Вы находитесь на странице: 1из 8

Review

TRENDS in Biotechnology

Vol.24 No.12

Bio-ethanol the fuel of tomorrow from the residues of today


n and G. Zacchi B. Hahn-Ha gerdal, M. Galbe, M.F. Gorwa-Grauslund, G. Lide
Lund University, PO Box 124, S-221 00 Lund, Sweden, Getingeva gen 60

The increased concern for the security of the oil supply and the negative impact of fossil fuels on the environment, particularly greenhouse gas emissions, has put pressure on society to nd renewable fuel alternatives. The most common renewable fuel today is ethanol produced from sugar or grain (starch); however, this raw material base will not be sufcient. Consequently, future large-scale use of ethanol will most certainly have to be based on production from lignocellulosic materials. This review gives an overview of the new technologies required and the advances achieved in recent years to bring lignocellulosic ethanol towards industrial production. One of the major challenges is to optimize the integration of process engineering, fermentation technology, enzyme engineering and metabolic engineering. Introduction One of the greatest challenges for society in the 21st century is to meet the growing demand for energy for transportation, heating and industrial processes, and to provide raw material for the industry in a sustainable way. An increasing concern for the security of the oil supply has been evidenced by increasing oil prices, which during 2006 approached US$80 per barrel. More importantly, the future energy supply must be met with a simultaneous substantial reduction of green house gas emissions. Actions towards this aim have been initiated. The European Commission plans to substitute progressively 20% of conventional fossil fuels with alternative fuels in the transport sector by 2020, with an intermittent goal set at 5.75% in 2010. In the USA, the Energy Policy Act of 2005 requires blending of 7.5 billion gallons of alternative fuels by 2012 [1], and recently the US President, in his state of the union address, set the goal to replace more than 75% of imported oil with alternative fuels by the year 2025 [2]. Liquid biofuels from renewable resources, particularly from lignocellulose materials, will have a substantial role in meeting these goals (Box 1). Ethanol has already been introduced on a large scale in Brazil, the US and some European countries, and we expect it to be one of the dominating renewable biofuels in the transport sector within the coming 20 years. Ethanol can be blended with petrol or used as neat alcohol in dedicated engines, taking advantage of the higher octane number and higher heat of vaporization; furthermore, it is an excellent fuel for future advanced exi-fuel hybrid
Corresponding author: Zacchi, G. (Guido.Zacchi@chemeng.lth.se). Available online 16 October 2006.
www.sciencedirect.com

vehicles. Currently, ethanol for the fuel market is produced from sugar (Brazil) or starch (USA) at competitive prices. However, this raw material base, which also has to be used for animal feed and human needs, will not be sufcient to meet the increasing demand for fuel ethanol; and the reduction of greenhouse gases resulting from use of sugaror starch-based ethanol is not as high as desirable [3]. Both these factors call for the exploitation of lignocellulose feedstocks, such as agricultural and forest residues as well as dedicated crops, for the production of ethanol. This review summarizes recent developments in the bioconversion processes aimed at fuel ethanol production, with emphasis on process integration. In particular, the concept that each individual unit operation has to be developed and optimized in relation to the preceding and subsequent process steps will be discussed. Overview of the conversion process With so many advantages, why are there still no production facilities using lignocellulosic materials? Ethanol is currently produced from sugar cane and starch-containing materials, where the conversion of starch to ethanol includes a liquefaction step (to make the starch soluble) and a hydrolysis step (to produce glucose). The resulting glucose is then readily fermented. Although there are similarities between the lignocellulosic and the starch process, the techno-economic challenges facing the former are large. There are several options for a lignocellulose-to-ethanol process but, regardless of which is chosen, the following features must be assessed in comparison with established sugar- or starch-based ethanol production. (i) Efcient de-polymerization of cellulose and hemicellulose to soluble sugars. (ii) Efcient fermentation of a mixed-sugar hydrolysate containing six-carbon (hexoses) and ve-carbon (pentoses) sugars as well as fermentation inhibitory compounds. (iii) Advanced process integration to minimize process energy demand. (iv) Cost-efcient use of lignin. The rst step in the conversion of biomass to ethanol is size reduction and pretreatment. In this review, only the enzymatic process (Figure 1) will be discussed because it is considered to be the most promising technology [46]. The hemicellulose and cellulose polymers are hydrolyzed with enzymes or acids to release monomeric sugars. The sugars from the pretreatment and enzymatic hydrolysis steps are fermented by bacteria, yeast or lamentous fungi,

0167-7799/$ see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.tibtech.2006.10.004

550

Review

TRENDS in Biotechnology Vol.24 No.12

Box 1. Advantages of lignocellulose-based liquid biofuels


 Biofuel sources are geographically more evenly distributed than the fossil fuels; thus, the sources of energy will, to a larger extent, be domestic and provide security of supply.  Lignocellulosic raw materials minimize the potential conflict between land use for food (and feed) production and energy feedstock production. The raw material is less expensive than conventional agricultural feedstock and can be produced with lower input of fertilizers, pesticides, and energy.  Biofuels from lignocellulose generate low net greenhouse gas emissions, reducing environmental impacts, particularly climate change.  Biofuels might also provide employment in rural areas

although the enzymatic hydrolysis and fermentation can also be performed in a combined step a so-called simultaneous saccharication and fermentation (SSF). After nal purication (by distillation and molecular sieves or other separation techniques), the ethanol is ready to be used as a fuel, either neat or blended with petrol. A part of the lignin, the principal solid part of the biomass remaining, can be burnt to provide heat and electricity for the process, whereas the rest is retained as a valuable coproduct. The most probable use today would be as an ash-free solid fuel, but various technologies are under development to convert it to a higher-value product, which could form the basis for a new branch of industrial chemistry [7]. First, the biomass is converted to sugars. . . It is only 40 years ago that the biodegradation of lignocellulosics was rst discussed [8]. Enzyme conversion is substrate-specic without by-product formation, which reduces inhibition of the following process steps. However, enzyme-catalysed conversion of cellulose to glucose is slow unless the biomass has been subjected to pretreatment, which is also required to reach high yields and to make the process commercially successful [9]: the pretreatment aims

to increase pore size and reduce cellulose crystallinity. In acid-catalyzed pre-treatment, the hemicellulose layer is hydrolyzed, whereas in alkali-catalyzed pretreatment, mainly, a part of the lignin is removed and hemicellulose has to be hydrolysed by the use of hemicellulases. Hence, pretreatment is necessary to expose the cellulose bres to the enzymes or to at least make the cellulose more accessible to the enzymes. An efcient pretreatment can substantially reduce the enzyme requirements, which make up a large part of the production cost. Pretreatment is usually assessed in a number of ways: by enzymatic hydrolysis (EH) of the solid material to determine the digestibility; by fermentation of the liquid to assess the effect of potential inhibitors towards the fermenting microorganism; and/or by simultaneous saccharication and fermentation (SSF) of the pretreated material. The conditions for the assessment can vary, particularly in terms of washed or non-washed solids and the concentration of solids and enzymes in the EH and SSF, making comparison of results from different investigations difcult. In an extensive study undertaken in the USA, where the same batch of corn stover was pretreated using various methods (e.g. dilute acid, AFEX, hot water treatment) and then subjected to standard evaluation techniques, the yields of sugars were found to be more or less the same [10]. Total sugar yields after pretreatment followed by enzymatic hydrolysis of around 90% or more were reached, demonstrating that corn stover is an easily degradable material. When corn stover was steam pretreated with small amounts of SO2, overall sugar yields close to the theoretical value were obtained; steam pretreatment without a catalyst also resulted in 90% glucose yield [11]. In countries such as Sweden, Canada and the USA, much of the available biomass is softwood, which is more difcult to hydrolyse than corn stover. For softwood, steam pretreatment with the addition of an acid catalyst such as

Figure 1. Schematic flowsheet for the conversion of biomass to ethanol.


www.sciencedirect.com

Review

TRENDS in Biotechnology

Vol.24 No.12

551

H2SO4 or SO2 is a prerequisite to reach high sugar yields. Acid increases the recovery of hemicellulose sugars and it improves the enzymatic hydrolysis of the solid fraction; the acid catalyst in steam pretreatment functions similar to acid pulp cooking but with less liquid. Steam pretreatment of SO2 impregnated spruce chips yields a material that is relatively easy to hydrolyze and ferment, whereas dilute acid impregnation results in a material harder to ferment because of the generation of inhibitory compounds [12]. With present bench-scale technology, it is possible to obtain around 300 litres of ethanol per metric ton spruce, which is 70% of the theoretical overall yield based on hexose sugars [12,13]. Steam pretreatment with the addition of a catalyst for hydrolysis and improved enzymatic digestibility is the closest to commercialisation. It has been widely tested in pilot-scale equipment, for example, in the Iogen pilot plant (Canada) [14], the Souston pilot plant (France) [15] rnsko ldsvik (Sweden), and is and in the pilot plant in O used in a demonstration-scale ethanol plant at Iogen (Canada). It is also to be used in Salamanca (Spain), in a plant constructed by the company Abengoa. Most probably, there will not be one general method because different types of raw material require different pretreatments. So far, methods such as ammonia ber explosion (AFEX), wet oxidation and liquid hot water (LHW) treatment seem to be more successful for agricultural residues [1618], whereas steam pretreatment has resulted in high sugar yields for both forestry and agricultural residues. Glucose yields >90% and xylose yields >80% were obtained after enzymatic hydrolysis, both with and without the addition of an acid catalyst [12,1921]. Acid-catalyzed pretreatment primarily solubilizes the hemicellulose fraction into the liquid phase. For softwood, the liquid mainly contains solubilized mannose in addition to small amounts of xylose, arabinose, galactose and glucose. The solid phase comprises lignin and cellulose, the latter of which is subjected to enzymatic hydrolysis. The maximum cellulase activity of most fungal-derived cellulases and bglucosidases is observed at 50 8C and at a pH of 4.05.0; however, the optimal conditions vary with the hydrolysis time and are dependent on the source of the enzymes [22]. Cellulases belong to two groups of enzymes known as endoglucanases (EG) and cellobiohydrolases (CBH), respectively. EG randomly attack the cellulose chain, creating free ends for CBH to cleave dimers of glucose (cellobiose) off [23]. A third type of enzyme, b-glucosidase, which hydrolyzes cellobiose into two glucose molecules, is also necessary: in the absence of b-glucosidase, end-product inhibition from cellobiose will occur. Furthermore, compounds generated during pretreatment might have an adverse effect on enzymatic hydrolysis. The enzymatic hydrolysis of spruce was greatly improved when the liquid fraction from the pretreatment step was replaced with a buffer solution [24]. This could not be entirely ascribed to the reduction in end-product inhibition, suggesting that inhibitory compounds had also been removed. In an initiative of the US Department of Energy, two companies, Genencor International (http://www.genencor. com/) and Novozymes Inc. (http://www.novozymes.com/), were awarded $17 million each, with the goal to reduce
www.sciencedirect.com

the enzyme cost 10-fold (http://www.eere.energy.gov/). However, the production cost of enzymes is still too high and requires further reduction. One way to achieve this is to use a fraction of the feedstock and/or the hydrolysate for in situ enzyme production by fungi or other microorganisms [25]. . . . and then the sugars are fermented to ethanol Contrary to sucrose- and starch-based ethanol production, lignocellulose-based production is a mixed-sugar fermentation in the presence of inhibiting compounds low molecular weight organic acids, furan derivatives, phenolics and inorganic compounds released and formed during pretreatment and/or hydrolysis of the raw material [26]. Lignocellulosic raw materials, in particular hardwood and agricultural raw materials, can contain 520% (or more) of the pentose sugars xylose and arabinose, which are not fermented to ethanol by the most commonly used industrial fermentation microorganism, the yeast Saccharomyces cerevisiae. Xylose is by far the most abundant pentose sugar, whereas arabinose can constitute as much as 1415% in corncob hulls and wheat bran, respectively. Consequently, most research efforts have been devoted to the development of efcient xylose-fermenting microorganisms [27,28]. Xylose-fermenting microorganisms are found among bacteria, yeast and lamentous fungi [29]. Anaerobic bacteria ferment pentoses, but are inhibited already at low sugar and ethanol concentrations. In addition, the ethanolic fermentation occurs with considerable by-product formation, which reduces the ethanol yield [30]. Natural xylose-fermenting yeast, notably Pichia stipitis CBS 6054, ferment xylose to ethanol with reasonable yield and productivity; however, these yeast strains are inhibited by compounds generated during pretreatment and hydrolysis of the lignocellulose material [31]. Filamentous fungi tolerate inhibitors but are too slow for a competitive industrial process. Therefore, efforts have predominantly been made to obtain recombinant strains of bacteria and yeast able to meet the requirements of industrial lignocellulose fermentation (Figure 2).

Figure 2. Strains that are metabolically engineered for ethanol production from pentoses. In essence, either the tail end, as in Escherichia coli and Klebsiella oxytoca, or the front end of metabolism, as for Saccharomyces cerevisiae and Zymomonas mobilis, have been engineered. Abbreviation: rec, recombinant.

552

Review

TRENDS in Biotechnology Vol.24 No.12

Pentose-fermenting Escherichia coli [32] and Klebsiella oxytoca [33] have been generated by introducing ethanologenic genes from Zymomonas mobilis (Figure 2). At the same time, the rst xylose-fermenting S. cerevisiae strain was generated through the introduction of genes for xylosemetabolizing enzymes from P. stipitis [34] (Figure 2). Later xylose-fermenting strains of S. cerevisiae were constructed by introducing the genes encoding xylose isomerase from the bacterium Thermus thermophilus [35] and the anaerobic fungus Piromyces sp. [36], respectively. For xyloseusing S. cerevisiae, high ethanol yields from xylose also require metabolic engineering strategies to enhance the xylose ux. This was independently demonstrated in an SSF set-up for corn stover, where glucose and xylose were fermented simultaneously [37], and in a recombinant strain with increased carbon ux [38]. Z. mobilis also efciently produces ethanol from the hexose sugars glucose and fructose but not from pentose sugars, although a xylose fermenting Z. mobilis was generated by introducing a xylose-metabolizing pathway from E. coli [39]. More recently, the obligatory anaerobic bacterium Thermoanaerobacterium saccharolyticum has been genetically engineered for improved ethanolic fermentation (Joe Shaw et al., oral presentation, Nashville, 2006). E. coli and K. oxytoca naturally metabolize arabinose, such that the ethanologenic strains ferment all lignocellulose-derived sugars [40]; furthermore, xylose- and arabinose-fermenting strains of Z. mobilis have been constructed [41]. Because yeast only ferment arabinose to ethanol in rich media [42,43], S. cerevisiae has been engineered for arabinose use by introducing both bacterial [44] and fungal genes [45] encoding arabinose-metabolizing enzymes, where the fungal approach did not result in appreciable arabinose fermentation. The functional arabinose-metabolizing pathway has recently been

integrated into the diploid xylose-fermenting S. cerevisiae strain TMB 3400, and co-usage of xylose and arabinose has been demonstrated [46]. In addition to being able to ferment both hexose and pentose sugars, the fermenting microorganisms must do this in the presence of inhibiting compounds such as weak acids, furan derivatives and phenolics [47]. Detoxication by chemical or physical methods before fermentation reduces the concentration of inhibitors and improves the performance in the fermentation step. Treatment with Ca(OH)2 (so-called overliming) and ion-exchange treatment increased the fermentability signicantly [48]; however, this was at the expense of increased process cost and sugar loss [49]. Furthermore, Ca(OH)2 is unlikely to be accepted in a full-scale ethanol plant, where precipitation of calcium salts might foul distillation columns, evaporators and heat-exchanger surfaces [50,51]. In ethanolic yeast fermentation, in-situ biological detoxication occurs when carbonyl compounds furans and phenolics are reduced to the corresponding alcohols [52,53], which are less inhibitory to yeast [54]. This phenomenon has been successfully exploited in online probing process control in fed-batch fermentation, by adjusting the feed rate of hydrolysate to the intrinsic capacity for inhibitor conversion of the yeast [55]. In such cases, the fermentation productivity will be a function of the inhibitor concentration, the conversion capacity of the yeast and the quality of the process control. Only recently it was demonstrated that the overexpression of a gene encoding alcohol dehydrogenase activity improved the fermentation of a 5-hydroxymethyl furfural-containing medium [56]. The competitiveness of fuel ethanol research makes strain comparisons difcult. Literature data are hampered by incomplete reporting, differences in fermentation conditions (including media composition and oxygenation),

Table 1. Latest fermentation results on hydrolysate from various recombinant and native xylose-fermenting organisms.
Organism Hydrolysate Detoxication Fermentation mode Yield Fermentation time Sp. Refs g/g initial sugar (hours) productivity g/g cells.h + n/a 0.49 n/a 0.37 [66] + Batch Batch Fed-batch n/a Batch Batch Batch n/a Continuous Batch 12 Batch 12 Batch 12 Fed batch Fed batch SSF, Batch 15 SSF, Fed batch 15 0.390.41 3 0.300.38 4 0.350.39 5 0.46 0.43 0.40 0.42 0.41 0.35 0.36 0.41 0.45 0.37 0.43 0.33 16 0.30 16 93102 2968 118 n/a 64 39 n/a n/a 48 24 55 n/a n/a 96 96 n/a [67]

Escherichia coli KO11

Bagasse hemicellulose 1 Corn ber 2

E. coli FBR5

Zymomonas mobilis 8b 6 Pichia stipitis 8 P. stipitis CBS 5773 Saccharomyces cerevisiae 424A (LNFST) S. cerevisiae TMB3006 S. cerevisiae TMB3400

Corn stover Rice hull Rice hull Corn stover 7 Wheat straw Spent sulte liquor Corn ber Corn stover Corn stover Spruce Spruce Corn stover 14

+ +3 +3 +3 + 10 +3 +3 11

0.214 n/a n/a n/a 0.44 0.05 0.072 13 0.044 13 0.048 13 0.66 0.25 0.036 0.076

[66] [68] [68] [66] [69] [66,70] 9 [66] [71] [71] [9] [66] [66] [37]

Supplemented with 2.5% (w/v) corn steep liquor; 2 supplemented with 5% (w/v) corn steep liquor; 3 overliming; 4 neutralization by anion exchange; 5 membrane pervaporation; 6 acid-tolerant derivative of ZM4/AcR (pZB5); 7 80% hydrolysate supplemented at 100 g/l glucose; 8 adapted strain of NRRL Y-7124; 9 calculated from Nigam 2001 [70]; 10 rotoevaporated spent sulte liquor; 11 optimized pH-controlled liquid hot water pretreatment; 12 Semi-anaerobic fermentations with 10 g/l yeast extract added; 13 Calculated with the given initial cell dry weight of 8.5 g/l; 14 0.05% water insoluble solid; 15 1 g/l yeast extract added; 16 % of theoretical based on the glucose and xylose content in the raw material, (i.e. taking into account both hydrolysis and fermentation yields). Abbreviations. n/a, not available; SHF, separate hydrolysis and fermentation; SSF, simultaneous saccharication and fermentation.
www.sciencedirect.com

Review

TRENDS in Biotechnology

Vol.24 No.12

553

vastly different raw materials, and pretreatment and hydrolysis conditions [29]. Table 1 summarizes the most recently published results on the fermentation of lignocellulose hydrolysate with natural and recombinant bacteria and yeast. In essence, so far only recombinant S. cerevisiae strains have been able to ferment xylose in non-detoxied hydrolysates, where the fed-batch technology also permits the fermentation of extremely inhibitory softwood hydrolysates. Moving on from research to a commercial process The estimated cost of producing ethanol from cellulosic materials varies widely between investigations, as shown both in some earlier review papers [57,58], with production costs in the range of 0.28 to 1.0 US$/l ethanol, and in more recent techno-economic evaluations [13,59,60]. However, most cost estimations are based on laboratory-scale and, to some extent, pilot-scale data for individual process steps and should be treated with caution. The cost of raw material, which varies considerably between different studies (US$22US$61 per metric ton dry matter), and the capital costs, which makes the total cost dependent on plant capacity, contribute most to the total production cost. The cost for hydrolysis, particularly for the enzymatic process, is also a major cost contributor. Raw material cost is reduced by using the whole crop for products and co-products. High ethanol yield requires complete hydrolysis of both cellulose and hemicellulose with a minimum of sugar degradation, followed by efcient fermentation of all sugars in the biomass. In the shortterm, co-products are likely to be used for the production of fuel, heat and electricity; however, in the long term, bioethanol technology will form the basis for the sustainable production of commodity chemicals and materials in future bioreneries. High co-product yield requires reduced energy demand for ethanol production. This is achieved when high solids concentrations (Figure 3) are combined with integration of energy-intensive process steps (e.g. pretreatment, distillation, evaporation and

drying). In SSF, 12 % WIS (water-insoluble solids) result in an ethanol concentration >4 wt-% (weight-%; kg ethanol per 100 kg solution), which is necessary to reduce the energy demand in the distillation steps (A. Wingren, PhD thesis, Lund University, 2005). Further reductions in the energy demand can be obtained by recycling certain process streams, to minimise the amount of fresh water used [61]. However, high solids concentrations and recycling of process streams increase the concentration of compounds that are inhibitory to enzymatic hydrolysis and fermentation, necessitating detoxication or fed-batch technology, as described previously. It also results in high viscosity, which limits mixing and pumping. Process integration reduces the capital costs. In the separate hydrolysis and fermentation (SHF) process, cellulose is rst hydrolyzed to glucose and then glucose is fermented to ethanol. The primary advantage of SHF is that hydrolysis and fermentation occur at optimum conditions; the disadvantage is that cellulolytic enzymes are end-product inhibited so that the rate of hydrolysis is progressively reduced when glucose and cellobiose accumulate [24]. Product inhibition was the rationale for the rst report on simultaneous saccharication and fermentation (SSF) of cellulose [62]: in SSF, hydrolysis and fermentation occur simultaneously in the same vessel, and the end-product inhibition of the enzymes is relieved because the fermenting organism immediately consumes the released sugars. Furthermore, the fermentation seems to decrease the inhibition of the enzymes by converting some of the toxic compounds present in the hydrolysate [24]. This increases the overall ethanol productivity, the ethanol concentration and the nal ethanol yield [12,63] (Figure 4). More recently, the SSF technology has proved advantageous for the simultaneous fermentation of hexose and pentose sugars (so called SSCF). In SSCF, the enzymatic hydrolysis continuously releases hexose sugars, which increases the rate of glycolysis such that the pentose sugars are fermented faster and with higher yield [37].

Figure 3. Ethanol production cost as a function of water-insoluble solids (WIS) in the SSF step (expressed as weight-% of total material) for production of ethanol from spruce, based on a production capacity of 200 000 DM raw material per year. Broken line: use of meachanical vapour recompression in evaporation (1 SEK = $0.13). This figure clearly shows the importance of working at high WIS.
www.sciencedirect.com

Figure 4. Overall yields of fermentable sugars and ethanol from steam-pretreated spruce. Abbreviations: EH, sugar yield after separate enzymatic hydrolysis; SSF, ethanol yield after SSF; SO2, after impregnation with sulfur dioxide; H2SO4, after impregnation with sulfuric acid. Adapted from data presented in [12].

554

Review

TRENDS in Biotechnology Vol.24 No.12

Figure 5. Biorefinery integration of a combined heat and power plant with an ethanol production plant.

Further process integration can be achieved by performing both hydrolysis and fermentation in a single reactor, using one or a mixture of microorganisms that produce all the required enzymes and ferment all sugars so-called consolidated bioprocessing (CBP) [64]. However, no such microorganisms are currently available, and the concept is subject to further research. The economic analysis [60] of the cellulosic bioethanol process shows that reliable cost estimations require laboratory results are veried in pilot and demonstration plants, where all steps are integrated into a continuous process. This also provides the possibility to explore the benets of process integration to reduce the number of process steps and the energy demand, and to recirculate process streams to eliminate the use of fresh water and to reduce the amount of waste streams. Currently, the Iogen Corp. (http://www.iogen.ca/) demo-plant is the only operating plant for the production of bioethanol from lignocellulose using the enzymatic hydrolysis process. The plant can handle up to 40 tonnes per day of wheat, oat and barley straw and is designed to produce up to 3 million litres of cellulose ethanol per year. Abengoa Bioenergy (http:// www.abengoabioenergy.com) is also constructing a pilot plant in York, USA, to convert residual starch, cellulose and hemicellulose mainly corn stover to bioethanol and high-protein feed. In Salamanca, the same company constructed a demonstration plant integrated with a fuelethanol-from-grain plant, producing 195 million liters. In the demonstration plant, an additional 5 million liters of ethanol per year will be produced from cellulose, mainly
Box 2. Major research challenges
 Improving the enzymatic hydrolysis with efficient enzymes, reduced enzyme production cost and novel technology for high solids handling.  Developing robust fermenting organisms, which are more tolerant to inhibitors and ferment all sugars in the raw material in concentrated hydrolysates at high productivity and with high ethanol concentration.  Extending process integration to reduce the number of process steps and the energy demand and to re-use process streams to eliminate the use of fresh water and to reduce the amount of waste streams.
www.sciencedirect.com

from wheat staw. In Sweden, a fully integrated pilot plant for ethanol production from softwood, comprising both twostage dilute acid hydrolysis and the enzymatic process, was taken into operation in mid 2004. The pilot has a maximum capacity of 2 ton (DM) wood per day (http://www.etek.se). The step from pilot- and demo-scale production of lignocellulosic ethanol to competitive full-scale production requires further reduction of the production cost. One approach to this is the integration of ethanol production with a combined heat and power plant (Figure 5) or with a pulp and paper mill. This has been estimated to reduce the ethanol production cost by up to 20 percent for conditions prevailing in Sweden [65] and it is the main strategy pursued in the Swedish cellulosic ethanol effort. Similar conclusions were reached in a study on co-production of ethanol and electricity from softwood, based on conditions in California. Another option is to integrate cellulosic ethanol production with starch-based ethanol production to use the whole agricultural crop. For the immediate future, we believe that these integrated plant concepts will be used in the rst successful industrial scale production of lignocellulosic fuel ethanol. The transition of lignocellulosic fuel ethanol production into a mature industrial technology requires research and development efforts in the areas summarized in Box 2.

References
1 Gray, K.A. et al. (2006) Bioethanol. Curr. Opin. Chem. Biol 10, 16 2 Herrera, S. (2006) Bonkers about biofuels. Nat. Biotechnol. 24, 755 760 3 Farrell et al. (2006) Ethanol can contribute to energy and environmental goals. Science 311, 506508 4 Ogier, J.C. et al. (1999) Ethanol production from lignocellulosic biomass. Oil Gas Sci Technol 54, 6794 5 Yu, Z. and Zhang, H.X. (2004) Ethanol fermentation of acid-hydrolyzed cellulosic pyrolysate with Saccharomyces cerevisiae. Bioresour. Technol. 93, 199204 6 Sheehan, J. (2001) The road to bioethanol. A strategic perspective of the US Department of Energys National Ethanol Program. In Glycosyl Hydrolases for Biomass Conversion (Himmel, M.E., Baker, J.O. and Saddler, J.N., eds), pp. 225, American Chemical Society 7 Fernando, S. et al. (2006) Bioreneries: current status, challenges, and future direction. Energy Fuels 20, 17271737 8 Keller, F.A. (1996) Integrated bioprocess development. In Handbook on Bioethanol: Production and Utilization (Wyman, C.E., ed.), pp. 351 380, (Chapter 16), Taylor & Francis 9 Mosier, N. et al. (2005) Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour. Technol. 96, 19861993 10 Wyman, C.E. et al. (2005) Coordinated development of leading biomass pretreatment technologies. Bioresour. Technol. 96, 19591966 hgren, K. et al. Effect of hemicellulose and lignin removal on 11 O enzymatic hydrolysis of steam-pretreated corn stover. Bioresour. Technol. (in press) http://www.sciencedirect.com/science/journal/ 09608524 derstro m, J. et al. (2005) Separate versus simultaneous 12 So saccharication and fermentation of two-step steam pretreated softwood for ethanol production. J. Wood Chem. Technol. 25, 187 202 13 Wingren, A. et al. (2005) Effect of reduction in yeast and enzyme concentrations in a simultaneous-saccharication-and-fermentationbased bioethanol process technical and economic evaluation. Appl. Biochem. Biotechnol. 121, 485499 14 Schell, D.J. and Duff, B. (1996) Review of pilot plant programs for bioethanol conversion. In Handbook on Bioethanol: Production and Utilization (Wyman, C.E., ed.), pp. 381394, (Chapter 17), Taylor & Francis

Review

TRENDS in Biotechnology

Vol.24 No.12

555

15 Ropars, M. et al. (1992) Large-scale enzymic hydrolysis of agricultural lignocellulosic biomass Part 1: pretreatment procedures. Bioresour. Technol. 42, 197204 16 Gollapalli, L.E. et al. (2002) Predicting digestibility of ammonia ber explosion (AFEX)-treated rice straw. Appl. Biochem. Biotechnol. 98 100, 2335 17 Varga, E. et al. (2002) Chemical pretreatments of corn stover for enhancing enzymatic digestibility. Appl. Biochem. Biotechnol. 98 100, 7387 18 van Walsum, G.P. et al. (1996) Conversion of lignocellulosics pretreated with liquid hot water to ethanol. Appl. Biochem. Biotechnol. 5758, 157170 19 Nguyen, Q.A. et al. (2000) Two-stage dilute acid pretreatment of softwoods. Appl. Biochem. Biotechnol. 8486, 561576 20 Sassner, P. et al. (2006) Bioethanol production based on simultaneous saccharication and fermentation of steam-pretreated Salix at high dry-matter content. Enzyme Microb. Technol. 39, 756762 21 Ohgren, K. et al. (2005) Optimization of steam pretreatment of S02impregnated corn stover for fuel production. Appl. Biochem. Biotechnol. 121, 10551067 22 Tengborg, C. et al. (2001) Inuence of enzyme loading and physical parameters on the enzymatic hydrolysis of steam-pretreated softwood. Biotechnol. Prog. 17, 110117 23 Enari, T.M. et al. (1987) Enzymic hydrolysis of cellulose: is the current theory of the mechanisms of hydrolysis valid? Crit. Rev. Biotechnol. 5, 6787 24 Tengborg, C. et al. (2001) Reduced inhibition of enzymatic hydrolysis of steam pretreated softwood. Enzyme Microb. Technol. 28, 835844 25 Szengyel, Z. et al. (2000) Cellulase production of Trichoderma reesei Rut C 30 using steam-pretreated spruce hydrolytic potential of cellulases on different substrates. Appl. Biochem. Biotechnol. 846, 679691 26 Larsson, S. et al. (2000) Inuence of lignocellulose-derived aromatic compounds on oxygen-limited growth and ethanolic fermentation by Saccharomyces cerevisiae. Appl. Biochem. Biotechnol. 8486, 617632 27 Dien, B.S. et al. (2003) Bacteria engineered for fuel ethanol production: current status. Appl. Microbiol. Biotechnol. 63, 258266 28 Jeffries, T.W. (2006) Engineering yeasts for xylose metabolism. Curr. Opin. Biotechnol. 17, 17 gerdal, B. (1988) Xylose fermentation. Enzyme 29 Skoog, K. and Hahn-Ha Microb. Technol. 10, 6680 30 Desai, S.G. et al. (2004) Cloning of the L-lactate dehydrogenase gene and elimination of lactic acid production via gene knockout in Thermoanaerobacterium saccharolyticum JW/SL-YS485. Appl. Microbiol. Biotechnol. 65, 600605 gerdal, B. et al. (1994) An interlaboratory comparison of the 31 Hahn-Ha performance of ethanol-producing microorganisms in a xylose-rich acid hydrolysate. Appl. Microbiol. Biotechnol. 41, 6272 32 Ingram, L.O. et al. (1987) Genetic engineering of ethanol production in Escherichia coli. Appl. Environ. Microbiol. 53, 24202425 33 Burchhardt, G. and Ingram, L.O. (1992) Conversion of xylan to ethanol by ethanologenic strains of Escherichia coli and Klebsiella oxytoca. Appl. Environ. Microbiol. 58, 11281133 tter, P. and Ciriacy, M. (1993) Xylose fermentation by 34 Ko Saccharomyces cerevisiae. Appl. Microbiol. Biotechnol. 38, 776783 35 Walfridsson, M. et al. (1996) Ethanolic fermentation of xylose with Saccharomyces cerevisiae harbouring the Thermus thermophilus xylA gene which expresses an active xylose (glucose) isomerase. Appl Environ Microb 62, 46484651 36 Kuyper, M. et al. (2003) High-level functional expression of a fungal xylose isomerase: the key to efcient ethanolic fermentation of xylose by Saccharomyces cerevisiae? FEMS Yeast Res. 4, 6978 hgren, K. et al. (2006) Simultaneous saccharication and co37 O fermentation of glucose and xylose in steam-pretreated corn stover at high bre content with Saccharomyces cerevisiae TMB3400. J Biotechnol, In press, Available online May, 12. 38 Karhumaa, K. et al. (2006) High activity of xylose reductase and xylitol dehydrogenase improves xylose fermentation by recombinant Saccharomyces cerevisiae. Appl. Microbiol. Biotechnol. DOI: 10.1007/s00253-006-0575-3 In: http://www.springerlink.com/) 39 Zhang, M. et al. (1995) Metabolic engineering of a pentose metabolism pathway in ethanologenic Zymomonas mobilis. Science 267, 240 243
www.sciencedirect.com

40 Lindsay, S.E. et al. (1995) Improved strains of recombinant Escherichia coli for ethanol production from sugar mixtures. Appl. Microbiol. Biotechnol. 43, 7075 41 Mohagheghi, A. et al. (2002) Cofermentation of glucose, xylose, and arabinose by genomic DNA-integrated xylose/arabinose fermenting strain of Zymomonas mobilis AX101. Appl. Biochem. Biotechnol. 98 100, 885898 42 McMillan, J.D. and Boyton, B.L. (1996) Arabinose utilization by xylosefermenting yeasts and fungi. Appl. Biochem. Biotechnol. 4546, 569 584 43 Dien, B.S. et al. (1996) Screening for L-arabinose fermenting yeasts. Appl. Biochem. Biotechnol. 5758, 233242 44 Becker, J. and Boles, E. (2003) A modied Saccharomyces cerevisiae strain that consumes L-arabinose and produces ethanol. Appl. Environ. Microbiol. 69, 41444150 45 Richard, P. et al. (2003) Production of ethanol from L-arabinose by Saccharomyces cerevisiae containing a fungal L-arabinose pathway. FEMS Yeast Res 3, 185189 46 Karhumaa, K. et al. (2006) Co-utilisation of L-arabinose and D-xylose by laboratory and industrial Saccharomyces cerevisiae strains. Microb Cell Factor 5, 8 47 Palmqvist, E. and Hahn-Hagerdal, B. (2000) Fermentation of lignocellulosic hydrolysates. II: inhibitors and mechanisms of inhibition. Bioresour. Technol. 74, 2533 48 Larsson, S. et al. (1999) Comparison of different methods for the detoxication of lignocelluloase hydrolyzates of spruce. Appl. Biochem. Biotechnol. 7779, 91103 49 Nilvebrant, N-O. et al. (2003) Limits for alkaline detoxication of dilute-acid lignocellulose hydrolysates. Appl. Biochem. Biotechnol. 105108, 615628 50 Schell, D.J. et al. (2004) A bioethanol process development unit: initial operating experiences and results with a corn bre feedstock. Bioresour. Technol. 91, 179188 51 Kadam, K.L. et al. (2000) Softwood forest thinnings as a biomass source for ethanol production: A feasibility study for California. Biotechnol. Prog. 16, 947957 nsson, L.J. (2003) Comparison of the resistance of 52 Martin, C. and Jo industrial and laboratory strains of Saccharomyces and Zygosaccharomyces to lignocellulose-derived fermentation inhibitors. Enz. Microbial. Technol. 32, 386395 53 Sarvari Horvath, I. et al. (2003) Effects of furfural on the respiratory metabolism of Saccharomyces cerevisiae in glucose-limited chemostats. Appl. Environ. Microbiol. 69, 40764086 54 Taherzadeh, M.J. et al. (2000) Inhibition effects of furfural on aerobic batch cultivation of Saccharomyces cerevisiae growing on ethanol and/ or acetic acid. J. Biosci. Bioeng. 90, 374380 55 Taherzadeh, M.J. et al. (2000) On-line control of fed-batch fermentation of dilute-acid hydrolyzates. Biotechnol. Bioeng. 69, 330338 56 Petersson, A. et al. (2006) A 5-hydroxymethyl furfural reducing enzyme encoded by the Saccharomyces cerevisiae ADH6 gene conveys HMF tolerance. Yeast 23, 455464 57 von Sivers, M. and Zacchi, G. (1996) Ethanol from lignocellulosics: A review of the economy. Bioresour. Technol. 56, 131140 58 Lynd, L. et al. (1996) Likely features and costs of mature biomass ethanol technology. Appl Biochem. Biotechnol. 5758, 741 760 59 Wooley, R.J. et al. (1999) Lignocellulosic biomass to ethanol processing design and economics utilizing co-current dilute acid prehydrolysis and enzymatic hydrolysis current and futuristic scenarios. National Renewable Energy Laboratory Technical Report NREL/TP-51026157 (http://www.nrel.gov/docs/fy99osti/26157.pdf) 60 Aden, A. et al. (2002) Lignocellulosic biomass to ethanol processing design and economics utilizing co-current dilute acid prehydrolysis and enzymatic hydrolysis for corn stover. National Renewable Energy Laboratory Technical Report NREL/TP-510-32438 (http://www.nrel. gov/docs/fy02osti/32438.pdf) 61 Alkasrawi, M. et al. (2002) Recirculation of process streams in fuel ethanol production from softwood based on simultaneous saccharication and fermentation. Appl. Biochem. Biotechnol. 98 100, 849861 62 Takagi, M. et al. (1977) A method for production of alcohol direct from cellulose using cellulase and yeast. In Proceedings of the Bioconversion Symposium IIT (Delhi) pp. 551571

556

Review

TRENDS in Biotechnology Vol.24 No.12

63 Wright, J.D. et al. (1988) Simultaneous saccharication and fermentation of lignocellulose: process evaluation. Appl. Biochem. Biotechnol. 18, 7590 64 Lynd, L. et al. (2005) Consolidated bioprocessing of cellulosic biomass: an update. Curr. Opin. Biotechnol. 16, 577583 ngpannefo reningen (1994) IPK System study techno/economic 65 A reviews of process combinations of ethanol processes and other relevant industrial processes. Report P23332-1, NUTEK, Stockholm, Sweden gerdal, B. and Pamment, N. (2004) Microbial pentose 66 Hahn-Ha metabolism. Appl. Biochem. Biotechnol. 113116, 12071209 67 OBrien, D.J. et al. (2004) Ethanol recovery from corn ber hydrolysate fermentations by pervaporation. Bioresour. Technol. 92, 1519

68 Saha, B.B. et al. (2005) Dilute acid pretreatment, enzymatic saccharication, and fermentation of rice hulls to ethanol. Biotechnol. Prog. 21, 816822 69 Mohaghegi, A. et al. (2004) Performance of a newly developed integrant of Zymomonas mobilis for ethanol production on corn stover hydrolysate. Biotechnol. Lett. 26, 321325 70 Nigam, J.N. (2001) Ethanol production from wheat straw hemicellulose hydrolysate by Pichia stipitis. J. Biotechnol. 87, 1727 71 Sedlak, M. and Ho, N.W. (2004) Production of ethanol from cellulosic biomass hydrolysates using genetically engineered Saccharomyces yeast capable of cofermenting glucose and xylose. Appl. Biochem. Biotechnol. 113116, 403416

Progress in Biophysics and Molecular Biology


January 2007: Vol. 93 Themed issue on

Effects of ultrasound and infrasound relevant to human health


Guest edited by Alastair McKinlay
26 articles, including: * Medical diagnostic applications and sources Tony A. Whittingham * Therapeutic applications of ultrasound Gail ter Haar * Medical ultrasound imaging Jrgen Arendt Jensen * Medical and non-medical protection standards for ultrasound and infrasound Francis A. Duck * Rapporteur report: Mechanisms and interactions Timothy G. Leighton * Quantification of risk from fetal exposure to diagnostic ultrasound Charles C. Church * Ultrasound, microbubbles and the blood-brain barrier Stephen Meairs * Shear stress in cells generated by ultrasound Junru Wu * The enhancement of bone regeneration by ultrasound Lutz Claes * Cardiac imaging: The biological effects of diagnostic cardiac ultrasound Maria Grazia Andreassi http://www.sciencedirect.com/science/journal/00796107
www.sciencedirect.com

Вам также может понравиться