8

Biomass pretreatment for consolidated bioprocessing (CBP)

V. Agbor, C. Carere, N. Cicek, R. Sparling and D. Levin,    University of Manitoba, Canada

Abstract:

Biomass pretreatment and subsequent downstream processing contribute to the final cost of biocommodities produced from lignocellulosic feedstocks. Strategies that employ fewer steps for the processing of biomass can reduce costs and produce fuels and value-added products more cost effectively. This chapter reviews the various types of physico-chemical pretreatments used for lignocellulosic biomass. It describes the methods and process conditions used, as well as the physical and chemical effects of the treatment on biomass structure. Different configurations of biomass processing are presented, highlighting the increasing trend towards consolidated bioprocessing as a path to low cost biorefining for biomass-based fuels and chemicals.

Key words

biomass; pretreatment; consolidated bioprocessing; cellulolytic bacteria; biofuels

8.1 Introduction

The depletion of ‘sweet crude’ oil reserves around the world and the increasing global effort to reduce dependence on petroleum-based fuels have intensified the development of biofuels as transportation and industrial fuels. The costs of production, transportation, preconditioning and pretreatment, and subsequent conversion of these feedstocks via microbial fermentation ultimately determine the costs of biomass-based (‘cellulosic’) fuels. The cost of biomass production drives the final cost of energy production from any given feedstock (Lynd et al., 2002; Chandra et al., 2007), but pretreatment of the biomass is the second most expensive unit cost in the conversion of lignocellulose to ethanol and other chemicals (Merino and Cherry, 2007).

Industrial bioethanol production is advancing beyond grain-based ethanol production because of the energy limitations and economic/environmental concerns associated with sugar/starch-based ethanol production (Brown, 2006; Groom et al., 2008; Searchinger et al., 2008; Simpson et al., 2008). With the annual global energy demand predicted to increase to 17 billion tonnes of oil by 2035, it is evident that fuels derived from lignocellulosic biomass are an attractive and less expensive alternative for local biofuel production, compared to sugars-, starch-, or oil-based feedstocks with higher economic value.

Lignocellulosic biomass is the most abundant organic material in nature with 10–50 billion tonnes annual worldwide production (Claassen et al., 1999). Lignocellulosic biomass consists of tightly knit polymers (cellulose, hemicelluloses, lignin, waxes, and pectin) synthesized by plants as they grow. Figure 8.1 shows the inter-relationship between the three major plant polymers targeted by the biorefinery.

image
8.1 Plant biomass polymers (cellulose, hemicelluloses and lignin). The source of polymers, the inter-relatedness in the heteromatrix and the monomeric units of the polymers from a woody plant showing approximate percentages.

As a result of its complex and highly ordered structure, lignocellulosic biomass is inherently recalcitrant to bioprocessing and requires deconstruction by various pretreatments to release sugars within the biopolymers for fermentation. The term pretreatment was coined to describe any process that converts lignocellulosic biomass from its native form (which is recalcitrant to hydrolysis with cellulases) to a form that is more digestible by hydrolytic enzymes (Lynd et al., 2002). Pretreatments range from simple size reduction to more advanced biological or physico-chemical processes designed to improve the digestibility of the biomass. Table 8.1 summarizes the various physical pretreatments and their effects on biomass structure, while Table 8.2 presents a summary of the major physico-chemical pretreatments. A combination of physical and physico-chemical pretreatments is often used to improved the digestibility of lignocellulosic biomass (Agbor et al., 2011).

Table 8.1

Various physical pretreatments and their effects on biomass structure

Method Particle size (mm) Main effect on biomass References
Coarse size reduction 10–50 Increase in available surface area (Cadoche and López, 1989; Palmowski and Muller, 1999)
Chipping 10–30 Decrease heat and mass transfer limitations (Palmowski and Muller, 1999)
Grinding 0.2–2 Shearing, reduce particle size, degree of polymerization and cellulose crystallinity (Sun and Cheng, 2002; Agbor et al., 2011)
Milling (disk, hammer and ball milling) 0.2–2 Shearing, reduce particle size, degree of polymerization and cellulose crystallinity (J. Y. Zhu et al., 2009)
Use of microwaves No change Effects compaction of biomass and specific energy required for compression when used with chemicals, e.g. NaOH and water (Kashaninejad and Tabil, 2011)
Use of gamma rays No change Cleave β-1,4-glycosidic bonds, thus increasing surface area and decrease
cellulose crystallinity
(Takacs et al., 2000)

Image

Table 8.2

Summary of physico-chemical pretreatment methods

Method Process conditions Mode of action References
Steam pretreatment or steam explosion Involves rapidly heating biomass with steam at elevated temperatures (190–240°C) and pressures between 0.7 and 4.8 MPa with residence times of 3–8 min followed by explosive decompression as the pressure is released. Hemicellulose hydrolysis is thought to be mediated by the acetic acid generated from acetyl groups associated with hemicellulose and other acids released during pretreatment. The pressure is held for several seconds to a few minutes to promote hemicellulose hydrolysis, and then released. (McMillan, 1994 ) (Mosier et al., 2005) (Weil et al., 1997)
Liquid hot water pretreatment Optimally operated between 180 and 190°C, (e.g., for corn stover) and at low dry matter (&~1–8%) content, leading to more poly- and oligosaccharide production. Temperatures of 160–190°C are used for pH controlled LHW pretreatment and 170–230°C have been reported depending on the severity of the pretreatment. Hemiacetal linkages are cleaved by hot water, liberating acids during biomass hydrolysis, which facilitates the breakage of ether linkages in the biomass. (Bobleter, 1994) (Wyman et al., 2005)
Dilute acid pretreatment Dilute sulfuric acid is mixed with biomass to increase the accessibility to the cellulose in the biomass by solubilizing hemicellulose. The mixture is heated directly with the use of steam as in steam pretreatment, or indirectly via the vessel walls of the reactor. The substrate is heated to the desired temperature in an aqueous solution and pretreated using preheated sulfuric acid (concentrations of < 4 wt%) in a stainless steel reactor. In this pretreatment the dilute acid releases oligomers and monomeric sugars by affecting the reactivity of the biomass carbohydrate polymers. (Esteghalian et al., 1997) (Torget et al., 1990)
Ammonia fiber/freeze explosion, ammonia recycle percolation and soaking aqueous ammonia By bringing biomass in contact with anhydrous liquid ammonia at a loading ratio of 1:1 to 1:2 (1–2 kg of ammonia/kg of dry biomass) for 10–60 min at 60–90°C and pressures
above 3 MPa, or 150–190°C for a few minutes.
The chemical effect of ammonia or ammonia under pressure causes the cellulosic biomass to swell, thus increasing the accessible surface area while decrystallizing cellulose as the ammonia penetrates the crystal lattice to yield a cellulose–ammonia complex. (Alizadeh et al., 2005) (Kim and Lee, 2005) (Mittal et al., 2011)
Organosolv pretreatment Organosolv pretreatments are conducted at high temperatures (100–250°C) using low boiling point organic solvents (methanol and ethanol) or high boiling point alcohols (ethylene glycol, glycerol, tetrahydrofurfuryl alcohol) and other classes of organic
compounds.
Pretreatments with organic solvents extract lignin and solubilized
Hemicelluloses by hydrolyzing the internal lignin bonds, as well as the ether and 4-O-methylglucuronic acid ester bonds between lignin and hemicellulose and also by hydrolyzing glycosidic bonds in hemicellulose, and partially in cellulose depending on process conditions.
(Thring et al., 1990) (Zhao et al., 2009)
Lime pretreatment Conducted over a wide temperature range 25–130°C using 0.1 g Ca(OH)2/g biomass at low pressures. Solubilize hemicelluloses and lignin by deactylation and partial delignification. (Chang et al., 1997)
Wet oxidative pretreatment Treatment of biomass with air, water, and oxygen at temperatures above 120°C with or without a catalyst such as an alkali. Oxidative factors come into play when oxygen is introduced at high
pressures.
(Chang et al., 2001) (Galbe and Zacchi, 2007)
Carbon dioxide explosion
Pretreatment
Involves the use of supercritical carbon dioxide at high pressures (1,000–4,000 psi) at a given temperature up to 200°C for a few minutes. Carbonic acid formed from the penetration of carbon dioxide into wet biomass at high pressure helps in hemicellulose hydrolysis while the release of the pressure results in disruption of biomass. (Kim and Hong, 2001) (Zheng et al., 1995)
Ionic-liquid pretreatment Using ionic liquids at temperatures < 100°C as non-derivatizing solvents to effect dissolution of cellulose. Ionic liquids disrupt the three-dimensional network of lignocellulosic components by competing with them for hydrogen bonding. (Moultrop et al., 2005) (Zavrel et al., 2009)
Fractionation solvents Cellulose and organic solvents lignocellulose fractionation. Solvents such as phosphoric acids, sulfite or ionic liquids enable disruption to fibrillar structure of biomass and effecting cellulose crystallinty. (Z. G. Zhu et al., 2009)

Image

Pretreaments that modify the biomass composition to make it more accessible vary from neutral, to acidic, to quite alkaline (Table 8.2). Dilute acidic pretreatments will hydrolyze mostly the hemicelluloses, leaving the cellulose and lignin intact. Alkaline pretreatments will solubilize less hemicellulose and lignin than acidic pretreatments, but will alter the structural/chemical nature of the lignin, producing a hydrated cellulose product, mixed with hemicelluloses and lignin. Solvent-based pretreatment such as Organosolv will solubilize almost all of the hemicellulose, precipitate the lignin, and leave behind a purer cellulose mesh (Mosier et al., 2005; Merino and Cherry, 2007; Zhao et al., 2009).

8.2 Process configurations for biofuel production

Industrial-scale cellulosic biofuels production requires efficient, low cost processes that will ensure economic viability. The current paradigm for bioprocessing of lignocellulosic biomass into bioethanol involves a four-step process: (i) cellulase production, (ii) hydrolysis of polysaccharides, (iii) fermentation of soluble cellulose hydrolysis products, (iv) fermentation of soluble hemicellulose hydrolysis products (Lynd et al., 2002). This process has been segmented in different combinations over time to design different configurations or processing strategies to reduce the cost of biofuel production, as shown in Fig. 8.2.

image
8.2 Evolution of biomass processing strategies featuring enzymatic hydrolysis. The horizontal arrow indicates the four primary treatment processes involved with cellulosic ethanol production. The vertical arrow indicates steps taken toward increased consolidation. SHF: separate hydrolysis and fermentation, SSF: simultaneous saccharification and fermentation, SSCF: simultaneous saccharification and co-fermentation, CBP: consolidated bioprocessing. (adapted from Lynd et al., 2002)

Separate hydrolysis and fermentation (SHF) is a four-stage process with a separate biocatalyst for each. Simultaneous saccharification and fermentation (SSF) combines hydrolysis and fermentation of hexose (C6) sugars, without the pentose (C5) sugars, while simultaneous saccharification and co-fermentation (SSCF) combines cellulose hydrolysis and fermentation of both hexose and pentose sugars in one step. Currently, the majority of pilot studies and industrial processes have proceeded with separate saccharification and fermentation options for bioprocessing, featuring enzymatic hydrolysis, e.g. bioethanol production.

Consolidated bioprocessing (CBP), on the other hand, combines cellulase production and substrate hydrolysis and fermentation of the hydrolysate (both hexose and pentose sugars) in one step, thus saving the cost of investing in a multi-step process (Lynd, 1996; Lynd et al., 2002; Xu et al., 2009). Of all the reported technological advances to reduce processing costs for cellulosic ethanol, CBP has been estimated to reduce production costs by as much as 41% (Lynd et al., 2008). A techno-economic evaluation for bioethanol production from softwood (spruce), hardwood (salix), and an agricultural residue (corn stover), concluded that the process configuration (SSF) had greater impact on the cost reduction compared to the choice of substrate (Sassner et al., 2008). Hence, direct microbial conversion (DMC) or CBP of lignocellulosic biomass utilizing bacteria, may be considered a preferred method for ethanol or biological hydrogen production (Levin et al., 2006; Carere et al., 2008).

Although CBP offers the greatest potential for reducing the costs of biofuel and co-product production, a great deal remains to be realized in the development of microbial biocatalysts that can utilize cellulose and other fermentable sugars to produce the products of interest with yields that are industrially relevant. In the rest of this chapter, we discuss the rationale for, and models of, CBP as well as strategies for development of microbial biocatalysts that may improve yields of desired products from CBP-based biorefineries.

8.2.1 The rationale for CBP

Bioenergy use is propagated by incentives and successful policy interventions that have been enacted to select and encourage industrial development of renewable energy sources. Bioethanol continues to be the dominant biofuel with increasing use as a transportation fuel in Brazil, USA, Canada, and the European Union (EU). The routine ways (SHF, SSF, and SSCF) for industrial ethanol production all involve a dedicated cellulase production step. Avoiding this step has been noted to be the largest potential energy-saving step. For a given amount of pretreated biomass, it has been shown that CBP offers cost savings that are not associated with the routine methods, even at the lowest amount of cellulase loading required to produce fermentation products. However, typical CBP yields of ethanol and other end-products are much lower than obtained via traditional processes (Lynd et al., 2005, 2008).

Hydrogen (H2), on the other hand, is considered a clean energy, having a gravimetric energy density of 122 KJ g− 1 and water as the only by-product of combustion. Currently H2 is produced via energy intensive, environmentally harmful processes such as catalytic steam reformation of methane, nuclear or fossil fuel-mediated electrolysis of water, or by coal gasification (Levin et al., 2004; Carere et al., 2008). Biological H2 production via anaerobic fermentation using cellulosic substrates is an attractive process for the following reasons: (i) it is less energy intensive, (ii) it utilizes a simple process design and feedstock processing, and (iii) it has the potential to utilize agricultural or agri-industrial by-product streams (Sparling et al., 1997; Valdez-Vazquez et al., 2005; Levin et al., 2006). However, rates and yields of H2 by direct microbial conversion are low, and increasing the rates and yields remains a challenge for biological H2 production by CBP (Levin et al., 2009).

8.3 Models for consolidated bioprocessing (CBP)

The trend in industrial bioprocessing for biofuels and other industrial products, such as lactic acid, glutamic acid, n-butanol, and pinene (Hasunuma et al., 2013) is toward increased consolidation of the different process steps (as described in Fig. 8.2). Lessons drawn from nature and industry can help further research and development of CBP for biofuels and co-products, given that CBP seeks to mimic natural microbial cellulose utilization for industrial applications.

8.3.1 Ruminant or natural CBP

Many animals and insects have evolved to feed on and digest raw biomass. Some well-known cellulolytic bacteria have been isolated from these organisms as their natural habitat, e.g. Ruminococcusalbus and Clostridium termitidis inhabit the gut of ruminants and termites, respectively. By taking a close look at the highly developed ruminal fermentation of cattle (i.e., ruminant CBP = rCBP), Weimer et al. (2009) proposed that breakthroughs developed by ruminants and other already existing anaerobic systems with cellulosic biomass conversion can guide future improvements in engineered CBP (eCBP) systems. Comparing the journey of the feed through the bovine digestive tract to the transformation process of a cellulosic feedstock in a biorefinery, Weimer et al. (2009) suggest that the sliding, longitudinal movement of bovine rumination is a better physical pretreatment than conventional grinding. This is because it results in substantial increase in surface area of the plant material available for microbial attack resulting in ‘effective fiber’ properties similar to burr mills. Burr mills consume two-thirds of the energy required by hammer mills, and although they have been considered less efficient in the grinding of grain, they could be efficient in the milling of lignocellulosic biomass for CBP as a result of the ‘effective fiber’ properties generated (Weimer et al., 2009).

A great amount of effort is invested chewing the feed into a fine physically pretreated state. The feed is masticated while it is moist, another strategy supported by recent studies which show that milling after chemical treatment will significantly reduce energy consumption, reduce cost of solid–liquid separation requirements, and reduce the energy required for mixing pretreated slurries (J. Y. Zhu et al., 2009; Zhu and Pan, 2010).

Another similarity is the fact that ruminal microflora (R. albus, R. flavefaciens, and fibrobactersuccinogens) found in high numbers in the rumen are capable of rapid growth on cellulose using cellulosomal complexes similar to the well-characterized cellulosomes of Clostridium thermocellum or C. phytofermentans that are being investigated for eCBP (Lynd et al., 2002; Weimer et al., 2009). Thus, the limitation of rCBP could be explored to develop better operating parameters for eCBP. In summary, it appears that ruminants have developed an efficient and elegant physical pretreatment process that could provide insight in developing a physical pretreatment process tailored for industrial CBP, as well as serve as a model for eCBP.

8.3.2 Engineered CBP

Unlike rCBP, eCBP seeks to utilize pure cultures of specialist native cellulolytic bacteria, or recombinant cellulolytic bacteria, to convert cellulose via direct fermentation to value-added end-products. Aerobic or anaerobic microorganisms could be used for eCBP; however, the use of a separate aerobic step for cell growth is not envisioned because it is not a characteristic feature of CBP. In eCBP, much attention is dedicated to the strategic development of cellulolytic and hemicellulolytic microorganisms both for substrate utilization and end-product formation. Among the many bacteria that have been considered as CBP-enabling microorganisms, anaerobic bacteria such as C. thermocellulum, C. phytofermentans, and the aerobic yeast Saccharomyces cerevisiae have been the most investigated as potential eCBP-enabling microorganisms. Figure 8.3 compares natural and engineered bioprocessing by differentiating the different unit operations.

image
8.3 A comparison of natural and engineered consolidated bioprocessing by differentiating the different unit operations. (adapted from Weimer et al., 2009)

8.4 Microorganisms, enzyme systems, and bioenergetics of CBP

8.4.1 CBP microorganisms

Based on substrate utilization, carbohydrate hydrolyzing species represent a wide range of specialist and non-specialist microbes, and specialized microbes capable of utilizing cellulose or hemicellulose-derived sugars are preferentially selected as CBP-enabling microorganisms. Cellulolytic bacteria belong to the phyla Actinobacteria, Proteobacteria, Spirochates, Thermotogae, Fibrobacteres, Bacteriodes, and Firmicutes, but approximately 80% of the cellulolytic bacteria are found within the Firmicutes and Actinobacteria (Bergquist et al., 1999). Many of these bacteria isolated from soil, insects, ruminants, compost, and sewage have the natural ability to hydrolyze cellulose and/or hemicellulose with the majority of the reported bacteria belonging to the phylum Firmicutes, and are within the class Clostridia and the genus Clostridium. For example, Clostridium thermocellum is a Gram positive, acetogenic, obligate anaerobe with the highest known growth rate on crystalline cellulose and the most investigated as a potential CBP-enabling bacteria (Lynd et al., 2002; Xu et al., 2009).

Although cellulolytic bacteria belong to aerobic or anaerobic groups of bacteria, for large-scale CBP, anaerobiosis is advantageous because of oxygen transfer limitations that are avoided when using anaerobic bacteria (Demain et al., 2005). However, because of the increasing tendency to consolidate steps for bioethanol production, Saccharomyces cerevisiae and the fungus Tricoderma reesei are being investigated as CBP-enabling candidates for bioethanol production (van Zyl et al., 2007; Xu et al., 2009). Other candidates into which saccharolytic systems have been engineered for CBP include Zymomonas mobilis, Escherichia coli, and Klebsiella oxytoca (van Zyl et al., 2007).

8.4.2 Carbohydrate active enzyme systems

To utilize plant biomass for growth, microorganisms produce multiple enzymes that hydrolyze the cellulose, hemicellulose, and pectin polymers found in plant cell walls (Warren, 1996). As a class, these carbohydrate active enzymes are referred to as glycoside hydrolases (GHs). Extracellular GHs can be secreted freely into the environment surrounding the cell (non-complexGH systems) or they can be cell-associated in large enzyme complexes (cellulosomes). Gycoside hydrolases that specifically target cellulose include:

• endoglucanases (1,4 ß-D-glucan-4-glucanohydrolases), which cleave random internal amorphous sites of a cellulose chain producing cellulodextrins of various lengths and thus new chain ends;

• exoglucanases (including 1,4-ß-D-glucanohydrolases or cellodextrinases and 1,4-ß-D glucancellobiohydrolases, or simply cellobiohydrolase), which act in a processive manner on either the reducing and non-reducing ends of cellulose chains liberating either D-glucose (glucanohydrolase) or D-cellobiose (cellobiohydrolase) or shorter cellodextrins; and

• ß-glucosidase (ß-glucoside glucohydrolases) which hydrolyze soluble cellodextrins and cellobiose to glucose.

The ability of cellulases to hydrolyse ß-1,4-glycosidic bonds between glucosyl residues distinguishes cellulase from other glycoside hydrolases (Lynd et al., 2002).

8.4.3 Non-complex glycoside hydrolase systems

Non-complex systems consist of secreted glycoside hydrolases and generally involve fewer enzymes. Aerobic fungi of the genera Trichoderma and Aspergillus have been the focus of research for non-complex cellulase systems. Trichoderma reesei, which is the most researched non-complex cellulase system, produces at least two exoglucanases (CBH I and CBH II), five endoglucanases (EG I, EG II, EG III, EG IV, and EG V) and two ß-glucosidases (BGL I and BGL II) that act synergistically in the hydrolysis of polysaccharides. However, CBH I and CBH II are the principal components of the T. reesei cellulase system representing 60% and 20%, respectively, on a mass basis of the total protein produced (Lynd et al., 2002). The ability to produce and secrete over 100 g of cellulase per liter of culture has established T. reesei as a commercial source of cellulase enzymes (Xu et al., 2009).

8.4.4 Complex glycoside hydrolase systems

Some anaerobic cellulolytic microorganisms possess a specialized macromolecular complex of carbohydrate active enzymes known as the cellulosome. First described for C. thermocellum by Lamed et al. (1983), the cellulosome is an exocellular, multicomponent complex of GHs that mediates binding to lignocellulosic biomass and subsequent hydrolysis of the cellulose and hemicellulose polymers (Lamed et al., 1983; Carere et al., 2008). Functionally, cellulosomes are assembled on the cell walls of bacteria and enable concerted enzyme activity by minimizing distances of enzyme substrate interactions and optimizing synergies among the catalytic components, thus enabling efficient hydrolysis of the polymers and uptake of the hydrolysis products (Lynd et al., 2002).

Although cellulosome compositions can differ in the number and variety of GHs from one species to another, they generally consist of catalytic components attached to a glycosylated, non-catalytic scaffold protein that is anchored to the cell wall. In C. thermocellum, the anchor protein, known as the cellulose integrating protein (CipA), or ‘scalffoldin’, is a large (1,850 amino acid long and 2–16 MDa) polypeptide, which is anchored to the cell wall via type II cohensin domains. The C. thermocellum cellulosome contains nine GHs with endoglucanase activity (CelA, CelB, CelD, CelE, CelF, CelG, CelH, CelN, and CelP), four GHs which exhibit exoglucanase activity (CbhA, CelK, CelO, CelS), five or six GHs which exhibit xylanase activity (XynA, XynB, XynV, XynY, XynZ), one enzyme with chitinase activity (ManA), and one or two with lichenase activity (LicB). CelS, the major exoglucanase, and CelA, the major endoglucanase associated with the C. thermocellum cellulosome generate oligocellulodextrins containing two (cellobiose) to five (cellopentose) glucose residues (Lynd et al., 2002; Demain et al., 2005). The cellulosomes of some strains of C. thermocellum have been shown to degrade pectin probably via pectin lyase, polygalacturonate hydrolase, or pectin methylesterase activities. Other minor activities include ß-xylosidase, ß-galactosidase, and ß-mannosidase (Lynd et al., 2002; Demain et al., 2005). These modules have dockerin moieties that can associate with the cohesins of the scaffoldin to form the cellulosome (Fig. 8.4).

image
8.4 Structural representation of the cellulosome as a macromolecular enzyme complex on the surface of a cellulolytic bacteria, displaying the various components of a complex cellulase system.

Cellobiosephosphorylase, which hydrolyzes cellobiose and longer chain oligocellulodextrins to glucose and glucose-1-phosphate via substrate level phosphorylation, and cellodextrinphosphorylase, which phosphorylates ß-1,4-oligoglucans via phosphorolytic cleavage, have also been associated with the C. thermocellum cellulase system. Unlike the fungal cellulases, the celllosome of C. thermocellum is able to completely solubilize crystalline cellulose such as Avicel, a characteristic referred to as Avicelase or ‘true cellulase’ activity (Demain et al., 2005). Moreover, the C. thermocellum cellulase system results in the oligosaccharide hydrolysis products that are different from those of aerobic cellulolytic fungi like T. reesei, which generates cellobiose as the primary hydrolysis product (Zhang and Lynd, 2005).

8.4.5 Mode of action

Polysaccharide hydrolyzing enzymes such as cellulases and xylanases are modular proteins consisting of at least two domains: the catalytic module, and the carbohydrate binding module (Gilkes et al., 1991; Horn et al., 2012). Common features of most GH systems that effect binding to cellulose surface and facilitate hydrolysis are the carbohydrate binding modules (CBMs), which are known to have the following functions:

• CBMs play a non-catalytic role by ‘sloughing-off’ cellulose fragments from the surface of cellulosic biomass by disrupting the non-hydrolytic crystalline substrate (Lynd et al., 2002);

• they help concentrate the enzymes on the surface of the substrate (i.e., the proximity effect or phase transfer); and

• they help in substrate targeting/selectivity.

CBMs specific for insoluble cellulose are categorized as Type A CBMs, which interact with crystalline cellulose and Type B, which interact with non-crystalline cellulose (Arantes and Saddler, 2010a).

Recent studies that have focused on bacterial and fungal GHs have identified two GH families that have flat substrate-binding surfaces with the capability of cleaving crystalline polysaccharides via an oxidative reaction mechanism that depends on the presence of divalent metal ions and an electron donor. These two families are the Family 33 carbohydrate binding module (CBM33) proteins identified in bacteria and Family 61 glycoside hydrolases (GH61) from fungi (Vaaje-Kolstad et al., 2010; Horn et al., 2012). CBM33 and GH61 GHs bind to cellulose via their flat CBM substrate binding sites, which disrupt the orderly packing of the crystalline cellulose chains, creating accessible points by both introducing cuts in the polymer chains and by generating charged groups at the cut sites.

Endoglucanases and exoglucanases act synergistically. The endoglucanases generate new reducing and non-reducing ends for exoglucanases, which in turn release soluble cellodextrins and cellobiose that are converted to glucose by ß-glucosidase (Wood and McCrae, 1979; Horn et al., 2012; Kostylev and Wilson, 2012). However, for GHs to efficiently hydrolyze cellulosic biomass, they must first be able to access the cellulose chains that are tightly packed in microfibrils trapped within a heteropolymer matrix (Arantes and Saddler, 2010a; Horn et al., 2012). Factors that increase accessibility have been identified and intensely investigated (Reese, 1956; Jeoh et al., 2007; Arantes and Saddler, 2010a, 2010b; Horn et al., 2012).

8.4.6 Bioenergetics of CBP

The bioenergetics of CBP will differ with the type (aerobic, aerotolerant, or anaerobic) and number (pure, co-, or mixed-cultures) of microorganisms being considered as CBP-enabling agents. This was thought to be even more challenging for anaerobes from a bioenergetic standpoint given that ATP available from catabolism is used to support both cell growth and cellulase production (Lynd et al., 2002). An assessment of the bioenergetic benefits associated with growth on cellulosic substrates in terms of net cellular energy currency (ATP, ADP, or AMP) available for growth and cellulase production is vital for eCBP-enabling microorganisms.

A comprehensive bioenergtic model validating the bioenergetic feasibility of employing C. thermocellum on crystalline cellulose was reported by Zhang and Lynd (2005), who determined that C. thermocellum assimilates oligo-cellodextrins (G2–G6) of mean chain length of n ≈ 4 (where n = degree of polymerization of glucose (G) moieties). The oligocellulodextrins are imported into the cell and then cleaved by substrate level phosphorylation by cellodextrin- and cellobiose-phosphorylases. Phosphorylation results in cleavage of ß-glucosidic bonds releasing glucose and glucose-6-phosphate, that undergo glycolysis via the Emden–Meyerhoff pathway to generate ATP. Assimilation of oligo-cellodextrins with an average of 4.2 glucose units more than compensates for higherATP expended on cellulase synthesis when C. thermocellum is grown on cellulose compared to cellobiose (Zhang and Lynd, 2005). Thus, the anaerobic fermentation of cellulose using C. thermocellum as a CBP-enabling microorganism is bioenergetically feasible, without the need for added saccharolytic enzymes.

8.5 Organism development

Although CBP of cellulosic biomass offers great potential for lower cost biofuels and fermentation products, robust, industrial microorganisms capable of both high rates of substrate conversion and high yields of the desired fermentation end-products are not available. Desirable characteristics of CBP-enabling microorganisms include production of highly active GH enzymes for rapid substrate hydrolysis, transport and utilization of the resulting hydrolytic products, high product selectivity and yield. Considerable efforts are underway to identify natural isolates with the desired characteristics and/or to develop stains with the desired characteristics via genetic engineering. The former strategy involves using naturally occurring cellulolytic microorganisms to improve end-product properties related to product yield, tolerance and titre. A classical approach is to metabolically influence end-product yield and solvent tolerance in anaerobic cellulolytic Clostridia. The latter strategy involves the use of genetic engineering of non-cellulolytic microorganisms. The best example of this is the engineering of S. cerevisiae, which naturally exhibits high product yields and solvent tolerance, to express a heterologous GH system that enables it to hydrolyze cellulose/hemicellulose or utilize sugars derived from hemicellulose hydrolysis (Lynd et al., 2002, 2005). The vast majority of R&D towards organism development is focused on either bacteria or yeast as primary candidates for CBP-enabling microbes. However, the use of cellulolytic non-unicellular fungi as CBP has also been proposed (Xu et al., 2009). This strategy can be classified under the native cellulolytic strategy from the proponents of CBP. Figure 8.5 shows the organism development strategies and commonly employed CBP-enabling microorganisms.

image
8.5 Organism development strategies employed in research development for a CBP-enabling microorganism.

8.5.1 Metabolic engineering

Bailey (1991) defined metabolic engineering (ME) as the improvement of cellular activities by manipulation of enzymatic, transport, and regulatory functions of the cell with recombinant DNA technology. Metabolic engineering tools being used in the quest for the development of ethanologenic and currently hydrogenic microorganisms include the following:

• Mutagenesis via homologous recombination involving the mutation of a target gene that encodes a native protein to downregulate the expression of another protein or results in the synthesis of an undesired or inactive protein.

• Heterologous gene expression as a means of manipulating the metabolic fluxes toward the synthesis of a desired end product. This is the most likely metabolic engineering strategy amenable to biofuels and overexpression of enzyme catalysis flow past forks to the desired end product is a common strategy (Carere et al., 2008). Pyruvate overflow in Clostridium cellulolyticum was established to be as a result of the inability of pyruvate-ferredoxinoxido-reductase to metabolize pyruvate to acetyl-CoA, resulting in reduced cell growth and increased lactate. However, heteologous expression of pyruvate decarboxylase and alcohol dehydrogenase from Zymomonas mobilis into Clostridium cellulolyticum resulted in 93% acetate, 53% ethanol and hydrogen yield increased by more than 75% thus showing that cellulose fermentation can be improved by using genetically engineered strains of cellulolytic Clostridia (Guedon et al., 1999).

• Antisense RNA (asRNA) attempts in redirecting metabolic flow by targeting the same genes as in mutagenesis but instead of completely abolishing protein activity as in mutagenesis, asRNA aims explicitly at downregulating the expression of a native protein by inhibiting translation due to duplex RNA structure blocking the ribosome binding site or rapid degradation of mRNA by RNases specific for RNA duplex, or by the inhibition of mRNA transcription due to premature termination. By so doing, asRNA avoids potentially lethal mutations and can be used to inducibly repress expression of proteins by using inducible promoters for asRNA. This strategy was used to reduce levels of enzymes responsible for butyrate formation in Clostridium acetobutylicum, demonstrating that asRNA can be used to downregulate specific protein, thus redirecting metabolic flux (Desai and Papoustakis, 1999; Carere et al., 2008).

8.5.2 Natural versus engineered GH systems

The development of bacteria and fungi for CBP has focused mostly on the use of microorganisms that naturally express GH systems to hydrolyze cellulose/hemicellulose and synthesize products of interest from the hydrolysis products. Metabolic engineering of anaerobic cellulolytic bacteria has been the primary approach for enhancing the yields of the desired products so that they can meet the requirements of an industrially consolidated bioprocess. Gene transfers systems, electrotransformation protocols, and recombinant strains with enhanced product synthesis profiles have been described for both C. cellulolyticum and C. thermocellum.

Previous studies have shown that cellulose utilization by the mesophilic C. cellulolyticum is strongly dependent on initial cellulose concentration, which ultimately affects carbon flow distribution leading to end products. And the cessation of early growth was as a result of pyruvate overflow during high carbon flux (Guedon et al., 1999; Desvaux et al., 2000). Increased levels of less reduced metabolite, ethanol and lactate were observed with high levels of carbon flux, whereas at a low carbon flux, pyruvate is oxidized preferentially to acetate and lactate, thus showing an innate capability to balance carbon and electron flow or generation of reducing equivalents (Guedon et al., 1999). However, a decrease in the accumulation of pyruvate at high carbon flux was achieved by heterologous expression of pyruvate decarboxylase (PDC) and alcohol dehydrogenase (ADH) from Zymomonas mobilis in a shuttle vector pMG8. Growth of recombinant strain resulted in a 150% increase in cellulose utilization, 180% increase in dry cell weight, 48% decrease in lactate production, 93% increase in acetate and 53% increase in ethanol over the wild C. cellulolyticum, proving that genetically engineered strains could be used to greatly increase yields of cellulose fermentation by C. cellulolyticum (Guedon et al., 2002).

To show the potential of using C. thermocellum as robust platform organism for CBP, Argyros et al. (2011) constructed a mutant with novel genetic engineering tools that allow for the creation of unmarked mutations while using a replicating plasmid. A counter selection strategy was used to delete genes for lactate dehydrogenase (Ldh) and phosphotransacetylase (Pta) resulting in a stable strain with 40:1 ethanol selectivity and a 4.2-fold increase in ethanol yield over the wild-type strain (Argyros et al., 2011).

Expression of heterologous cellulases in non-cellulolytic microorganisms that are known to possess desired product formation characteristics, such as faster sugar consumption, higher ethanol yield and high resistance to ethanol and fermentation inhibitors, has also been accomplished (Hasunuma and Kondo, 2012). For example, genes encoding endoglucanse II (EG II) cellobiohydrolase II (CB II) from T. reesei and beta-glucosidase BGL 1 from Aspergilus aculeatus were integrated into the chromosome of wine yeast strain using a single vector conferring resistance to antibiotics G418. The mutant strain was able to hydrolyze corn stover cellulose and produced ethanol without the addition of exogenous saccharolytic enzymes (Khramtsov et al., 2011). Significant advances related to recombinant enzyme expression support the potential of S. cerevisiae as CBP host, and the number of genes expressed is not probably as important as the metabolic burden and stress responses associated with such high-level expression (van Zyl et al., 2007).

Heterologous expression of cellulolytic enzymes for the development of a cell surface which provides display of cellulolytic enzymes or cellulases that are secreted is currently being investigated in other non-cellulolytic, ethanologenic bacteria such as E. coli, Zymomonas mobilis and Klebsiella oxytoca to enable growth and fermentation of pretreated lignocellulosic biomass (Jarboe et al., 2007; van Zyl et al., 2007).

8.6 Conclusion

CBP is a less energy intensive method and a potential low cost route for production of cellulosic ethanol, as well as other industrially important products, because of the avoided cost of exogenous enzymes required for cellulose hydrolysis in SHF, SSF and SSCF (Lynd et al., 2008; Weimer et al., 2009; Xu et al., 2009). The saccharification and fermentation steps in SHF and SSF have large differences in operating temperatures which complicate development of pilot- and industrial-scale processes compared to CBP, which is conducted in a single vessel at a single optimized temperature. CBP offers simplification of the total operation process for ethanol production from cellulosic biomass compared to SHF and SSF (Hasunuma and Kondo, 2012; Hasunuma et al., 2013). CBP also has the added benefit of requiring minimal pretreatment of lignocellulosic biomass, because pretreated feedstocks for CBP do not need to be completely saccharified with costly, huge volumes of exogenous enzymes, thus the cost of pretreatments, which is a key bottleneck in lowering the net cost of production of cellulosic bioethanol, is kept very low (Xu et al., 2009; Agbor et al., 2011).

Although suitable for the production of high value products and low cost fuels, the quest for a suitable industrial CBP-enabling microorganism limits the impact of the process technology design for industrial purposes compared to sequential step processes. While the use of industrial yeast and bacterial strains used in conventional SHF and SSF processes is well established, the use of natural and/or engineered microorganisms in CBP is not yet mature and hence industrial uptake has been slow. However, research interest in CBP is growing and production of different products via CBP is under investigation (Lynd et al., 2008; Xu et al., 2009; Hasunuma and Kondo, 2012).

8.7 References

1. Agbor VB, Sparling R, Cicek N, Berlin A, Levin DB. Biomass pretreatments: fundamentals toward application. Biotechnology Advances. 2011;5:1–11.

2. Alizadeh H, Teymouri F, Gilbert TI, Dale BE. Pretreatment of switchgrass by ammonia fibre explosion (AFEX). Applied Biochemistry and Biotechnology. 2005;121–124:1133–1141.

3. Arantes V, Saddler JN. Access to cellulose limits the efficiency of enzymatic hydrolysis: the role of amorphogenesis. Biotechnology for Biofuels. 2010a;3:4.

4. Arantes V, Saddler JN. Cellulose accessibility limits the effectiveness of minimum cellulase loading on the efficient hydrolysis of pretreated lignocellulosic substrates. Biotechnology for Biofuels. 2010b;4:3.

5. Argyros DA, Tripathi SA, Barrett TF, et al. High ethanol titers from cellulose by using metabolically engineered thermophilic, anaerobic microbes. Applied and Environmental Microbiology. 2011;77:8288–8294.

6. Bailey JE. Toward a science of metabolic engineering. Science. 1991;252:1668–1674.

7. Bergquist PL, Gibbs MD, Morris DD, Te’o VS, Saul DJ, Morgan HW. Molecular diversity of thermophillic cellulolytic and hemicellulolytic bacteria. FEMS Microbiology Ecology. 1999;28:99–147.

8. Bobleter O. Hydrothermal degradation of polymers derived from plants. Progress in Polymer Science. 1994;19:797–841.

9. Brown L. Exploding US grain demand for automotive fuel threatens world food security and political stability Washington, DC: Earth Policy Institute; 2006.

10. Cadoche L, López GD. Assesment of size reduction as a preliminary step in the production of ethanol from lignocellulosic wastes. Biological Wastes. 1989;30:153–157.

11. Carere CR, Sparling R, Cicek N, Levin DB. Third generation biofuels via direct cellulose fermentation. International Journal of Molecular Sciences. 2008;9:1342–1360.

12. Chandra RP, Bura R, Mabee WE, Berlin A, Pan X, Saddler JN. Substrate pretreatment: the key to effective enzymatic hydrolysis of lignocellulosics? Advances in Biochemical Engineering/Biotechnology. 2007;108:67–93.

13. Chang VS, Burr B, Holtzapple MT. Lime pretreatment of switchgrass. Applied Biochemistry and Biotechnology. 1997;63–65:3–19.

14. Chang VS, Nagwani M, Kim CH, Holtzapple MT. Oxidative lime pretreatment of high-lignin biomass – poplar wood and newspaper. Applied Biochemistry and Biotechnology. 2001;94:1–28.

15. Claassen P, Van Lier J, Lopez Contreras A, et al. Utilisation of biomass for the supply of energy carriers. Applied Microbiology and Biotechnology. 1999;52:741–755.

16. Demain AL, Newcomb M, Wu JHD. Cellulase, clostridia, and ethanol. Microbiology and Molecular Biology Reviews. 2005;69:124–154.

17. Desai RP, Papoustakis ET. Antisense RNA strategies for metabolic engineering of Clostridium acetobutylicum. Applied and Environmental Microbiology. 1999;65:936–945.

18. Desvaux M, Guedon E, Petitdemange H. Cellulose catabolism by Clostridium cellulolyticum growing in batch culture on defined medium. Applied and Environmental Microbiology. 2000;66:2461–2470.

19. Esteghalian A, Hashimoto AG, Fenske JJ, Penner MH. Modelling and optimization of dilute-sulfuric-acid pretreatment of corn stove, poplar and switchgrass. Bioresource Technology. 1997;59:129–136.

20. Galbe M, Zacchi G. Pretreatment of lignocellulosic materials for efficient bioethanol production. Advances in Biochemical Engineering/Biotechnology. 2007;108:41–65.

21. Gilkes NR, Henrissat B, Kilburn DG, Miller RCJ, Warren RAJ. Domains in microbial β-1,4-glycanases: sequence conservation, function, and enzyme families. Microbiological Reviews. 1991;55:303–315.

22. Groom MJ, Gray EM, Townsend PA. Biofuels and biodiversity: principles for creating better policies for biofuel production. Conservation Biology. 2008;22:602–609.

23. Guedon E, Payot S, Desvaux M, Petitdemange H. Carbon and electron flow in Clostridium cellulolyticum grown in chemostat culture on synthetic medium. Journal of Bacteriology. 1999;181:3262–3269.

24. Guedon E, Desvaux M, Petitdemange H. Improvement of cellulolytic properties of Clostridium cellulolyticum by metabolic engineering. Applied and Environmental Microbiology. 2002;68:53–58.

25. Hasunuma T, Kondo A. Consolidated bioprocessing and simultaneous saccharification and fermentation of lignocellulose to ethanol with thermotolerant yeast strains. Process Biochemistry. 2012;47:1287–1294.

26. Hasunuma T, Okazaki F, Okai N, Hara KY, Ishii J, Kondo A. A review of enzymes and microbes for lignocellulosic biorefinery and the possibility of their application to consolidated bioprocessing technology. Bioresource Technology. 2013;135:513–522.

27. Horn SJ, Vaaje-Kolstad G, Westereng B, Eijsink VGH. Novel enzymes for the degradation of cellulose. Biotechnology for Biofuels. 2012;5:45.

28. Jarboe L, Grabar T, Yomano L, Shanmugan K, Ingram L. Development of ethanologenic bacteria. Advances in Biochemical Engineering/Biotechnology. 2007;108:237–261.

29. Jeoh T, Ishizawa CI, Davis MF, Himmel ME, Adney WS, Johnson DK. Cellulase digestibility of pretreated biomass is limited by cellulose accessibility. Biotechnology and Bioengineering. 2007;98:112–122.

30. Kashaninejad M, Tabil LG. Effect of microwave–chemical pre-treatment on compression characteristics of biomass grinds. Biosystems Engineering, pp. 2011;36–45.

31. Khramtsov N, McDade L, Amerik A, et al. Industrial yeast strain engineered to ferment ethanol from lignocellulosic biomass. Bioresource Technology. 2011;102:8310–8313.

32. Kim HK, Hong J. Supercritical CO2 pretreatment of lignocellulose enhances enzymatic cellulose hydrolysis. Bioresource Technology. 2001;77:139–144.

33. Kim TH, Lee YY. Pretreatment of corn stover by ammonia recycle percolation process. Bioresource Technology. 2005;96:2007–2013.

34. Kostylev M, Wilson DB. Synergistic interactions in cellulose hydrolysis. Biofuels. 2012;3:61–70.

35. Lamed R, Setter E, Bayer EA. Characterisation of a cellulosebinding, cellulose-containing complex in Clostridium thermocellum. Journal of Bacteriology. 1983;156:828–836.

36. Levin DB, Pitt L, Love M. Biohydrogen production: prospects and limitations to practical application. International Journal of Hydrogen Energy. 2004;29:173–185.

37. Levin DB, Islam R, Cicek N, Cicek R. Hydrogen production by Clostridium thermocellum 27405 from cellulosic biomass substrates. International Journal of Hydrogen Energy. 2006;31:1496–1503.

38. Levin DB, Carere CR, Cicek N, Sparling R. Challenges for biohydrogen production via direct lignocellulose fermentation. International Journal of Hydrogen Energy. 2009;34:7390–7403.

39. Lynd LR. Overview and evaluation of fuel ethanol from cellulosic biomass: technology, economics, the environment, and policy. Annual Review of Energy and the Environment. 1996;21:403–465.

40. Lynd LR, Weimer PJ, Zyl WHV, Pretorius IS. Microbial cellulose utilization: fundamentals and biotechnology. Microbiology and Molecular Biology Reviews. 2002;66:506–577.

41. Lynd LR, Van Zyl WH, McBride JE, Laser M. Consolidated bioprocessing of cellulosic biomass: an update. Current Opinion in Biotechnology. 2005;16:577–583.

42. Lynd LR, Laser MS, Bransby D, et al. How biotech can transform biofuels. Nature Biotechnology. 2008;26:169–172.

43. McMillan JD. Pretreatment of lignocellulosic biomass. Enzymatic Conversion of Biomass for Fuels Production. 1994;566:292–324.

44. Merino ST, Cherry J. Progress and challenges in enzyme development for biomass utilization. Advances in Biochemical Engineering/Biotechnology. 2007;108:95–120.

45. Mittal A, Katahira R, Himmel ME, Johnson DK. Effects of alkaline pretreatment or liquid ammonia treatment on crystalline cellulose: changes in crystalline structures and effects. Biotechnology for Biofuels. 2011;4:41.

46. Mosier N, Wyman CE, Dale BE, et al. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresource Technology. 2005;96:673–686.

47. Moultrop JS, Swatloski RP, Moyna G, Rogers RD. High resolution 13C NMR studies of cellulose and cellulose oligomers in ionic liquid solutions. Chemical Communications. 2005;12:1557–1619.

48. Palmowski L, Muller J. Influence of the size reduction of organic waste on their anaerobic digestion. In: II International Symposium on Anaerobic Digestion of Solid Waste. Barcelona, Spain 15–17 June 1999:137–144.

49. Reese ET. Enzymatic hydrolysis of cellulose. Applied Microbiology. 1956;4:39–45.

50. Sassner P, Galbe M, Zacchi G. Techno-economic evaluation of bioethanol production from three different lignocellulosic materials. Biomass and Bioenergy. 2008;32:422–430.

51. Searchinger T, Heimlich R, Houghton RA, et al. Use of US croplands for biofuels increases greenhouse gases through emissions from land-use change. Science. 2008;319:1238–1240.

52. Simpson TW, Sharpley AN, Howarth RW, Paerl HW, Mankin KR. The new gold rush: fueling ethanol production while protecting water quality. Environmental Quality. 2008;37:318–324.

53. Sparling R, Risbey D, Poggi-Varaldo HM. Hydrogen production from inhibited anaerobic composters. International Journal of Hydrogen Energy. 1997;22:563–566.

54. Sun Y, Cheng J. Hydrolysis of lignocellulosic materials for ethanol production: a review. Bioresource Technology. 2002;83:1–11.

55. Takacs E, Wojnarovits L, Foldavary C, Hargagittai P, Borsa J, Sajo I. Effect of combined gamma-radiation and alkali treatment on cotton-cellulose. Radiation Physics and Chemistry. 2000;57:399–403.

56. Thring RW, Chornet E, Overend R. Recovery of a solvolytic lignin: effects of spent liquor/acid volume ratio, acid concentrated and temperature. Biomass. 1990;23:289–305.

57. Torget RW, Werdene P, Grohmann K. Dilute acid pretreatment of two short-rotation herbaceous crops. Applied Biochemistry and Biotechnology. 1990;24(25):115–126.

58. Vaaje-Kolstad G, Westereng B, Horn SJ, et al. An oxidative enzyme boostying enzymatic conversion of recalcitrant polysaccharides. Science. 2010;330:219–222.

59. Valdez-Vazquez I, Sparling R, Risbey D, Rinderknecht-Seijas N, PoggiVaraldo HM. Hydrogen production via anaerobic fermentation of paper mill waste. Bioresource Technology. 2005;96:1907–1913.

60. Van Zyl WH, Lynd LR, Den Haan R, McBride JE. Consolidated bioprocessing for bioethanol production using Saccharomyces cerevisiae. Biofuels. 2007;108:205–235.

61. Warren RAJ. Microbial hydrolysis of polysaccharides. Annual Reviews Microbiology. 1996;50:183–212.

62. Weil JR, Sariyaka A, Rau SL, et al. Pretreatment of yellow poplar wood sawdust by pressure cooking in water. Applied Biochemistry and Biotechnology. 1997;68:21–40.

63. Weimer PJ, Russell JB, Muck RE. Lessons from the cow: what the ruminant animal can teach us about consolidated bioprocessing of cellulosic biomass. Bioresource Technology. 2009;100:5323–5331.

64. Wood TM, McCrae SI. Synergism between enzymes involved in the solubilization of native cellulose. Advances in Chemistry. 1979;181:181–209.

65. Wyman CE, Dale BE, Elander RT, Holtzapple MT, Ladisch MR, Lee YY. Coordinated development of leading biomass pretreatment technologies. Bioresource Technology. 2005;96:1959–1966.

66. Xu Q, Singh A, Himmel ME. Perspectives and new directions for the production of bioethanol using consolidated bioprocessing of lignocellulose. Current Opinion in Biotechnology. 2009;20:364–371.

67. Zavrel M, Bross D, Funke M, Buchs J, Spiess AC. High-throughput screening for ionic liquids dissolving (ligno-)cellulose. Bioresource Technology. 2009;100:2580–2587.

68. Zhang YHP, Lynd LR. Cellulose utilization by Clostridium thermocellum: bioenergetics and hydrolysis product assimilation. National Academy of Science. 2005;102:7321–7325.

69. Zhao X, Cheng K, Liu DH. Organosolv pretreatment of lignocellulosic biomass for enzymatic hydrolysis. Applied Microbiology and Biotechnology. 2009;82:815–827.

70. Zheng YZ, Lin HM, Tsao GT. Supercritical carbon-dioxide explosion as a pretreatment for cellulose hydrolysis. Biotechnology Letters. 1995;17:845–850.

71. Zhu JY, Pan XJ. Woody biomass pretreatment for cellulosic ethanol: technology and energy consumption evaluation. Bioresource Technology. 2010;100:4992–5002.

72. Zhu JY, Wang GS, Pan XJ, Gleisner R. Specific surface to evaluate the efficiencies of milling and pretreatment of wood for enzymatic saccharification. Chemical Engineering Science. 2009a;64:474–485.

73. Zhu ZG, Sathitsuksanoh N, Vinzant T, Shell DJ, McMillan JD, Zhang YH. Comparaitive study of corn stover preteated by dilute acid and cellulose solvent-based lignocellulose fractionation: enzymatic hydrolysis, supramolecular structure, and substrate accessibility. Biotechnology and Bioengineering. 2009b;103:715–724.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset