19

Bio-based chemicals from biorefining: carbohydrate conversion and utilisation

K. Wilson,    European Bioenergy Research Institute, Aston University, UK

A.F. Lee,    University of Warwick, UK and Monash University, Australia

Abstract:

The quest for sustainable sources of fuels and chemicals to meet the demands of a rapidly rising global population represents one of this century’s grand challenges. Biomass offers the most readily implemented, and low cost, solution for transportation fuels, and the only non-petroleum route to organic molecules for the manufacture of bulk, fine and speciality chemicals and polymers. Chemical processing of such biomass-derived building blocks requires catalysts compatible with hydrophilic, bulky substrates to facilitate the selective deoxygenation of highly functional bio-molecules to their target products. This chapter addresses the challenges associated with carbohydrate utilisation as a sustainable feedstock, highlighting innovations in catalyst and process design that are needed to deliver high-value chemicals from biomass-derived building blocks.

Key words

heterogeneous catalysis; lignocellulose; platform chemicals; biofuels; porous materials

19.1 Introduction

Mounting concerns over dwindling petroleum oil reserves in concert with growing governmental and public acceptance of the anthropogenic origin of rising CO2 emissions and associated climate change, are driving academic and commercial routes to utilise renewable feedstocks as sustainable sources of fuel and chemicals. The quest for such sustainable resources to meet the demands of a rapidly rising global population represents one of this century’s grand challenges.1 Biomass offers the most readily implemented, and low cost, solution for transportation fuels,2 and the only non-petroleum route to organic molecules for the manufacture of bulk, fine and speciality chemicals and polymers3 required to meet future societal demands.4,5

In order to be considered truly sustainable, biomass feedstocks must be derived from sources which do not compete with agricultural land use for food production, or compromise the environment, e.g. via deforestation.6 Potential feedstocks include cellulosic or oil-based materials derived from plant or aquatic sources, with the so-called biorefinery concept offering the co-production of fuels, chemicals and energy,7 analogous to today’s petroleum refineries which deliver high volume/low value (e.g., fuels and commodity chemicals) and low volume/high value (e.g., fine/speciality chemicals) products, maximising biomass valorisation.8 Unlike fossil fuels, which comprise predominantly unfunctionalised hydrocarbons, carbohydrates derived from sugars, starches or lignocellulose are highly functionalised, thus requiring new conversion technologies to yield useful chemicals.9,10 This chapter addresses the challenges associated with carbohydrate utilisation as a sustainable feedstock,11 highlighting recent developments in heterogeneous catalysis for the production of platform chemicals.

19.2 Sustainable carbohydrate sources

Feedstock selection for any bio-based process requires life-cycle and socio-economic assessments of attendant energy, resource and land use requirements to ensure sustainability. Polysaccharides sourced from starch (C6H10O5)n derived from wheat, corn or tuberous plants, or sugar cane12 and sorghum13, are most easily processed to glucose via hydrolysis, but such food crops are not considered sustainable. In this respect, non-food sources of grass sugars (e.g., ryegrass)14 and sugar beet pulp15 are more attractive carbohydrate sources, with the latter composed of pectin (a hetero-polysaccharide comprising galactose, galacturonic acid, arabinose and xylose units) which is abundant in the cell walls of non-woody biomass (e.g., apple pomace and citrus waste).16

Lignocellulose derived from waste agricultural or forestry materials (e.g., logging or mill and manufacturing residues), or perennial herbaceous plants and short rotation woody crops (e.g., miscanthus, eucalyptus or willow), is considered a particularly viable option for sustainable fuels and chemicals production. 17,18 The use of waste biomass residues19 from food processing, such as bagasse and rice husk and straw, offers a very attractive means to valorise waste materials that would otherwise be left to decompose, or in the case of rice straw, simply burned thereby releasing atmospheric CO2.20

Lignocellulose is a biopolymer comprising cellulose (30–50% of total lignocellulosic dry mass) and hemicellulose (20–40% of total dry mass), themselves assembled from C6 and C5 sugars such as glucose, xylose and amylose, bound together by poly-phenolic lignin which makes up the remaining 15–25% of dry biomass and imparts rigidity to plants and trees (Fig. 19.1).

image
19.1 Structure of lignin, hemicelluloses and cellulose contained within lignocellulose.

Cross-linking of the cellulose and hemicellulosic components with lignin via ester and ether linkages renders lignocellulose resistant towards hydrolysis, hampering its chemical conversion.19 Figure 19.2 illustrates popular approaches to lignocellulosic biomass utilisation for fuels and chemicals synthesis, encompassing sugar fermentation to ethanol, gasification to syngas (CO/H2), and liquefaction or pyrolysis to bio-oils.

image
19.2 Biochemical and thermochemical routes for lignocellulose conversion to chemicals and fuels.

A limitation of lignocellulose is that it cannot be used directly in biochemical processes, thus while the conversion of sugars such as glucose to chemical feedstocks via fermentation appears an attractive prospect, it requires extensive pretreatment of the raw materials.21 Lignin and C5 sugars from hemicellulose cannot be used in fermentation processes (to produce, e.g., ethanol),22 hence lignocellulosic biomass first requires purification of the cellulosic component, typically through acid or base hydrolysis, steam explosion23 or organosolv treatments24 to separate the polysaccharide from lignin components.25 Once separated, the cellulose fractions are typically hydrolysed to fermentable sugars for further processing into fuels and/or chemicals via enzymatic,26 chemical (acid or base),27,28 supercritical water29 or more recently ionic liquid (IL)-based30 treatments (Fig. 19.3).

image
19.3 Conversion of lignocellulose to cellulose via fractionation.

Since hemicellulose (C5H8O5)n is relatively amorphous, it is easier to chemically or thermally decompose than cellulose, but yields a mixture of C6 and C5 sugars. Current pretreatment steps are energy intensive, and generate significant quantities of waste during acid/base neutralisation steps. Hence there is great scope for improved technologies to enhance energy and atom economies. Furthermore, complete fractionation of lignocelluloses often sacrifices one or more of the components (e.g., lignin31 or hemicellulose,32). For example, sulfuric acid treatments during the Kraft process result in a particularly intractable Kraft lignin material.33 Likewise, decomposition of hemicelluloses during high temperature/pressure steam explosion leads to an aqueous fraction rich in furfural and C1-C2 acids34 which are problematic for subsequent fermentation since they can inhibit yeast growth.

19.3 Chemical hydrolysis of cellulose to sugars

19.3.1 Acid hydrolysis

Cellulose is composed of microcrystalline fibres, hydrogen bonded to each other by a charged water boundary layer formed from dipole–dipole interactions between aligned water molecules and the polar surface of the cellulose fibres.35 To initiate acid hydrolysis, protons must penetrate this charged water boundary layer, and thereby catalyse hydrolysis of the β-glycosidic linkages to release glucose or cellobiose (glucose dimers) from the outermost layer of the cellulose crystallites. However, the initial hydrolysis products tend to remain in close proximity to the cellulose surface due to hydrogen bonding interactions, resulting in a viscous boundary layer that slows further hydrolysis of the underlying cellulose.36 Efficient mixing is thus essential to displace reactively formed glucose from the surface, while the addition of Li+, Na+, K+, Ca2 + and NH4+ salts has also been reported to help disrupt the crystalline water matrix. While elevated reaction temperatures (via hot compressed water) can accelerate hydrolysis, they may also promote undesired side reactions if products are not continuously removed, with glucose readily undergoing subsequent reaction to by-products such as furfural and 5-HMF.37 Without continuous removal of hydrolysis products from the ‘boundary layer’at the surface of cellulose particles, conventional acid hydrolysis routes can only achieve glucose yields of around 70%, due to glucose degradation under these forcing reaction conditions.36 Furthermore, sulfuric acid-initiated cellulose hydrolysis at acid concentrations ranging from 0.4 to 2 wt%3840 poses additional problems due to the requirement for expensive corrosion-resistant reactors and co-production of vast quantities of gypsum waste formed during acid neutralisation via lime addition. There is thus an urgent need for low energy technologies for cellulose conversion to sugars and platform chemicals.

Catalytic aqueous-phase reforming (APR) of cellulose is one such approach which allows the direct conversion of carbohydrates into hydrogen and alkanes (C1 to C15) and could form a platform for fuels production within an integrated biorefinery.41 APR typically involves the reaction of cellulose under acid conditions at ∼ 200°C to yield C1 to C6 alkanes by aqueous-phase dehydration/hydrogenation (APD/H) of sugars, or C7 to C15 alkanes by combining aldol condensation with dehydration/hydrogenation (Fig. 19.4).

image
19.4 Reaction scheme for aqueous phase reforming of cellulose to alkanes. Reproduced from Ref. 41 with permission from Elsevier.

19.3.2 Heterogeneous catalysts for cellulose conversion to platform chemicals

Cellulose depolymerisation to sugars through acid hydrolysis is the first step in a biorefinery,42 hence there is much interest in developing heterogeneous catalysts to replace the conventional mineral acids (e.g., H2SO4) currently employed to achieve this.43,44 Solid acids explored to date include sulfonic acid mesoporous silicas,45 sulfonated porous carbons4648 and carbon-silicas,49 as well as sulfonic acid polystyrene resins.50 Sulfonated mesoporous carbons48 are reported to produce remarkably high glucose yields (74.5%) following the 150°C hydrolysis of amorphous cellulose obtained from pretreating cellulose in a planetary ball mill at 500 rpm for 48 h. For cellulose hydrolysis by solid acids, efficient solid–solid interfacial contact is critical,51 with the amount of water playing a key role in controlling the reaction kinetics over carbon-based catalysts. Glucose yields were optimal when the quantity of water was comparable to the weight of solid carbon catalyst; high water content is suggested to hydrate acid sites, decreasing the catalyst Brönsted acidity.

The use of magnetic sulfonated materials affords a novel and facile means to separate catalysts from aqueous media containing water-soluble sugars and solid cellulose. For example, magnetic silica nanoparticulate catalysts comprising a CoFe2O4 core with a sulfonic acid derivatised SiO2 shell52 are active for cellulose hydrolysis, while offering facile separation from the reaction media by a magnet, improving their reusability.

Cellulose is only soluble in water at temperatures in excess of > 320°C. However, such temperatures, required to interrupt the hydrogen-bonding between the fibres,53 result in low glucose yields due to subsequent reactions under these aggressive conditions. The hydrolytic hydrogenation of cellulose to sorbitol and sorbitan by mineral acids and supported Ru catalysts under H2 pressures of 70 bar offer improved sugar yields by hydrogenating glucose as it is formed to sugar alcohols, which have higher chemical stability (Fig. 19.5).54 Although promising, the use of mineral acids is undesirable, and the developments of processes employing only solid catalysts are preferred from safety, product recovery and catalyst recycle perspectives.55 Ru/CMK-356 and Ru/ZSM-557 have been reported to catalyse cellulose hydrolysis via hydrolytic hydrogenation, with the active Ru species on CMK-3 proposed as a hydrated Ru oxide (RuO2.2H2O) postulated to act as a Lewis or Brönsted acid depending on the degree of hydration.

image
19.5 Hydrolytic hydrogenation of cellulose to sugar polyols.

Supported metal catalysts such as Pt/γ-Al2O3, Pt/carbon and Ru/carbon are also reported to convert cellulose into sugar alcohols (sorbitol and mannitol) in the presence of hydrogen.55 High selectivity to glucose was achieved by choosing appropriate reaction conditions, which include rapid heating (from 25°C to 230°C in 15 min) to hydrolyse cellulose followed by cooling to inhibit glucose degradation.

19.3.3 Cellulose conversion in ionic liquids

Ionic liquids (IL) have shown huge promise for applications in biomass fractionation, and attracted significant academic and industrial interest.58 Low temperature ILs are low melting point salts, which form liquids comprising only cations and anions, having low vapour pressures and exceptional solvating properties for diverse compounds. The miscibility of ionic liquids with water or organic solvents can be controlled by varying the side chain lengths on the cation, and by careful anion choice. Some common cation and anion combinations used as ionic liquids are shown in Fig. 19.6. Organic groups on ILs can also be functionalised to impart acid or base character. While ILs are often described as ‘greener’ than conventional solvents, they are not all benign. The toxicity of ILs is mainly ascribed to the alkyl chain, with the toxicity of imidazolium and pyridinium ILs increasing with cation chain length.59 Furthermore, bmim (BF4, PF6, NTf2 and N(CN)2) ILs exhibit poor biodegradability (less than 5% after 28 days), hence large-scale use of such solvents in fuel production would require careful regulation.

image
19.6 Common cations and anions used in ionic liquids.

As outlined above, the β-glycosidic linkages in cellulose are protected against hydrolysis by tight packing of the cellulose chains into microfibrils, hence cellulose hydrolysis requires severe conditions, such as the use of dilute sulfuric acid at high temperatures. However, cellulose58 and wood60,61 dissolve in alkylmethylimidazolium ILs leaving the cellulose chains accessible to chemical transformations.62 ILs have been utilised as reaction media for the synthesis of cellulose derivatives such as carboxymethyl cellulose and cellulose acetate, while the regeneration of cellulose from ionic liquid solutions has been employed in the fabrication of films, gels and composite materials. Even though facile cellulose dissolution in IL solvents represents a significant breakthrough, the use of ILs for cellulose hydrolysis or degradation has only recently been explored.63

Early reports58 demonstrated that room temperature ILs, such as 1-n-butyl-3-methylimidazolium (C4mim+) salts with Cl, Br, and SCN anions, are capable of dissolving cellulose. High molecular weight pulp cellulose (5–10 g -DP 1000) slowly dissolved in ∼ 100 g ionic liquid when heated to 100°C, yielding viscous solutions. Subsequent studies have attempted to improve upon the initial discovery of this solubility. Interactions between the carbohydrate and anion of ILs appear the dominant factor in controlling cellulose dissolution, with Cl reported to have stronger hydrogen-bonding basicity than Br, SCN, PF6 and BF4. However, the chemical structure of cations also affects carbohydrate dissolution, with cellulose solubility decreasing with increasing alkyl chain length in the imidazolium cation, e.g. use of short alkyl chains in the AMIM cation enhances solubility relative to BMIM. The use of oxygenated side chains can also enhance carbohydrate solubility via hydrogen bonding between the oxygen and the carbohydrate. High throughput screening64 indicates that EMIM-Ac was the most efficient IL for dissolving cellulose, while AMIM-Cl (1-allyl-3methyl-imidazolium chloride) was the most effective for dissolving wood chips. A proposed mechanism by which ionic liquids break up the hydrogen bonding network in cellulose is shown in Fig. 19.7.65 Separation of the resulting sugar from the IL was achieved using ion exclusion chromatography, enabling over 95% recovery of the ionic liquid and 94% recovery of glucose. A commercial process for industrial-scale cellulose conversion to glucose would, however, require alternative separation technologies.

image
19.7 Disruption of the hydrogen bonding network in cellulose by ILs.

In a recent study by Li et al.,66 cellulose hydrolysis was initiated by adding catalytic amounts of H2SO4 to a cellulose–C4min+Cl solution, with a H2SO4:cellulose mass ratio of 0.92 producing total reducing sugars (TRS) and glucose in 59% and 36% yields, respectively, within 3 min. Lowering the acid:cellulose mass ratio to 0.46 produced higher yields after 42 min reaction, and for a mass ratio of only 0.11, TRS and glucose yields of 77% and 43%, respectively, were achieved after 9 h. These mild operating conditions and the use of a catalytic amount of H2SO4 offer exciting prospects for cellulose conversion.

Incorporation of an acidic function into the ILs has also been investigated through the synthesis of Brönsted acid ILs for the simultaneous dissolution and hydrolysis of cellulose.67 Such Brönsted acidic ILs can act as both solvent and catalyst, eliminating waste conventionally produced during mineral acid neutralisation and separation steps. Furthermore, the high concentration of –SO3H active sites that can be introduced to ILs is expected to accelerate reactions and thus enable lower operating temperatures, facilitating cellulose dissolution and hydrolysis under moderate reaction temperatures and atmospheric pressure. Three types of Brönsted acidic ILs based on methyl imidazolium (1a,b), pyridinium (2), and triethanolammonium (3) have been studied for their ability to dissolve and hydrolyse cellulose at mild reaction temperatures (Fig. 19.8). Cellulose dissolution in Brönsted acidic ionic liquids such as 1-(1-propylsulfonic)-3-methylimidazolium chloride and 1-(1-butylsulfonic)-3-methylimidazolium chloride is achievable up to 20 g/100 g IL upon gentle room temperature mixing.

image
19.8 Acidic ILs explored in cellulose hydrolysis.

Solid acids have been reported to selectively catalyse the depolymerisation of cellulose solubilised in 1-butyl-3-methylimidazolium chloride (BMIMCl) at 100°C. Acid strength plays a key role in cellulose depolymerisation,50 wherein the resin-based solid acid Amberlyst gave impressive results for cellulose hydrolysis, despite diffusional limitations and poor accessibility of the attendant acid sites. Reactions using solid catalysts preferentially cleave longer cellulose chains to produce oligomers consisting of approximately ten anhydroglucose units (AGU). However, care must be taken with the use of some resin-based catalysts, as these may degrade under reaction conditions leading to homogeneous catalysis.68 For example, Amberlyst-15 is reported to be very stable in BMIMCl, but a simple change to BMIM(CH3COO), a potentially more desirable solvent because of its lower viscosity, resulted in rapid Amberlyst-15 degradation. The poor activity of traditional inorganic solid acids (e.g., Zeolite-Y and ZSM-5) has been attributed to their microporous nature and consequent poor accessibility to bulky cellulose fibres.

Cellulose has been directly converted into environmentally friendly alkyl glycoside surfactants in a one-pot transformation.Utilising BMIMCl in conjunction with an Amberlyst 15Dry (A15) catalyst, and coupling the rates of cellulose hydrolysis and glycosidation of the resultant monosaccharides with C4-C8 alcohols, an 82% mass yield of octyl-α-β-glucoside and octylα-β-xyloside was obtained.69 The IL alone was unable to effect any cellulose conversion, possibly reflecting the low reaction temperature employed, since ILs are not known to hydrolyse cellulose below 90°C.70

Solid acid-promoted hydrolysis of cellulose in ionic liquids can be accelerated by microwave heating, affording similar yields of TRS and glucose to those obtained using liquid mineral acids. Protonated zeolites with a low Si/Al molar ratio and large surface area showed the highest catalytic activity, outperforming the acidic sulfated ion-exchange resin NKC-9.71 Despite these initial successes, significant breakthroughs in catalyst and process design are essential in order to improve the overall efficiency of glucose production for subsequent bio refining.

19.4 Types and properties of carbohydrate-based chemicals

The US Department of Energy (DoE) has identified 12 platform chemicals obtainable through sugars via the chemical or biochemical transformation of lignocellulosic biomass (Fig. 19.9).5 Akin to petrochemical refineries, the co-production of high value chemicals and fuels by thermo-72 and/or biochemical 73 routes, along with heat and power in an integrated biorefinery, offers the most economically viable means to utilise biomass.74 The idea of utilising bio-oils produced from biomass pyrolysis is an emerging concept which could potentially offer a feed stream that could be directly distilled75 or co-refined alongside fossil fuel-derived oils in conventional petroleum refineries.76,77 Hydrodeoxygenation (HDO) of crude bio-oils can yield hydrocarbons with similar properties to crude petroleum oil.

image
19.9 Possible platform chemicals produced from biomass.78

19.4.1 Biochemical conversion of carbohydrates

The use of enzymes in biochemical routes allows for selective conversion of sugars via fermentation, but faces several limitations that hinder economic biomass processing. Notably, conventional enzyme catalysed processes are unable to process pentoses, which must therefore be removed from biomass sugar feeds, requiring extra pretreatment steps in order to purify the glucose feedstock. Furthermore, the platform molecules derived from fermentation are often present at low concentrations (typically < 10%) in aqueous solutions, alongside other polar molecules. Purification of such fermentation broths is particularly difficult,79,80 and energetically unfavourable, hence an ability to directly transform these aqueous solutions would be desirable.81 Catalysts capable of driving organic chemistry in water to selectively transform platform molecules into useful chemical feedstocks,5,9,74,82,83 and are resistant to impurities present in the fermentation broth,84 therefore require development.

The utilisation of biomass-derived chemicals, whether from fermentation broths or sugars themselves, represents an area with extensive R&D potential for a renewable feedstock-based technology platform. Approaches to handling biomass-derived building blocks will be very different from petroleum processing, e.g. requiring reverse chemical transformations wherein highly functional bio-molecules are deoxygenated to their target product, instead of oxygenated as is usual when starting from crude oil.85 Figure 19.10 illustrates a possible biomass synthesis of adipic acid (currently manufactured via the selective oxidation of cyclohexane) involving the selective reduction of glucose. New classes of catalyst are urgently required which are compatible with hydrophilic, bulky substrates to facilitate the move away from existing short-chain hydrocarbon supplies. Improvements and innovations in catalyst and processes design are needed in order to deliver high-value chemicals from biomass-derived building blocks.

image
19.10 Alternative routes to adipic acid from biomass or petroleum feedstocks.

These new catalysts will be hydrophilic, stable over a wide pH range and resistant to in-situ leaching.86 Catalyst porosity will also be important to enable diffusion of bulky, viscous reactants to active sites; support materials with larger pores (compared to zeolites) are thus a promising starting point. Organic-inorganic hybrid catalysts may also prove interesting, as these allow catalyst hydrophobicity to be readily tuned, and in turn the adsorption strengths of polar molecules.87 Mesoporous carbons88 are well-suited for biomass conversion since they tend to be highly resistant to acidic and chelating media. Transforming the functional groups on platform chemicals will require catalysts capable of dehydration, hydrogenolysis and hydrogenation chemistry. Suitable catalysts will thus contain acid sites, or exhibit bifunctional character, possessing acid sites for dehydration and metal sites for (de)hydrogenation. Corma et al. have extensively reviewed proposed methods for transforming platform molecules into chemicals,9 many of which employ conventional homogeneous reagents or commercial catalysts. There is thus enormous scope for rationally designing improved heterogeneously catalysed processes tailored towards biomass-derived feedstocks.

A number of reports have described pathways to important chemical intermediates from platform molecules.9,82,83,89 Succinic acid is proposed as a valuable platform chemical, from which a range of chemical intermediates can be derived, as illustrated in Fig. 19.11 for acid catalysed esterification, or metal catalysed reductions. Carbon-based solid acid catalysts have proven effective for succinic acid esterification with ethanol.90,91 Succinic acid may also afford new biopolymers based on polyesters, polyamides and polyesteramides.92

image
19.11 Selected acid catalysed or hydrogenation products of succinic acid.

The purity of crude biorefinery feeds presents a major challenge for catalyst development, with the fermentation broth typically produced as a salty medium containing diammonium succinate rather than pure succinic acid. However, this could be exploited in the production of 2-pyrrolidone or N-methyl-pyrrolidone via hydrogenation of diammonium succinate 93 using catalysts such as Pd/ZrO2/C and 2.5% Rh-2.5%Re on C, with reaction in the presence of methanol favouring N-methyl-pyrrolidone formation.

19.4.2 Thermochemical conversion of carbohydrates

Carbohydrate conversion by thermochemical routes such as fast pyrolysis, aqueous phase reforming or gasification offers alternative building blocks for chemicals or fuels production. Thermal decomposition of biomass in the absence of oxygen by pyrolysis yields a range of feedstocks depending on the reaction temperature and residence time,77 with low temperatures and long residence times favouring charcoal formation, while moderate temperatures and short vapour residence time are optimal for liquids production. High temperatures and long residence times favour gasification,94 with the resulting syngas available for well-established catalytic processes such as Fischer–Tropsch and methanol synthesis routes to convert CO/H2 mixes to fuels and methanol. While complete gasification may be energetically costly, less initial biomass processing is required, so with efficient heat recovery modules, this approach is attractive for its compatibility with existing industrial processes. Low temperature (< 200°C), thermochemical aqueous phase processing of sugars is of particular interest, as this offers a viable method to generate highly functional intermediates such as aldehydes, alcohols and esters via dehydration, hydrogenolysis and isomerisation (Fig. 19.12).55,74,95

image
19.12 Major components formed via thermochemical processing of hemicellulose and cellulose.

Solid acids or bases are commonly employed in aqueous phase processing, although the application of bi-functional metal-doped catalysts is attracting interest for combined dehydration/hydrogenation of sugar intermediates (HDO). Some examples of key catalytic transformations of platform molecules will now be described.

19.4.3 Conversion of C6 sugars to 5-HMF

5-Hydroxymethylfurfural (HMF) is a popular platform molecule with huge potential as an important bio-based commodity chemical96,97 for the synthesis of various commercially useful acids, aldehydes, alcohols, and amines, as well as the promising fuel 2,5-dimethylfuran (DMF)98 and renewable monomer furan dicarboxylic acid (FDCA)99 as shown in Fig 19.13. HMF also has potential as a building block for the manufacture of commodity chemicals such as caprolactam, the precursor to Nylon 6,6,100 following hydrogenation and hydrogenolysis to 1,6-hexanediol.

image
19.13 Potential chemical feedstocks derived from 5-HMF.

While dehydration of C6 sugars to 5-hydroxymethylfurfural (5-HMF) can readily be achieved by acid-catalysed dehydration of three water molecules,101 glucose conversion by Brönsted acids often proceeds with low selectivity to 5-HMF due to competing side reactions which form humins.102,103 In contrast, fructose conversion proceeds with higher selectivity to 5-HMF, hence bifunctional catalysts capable of isomerising glucose to fructose prior to a subsequent acid-catalysed dehydration would be desirable. For example, under high temperature conditions of hot compressed water (200°C), ZrO2 can promote glucose and fructose isomerisation via a base-catalysed route. Anatase TiO2, which possesses both acid and base character, promotes glucose isomerisation and dehydration into HMF (Fig. 19.14).104 For more selective chemistry, tailored solid catalysts capable of working at lower temperatures are sought.

image
19.14 Proposed acid- and base-catalysed pathways in the dehydration of glucose to 5-HMF over ZrO2 and TiO2 in hot compressed water.

Initial reports of sucrose dehydration over solid acids, including acidic resins,105 HY-zeolite,106 aluminium-pillared montmorillonite, MCM-20 and MCM-41,107 have sparked significant interest in 5-HMF production. Fructose dehydration to 5-HMF can be initiated by Brönsted acids in polar solvents and a range of aprotic polar solvents, such as dimethyl sulfoxide (DMSO), N,N-dimethylformamide (DMF), N,N-dimethylacetamide (DMA) and sulfolane. A variety of solid acids, including ion-exchange resins,108 zeolites109,110 metal oxides,111 heteropoly acids,112 niobic acid,113 niobium phosphate,114 sulfated zirconia,115 and sulfonic acid-functionalised mesoporous silicas116 have been explored for 5-HMF production from fructose. Detailed studies of WOx/ZrO2 catalysts reveal111 the importance of catalyst amphoteric character in achieving high selectivity to HMF during fructose dehydration. While the overall fructose conversion correlates with acid site density, optimum selectivity towards 5-HMF of ∼ 40% was obtained for catalysts with a base:acid site ratio of 0.3. Solid acid-catalysed conversion of fructose to 5-HMF has also been explored in ILs.117121

As discussed above, the direct conversion of glucose to 5-HMF is difficult, requiring isomerisation to fructose, either via proton transfer or an intramolecular hydride shift, respectively base or Lewis acid catalysed. Base-catalysed glucose to fructose isomerisation occurs via deprotonation of the α-carbonyl carbon of glucose to form a series of enolate intermediates, and can be conducted by cation-exchanged zeolites or Mg-Al hydrotalcites.122,123 High 5-HMF yields are obtained upon subsequent acid-catalysed dehydration of the resulting fructose.

The Lewis-Brönsted acid ratio also plays an important role in directing selectivity during glucose dehydration.124 Davis and coworkers showed that tin-containing zeolites are highly active catalysts for glucose isomerisation in water, wherein Sn behaves as a Lewis acid catalysing an intramolecular hydride shift.125,126 In order to minimise side reactions, biphasic systems, employing Lewis and Brönsted catalysts in conjunction with reactive extraction, have been proposed to improve 5-HMF yields (Fig. 19.15).97,127 The combination of Lewis and Brönsted acidity is beneficial for 5-HMF formation from glucose, with Sn-β (a Lewis acid) and HCl (a homogeneous Brönsted acid) offering good 5-HMF yield from glucose in a water-tetrahydrofuransolvent mix.128 A similar approach was adopted using homogeneous Lewis acid metal chlorides (e.g., AlCl3) and HCl in a biphasic reactor comprising water and 2-sec butylphenol.129 In this respect, amorphous niobium oxide hydrate (Nb2O5 · nH2O; niobic acid) is an interesting candidate for HMF production from glucose in water, as it possesses water-tolerant Lewis acid sites co-existing alongside Brönsted acid sites.130

image
19.15 Biphasic system for reactive extraction of 5-HMF during acid-catalysed fructose (or glucose) dehydration. Reprinted with permission from Reference 97. Copyright (2010) American Chemical Society.

5-HMF yields are also enhanced over mesoporous catalysts with pore diameters of 1–3 nm. Porosity was observed to have a significant effect on 5-HMF and levulinic acid yields obtained from sucrose dehydration over acid exchange resins.105 Larger pores favour 5-HMF production, with slow 5-HMF diffusion out of smaller pores appearing to promote subsequent reaction and higher selectivity to levulinic acid.

19.4.4 Conversion of C5 sugars to furfural

Xylose, derived from hemicelluloses, can be readily dehydrated to furfural using Brönsted acids at high temperature. A variety of Brönsted acid catalysts have been examined for furfural synthesis including heteropoly acids131,132 and mesoporous sulfonic acid silicas,133,134 H-type zeolites such as H-mordenite and H-Y faujasite,109 ion-exchange resins,135 niobium silicate,136 SO4/ZrO2134 and SO4/SnO2.138

Oxidative ring opening of furfural using H2O2 in the presence of a solid acid offers an opportunity to produce succinic acid along with maleic acid as a by-product (Fig. 19.16).139 The yield of succinic acid was optimal using Amberlyst-15 when compared to other resin-based solid acids such as Nafion NR50 and Nafion SAC13, or other inorganic solid acid catalysts such as Nb2O5, H-ZSM-5 and ZrO2.

image
19.16 Oxidative ring opening of furfural to succinic acid.

The catalytic aerobic oxidation of furfural to maleic acid has also been explored using phosphomolybdic acid in biphasic aqueous/organic systems (Fig. 19.17), in which oxidation occurs in the aqueous phase, while the organic phase serves as the reservoir for furfural. Using this approach, maleic acid was obtained with 69% selectivity at a furfural conversion of 50%. Product separation was aided by the fact that furfural and maleic acid predominantly phase separate.140

image
19.17 Biphasic system for the reactive ring opening of furfural to maleic acid. Reprinted with permission from Ref. 140. Copyright (2011) American Chemical Society.

19.4.5 Levulinic acid production

Levulinic acid is another valuable precursor to a range of chemical intermediates (Fig. 19.18), which can be used in chiral reagents, inks, coatings and batteries, and can be generated by a combination of acid-catalysed dehydration/esterification or metal-catalysed reduction.141 While a number of studies have investigated reductions, there is surprisingly little work concerning the esterification of platform molecules using solid acid catalysts. To date, levulinic acid esterification is mostly performed using H2SO4,142 with only a few studies employing solid acids such as sulfated TiO2 and SnO2,143,144or heteropoly acids.145,146

image
19.18 Selected acid-catalysed or hydrogenation products from levulinic acid.

Amberlyst-70 has also been reported to yield 21% of levulinic acid following dehydration of the water-soluble organics obtained via hydrothermal cellulose decomposition at 160°C.147 When converting saccharides to levulinic acid, resin pore size was also demonstrated to have a significant effect on product selectivity.105 There is clearly scope for the development of new catalytic systems for the conversion of biorefinery feedstocks to chemicals. However, efforts must focus on water-tolerant catalysts for the direct reaction of aqueous feeds, and the development of tailored porous-solids capable of operating under continuous flow.

19.4.6 Lactic acid synthesis

Sugar fermentation to lactic acid has received much attention in the context of polylactic acid (PLA) synthesis, considered the ‘gold standard’ for renewable polymers and applications including solvents and coatings.148 Fermentation routes to LA proceed through homolactic or heterolactic fermentation of glucose, via glycolysis to pyruvate,149 typically yielding > 90% lactic acid as a 100 g.l− 1 aqueous solution of Ca-lactate. Subsequent H2SO4 recovery of lactic acid generates 1 kg CaSO4 waste per kg lactic acid produced. Homolactic routes are less efficient, with ∼ 44% lactate yields reported150 at concentrations up to 55–60 g.l− 1.151 Extraction costs account for 50% of such process economics,152 due to the high energy consumption for separation and purification. Process improvements have been proposed employing ammonia to isolate an ammonium L-lactate product. 153 However, for large-scale processing, a catalytic route would be a desirable alternative.

Lactic acid synthesis has also been proposed via the retro-aldol condensation of fructose or glucose to form glyceraldehyde (GLA) and dihydroxyacetone (DHA).154 These require Lewis acid or solid base catalysts to achieve high selectivity, with Sn-β,155,156 Sn-SiO2,157 Sn-SiO2-carbon composite,158 and Brönsted base hydrotalcite catalysts,159 reported as potential candidates for lactic acid or lactate synthesis from sugars. Conversion of GLA and DHA is proposed to yield pyruvaldehyde, which is subsequently hydrated to lactic acid via an internal Cannizzaro reaction, Meerwein–Ponndorf–Verley reduction or Oppenauer oxidation (Fig. 19.19).

image
19.19 Proposed reaction scheme for converting monosaccharides such as trioses and aldo- and ketohexoses like fructose into lactic acid in aqueous medium (R = H) or into alkyl lactate in alcoholic solvents (R = alkyl). The side-reaction leading to the formation of pyruvic aldehyde dialkylacetal (b) is undesired; LA = Lewis acid and BA = Brönsted acid. Reprinted with permission from Ref. 158. Copyright (2012) American Chemical Society.

19.5 Routes to market for bio-based feedstocks

19.5.1 Challenges and opportunities for introduction of bio-based products

The most significant driver for the production of renewable chemicals is the price of crude oil. When oil prices exceed US$100 per barrel, bio-based feedstocks may compete with petroleum sources.160 However, there are other motivators for renewable feedstocks such as green marketing initiatives, individual preferences for more sustainable products, and critically their potential to mitigate pollution and CO2 emissions. There are a number of potential markets for bio-derived chemicals, most notably as identical replacements for current commodity chemicals, so called ‘drop-in’ chemicals (e.g., bio-ethene, nylon or PET) or as completely new types of platform chemical (e.g., PLA, or PEF from FDCA).

Current and potential commercial applications of bio-based chemicals have been reviewed by Erickson et al.,161 and some near-term commercial and future bio-products are summarised in Table 19.1. Succinic acid and 1,4-butanediol are identified as near-term, bio-based chemicals, while longer term targets include levulinic acid and adipic acid, the latter an important precursor to nylon.

Table 19.1

(a) Near commercial and (b) future commercial bio-based chemicals

(a)
Chemical Companies Applications
Succinic acid Myriant BioAmber, DSM Flavourings, dyes, perfumes, lacquers
1,4-Butanediol Genomatica Spandex,
Isoprene Genencor Amyris Rubber, adhesives
Isobutanol Gevo Solvents, paint, biofuels
(b)
Chemical Companies Applications
Levulinic acid Dupont Tate & Lyle Plasticisers, solvents,
Adipic acid Verdezyne Fibres, plastics
Itaconic acid Itaconix Pigments, stabilisers

Image

Source: Adapted from Ref. 161.

Using polyamide synthesis as a case study, a number of building blocks have been developed around natural products using biotechnology. For example, the unsaturated fatty acids sebacic or undecylenic acid derived from castor oil are employed in nylon 6,10 and 10,10 manufacture for use in automobile components (e.g., by Rhodia and Dupont),162 while Cognis reportedly employ an enzymatic oxidative scission of oleic acids to produce muconic acid (a key precursor to adipic acid).163 While these routes produce monomers for new materials, ‘drop in’ chemicals as direct replacements for those currently obtained via crude oil feedstocks are also essential for a sustainable chemicals industry. Promising biochemical routes to adipic acid via fermentation routes to muconic acid, were reported in the early 1990s. However, such approaches remain uneconomic, operating at only a 22% carbon balance, and producing dilute ∼ 3.68 wt% aqueous solutions of muconic acid which impose excessive separation and purification costs.164 Despite these hurdles, US biotech companies (e.g., Verdezyne) are attempting to scale-up new biocatalytic production routes for commercialisation. These efforts face stiff competition from a recent, cheaper petrochemical route to nylon 6 via butadiene carbonylation, developed by DSM, DuPont and Shell.165 Selective catalytic routes to bionylon from a dipic acid or caprolactam (that avoid costly separation processes) would thus be particularly attractive.100

19.5.2 Challenges for process design

Conversion of bio-based feedstocks presents new challenges to the catalytic scientist, as the attendant reaction conditions are very different from those typical of petroleum processing, which occurs mainly through vapour phase processes > 400°C. Biomass processing will be characterised by liquid phase, lower temperature41 pathways such as hydrolysis, dehydration, isomerisation, oxidation, aldol, condensation, and hydrogenation.166 The design of catalysts for such biomass transformations requires careful tailoring of pore structure to minimise mass transport limitations, hydrothermal stability under aqueous operation, and tunable hydrophobicity to aid product/reactant adsorption. 167 Catalyst development should thus focus on the use of tailored porous solids as high area supports to enhance reactant accessibility to active acid/basic groups. The preparation of such templated porous solids has been extensively reviewed168171 but generally involves the use of micellar templates to direct the growth of metal oxide frameworks as shown in Fig. 19.20. Subsequent calcination to burn out the organic template, or solvent extraction, yields materials with well-defined meso-structured pores of 2–10 nm and surface areas up to 1000 m2.g− 1.

image
19.20 (a) Liquid crystal templating route to form mesoporous silica and combined physical templating method using polystyrene microspheres to introduce a macropore network; (b) SEM of macroporous-mesoporous SBA-15 showing macropore network and TEM showing interconnecting mesopores.

Macropores can also be introduced if an additional physical template, such as polystyrene microspheres, is added during templating of the mesopore network.172 Hierarchical macroporous-mesoporous sulfonic acid SBA-15 silicas and aluminas have been synthesised via such dual-templating routes, employing liquid crystalline surfactants and polystyrene beads.173,174 These materials offer high surface areas and well-defined, interconnected macro- and mesopore networks, with respective narrow size distributions tunable over the range 100–300 nm and 3–5 nm. Such bimodal solid acid architectures offer significantly enhanced activity over mesoporous analogues. A second challenge centres around the need to better understand the selective deoxygenation and hydrogenation of polyols under acid- or base-catalysed conditions. Bifunctional catalysts need to be capable of low temperature dehydration and hydrogenation, and efficient use of H2 during in situ hydrogenation and/or hydrogenolysis. Surface polarity is another important parameter to control and thereby permit catalyst operation in biphasic systems. As discussed in relation to Fig. 19.15, partitioning of products between aqueous and organic phase solvents via reactive extraction is a promising means to minimise side reactions.175,176

Hydrophobic catalysts operating in biphasic systems have been explored for ‘bio-oil’ obtained via biomass pyrolysis; bio-oil is a complex liquid that is only partially soluble in water or hydrocarbon solvents. In order to overcome problems of working with such mixtures, materials with tunable hydrophobicity, or even the use of micellar catalyst systems should be considered. This has recently been exploited by the group of Resasco,176 in which a phase transfer system based upon Pd nanoparticles immobilised on carbon nanotube/silica composite support was employed. Such materials are able to catalyse the transformation of both hydrophilic and hydrophobic substrates to fuels at the oil:water interface, without the need for multiple separation steps or addition of surfactants.177

The development of new reactor technology will also be critical for efficient biomass processing.178 Conventional stirred batch reactors are not suited to commodity chemical production scales, hence there is a need to devise slurry reactors that operate continuously. Intensive processing179 could offer exciting means to improve overall process efficiency. For example, process engineering solutions available for continuous reactions include the use of fixed-bed180 or microchannel-flow reactors,181 pervaporation methods,182,183 or reactive distillation.184186 Continuous reactors must be carefully designed to utilise the full potential of the associated heterogeneous catalyst; plug flow is a desirable reactor characteristic, as it allows tighter control of product composition and thereby reduces downstream separation processes and capital and running costs. However, conventional plug flow reactors are unsuitable for slow reactions as they necessitate very high length:diameter ratios to achieve the required mixing. Such designs are problematic due to their large footprints, pumping duties and control difficulties.

The oscillatory baffled reactor (OBR) circumvents these problems, offering good mixing and plug flow by oscillating the reaction fluid through orifice plate baffles.187 In an OBR, mixing is decoupled from the net flow: this allows reactor designs that are very different from conventional plug flow reactors. OBR mixing is also scalable, meaning that longer reaction times can be achieved on industrial scales. Previous work188,189 has shown that vortical mixing in the OBR is an effective, controllable method of uniformly suspending solid (e.g., catalyst) particles. In this respect, the first such demonstration of a solid acid catalyst incorporated within an OBR for the esterification of organic acids to short chain and fatty methyl esters, pertinent to fine chemicals synthesis and biofuels production, was recently reported. 190

The development of more efficient separation methods will also be crucial191 with advances in distillation, extraction, adsorption with molecular sieves, filtration, crystallisation and osmosis holding great promise.

19.6 Conclusion and future trends

This chapter has highlighted the significant progress made in recent years towards the conversion of renewable feedstocks into chemicals and fuels. Heterogeneous catalysis and process engineering hold the key to realising the potential of lignocellulosic biomass for the production of such renewable chemicals. Ultimately, catalytic chemists and engineers need to emulate the successes of heterogeneous catalysis in petroleum refining. Advances in chemistry, nanotechnology and spectroscopy will aid catalyst design, but overall process development requires an improved understanding of biomass properties and its impact on catalyst deactivation in order to accelerate biomass-to-chemicals and fuels production. Commercial heterogeneously catalysed processes will require a better understanding of individual reactant interactions with the active catalyst phase, particularly when dealing with bulky polar molecules such as those found in biorefinery feeds.

Most crucially however, widespread uptake and the development of next-generation biofuels and chemical feedstocks requires progressive government policies and incentive schemes to place biomass-derived chemicals on a comparable footing with cheaper fossil fuel-derived resources.4,192 Biomass pretreatment to obtain sugars is one of the most wasteful steps in biorefineries, and new approaches are required to improve the processing of lignocellulose such that the initial acid hydrolysis/extraction step to form lignocellulose can be performed more efficiently. Cellulose stability itself presents a major hurdle, with environmentally friendly and energy efficient means to break up this biopolymer an ongoing challenge. In contrast to petroleum-derived oil, conventional heterogeneous catalysts cannot be employed alone in solid–solid mixtures, although some recent reports propose ball milling as an effective means to induce ‘mechanocatalysis’ between cellulose and clay-based catalysts with layered structures.193 Alternative approaches are building upon the exciting discovery that ionic liquids can dissolve cellulose, and when coupled with acidic reagents can also generate selected platform chemicals.61,194 The latter approach has been coupled with solid catalysts50,195 to combine the ease of separation of a solid catalyst, with the dissolution strength of ILs, offering an exciting prospect for the cleaner conversion of cellulose to chemicals.

Development of new catalysts and overall process optimisation requires collaboration between catalytic chemists, chemical engineers and experts in molecular simulation to take advantage of innovative reactor designs; the future of renewable feedstock utilisation requires a concerted effort from chemists and engineers to develop catalysts and reactors in tandem. Current political concerns over the ‘food versus fuel’ debate also require urgent development of non-edible oil feedstocks, as well as necessary technical advances in order to ensure that catalytic routes to convert biomass to chemicals and fuels become viable processes in the renewables sector during the twenty-first century.

19.7 Sources of further information and advice

A number of global organisations exist performing research or acting in a consulting role concerning biomass utilisation. Useful websites include:

• Bioproducts, Sciences, and Engineering Laboratory (BSEL) at Pacific Northwest National Laboratory: http://www.pnnl.gov/biobased/bsel.stm

• NREL National Renewables Energy Laboratory: http://www.nrel.gov/

• National Non-Food Crops Centre: http://www.nnfcc.co.uk/NNFCC

• European Biomass Association: http://www.aebiom.org/

19.8 References

1. Walter B, Gruson JF, Monnier G. Oil & Gas Science and Technology – Revue D Ifp Energie Nouvelles. 2008;63:387–393.

2. Armaroli N, Balzani V. Angew Chem.-Int Edit. 2007;46:52–66.

3. Chen G-Q, Patel MK. Chemical Reviews. 2011;112:2082–2099.

4. Azadi P, Inderwildi OR, Farnood R, King DA. Renewable and Sustainable Energy Reviews. 2013;21:506–523.

5. Bozell JJ, Petersen GR. Green Chemistry. 2010;12:539–554.

6. Danielsen F, Beukema H, Burgess ND, et al. Conservation Biology. 2009;23:348–358.

7. Kamm B, Kamm M. Chemie Ingenieur Technik. 2007;79:592–603.

8. Kamm B. Angew Chem.-Int Edit. 2007;46:5056–5058.

9. Corma A, Iborra S, Velty A. Chemical Reviews. 2007;107:2411–2502.

10. Gallezot P. Catalysis Today. 2007;121:76–91.

11. Lichtenthaler FW, Peters S. Comptes Rendus Chimie. 2004;7:65–90.

12. Kim M, Day DF. J Ind Microbiol Biotechnol. 2011;38:803–807.

13. Tew TL, Cobill RM, Richard EP. BioEnergy Res. 2008;1:147–152.

14. Charlton A, Elias R, Fish S, Fowler P, Gallagher J. Chemical Engineering Research and Design. 2009;87:1147–1161.

15. Zheng Y, Yu CW, Cheng YS, et al. Appl Energy. 2012;93:168–175.

16. Edwards MC, Doran-Peterson J. Appl Microbiol Biotechnol. 2012;95:565–575.

17. Cherubini F. Energy Conversion and Management. 2010;51:1412–1421.

18. Menon V, Rao M. Progress in Energy and Combustion Science. 2012;38:522–550.

19. Hoogwijk M, Faaij A, van den Broek R, Berndes G, Gielen D, Turkenburg W. Biomass and Bioenergy. 2003;25:119–133.

20. Kadam KL, Forrest LH, Jacobson WA. Biomass and Bioenergy. 2000;18:369–389.

21. Mosier N, Wyman C, Dale B, et al. Bioresource Technology. 2005;96:673–686.

22. Garde A, Jonsson G, Schmidt AS, Ahring BK. Bioresource Technology. 2002;81:217–223.

23. Glasser WG, Wright RS. Biomass and Bioenergy. 1998;14:219–235.

24. Zhao X, Cheng K, Liu D. Appl Microbiol Biotechnol. 2009;82:815–827.

25. Fatih Demirbas M. Appl Energy. 2009;86(Supplement 1):S151–S161.

26. Alvira P, Tomás-Pejó E, Ballesteros M, Negro MJ. Bioresource Technology. 2010;101:4851–4861.

27. Mok WS, Antal MJ, Varhegyi G. Industrial & Engineering Chemistry Research. 1992;31:94–100.

28. Rinaldi R, Schüth F. ChemSusChem. 2009;2:1096–1107.

29. Sasaki M, Kabyemela B, Malaluan R, et al. The Journal of Supercritical Fluids. 1998;13:261–268.

30. Binder JB, Raines RT. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:4516–4521.

31. Suhas, Carrott PJM, Ribeiro Carrott MML. Bioresource Technology. 2007;98:2301–2312.

32. Gírio FM, Fonseca C, Carvalheiro F, Duarte LC, Marques S, Bogel- Łukasik R. Bioresource Technology. 2010;101:4775–4800.

33. Chakar FS, Ragauskas AJ. Industrial Crops and Products. 2004;20:131–141.

34. Ruiz E, Cara C, Manzanares P, Ballesteros M, Castro E. Enzyme and Microbial Technology. 2008;42:160–166.

35. Dudley RL, Fyfe CA, Stephenson PJ, Deslandes Y, Hamer GK, Marchessault RH. Journal of the American Chemical Society. 1983;105:2469–2472.

36. Torget RW, Kim JS, Lee YY. Industrial & Engineering Chemistry Research. 2000;39:2817–2825.

37. Yu Y, Lou X, Wu H. Energy & Fuels. 2008;22:46–60.

38. Saeman JF. Industrial and Engineering Chemistry. 1945;37:43–52.

39. Conner AH, Wood BF, Hill CG, Harris JF. Journal of Wood Chemistry and Technology. 1985;5:461–489.

40. Bouchard J, Abatzoglou N, Chornet E, Overend RP. Wood Science and Technology. 1989;23:343–355.

41. Huber GW, Dumesic JA. Catalysis Today. 2006;111:119–132.

42. Rinaldi R, Schuth F. Chemsuschem. 2009;2:1096–1107.

43. Van de Vyver S, Geboers J, Jacobs PA, Sels BF. Chemcatchem. 2011;3:82–94.

44. Kobayashi H, Ohta H, Fukuoka A. Catalysis Science & Technology. 2012;2:869–883.

45. Dhepe PL, Ohashi M, Inagaki S, Ichikawa M, Fukuoka A. Catalysis Letters. 2005;102:163–169.

46. Suganuma S, Nakajima K, Kitano M, et al. Journal of the American Chemical Society. 2008;130:12787–12793.

47. Onda A, Ochi T, Yanagisawa K. Green Chemistry. 2008;10:1033–1037.

48. Pang J, Wang A, Zheng M, Zhang T. Chemical Communications. 2010;46:6935–6937.

49. Van de Vyver S, Peng L, Geboers J, et al. Green Chemistry. 2010;12:1560–1563.

50. Rinaldi R, Palkovits R, Schuth F. Angew Chem.-Int Edit. 2008;47:8047–8050.

51. Yamaguchi D, Kitano M, Suganuma S, Nakajima K, Kato H, Hara M. Journal of Physical Chemistry C. 2009;113:3181–3188.

52. Takagaki A, Nishimura M, Nishimura S, Ebitani K. Chemistry Letters. 2011;40:1195–1197.

53. Deguchi S, Tsujii K, Horikoshi K. In: Chemical Communications. 2006;3293–3295.

54. Balandin AA, Vasunina NA, Chepigo SV, Barysheva GS. Doklady Akademii Nauk Sssr. 1959;128:941–944.

55. Ruppert AM, Weinberg K, Palkovits R. Angew Chem.-Int Edit. 2012;51:2564–2601.

56. Kobayashi H, Komanoya T, Hara K, Fukuoka A. Chemsuschem. 2010;3:440–443.

57. Geboers J, Van de Vyver S, Carpentier K, Jacobs P, Sels B. Chemical Communications. 2011;47:5590–5592.

58. Swatloski RP, Spear SK, Holbrey JD, Rogers RD. Journal of the American Chemical Society. 2002;124:4974–4975.

59. Zakrzewska ME, Bogel-Lukasik E, Bogel-Lukasik R. Energy & Fuels. 2010;24:737–745.

60. Fort DA, Remsing RC, Swatloski RP, Moyna P, Moyna G, Rogers RD. Green Chemistry. 2007;9:63–69.

61. Brandt A, Hallett JP, Leak DJ, Murphy RJ, Welton T. Green Chemistry. 2010;12:672–679.

62. El Seoud OA, Koschella A, Fidale LC, Dorn S, Heinze T. Biomacromolecules. 2007;8:2629–2647.

63. Pinkert A, Marsh KN, Pang S, Staiger MP. Chemical Reviews. 2009;109:6712–6728.

64. Zavrel M, Bross D, Funke M, Buchs J, Spiess AC. Bioresource Technology. 2009;100:2580–2587.

65. Feng L, Chen ZI. Journal of Molecular Liquids. 2008;142:1–5.

66. Li C, Wang Q, Zhao ZK. Green Chemistry. 2008;10:177–182.

67. Amarasekara AS, Owereh OS. Industrial & Engineering Chemistry Research. 2009;48:10152–10155.

68. Rinaldi R, Meine N, vom Stein J, Palkovits R, Schueth F. Chemsuschem. 2010;3:266–276.

69. Villandier N, Corma A. Chemical Communications. 2010;46:4408–4410.

70. Zhang YT, Du HB, Qian XH, Chen EYX. Energy & Fuels. 2010;24:2410–2417.

71. Zhang ZH, Zhao ZBK. Carbohydrate Research. 2009;344:2069–2072.

72. Bridgwater AV. Biomass & Bioenergy. 2012;38:68–94.

73. Rabinovitch-Deere CA, Oliver JWK, Rodriguez GM, Atsumi S. Chemical Reviews. 2013;113:4611–4632.

74. Fernando S, Adhikari S, Chandrapal C, Murali N. Energy & Fuels. 2006;20:1727–1737.

75. Zhang X-S, Yang G-X, Jiang H, Liu W-J, Ding H-S. Sci Rep. 2013;3:1120.

76. De Miguel Mercader F. PhD Thesis Pyrolysis oil upgrading for co-processing in standard refinery units The Netherlands: University of Twente; 2010.

77. Czernik S, Bridgwater AV. Energy & Fuels. 2004;18:590–598.

78. Werpy T, Petersen G. Top Value Added Chemicals From Biomass Volume I: Results of Screening for Potential Candidates from Sugars and Synthesis Gas Pacific Northwest National Laboratory (PNNL) and National Renewable Energy Laboratory (NREL) August 2004.

79. Cheng KK, Zhao XB, Zeng J, et al. Appl Microbiol Biotechnol. 2012;95:841–850.

80. Huang HJ, Ramaswamy S, Tschirner UW, Ramarao BV. Sep Purif Technol. 2008;62:1–21.

81. Huh YS, Jun YS, Hong YK, Song H, Lee SY, Hong WH. Process Biochemistry. 2006;41:1461–1465.

82. Corma A, Renz M, Susarte M. Topics in Catalysis. 2009;52:1182–1189.

83. Climent MJ, Corma A, Iborra S. Green Chemistry. 2011;13:520–540.

84. Clark JH. Journal of Chemical Technology and Biotechnology. 2007;82:603–609.

85. Schlaf M, Dalton T. In: 2006;4645–4653.

86. Rinaldi R, Schueth F. Energy & Environmental Science. 2009;2:610–626.

87. Dacquin JP, Cross HE, Brown DR, et al. Green Chemistry. 2010;12:1383–1391.

88. Li WC, Lu AH, Schuth F. Chemistry of Materials. 2005;17:3620–3626.

89. Climent MJ, Corma A, Iborra S. Chemical Reviews. 2011;111:1072–1133.

90. Zhang B, Ren J, Liu X, et al. Catalysis Communications. 2010;11:629–632.

91. Clark JH, Budarin V, Dugmore T, Luque R, Macquarrie DJ, Strelko V. Catalysis Communications. 2008;9:1709–1714.

92. Bechthold I, Bretz K, Kabasci S, Kopitzky R, Springer A. Chemical Engineering & Technology. 2008;31:647–654.

93. Delhomme C, Weuster-Botz D, Kuehn FE. Green Chemistry. 2009;11:13–26.

94. Mohan D, Pittman Jr CU, Steele PH. Energy & Fuels. 2006;20:848–889.

95. Catoire L, Yahyaoui M, Osmont A, Goekalp I. Energy & Fuels. 2008;22:4265–4273.

96. van Putten R-J, van der Waal JC, de Jong E, Rasrendra CB, Heeres HJ, de Vries JG. Chemical Reviews. 2013;113:1499–1597.

97. Zakrzewska ME, Bogel-Łukasik E, Bogel-Łukasik R. Chemical Reviews. 2010;111:397–417.

98. Hu L, Zhao G, Hao WW, et al. RSC Adv. 2012;2:11184–11206.

99. Dutta S, De S, Saha B. ChemPlusChem. 2012;77:259–272.

100. Buntara T, Noel S, Phua PH, Melián-Cabrera I, de Vries JG, Heeres HJ. Angewandte Chemie International Edition. 2011;50:7083–7087.

101. Roman-Leshkov Y, Barrett CJ, Liu ZY, Dumesic JA. Nature. 2007;447:982–985.

102. Yang G, Pidko EA, Hensen EJM. J Catal. 2012;295:122–132.

103. Patil SKR, Lund CRF. Energy & Fuels. 2011;25:4745–4755.

104. Watanabe M, Aizawa Y, Iida T, Nishimura R, Inomata H. Applied Catalysis A – General. 2005;295:150–156.

105. Schraufnagel RA, Rase HF. Industrial & Engineering Chemistry Product Research and Development. 1975;14:40–44.

106. Lourvanij K, Rorrer GL. Industrial & Engineering Chemistry Research. 1993;32:11–19.

107. Lourvanij K, Rorrer GL. Journal of Chemical Technology and Biotechnology. 1997;69:35–44.

108. Nakamura Y, Morikawa S. Bulletin of the Chemical Society of Japan. 1980;53:3705–3706.

109. Moreau C, Durand R, Peyron D, Duhamet J, Rivalier P. Industrial Crops and Products. 1998;7:95–99.

110. Moreau C, Durand R, Alies FR, Cotillon M, Frutz T, Theoleyre MA. Industrial Crops and Products. 2000;11:237–242.

111. Kourieh R, Rakic V, Bennici S, Auroux A. Catalysis Communications. 2013;30:5–13.

112. Shimizu K-I, Furukawa H, Kobayashi N, Itaya Y, Satsuma A. Green Chemistry. 2009;11:1627–1632.

113. Carlini C, Giuttari M, Galletti AMR, Sbrana G, Armaroli T, Busca G. Applied Catalysis A – General. 1999;183:295–302.

114. Armaroli T, Busca G, Carlini C, Giuttari M, Galletti AMR, Sbrana G. Journal of Molecular Catalysis A – Chemical. 2000;151:233–243.

115. Qi X, Watanabe M, Aida TM, Smith Jr RL. Catalysis Communications. 2009;10:1771–1775.

116. Crisci AJ, Tucker MH, Lee M-Y, Jang SG, Dumesic JA, Scott SL. ACS Catalysis. 2011;1:719–728.

117. Qi X, Watanabe M, Aida TM, Smith Jr RL. Green Chemistry. 2009;11:1327–1331.

118. Qi X, Watanabe M, Aida TM, Smith Jr RL. Chemsuschem. 2009;2:944–946.

119. Lansalot-Matras C, Moreau C. Catalysis Communications. 2003;4:517–520.

120. Jadhav H, Taarning E, Pedersen CM, Bols M. Tetrahedron Letters. 2012;53:983–985.

121. Guo X, Cao Q, Jiang Y, Guan J, Wang X, Mu X. Carbohydrate Research. 2012;351:35–41.

122. Takagaki A, Ohara M, Nishimura S, Ebitani K. In: Chemical Communications. 2009;6276–6278.

123. Ohara M, Takagaki A, Nishimura S, Ebitani K. Applied Catalysis A – General. 2010;383:149–155.

124. Weingarten R, Tompsett GA, Conner WC, Huber GW. J Catal. 2011;279:174–182.

125. Moliner M, Roman-Leshkov Y, Davis ME. Proc Natl Acad Sci U S A. 2010;107:6164–6168.

126. Roman-Leshkov Y, Moliner M, Labinger JA, Davis ME. Angewandte Chemie. 2010;49:8954–8957.

127. Pagán-Torres YJ, Wang T, Gallo JMR, Shanks BH, Dumesic JA. ACS Catalysis. 2012;2:930–934.

128. Nikolla E, Roman-Leshkov Y, Moliner M, Davis ME. ACS Catalysis. 2011;1:408–410.

129. Pagán-Torres YJ, Wang TF, Gallo JMR, Shanks BH, Dumesic JA. ACS Catalysis. 2012;2:930–934.

130. Nakajima K, Baba Y, Noma R, et al. Journal of the American Chemical Society. 2011;133:4224–4227.

131. Dias AS, Lima S, Pillinger M, Valente AA. Carbohydrate Research. 2006;341:2946–2953.

132. Dias AS, Pillinger M, Valente AA. Applied Catalysis A – General. 2005;285:126–131.

133. Jeong GH, Kim EG, Kim SB, Park ED, Kim SW. Microporous and Mesoporous Materials. 2011;144:134–139.

134. Dias AS, Pillinger M, Valente AA. J Catal. 2005;229:414–423.

135. Lam E, Majid E, Leung ACW, Chong JH, Mahmoud KA, Luong JHT. Chemsuschem. 2011;4:535–541.

136. Dias AS, Lima S, Brandao P, Pillinger M, Rocha J, Valente AA. Catalysis Letters. 2006;108:179–186.

137. Dias AS, Lima S, Pillinger M, Valente AA. Catalysis Letters. 2007;114:151–160.

138. Suzuki T, Yokoi T, Otomo R, Kondo JN, Tatsumi T. Applied Catalysis A – General. 2011;408:117–124.

139. Choudhary H, Nishimura S, Ebitani K. Chemistry Letters. 2012;41:409–411.

140. Guo HJ, Yin GC. Journal of Physical Chemistry C. 2011;115:17516–17522.

141. Leonard RH. Industrial and Engineering Chemistry. 1956;48:1331–1341.

142. Bart HJ, Reidetschlager J, Schatka K, Lehmann A. Industrial & Engineering Chemistry Research. 1994;33:21–25.

143. Li Z, Wnetrzak R, Kwapinski W, Leahy JJ. ACS Applied Materials & Interfaces. 2012;4:4499–4505.

144. Fernandes DR, Rocha AS, Mai EF, Mota CJA, Teixeira da Silva V. Applied Catalysis A – General. 2012;425:199–204.

145. Pasquale G, Vazquez P, Romanelli G, Baronetti G. Catalysis Communications. 2012;18:115–120.

146. Dharne S, Bokade VV. Journal of Natural Gas Chemistry. 2011;20:18–24.

147. Weingarten R, Conner Jr WC, Huber GW. Energy & Environmental Science. 2012;5:7559–7574.

148. Fan YX, Zhou CH, Zhu XH. Catal Rev.-Sci Eng. 2009;51:293–324.

149. Auras R, Lim L-T, Selke SEM, Tsuji H. Polylactic Acid, Synthess Structures Properties Processing and Application New York: Wiley; 2010.

150. Hang YD. Biotechnology Letters. 1989;11:299–300.

151. John RP, Nampoothiri KM, Pandley A. Journal of Basic Microbiology. 2007;47:25–30.

152. Wasewar KL, Yawalkar AA, Moulijn JA, Pangarkar VG. Industrial & Engineering Chemistry Research. 2004;43:5969–5982.

153. Miura S, Dwiarti L, Arimura T, Hoshino M, Tiejun L, Okabe M. Journal of Bioscience and Bioengineering. 2004;97:19–23.

154. Onda A, Ochi T, Kajiyoshi K, Yanagisawa K. Applied Catalysis A – General. 2008;343:49–54.

155. Holm MS, Saravanamurugan S, Taarning E. Science. 2010;328:602–605.

156. Roman-Leshkov Y, Davis ME. ACS Catalysis. 2011;1:1566–1580.

157. Liu Z, Feng G, Pan CY, et al. Chin J Catal. 2012;33:1696–1705.

158. de Clippel F, Dusselier M, Van Rompaey R, et al. Journal of the American Chemical Society. 2012;134:10089–10101.

159. Onda A, Ochi T, Kajiyoshi K, Yanagisawa K. Catalysis Communications. 2008;9:1050–1053.

160. NEXANT Inc., CHEMSYSTEMS® Prospectus Bio-Based Chemicals: Going Commercial. 2012.

161. Erickson B, Nelson JE, Winters P. Biotechnology Journal. 2012;7:176–185.

162. Naughton FC. Journal of the American Oil Chemists Society. 1974;51:65–71.

163. Bozell JJ, Patel MK. Feedstocks for the Future, Renwables for the Production of Chemicals & Materials. Washington, DC: American Chemical Society; 2004.

164. Draths KM, Frost JW. Journal of the American Chemical Society. 1994;116:399–400.

165. de Guzman D. Green Chemicals: DSM adds adipic acid to bio-based chemicals portfolio. Available at: http://www.icis.com/Articles/2011/10/10/9498186/Green-Chemicals-DSM-adds-adipic-acid-to-bio-based-chemicals.html; (accessed March 2013).

166. Lin Y-C, Huber GW. Energy & Environmental Science. 2009;2:68–80.

167. Dacquin J-P, Lee AF, Wilson K. Heterogeneous Catalysts for Converting Renewable Feedstocks to Fuels and Chemicals New York: Springer; 2012.

168. Davidson A. Current Opinion in Colloid & Interface Science. 2002;7:92–106.

169. Galarneau A, Iapichella J, Bonhomme K, et al. Advanced Functional Materials. 2006;16:1657–1667.

170. Linssen T, Cassiers K, Cool P, Vansant EF. Advances in Colloid and Interface Science. 2003;103:121–147.

171. Ying JY, Mehnert CP, Wong MS. Angew Chem.-Int Edit. 1999;38:56–77.

172. Parlett CMA, Wilson K, Lee AF. Chemical Society Reviews. 2013;42:3876–3893.

173. Dacquin J-P, Dhainaut J, Duprez D, Royer S, Lee AF, Wilson K. Journal of the American Chemical Society. 2009;131:12896–12897.

174. Dhainaut J, Dacquin J-P, Lee AF, Wilson K. Green Chemistry. 2010;12:296–303.

175. Torres AI, Daoutidis P, Tsapatsis M. Energy & Environmental Science. 2010;3:1560–1572.

176. Crossley S, Faria J, Shen M, Resasco DE. Science. 2010;327:68–72.

177. Ruiz MP, Faria J, Shen M, Drexler S, Prasomsri T, Resasco DE. Chemsuschem. 2011;4:964–974.

178. Dapsens PY, Mondelli C, Perez-Ramirez J. ACS Catalysis. 2012;2:1487–1499.

179. Sanders JPM, Clark JH, Harmsen GJ, et al. Chem Eng Process. 2012;51:117–136.

180. Cheng Y, Feng Y, Ren Y, et al. Bioresource Technology. 2012;113:65–72.

181. Kulkarni AA, Zeyer K-P, Jacobs T, Kienle A. Industrial & Engineering Chemistry Research. 2007;46:5271–5277.

182. de la Iglesia O, Mallada R, Menendez M, Coronas J. Chemical Engineering Journal. 2007;131:35–39.

183. Assabumrungrat S, Kiatkittipong W, Praserthdam P, Goto S. Catalysis Today. 2003;79:249–257.

184. Buchaly C, Kreis P, Gorak A. Industrial & Engineering Chemistry Research. 2012;51:896–904.

185. Lai IK, Liu Y-C, Yu C-C, Lee M-J, Huang H-P. Chemical Engineering and Processing. 2008;47:1831–1843.

186. Singh A, Hiwale R, Mahajani SM, Gudi RD, Gangadwala J, Kienle A. Industrial & Engineering Chemistry Research. 2005;44:3042–3052.

187. Ni X, Mackley MR, Harvey AP, Stonestreet P, Baird MHI, Rao NVR. Chemical Engineering Research & Design. 2003;81:373–383.

188. Harvey AP, Mackley MR, Stonestreet P. Industrial & Engineering Chemistry Research. 2001;40:5371–5377.

189. Fabiyi ME, Skelton RL. Process Safety and Environmental Protection. 2000;78:399–404.

190. Eze V, Phan AN, Pirez C, Harvey AP, Lee AF, Wilson K. Catalysis Science & Technology. 2013;3:2373–2379.

191. Huang H-J, Ramaswamy S, Tschirner UW, Ramarao BV. Sep Purif Technol. 2008;62:1–21.

192. Pinzi S, Garcia IL, Lopez-Gimenez FJ, Luque de Castro MD, Dorado G, Dorado MP. Energy & Fuels. 2009;23:2325–2341.

193. Hick SM, Griebel C, Restrepo DT, et al. Green Chemistry. 2010;12:468–474.

194. Binder JB, Raines RT. Journal of the American Chemical Society. 2009;131:1979–1985.

195. Zhao H, Holladay JE, Brown H, Zhang ZC. Science. 2007;316:1597–1600.

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

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