9

Developments in bioethanol fuel-focused biorefineries

S. Mutturi, B. Palmqvist and G. Lidén,    Lund University, Sweden

Abstract:

The production of ethanol from lignocellulose is on its way to industrial realization as evidenced by the current completion of several commercial scale plants. In these plants, biomass can be utilized for a range of different products apart from ethanol, such as electricity, biogas and heat. In this chapter, we describe the technology for production of ethanol within these biorefineries and discuss the challenges and development trends. The focus of this chapter is on the biochemical conversion routes.

Key words

pretreatment; hydrolysis; fermentation; lignocellulose; co-products

9.1 Introduction

Ethanol is clearly a significant product in the world of biorefineries. Alternatively, you could also say that biorefineries are increasingly important in the production of ethanol. In the former case, ethanol is one of a range of products in the processing of biomass. In the latter case, biorefineries are used to increase the overall profitability (or feasibility) of ethanol production from biomass (Pham and El-Halwagi, 2012). In this review we will start from the latter perspective, i.e. we already know (or have decided) that ethanol is a desired product and we want to enable the feasibility, improve the sustainability or maximize the profitability of biorefineries. All product streams and substrate streams are clearly of interest, just as in a petrochemical refinery, from which we borrow ‘refinery’ in the term ‘biorefinery’.

Ethanol is today mainly produced from sugars (obtained from sugar cane primarily) or starch (obtained from corn primarily). The growth in ethanol production worldwide has been impressive since the mid-1970s, driven almost exclusively by the two dominating producers, the US and Brazil, and with the purpose of producing ethanol as a fuel (Fig. 9.1). The first phase of the increase in ethanol production was a result of the Brazilian Pro-Alcohol program launched in 1975, which caused an impressive growth in ethanol production between 1975 and 1985 (Goldemberg, 2006). The starting point of the program was the oil embargo in 1973, which was reinforced by the realization that Brazil had a huge unused capacity as a sugar cane producer. The first decade of rapid ethanol production growth was followed by a period of somewhat slower growth – and even a short period of decreased world production – until about the year 2000. At this point in time, the expansion of US ethanol production started. Ethanol production in the US surpassed that of Brazil in 2005 and the production gap has widened since then.

image
9.1 World total ethanol production since 1975 together with the two major fuel ethanol producers, US and Brazil. (Compiled from: Dr Berg, F. O. Litchs, the Renewable Fuels Association (http://www.ethanolrfa.org/) and Goldemberg, 2006).

The history of ethanol as a fuel, however, dates much further back than the 1970s. The first use of ethanol was likely as a mixture with pine-derived turpentine in oil lamps in the 1860s (Songstad et al., 2009). Whale oil was running out and an alternative fuel was needed. At about this time, Nicholas Otto was working on his combustion engine using a mixture containing ethanol. Decades later, Henry Ford had a vision of using ethanol derived from biomass as a motor fuel. Henry Ford allegedly shared the views of the so-called ‘Chemurgy’ movement1 in the 1930s, which promoted the production of chemicals from agriculture (Giebelhaus, 1980). The people in the Chemurgy movement were, in a sense, early proponents of biorefineries. However, petroleum-derived fuels could be produced more cheaply and ethanol was soon outcompeted as a fuel. So with the exception of the two world wars, the interest in ethanol as a vehicle fuel stayed low until the oil crisis in 1973. Apart from previous periodic shortages of petroleum on the market (and the risk of a more permanent future shortage), there are two other major drivers behind production of ethanol and other chemicals from renewable resources, namely environmental concerns and agricultural policies. These driving forces are interconnected and difficult to resolve.

It is probably fair to conclude that environmental concerns for climate changes have had a significant impact in the introduction of renewable fuel targets in both Europe and the US. These targets have been formulated in the Renewable Energy Directive (RED) in the European Union (Commission Directive 2009/28/EC), and the Energy Independence and Security Act (EISA, 2007) in the US. The former puts forward the so-called 20-20-20 targets, i.e. a 20% reduction in EU greenhouse gas emissions (compared to 1990 levels); an increased share of EU energy consumption produced from renewable resources to 20%, and finally a 20% improvement in energy efficiency within the EU – all to be reached by 2020. Specifically, a target of 10% renewable fuels is stated. The EISA has instead a very specific absolute production target, calling for 36 billion gallons (136 billion liters) per year of renewable biofuels by 2022, out of which a maximum of 15 billion gallons can be corn-based ethanol.

The EU targets on greenhouse gas (GHG) emissions have led to much activity on standardization of calculation of ‘field-to-exhaust pipe’ GHG emissions using life cycle analysis (LCA), and there are now ISO standards for LCA (14040 and 14044). The requirements for a net reduction of GHG emissions are gradually increased in the coming decade and suppliers of renewable fuels must make calculations of GHG emissions using the standards. A net GHG reduction of 60% in comparison to fossil fuels should be reached by 2020, which cannot typically be met by today’s starch-based ethanol production, and will favor production from lignocellulose. There is thus presently a strong regulatory incentive to push ethanol production from starch-based into lignocellulose-based production. Such a production will necessarily require efficient use of all parts of the raw material, i.e. a biorefinery approach is called for.

9.2 Ethanol biorefineries

A biorefinery can be depicted as in Fig. 9.2, i.e. it is a processing facility for producing a multitude of product based on a renewable carbon source. The International Energy Agency (IEA) states that ‘Biorefinery is the sustainable processing of biomass into a spectrum of marketable products (food, feed, materials, chemicals) and energy (fuels, power, heat)’. The span of products and processes in a biorefinery is therefore very large and a sub-classification is often made (Kamm and Kamm, 2007). This classification can be based on:

• the type of substrate used (e.g., agricultural material, forest materials, waste and residues, or algae)

• the type of product obtained (e.g., fuel, heat, or electricity)

• the type of conversion processes used (thermochemical, biochemical or pyrolysis).

image
9.2 Conceptual picture of a biorefinery, showing the main technology choices.

The IEA through its Working Group 42 introduced a fourth ground for classification, namely the so-called platforms, which are based on core intermediates in the processing, e.g. C5/C6 carbohydrates, syngas, lignin or pyrolysis oil (http://www.iea-bioenergy.task42-biorefineries.com).

The variation in technology can be quite large, but there are still enough common features to justify the use of the term biorefinery. Similar to the oil refinery, fuels and heat are prime products and out of the fuels, ethanol is a major target product. Ethanol is of course also a chemical per se, and a potential platform chemical, which can be used for production of, for example, ethylene to be further used in the polymer industry. The Brazilian company, Braskem has launched the production of ‘green polyethylene’ using this approach (http://www.braskem.com.br/plasticoverde/eng/default.html).

By ‘ethanol biorefinery’ we mean here a biomass-based refinery, in which ethanol is a major product. There are several options in terms of feedstocks, technology, products and platforms. Ethanol can be a product of the syngas route, where the syngas is obtained through thermochemical processing, followed by either Fischer–Tropsch synthesis using the syngas or fermentative production from syngas using various kinds of Clostridia, such as Clostridium ljungdahlii, Clostridium carboxidivorans, or other proprietary strains, e.g. Clostridium P11 (Kundiyana et al., 2010; Wilkins and Atiyeh, 2011; Köpke et al., 2011). This route is currently pursued by companies such as Coskata and Ineos Bio. A principal advantage of the syngas route is the potential to use also the lignin fraction for ethanol production (of course at the same time withdrawing lignin from other potential uses). A drawback is that ethanol is often formed together with acetate, and scale-up hinges on efficient gas–liquid mass transfer (Köpke et al., 2011).

In this chapter we will focus on the biochemical conversion route via the so-called sugar platform (i.e., the option which is within the dashed box in Fig. 9.2). The choice of feedstock is obviously critical. The availability of the feedstock will define and limit the production capacity and will also be geographically specific (Table 9.1). Furthermore, the net GHG emissions are to a very large extent determined by the feedstock itself – or rather the agricultural practices associated with it – and to a lesser degree by the conversion process.

Table 9.1

Classification and availability of various lignocellulose sources

Biomass type Species Typical growth region Reported productivities
Hardwood Birch (Betula spp.)
Eucalyptus (Eucalyptus spp.)
Northern hemisphere
Australia, New Guinea, Indonesia, Europe
Spain:
13.9–14.6 T/ha/year (E. globulus)
20.4–21.5 T/ha/year (E. nitens)
(Pérez-Cruzado et al., 2011)
Willow (Salix spp.) Northern hemisphere Worldwide average chip production:
10 T/ha/year (González-García et al., 2012b)
Denmark: 11–22 T/ha/year (Callesen et al., 2010)
Poplar (Populus spp.)  Europe and North Central USA:
2–11 T/ha/year (Amichev et al., 2010)
Aspen (Populus tremula L.)  Scandinavia average:
7.9–9.5 T/ha/year (Tullus et al., 2009)
Softwood Douglas Fir (Pseudotsuga menziesii) North America  
Norway Spruce (Picea abies) Northern Europe Sweden: 5–9 T/ha/year (Bergh et al., 2005)
Pine (Pinus spp.) Northern hemisphere, Chile, Australia, New Zealand Australia: 17–39 T/ha/year (Snowdon and Benson, 1992)
Dedicated crops Switch grass (Panicum virgatum) North America  
Miscanthus (Miscanthus giganteus) Europe, Africa, South Asia  
Giant Reed (Arundo donax L.) Mediterranean, South Asia Italy: 37.7 T/ha/year (Angelini et al., 2009)
Cassava pulp (Manihot esculenta) Africa (Nigeria), South East Asia  
Hemp (Cannabis sativa L) Canada, Europe  
Bamboo (Phyllostachys spp.,
Bambusa bambos, Thyrsostachys siamensis)
South Asia (India, Thailand)
East Asia (China, Japan)
 
Crop residues Wheat straw Asia, Europe, USA  
Rice straw Asia  
Corn stover USA, China  
Corn cobs USA, China  
Oat straw North-western Europe, Mid-eastern Africa  
Sugarcane bagasse Brazil, India, China  
Barley straw Europe  
Cotton stalk Asia  
Sweet sorghum bagasse South Asia Central America, Africa  

Image

9.3 The lignocellulose to ethanol process

The main process steps in any biochemical ethanol refinery are (Fig. 9.3):

• pretreatment

• hydrolysis

• fermentation

• product recovery

• wastewater purification.

image
9.3 Schematic process overview of a bioethanol-focused biorefinery. After pretreatment, the two conversion steps take place: first, degradation of cellulose into monomeric sugars, and second, fermentation of the sugars to ethanol. Traditionally this is done in either a separate hydrolysis and fermentation (A) or a simultaneous saccharification and fermentation (B) set-up.

Co-products will vary a lot depending on feedstock and market conditions, but in an energy focused biorefinery, co-products will be heat, electricity, pellets and biogas, as shown in Fig. 9.3. The relative proportion between these co-products can be changed – within certain limits – and this flexibility is an important feature of a biorefinery. We will now go through the basic process steps in more detail.

Although the same main process steps in a bioethanol biorefinery are always needed, the way to operate each individual step is highly dependent on desired products and to a large extent on the selected feedstock. Biomass is composed of three major macromolecules, namely cellulose, hemicelluloses and lignin, with smaller fractions of proteins, extractives and ash. However, the relative ratios and the composition of, in particular, hemicellulose and lignin differ greatly between different types of biomass, as indicated in Table 9.2. The main sugar in the branched hemicellulose polymer is, for example, xylose in most hardwoods and agricultural crops, whereas mannose is the dominant sugar in softwood hemicellulose. The cellulose polymer, in contrast, is always comprised of repeating cellobiose units (i.e., glucose dimers) regardless of biomass type.

Table 9.2

Composition of various lignocellulose feedstocks

Type Plant Glucan Xylan Arabinan Mannan Galactan Acetyl Lignina Extractivesb Reference
Hardwood Birch 38.2 18.5 NR 1.2 NR NR 22.8 2.3 Hayn et al., 1993
Willow 43.0 14.9 1.2 3.2 2.0 2.9 24.2 NR Sassner et al., 2006
Poplar 49.9 17.4 1.8 4.7 1.2 NR 18.1 NR Wiselogel et al., 1996
Red Maple 41.9 19.3 0.8 NR NR NR 24.9 NR Jae et al., 2010
Eucalyptus 42.9c 12.7c 2.3c 0.9c 2.2c NR 16.7 19.2 Vázquez et al., 2007
 46.1 17.1 0.8 0.4 1.5 NR 19.8 0.6 Rencoret et al., 2010
Aspen 45.9 16.7 0.0 1.2 0.0 NR 23.0 NR Youngblood et al., 2010
Softwood Douglas Fir 43.0 3.0 1.0 13.0 2.0 NR 28.0 NR Mabee et al., 2006
 45.5 3.1 0.7 12.7 4.3 NR 30.6 3.8 Johansson, 2010
Spruce 43.4 4.9 1.1 12.0 1.8 NR 28.1 1.0 Tengborg et al., 1998
 41.9 6.1 1.2 14.3 NR NR 27.1 3.8 Hayn et al., 1993
Pine 46.4 8.8 2.4 11.7 NR NR 29.4 NR Wiselogel et al., 1996
 37.7 4.6 0.0 7.0 NR NR 27.5 5.4 Hayn et al., 1993
 41.7 6.3 1.8 10.8 3.9 NR 26.9 NR Youngblood et al., 2010
Western Hemlock 41.4 3.3 1.0 12.0 1.8 NR 31.4 0.8 Johansson, 2010
Crop residues Wheat straw 38.2 21.2 2.5 0.3 0.7 NR 23.4 13.0 Wiselogel et al., 1996
 35.2 30.5 4.4 0.0 0.0 NR 18.5 NR Foyle et al., 2007
 36.5 18.4 2.2 0.0 NR NR 17.6 3.6 Hayn et al., 1993
Rice straw 34.2 24.5 NR NR NR NR 11.9 17.9 Wiselogel et al., 1996
 38.9 20.4 3.4 0.0 0.5 NR 13.5 5.3 Kadam et al., 2000
Corn stover 35.6 18.9 2.9 0.3 NR NR 12.3 5.5 Hayn et al., 1993
 38.9 23.0 3.4 0.4 1.8 2.6 16.2 NR Templeton et al., 2009
 36.4 18.0 3.0 0.6 1.0 NR 16.6 7.3 Wiselogel et al., 1996
Corn cobs 37.0 27.8 2.2 NR 0.6 NR 13.9 NR Wang et al., 2011
Sugarcane bagasse 39.0 22.1 2.1 0.4 0.5 NR 23.1 NR DOE, USA
Barley straw 38.1 18.7 3. 9 0.0 0.0 NR 20. 5 NR Kim et al., 2011
Cotton stalk 35.6 21.4 0.0 0.0 0.0 NR 27.8 NR Akpinar et al., 2007
Sweet sorghum bagasse 41.3 18.0 1.94 0.85 1.26 NR 16.5 NR Goshadrou et al., 2011
Dedicated crops Switch grass 31.0 20.4 2.8 0.3 0.9 NR 17.6 17.0 Wiselogel et al., 1996
 34.2 22.8 3.1 0.3 1.4 NR 19.1 NR DOE, USA
Miscanthus 39.5c 19.0c 1.8c NR 0.4 NR 24.1 4.2 Vrije et al., 2002
Arundo donax L. 39.3 18.4 1.2 0.2 0.4 NR 26.2 NR Bura et al., 2012
Cassava pulp 19.1 4.2 1.4 0.7 0.5 NR 2.2 NR Kosugi et al., 2009
Bamboo 42.6 15.0 0.0 0.0 0.0 NR 26.2 NR Sathitsuksanoh et al., 2010
 40.7 23.6 1.1 0.6 1.2 NR 27.1 NR Tippayawong and Chanhom, 2011
Hemp (Cannabis sativa L) 37.4 21.1 2.9 NR NR 2.9 18.0 NR González-García et al., 2012a
Sorghum fiber 28.7 15.8 2.03 0.4 0.4 NR NR NR Godin et al., 2011
Secondary and tertiary Newspaper 35.1c 5.0c 3.9c 10.7c 2.3c NR 39.1c NR Foyle et al., 2007
White office paper 65.4c 14.4c 0.0 0.0 0.0 NR 9.5c NR Foyle et al., 2007

Image

NR: Not reported.
DOE: Department of Energy, USA (http://www.afdc.energy.gov/biomass/progs/search1.cgi).

aThe values denote acid insoluble or klason lignin content.

bThe values denote organic solvent extractives (see corresponding reference for more specific details).

cThe values denote the monomeric form of the respective carbohydrate.

9.3.1 Pretreatment

Pretreatment is necessary to overcome the recalcitrant nature of native lignocellulose. In contrast to the polysaccharides starch and glycogen, the function of cellulose is not to serve as an energy store, but rather to serve as a construction material. Protection of the cellulose by lignin and hemicellulose reinforces that structure, and the degradation rate in nature is therefore lower than desirable for technical application (but well suited for a functioning eco-system). A pretreatment of the material makes the fibers more accessible to enzymatic attacks. The treatment aims to open up the structure of fibers, solubilize parts of the material (i.e., lignin or hemicelluloses), reduce particle size and degree of polymerization (DP), and increase the surface area of the material.

The pretreatment in many ways determines the bioethanol process, since the properties of the material after pretreatment will have an impact on all the subsequent steps in the production. In general, it is important to design the pretreatment so that not too much degradation products are formed, since these will impact primarily the fermentation, but also the hydrolysis (Almeida et al., 2007).

Pretreatment methods can be divided into physical or chemical methods, although a combination of the two is often used. One can say that there are two principal strategies of the pretreatment; either you aim to remove mainly the lignin or you aim to remove mainly the hemicellulose fraction from the material. No method will completely succeed with any of these two aims, but hemicellulose is mostly removed with dilute acid catalyzed steam pretreatment, and lignin is to a larger extent removed by alkaline pretreatment, and oxidation improves the efficiency of lignin removal (Alvira et al., 2010; Galbe and Zacchi, 2007). Two common pretreatment abbreviations used are STEX for steam explosion and AFEX for ammonia fiber explosion. A compilation of some of the most common technologies is given in Table 9.3.

Table 9.3

Summary of different methods for lignocellulose biomass pretreatment

Pretreatment method Main action Advantage Disadvantage
Biological Lignin and hemicellulose degradation Low energy consumption Slow process
Milling Particle size reduction Reduces cellulose crystallinity High energy consumption
Steam explosion (STEX) Hemicellulose removal High glucose yields and cost effective Partial hemicelluloses removal Some inhibitor generation
Ammonia fiber explosion (AFEX) Lignin removal Increased accessible surface area
Low formation of inhibitors
Not efficient for lignin rich materials
Large amounts of ammonia
Wet oxidation Lignin removal Minimizes energy demand (exotermic) Low formation of inhibitors High cost of oxygen and alkaline catalyst
Organosolv Lignin and hemicelluloses hydrolysis Targets both hemicellulose and lignin High cost Solvents need to be drained and recycled
Dilute acid Hemicellulose removal Less corrosion problems and less formation of inhibitors compared to concentrated acid More inhibitor generation than STEX
Concentrated acida Hemicellulose and cellulose degradation High glucose yield
Ambient temperatures
Formation of inhibitors
High cost of acid and acid recirculation
   Reactor corrosion

Image

aCan also be considered as a method for complete hydrolysis to monomeric sugars, as opposed to a pretreatment step.

9.3.2 Hydrolysis

The pretreatment will leave most of the cellulose in polymeric form, and, depending on the pretreatment, some hemicelluloses in polymeric or oligomeric form. Enzymatic hydrolysis of cellulose (and hemicelluloses) into monomeric sugars is thus needed since most microorganisms utilize only monomeric (or possibly dimeric) sugars during fermentation. The enzymatic hydrolysis was long regarded as the principal bottleneck in bioethanol production from lignocellulose due to the rather slow action of cellulases and the need for large amounts of expensive enzymes. However, impressive progress in the past decade has resulted from major R&D efforts by enzyme developers, such as Novozymes, Genencor/DuPont, DSM and others, to reduce both enzyme loadings and production costs. The US Department of Energy (DOE) set a target that enzyme costs should not exceed US$0.12/gallon ethanol by 2012 in their grants to some major enzyme companies. A benchmarking of enzymes from these companies was recently conducted by the National Renewable Energy Laboratory (NREL) (McMillan et al., 2011).

Traditionally, enzymatic hydrolysis has been regarded as a synergetic reaction between three major classes of enzymes, i.e. endo-1,4-β-glucanases, exo-1,4-β-glucanases and β-glucosidases (Van Dyk and Pletschke, 2012). Endo-1,4-β-glucanases randomly cleave internal bonds in the cellulose polymer which results in the formation of two new chain ends. Exo-1,4-β-glucanases (mainly cellobiohydrolases, CBH) are the most abundant component in both natural and commercial enzyme mixtures and they mainly work in a processive manner, cutting (mainly) cellobiose units from either the reducing or non-reducing ends of the glucan chain. β-Glucosidases catalyze the hydrolytic splitting of cellobiose into glucose. The enzymes are end-product inhibited by glucose and especially cellobiose. Therefore it is important to have an enzyme blend with sufficient β-glucosidase activity in order to avoid cellobiose inhibition of the cellobiohydrolases and to be able to cope with the activity loss due to the accumulated high amount of glucose.

The synergistic effects between the different cellulase components have been extensively studied throughout the years and more recently also the synergism between cellulases and other auxiliary enzymes, e.g. xylanases, xylosidases, mannanases and esterases have been studied (Van Dyk and Pletschke, 2012). In particular it has been shown that xylobiose, and larger xylo-oligomers, exhibit a very strong inhibition on the cellulases, which can be reduced/minimized by adequate supplementation of hemicellulases (Qing and Wyman, 2011). This xylo-oligomer inhibition can be quite significant when working with xylan-rich lignocellulosic substrates (e.g., agricultural residues and hardwood; Table 9.2), especially when a rather mild pretreatment method has been used, which gives solubilization of xylan into oligomeric rather than monomeric xylose.

Another complication when hydrolyzing lignocellulosic materials is the presence of large amounts of lignin in the material (if a lignin dissolving pretreatment method has not been used). Lignin could act as a physical barrier on the cellulose surface, limiting the accessibility of the enzymes, and it has been shown to unproductively bind enzymes which lower the free amount of enzymes able to hydrolytically degrade the cellulose/cellobiose (Palonen et al., 2004).

Recently a new type of enzyme and enzyme action was identified as an important component for efficient enzymatic hydrolysis of cellulosic material (Horn et al., 2012). The enzyme belongs to the GH61 family (the abbreviation GH stands for glycoside hydrolases) but the enzyme is in fact an oxidizing enzyme that oxidizes the C1/C4 carbon on the glucan chain. Hence it has a completely different way of cleaving intermolecular glucose bonds compared to the hydrolytic cellulases mentioned earlier. Unlike the cellobiohydrolases, the GH61 enzyme does not work in a processive manner, and does not appear to need a specific binding site on the polymer chains. Therefore it can attack and cut the cellulose chain at any location, creating two new free ends for the cellobiohydrolases to act on. As such, it has the potential to break crystalline cellulose structures and enhance the hydrolysis rate by creating more reactive sites for the CBH enzymes.

9.3.3 Fermentation

Once the fermentable sugars have been obtained in the hydrolysis, the next issue is the conversion of these sugars (all sugars, or only a selected stream) by fermentation. The ideal reaction for conversion of glucose (the main hexose sugar) to ethanol is:

C6H12O22C2H5OH+2CO2

image

i.e., one mole of glucose gives a maximum of two moles of ethanol, which corresponds to 0.51 g ethanol/g glucose. In practice, one should not count on more than about 90% efficiency in the fermentation, i.e. about 0.46 g/g (Öhgren et al., 2007). The other main hexose sugars are mannose and galactose (see Table 9.2). Out of these sugars, mannose is normally easily fermented, whereas the fermentation of galactose will depend on the microorganism used. The fermentation does not need to be done after the hydrolysis, but can be done concomitant with the enzymatic hydrolysis – a process option called SSF (simultaneous saccharification and fermentation; see further discussion later on). The microorganism used for the fermentation must be able to work in the process streams with as little conditioning of the stream as possible. The microorganism should therefore be:

• inhibitory tolerant (to compounds present in the hydrolysate)

• able to utilize multiple sugars (hexoses and pentoses)

• ethanol tolerant

• able to ferment at low pH to minimize contamination.

There are many microorganisms that are able to ferment sugars, including yeasts, e.g. Saccharomyces cerevisiae, Scheffersomyces stipitis,2 Kluyveromyces marxianus, Dekkera bruxellensis (Blomqvist et al., 2010), bacteria such as Zymomonas mobilis (Rogers et al., 1982), Escherichia coli (reviewed by, e.g., Jarboe et al., 2007) and filamentous fungi such as Mucor indicus, and Rhizopus oryzae (Abedinifar et al., 2009). The organisms have different pros and cons in terms of ethanol tolerance, fermentation rate, sugar utilization range, by-product formation pattern, and tolerance to inhibitors. Genetic engineering of the organism is often needed to improve properties. For instance E. coli produces only little ethanol in its native state, but can be transformed into an ethanologenic organism (Gonzalez et al., 2003).

The process design, or the entire refinery concept, will influence the choice of organism. The organism will need to be able to use all sugars if a maximum ethanol production is desired (which is not the case in all biorefinery concepts). Alternatively, if the SSF configuration is chosen, a high thermo-tolerance of the microorganism will be an advantage (apart from the above four) as the temperatures optima for enzymatic hydrolysis (typically 45–50°C) and fermentation (typically 30–35°C) differ significantly. The thermo-tolerant yeast Kluyveromyces marxianus may be of interest in these processes (Pessani et al., 2011). However, the prime choice in the current sugar and starch-based fermentation industry has been the common Baker’s yeast, Saccharomyces cerevisiae and there is good cause to believe that this will be the case also in lignocellulose-based biorefinery approaches. There is well-established process technology for large-scale production of the yeast (Østergaard et al., 2000), and it is fully accepted in the fermentation industry today and enjoys GRAS status (i.e., it is generally regarded as safe). S. cerevisiae has an excellent ethanol tolerance, some strains tolerate up to 20% (v/v) (Verduyn et al., 1990; Casey and Ingledew, 1986). Not least important is that the organism is relatively robust to inhibitors from the pretreatment step of lignocellulose (Hahn-Hägerdal et al., 1994; Olsson and Hahn-Hägerdal, 1993).

9.3.4 Pentose utilization

In biorefinery concepts aiming at a maximum ethanol yield, all sugar streams should be fermented to ethanol, which requires the conversion also of pentoses. However, wild-type S. cerevisiae is not able to ferment pentoses, and genetic engineering of the organism is needed to enable pentose conversion. The main pentose sugar in most biomass is xylose (D-xylose) (Table 9.2) and metabolic engineering of the yeast has therefore aimed at enabling xylose fermentation in S. cerevisiae (reviewed by, e.g., Almeida et al., 2011; Van Vleet and Jeffries, 2009; Matsushika et al., 2009; Hahn-Hägerdal et al., 2007). There are two main options to introduce the enzymatic pathway which will enable conversion of the pentose xylose. One option is the two-step conversion via the enzymes xylose reductase (XR) and xylitol dehydrogenase (XDH). The other option is to use the one-step isomerization catalyzed by xylose isomerase (XI). In both cases the sugar xylulose is formed, which later is phosphorylated by the native enzyme xylulokinase (XK) to xylulose-5-phosphate, and enters into the main glycolysis via the pentose phosphate pathway (PPP) (see Fig. 9.4). The mere introduction of the pathway is, however, not sufficient for obtaining efficient yeast strains and additional improvements are necessary (as discussed in, e.g., Almeida et al., 2011). In the XR/XDH pathway, a main issue has been the co-factor dependencies of the two reactions, which may result in excretion of the by-product xylitol. This co-factor problem is avoided when using the XI pathway, but a main challenge has been to find an isomerase showing sufficient activity at allowable temperature for the yeast. Successful results were not really obtained until a XI from the anaerobic fungus Piromyces was expressed in S. cerevisiae (Kuyper et al., 2003, 2005). More recently also other useful XIs, from the anaerobic bacterium Clostridium phytofermentans (Brat et al., 2009), the fungus Orpinomyces (Madhavan et al., 2009), and from Burkholderia cenocepacia (de Figueiredo Vilela et al., 2013), have been found and expressed in S. cerevisiae.

image
9.4 Xylose and arabinose pathways expressed in recombinant Saccharomyces cerevisiae. The utilization of co-factors and ATP in the central carbon metabolism is depicted schematically. Abbreviations: XR, xylose reductase; XDH, xylitol dehydrogenase; XI, xylose isomerase; XK, xylulokinase; AI, arabinose isomerase; RK, ribulokinase; RE, ribulose-5-P 4-epimerase; LAD, L-arabitol dehydrogenase; LXR, L-xylulosereductase; PPP, pentose phosphate pathway; Fru-6P, fructose 6-phosphate; Fru-1,6 bisP, fructose 1,6 bisphosphate; GA3P: glyceraldehyde-3 phosphate; DHAP, dihydroxyacetone phosphate; L-Rib, L-ribulose; L-Rib-5P, L-ribulose 5-phosphate; HAc, undissociated weak acid; Ac-, dissociated weak acid. (Reprinted with permission from J. Almeida, D. Runquist, V. Sànchez i Nogué, G. Lidén and M. F. Gorwa-Grauslund. ‘Stress-related challenges in pentose fermentation to ethanol by the yeast Saccharomyces cerevisiae’, Biotechnol. J, 6, 286–299, 2011.)

In addition to the direct genetic engineering, i.e. introduction of the ‘missing’ heterologous genes, it has been shown to be very important to evolve the strains on the xylose-rich substrate. In particular, this was crucial in obtaining the first successful XI strain (Kuyper et al., 2003). During the evolution, the relative expression levels are ‘tuned’ for optimal performance by the yeast itself through spontaneous mutation and selection mechanisms (Zhou et al., 2012; Lee et al., 2012). Some of the currently available S. cerevisiae strains designed for utilization of xylose are given in Table 9.4.

Table 9.4

Some of the xylose fermenting engineered strains of S. cerevisiae

Strain Xylose utilizationa Description Reference
TMB3400 XR/XDH/XK + RM Xyl1, Xyl2-P. stipitis
XK-S. cerevisiae
Wahlbom et al., 2003
MA-R5 XR/XDH/XK Xyl1, Xyl2-P. stipitis
XK-S. cerevisiae (ScXK)
Matsushika et al., 2009
424A (LNH-ST) XR/XDH/XK US patent application #08/148, 581; patent no. 5789210 Sedlak and Ho, 2004
XI XylA-Piromyces sp. strain E2 Kuyper et al., 2003
RWB218 XI + EE XylA-Piromyces sp. strain E2 Kuyper et al., 2005
INVSc1/pRS406XKS/ pILSUT1/pWOXYLA XI XylA-Orpinomyces
SUT1-P. stipitis
XKS-S. cerevisiae
Madhavan et al., 2009
BWY10Xyl/YEp-opt.XI-Clos-K XI + EE XylA-Clostridium phytofermentans Brat et al., 2009
BY4741-S1 XI + RM + EE XylA-Piromyces sp. Lee et al., 2012
H131-A3-ALCS XI/XK/NOPPP + EE XylA-Piromyces sp.
Xyl3-P. stipitis
Zhou et al., 2012

Image

aXR-xylose reductase; XDH, xylose dehydrogenase; XI, xylose isomerase; XK, xylose kinase; RM, random mutagenesis; EE, evolutionary engineering; NOPPP, overexpression of non-oxidative pentose phosphate pathway genes.

A second pentose found at relatively high levels in some materials is arabinose (L-arabinose). The problem of introducing arabinose utilization into yeast is similar to that for xylose utilization as both pathways end with the formation of xylulose (Fig. 9.4; Almeida et al., 2011). Also similarly there is an isomerase pathway and a reduction/oxidation pathway. Arabinose utilization on its own is less interesting, and work has therefore focused on combining utilization of L-arabinose and D-xylose (Bettiga et al., 2009; Sanchez et al., 2010). Whereas there are currently several xylose fermenting yeast strains available, this is not yet quite true for arabinose co-utilizing yeasts.

9.3.5 Inhibitor tolerance

The challenge in lignocellulose fermentation is not only to develop organisms which use more sugars, but to have them do that in a complex medium containing several inhibitors (Fig. 9.5). The pretreatment process generates several compounds, which affect the performance of the fermenting microorganism (Klinke et al., 2004; Palmqvist and Hahn-Hägerdal, 2000). These inhibitors can be divided into:

• furans – most important 2-furaldehyde (or furfural) and 5-hydroxymethyl-2-furaldehyde (or HMF);

• weak acids – acetic acid, levulinic and formic acid formed in degradation of furans); and

• phenolics – derived from the lignin fraction.

image
9.5 Challenges faced by the yeast Saccharomyces cerevisiae during the production of ethanol from lignocellulosic feedstocks. (Reprinted with permission from J. Almeida, D. Runquist, V. Sànchez i Nogué, G. Lidén and M. F. Gorwa-Grauslund. ‘Stress-related challenges in pentose fermentation to ethanol by the yeast Saccharomyces cerevisiae’, Biotechnol. J, 6, 286–299, 2011.)

The furans are formed in the degradation of monosaccharides, and levulinic and formic acid come from further degradation of the furans. Acetic acid is different in the sense that it is already present in the material itself in terms of acetyl groups on hemicellulose.

One can handle the inhibition problem in different ways (Fig. 9.6). Obviously, the best situation would be not to have any inhibitors in the medium. That would require either a process that does not generate any inhibitors, or the removal of inhibitors. Alternatively, one can screen for strains that are tolerant, or develop more tolerant strains. Also the fermentation process in itself may be important in avoiding problems with inhibition. The reason is that several of the inhibitors, notably several aldehyde compounds such as furfural and HMF, are converted by the yeast itself into less toxic alcohols (Almeida et al., 2007, 2009) and an in-situ detoxification can be obtained during the process if suitably tuned.

image
9.6 A schematic representation on how to deal with inhibition. (Reprinted with permission from J. R. M. Almeida, M. Bertilsson, M. F. Gorwa-Grauslund, S. Gorsich and G. Lidén. ‘Metabolic effects of furaldehyde and impacts on biotechnological processes’, J Appl Microbiol Biotechnol, 82, 625–638, 2009.)

By understanding this particular mechanism of inhibition, several genetic targets have been identified. Overexpression of the gene encoding for alcohol dehydrogenase (ADH6) was, for instance, found to increase the ability of S. cerevisiae to convert HMF (Petersson et al., 2006) and also increased the ethanol formation from a hydrolysate. Alriksson et al. (2010) observed that overexpression of three native genes (ATR1, FLR1, YAP1) involving multidrug resistance and stress response in S. cerevisiae resulted in enhanced resistance to coniferyl aldehyde, HMF, and spruce hydrolysate. More targets for improved resistance are continuously reported. Very promising results have also recently been reported from adding sulfite or dithionite to the medium, which gave detoxifying effects on spruce hydrolysates (Alriksson et al., 2011). The mechanism of detoxification may partly be reduction, but also sulfonation of inhibitors (Cavka et al., 2011).

Acetic acid is a different story in comparison to many other inhibitors in that it is linked to the material composition per se since the hemicellulose is acetylated. Thus acetic acid is very difficult to avoid if the hemicellulose is to be hydrolyzed. The effects of weak acids are pH dependent, since it is the undissociated form of the acids which diffuses across the plasma membrane. There it dissociates and releases a proton due to a higher intracellular pH. To maintain the intracellular pH, the cell has to export the proton which costs ATP. The pKa value of the acid will determine at what pH the undissociated form of the acid dominates. Since many carboxylic acids have pKa in the range 4.5–5, large changes in toxicity are likely to occur for pH changes in this range. Xylose utilization appears to be extra sensitive to acetic acid (Bellissimi et al., 2009; Casey et al., 2010).

9.3.6 Thermo-tolerance

Thermo-tolerant S. cerevisiae would be an advantage in the SSF process since the temperature optima for hydrolysis (45–50°C) and fermentation (30–35°C) are different. Exposure of S. cerevisiae cells to high temperatures induces expression of several heat stress–response genes, including those encoding heat shock proteins (HSPs) and enzymes involved in trehalose and glycogen metabolism (Morano et al., 1998; Singer and Lindquist, 1998). A high-temperature growth phenotype (Htg +) was categorized in thermo-tolerance of S. cerevisiae (Steinmetz et al., 2002). Results of classical genetic analysis suggested that the Htg phenotype is dominant and approximately six genes, designated HTG1 to HTG6, are responsible for conferring this phenotype. HTG6, one of the six genes was recently identified to be the gene RSP5, which encodes a ubiquitin ligase (Shahsavarani et al., 2012) and overexpression of RSP5 ubiquitin ligase improved thermo-tolerance in S. cerevisiae. Zhang et al. (2012) isolated thermo-tolerant S. cerevisiae with a high-energy pulse electron (HEPE) beam, and obtained a strain which could produce more than 80 g/L of ethanol at as high a temperature as 43°C.

9.4 Design options for biorefining processes

In designing a process, there is a choice between completing the hydrolysis before starting the fermentation – separate hydrolysis and fermentation (SHF) – or running the hydrolysis together with the fermentation – simultaneous saccharification and fermentation (SSF).

9.4.1 Separate hydrolysis and fermentation (SHF) or simultaneous saccharification and fermentation (SSF)?

The two basic choices have different pros and cons (Table 9.5). The fact that hydrolysis and fermentation are carried out in separate vessels in SHF introduces a degree of freedom to operate each process at its respective optimal conditions, however, at the cost of one more vessel. The extra capital cost may be quite substantial as indicated by Wingren et al. (2003). If the solid fraction (mainly lignin) is removed after the enzymatic hydrolysis, it is possibly to recycle the yeast in an SHF process. This is not feasible in SSF, since the yeast cannot readily be separated from the remaining lignin. However, in the removal of the lignin fraction in SHF, there may be non-negligible sugar losses, especially if a high solids loading is used in the enzymatic hydrolysis. In contrast, if the entire slurry is sent to fermentation, all monosaccharides liberated in the pretreatment can potentially be fermented. The removal of sugars as they are released in the hydrolysis in SSF is furthermore an important advantage since end-product inhibition of cellulases can be minimized. An increased overall yield for these two reasons was the main benefit stated by the original inventors of SSF (Gauss et al., 1976). The enzymes are to some extent inhibited by the ethanol produced, but this inhibition is relatively low compared to that of cellobiose and glucose (Wu and Lee, 1997). Decreased end-product inhibition of cellulases, in particular cellobiose inhibition on CBH, is, however, a major target in development of more efficient enzymes, which may tilt the balance in favor of SHF processes.

Table 9.5

Comparison of SHF and SSF

SHF SSF
Advantages Disadvantages Advantages Disadvantages
Different temperature and pH for enzymatic hydrolysis and fermentation possible Capital cost – one more reactor
Loss of sugars during solid/liquid separation
Decreased sugar end-product inhibition
In situ detoxification
Decreased capital cost
Same temperature and pH for both hydrolysis and fermentation
No yeast recirculation
Yeast recirculation possible End-product inhibition
Inhibition by other inhibitors
Improvement of xylose fermentation by increased ratio of xylose to glucose  

Image

Co-fermentation of pentoses and hexoses in SSF is sometimes dignified with the abbreviation SSCF, standing for simultaneous saccharification and co-fermentation. SSCF may offer an advantage over SHF in these cases since the ratio between xylose and glucose can be higher than in SHF (for a more thorough discussion, see Olofsson et al., 2008). This is true for the cases in which the pretreatment results in the release of monosaccharides from hemicellulose, e.g. in acid catalyzed steam pretreatment, but not in cases where the pretreatment does not release monosaccharides.

Since the yeast cannot be reused in the SSF process, it has to be produced for each batch. Efficient aerobic cultivation of S. cerevisiae has to be performed in a fed-batch mode to avoid glucose repression which would give a low biomass yield. The fed-batch cultivation of the yeast should use lignocellulose hydrolysate (derived from the pretreatment stage) to adapt the yeast since this improves the performance significantly (Alkasrawi et al., 2006). The cultivation must be designed not only to avoid glucose repression but also inhibition by the medium. RQ (respiratory quotient), DOT (dissolved oxygen tension), or ethanol concentration can be used as measured variables for feedback control during the propagation of the yeast (as discussed by Rudolf, 2007).

The choice between SHF and SSF will depend on both feedstock and the desired product spectrum in the biorefinery. There are many different variants of both these concepts (Fig. 9.7), in particular with respect to feeding of substrates and enzymes. Feeding will be a means to minimize effects of inhibition, increase xylose conversion, or handle high viscosities (Olofsson et al., 2008).

image
9.7 A schematic representation of different fermentation strategies for production of bioethanol using lignocelluloses.

9.5 Process intensification: increasing the dry-matter content

A key aspect to attain a good economy in the bioethanol process is to reach a sufficiently high ethanol concentration after the fermentation step, since this reduces the cost (specific energy demand) of the subsequent distillation in a non-linear fashion (Galbe et al., 2007). The cost curve gradually flattens out and a concentration above 5 wt% is relatively satisfactory. High ethanol titers will require that the process operate at high biomass loadings to give high concentrations throughout the production. This also decreases the use of process water and capacity needed for water treatment. Increasing the biomass concentration in the hydrolysis and fermentation step is, however, not without problems since enzymatic conversion yields tend to decrease dramatically at elevated biomass loadings even at the same specific enzyme loading (Kristensen et al., 2009; Mohagheghi et al., 1992).There will therefore be an economic optimum at some intermediate biomass loading, which will vary with the type of biomass used, the price of enzymes, and the kind of by/co-products obtained.

9.5.1 Hybrid processes and novel concepts

The desire to increase the biomass content in the process has led to different ‘hybrid’ process configurations which employ a combination of the SHF and SSF approach – aiming at getting the best of both worlds. One such option is the inclusion of a so-called viscosity reduction step, which in essence is an enzymatic hydrolysis step operated at optimal hydrolysis temperature for a short time, but not to complete hydrolysis. A high initial hydrolysis rate can be achieved, and the process can continue as an SSF process, much like in corn-based ethanol production. One can also combine this with various fed-batch strategies in which material, enzymes and/or yeast are fed to the fermentation vessel during the process. Another option is to apply a changing temperature during the process. This is of particular interest in the SSF process, where a non-isothermal operation may give overall process improvements (Kang et al., 2012; Mutturi and Lidén, 2013). The relative rates of enzymatic hydrolysis, glucose consumption, yeast growth and decay are affected during such non-isothermal operation. Process benefits will depend on the recalcitrance of the pretreated material, and optimal and permissible temperature ranges of enzymes and yeast.

Another process concept is the so-called consolidated bioprocessing (CBP). In a CBP process, four steps (i.e., production of hydrolytic enzymes (cellulases and hemicellulases), hydrolysis of polymeric carbohydrates in the pretreated substrate, fermentation of hexoses, and fermentation of pentoses) are carried out in a single reactor (Lynd et al., 2005). One can either use a wild-type organism, such as Clostridium phytofermentans (Jin et al., 2011) able to both produce cellulases and ferment, or use genetic engineering to introduce genes encoding one or several cellulases into a fermenting microorganism, or alternatively engineer a good cellulase producing organism into an ethanol producer. The main approach has been to introduce cellulase encoding genes into a fermenting organism, e.g. S. cerevisiae. CBP is slowly gaining recognition as a potential low cost biomass processing methodology (van Zyl et al., 2007). Also here one can foresee various hybrid concepts, in which perhaps a ‘base mixture’ of enzymes is added, complemented by some specific enzyme activities that are expressed by tailored yeasts.

9.6 Different types of ethanol biorefineries

The ethanol biorefinery can be set up with focus on different combinations of co-products. To some extent there is flexibility in the operation, but the basic layout will set limits to this flexibility. With ethanol as the main product from the cellulosic part of the biomass, we can classify the biorefinery as either energy-, lignin- or C5-driven with respect to how the parts of the biomass not giving ethanol are treated (see Figs 9.89.10). In a recent study by Ekman et al. (2013), three different energy product scenarios were compared for a tentative straw-based ethanol biorefinery in Sweden. Ethanol was produced from either only the C6 fraction (C5 was taken for biogas production) or the C6 + C5 fractions. Surplus lignin (not needed for process heat) was used for electricity production. For the overall process economy it was essential to include a heat sink, i.e. a district heating system, for the low grade heat produced. The overall energy yield – counting all ‘energy’ products (ethanol, biogas, electricity and district heating) – could reach almost 70% for the best case which was when ethanol was produced from C6 sugars only. However, the economic optimization favored an ethanol production from both C6 and C5 fractions. This was in fact favored even in the absence of revenues for sale of low value heat.

image
9.8 Conceptual figure for an ‘energy-driven’ bioethanol-based biorefinery. The focus is to produce several different energy carriers to be able to make as much use as possible of the total energy content in the raw material.
image
9.9 Conceptual figure for a ‘lignin-driven’ bioethanol-based biorefinery. The focus is to separate lignin early on in the process and use all or some of it to produce high value co/by-products in order to increase the economical gain of the process.
image
9.10 Conceptual figure for a ‘C5-driven’ bioethanol-based biorefinery. The focus is to solubilize and separate the hemicelluloses early on in the process in order to have a separate C5-fermentation to high value co/by-products in order to increase the economical gain of the process.

The energy biorefinery is the concept which dominates the lignocellulose ethanol plants currently in operation, in demonstration scale or in construction phase (Table 9.6). The plant operated by Borregaard in Sarpsborg, Norway, can be said, however, to be a lignin biorefinery, with its major production of lignosulfonates, vanillin and also cellulose (Rødsrud et al., 2012).

Table 9.6

Overview of operational demonstration plants and commercial plants currently under construction

Company Location Process Feedstock Capacity
Operational demonstration plants
Borregaard Industries Ltda Sarpsborg, Norway Biochemical Spruce pulp 15,800 t/a
Abengoa Bioenergy Salamanca, Spain Biochemical Barley and wheat straw 4,000 t/a
Inbicon (DONG Energy) Kalundborg, Denmark Biochemical Wheat straw 4,300 t/a
SEKAB/EPAP Örnsköldsvik, Sweden Biochemical Primary wood chips 160 t/a
DuPont Tennessee, USA Biochemical Corn stover, cobs and fibre; switchgrass 750 t/a
Mascoma Corporation Rome, NY, USA Biochemical Wood chips, switchgrass 500 t/a
Iogen Corporation Saskatoon, Canada Biochemical Wheat, barley and oat straw; corn stover, sugar cane bagasse 1,600 t/a
Coskata Pennsylvania, USA Syngas fermentation Wood chips, natural gas 120 t/a
Verenium Jennings, LA, USA Biochemical Sugarcane bagasse, dedicated energy crops, wood products 4,200 t/a
Sued-Chemie AG Straubing, Germany Biochemical Wheat straw 1,000 t/a
Commercial plants under construction
Beta Renewables (Chemtex)b Crescentino, Italy Biochemical Arundo Donax, wheat straw 60,000 t/a
Abengoa Bioenergy Kansas, USA Biochemical Corn stover, wheat straw, switch grass 75,000 t/a
POET-DSM Advanced Biofuels Iowa, USA Biochemical Agricultural residues 75,000 t/a
INEOS Bio Florida, USA Syngas fermentation Waste (vegetative, wood and garden) 24,000 t/a

Image

Data taken from IEA Task 39 database (http://demoplants.bioenergy2020.eu/, 2012-11-20) and cross-checked with the Advanced Bio-fuels & Biobased Materials Project Database (Released Q3 2012 (07/27/12 – revision 1.1)).

aThe Borregard plant is actually a full-scale biorefinery with ethanol as one of their smaller co/by-products.

bThe Beta Renewables plant has finished construction and is under start-up during summer 2013.

*The capacities are those stated by the respective companies for the IEA database.

The potential of (co-)production of other chemicals from (part of) the sugar streams is large and many tentatively interesting compounds have been suggested. As pointed out by Bozell and Petersen (2010), this is a diverging problem since the choices of products are so many. The US Department of Energy, in a highly cited study, produced a list of a number of particularly interesting bio-based platform chemicals (Werpy and Petersen, 2004). These were selected based on criteria such as available biomass precursors (carbohydrates, lignin, fats, and proteins), process platforms, useful building blocks for further processing, secondary chemicals obtainable, intermediates, final products and applications. The 12 most interesting sugar-based building blocks were 1,4-diacids (succinic, fumaric, and malic), furan-2,5-dicarboxylic acid, 3-hydroxypropionic acid, aspartic acid, glucaric acid, glutamic acid, itaconic acid, levulinic acid, 3-hydroxybutyrolactone, glycerol, sorbitol, and xylitol/arabitol. A spectrum of high-value and commodity chemicals can be obtained from these platform chemicals as shown in Table 9.7. The original list was revised by Bozell and Petersen (2010) and some compounds, such as glutamic acid and glucaric acid, were taken off the list of most interesting compounds. However, the majority of the compounds originally identified stayed on the list and many are targets for current development projects. A comprehensive overview of recent developments on the fermentation routes and genetic engineering strategies toward many platform chemicals can be found in Jang et al. (2012).

Table 9.7

Commodity chemicals potentially derived from lignocellulose feedstocks

Source (carbon) Platform compound Biochemicals
Cellulose and hemicellulose (C3) Glycerol Fermentation products, propylene glycol, malonic, 1,3-PDO, diacids, propyl alcohol, dialdehyde, epoxides
Lactic acid Acrylates, L-propylene glycol, dioxanes, polyesters, lactide
3-Hydroxypropionate Acrylates, acrylamides, esters, 1,3-propanediol, malonic acid
Propionic acid Reagent, propionol, acrylate
Malonic acid Pharma, intermediates
Serine 2-amino-1,3-PDO, 2-aminomalonic acid
Cellulose and hemicellulose (C4) Succinic acid THF, 1,4-butanediol, γ-butyrolactone, pyrrolidones, esters, diamines, 4,4-bionelle, hydroxybutyric acid
Fumaric acid Unsaturated succinate derivatives
Malic acid Hydroxy succinate derivatives
Aspartic acid Amino scuccinate derivatives
3-Hydroxybutyrolactone Hydroxybutyrates, epoxy-γ-butyrolactone, butenoic acid
Acetoin Butanediols, butenols
Threonine Diols, ketone derivatives
Cellulose and hemicellulose (C5) Itaconic acid Methyl succinate derivatives, unsaturated esters
Furfural Many furan derivatives
Levulinic acid δ-aminolevulinate, 2-methyl
 THF, 1,4-diols, esters, succinate
Glutamic acid Amino diols, glutaric acid, substituted pyrrolidones
Xylonic acid Lactones, esters
Xylitol/Arabitol EG, PG, glycerol, lactate, hydroxyl furans, sugar acids
Citric/Aconitic acid 1,5-Pentanediol, itaconic derivatives, pyrrolidones, esters
5-hydroxymethylfurfural Numerous furan derivatives, succinate, esters, levulinic acid
Lysine Caprolactam, diamino alcohols, 1,5-diaminopentane
Gluconic acid Gluconolactones, esters
Glucaric acid Dilactones, monolactones, other products
Sorbitol Glycols (EG, PG), glycerol, lactate, isosorbide
Lignin (C6) Syngas products Methanol/dimethyl ether, ethanol, mixed liquid fuels
Hydrocarbons Cyclohexanes, higher alkylates
Phenols Cresols, eugenol, coniferols, syringols
Oxidized products Vanillin, vanillic acid, DMSO, aldehydes, quinones,aromatic and aliphatic acids
Macromolecules Carbon fibers, activated carbon, polymer alloys, polyelectrolites, substituted lignins, wood preservatives, neutraceuticals/drugs, adhesives and resins

Image

The 12 platform chemicals are represented in italics. The table was partially

adapted from Menon and Rao (2012).

Zhang et al. (2011) specifically looked at products from the lignin and hemicellulose streams. The main products from lignin today (when not used as a fuel) are lignosulfonates, which are used as adhesives, binders or plasticizers in, for example, concrete. However, lignin is also a potential source for various aromatic compounds. A few specific products produced from lignin, like vanillin, are already on the market. The challenge is to find economic methods for depolymerizing the lignin, and subsequently separating the various monomeric compounds. The C5-stream contains predominantly xylose. Products to be made in the short term from xylose, proposed by Zhang et al. (2011), include the sweetener xylitol, obtainable by reduction of xylose, and furfural which is obtained through (acid catalyzed) dehydration of xylose. Fermentation to lactic acid is another option of interest in the short term.

In the context of chemical production, it deserves to be repeated that ethanol in itself can be used not only as a fuel but also as a platform chemical. About 75% of organic chemicals were produced from only a few base chemicals; ethylene, propylene, benzene, toluene and xylene in 1987 (Coombs, 1987). Of these, ethylene can be obtained by dehydration of ethanol, but also acetaldehyde and acetic acid are relatively easily obtained from ethanol. Furthermore, it has been reported that production of chemicals, such as acetic acid, from bioethanol can actually lead to higher CO2 savings than by using it as a fuel (Rass-Hansen et al., 2007).

9.7 Future trends

9.7.1 Industrialization and process development

The development of lignocellulose-based ethanol biorefineries has gone through several ups and downs in the past decades. Steadily, however, research has come closer to industrial realization and there are today at least 10 demonstration-scale facilities in Europe and the US alone (see Table 9.6). A very significant step forward is the current completion of the first commercial (or semi-commercial) scale ethanol biorefinery plants. The plant in Crescentino, Italy, built by Beta Renewables leads the way, closely followed by plants built by Abengoa and POET in the US. The completion of commercial plants shows that the technology has reached a sufficient level of maturity for implementation. This, however, does not mean that there are no development needs, but rather that these now shift into more directly process-related issues and optimization. Some of the main issues worked on are summarized in Table 9.8. The number of biofuel-related patent filings is rising. Some recent innovations (Table 9.9) give a flavor of the current development efforts on process efficiency, pretreatment improvement, novel enzymes and fermenting strains as well as the ‘biorefinery’ issues on valorization of by-products.

Table 9.8

List of critical process development issues at various stages of biorefinery operation

 Process development issues
Biomass handling Harvest–transportation–storage
Pretreatment Should be mild yet efficient so as to achieve higher sugar release along with lower degradation products which inhibit fermentation
Enzymatic hydrolysis Cheap and efficient enzymes
Thermostable stable enzymes in case of SHF process
Discovery of novel enzymes
Fermentation Strains capable of converting all sugars released from lignocellulosics in the presence of inhibitors
Solids handling Higher solids loadings throughout the process so as to minimize the water usage
Co-products Development of viable technologies to produce co-products and by-products thereby reducing the residual waste
Energy Strategic regulation of power requirements during the operation of process plants (steam generation and recycle)

Table 9.9

Recent patent applications related to the biorefinery concept

Patent Number Description Assignee Inventor Publication date
WO 2012126099 A solvent extraction of the lignaceous residue of a biorefining process. Lignol Innovations Ltd, (British Columbia, CA) Robert SC, Yurevich BM, Ewellyn C 27/09/2012
WO 2012125925 A method for hydrolyzing biomass and viscosity reduction of biomass mixture using a composition comprising a polypeptide having glycosyl hydrolase family 61/endoglucanase. Danisco US Inc, (Palo Alto, CA, USA) Colin M, Mian LI, Bradley KR, Suzanne LE 20/09/2012
WO 2012085860 Delivery of steam produced by an electric power generation plant to a lignocellulosic biomass refinery. Inbicon A/S (Skærbæk Denmark), Boye LH, Henning A 28/06/2012
US 20120165582 A method for petrochemical products from direct liquefaction of the dried biomass. Nexxoil AG (Zurich, CH) Willner T 28/06/2012
WO 2012088429 A method for treating biomass by allowing gaseous ammonia to condense on biomass and react with water present in the biomass thereby increasing the reactivity of polysaccharides in the biomass. Board of Trustees of Michigan State University (MI, USA) Balan V, Bruce DE, Shishir C, Leonardo S 28/06/2012
WO 2012059105 A method for producing fermentation product such as ethanol using novel anaerobic, extreme thermophilic, ethanol high-yielding bacterium‘DTU01’ Technical University of Denmark, (Lyngby, DK) Irini A, Ana FT, Borisov KD 10/05/2012
US 20120071591 Method to manufacture plastic materials comprising lignin and polybutylene succinate from renewable sources. NA Mohanty AK, Misra M, Sahoo S 22/04/2012
WO 2012049054 A method for producing bioethanol, by pretreating the lignocellulosic vegetable raw materials in order to separate the cellulose Compagnie Industrielle (CIMV), (Rue Danton, France) Michel D, Bouchra BM 19/04/2012
US 20120052543 A method for pretreating biomass using combination of physical and chemical methods and thereby removing the detoxification and acid reconcentration steps. Keimyung University (Daegu, KR) Yoon KP 01/03/2012
US 20120023000 A method for accounting carbon flows and determining a regulatory value for a biofuel. NA Rhodes JS 26/01/2012
WO 2012012306 A methodology for reducing cost of enzymes in biorefinery. NA Andrew D, Michael E, Vince Y 26/01/2012
US 007973199 A method for producing acetone from hydrated ethanol derived from biomass using Zr-Fe catalyst. Metawater Co., Ltd (Tokyo, Japan) Masuda T, Tago T,Yanase T, Tsuboi, H 05/07/2011

Image

Source: www.freepatentsonline.com.

9.7.2 A future transition of paper industries?

One may argue that paper industries are in fact full-scale biorefineries. This is true in a sense, but the product focus has been very strong on pulp and paper only. The development of integrated forest biorefineries with also other products may open up for a new kind of business (Marinova et al., 2009). The major advantage of such integrated biorefineries would be to use the existing infrastructure for transport and material handling. Some options for the transition of a traditional pulp and paper industry into an integrated forest biorefinery are:

• extraction of hemicellulose prior to pulping and to be used for conversion into ethanol, organic acids, furfural, xylitol, polymers, or chemical intermediates

• recovery of lignin from black liquor stream for production of lignin-based commodity chemicals

• usage of lignin and wood residues in gasification processes for combined heat and power, chemicals, and fuel-production.

The black liquor in traditional pulp mills is used in recovery boilers for steam generation. However, alternative technologies are being developed for black liquor gasification, wherein the lignin present in the black liquor is converted to synthesis gas (Pettersson and Harvey, 2012). The synthesis gas (or syngas) is in turn used for production of dimethyl ether, methanol or other fuel compounds. A schematic representation of an integrated forest biorefinery based on a Kraft mill is shown in Fig. 9.11

image
9.11 An integrated forest biorefinery based on a Kraft pulp mill (adapted from Marinova et al., 2009)

9.7.3 Lessons from LCA studies

The future success of bioethanol refineries will be determined not only by the technical and economic performance of the plants, but also the environmental performance of the plants. Lowering of net greenhouse gas emissions (GHG) is, as mentioned in the introduction, particularly critical. Life cycle assessment (LCA) is necessary to monitor the above mentioned benefits when compared to that of fossil fuels, and LCA is now an accepted tool for guiding decision-making towards sustainability. LCA involves a holistic cradle-to-grave assessment approach to evaluate environmental performance by considering the potential impacts from all stages of manufacture, product use (including maintenance and recycling), and end-of-life management (von Blottnitz and Curran, 2007). Moreover, the International Organization for Standardization (ISO), a worldwide federation of national standards bodies, has standardized this framework within the ISO 14040 series on LCA.

The results from recent studies on LCA of biofuel production plants have established that global warming emissions and fossil energy consumption can be reduced significantly when conventional diesel and gasoline are substituted with biofuels (Punter et al., 2004; Kim and Dale, 2005; von Blottnitz and Curran, 2007; Liska et al., 2009; Cherubini and Jungmeier, 2010). However, in terms of other environmental impacts related to, for example, land usage, eutrophication of surface water, acidification and water pollution by pesticides, negative effects can result and must be carefully considered (Cherubini and Jungmeier, 2010). The environmental impact of a biorefinery depends on factors such as biomass feedstocks used, conversion routes, process configurations, and end-use applications. The LCA analysis can be used to identify further steps to be taken to, for example, reduce GHG emissions and improve overall energy efficiency of biorefineries. Examples include use of thermo-compressors for steam condensation, waste heat recovery and reuse of heat, and combustion of volatile organics using thermal oxidizers (Liska et al., 2009).

9.7.4 Future crops

The availability of feedstocks is a strategic question, and development on future dedicated feedstocks for biofuel production has also taken a prominent place in the agenda of bioenergy research. It is not possible to point out one single new feedstock which should be used for bioethanol production, as local factors related to regio-specific agricultural practices, market forces, political directives, social and biological issues will define the suitability (Sticklen, 2008). However, factors to consider include:

• identification of indigenous woody and grass species best suited to local conditions

• yields of such lignocellulose sources and scope for improvement

• production costs in terms of fertilizer input and comparison to traditional crops

• applicability of genetic engineering tools for drought tolerance and higher biomass productivity

• farm-to-refinery logistics in terms of storage and handling

• feedstock value from by-products

• environmental sustainability.

Technologies for the conversion of lignocellulosic feedstocks, such as agriculture residues from corn, rice, sugarcane, perennial grasses (switchgrass and giant miscanthus) or short rotation woody crops (fast-growing poplar and shrub willow), which contain higher cellulose fractions, should be given more prominence (Ravindranath et al., 2009; Sticklen, 2008). Newer and emerging tools in plant genetic engineering offer great potential in reducing the processing costs of these feedstocks to bioethanol. Techniques such as production of cellulases and hemicellulases within the crop biomass, genetic manipulation to modify lignin content, upregulation of cellulose and hemicellulose biosynthesis, and expression of drought tolerant traits are some of the potential research areas for improving biomass feedstocks for bioethanol production (Sticklen, 2008).

9.8 Conclusion

The worldwide growth in ethanol production has been spectacular in the past three decades, but so far has been based solely on starch or sugar. Continued expansion will have to be based on a broader raw material base, such as lignocellulose, and for both economic and environmental reasons a biorefinery approach will have to be used. The conceptual idea of a biorefinery is not new, but dates back to ‘pre-petroleum’ times. Today, however, the concept of biorefineries is about to be tested by industrial implementation, bringing together actors in agriculture, biotechnology and chemical process technology in a novel close collaboration. The technical and economic performance of the industrial pioneering efforts currently made will have a huge impact on the pace of development of the bioeconomy in the coming decade.

9.9 Sources of further information and advice

There is a wealth of information available on both biorefineries and bioethanol. The review paper by Kamm and Kamm (2007) is a good introduction to the concept of biorefineries, whereas the extensive report on ‘Top value added chemicals from biomass’ (Werpy and Petersen, 2004) gives a flavor of the chemical potential of biomass. The development of the bioethanol industry is well covered in several papers by Goldemberg (e.g., Goldemberg, 2006) which describe the development of the Brazilian ethanol industry.

Several organizations provide useful information on the internet. The International Energy Agency (IEA) has a working task force 42 entitled ‘Biorefineries: Co-production of Fuels, Chemicals, Power and Materials from Biomass’ (http://www.ieabioenergy.com). Useful information is also supplied by the US Department of Energy, in particular under its biomass programs (http://www.eere.energy.gov/biomass/). Links to maps of on-going biorefinery projects can be reached from this site. The renewable fuels association (RFA) (http://www.ethanolrfa.org/) is another organization which provides a wide range of information on ethanol.

9.10 References

1. Abedinifar S, Karimi K, Khanahmadi M, Taherzadeh MJ. Ethanol production by Mucor indicus and Rhizopus oryzae from rice straw by separate hydrolysis and fermentation. Biomass Bioenerg. 2009;33:828–833.

2. Akpinar O, Ak O, Kavas A, Bakir U, Yilmaz L. Enzymatic production of xylooligosaccharides from cotton stalks. J Agric Food Chem. 2007;55:5544–5551.

3. Alkasrawi M, Rudolf A, Lidén G, Zacchi G. Influence of strain and cultivation procedure on the performance of simultaneous saccharification and fermentation of steam pretreated spruce. Enzyme Microb Technol. 2006;38:279–287.

4. Almeida JRM, Modig T, Petersson A, Hähn-Hägerdal B, Lidén G, Gorwa-Grauslund MF. Increased tolerance and conversion of inhibitors in lignocellulosic hydrolysates by Saccharomyces cerevisiae. J Chem Technol Biotechnol. 2007;82:340–349.

5. Almeida JRM, Bertilsson M, Gorwa-Grauslund MF, Gorsich S, Lidén G. Metabolic effects of furaldehyde and impacts on biotechnological processes. Appl Microbiol Biotechnol. 2009;82:625–638.

6. Almeida JR, Runquist D, Sanchez i Nogué V, Lidén G, Gorwa-Grauslund MF. Stress-related challenges in pentose fermentation to ethanol by the yeast Saccharomyces cerevisiae. Biotechnol J. 2011;6:286–299.

7. Alriksson B, Horváth IS, Jönsson LJ. Overexpression of Saccharomyces cerevisiae transcription factor and multidrug resistance genes conveys enhanced resistance to lignocellulose derived fermentation inhibitors. Process Biochem. 2010;45:264–271.

8. Alriksson B, Cavka A, Jönsson LJ. Improving the fermentability of enzymatic hydrolysates of lignocellulose through chemical in-situ detoxification with reducing agents. Bioresour Technol. 2011;102:1254–1263.

9. Alvira P, Tomás-Pejó E, Ballesteros M, Negro MJ. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis: a review. Bioresour Technol. 2010;101:4851–4861.

10. Amichev BY, Johnston M, van Rees KCJ. Hybrid poplar growth in bioenergy production systems: biomass prediction with a simple process-based model (3PG). Biomass Bioenerg. 2010;34:687–702.

11. Angelini LG, Ceccarini L, Nassi N, Bonari E. Comparison of Arundo donax L and Miscanthus x giganteus in a long-term field experiment in Central Italy: analysis of productive characteristics and energy balance. Biomass Bioenerg. 2009;33:635–643.

12. Bellissimi E, van Dijken JP, Pronk JT, van Maris AJA. Effects of acetic acid on the kinetics of xylose fermentation by an engineered, xylose-isomerase-based Saccharomyces cerevisiae strain. FEMS Yeast Res. 2009;9:358–364.

13. Bergh J, Linder S, Bergstrom J. Potential production of Norway spruce in Sweden. Forest Ecology and Management. 2005;204:1–10.

14. Bettiga M, Gorwa-Grauslund MF, Hahn-Hägerdal B. Metabolic engineering in yeasts. In: Smolke C, ed. The Metabolic Pathway Engineering Handbook. London: Taylor and Francis/CRC Press; 2009;22.21–22.46.

15. Blomqvist J, Eberhard T, Schnürer J, Passoth V. Fermentation characteristics of Dekkera bruxellensis strains. Appl Microbiol Biotechnol. 2010;87:1487–1497.

16. Bozell JJ, Petersen GR. Technology development for the production of biobased products from biorefinery carbohydrates; the US Department of Energy’s Top 10 revisited. Green Chem. 2010;12:539–554.

17. Brat D, Boles E, Wiedemann B. Functional expression of a bacterial xylose isomerase in Saccharomyces cerevisiae. Appl Environ Microbiol. 2009;75:2304–2311.

18. Bura R, Ewanick S, Gustafson R. Assessment of Arundo donax (giant reed) as feedstock for conversion to ethanol. TAPPI Journal. 2012;11:59–66.

19. Callesen I, Grohnheit PE, Østergård H. Optimization of bioenergy yield from cultivated land in Denmark. Biomass Bioenerg. 2010;34:1348–1362.

20. Casey E, Sedlak M, Ho NWY, Mosier NS. Effect of acetic acid and pH on the cofermentation of glucose and xylose to ethanol by a genetically engineered strain of Saccharomyces cerevisiae. FEMS Yeast Res. 2010;10:385–393.

21. Casey GP, Ingledew WM. Ethanol tolerance in yeasts. Crit Rev Microbiol. 1986;13:219–280.

22. Cavka A, Alriksson B, Ahnlund M, Jönsson LJ. Effect of sulfur oxyanions on lignocellulose-derived fermentation inhibitors. Biotechnol Bioeng. 2011;108:2592–2599.

23. Cherubini F, Jungmeier G. LCA of a biorefinery concept producing bioethanol, bioenergy, and chemicals from switchgrass. Int J LCA. 2010;15:53–66.

24. Commission Directive 2009/28/EC. On the promotion of the use of energy from renewable sources and amending and subsequently repealing Directives 2001/77/EC and 2003/30/EC.2009.04.23, OJ L, 140, 16–62.

25. Coombs J. EEC resources and strategies. Phil Trans R Soc Lond Ser A. 1987;321:405–422.

26. de Figueiredo Vilela L, de Mello VM, Reis VCB, et al. Functional expression of Burkholderia cenocepacia xylose isomerase in yeast increases ethanol production from a glucose–xylose blend. Bioresour Technol. 2013;128:792–796.

27. Ekman A, Wallberg O, Joelsson E, Börjesson P. Possibilities for sustainable biorefineries based on agricultural residues – a case study of potential straw-based ethanol production in Sweden. Appl Energy. 2013;102:299–308.

28. Energy Independence and Security Act of 2007. Public Law 110–140, 2007; http://frwebgate.access.gpo.gov/cgi-bin/getdoc.cgi?dbname)110_cong_public_laws&docid)f:publ140.110.pdf.

29. Finlay MR. Old efforts at new uses: a brief history of chemurgy and the American search for biobased materials. J Ind Ecol. 2004;7:33–46.

30. Foyle T, Jennings L, Mulcahy P. Compositional analysis of lignocellulosic materials: evaluation of methods used for sugar analysis of waste paper and straw. Biores Technol. 2007;98:3026–3036.

31. Galbe M, Zacchi G. Pretreatment of lignocellulosic materials for efficient bioethanol production. Biofuels. 2007;108:41–65.

32. Galbe M, Sassner P, Wingren A, Zacchi G. Process engineering economics of bioethanol production. Adv Biochem Eng Biotechnol. 2007;108:303–327.

33. Gauss WF, Suzuki S, Takagi M. Manufacture of alcohol from cellulosic materials using plural ferments. 1976; US Patent US3990944.

34. Giebelhaus AW. Farming for fuel: the alcohol motor fuel movement of the 1930s. Agricultural History. 1980;54:173–184.

35. Godin B, Agneessens R, Gerin P, Delcarte J. Composition of structural carbohydrates in biomass: precision of a liquid chromatography method using a neutral detergent extraction and a charged aerosol detector. Talanta. 2011;85:2014–2026.

36. Goldemberg J. The ethanol program in Brazil. Environ Res Lett. 2006;1:1–5.

37. Gonzalez R, Tao H, Purvis JE, York SW, Shanmugam KT, Ingram LO. Gene array-based identification of changes that contribute to ethanol tolerance in ethanologenic Escherichia coli: comparison of KO11 (Parent) to LY01 (resistant mutant). Biotechnol Prog. 2003;19:612–623.

38. González-García S, Luo L, Moreira MT, Feijoo G, Huppes G. Life cycle assessment of hemp hurds use in second generation ethanol production. Biomass Bioenerg. 2012a;36:268–279.

39. González-García S, Iribarren D, Susmozas A, Dufour J, Murphy RJ. Life cycle assessment of two alternative bioenergy systems involving Salix spp biomass: Bioethanol production and power generation. Appl Energ. 2012b;95:111–122.

40. Goshadrou A, Karimi K, Taherzadeh MJ. Improvement of sweet sorghum bagasse hydrolysis by alkali and acidic pretreatments. In: Linköping, Sweden: World Renewable Energy Congress, Bioenergy Technology; 2011;374–380. 8–13 May.

41. Hahn-Hägerdal B, Jeppsson H, Olsson L, Mohagheghi A. An inter laboratory comparison of the performance of ethanol-producing microorganisms in a xylose-rich acid hydrolysate. Appl Microbiol Biotechnol. 1994;41:62–72.

42. Hahn-Hägerdahl B, Karhumaa K, Fonseca C, Spencer-Martins I, GorwaGrauslund MF. Towards industrial pentose fermenting yeast strains. Appl Microbiol Biotechnol. 2007;74:937–953.

43. Hayn M, Steiner W, Klinger R, Steinmüller H, Sinner M, Esterbauer H. Basic research and pilot studies on the enzymatic conversion of lignocellulosics. In: Saddler JN, ed. Bioconversion of Forest and Agricultural Plant Residues. Wallingford: CAB International; 1993;33–72.

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

45. Jae JH, Tompsett GA, Lin YC, et al. Depolymerization of lignocellulosic biomass to fuel precursors: maximizing carbon efficiency by combining hydrolysis with pyrolysis. Energy Environ Sci. 2010;3:358–365.

46. Jang Y-S, Kim B, Shin JH, et al. Bio-based production of C2–C6 platform chemicals. Biotechnol Bioeng. 2012;109:2437–2459.

47. Jarboe LR, Grabar TB, Yomano LP, Shanmugan KT, Ingram LO. Development of ethanologenic bacteria. Adv Biochem Eng Biotechnol. 2007;108:237–261.

48. Jin M, Balan V, Gunawan C, Dale BE. Consolidated bioprocessing (CBP) performance of Clostridium phytofermentans on AFEX-treated corn stover for ethanol production. Biotechnol Bioeng. 2011;108:1290–1297.

49. Johansson J. Pretreatment of prevalent Canadian west coast softwoods using the ethanol organosolv process: assessing robustness of the ethanol organosolv process Sweden: Masters thesis, Lund University; 2010.

50. Kadam KL, Forrest LH, Jacobson WA. Rice straw as a lignocellulosic resource: collection, processing, transportation, and environmental aspects. Biomass Bioenerg. 2000;18:369–389.

51. Kamm B, Kamm M. Biorefineries – multi-product processes. Adv Biochem Eng Biotechnol. 2007;105:175–204.

52. Kang H-W, Kim Y, Kim S-W, Choi G-W. Cellulosic ethanol production on temperature-shift simultaneous saccharification and fermentation using the thermostable yeast Kluyveromyces marxianus CHY1612. Bioprocess Biosyst Eng. 2012;35:115–122.

53. Kim S, Dale BE. Life cycle assessment of various cropping systems utilized for producing biofuels: bioethanol and biodiesel. Biomass Bioenerg. 2005;29:426–439.

54. Kim Y, Yu A, Han M, Choi G, Chung B. Enhanced enzymatic saccharification of barley straw pretreated by ethanosolv technology. Appl Microbiol Biotechnol. 2011;163:143–152.

55. Klinke HB, Thomsen AB, Ahring BK. Inhibition of ethanolproducing yeast and bacteria by degradation products produced during pretreatment of biomass. Appl Microbiol Biotechnol. 2004;66:10–26.

56. Köpke M, Mihalcea C, Bromley JC, Simpson SD. Fermentative production of ethanol from carbon monoxide. Curr Opin Biotechnol. 2011;22:320–325.

57. Kosugi A, Kondo A, Ueda M, et al. Production of ethanol from cassava pulp via fermentation with a surface engineered yeast strain displaying glucoamylase. Renew Energy. 2009;34:1354–1358.

58. Kristensen JB, Felby C, Jorgensen H. Yield-determining factors in high-solids enzymatic hydrolysis of lignocellulose. Biotechnol Biofuels. 2009;2:11.

59. Kundiyana DK, Bellmer DD, Huhnke RL, Wilkins MR, Claypool PL. Influence of temperature, pH and yeast on in-field production of ethanol from unsterilized sweet sorghum juice. Biomass Bioenerg. 2010;34:1481–1486.

60. Kuyper M, Harhangi HR, Stave AK, et al. High level functional expression of a fungal xylose isomerase: the key to efficient ethanolic fermentation of xylose by Saccharomyces cerevisiae. FEMS Yeast Res. 2003;4:69–78.

61. Kuyper M, Toirkens MJ, Diderich JA, Winkler AA, van Dijken JP, Pronk JT. Evolutionary engineering of mixed-sugar utilization by a xylose-fermenting Saccharomyces cerevisiae strain. FEMS Yeast Res. 2005;5:925–934.

62. Lee S-M, Jellison T, Alper HS. Directed evolution of xylose isomerase for improved xylose catabolism and fermentation in the yeast Saccharomyces cerevisiae. Appl Environ Microbiol. 2012;78:5708–5716.

63. Liska AJ, Yang HS, Bremer VR, et al. Improvements in life cycle energy efficiency and greenhouse gas emissions of corn-ethanol. J Ind Ecol. 2009;13:58–74.

64. Lynd LR, van Zyl WH, McBride JE, Laser M. Consolidated bioprocessing of cellulosic biomass: an update. Curr Opin Biotechnol. 2005;16:577–583.

65. Mabee WE, Gregg DJ, Arato C, et al. Updates on softwood-to-ethanol process development. Appl Biochem Biotechnol. 2006;129–132:55–70.

66. Madhavan A, Tamalampudi S, Ushida K, et al. Xylose isomerase from polycentric fungus Orpinomyces: gene sequencing, cloning, and expression in Saccharomyces cerevisiae for bioconversion of xylose to ethanol. Appl Microbiol Biotechnol. 2009;82:1067–1078.

67. Marinova M, Mateos-Espejel E, Jemaa N, Paris J. Addressing the increased energy demand of a Kraft mill biorefinery: the hemicellulose extraction case. Chem Eng Res Des. 2009;87:1269–1275.

68. Matsushika A, Inoue H, Kodaki T, Sawayama S. Ethanol production from xylose in engineered Saccharomyces cerevisiae strains: current state and perspectives. Appl Microbiol Biotechnol. 2009;84:37–53.

69. McMillan J, Jennings EW, Mohagheghi A, Zuccarello M. Comparative performance of precommercial cellulases hydrolyzing pretreated corn stover. Biotechnol Biofuels. 2011;4:29.

70. Menon V, Rao M. Trends in bioconversion of lignocelluloses: biofuels, platform chemicals and biorefinery concept. Prog Energ Combust. 2012;38:522–550.

71. Mohagheghi A, Tucker M, Grohmann K, Wyman C. High solids simultaneous saccharification and fermentation of pretreated wheat straw to ethanol. Appl Biochem Biotechnol. 1992;33:67–81.

72. Morano KA, Liu PC, Thiele DJ. Protein chaperones and the heat shock response in Saccharomyces cerevisiae. Curr Opin Microbiol. 1998;1:197–203.

73. Mutturi S, Lidén G. Effect of temperature on simultaneous saccharification and fermentation of pretreated spruce and arundo. Ind Eng Chem Res. 2013;52:1244–1251.

74. Öhgren K, Bura R, Lesnicki G, Saddler J, Zacchi G. A comparison between simultaneous saccharification and fermentation and separate hydrolysis and fermentation using steam-pretreated corn stover. Process Biochem. 2007;42:834–839.

75. Olofsson K, Bertilsson M, Lidén G. A short review on SSF – an interesting process option for ethanol production from lignocellulosic feedstocks. Biotechnol Biofuels. 2008;1:7.

76. Olsson L, Hahn-Hägerdal B. Fermentative performance of bacteria and yeasts in lignocellulose hydrolysate. Proc Biochem. 1993;28:249–257.

77. Østergaard S, Olsson L, Nielsen J. Metabolic engineering of Saccharomyces cerevisiae. Microbiol Mol Biol Rev. 2000;64:34–50.

78. Palmqvist E, Hahn-Hägerdal B. Fermentation of lignocellulosic hydrolysates II: Inhibitors and mechanisms of inhibition. Bioresour Technol. 2000;74:25–33.

79. Palonen H, Tjerneld F, Zacchi G, Tenkanen M. Adsorption of Trichoderma reesei CBH I and EG II and their catalytic domains on steam pretreated softwood and isolated lignin. J Biotechnol. 2004;107:65–72.

80. Pérez-Cruzado C, Merino A, Rodríguez-Soalleiro R. A management tool for estimating bioenergy production and carbon sequestration in Eucalyptus globulus and Eucalyptus nitens grown as short rotation woody crops in north-west Spain. Biomass Bioenerg. 2011;35:2839–2851.

81. Pessani NK, Atiyeh HK, Wilkins MR, Bellmer DD, Banat IM. Simultaneous saccharification and fermentation of Kanlow switchgrass by thermotolerant Kluyveromyces marxianus IMB3: the effect of enzyme loading, temperature and higher solid loadings. Bioresour Technol. 2011;102:10618–10624.

82. Petersson A, Almeida JR, Modig T, Karhumma K, Hahn-Hägerdal B, Gorwa-Grauslund MF. A 5-hydroxymethylfurfural reducing enzyme encoded by the Saccharomyces cerevisiae ADH6 gene conveys HMF tolerance. Yeast. 2006;23:455–464.

83. Pettersson K, Harvey S. Comparison of black liquor gasification with other pulping biorefinery concepts – systems analysis of economic performance and CO2 emissions. Energy. 2012;37:136–153.

84. Pham V, El-Halwagi MM. Process synthesis and optimization of biorefinery configurations. AIChE J. 2012;58:1212–1221.

85. Punter G, Rickeard D, Larivé JF, et al. Well-to-wheel evaluation for production of ethanol from wheat A report by the Low CVP fuels working group. 2004; WTW sub-group, FWGP-04–024, October.

86. Qing Q, Wyman C. Supplementation with xylanase and β-xylosidase to reduce xylo-oligomer and xylan inhibition of enzymatic hydrolysis of cellulose and pretreated corn stover. Biotechnol Biofuels. 2011;4:18.

87. Rass-Hansen J, Falsig H, Jorgensen B, Christensen CH. Bioethanol: fuel or feedstock? J Chem Technol Biotechnol. 2007;82:329–333.

88. Ravindranath NH, Manuvie M, Fargione J, et al. Greenhouse gas implications of land use and land conversion to biofuel crops. In: Howarth R, Bringezu S, eds. Biofuels:Environmental Consequences and Interactions with Changing Land Use. Proceedings of the Scientific Committee on Problems of the Environment (SCOPE) International Biofuels Project Rapid Assessment, 22–25 September 2008, Gummersbach, Germany. NY: Cornell University; 2009;11–125.

89. Rencoret J, Gutiérrez A, Nieto L, et al. Lignin composition and structure in young versus adult Eucalyptus globulus plants. Plant Physiol. 2010;155:667–682.

90. Rødsrud G, Lersch M, Sjöde A. History and future of world’s most advanced biorefinery in operation. Biomass Bioenerg. 2012;46:46–59.

91. Rogers P, Lee K, Skotnicki M, Tribe D. Ethanol production by Zymomonas mobilis. In: Microbial Reactions. New York: Spinger-Verlag; 1982;37–84.

92. Rudolf A. Fermentation and cultivation technology for improved ethanol production from lignocellulose. PhD thesis Lund University 2007.

93. Sanchez RG, Karhumaa K, Fonseca C, et al. Improved xylose and arabinose utilization by industrial recombinant Saccharomyces cerevisiae strain using evolutionary engineering. Biotechnol Biofuels. 2010;3:13.

94. Sassner P, Galbe M, Zacchi G. Bioethanol production based on simultaneous saccharification and fermentation of steam-pretreated Salix at high dry-matter content. Enzyme Microb Technol. 2006;39:756–762.

95. Sathitsuksanoh N, Zhu Z, Ho T-J, Bai M-D, Zhang Y-HP. Bamboo saccharification through cellulose solvent-based biomass pretreatment followed by enzymatic hydrolysis at ultra-low cellulase loadings. Bioresour Technol. 2010;101:4926–4929.

96. Sedlak M, Ho NW. Production of ethanol from cellulosic biomass hydrolysates using genetically engineered Saccharomyces yeast capable of cofermenting glucose and xylose. Appl Biochem Biotechnol. 2004;113–116:403–416.

97. Shahsavarani H, Sugiyama M, Kaneko Y, Chuenchit B, Harashima S. Superior thermotolerance of Saccharomyces cerevisiae for efficient bioethanol fermentation can be achieved by overexpression of RSP5 ubiquitin ligase. Biotechnol Adv. 2012;30:1289–1300.

98. Singer MA, Lindquist S. Thermotolerance in Saccharomyces cerevisiae: the yin and yang of trehalose. Trends Biotechnol. 1998;16:460–468.

99. Snowdon P, Benson ML. Effects of combinations of irrigation and fertilization on the growth and above-ground biomass production of Pinus radiata. For Ecol Manag. 1992;52:87–116.

100. Songstad DD, Lakshmanan P, Chen J, Gibbons W, Hughes S, Nelson R. Historical perspective of biofuels: learning from the past to rediscover the future. In Vitro Cell Dev Biol – Plant. 2009;45:189–192.

101. Steinmetz LM, Sinha H, Richards DR, et al. Dissecting the architecture of a quantitative trait locus in yeast. Nature. 2002;416:326–330.

102. Sticklen MB. Plant genetic engineering for biofuel production: towards affordable cellulosic ethanol. Nat Rev Genet. 2008;9:433–443.

103. Templeton DW, Sluiter AD, Hayward TK, Hames BR, Thomas SR. Assessing corn stover composition and sources of variability via NIRS. Cellulose. 2009;16:621–639.

104. Tengborg C, Stenberg K, Galbe M, et al. Comparison of SO2 and H2SO4 impregnation of softwood prior to steam pretreatment on ethanol production. Appl Biochem Biotechnol. 1998;70–72:3–15.

105. Tippayawong N, Chanhom N. Conversion of bamboo to sugars by dilute acid and enzymatic hydrolysis. Int J Renewable Energy Res. 2011;1:240–244.

106. Tullus A, Tullus H, Soo T, Pärn L. Above-ground biomass characteristics of young hybrid aspen (Populus tremula L x P. tremuloides Michx.) plantations on former agricultural land in Estonia. Biomass Bioenerg. 2009;33:1617–1625.

107. van Dyk JS, Pletschke BI. A review of lignocellulose bioconversion using enzymatic hydrolysis and synergistic cooperation between enzymes – factors affecting enzymes, conversion and synergy. Biotechnol Adv. 2012;30:1458–1480.

108. Van Vleet JH, Jeffries TW. Yeast metabolic engineering for hemicellulosic ethanol production. Curr Opin Biotechnol. 2009;20:300–306.

109. van Zyl WH, Lynd LR, Den Haan R, McBride JE. Consolidated bioprocessing for bioethanol production using Saccharomyces cerevisiae. Adv Biochem Eng Biotechnol. 2007;108:205–235.

110. Vázquez G, Freire MS, Antorrena G. Valorisation of lignocellulosic waste materials: tannins as a source of new products. In: Proceedings of European Congress of Chemical Engineering (ECCE-6), Copenhagen, 16–20 September. 2007.

111. Verduyn C, Postma E, Scheffers WA, van Dijken JP. Physiology of Saccharomyces cerevisiae in anaerobic glucose-limited chemostate cultures. J Gen Microbiol. 1990;136:395–403.

112. von Blottnitz H, Curran MA. A review of assessments conducted on bio-ethanol as a transportation fuel from a net energy, greenhouse gas, and environmental life cycle perspective. J Clean Prod. 2007;15:607–619.

113. Vrije TD, Haas GGD, Tan GB, Keijsers ERP, Claassen PAM. Pretreatment of Miscanthus for hydrogen production by Thermotoga elfii. Int J Hydrogen Energ. 2002;27:1381–1390.

114. Wahlbom CF, Cordero Otero RR, van Zyl WH, Hahn-Hägerdal B, Jonsson LJ. Molecular analysis of a Saccharomyces cerevisiae mutant with improved ability to utilize xylose shows enhanced expression of proteins involved in transport, initial xylose metabolism, and the pentose phosphate pathway. Appl Environ Microbiol. 2003;69:740–746.

115. Wang GS, Lee J-W, Zhu JY, Jeffries TW. Dilute acid pretreatment of corncob for efficient sugar production. Appl Microbiol Biotechnol. 2011;163:658–668.

116. Top value added chemicals from biomass. In: Werpy T, Petersen G, eds. Pacific Northwest National Laboratory and National Renewable Energy Laboratory 2004; Report 8674.

117. Wilkins MR, Atiyeh HK. Microbial production of ethanol from carbon monoxide. Curr Opin Biotechnol. 2011;22:326–330.

118. Wingren A, Galbe M, Zacchi G. Techno-economic evaluation of producing ethanol from softwood: comparison of SSF and SHF and identification of bottlenecks. Biotechnol Prog. 2003;19:1109–1117.

119. Wiselogel A, Tyson S, Johnson D. Biomass feedstock resources and composition. In: Wyman CE, ed. Handbook on Bioethanol: Production and Utilization. Washington, DC: Taylor & Francis; 1996;105–118.

120. Wu Z, Lee YY. Inhibition of the enzymatic hydrolysis of cellulose by ethanol. Biotechnol Lett. 1997;19:977–979.

121. Youngblood A, Zhu J, Scott CT. Ethanol production from woody biomass: silvicultural opportunities for surpressed western conifers. In: USDA Forest Service Proceedings RMRS-P-61. 2010.

122. Zhang Q, Fu Y, Wang Y, Han J, Lv J, Wang S. Improved ethanol production of a newly isolated thermotolerant Saccharomyces cerevisiae strain after high-energy-pulse electron beam. J Appl Microbiol. 2012;112:280–288.

123. Zhang X, Tu M, Paice MG. Routes to potential bioproducts from lignocellulosic biomass lignin and hemicelluloses. Bioenerg Res. 2011;4:246–257.

124. Zhou H, Cheng JS, Wang BL, Fink GR, Stephanopoulos G. Xylose isomerase overexpression along with engineering of the pentose phosphate pathway and evolutionary engineering enable rapid xylose utilization and ethanol production by Saccharomyces cerevisiae. Metab Eng. 2012;14:611–622.


1The latter part of the word is derived from ‘ergon’ meaning ‘work’ (Finlay, 2004).

2This yeast was previously called Pichia stipitis and most publications will be found under this name.

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

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