodity engineering. Biotechnol Progr 15:777793 (1995). 33 Selig MJ, Viamajala S, Decker SR, Tucker MP, Himmel ME and Vinzant TB. Deposition of lignin droplets produced during dilute acid pretreatment of maize stems retards enzymatic hydrolysis of cellulose. Biotechnol Progr ACS ASAP. 34. Berlin A, Balakshin M, Gilkes N, Kadla J, Maximenko V, Kubo S and Saddler J, Inhibition of cellulase xylanase and -glucosidase activities by softwood lignin preparations. J Biotechnol 125:198209 (2006). 35. Pan X, Xie D, Gilkes N, Gregg D J and Saddler JN, Strategies to enhance the enzymatic hydrolysis of pretreated softwood with high residual lignin content. Appl Biochem Biotechnol 121124:10691079 (2005). 36. Yang B and Wyman CE, BSA treatment to enhance enzymatic hydrolysis of cellulose in lignin containing substrates. Biotechnol Bioeng 94:611617 (2006). 37. Davis MF, Ishizawa C, Jeoh T, Adney WS, Himmel ME and Johnson DK, Chemical and physical properties of pretreated biomass that affect enzyme accessibility and digestibility 233rd ACS National Meeting, Chicago, IL (2007). 38. Pu Y, Ziemer C and Ragauskas AJ, CP/MAS 13C NMR analysis of cellulase treated bleached softwood kraft pulp. Carbohydr Res 341: 591597 (2006). 39. Hayashi N, Kondo T and Ishihara M, Enzymatically produced nanoordered short elements containing cellulose I crystalline domains. Carbohydr Polym 61:191197 (2005). 40. Manseld SD, Mooney C and Saddler JN, Substrate and enzyme characteristics that limit cellulose hydrolysis. Biotechnol Progr 15: 804816 (1999). 41. Lynd LR, Weimer PJ, van Zyl WH and Pretorius IS, Microbial cellulose utilization: Fundamentals and biotechnology. Microbiol Molecular Biol R 66(3):506577 (2002). 42. Zhang YHP, Cui J, Lynd LR and Kuang LR, A transition from cellulose swelling to cellulose dissolution by o-Phosphoric acid: evidence from enzymatic hydrolysis and supramolecular structure. Biomacromolecules 7:644648 (2006). 43. Kim S and Holtzapple MT, Effect of structural features on enzyme digestibility of corn stover. Biores Technol 97:583591(2005). 44. Schell DJ, Riley CJ, Dowe N, Farmer J, Ibsen KN, Ruth MF, Toon ST and Lumpkin RE, A bioethanol process development unit: initial operating experiences and results with a corn ber feedstock. Biores Technol 91:179188 (2004). 45. Eggeman Tim and Elander Richard T, Process and economic analysis of pretreatment technologies. Biores Technol 96:20192025 (2005). 46. Yang B, Gray MC, Liu C, Lloyd TA, Stuhler SL, Converse AO and Wyman CE, Unconventional relationships for hemicellulose hydrolysis and subsequent cellulose digestion, in Lignocellulose Biodegradation, ACS Symposium Series 889, ed by Saha BS and Hayashi K. American Chemical Society, Washington, pp.100125 (2004). 47. Glasser WG and Wright RS, Steam-assisted biomass fractionation II fractionation behavior of various biomass resources. Biomass Bioenerg 14:219235 (1998).

48. Ballesteros I, Negro, MJ, Oliva JM, Cabanas A, Manzanares P and Ballesteros M, Ethanol production from steam-explosion pretreated wheat straw. Appl Biochem Biotechnol 129132: 496508 (2006). 49. Negro MJ, Manzanares P, Oliva JM, Ballesteros I and Ballesteros M, Changes in various physical/chemical parameters of Pinus pinaster wood after steam explosion pretreatment. Biomass Bioenerg 25:301308 (2003). 50. Cantarella M, Cantarella L, Gallifuoco A, Spera A and Alfani F, Effect of inhibitors released during steam-explosion treatment of poplar wood on subsequent enzymatic hydrolysis and SSF. Biotechnol Progr 20:200206 (2004). 51 Boussaid AL, Esteghlalian AR, Gregg DJ, Lee KH and Saddler JN, Steam pretreatment of douglas-r wood chips: can conditions for optimum hemicellulose recovery still provide adequate access for efcient enzymatic hydrolysis? Appl Biochem Biotechnol 8486:693705 (2000). 52. Cara C, Ruiz E, Ballesteros I, Negro MJ and Castro E, Enhanced enzymatic hydrolysis of olive tree wood by steam explosion and alkaline peroxide delignication. Process Biochem 41:423429 (2006). 53. Liu C and Wyman CE, Impact of uid velocity on hot water only pretreatment of corn stover in a owthrough reactor. Appl Biochem Biotechnol 113116:977987 (2004). 54. Allen SG, Spencer MJ, Antal MJ Jr, Laser MS and Lynd LR, Hot liquid water pretreatment of lignocellulosics at high solids concentrations, in Developments in Thermochemical Biomass Conversion, Vol.1, ed by Bridgwater AV and Boocock DGB. Blackie, London, pp. 765772 (1997). 55. van Walsum GP, Allen SG, Spencer MJ, Laser MS, Antal MJ and Lynd LR, Conversion of lignocellulosics pretreated with liquid hot water to ethanol. Appl Biochem Biotechnol 57/58:157170 (1996). 56. Allen SG, Schulman D, Lichwa J, Antal MJ Jr, Jennings E and Elander R, A comparison of aqueous and dilute-acid single-temperature pretreatment of yellow poplar sawdust. Ind Eng Chem Res 40(10): 23522361 (2001). 57. Lloyd TA and Wyman CE, Combined sugar yields for dilute sulfuric acid pretreatment of corn stover followed by enzymatic hydrolysis of the remaining solids. Biores Technol 96:19671977 (2005). 58. Yang B, Gray MC, Liu C, Lloyd TA, Stuhler SL, Converse AO and Wyman CE Unconventional relationships for hemicellulose hydrolysis and subsequent cellulose digestion, in Lignocellulose Biodegradation, ACS Symposium Series 889, ed by Saha BS, Hayashi K. American Chemical Society, Washington, pp. 100125 (2004). 59. Kim SB and Lee YY, Diffusion of sulfuric acid within lignocellulosic biomass particles and its impact on dilute-acid pretreatment. Biores Technol 83:165171 (2002). 60. Soederstroem J, Galbe M and Zacchi G, Separate versus simultaneous saccharication and fermentation of two-step steam pretreated softwood for ethanol production. J Wood Chem Technol 25:187202 (2005). 61. Kenealy W, Horn E, Davis M, Swaney R and Houtman C, Vaporphase diethyl oxalate pretreatment of wood chips: Part 2 Release of hemicellulosic carbohydrates. Holzforschung 61:230235 (2007). 62. Chang VS, Nagwani M and Holtzapple MT, Lime pretreatment of crop residues bagasse and wheat straw. Appl Biochem Biotechnol 74:135159 (1998). 63. Chang VS, Nagwani M, Kim CH and Holtzapple MT, Oxidative lime pretreatment of high-lignin biomass: poplar wood and newspaper. Appl Biochem Biotechnol 94:128 (2001).

70

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

Review: New forestry biofuels sector

Y Pu et al.

64. Lissens G, Klinke H, Verstraete W, Ahring B and Thomsen AB, Wet oxidation pre-treatment of woody yard waste: parameter optimization and enzymatic digestibility for ethanol production. J Chem Technol Biotechnol 79:889895 (2004). 65. Kim TH and Lee YY, Pretreatment and fractionation of corn stover by ammonia recycle percolation process. Biores Technol 96:20072013 (2005). 66. Teymouri F, Laureano-Perez L, Alizadeh H, Dale BE, Optimization of the ammonia ber explosion (AFEX) treatment parameters for enzymatic hydrolysis of corn stover. Bioresour Technol 96:20142018 (2005). 67. Oh KK; Kim YS, Yoon HH, Tae, BS, Pretreatment of lignocellulosic biomass using combination of ammonia recycled percolation and diluteacid process. J Ind Eng Chem 8:6470 (2002). 68. Arato C, Pye EK and Gjennestad G, The lignol approach to biorening of woody biomass to produce ethanol and chemicals. Appl Biochem Biotechnol 121124: 871882 (2005). 69. Pan X, Kadla JF, Ehara K, Gilkes N and Saddler JN, Organosolv ethanol lignin from hybrid poplar as a radical scavenger: relationship between lignin structure extraction conditions and antioxidant activity. J Agric Food Chem 54:58065813 (2006). 70. Hasegawa I, Tabata K, Okuma O and Mae K, New pretreatment methods combining a hot water treatment and water/acetone extraction for thermo-chemical conversion of biomass. Energy Fuels 18:755760 (2004). 71. Bozell JJ, Black SK and Myers M, Clean fractionation of lignocellulosics a new process for preparation of alternative feedstocks for the chemical industry, 8th International Symposium on Wood and Pulping Chemistry, Helsinki, Finland Vol. 1, pp. 697704 (1995). 72. Pan X, Gilkes N, Kadla J, Pye K, Saka S, Gregg D, Ehara K, Xie D, Lam D and Saddler J, Bioconversion of hybrid poplar to ethanol and Co-products using an organosolv fractionation process: optimization of process yields. Biotechnol Bioeng 94: 851861 (2006). 73. Pan X, Xie D, Yu RW, Lam D and Saddler JN, Pretreatment of lodgepole pine killed by mountain pine beetle using the ethanol organosolv process: fractionation and process optimization. Ind Eng Chem Res 46:26092617 (2007). 74. Pan X, Arato C, Gilkes N, Gregg D, Mabee W, Pye K, Xiao Z, Zhang X and Saddler J, Biorening of softwoods using ethanol organosolv pulping: preliminary evaluation of process streams for manufacture of fuel-grade ethanol and co-products. Biotechnol Bioeng 90:473481 (2005). 75. Pye EK, Industrial lignin production and applications, in Bioreneries, Industrial Processes and Products, Vol. 2, ed by Kamm B, Gruber PR and Kamm M. Wiley-VCH, Weinheim, pp.165200 (2006). 76. Mabee WE, Gregg DJ, Arato C, Berlin A, Bura R, Gilkes N, Mirochnik O, Pan X, Pye EK and Saddler JN, Updates on softwood-to-ethanol process development. Appl Biochem Biotechnol 129132: 5570 (2006). 77. Zhang Y-HP, Ding S-Y, Mielenz JR, Cui J-B, Elander RT, Laser M, Himmel ME, McMillan JR and Lynd LR, Fractionating recalcitrant lignocellulose at modest reaction conditions. Biotechnol Bioeng 97(2): 214223(2007). 78. Galbe M and Zacchi G, A review of the production of ethanol from softwood. Appl. Microbiol Biotechnol 59:618628 (2002). 79. Mosier N, Wyman C, Dale B, Elander R, Lee YY, Holtzapple M and Ladisch M, Features of promising technologies for pretreatment of lignocellulosic biomass. Biores Technol 96: 673686 (2005).

80. Wyman CE, What is (and is not) vital to advancing cellulosic ethanol. Trends Biotechnol 25:153157 (2007). 81. Chiang VL, Monolignol biosynthesis and genetic engineering of lignin in trees, a review. Environ Chem Lett 4:143146 (2006). 82. Huntley SK, Ellis D, Gilbert M, Chapple C and Manseld SD, Signicant increases in pulping efciency in C4H- F5H-transformed poplars: improved chemical savings and reduced environmental toxins. J Agr Food Chem 51:61786183 (2003). 83. Pilate G, Guiney E, Holt K, Petit-Conil, M, Lapierre C, Leple JC, Pollet B, Mila I, Webster EA, Marstorp HG, Hopkins DW, Jouanin L, Boerjan W, Schuch W, Cornu D and Halpin C, Field and pulping performances of transgenic trees with altered lignication. Nat Biotechnol 20:607612 (2002). 84. Jeoh T, Ishizawa CI, Davis MF, Himmel ME, Adney WS and Johnson DK, Cellulase digestibility of pretreated biomass is limited by cellulose accessibility. Biotechnol Bioeng 98:112122 (2007). 85. Himmel ME, Ding SY, Johnson DK, Adney WS, Nimlos MR, Brady JW and Foust TD, Biomass recalcitrance: engineering plants and enzymes for biofuels production. Science 315:804807 (2007). 86. Davison BH, Drescher SR, Tuskan GA, Davis MF and Nghiem NP, Variation of S/G ratio and lignin content in a populus family inuences the release of xylose by dilute acid hydrolysis. Appl Biochem Biotechnol 129132:427435 (2006). 87. Chen F and Dixon RA, Lignin modication improves fermentable sugar yields for biofuel production. Nat Biotechnol 25:759761 (2007). 88. See http://www.eere.energy.gov/ biomass/cellulase_cost.html 89. Wyman CE, Ethanol from lignocellulosic biomass: technology, economics, and opportunities. Biores Technol 50:315 (1994). 90. Wingren A, Galbe M and Zacchi G, Techno-economic evaluation of producing ethanol from softwood: Comparison of SSF and SHF and identication of bottlenecks. Biotechnol Progr 19:11091117 (2003). 91. Ewanick SM, Bura R and Saddler JN, Acid-catalyzed steam pretreatment of lodgepole pine and subsequent enzymatic hydrolysis and fermentation to ethanol. Biotechnol Bioeng 98: 737746 (2007) 92. Soderstrom J, Pilcher L, Galbe M and Zacchi G, Combined use of H2SO4 and SO2 impregnation for steam pretreatment of spruce in ethanol production. Appl Biochem Biotechnol 105108:127140 (2003). 93. Pan X, Arato C, Gilkes N, Gregg D, Mabee W, Pye K, Xiao Z, Zhang X and Saddler J. Biorening of softwoods using ethanol organosolv pulping: Preliminary evaluation of process streams for manufacture of fuel-grade ethanol and co-products. Biotechnol Bioeng 90:473481 (2005). 94. Oliva JM, Saez F, Ballesteros I, Gonzalez A, Negro MJ, Manzanares P and Ballesteros M, Effect of lignocellulosic degradation compounds from steam explosion pretreatment on ethanol fermentation by thermotolerant yeast Kluyveromyces marxianus. Appl Biochem Biotechnol 105108: 141153 (2003). 95. Nagle NJ, Elander RT, Newman MM, Rohrback BT, Ruiz RO and Torget RW, Efcacy of a hot washing process for pretreated yellow poplar to enhance bioethanol production. Biotechnol Progr 18:734738 (2002). 96. Cantarella M, Cantarella L, Gallifuoco A, Spera A and Alfani F, Effect of inhibitors released during steam-explosion treatment of poplar wood on subsequent enzymatic hydrolysis and SSF. Biotechnol Progr 20:200206 (2004).

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

71

Y Pu et al.

Review: New forestry biofuels sector

97. De Bari I, Viola E, Barisano D, Cardinale M, Nanna F, Zimbardi F, Cardinale G and Braccio G, Ethanol production at ask and pilot scale from concentrated slurries of steam-exploded aspen. Ind Eng Chem Res 41:17451753 (2002). 98. De Bari I, Nanna F and Braccio G, SO2-catalyzed steam fractionation of aspen chips for bioethanol production: Optimization of the catalyst impregnation. Ind Eng Chem Res 46:77117720 (2007). 99. Berlin A, Gilkes N, Kilburn D, Maximenko V, Bura R, Markov A, Skomarovsky A, Gusakov A, Sinitsyn A, Okunev O, Solovieva I and Saddler JN. Evaluation of cellulase preparations for hydrolysis of hardwood substrates. Appl Biochem Biotechnol 129132: 528545 (2006). 100. Tu M, Chandra RP and Saddler JN, Evaluating the distribution of cellulases and the recycling of free cellulases during the hydrolysis of lignocellulosic substrates. Biotechnol Progr 23:398406 (2007). 101. Yang B and Wyman CE. Lignin blocking proteins and their use in treating lignocellulosics. US Pat. Appl. Publ. US 2006088922 (2006). 102. Boerjesson J, Engqvist M, Sipos B and Tjerneld F, Effect of poly(ethylene glycol) on enzymatic hydrolysis and adsorption of cellulase enzymes to pretreated lignocellulose. Enzyme Microb Technol 41:186195 (2007). 103. Sonderegger M, Jeppsson M, Larsson C, Gorwa-Grauslund M-F, Boles E, Olsson L, Spencer-Martins I, Hahn-Haegerdal B and Sauer U, Fermentation performance of engineered and evolved xylosefermenting Saccharomyces cerevisiae strains. Biotechnol Bioeng 87:9098 (2004). 104. Taherzadeh MJ and Niklasson C, Ethanol from lignocellulosic materials: Pretreatment, acid and enzymatic hydrolyses, and fermentation. ACS Symposium Series 889, pp. 4968 (2004). 105. Dien BS, Cotta MA and Jeffries TW, Bacteria engineered for fuel ethanol production: current status. Appl Microbiol Biotechnol 63:258266 (2003). 106. Sommer P, Georgieva T and Ahring BK, Potential for using thermophilic anaerobic bacteria for bioethanol production from hemicellulose. Biochem Soc T 32: 283289 (2004). 107. Olsson L and Hahn-Haegerdal B, Fermentation of lignocellulosic hydrolyzates for ethanol production. Enzyme Microb Technol 18: 31231 (1996). 108. Helle SS, Murray A, Lam J, Cameron DR and Duff SJB, Xylose fermentation by genetically modied Saccharomyces cerevisiae 259ST in spent sulte liquor. Biores Technol 92:163171 (2004). 109. Lawford HG and Rousseau JD, Performance testing of Zymomonas mobilis metabolically engineered for cofermentation of glucose, xylose, and arabinose. Appl Biochem Biotechnol 98100:429448 (2002). 110. Hahn-Haegerdal B, Galbe M, Gorwa-Grauslund MF, Liden G and Zacchi G, Bio-ethanol the fuel of tomorrow from the residues of today. Trends Biotechnol 24:549556 (2006). 111. Sues A, Millati R, Edebo L and Taherzadeh MJ, Ethanol production from hexoses, pentoses, and dilute-acid hydrolyzate by Mucor indicus. FEMS Yeast Res 5:669676 (2005). 112. Karimi K, Brandberg T, Edebo L and Taherzadeh MJ, Fed-batch cultivation of Mucor indicus in dilute-acid lignocellulosic hydrolyzate for ethanol production. Biotechnol Lett 27:13951400 (2005). 113. See http://www.eere.energy.gov/ biomass/cellulase_cost.html

114. Lynd LR, van Zyl WH, McBride JE and Laser M, Consolidated bioprocessing of cellulosic biomass: an update. Curr Opin Biotechnol 16:577583 (2005). 115. Den HR, McBride JE, La Grange DC, Lynd R and Van Zyl WH, Functional expression of cellobiohydrolases in Saccharomyces cerevisiae towards one-step conversion of cellulose to ethanol. Enzyme Microb Technol 40:12911299 (2007). 116. Weimer PJ, Ethanol and co-products from cellulosic biomass. Int Sugar J 108:630633 (2006). 117. Mohan D, Pittman CU Jr and Steele PH, Pyrolysis of wood/ biomass for bio-oil: a critical review. Energy Fuels 20:848889 (2006). 118. Lesueur D, Castola V, Bighelli A, Conti L and Casanova J, Quantication of hydroxyacetaldehyde in a biomass pyrolysis liquid using 13C NMR spectroscopy. Spectrosc Lett 40:591602 (2007). 119. Debdoubi A, El Amarti A, Colacio E, Blesa MJ and Hajjaj LH, The effect of heating rate on yields and compositions of oil products from esparto pyrolysis. Int J Energ Res 30:12431250 (2006). 120. Czernik S and Bridgwater AV, Overview of applications of biomass fast pyrolysis oil. Energy Fuels 18:590598 (2004). 121. Oasmaa A, Peacocke C, Gust S, Meier D and McLellan R, Norms and standards for pyrolysis liquids end-user requirements and specications. Energy Fuels 19:21552163 (2005). 122. Chiaramonti D, Bonini M, Fratini E, Tondi G, Gartner K, Bridgwater A V, Grimm H P, Soldaini I, Webster A and Baglioni P, Development of emulsions from biomass pyrolysis liquid and diesel and their use in engines Part 2: tests in diesel engines Biomass Bioenerg 25:101111 (2003). 123. Horne PA and Williams PT, Reaction of oxygenated biomass pyrolysis model compounds over a ZSM-5 catalyst. Renew Energ 7:131144 (1996). 124. Adam J, Antonakou E, Lappas A, Stoecker M, Nilsen MH, Bouzga A, Hustad JE and Oye G, In situ catalytic upgrading of biomass derived fast pyrolysis vapours in a xed bed reactor using mesoporous materials. Micropor Mesopor Mat 96:93101 (2006). 125. Antonakou E, Lappas A, Nilsen MH, Bouzga A and Stoecker M, Evaluation of various types of Al-MCM-41 materials as catalysts in biomass pyrolysis for the production of bio-fuels and chemicals. Fuel 85:22022212 (2006). 126. Gayubo AG, Aguayo AT, Atutxa A, Prieto R and Bilbao J, Deactivation of a HZSM-5 zeolite catalyst in the transformation of the aqueous fraction of biomass pyrolysis oil into hydrocarbons. Energy Fuels 18:16401647 (2004). 127. Huber GW and Dumesic JA, An overview of aqueous-phase catalytic processes for production of hydrogen and alkanes in a biorenery. Catal Today 111:119132 (2006). 128. Huber GW, Chheda JN, Barrett CJ and Dumesic JA, Production of liquid alkanes by aqueous-phase processing of biomass-derived carbohydrates. Science 308:14461450 (2005). 129. Chiaramonti D, Bonini M, Fratini E, Tondi G, Gartner K, Bridgwater AV, Grimm HP, Soldaini I, Webster A and Baglioni P, Development of emulsions from biomass pyrolysis liquid and diesel and their use in engines Part 2: tests in diesel engines. Biomass Bioenerg 25:101111 (2003). 130. Dasappa S, Paul PJ, Mukunda, HS, Rajan NKS, Sridhar G and Sridhar HV, Biomass gasication technology a route to meet energy needs. Curr Sci 87:908916 (2004).

72

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

Review: New forestry biofuels sector

Y Pu et al.

131. Hrabovsky M, Konrad M, Kopecky V, Hlina M, Kavka T, Chumak O, van Oost G, Beeckman E and Defoort B, Pyrolysis of wood in arc plasma for syngas production. High Temp Mater Processes 10:557570 (2006). 132. Jeczmionek L, Biomass as raw material for production of hydrocarbon fuels biomass gasication to produce synthesis gas and its use. Nafta-Gaz 62:7282 (2006). 133. Nordgreen T, Liliedahl T and Sjoestroem K, Metallic iron as a tar breakdown catalyst related to atmospheric uidised bed gasication of biomass. Fuel 85:689694 (2006). 134. Ma L, Verelst H and Baron GV, Integrated high temperature gas cleaning: tar removal in biomass gasication with a catalytic lter. Catal Today 105:729734 (2005). 135. Devi L, Nair SA, Pemen AJM, Yan K, van Heesch EJM, Ptasinski KJ and Janssen FJJG, Tar removal from biomass gasication processes, in Biomass and Bioenergy, ed by Brenes MD, Nova Science Publishers Inc., Hauppauge, NY, pp. 249274 (2006). 136. Kelley SS, Lignocellulosic bioreneries: reality hype or something in between? in Materials Chemicals and Energy from Forest Biomass, ACS Symposium Series 954, ed by Argyropoulos DS, American Chemical Society, Washington, pp. 3147 (2007). 137. van Heiningen A, Converting a kraft pulp mill into an integrated forest biorenery Pulp Pap Canada 107:3843 (2006). 138. Farmer MC, Adaptable biorenery: some basic economic concepts to guide research selection. TAPPI Engineering Pulping and Environmental Conference Proceedings, Philadelphia, PA (2005). 139. Kim DH, Allison L, Carter B, Hou Q, Courchene C, Ragauskas AJ and Sealey J, Proling the wood and pulping properties of southern pine thinning resources. Tappi J 4:2125 (2005). 140. Amidon TE, Francis R, Scott GM, Bartholomew J, Ramarao BV and Wood CD, Pulp and pulping processes from an integrated forest biorenery. Appl. No. PCT/US2005/013216 (2007). 141. Thorp B, Transition of mills to biorenery model creates new prot streams. Pulp and Paper, Nov., 3539 (2005). 142. Amidon TE, Bolton TS, Francis RC and Gratien K, Effect of hot water pre-extraction on alkaline pulping of hardwoods. TAPPI Engineering Pulping & Environmental Conference Proceedings. Atlanta, GA (2006).

143. Yoon SH, MacEwan K and Heiningen AV, Pre-Extraction of southern pine chips with hot water followed by kraft cooking. TAPPI Engineering Pulping & Environmental Conference Proceedings. Atlanta, GA (2006). 144. Kim SB and Jin WC, Pretreatment for enzymatic hydrolysis of used newspaper, in Lignocellulose Biodegradation, ACS Symposium Series 889, ed by Saha BS, Hayashi K. American Chemical Society, Washington, pp. 3648 (2004). 145. Agblevor FA, Method for producing bioethanol from a lignocellulosic biomass and recycled paper sludge. US Pat. Appl. Publ. US 2007134781(2007). 146. Kadar Zs, Szengyel Zs and Reczey K, Simultaneous saccharication and fermentation (SSF) of industrial wastes for the production of ethanol. Ind Crop Prod 20:103110 (2004). 147. Fan Z and Lynd LR, Conversion of paper sludge to ethanol II: Process design and economic analysis. Bioproc Biosyst Eng 30:3545 (2007). 148. See http://www.energy.gov/news/4827.htm 149. Magdzinski L, Tembec Temiscaming integrated biorenery. Pulp Pap Can 107: 4446 (2006). 150. Perine L, Bio-reneries on the brink, Paper 360 o Dec., 31 (2006); Raymond D and Closset G, Forest Product Biorenery: technology for a new future. Solutions, Sept., 4953 (2004). 151. See http://www.idahoforests.org/re_20020033.htm 152. See http://www.tappi.org/s_tappi/doc.asp?CID=183&DID=551863 153. Solomon BD, Barnes JR and Halvorsen KE, Grain and cellulosic ethanol: history, economics and energy policy. Biomass Bioenerg 31:416425 (2007). 154. See http://www.tappi.org/s_tappi/doc.asp?CID=183&DID=554165 155. American Petroleum Institute Technical Report 939-D, Second Edition, Stress corrosion cracking of carbon steel in fuel grade ethanol: review, experience, survey, eld monitoring, and laboratory testing, May (2007). 156. Sridhar N, Price K, Buckingham J and Dante J, Stress corrosion cracking of carbon steel in ethanol. Corrosion 62:687702 (2006). 157. Maldonado JG and Sridhar N, SCC of carbon steel in fuel ethanol service: effect of corrosion potential and ethanol processing source. Corrosion/2007, Paper No. 07574, Houston, TX: NACE International (2007). 158. Morris D, The next economy: from dead carbon to living carbon. J Sci Food Agr 86: 17431746 (2006).

2007 Society of Chemical Industry and John Wiley & Sons, Ltd | Biofuels, Bioprod. Bioref. 2:5873 (2008); DOI: 10.1002/bbb

73

Anda mungkin juga menyukai