Anda di halaman 1dari 30

Chem Soc Rev

Cite this: Chem. Soc. Rev., 2011, 40, 55885617 www.rsc.org/csr

Dynamic Article Links

View Online

CRITICAL REVIEW

Catalytic conversion of lignocellulosic biomass to ne chemicals and fuels


Chun-Hui Zhou,*a Xi Xia,a Chun-Xiang Lin,b Dong-Shen Tonga and Jorge Beltraminib
Received 5th May 2011 DOI: 10.1039/c1cs15124j
Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Lignocellulosic biomass is the most abundant and bio-renewable resource with great potential for sustainable production of chemicals and fuels. This critical review provides insights into the state-of the-art accomplishments in the chemocatalytic technologies to generate fuels and value-added chemicals from lignocellulosic biomass, with an emphasis on its major component, cellulose. Catalytic hydrolysis, solvolysis, liquefaction, pyrolysis, gasication, hydrogenolysis and hydrogenation are the major processes presently studied. Regarding catalytic hydrolysis, the acid catalysts cover inorganic or organic acids and various solid acids such as sulfonated carbon, zeolites, heteropolyacids and oxides. Liquefaction and fast pyrolysis of cellulose are primarily conducted over catalysts with proper acidity/basicity. Gasication is typically conducted over supported noble metal catalysts. Reaction conditions, solvents and catalysts are the prime factors that aect the yield and composition of the target products. Most of processes yield a complex mixture, leading to problematic upgrading and separation. An emerging technique is to integrate hydrolysis, liquefaction or pyrolysis with hydrogenation over multifunctional solid catalysts to convert lignocellulosic biomass to value-added ne chemicals and bio-hydrocarbon fuels. And the promising catalysts might be supported transition metal catalysts and zeolite-related materials. There still exist technological barriers that need to be overcome (229 references).
a

Research Group for Advanced Materials & Sustainable Catalysis (AMSC), Breeding Base of State Key Laboratory of Green Chemistry Synthesis Technology, College of Chemical Engineering and Materials Science, Zhejiang University of Technology, Hangzhou, 310032, China. E-mail: catalysis8@yahoo.com.cn, chc.zhou@yahoo.com.cn b The School of Chemical Engineering, ARC Centre of Excellence for Functional Nanomaterials, AIBN, The University of Queensland, St. Lucia, QLD 4072, Australia

Introduction

The depletion of fossil fuel resources and the resulting adverse eects on the global environment and climate are of major academic, economic and political concern worldwide.13 As supplies of fossil fuels and related petrochemicals may soon be limited, alternative solutions are sought. One alternative is to

Dr Chun-Hui (Clayton) Zhou is a Professor of Chemical Engineering and Group Leader of the Research Group for Advanced Materials and Sustainable Catalysis (AMSC) at Zhejiang University of Technology as well as a Member of the Editorial Board of Applied Clay Science (Elsevier). He worked as a visiting Academic at the ARC Center of Excellence for Functional Nanomaterials, AIBN, the University of Queensland Chun-Hui Zhou in 20062007 and as Visiting Professor at the Centre for Strategic Nano-fabrication, the University of Western Australia in 2010. His research interests include clay-based materials for miscellaneous applications, sustainable catalysis for fuels and chemicals from biorenewable resources, process intensication and advanced biomaterials.
5588 Chem. Soc. Rev., 2011, 40, 55885617

Xi Xia has been a postgraduate under the supervision of Prof. Chun-Hui Zhou in the Research Group for Advanced Materials & Sustainable Catalysis, Zhejiang University of Technology (ZJUT) since September, 2008. He obtained his Bachelors Degree in Chemistry from Nanyang Normal University in 2008 and his Masters degree of Industrial Catalysis at ZJUT in 2011. His research interests are in advanced clay-based Xi Xia materials, catalysts and their potential for the catalytic conversion of biomass into liquid fuels and chemicals.

This journal is

The Royal Society of Chemistry 2011

View Online

develop a series of novel chemical processes based on renewable feedstocks, typically biomass4 and biomass-derived chemicals.57 Biomass generally refers to organic materials such as wood, grass, algae, agricultural crops and their residues and wastes, including some animal waste.8 All of these materials originally result from biological photosynthesis from readily available atmospheric CO2, water and sunlight. Therefore, biomass is a sustainable and green feedstock for the production of fuels and ne chemicals with net zero carbon emission. Biochemically, living systems produce a large number of organic molecules such as carbohydrates, fats and oils, phospholipids and glycolipids, waxes, steroids, proteins and nucleotides. Among them, carbohydrates are the primary energy-storage molecules. Carbohydrates are the main structural components of the cell walls of plants, in which the carbohydrate polymers, namely cellulose and hemicellulose, are tightly bound to the lignin.9 Any materials rich in cellulose, hemicelluloses, and lignin are commonly referred to as lignocellulosic biomass.10 For example, wood, bamboo, grass and their derived pulp and paper, and agricultural residues like corn stover and sugarcane bagasse are typical sources of lignocellulosic biomass. Lignocellulosic biomass has received intensive attention over the past decade. Many studies have shown that lignocellulosic biomass oers great potential to be used as a renewable feedstock in abundance with easy availability.1114 Fig. 1 displays the weight percentage of cellulose, hemicellulouse and lignin in some typical biological sources such as pinaster,15 eucalyptus globules,15 wheat straw,16 sorghum stalks,17 bamboo18 and banana pseudo-stems.19 Generally, most of the lignocellulosic biomass contains 3550% of cellulose, 2035% of hemi-cellulose, and 1025% of lignin. Clearly, the major component of lignocellulosic biomass is cellulose. It is estimated that nearly half of the organic carbon in the biosphere is present in the form of cellulose. Therefore, in the context of processing lignocellulosic biomass into fuels and valuable chemicals, the conversion of cellulose is of paramount importance and worth being given priority.1114 The development of the processes to convert lignocellulosic biomass to fuels and value-added chemicals, however, remains a big challenge. Firstly, the complex chemical composition of lignocellulosic biomass makes it dicult to yield target fuels and chemicals in a high yield and quality. Comparatively, as to the three main components in lignocellulosic biomass, the structure of cellulose is the simplest and most ordered. Cellulose consists of only anhydrous glucose units, while hemicelluloses contain many dierent sugar monomers. Consequently, cellulose is crystalline and hemicellulose just has a random, amorphous structure. Dierent from carbohydratebased cellulose and hemicellulose, lignin is a class of complex, cross-linked, three-dimensional biopolymers with phenylpropane units with relative hydrophobic and aromatic properties. Those dierences in the chemical composition and structure in lignin, hemicellulose and cellulose lead to their much dierent reactivity. On one hand, for example, the decomposition of lignin and hemicellulose is generally easier than that of cellulose.20,21 In other words, the decomposition and degradation of cellulose is the most dicult and energy-consuming step in the conversion of lignocellulosic biomass. As a result, to improving the reactivity of
This journal is
c

cellulose is a rst critical point. On the other hand, based on their dierent reactivity, it is feasible to separate cellulose and hemicellulose from lignin.22,23 Therefore, as reported in the literature and discussed below, besides using the natural lignocellulosic biomass as the reactants, many researchers often choose pure cellulose, in particular microcrystalline cellulose, as a model of lignocellulosic biomass. Over decades, many researchers have endeavored to explore the production of fuels and chemicals from cellulose, which is representative of lignocellulosic biomass.24,25 A cellulose molecule has the generic chemical formula (C6H12O5)n. It consists of a skeletal linear polysaccharide, in which glucosebased monomer units are jointed together through b-1,4-glycosidic linkages.26 The glucose units are further tightly bound by extensive intramolecular and intermolecular hydrogen bonding networks (Scheme 1a). The chain length of a cellulose molecule ranges from about 100 to 14 000 units. Accordingly, cellulose has an average molecular weight in the range of 3 00 0005 00 000.27 Noticeably, cellulose exists in the form of a robust crystalline structure. And it is the principle scaolding component of all plant cell walls. Such a cell wall polymer is neither soluble in water nor easily digestible in the gastrointestinal tract of humans. The changes in the types of chemical bonding and the crystalline structure in polysaccharides, however, have signicant impacts on their physical properties and chemical reactivity.23 For example, starch is also a polysaccharide with the same general formula (C6H12O5)n as cellulose. Nevertheless, in each starch(amylose) molecule the neighboring glucose units are linked through the 1,4 0 -a-glycosidic bond (Scheme 1b) or the 1,6 0 -a-glycosidic bond (amylopectin), rather than the 1,4 0 -b-glycosidic bond in each cellulose molecule.28 This dierence has a profound eect on the three dimensional structure and reactivity of the biopolymer molecule. Such 1,40 -a-glycosidic linkages in starch are more easily attacked by acids or enzymes. Thereby in the presence of acids or enzymes as catalysts starch is more easily deconstructed into glucose monomers than cellulose. For instance, starch is more easily fermented to yield ethanol than cellulose. This is one of the major reasons that starches are currently used as the starting materials for commercial bioethanol production, instead of cellulose.29,30 Nonetheless, any large-scale processing of starch to fuels and chemicals readily brings about strong competition between chemical production and food demands.31 Therefore, in terms of both the abundance

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Fig. 1 The weight percentage of cellulose, hemicelluloses and lignin in some typical biological sources: pine pinaster,15 eucalyptus globules,15 wheat straw,16 sorghum stalks,17 bamboo18 and banana pseudo-stems.19

The Royal Society of Chemistry 2011

Chem. Soc. Rev., 2011, 40, 55885617

5589

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 1 Comparison of the chemical structure of polysaccharides. (a) cellulose with 1,4 0 -b-glycosidic bonds and intra- and inter-chain hydrogen bonding,27 Adapted and reprinted with permission from ref. 27. Copyright 2006 American Chemical Society; and (b) starch (amylose) with 1,4 0 -a-glycosidic bonds.28 Adapted and reprinted from ref. 28, Copyright 2003, with permission from Elsevier.

and availability from bioresources and little impact on food supply, cellulosic materials as feedstocks are more practicable and sustainable for the production of fuels and commodity chemicals. In this context, the catalytic conversion of cellulose to fuels and chemicals is more attractive and promising than that of any other form of biomass like starch to bioethanol and vegetable oil to biodiesel.3235 Still, lignocellulosic biomass such as wood, grass and agricultural residues are conventionally used as energy sources typically by direct combustion.36 The principal utilization of cellulose in industry is currently limited to textiles and paper manufacturing.37 In some biological processes such as fermentation and enzymic catalysis, peculiar enzymes, bacteria and other microorganisms are used to break down cellulose molecules and thus a few commodity chemicals can be obtained.38,39 Nevertheless, such biological processes generally suer from unsolvable problems such as low eciencies, narrow reaction conditions and limited scale of production.39,40 Chemocatalytic conversion of cellulose has been around for some time, but it only receives serious attention with the advent of a series of novel chemocatalytic reaction routes since the fossil oil crisis in the 1970s. However, besides the complexity of biological sources, the inert chemical structure and the compositional ratio of carbon, hydrogen and oxygen in molecules in lignocellulosic biomass produce additional diculties in the chemocatalytic
5590 Chem. Soc. Rev., 2011, 40, 55885617

conversion of cellulose to valuable fuels and commodity chemicals. In this context, the development of a new family of highly active and selective catalyst systems is an essential prerequisite for chemoselectively catalytic conversion of lignocellulosic biomass to desired products.41 As such, this article attempts to summarize and review recent advancements in the dierent catalytic processes for the conversion of lignocellulosic biomass, in particular cellulose, into potential fuels and commodity chemicals. Scheme 2 lists some typical ne chemicals and fuels which can be produced by chemocatalytic conversion of cellulose by dierent chemical processes. Clearly, a variety of fuels, including ethanol, hydrogen, methane and chemicals such as glucose, fructose, sorbitol, levulinic acid and lactic acid can be obtained from catalytic conversion of lignocellulosic biomass. Lignocellulosic biomass can also be used to produce syngas (CO + H2) which can then be transformed into fuels and myriad chemicals. In many instances, depolymerization and hydrolysis of cellulose to monomer glucose is regarded a necessary rst step. Then glucose is further catalytically degraded into various intermediates, chemicals and fuels (Scheme 2, Route A). Recent research has paid much heed to accelerating the conversion of lignocellulosic biomass in the presence of multifunctional catalysts. The typical processes can be categorized as direct liquefaction, gasication and aqueous-phase reforming,
This journal is
c

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 2 Potential chemicals and fuels from the catalytic conversion of cellulose. (Route A: chemocatalytic conversion of cellulose into various chemicals and fuels through glucose, Route B: Conversion of cellulose into various chemicals and fuels by a one-pass catalytic process in the presence of a multifunctional catalyst).

pyrolysis. An emerging trend is to integrate catalytically thermochemical conversion with selectively catalytic hydrogenolysis and hydrogenation with the addition of hydrogen.4244 Because of the advantages of easy recovery and reusability of solid catalysts, heterogeneous processes have been taking priority over homogeneous catalytic processes. In addition, increasing endeavors are being made to integrate the two- or multi-step batch processes into a one-pass continuous conversion by using well-designed multifunctional catalysts with an expectation to transform lignocellulosic biomass into various commodity chemicals in an ecient and environmentally friendly way (Scheme 2, Route B).45

obtained. Interests in such acid-catalyzed hydrolysis have been reignited because many ne and even platform chemicals can be obtained from further conversion of reducing sugars. Moreover, using a multistep fractionation process, various ne chemicals and potential fuels can also be directly obtained from acid processing of lignocellulosic biomass.46 Nevertheless, instead of the traditional use of liquid acids, nowadays much attention is given to the use of solid catalysts such as solid-supported Brnsted or Lewis acid reagent catalysts, acidmodied amorphous carbon, layered transition metal oxides, and acidic resins for the depolymerization of lignocellulosic biomass by hydrolysis (Tables 1 and 2). 2.1 Liquid acid-catalyzed hydrolysis Acid-catalyzed hydrolysis of cellulose is a reaction in which hydrogen cations (H+) and hydroxide anions (OH) from the splitting of molecules of water react with polymeric cellulose molecules, thereby yielding hydrolytic products by depolymerization. The discovery of such an acid-catalyzed reaction of cellulose hydrolysis has a relatively long history.5577 Commercial microcrystalline cellulose is just a slightly depolymerized
Chem. Soc. Rev., 2011, 40, 55885617 5591

Catalytic hydrolysis of lignocellulosic biomass

Lignocellulosic biomass with cellulose as a main component is generally resistant to hydrolysis in water. As such, the hydrolysis of cellulose in lignocellulosic biomass usually involves the use of strong liquid acids as catalysts. Cellulose hydrolysis in the presence of acids primarily yields reducing sugars, which are a class of compounds that have an open chain with an aldehyde or a ketone group. Typically, glucose and xylose are
This journal is
c

The Royal Society of Chemistry 2011

View Online

Table 1 Catalyst

Typical hydrolysis of lignocellulosic biomass or pure cellulose with H2O over liquid acid catalysts Feedstock Wood chips Filter paper Cotton Corn cobs Sugar maple wood extract Reaction conditions 483 K, 2 min B1.9 MPa 488 K, 35 min 4 MPa (H2O) 473 K, 6 min 0.1 MPa 398 K, 165 min 368 K, 50 min 0.1MPa Typical products Xylose Reducing sugar Levoglucosenone Furfural 5-HMF Xylose Xylose Glucose Rhamnose Mannose Galactose Glucose Xylose Levulinic acid Glucose HMF Glucose Reducing sugar Reducing sugar Glucose Yield (%) 47.054.0 46.6 42.2 26.9 8.8 25.0 161.6 g L 13.3 g L1 12.4 g L1 12.9 g L1 5.6 g L1 37.0 (wt%) 10.3 (wt%) 12.0 B90.0 57.0 B11.0 35.0 82.5 2.83.9
1

Researchers/Year Emmel et al./2003 Zhuang et al./2006 Kawamoto et al./2007 Rivas et al./2008 Hu et al./2010

Ref. 57 53 144 55 56

H2SO4 (0.175 wt%) H2SO4 (0.05 wt%) H2SO4 (0.1 wt%) H2SO4 (0.5 wt%) H2SO4 (6.2 wt%)

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

SO2 impregnation HCl (5 wt%) HCl (20 wt%)/ + [EMIM]Cl CO2 + H2O (100% saturation) p-toluenesulfonic acid + C4MIMCl H3PO4 (10 wt%) Oxalic(0.1M) + NaCl

Aspen chips Walnut shells Cellulose Cellulose Cellulose Potato peel Cellulose

478 K, 3 min 1.62 MPa (H2O) 523 K, 430 min 1.58.6 MPa (H2O) 378 K, 24 h 533 K, B2 min 2025 MPa CO2, 373 K, 5 h 408 K, 8 min 403 K, 6 h 3 MPa CO2

De Bari et al./2007 Liu et al./2006 Binder et al./2010 Rogalinski et al./2008 Rinaldi et al./2008 Lenihan et al./2010 Vom Stein et al./2010

58 59 68 64 90 60 67

5-HMF: 5-hydroxymethyl-2-furfural. Reducing sugar: mainly mono- and disaccharides with an aldehyde or a ketone group. [EMIM]Cl: 1-ethyl-3methylimidazolium chloride. C4MIMCl: 1-butyl-3-methylimidazolium chloride.

product of cellulose by liquid acid-catalyzed hydrolysis.47,48 More importantly, various products such as glucose, xylose, arabinose and cellobiose and oligosaccharides can be readily formed from liquid acid-catalyzed hydrolysis of cellulose (Scheme 3). Both inorganic liquid mineral acids such as H2SO4 and HCl, and organic acids such as various carboxylic acids and p-toluenesulfonic acid can be used as catalysts for the hydrolysis of cellulose. Various types of lignocellulosic biomass, for example wood chips, sawdust, corn-cobs, and walnut shells, have been tentatively processed by the liquid acid-catalyzed hydrolysis (Table 1). However, it should be pointed out that the hydrolysis mainly occurred in the cellulosic components in these lignocellulosic sources with glucose and xylose as the main products in most cases. As summarized in Table 1, thus far H2SO4 has been the most frequently used liquid acid catalyst in the hydrolysis of lignocellulosic biomass. The reaction is usually conducted at temperatures in the range of 363533 K and under atmospheric or higher pressure. In general, when a dilute acid solution is used, higher temperature and longer reaction time are necessary.49 In turn, when the hydrolysis reaction of cellulose is conducted at low temperature and in a short reaction time, concentrated acid is favorable. The treatment of cellulose in a concentrated acid solution remarkably accelerates such biopolymer molecules to be depolymerized to monomers.50,51 Usually, the yield of reducing sugars increases with increasing acid concentration. However, the use of concentrated acid solutions could lead to a decrease of the yield of glucose and xylose due to further degradation.
5592 Chem. Soc. Rev., 2011, 40, 55885617

Consequently, there is a substantial increase in the amount of products of further degradation of glucose or xylose, for example 5-hydroxymethyl-2-furfural (5-HMF).52 In addition, taking into account the operation safety, equipment corrosion and waste disposal, dilute acid is preferable for use in the hydrolysis of cellulose.53 Furthermore, as for catalytic performances in the hydrolysis of lignocellulosic biomass with a high content of hemicelluloses, dilute acid could be more eective than concentrated acid.54 The deep insights into the eect of pressure on the conversion and distribution of products are scarcely reported. However, it has been revealed that under atmospheric pressure, catalytic hydrolysis of cellulose is also feasible with both hydrolytic and dehydrated products.56 Clearly, under atmospheric pressure, higher temperature, concentrated acid and longer residence time are favorable to eectively depolymerize and degrade lignocellulosic biomass. The impregnation of lignocellulosic biomass with acid solution can lead to partial destruction of the cellulosic parts in lignocellulosic biomass. Thereby such a pretreatment can facilitate their further hydrolytic reaction. For example, Emmel and co-workers57 reported that the pretreatment of wood chips by impregnating them in 0.175 wt% H2SO4 led to the remarkable hydrolysis of hemicellulose after the dilute acid-catalyzed hydrolysis at 483 K for 2 min in steam. Thus a yield of xylose up to 70% of its theoretical yield was obtained in the water-soluble fraction. In this case, the pretreatment temperature was found to inuence the yield of xylose. Interestingly, impregnation with gaseous SO2 proved to be more eective than H2SO4 since a higher yield of
This journal is
c

The Royal Society of Chemistry 2011

View Online

Table 2 Catalyst

Typical hydrolysis of cellulose with H2O over solid acid catalysts Feedstock Cellulose Cellulose Cellulose Cellulose Cellulose Cellulose Cellulose Cellulose +CH3OH Cellulose Cellulose Cellouse Reaction conditions 373 K, 3 h 373 K, 3 h 423 K, 24 h 423 K, 24 h MW (350 W), 373 K, 1 h 403 K, 12 h 453 K, 2 h r473 K, 0.5 h 3 MPa 433 K,10 min 463 K, 24 h 423 K, 24 h Typical products Glucose b-1,4-glucan Glucose b-1,4-glucan Glucose Glucose Glucose Glucose Glucose Levulinic acid HMF Methyl Glucosides 5-HMF Glucose Levulinic acid Glucose Yield (%) 4.0 64.0 B8.1b 0.93 40.5 (C-%) 61.0 19.8 21.0 50.5 0.14 1.6 50.060.0 62.3 9.0 2.0 50.0
a

Researchers/Year Suganuma et al./2008 Yamaguchi et al./2009 Onda et al./2008 Pang et al./2010 Wu et al./2010 Takagaki et al./2008 Tian et al./2010 Deng et al./2010 Wang et al./2011 Hegner et al./2010 Van de Vyver et al./2010

Ref. 82 78 79 83 81 88 86 94 89 76 73

ACSO3H + COOH + OH ACSO3H + COOH + OH ACSO3H ACSO3H (Sulfonation T = 523 K) BCSO3H HNbMoO6

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

H3PW12O40 H3PW12O40 CrCl3/LiCl + [C4MIM]Cl Naon/silica Sulfonated silica/carbon nanocomposites

AC: Active carbon. BC: Biomass char. [C4MIM]Cl: 1-butyl-3-methylimidazolium chloride. MW: Microwave-assisted. a catalyst: 0.300 g; cellulose: 0.025 g; distilled water: 0.7 g. b catalyst: 0.300 g; cellulose: 0.25 g; distilled water: 2.25 g.

Scheme 3 Possible products from acid-catalyzed hydrolysis of cellulose.90 Adapted and reprinted with permission from ref. 90. Copyright 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

glucose could be obtained.58 It was suggested that the addition of SO2 catalyst (0.9% w/w dry raw material) reduced the degree of polymerization of the cellulose by an average of 50%. In addition to the catalytic performances, another noteworthy feature is that SO2 is less corrosive than the H2SO4 solution. In addition to reaction parameters such as temperature, pressure, time and the concentration of acids, the type of acid catalysts is also an inuential parameter for the distribution of products in the catalytic hydrolysis of lignocellulosic biomass. For example, in the presence of HCl acid in the hydrolysis of walnut shells, the yield of levulinic acid could reach 12%.59 This result might indicate that HCl acid as a catalyst could promote further catalytic dehydration of hydrolytic compounds, thereby resulting in the remarkable generation of levulinic acid. Dilute phosphoric acid is also an alternative catalyst.60 In a phosphate buer solution at pH 2, the hydrolysis and
This journal is
c

dissolution of Japanese red pine wood chips under subcritical water conditions yielded hydrolytic saccharides and such further dehydrated products as 5-HMF and furfural.61 It might indicate that the presence of phosphate buer could play a pivotal role in deep degradation of such a lignocellulosic biomass. However, in the absence of a catalyst, water-soluble saccharides can also be directly produced from the hydrolysis of lignocellulosic biomass, for example sugi wood powder, in supercritical and subcritical water.62 Despite the relatively high catalytic activity of H2SO4, HCl and H3PO4 in the hydrolysis of cellulosic materials, by and large their uses are still uneconomical because the process suers from severe corrosion, costly separation and neutralization of waste acids. These problems could be signicantly addressed provided that carbonic acid (H2CO3), present in the form of a solution of CO2 in water, can be used as a catalyst.63,64 First of all, carbonic acid has little corrosive eect. Secondly, theoretically, it can be
Chem. Soc. Rev., 2011, 40, 55885617 5593

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

easily removed by depressurization and thus it does not leave any waste acid catalysts after reaction. Therefore, such a carbonic acid-catalyzed process is environmentally friendly. Rogalinski et al.64 reported that for the hydrolysis of cellulose in subcritical water with pressurized CO2 at 533 K, the yield of glucose could be signicantly increased by the acidication of CO2 compared with that in pure water. Moreover, the onset of the formation of glucose is shifted to a shorter residence time due to the faster cleavage of 1,4 0 -b-glycosidic bonds facilitated by carbonic acid. As for the catalytic mechanism, CO2 is believed to promote the hydrolysis reaction of cellulose by the formation and then dissociation of carbonic acid in water to give active H+ cations, which then attack and break 1,40 -b-glycosidic bonds in cellulose (eqn (1)).63,64
+ CO2 + H2O ! H2CO3 ! H+ + HCO + CO2 3 ! 2H 3 (1)

after catalytic hydrolysis.69 In addition, even the size of biomass feedstock particles may also be one of the important parameters. These are also common points worth consideration in any other processes discussed below. 2.2 Solid acid-catalyzed hydrolysis Compared with homogeneous catalysts, heterogeneous solid catalysts have several advantages. In particular, solid catalysts can be easily separated from the liquid mixture after the reaction, thereby allowing the possible reuse of catalysts, along with the minimization of corrosion. Moreover, solid acid catalysts have the potential to be easily applied to a continuous ow xed-bed reactor. These merits have stimulated the research and development of recyclable solid acids as replacements for the unrecyclable liquid acid catalysts in the catalytic hydrolysis of lignocellulosic biomass.7072 As such, several types of solid acids, such as Naon, Amberlyst, SO3H functionalized amorphous carbon or mesoporous silica, H-form zeolites like HZSM-5, heteropolyacids and even metal oxides (for example, g-Al2O3) have been explored on their catalytic performances in the catalytic hydrolysis of lignocellulosic biomass. It has been shown that solid Brnsted acid catalysts are indeed applicable for the ecient hydrolysis of lignocellulosic biomass.74,75 Over solid acid catalysts, the typical hydrolytic products like glucose and xylose were also obtained, together with other chemicals such as b-1,4-glucan, furfural, 5-HMF, and levulinic acid. Furthermore, solid surface supported Lewis acid catalysts, for example FeCl3 supported onto amorphous silica, can also catalyze the hydrolysis of cellulose to glucose and then to levulinic acid even under relatively mild conditions.76 However, in most cases, over solid catalysts the catalytic activity and the selectivity to the desired products are quite low. Therefore, a longer reaction time is needed to acquire the high yield of glucose (Table 2), compared with liquid acid catalysts (Table 1). As shown in Scheme 4, a commonplace insight into the hydrolysis reaction is that it involves the splitting of the water molecule into hydrogen cations (H+) and hydroxide anions (OH). In the presence of cellulose, the H+ attacks the oxygen atom in the 1,4 0 -b-glycosidic linkage. Thus the 1,4 0 -bglycosidic linkage is broken to form a cyclic carbonium cation with a chair shape. This step is considered as the ratedetermining step.77 Finally, glucose is formed by rapid ion transfer of OH from the dissociation of water molecules to the glucose unit-based carbonium cation. According to such a reaction mechanism, that a solid acid catalyst and reaction conditions favor the splitting of water will favor the hydrolysis of cellulose. The splitting of a water molecule into hydrogen cations (H+) and hydroxide anions (OH) can be regarded as deprotonation. It was discovered that the catalytic activity of a catalyst in the hydrolysis of cellulose increases with a decrease in the deprotonation enthalpy (DPE) of water on the surface of the solid acid catalyst.85 These ndings suggested that stronger Brnsted acidity is more favorable to the catalytic hydrolysis of cellulose. In addition, to some extent, an increase in the amount of water had a positive eect on the breakage of 1,4 0 -b-glycosidic bonds and intramolecular hydrogen bonds in
This journal is
c

The avoidance of the use of strong liquid acids oers an alternative approach to develop a green acid-catalyzed process. However, the insolubility of cellulose is another big barrier that needs to be tackled. In this context, one possible solution is to choose a suitable solvent. For instance, Recently Lin et al.65 investigated the hydrolysis and degradation of cotton cellulose treated with 1.39% HCl acid either in water or in ethanol at 338 K for more than 1 h. A higher yield of hydrolytic products was obtained in ethanol (2.445.04%) than in water (1.121.50%). Moreover, the eect of the reaction temperature on the yield of hydrolytic products in ethanol was much greater than that in water. The phenomena could be ascribed to the fact that the solubilization of cellulose is enhanced by ethanol. This might well explain why cellulose was more easily deconstructed in ethanol than in water. The other alternative solution is to use a mixture of inorganic and organic acids as catalysts.66 It has been found that dicarboxylic acids, like oxalic and maleic acid, are able to catalytically depolymerize cellulose, producing oligomers and glucose.67 The third solution is to use ionic liquids as the solvent and/or as the catalyst. Recently Binder and Raines68 reported a much eective process, in which water was gradually added to a catalytic system consisting of chloride ionic liquid and acid catalysts (1-ethyl-3-methylimidazolium chloride, [EMIM]Cl + H2SO4 or HCl). This method led to a nearly 90% yield of glucose from cellulose and 7080% yield of sugars from corn stover in a few hours. In comparison to other conventional solvents and organic acids, however, the high cost of ionic liquids and the problematic reusability might still hinder its scale-up to commercial use. Finally, in the liquid acid-catalyzed hydrolysis, besides acidic catalysts, solvents and reaction variables, the source of lignocellulosic biomass is also an important factor. Obviously, the molecular composition and arrangements of cellulose, hemicellulose and lignin in the plant cell wall changes throughout the life of the plant. Moreover, they dier among plant species, among tissues of a single species and even among geographical regions of the plant. The dierences of chain length and hydrogen bonding pattern between amorphous hemicellulose and crystalline cellulose, for example, greatly aected the distribution of glucose and oligomers in the liquid products
5594 Chem. Soc. Rev., 2011, 40, 55885617

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 4

Schematic mechanism of breakage of 1,4 0 -b-glycosidic bonds and formation of glucose in the hydrolysis of cellulose.

insoluble cellulose, thereby leading to more hydrolytic products. Clearly, the formation of glucose requires more water than the formation of b-1,4-glucan. Yamaguchi and co-workers78 revealed that the amount of water comparable to the solid catalyst weight could lead to a maximum yield of glucose in the heterogeneous catalytic hydrolysis reaction of cellulose. On the one hand, the conditions conducive to splitting of water molecules are needed. On the other hand, the conditions should help to avoid deep degradation. Therefore, mildly hydrothermal conditions were found favorable to the production of glucose from the hydrolysis of cellulose over solid acid catalysts.78,79 Among a series of typical solid acid catalysts like H-form zeolite, sulfated zirconia, sulfonated activated carbon and Amberlyst polymer- based materials (Amberlyst, a commercially available sulfonated resin), as reported by Onda and co-workers,79 sulfonated activated carbon showed a remarkably high yield of 40.5% of glucose (Fig. 2). Over such sulfonated activated carbon catalysts, a selectivity of higher than 90% to glucose was obtained. The catalysts could act in a way similar to sulfonic acids containing the group SO3H, which has fairly strong acidity. The remarkably catalytic properties can be ascribed to the strong acidity of SO3H functional groups and the hydrophobic planes on the surfaces of such a catalyst, as well as its high hydrothermal stability. Such systematic evaluation of a series of catalysts provided valuable information and clues for the design and preparation of a class of eective solid acid catalyst for catalytic hydrolysis of cellulose by adjusting its surface acidity and hydrophilicity/ hydrophobicity. It is worth noting that lignocellulosic biomass itself can be made into carbon materials by carbonization.80 Bio-char or activated carbon from lignocellulosic biomass such as bamboo, wood, and coconut shells are well-documented catalyst supports. Microwave-assisted hydrolysis of cellulose in water over a sulfonated bio-char acid catalyst (BC-SO3H) even showed a much higher turnover number (TON, 1.331.73) compared to that in dilute H2SO4 solution (TON, 0.02).81 The results were likely due to the strong anity to 1,40 -b-glycosidic bonds of cellulose on the surfaces of such catalysts, thereby accelerating the activation of cellulose molecules. Microwave irradiation could strengthen the collision of cellulose particles in the reaction system. In addition, the surface of carbon solids can be functionalized with several functional groups. For example, Suganuma et al.82 described that a modied amorphous carbon
This journal is
c

Fig. 2 Catalytic hydrolysis of cellulose over various solid acid catalysts at 423 K. Reaction conditions: milled cellulose 45 mg, catalyst 50 mg, distilled water 5.0 mL, 24 h.79 Ref. 79Reproduced by permission of the PCCP Owner Societies.

material can simultaneously bear SO3H, COOH, and OH groups by means of partial carbonization of cellulose, followed by sulfonation of the resulting amorphous carbon (Fig. 3).80 This strategy of catalyst preparation could eciently enhance the performance of catalysts. Over such a catalyst, for example, when the hydrolysis of microcrystalline cellulose was conducted in a Pyrex reactor (catalyst = 0.300 g; cellulose = 0.025 g; distilled water = 0.700 g) at 373 K for 3 h, the yield of glucose and b-1,4-glucan reached 4% and 64%, respectively.82 The yield of glucose was increased to 8.08% under the optimized conditions (catalyst = 0.300 g; cellulose = 0.25 g; distilled water = 2.25 g).78 The enhanced catalytic performance is attributed to the good ability of such a material with a multi-functional surface to adsorb b-1,4-glucan. This view might be partly supported by the result that sulfonated silica/carbon nanocomposites exhibited high performance in the selective hydrolysis of cellulose into glucose.73 In this case, the hybrid surface structure constituted interpenetrated silica and carbon components which can greatly facilitate the adsorption of b-1,4-glucan on the solid catalyst. More recently, it was suggested that elevating the temperature of sulfonation during the preparation of sulfonated carbon catalysts enhanced their catalytic properties in the hydrolysis of cellulose.83 No matter what methods were used, in essence, the presence of strong, accessible
Chem. Soc. Rev., 2011, 40, 55885617 5595

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Fig. 3 Proposed schematic structure of the prepared carbon material bearing SO3H, COOH, and OH functions.82 Reprinted with permission from ref. 82. Copyright 2008 American Chemical Society.

Brnsted acid sites on catalysts played a crucial role in the catalytic hydrolysis of cellulose. In addition to the surface-grafted or supported solid acids, heteropolyacids such as tungstophosphoric acid H3PW12O408486 and HNbMoO687,88 can also be used as heterogeneous-like catalysts for the catalytic hydrolysis of cellulose. For the hydrolysis of cellulose samples, H3PW12O40 and Sn0.75PW12O40 catalysts gave a higher yield of total reducing sugars than H2SO4. A 50.5% yield of glucose with a selectivity higher than 90% were achieved over H3PW12O40 catalysts at 453 K for 2 h with a mass ratio of 0.42 of cellulose to H3PW12O4086 Over a series of catalysts like H3PW12O40, H4SiW12O40 and polyvalent transition metal salts of PW12O403 (Ag+, Ca2+, Co2+, Y3+, Sn4+, Sc3+, Ru3+, Fe3+, Hf4+, Ga3+ and Al3+), the highest reaction rates were observed for the catalysts with moderate Lewis acidity, such as Sn4+ and Ru3+-containing catalysts.86 In addition, layered transition metal oxides, for example HNbMoO6, have also been found to be a class of eectively solid acid catalyst for the hydrolysis of cellulosic materials to glucose.88 A 21% of the yield of glucose was obtained over layered HNbMoO6 catalysts, even higher than that over Amberlyst-15 (3.4%) under the same reaction conditions (Table 2). The peculiar layered structure might make facile the accessibility of saccharides into the interlayer gallery of HNbMoO6, where the strong acidic sites are located. Notably, such a layered structure exhibited good properties of water tolerance, which is of particular importance in a hydrolysis reaction. Nevertheless, in the case of solid acid catalysts, there is a signicantly poorer solidsolid reaction contact between catalysts and lignocellulosic biomass as both lignocellulosic biomass and solid cannot be dissolved in water. As such, either the H+ ions from solid Brnsted acids or the active sites from solid Lewis acid are dicult to access by reactants. Therefore, the 1,4 0 -b-glycosidic bonds suer from diculty to be activated, attacked and then opened. Consequently, it results in a low reaction activity. Moreover, in practice, there is another diculty in the use of solid catalysts. To determine the true conversion of lignocellulosic biomass remains an essential task because it is not easy to separate unreacted cellulose from sticky residues and various insoluble products that might readily be adsorbed onto the solid acid catalysts.79 To tackle these challenges, integrating the use of ionic liquids for dissolving microcrystalline cellulose along with solid acids
5596 Chem. Soc. Rev., 2011, 40, 55885617

for catalytic splitting of cellulose into sugars may provide a possible solution.89 For example, Rinaldi et al.90 showed that the cellulose biopolymers were broken down into oligomers in the form of averaging ten glucose units catalyzed by Amberlyst catalysts in 1-butyl-3-methylimidazolium chloride (C4MIMCl) ionic liquid after 5 h. It was postulated that the easy release of H+ ions into the solvent, the high surface area and large pores in Amberlyst catalysts were probably crucial factors to enable the long cellulose chains to be approached and activated, thereby increasing the reactivity. Furthermore, to further convert in situ hydrolytic glucose to value-added commodity chemicals in a one-pot reaction system is promising. So far the acidic solids have proved to be eective in many reactions in organic synthesis. For example, alkylation is also an acid-catalyzed reaction.91 After acidic resin Amberlyst catalyzed the hydrolysis of the 1,4 0 -b-glycosidic linkages in the cellulose polymeric chain to give glucose, such acid catalysts could further catalyzed the in situ alkylation of the hydroxyl groups at the C1 position of the glucose as an intermediate. Consequently, cellulose was directly converted into an environmentally friendly alkyl glycoside surfactant in a one pot transformation.92 This stateof-the-art integration of consecutive reactions has also been demonstrated in a recent work by Ignatyev et al.93 that cellulose was fully converted into alkylglycosides under mild conditions in the C4MIMCl ionic liquid in the presence of an acidic catalyst. Primary alcohols like n-butanol and n-octanol were used as alkylating reagents. In the reaction with n-butanol, the yield of butylglucopyranoside isomers reached 86%. According to a well-established concept, the removal of intermediate glucose by such alkylation can facilitate chemical equilibrium to move forward, thereby leading to a higher conversion of cellulose.

Catalytic solvolysis of lignocellulosic biomass

Apart from organic ionic liquid, certain organic reagents can also physically and chemically modify cellulosic materials by means of solvolysis, thereby promoting the contact between cellulose molecules and surfaces of a solid acid catalyst. In this context, the organic molecules act both as the solvents and as the reactants. Consequently, they can signicantly facilitate the catalytic conversion of lignocellulosic biomass over a solid catalyst. Clearly, there is a chemical interaction between the solvent and the solute lignocellulosic biomass. Therefore, the inherent properties of solvents have a signicant eect on products distribution. Moreover, it is also possible that organic solvents react with intermediates, for example oligosaccharides or monomers, to yield secondary products. Deng et al.94 reported that for the conversion of cellulose, the use of methanol as a solvent led to the formation of methyl glucosides in the presence of dilute H2SO4, heteropolyacids or sulfonated solid acid catalysts under mild conditions (r473 K). H3PW12O40 catalysts gave the highest TON (B73 in 0.5 h) to methyl glucosides. Additionally, methyl glucosides formed in methanol are more stable against further degradation than glucose in water. Recently, it was revealed that alcohol (methanol or ethanol) and water showed a synergistic eect on direct liquefaction of biomass.95
This journal is
c

The Royal Society of Chemistry 2011

View Online

The 50 wt% co-solvent of either methanol-water or ethanolwater was found to be an eective mixed solvent for the liquefaction of eastern white pine sawdust. Comparatively, the dissolution of lignin and hemicellulose is easier than that of cellulose owing to the abovementioned dierences in their chemical structure. Therefore, the solvolysis and degradation of cellulose might be worth being emphasized in the catalytic solvolysis of lignocellulosic biomass. Yamada et al.96 conducted the acid-catalyzed solvolysis of cellulose using ethylene carbonate (EC) as an organic reagent and 97% H2SO4 as catalyst. In comparison with solvolysis by using polyethylene glycol (PEG), using EC as a solvolysis reagent lead to a faster reaction rate. It was suggested that cellulose with EC led rst to the formation of glucosides, which then decomposed to form HMF derivatives or levulinic acid as intermediates.96 In particular, HMF derivatives could be either decomposed or polymerized to form an insoluble fraction of products. Thereby, to obtain a desired levulinic acid, how to control the formation of HMF derivatives might be a critical issue. The solvolysis of lignocellulosic biomass can even proceed in the absence of catalyst under hydrothermal conditions. The results showed that the solvent such as acetone, ethylene glycol and toluene have a profound eect on solvolytic product.97 In addition, solvolysis can also be used to deal with intermediates from deploymerization or degradation of lignocellulosic biomass, for example 5-HMF, and levulinic acid.98 Thereby it might be applied to acceleration of a catalytic solvolysis process mainly through the solvolysis of intermediates. Recently, Liu et al.99 investigated microwave-assisted organosolvolysis by a pretreatment of recalcitrant softwood in aqueous glycerol with a series of organic and inorganic acids with dierent pKa values including sulfuric, hydrochloric acid, phosphoric, malonic, acetic, citric and lactic acids. It was found that the eciency of pretreatment correlated linearly with the pKa of the acids except malonic and phosphoric acids. Noticeably, the organic acids, for example carboxylic acids, could act simultaneously as the solvent, the reactant and the catalyst. Moreover, organic acids can be possibly recovered and reused by distillation or extraction. In addition to common organic acids like the carboxylic acids and sulfonic acids, organic compounds bearing an OH or SH group might also have certain acidity. Thus specic organic compounds can be judiciously selected for both catalysis and solvolysis of lignocellulosic biomass. The method could provide an economical alternative route to the selective conversion of cellulose to small organic molecules.

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

materials instead. Clearly, the objective of the liquefaction of lignocellulosic materials is to produce low molecular-weight liquid products, instead of glucose and xylose by hydrolysis.100 The catalytic liquefaction of lignocellulosic biomass is generally performed in hydrothermal water similar to conditions for hydrolysis at the temperature of 393653 K (Table 3). An obvious advantage of the hydrothermal liquefaction process is that it does not require the lignocellulosic biomass feedstock to be pre-dried. Dierent from catalytic hydrolysis, however, the liquefaction process of lignocellulosic biomass targets deep degradation and decomposition of glucose through dehydration, isomerization, decomposition and rearrangement, and so forth.101 The products are greatly changed when the type of catalyst is substantially changed. For example, over halide catalysts, as shown in Scheme 5, glucose, 5-HMF and furfural could respectively decompose to liquid products, such as levulinic acid, 5-HMF, furan and formic acid, in a parallel reaction.102 However, over alkaline catalysts bio-oils and phenolic compounds are dominantly produced (Table 3). 4.1 Hydrothermal liquefaction over halide and sulfate catalysts Recent literature reports that catalytic liquefaction of lignocellulosic biomass was often carried out over sulfate, halide and oxide-based catalysts. Lewis acid catalysts could exhibit good catalytic properties in hydrothermal liquefaction of lignocellulosic biomass while catalytic hydrolysis is frequently conducted in the presence of Brnsted acid catalysts (Table 3). Several studies suggested, for example, that alkali and alkaline earth metal chlorides were not eective in the conversion of cellulose, while transition metal chlorides such as CrCl3, FeCl3, CuCl2 and AlCl3,105,106 including a pair of these metal chlorides (for example CuCl2 and CrCl2),107 exhibited high catalytic activity. In particular, CrCl3 was found to be highly active in the conversion of cellulose to levulinic acid, reaching a yield of 67 mol% after 180 min at 473 K with a catalyst dosage of 0.02 M and substrate concentration of 50 wt%.104 In particular, the use of ionic liquids such as [C4MIM]Cl105,106 and [EMIM]Cl107 can eectively promote cellulose depolymerization, thereby quickening the hydrothermal liquefaction. Recently, a one-step process for the conversion of lignocellulosic biomass into valuable furanic compounds in ionic liquids over CrCl3 under microwave irradiation was reported.105 The results showed that CrCl3 could eectively catalytically convert lignocellulosic biomass into 5-HMF and furfural with yields up to 52% and 31%, respectively, within a few minutes without any pretreatment. It was postulated that CrCl3 in [C4MIM]Cl might form complexes [C4MIM]n[CrCl3+n]. Thus, in the presence of ionic liquid, the 1,4-glucosidic bonds were weakened partially at the cellulose hydrolysis step under microwave irradiation (400 W) because of coordination with [CrCl3+n]n. As a result, it was more easily attacked by water to form glucose and oligomers (Scheme 6).106 This coordination then promoted rapid mutarotation of the a-glucopyranose anomer of glucose to the b- glucopyranose anomer of glucose through hydrogen bonds of chloride anions with the hydroxyl groups of glucose. The hemiacetal portion of b-glucopyranose with [CrCl4] then formed a Cr3+ enolate anion complex and
Chem. Soc. Rev., 2011, 40, 55885617 5597

4 Catalytically hydrothermal liquefaction of lignocellulosic biomass


As discussed above, over certain acid catalysts, longer reaction time, higher reaction temperature and pressure can readily lead to deep degradation and decomposition of glucose, xylose, arabinose and cellobiose and oligosaccharides which are yielded from the hydrolysis of lignocellulosic biomass. Thereby a liquid mixture product mainly consisting of C2C6 organic compounds are obtained. In this respect, such a process is usually called the liquefaction of lignocellulosic
This journal is
c

The Royal Society of Chemistry 2011

View Online

Table 3 Catalyst

Catalytically hydrothermal liquefaction of lignocellulosic biomass or pure cellulose in an aqueous medium Feedstock Cellulose Reaction conditions Main products Yield (%) B6.6 5.8 (wt%) 19.2 (wt%) 2.3 (wt%) 4.0 (wt%) 8.0 (wt%) B9.5 62.0 20.0 52.0 31.0 57.5 1.27 67.0 N/a N/a 47.2 (wt%) 53.3 9.3 (wt%) N/a B53
a

Researchers/Year Kong et al./2008 Seri et al./2002 Schutt et al./2002 Kong et al./2008 Li et al./2009 Zhang et al./2010 Su et al./2010 Peng et al./2011 Kumar et al./2008 Akhtar et al./2008 Song et al./2004 Qian et al./2007 Karagoz et al./2004 Liu et al./2006

Ref. 109 108 103 109 106 105 107 104 124 119 123 122 115 59

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

573 K, 120 s Lactic acid Cellulose 523 K, 180 s (in N2) Glucose LaCl3 (0.01 M) 5-HMF Levulinic acid Cellulose 423 K, 5 h Acetic Acid La/Al2O3 (0.34 wt%) Malic Acid Cellulose 573 K, 120 s Lactic Acid CoSO4 (400 ppm) MW (400 W) 5-HMF CrCl3 (3.6 wt%) + [C4MIM]Cl Cellulose 2 min, 0.1 MPa Reducing sugar Pine wood MW (400 W) 5-HMF CrCl36H2O+ [C4MIM]Cl 3 min, 0.1 MPa Furfural Cellulose 393 K, 8 min 5-HMF CuCl2/CrCl2 + EMIM]Cl Sorbitol Cellulose 473 K, 3 h Levulinic acid CrCl3 (0.02 M) Cellulose 608 K, 4.7 s bio-oil K2CO3 (0.5 wt%) 27.6 MPa Empty palm 543 K, 20 min Phenolic compounds K2CO3 (1.0 M) fruit bunch 2 MPa (H2O) Corn stalk 647 K bio-oil Na2CO3 (1.0 wt%) 25 MPa (H2O) Woody biomass 653 K, 1620 min Heavy oil Na2CO3 8 MPa (H2) Sawdust 553 K, 15 min Oil Ca(OH)2 (0.0243 M) KOH (0.5 M) Walnut shells 523 K, 430 min, Phenolic compounds 1.58.6 MPa Lignin 573 K, 1 h Phenolic compounds Ba(OH)2 or Rb2CO3 2 MPa, H2 and oils Ba(OH)2 or Rb2CO3 Sawdust/cornstalks 573 K, 1 h Phenolic compounds 2 MPa, H2 and oils Waste biomass 673 K, 10 min Oil K2CO3 + ZrO2 22.1 MPa

NiSO4 (400 ppm)

Tymchyshyn et al./2010 126

B32 + 30/ Tymchyshyn et al./2010 126 32a + 25 913% Hammerschmidt 125 et al. /2011

[EMIM]Cl: 1-ethyl-3-methylimidazolium chloride. [C4MIM]Cl: 1-n-butyl-3-methylimidazolium chloride. N/a: not available. MW: Microwave irradiation. a data from reaction without catalysts, and the sum is the data with addition of catalysts.

it aroused the isomerization of glucose into fructose. Finally, fructose is dehydrated to 5-HMF.106 Nonetheless, the heating by microwave irradiation could also have an eect on cellulose decomposition and 5-HMF formation. Though microwave irradiation is facile at the laboratory scale, it might still be dicult at large scale-up in industry. Furthermore, it cannot be ignored that ionic liquids may also play a role in catalysis, in addition to their role in facilitating the dissolution of cellulose.107 Rare-earth metal salts or ions, for example lanthanum(III) ions, which retain Lewis acid reactivity even in aqueous solutions, have been used to catalyze the degradation of lignocellulosic biomass, typically cellulose.103 Earlier, Seri et al.108 showed that the conversion of cellulose degradation in water at 523 K in the presence of LaCl3 reached 80.3% after 180 s. A 54.2 wt% yield of a water-soluble product mainly containing 5-HMF glucose and levulinic acid was obtained. In addition to metal halides, sulfates can also be used as catalysts for the catalytic liquefaction of cellulose. Furthermore, the reaction can be directed towards a specic target product other than 5-HMF, glucose and levulinic acid. Kong et al.109 revealed, for example, that lactic acid can be produced from the catalytic hydrothermal liquefaction of lignocellulosic biomass in the presence of dierent transition metal ions like ZnSO4, NiSO4, CoSO4 or Cr2(SO4)3. In comparison to a non-catalytic process, the addition of 400 ppm
5598 Chem. Soc. Rev., 2011, 40, 55885617

Ni2+ catalyst evidently increased the yield of lactic acid from 3.25% to 6.62% at 573 K for 120 s in the hydrothermal liquefaction microcrystalline cellulose samples in subcritical water. It is noteworthy that lactic acid can also act as a platform chemical from which lots of ne chemicals can be obtained.6 4.2 Hydrothermal liquefaction over alkaline catalysts One of the important features of lignocellulosic biomass is that it mainly contains two types of biopolymers, namely cellulosic polymer and lignin polymer. Relatively speaking, the former readily interacts with acid while the latter is apt to interact with alkali. Therefore, in the presence of alkaline catalysts, liquefaction of lignocellulosic biomass mainly leads to oil-like products, although this process is usually conducted thermally in an aqueous medium.110,111 In a sense, this process mimics the conversion of ancient plant material into the crude oil.112 So far, various carbonates like Na2CO3, K2CO3, and hydroxides such as KOH or Ca(OH)2 as catalysts have been evaluated in the hydrothermal liquefaction for wood, agricultural and civic wastes such as sugar cane bagasse and corn stalk as sources of lignocellulosic biomass (Table 3). The so-called bio-oil product is actually a complicated liquid mixture with a wide range of compositions. It typically consists of glycoaldehyde dimers, 1,3-dihydroxyacetone dimers, anhydroglucose,
This journal is
c

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 5 Main reaction pathways of degradation of major product, glucose, from cellulose hydrolysis over dierent catalysts.90,102 Adapted and reprinted with permission from ref. 90. Copyright 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim; Adapted and reprinted from ref. 102, Copyright 2008, with permission from Elsevier.

Scheme 6 Direct conversion of cellulose to 5-HMF catalyzed with CrCl3 in [C4MIM]Cl (n = 13).106 [C4MIM]Cl: 1-n-butyl-3-methylimidazolium chloride. Reprinted from ref. 106, Copyright 2009, with permission from Elsevier.

soluble polyols, 5-HMF, furfural, organic acids, phenolic compounds and even hydrocarbons. Therefore, an ecient hydrothermal liquefaction process of lignocellulosic biomass over alkaline catalysts oers remarkable potential for simultaneously producing biofuels and value-added chemicals by integrating it with proper separation and extraction techniques. Clearly, the reaction temperature and time and the type of alkaline medium have a signicant eect on the reaction rate
This journal is
c

and product distribution.113,128 As for the eect of temperature on the reaction rate and the product distribution, for example, earlier Minowa and co-workers114 showed that in the presence of 0.826 wt% Na2CO3 under 3 MPa, the decomposition of cellulose started at a reaction temperature lower than 453 K. Only water-soluble products were obtained at temperatures below 533 K. The cellulose decomposed quickly at 533573 K. Bio-oil was formed at temperatures above 533 K and its yield reached the highest values at 593613 K.
Chem. Soc. Rev., 2011, 40, 55885617 5599

The Royal Society of Chemistry 2011

View Online

No cellulose remained after reaction at temperatures greater than 573 K. Later, Karagoz and colleagues115 used a 0.0243 M Ca(OH)2 solution as a solvent instead of pure water a in liquefaction reaction of wood biomass. The results also revealed that increasing the reaction temperature increased the conversion of wood biomass and the total bio-oil yield. Maximum bio-oil yield was found to be 8.5 wt% for a reaction at 553 K for 15 min. Regarding these ndings, in general, an obvious reason is that decomposition temperature for cellulose and lignin is inherently dierent. Moreover, decomposition of cellulose tends to yield polyols, alcohols and carboxylic acids while decomposition of lignin principally leads to phenolic compounds. However, at elevated temperatures or over longer reaction times, the occurrence of secondary reactions of bio-oil products is increased so much that the yield of the bio-oil decreases. In particular, for most of the lignocellulosic biomass treated at a temperature higher than 573 K a carbonization process tends to readily take place.116 The carbonization could be eectively avoided at lower reaction temperature in the presence of an appropriate catalyst. With regard to low-temperature catalytic hydrothermal treatment of wood biomass, several studies showed that the use of alkaline catalysts increased the formation of the compounds such as 4-methyl-2-methoxy-phenol, 4-ethyl-2-methoxy-phenol and benzenediol derivatives and decreased the solid residues, tar and biochar.117,118 The role of catalysts in directing product distribution is clearly proved by the observation that the furan derivatives obtained from the thermal run were not observed for the catalytic run. Moreover, the conversion and yield of liquid products decreases in the following order: K2CO3 4 KOH 4 Na2CO3 4 NaOH.117119 It is most likely that the decomposition of lignocellulosic biomass in water inhibits char formation. In particular, such a process is conducted under subcritical or supercritical water conditions. It is well-documented that near the critical point of water (647 K, 22.1 MPa), the ionic concentration (Kw = [H+] [OH]; [H+] = [OH]) increases and water behaves as a weakly polar solvent. Therefore, subcritical and supercritical water conditions are often considered to be applied to the liquefaction of lignocellulosic biomass. Earlier, Demirbas120 found that the yield of bio-oil from catalytic supercritical uid extraction of oriental beech reached 68.6% at 573 K after 30 min in the presence of 20% wt NaOH. However, there are also disadvantages in the use of subcritical or supercritical water. Typically, the equipment corrosion by caustic hydroxides is severely enhanced under subcritical and supercritical water conditions. Therefore, in this aspect, alkali and alkaline earth carbonate salts are favorable for use as catalysts. When cellulose liquefaction is carried out in the presence of Na2CO3 under subcritical water and at a reaction temperature of 623 K, bio-oil and gas were the main products, with little solid residue formed.121 Qian et al.122 reported that when woody biomass was liqueed at 653 K in the presence of Na2CO3 as a catalyst, a yield of 53.3% of bio-oil, mainly composed of hydrocarbon, aldehyde, ketone, hydroxybenzene and ester, was obtained. However, Song et al.123 showed that in the case of isothermal liquefaction of corn stalks in water without a catalyst, the conversion was also very high, reaching 95.4 wt% with a liquid
5600 Chem. Soc. Rev., 2011, 40, 55885617

yield of 77.6% and gas production of 17.9%. When 1.0 wt% Na2CO3 was added as a catalyst, the conversion rate remained unchanged (95.7%), but the catalytic reaction favored the yield of liquid products which increased to 88.0% and the gas production decreased to 6.8%. Moreover, in the presence of a catalyst, the bio-oil content in the liquid products sharply increased from 33.4 wt% to 47.2 wt%. Clearly, not only does the catalyst change the rate of the liquefaction reaction, but it can also change the direction of the liquefaction reaction towards dierent types of products. The hydrothermal liquefaction is almost all conducted in a batch reactor. However, it can be carried out in a continuous ow reactor. For example, Kumar et al.124 revealed that when using K2CO3 as a catalyst in a continuous ow reactor, increasing the reaction temperature to the supercritical region, or increasing the reaction time, hydrolytic oligomers and monomers of cellulose samples were further degraded into glycoaldehyde dimer, D-fructose, 1,3-dihydroxyacetone dimer, anhydroglucose, 5-HMF, and furfural, thereby yielding biooils. To improve the selectivity to target compounds in the bio-oil, Hammerschmidt and co-workers125 recently described a continuous one-step process making use of two kinds of catalysts, a homogeneous K2CO3 catalyst and a heterogeneous ZrO2 catalyst. Such combination of a homogeneous system with a heterogeneous system provides new ways to realize a more ecient continuous process. It should be noted that thermal liquefaction in an aqueous solution might also undergo hydrolysis, pyrolysis and steam gasication, in particular in super- or subcritical water. Such complex reaction networks for cellulose decomposition in homogeneous and heterogeneous reaction conditions under subcritical water using Na2CO3, and Ni as catalysts and without catalysts, have been proposed by Fang and co-workers, as shown in Scheme 7.121 In the course of the reaction, dissolved compounds, for example part of cellulose, oligomers, glucose, and levoglucosan and non-dissolved feedstock cellulose and compounds, might undergo dierent reaction mechanisms, thereby resulting in dierent intermediates and nal products. At a high heating rate, a Na2CO3 catalyst resulted in more oil while a Ni catalyst produced more gas. The dissolved reaction paths in the homogenous phase were homogeneously hydrolyzed to glucose and other soluble pyrolytic products (e.g. levoglucosan), which then further decomposed to acids, aldehydes, and alcohols of C1C3. Part of the glucose degraded to furfurals and 5-HMF, which ultimately condensed to phenols and dehydrated to levulinic and formic acids. The nondissolved reaction fraction in the heterogeneous phase possibly underwent pyrolysis to yield a mixture of hydrocarbons, aromatic derivatives and tar, randomly linked oligosaccharides, light volatiles and gases via dehydration, cracking and ssion, and rearrangement of the sugar units. Even for single cellulose as a feedstock, the reaction mechanism is extraordinarily complicated. For lignocellulosic biomass, the presence of lignin and hemicellulose certainly make it much more perplexing. Among the components in bio-oils, phenolic compounds are one type of typical, important product from the liquefaction of lignocellulosic biomass after treatment with alkali. It is welldocumented that alkalis are exceptionally eective to extract and convert lignin lignocellulosic biomass. Liu et al.59 showed that
This journal is
c

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 7 Mechanism of cellulose decomposition promoted in homogeneous and heterogeneous environments in subcritical water. High heating rates result in glucose char without catalyst, but more oil with Na2CO3 catalyst, and more gas with Ni catalyst.121 Reprinted with permission from ref. 121. Copyright 2004 American Chemical Society.

the products from the hydrothermal conversion of walnut shells catalyzed by 0.5 M KOH at 573 K were mainly phenol derivatives, especially the methoxy phenolic compounds. Clearly, these chemicals are readily derived from lignin. Owing to the inherent multicomponent plant cells, small amounts of cyclopentene derivatives and C1218 fatty acids were also detected. Recently Tymchyshyn and colleagues126 described a series of direct liquefaction of lignocellulosic wastes like sawdust and cornstalks, and model biomass compounds like lignin and cellulose in hot-compressed water at temperatures from 523 to 623 K in the presence of 2 MPa H2, with the addition of Ba(OH)2 and Rb2CO3 catalysts. Signicant quantities of phenolic compounds such as 2-methoxy-phenol, 4-ethyl-2methoxyphenol, and 2,6-dimethoxy-phenol were obtained from the two lignocellulosic wastes and pure lignin. Yet the liquid products from cellulose contained essentially carboxylic acids and neutral compounds. Such experimental data evidently revealed that lignin and cellulose inherently tend to result in dierent categories of products. To bear this in mind is of much importance in designing and developing a catalytic system for hydrothermal liquefaction of lignocellulosic biomass to the target products based on dierent goals. As discussed above, catalytically hydrothermal liquefaction of lignocellulosic biomass produces a very complex mixture of liquid products along with some gases and solid residues and tars. It makes the production of bio-oil from hydrothermal liquefaction of lignocellulosic biomass very uneconomical at present when compared to the costs of diesel or gasoline production from fossil oils. Scheme 8 shows the separation and extraction procedure proposed by Karagoz et al.117 It involves the complicated multistep extraction by several
This journal is
c

types of solvents. Moreover it needs inorganic acids for neutralization and the separation of water-soluble organic compounds in the water phase. Therefore, in addition to the technical need for the improvement of the activity and selectivity of catalysts, the novel technology for separation and extraction of downstream products from hydrothermal liquefaction of lignocellulosic biomass need to be developed.127,128 On the other hand, such complexity in products is a reection of the complex reaction mechanism in catalytic liquefaction of lignocellulosic biomass. Thus far, researchers have not acquired in-depth understanding of the reaction mechanism yet. To disclose it certainly is conducive to developing such a process for the commercial use. Finally, generally speaking, hydrothermal liquefaction is relatively time-consuming and energy-intensive. In this context, comparatively, fast pyrolysis, as discussed below, is more promising.

Catalytic pyrolysis of lignocellulosic biomass

Pyrolysis of lignocellulosic biomass is generally a process of thermal chemical decomposition in the absence of oxygen to convert biomass into liquid products (bio-oils) together with some gases and solid chars. Such a thermolysis process simultaneously involves various reactions such as depolymerization, dehydration, decarboxylation, esterication, condensation, cyclization, and so forth.27,129,130 However, dierent from catalytic liquefaction at lower temperature in the aqueous phase, pyrolysis is usually conducted at temperatures greater than 773 K (Table 4). The liquid products are commonly referred to as pyrolysis oils, which have been identied as promising alternative renewable liquid fuels.131,132
Chem. Soc. Rev., 2011, 40, 55885617 5601

The Royal Society of Chemistry 2011

View Online

5.1 Fast pyrolysis mechanism The fast pyrolysis reaction of lignocellulosic biomass undergoes a complex mechanism in oxygen-decient conditions at high temperature, even though the reaction is conducted in an inert atmosphere. Therefore, presently the investigation into the mechanism could be simplied through the single use of cellulose or lignin in a non-catalytic system. As shown in Scheme 9, earlier Luo et al.137 proposed a fast pyrolysis model of cellulose in an inert nitrogen atmosphere. Hydroxyacetaldehyde and 1-hydroxy-2-propanone was thought to be produced from the decomposed fragments of activated cellulose by competing against levoglucosan formation. Meanwhile, the intermediate fragments might also decompose to yield gases (COx). At low temperature, activated cellulose would also directly decompose to produce char, CO2, and H2O. A long residence time of intermediates like levoglucosan in the reactor could lead to the secondary cracking reaction, thereby leading to the formation of methanol, aldehyde, allyl alcohol, glyoxal, furfural, 5-HMF and gases. Regarding the gas products, a possible mechanism for fast pyrolysis of cellulose in an inert gaseous atmosphere was proposed by Lanza and co-workers recently,138 with the following simplied decomposition sequence: Cellulose - Gm, H2O where Gm = short oligomers. Gm, H2O - Cn, CO, H2O (+CO2, H2), with n r 4 Cn - CH4 CH4 + H2O - CO + 3H2 Similar to the liquid products from thermal liquefaction, the pyrolysis oil is a complicated mixture of oxygenated and deoxygenated compounds, typically consisting of organic acids, aldehydes, ketones, esters, aromatic and phenolic derivatives, and even hydrocarbons. Accordingly, liquid pyrolysis oils are normally acidic and emulsion-like uids containing both aqueous phases and phenolic phases. Thus it is imperative to further upgrade pyrolysis oil so as to get useful liquid hydrocarbon-like fuels and fuel additives. Pyrolytic processing of lignocellulosic biomass can be roughly classied into steam pyrolysis, fast pyrolysis in the ow of an inert gas and vacuum pyrolysis. For all of these approaches, a major, common objective is to improve the selectivity of low molecular-weight substances with great potential as fuels. As seen in Table 4, fast pyrolysis has been the attractive subject of a lot of research over the past decade because it is proven to be two to three times more economical than liquefaction and gasication processes.133 Generally, four parameters play a pivotal role in the catalytic activity and selectivity of targeted pyrolysis oils. They are (1) the catalysts, (2) heating rate of lignocellulosic biomass, (3) residence time of reactants and the products inside the reactor, and (4) reaction temperature.130 In addition, the atmosphere in the reactor is inuential. For example, if the pyrolysis is conducted in the presence of hydrogen, more hydrogenated products will be obtained. Such an integrated pyrolysis with hydrogenation process will be discussed in detail in Section 7.
5602 Chem. Soc. Rev., 2011, 40, 55885617

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 8 Separation and extraction procedure.117 Reprinted with permission from ref. 117, Copyright 2005, with permission from Elsevier. WSH: Water soluble hydrocarbons (WSH and others in wt% = 100 (oil1 + oil2 + oil3 + solid + gas). (Typical products in oil 1: 2-Furancarboxaldehyd, 2-Methoxy-phenol, 4-Methyl-phenol, 4-Hydroxy-3-methoxy-benzeneacetic acid, 2-Hydroxy-3-methyl-2cyclopenten-1-one and Butylated hydroxytoluene; oil 2: 3-Methoxy-1propene, 4-oxo-Pentanoic acid, 5-Ethyldihydro-2(3H)-furanone; oil 3: 2-Furancarboxaldehyde,2-Methoxy-phenol, 3-Methyl-phenol or 4-methyl-phenol, Eugenol, 4-Hydroxy-3-methoxy-benzeneaceticacidmethyl ester, 3-Acetyl-7,8-dimethoxy-2-methyl-1H-naptho[2,1-b]pyran-1one.

(2)

(3) (4) (5)

As depicted by the reaction paths in eqn (2)(5), cellulose decomposes primarily to H2O and short oligomers (Gm), which is then broken down to levoglucosan. The levoglucosan easily breaks down to mainly form CO, H2O and hydrocarbons up to C4, along with minor CO2 and H2. In these sequential reactions, butadiene was identied as a key intermediate. The hydrocarbons undergo further thermal cracking reactions to yield CH4. The presence of H2O steam and hydrocarbons results in a reforming reaction, nally leading to chemical equilibrium composition among syngas and CH4. According to this mechanism, H2 and CO can form in two dierent reaction routes in eqn (2) and (5), respectively. And the dierent pathways and rates in the formation of CO and H2 explain why the ratio of CO to H2 is not constant under dierent conditions, particularly at short residence time. Therefore, it could be reasonably deduced that a longer contact time is required if the aim is to produce syngas from the fast pyrolysis of cellulose. Although all of the abovementioned mechanisms have proposed relatively simple reaction routes and identied key intermediates in the non-catalyst system,137,138 it is evident that fast pyrolysis involves a series of consecutive and parallel reactions competing against each other among liquid products and gaseous products. Nevertheless, as such reactions are frequently conducted either in the gas or liquid phase in the presence of homogenous and heterogeneous catalysts,139,140 the reaction mechanism might undergo far more complicated
This journal is
c

The Royal Society of Chemistry 2011

View Online

Table 4

Typical pyrolysis of lignocellulosic biomass in the presence or absence of catalysts Catalyst No No No No 0.1wt% sulfuric or polyphosphoric acid Pretreated by 1wt% phosphoric acid Pretreated by 1% phosphoric acid NaOH Na2CO3 Fe2(SO4)3 ZSM-5 H-ZSM-5 Ni-ZSM-5 Reaction conditions Vacuum pyrolysis 773 K Flash pyrolysis 11231473 K, 3575 ms. Pyrolysis, 773 K Under N2 ow Fast pyrolysis 1173 K, Z 3 h Pyrolysis in tetramethylene sulfone, 473 K, 6 min, Under N2 ow Fast pyrolysis, 773 K Under He ow Fast pyrolysis, 773 K Under He ow Fast pyrolysis, 773 K Microwave heating Fast pyrolysis, 773 K Microwave heating Fast pyrolysis, 773 K Microwave heating Fast Pyrolysis, 873 K, 240 s Under He ow Flash pyrolysis, 673773 K, 50 ms, under N2 ow Fast pyrolysis, 873 K Under He ow Typical products Parans, sterols Fatty acid methyl esters Soluble solids Liquid product Liquid product CO, H2, CH4, hydrocarbons Levoglucosenone, furfural 5-HMF Levoglucosane Levoglucosenone Levoglucosane Levoglucosenone Acetol Furfural 2-furanmethanol Acetol Furfural 2-furanmethanol Acetol Furfural 2-furanmethano Aromatic compounds CO C4-hydrocarbons Hydrocarbons Toluene Yield (%) N/a 20.045.9 12.5-27.4 60 N/a 42.2 26.9 8.8 33.6 14.1 25.2 29.8 53.3 0 2.08(area%) 45.28 1.03 2.58 0 50 0 31.1 (C%) 15.2 (C%) 15.9 16 3.5 Researchers/Year Garcia-Perez et al./2007 Piskorz et al./2000 Karagoz/2009 Lanza et al./2009 Kawamoto et al./2007 Dobele et al./2005 Dobele et al./2005 Chen et al./2008 Chen et al./2008 Chen et al./2008 Carlson et al./2008 Olazar et al./2000 French et al./2009 Ref. 134 135 136 138 144 142 142 146 146 146 152 150 151

Feedstock Softwood bark and Hardwood Cellulose Soybean oil cake Cellulose Cellulose

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Wood Cellulose Pine wood sawdust Pine wood sawdust Pine wood sawdust Cellulose Pine sawdust Wood

n/a: no available. No. without catalyst.

Scheme 9 Modied model of fast pyrolysis of cellulose.137 Adapted and reprinted with permission from ref. 137. Copyright 2004 American Chemical Society.

steps and reaction networks. Furthermore, many intermediate products are more active than cellulose itself and polyols, thereby making the reaction more perplexing. Such complex reaction networks are still a puzzle and a barrier on the way to understanding and utilizing cellulose by pyrolysis conversion in a more ecient, economical way. 5.2 Catalytic fast pyrolysis

To increase the pyrolysis eciency and the content of the target compounds in bio-oils, liquid acid such as H2SO4, hydrochloric acid, phosphoric acid and Lewis acids can be used as a specic pretreatment for pyrolysis of lignocellulosic biomass.141
This journal is
c

For example, Dobele et al.142 reported that the pretreatment of pine wood sawdust and microcrystalline cellulose samples with phosphoric acid or iron(III) sulfate drastically increased the contents of levoglucosenone and levoglucosane in volatile products obtained by a subsequent fast pyrolysis (Table 4). These results were attributed to the catalysis of the cellulose depolymerization and dehydration in the presence of acids from the pretreatment. As for the use of iron ions, possibly the cleavage of glycoside bonds were catalyzed by iron oxihydroxide which formed during pretreatment of lignocellulosic biomass. As a result, such a pretreatment had a positive eect on the yield of 1,6-anhydrosaccharides. With regards to lignocellulosic biomass with a higher content of
Chem. Soc. Rev., 2011, 40, 55885617 5603

The Royal Society of Chemistry 2011

View Online

lignin, higher concentration of phosphoric acid or iron(III) sulfate was needed for pretreatment. Clearly, directly employing catalysts in the fast pyrolysis of lignocellulosic biomass can remarkably improve the conversion and selectivity to targeted products. As for the catalytic fast pyrolysis of microcrystalline cellulose at 7731073 K by on-line analysis of the pyrolysis vapors, Lu et al.143 revealed that SO42/SnO2 was the most eective catalyst to yield 5-methyl furfural. The selectivity was evidently changed with the change of catalyst support. For example, the SO42/TiO2 catalyst favored the formation of furfural, while SO42/ZrO2 favored the formation of furan.143 In addition, it is noteworthy that the type of solvent used during the pyrolysis reaction also has a signicant impact on the distribution of products. Kawamoto et al.144 depicted the catalytic pyrolysis of cellulose catalyzed by H2SO4 in tetramethylene sulfone (C4H8O2S) and found that the yield of levoglucosenone, furfural, and 5-HMF reached up to 42.2%, 26.9%, and 8.8% (as mol% yield based on the glucose unit), respectively, at 473 K for 6 min.144 Remarkably, the cellulose pyrolysis in tetramethylene sulfone suered so little from carbonization reactions that the residues obtained from pyrolysis in sulfolane were colorless. There, levoglucosan was identied as the major anhydromonosaccharide formed during the pyrolysis of cellulose.145 Tetramethylene sulfone is an aprotic polar solvent and it could dissolve anhydromonosaccharides, therefore it could prevent the polymerization and carbonization reaction. As such, cellulose completely decomposed into soluble products in tetramethylene sulfone.145 In contrast, pyrolysis under nitrogen or in dioctyl phthalate (a poor solvent for levoglucosan) produced brown/black solid residues. The ndings indicated that the solvent is also inuential in pyrolysis and it can be properly selected in view of promoting dissolution of lignocellulosic biomass and/or intermediates. At present, the catalytic pyrolysis of lignocellulosic biomass is often conducted in the presence of acid or alkaline catalysts such as carbonates and hydroxides. Because the pyrolysis oils yield various organic acids, the presence of alkaline species in the reaction system has an important eect on the stability of bio-oil. Phosphates, sulfates and chlorides also have certain catalytic activity in the catalytic pyrolysis of lignocellulosic biomass. With regard to the catalytic fast pyrolysis of pine wood sawdust by microwave heating in a uidized reactor, alkaline sodium compounds like NaOH, Na2CO3 and Na2SiO3 catalytically resulted in bio-oils rich in acetol, and to some extent favored H2 formation while the use of the Fe2(SO4)3 catalyst favored the formation of fufural and 4-methyl-2methoxy-phenol.146 In addition, phosphates, sulfates and chlorides could be used to replace corrosive alkaline catalysts or corresponding acid catalysts. Other than catalysis, diammonium phosphate (NH4)2HPO4 and diammonium sulfate (NH4)2SO4 could act as dehydrating agents upon heating during pyrolysis. Therefore, at reaction temperatures of about 650750 K, a pretreatment of water-extracted re wood by the impregnation in about 1% of (NH4)2HPO4 or (NH4)2SO4 solution before pyrolysis is able to increase the yields of levoglucosan, 5-methyl-2-furaldehyde and 2-acetylfuran in the range of 25%.147 Furthermore, alkali and alkaline earth metal chlorides have also been investigated as catalysts. Shimada and co-workers148 showed that the use of alkaline earth metal
5604 Chem. Soc. Rev., 2011, 40, 55885617

chlorides could substantially reduce the reaction temperature from 673 to 423 K. Recently, the catalytic pyrolysis of lignocellulosic biomass over zeolites and supported solid catalysts have received serious attention. In particular, it is attractive to employ continuous fast pyrolysis of lignocellulosic biomass catalyzed by solid catalysts which has obvious advantages over the use of liquid acid catalysts in a batch reactor. More importantly, the most desired product from the catalytic pyrolysis of lignocellulosic biomass should be directed to hydrocarbonscontaining pyrolysis oils with high yield, similar to petrol presently in use with easy storage and transportation. In response, recent advances have shown that hydrocarbons can indeed be produced in considerable quantities by fast pyrolysis of biomass over zeolite and mesoporous catalysts such as ZSM-5, Al-MCM-41.149 Such a catalytic fast pyrolysis demonstrated an optimistic prospect for direct production of high-octane pyrolysis oils from biomass for practical applications. For instance, Olazar et al.150 studied the fast pyrolysis of pine sawdust with ZSM-5 in a spouted-bed reactor using nitrogen as the carrier gas and observed a yield of aromatic compounds of 12% (carbon). By using a semi-continuous ow reactor with helium as carrier gas, French et al.151 revealed that lignocellulosic biomass could be pyrolyzed at temperatures ranging from 673 to 873 K with a catalyst-to-biomass ratio of 510 by weight over nickel, cobalt, iron, or galliumsubstituted ZSM-5, leading to an approximately 16 wt% yield of hydrocarbons. Aromatic compounds like toluene151 and even gasoline-like products152 can be obtained in remarkable quantities even without using the hydrogen. However, over solid catalysts at high temperature, the cracking and deoxygenating activity might decrease with time because of the coke deposits formed on the solid catalyst.152 Finally, it should be noted that the distribution of the pyrolysis products such as charcoal, liquids and gaseous products are also strongly inuenced by the physical and chemical characteristics of the raw materials of lignocellulosic biomass. Therefore, the cellulosic sources with adequate chemical composition should be selected for the specic desired product, together with the search for optimal reaction conditions and the catalysts. In this context, the growth and productivity, ecological inuence and cost of plant sources of lignocellulosic biomass should be taken into consideration.

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Catalytic gasication of lignocellulosic biomass

Gasication is usually referred to as a process of oxygendecient thermal decomposition of carbonaceous matter such as coal, petroleum, biofuel, or lignocellulosic biomass with a major objective to produce valuable gaseous products like hydrogen or syngas.153155 In particular, the production of hydrogen, an environmentally clean fuel, from cellulosic biomass gasication is of much interest to scientists and engineers. Through FischerTropsch synthesis techniques, bio-syngas can also be converted into synthetic liquid fuels and classied as clean and renewable transportation fuels.156 In this context, gasication plus FischerTropsch synthesis are denoted as indirect liquefaction processes of ignocellulosic biomass. Gasication of lignocellulosic biomass can also be
This journal is
c

The Royal Society of Chemistry 2011

View Online

performed at high temperatures with a controlled amount of oxygen like catalytic partial oxidation.157 Provided that air is used, the resultant products will contain nitrogen, thereby leading to an additional problem of separation and purication. In many instances, gasication also produces small amounts of tar, mainly composed of highly condensed polyaromatic hydrocarbons.158,159 To avoid such a problem and to increase the eciency, biomass gasication in supercritical water is of growing scientic and technical interest.160162 This so-called steam gasication of cellulose, can be described by the following reaction (eqn (6)).163165
Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Org+ refers to an intermediate from the decomposed lignocellulosic biomass. According to such a proposition, CO was formed through the cyclic oxidation-reduction reaction of RuO2 with intermediate organics compounds (Org) and then hydrogen was generated from the reduction of water by Ru2+ (eqn (7) and (8)). The hydrogen and CO could reacted to form methane (eqn (9) and (10)). CO + 3H2 - CH4 + H2O 2CO + 2H2 - CH4 + CO2 (9) (10)

0.59C6H12O6 + 0.7H2O - 1.42CH4 + 2.12CO2 + 1.4H2 (6) Table 5 lists some typical reaction conditions and results in catalytic gasication of lignocellulosic biomass in the presence of steam. The possibly improved solubility of cellulosic materials in the supercritical water employed can greatly reduce mass-transfer limitation. In addition, it can prevent catalyst deactivation due to coke deposition by means of extracting the coke precursor from the catalyst surface in the supercritical water.166 Moreover, the use of supercritical water might improve the hydrolysis and saccharization of cellulose in extremely short residence times. Furthermore, Fushimi et al. recently revealed that the evolutions of water-soluble tar and gaseous products (CO, H2, CH4 and C2H4) are signicantly suppressed by the interaction between cellulose and lignin.167 6.1 Gasication over supported noble metal catalysts

Supported Ru, Pt or Pd catalysts appear promising in the catalytic gasication of lignocellulosic biomass under supercritical conditions (Table 5). In addition, these catalysts are commonly active for hydrogenation. Therefore, the H2 produced from gasication in turn might be a reactant for in situ consecutive hydrogenation. Various Ru-based catalysts such as RuO2, Ru/TiO2 and Ru/C have been explored as catalysts in the catalytic gasication of lignocellulosic biomass. Earlier, Park et al.169 reported the production of hydrogen from sewage disposals in supercritical water using RuO2 as the catalyst. The reaction was more eective than conventional steam reforming to produce hydrogen. Izumizaki and co-workers170 chose RuO2 as a catalyst for the production of hydrogen from biomass in supercritical water under argon atmosphere of 44 MPa at 723 K for 120 min. The amount of hydrogen in gaseous mixtures was 15.0% from cellulose, 14.1% from pulp, 21.0% from the mixture of cellulose and lignin, 16% from waste paper and 27% from paper sludge, respectively. RuO2 evidently accelerated the thermal decomposition for cellulose and pulp. However, for lignin, the catalytic performance could be greatly suppressed because of the quick deactivation of the RuO2 catalyst by coke and char formation. A possible reaction model over such a RuO2 catalyst is proposed as below.170 RuO2 + Org - Ru
2+

According to this reaction mechanism, a higher hydrogen ratio in the product stream is subject to low methane formation. In this case, the desired catalyst is the one which can signicantly enhance the overall conversion of lignocellulosic biomass whilst hindering reduced methane formation so that the yield of hydrogen can be increased. Osada et al.171 showed that the gasication of cellulose over Ru/TiO2 in supercritical water conducted at 673 K for 15 min led to a high gas yield of 74.4 C% (calculation based on the carbon balance). Nevertheless, using the same reaction conditions without the presence of a catalyst, the gas yield and the water-soluble yield were 11.3 C% and 39.0 C%, respectively. In this case, it is reported that in the catalytic reaction no solid product was formed. In addition to the use of supercritical water conditions, to tackle the insolubility of cellulosic materials in gasication, there are a few reports suggesting that the derivative of cellulose, carboxymethylcellulose (CMC) and acetyl cellulose, can be mixed with particulate biomass and water to form a uniform and stable viscous paste. Such a mixture can then be gasied in a continuous ow system more eciently.172,173 Hao et al.173 found that the 10 wt% cellulose or sawdust sample mixed with CMC (23 wt%) can be completely gasied in supercritical water at 773 K, 27 MPa and for 20 min residence time in the presence of a Ru/C catalyst. And the experimental results suggested the catalytic activities in the order as follows: Ru/C 4 Pd/C 4 nano-(CeZr)xO2 4 nanoCeO2 4 CeO2. When acidic solid support, for example in a Pt/Al2O3 catalyst, was used, the catalytic gasication of lignocellulosic biomass could consist of two main steps: (1) Al2O3 with surface acidity could eectively act as a catalyst in the rst step for the depolymerization of a lignocellulosic biomass feedstock, similar to acid-catalyzed hydrolysis. As mentioned previously, in supercritical water this depolymerization can also occur to some extent even without catalysts. (2) The next step is the gasication of water-soluble depolymerized products over Pt/Al2O3 in a catalytic reactor, producing H2, CO2, and CH4 (eqn (11)).174,175 Similarly, in the case of Pt/C catalyst in conversion of cellulose into hydrogen by aqueous phase reforming process, it was proposed that this process might involve slow hydrolysis of cellulose to glucose which was catalyzed by the H+ reversibly formed in water during the reaction and followed by rapid reforming of glucose to H2.176 Cellulose ! C6 H12 O6 ! 6nC3n H2 O
H2 O 3nH2 O

+ Org

11

+ 2CO

(7) (8)

Ru2+ + 2H2O - RuO2 + 2H+ + H2


This journal is
c

Gasication of cellulose by the reforming reaction with H2O under high temperature can be regarded as catalytic wet oxidation.103,177
Chem. Soc. Rev., 2011, 40, 55885617 5605

The Royal Society of Chemistry 2011

View Online

Table 5 Catalyst

Some typical reaction conditions and results in catalytic gasication of cellulose or lignocellulosic biomass in the presence of steam Reactants Cellulose + H2O Sawdust + H2O Cellulose + H2O Cellulose + H2O Cellulose + H2O Cellulose + H2O Cellulose + H2O Wood Palm empty fruit bunch Reaction conditions 673 K, 15 min 1 MPa (argon gas) 773 K, 20 min 27 MPa (N2) 723 K, 2 h 44 MPa (argon gas) 623 K, 10 min 30 MPa (N2) 623 K, 3 h 30 MPa (N2) 1106 Ka, 60 ms 0.1 MPa (N2/O2 = 3.76) 623 K, 30 min or 1 h 18 MPa (N2) 973 K, in N2 10231063 K Air-blown bubbling Typical products CH4 H2 H2 CH4 H2 H2 CH4 H2 CH4 H2 CO CO2 + H2 + CH4 H2-rich gas Gas
b

Yield (%) 44.0 9.0 24.0 (wt%) 32.1 13.7b 31.3 9.2 (vol%) 42.3 7.7 (vol%) B52.0 B49.0 88.1 N/a 1.76 m3 gas kg1 biomass
b

Researchers/Year Osada et al./2004 Hao et al./2005 Izumizaki et al./2005 Usui et al./2000 Usui et al./2000 Dauenhauer et al./2007 Minowa et al./1999 Richardson et al./2010 Lahijani et al./2011

Ref. 171 173 170 178 178 188 192 168 155

Ru/TiO2 (2 wt%) Ru/C (5 wt%) RuO2 Pt/Al2O3 (5 wt%) Pd/Al2O3 (5 wt%) Rh catalysts

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Ni/SiO2Al2O3 (50 wt%) Ni NP


a

Temperature measured at 10 mm from the top of the catalytic bed.

L/per 100 g feedstock. NP: nanoparticles.

Usui et al.178 studied how cellulose is gasied in hot-compressed water at 623 K in the presence of a series of supported catalysts (2 wt% loading) using Zr(OH)4, (CH3COCHQC(O)CH3)3Fe, ferrocene, Ru3(CO)12, (CH3COCHQC(O)CH3)2Co, NiC2O4, NiO, Ni(OH)2, 2NiCO33Ni(OH)2, PdI2, CuCO3Cu(OH)2, (CH3CO2)2CuH2O, (CH3COCHQC(O)CH3)2Cu, and Cu(OH)2. Among these catalysts, after reaction for 3 h, 5 wt% Pd supported on Al2O3 particularly showed the highest catalytic activity, leading to a yield of H2 at 42.3 vol% and CH4 at 7.7 vol%, respectively (Table 5). A possible two-step mechanism is also proposed. Namely, rst the C2C10 unit aldehydes, ketones, and organic acids were formed from cellulose in hot compressed water and then these intermediates were catalytically decomposed to give gaseous products mainly consisting of H2, CH4, CO, and CO2. It is noteworthy that using transition metal complexes as catalysts for the gasication of cellulose, oil and char were mainly produced with few gaseous products. Many studies have shown that the deactivation of a catalyst in gasication of lignocellulosic biomass is often related to the amount of coke deposition.179 In addition to catalyst deactivation, the formation of tar readily causes trouble in downstream equipment such as blocking and fouling. Therefore, a good catalyst should be multifunctional, including the conducive roles in further catalytic steam reforming tar to syngas and in the removal of the inactive biochar by catalytic combustion. For example, Miyazawa et al.180 demonstrated catalytic partial oxidation of tar over a multicomponent Rh/CeO2/SiO2 catalyst for syngas production from lignocellulosic biomass. In addition, Asadullah and coworkers181 conducted a series of experiments on Rh-based catalysts for gasication of cellulose and found that the Rh loading of 1.2 104 mol g1 catalyst supported on CeO2 exhibited multiple functions during the gasication of microcrystalline cellulose. The Rh/CeO2 catalysts remarkably increased the conversion and inhibited the formation of methane, thereby enhancing the yields of CO and H2, compared to Ru/CeO2, Pd/CeO2, Ni/CeO2, Rh/Al2O3 and Rh/TiO2. These studies suggested that Rh/CeO2 has great potential to be developed as a catalyst for complete gasication of cellulose.
5606 Chem. Soc. Rev., 2011, 40, 55885617

Finally, in addition to the catalysts and reaction variables, the reactor also plays a role in gasication eciency and product distribution.182186 Kaewluan et al.187 depicted a bubbling uidized bed gasier having a high carbon conversion eciency (97.3%) and gasication eciency (80.2%) to syngas from rubber wood chip. Dauenhauer and co-workers188 reported millisecond reforming, by which cellulose particles can be converted extremely quickly into syngas on a hot Rh surface without detectable deactivation from carbon formation. 6.2 Gasication over non-noble metal catalysts It is well-documented that Ni, Fe, Co, and Cu-based catalysts are eective for the water-gas shift reaction in which carbon monoxide reacts with water to form CO2 and hydrogen (eqn (12)).189 Therefore, when such kinds of solid catalysts exist in the gasication of cellulose in the presence of steam, the yield of hydrogen can be increased (eqn (13)), possibly following a model involving in the watergas shift reaction (eqn (12)).190 CO + H2O - CO2 + H2 Cellulose ! water-soluble products ! gasesH2 CO
Ni gasfication decomposition

(12)

13

! gasesCH4 CO2
Ni

methanation

However, the H2 is readily released from the liquid phase to the gas phase. As a result, the catalyst cannot be in contact with a large concentration of hydrogen. This might prevent the methanation reaction from occurrence although methanation is readily catalyzed by such types of catalysts (eqn (9) and (10)). Recently, Sinag et al.191 compared the eect of nanosized and bulky ZnO and SnO2 on the water-gas shift reaction in gasication of cellulose. The results showed that at 573 K the watergas shift reaction proceeded faster over ZnO catalysts than that over SnO2 catalysts. Therefore, a higher yield of hydrogen was obtained in the presence of ZnO.
This journal is
c

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

The methanation reaction could be also suppressed under subcritical and supercritical conditions. Minowa et al.192 reported that cellulose was gasied to hydrogen-rich gas over a nickel catalyst supported on kieselguhr in hot-compressed water at around 623 K and 18 MPa (subcritical conditions). During gasication, the steam reforming reaction and methanation reaction both occurred. The gas phase equilibrium composition was 4146 mol% CH4, 4749 mol% CO2 and 611 mol% H2. Recently, a series of catalysts such as nickel, iron, copper, zinc, zirconium, ruthenium and RANEYs nickel slurry were evaluated by Resende and Savage in the gasication of cellulose and lignin in supercritical water.193 The results revealed that the metallic catalysts played a critical role in the formation of gaseous product and the selectivity. The yield of hydrogen gas could be increased to some degree over nickel or copper by optimizing catalyst loading and the catalyst surface area/biomass weight ratio. However, it is worth noting that, in addition to the active component in a catalyst, usually the acidity and basicity of a support is also an inuential factor on product distribution and coke formation. Interestingly, Tasaka and co-workers disclosed that steam reforming of tar derived from cellulose gasication was eciently catalyzed by 12 wt% Co/MgO catalyst at 873 K in a uidized bed reactor. Over such a type of MgO supported catalysts, the conversion of tar was found to increase with the amount of Co loading.194,195 Those ndings indicate that the eciency of gasication of lignocellulosic biomass could be greatly enhanced by endowing the catalysts an additional function of in situ removal of tar by catalytic steam reforming.

supported metal catalysts have relatively large surface areas with well-dispersed active metal sites. In addition, metalsupport interactions might signicantly contribute to the activity, selectivity and stability of a catalyst. The introduction of metal ions to a catalyst support can also modulate the acidity/basicity on the surfaces, thereby possibly facilitating the degradation of lignocellulosic biomass and reducing carbon deposition by both the support and the active metal sites. Over these supported catalytic metal catalysts, the eciency of converting lignocellulosic biomass into hydrogenated products can be greatly enhanced through complicated reaction networks consisting of dehydration, dehydrogenation, hydrogenation, isomerization, oligomerization, decarbonylation and decarboxylation, and so forth. 7.1 Integrated hydroprocessing to produce commodity chemicals As summarized in Table 6, presently many researchers use cellulose as a model of lignocellulosic biomass for investigation of the integrated hydroprocessing with hydrolysis to produce commodity chemicals. A typical method is the combination of supported Pt, Pd or Ru catalysts for hydrogenation with liquid acid catalysts for hydrolysis. It oers a more ecient route to the conversion of cellulose in a one-pot process reaction in the aqueous phase to form C4 to C6 sugar alcohols, in particular sorbitol, mannitol, and even ethylene glycol (Scheme 10, Table 6). In particular, sorbitol can be transformed to fuels (hydrogen, light alkanes and liquid alkanes) by aqueous-phase reforming or to a number of high value-added chemicals by aqueous-phase hydrodeoxygenation.200 The type of noble metal and support, the type of acid, acid concentration and reaction time signicantly inuence cellulose conversion, carbon eciency and product distribution.201 Notably, the breakage of cellulose involves the cleavage of two types of COC bonds: the 1,4-b-glycosidic bond connecting two adjacent glucose monomers and the COC bond within the glucose ring. Scheme 11 illustrates that over a ruthenium nanocluster catalyst the cleavage of the 1,4 0 -b-glycosidic bond at a dierent position will result in two dierent products:202 Cleavage at the C1O bond (position a) leads to dehydrated sorbitol, while cleavage at the position b gives dehydroxyed glucose. Therefore, to design a catalyst for a precise COC cleavage is much conducive to the selectivity of the desired product. In addition, when noble metals are supported onto acidic solids, possibly there is no need for the use of liquid acid. For example, over Pt/g-Al2O3 catalysts, the surfaces of g-Al2O3 support have acidic sites, which can catalyze the hydrolysis of cellulose to glucose, and then the CQO group in the glucose molecule can be catalytically reduced by H2 over Pt species on the surface of g-Al2O3, thereby leading to the formation of hydrogenated products like sorbitol or mannitol.203,204 Nevertheless, g-Al2O3 itself is not acidic enough to eciently hydrolyze cellulose.118 The additional acidity can be produced from the interaction between Pt/Al2O3 and hydrogen.28 Under the reaction conditions H2 may undergo heterolytic dissociative chemisorption of hydrogen on platinum atoms. Moreover, hydrogen molecules activated on the platinum surface are able
Chem. Soc. Rev., 2011, 40, 55885617 5607

Integrated hydroprocessing

Because the dominant components in lignocellulosic biomass are oxygen-rich cellulosic carbohydrates, the deploymerization and degradation of lignocellulosic biomass by catalytic hydrolysis, solvolysis, liquefaction and pyrolysis primarily leads to oxygenated molecules with carboxyl, keto, and/or hydroxyl groups. Therefore, to yield targeted value-added chemicals and liquid hydrocarbons as transportation fuels, most oxygenated molecules should be further deoxygenated along with the formation of new CC bonds. In other words, an additional hydrogenation or hydroprocessing is needed.196 When hydrogen is present in the reaction system over multifunctional catalysts, the partial or thorough removal of oxygen from the oxygenated molecules and the formation of CC bonds by hydrogenation and aldol condensation, alkylation, or ketonization could occur simultaneously or in situ consecutively.197 In addition, the catalytic hydrogenolysis of the glucosic carbonoxygen (CO and CQO) bonds in cellulose and hemicelluloses and aromatic carbonoxygen (CO) bonds in lignin may also take place in the presence of hydrogen.198 Therefore, the technology of integrating hydroprocessing with hydrolysis, solvolysis, liquefaction and pyrolysis has increasingly received attention over the past decade. And it has shown great potential for the production of the commodity chemicals and fuels from lignocellulosic biomass.199 In the context of catalytic hydrogenation, supported metal catalysts, for example Pt, Ru, Pd or Ni catalysts conventionally possess superior catalytic properties. Moreover, these
This journal is
c

The Royal Society of Chemistry 2011

View Online

Table 6 Catalyst

Integrated hydrolysis with hydroprocessing in the conversion of cellulose over supported catalysts Reactants Cellulose + H2O + H2 Cellobiose + H2O + H2 Cellulose + H2O + H2 Cellulose + H2O + H2 Cellulose + H2O + H2 Cellulose + H2O + H2 Cellobiose + H2O + H2 Cellulose + H2O + H2 Cellulose + H2O + H2 Cellulose + H2O + H2 Cellulose + H2O + H2 Cellulose + H2O + H2 Reaction conditions 463 K, 24 h 5 MPa (H2) 393 K, 12 h 4 MPa (H2) 518 K, 30 min 6 MPa (H2) 498 K, 5 min 6 MPa (H2) + CH3OH 463 K 5MPa (H2) 463 K, 24 h 5 MPa (H2) 393 K, 12 h 4 MPa (H2) 518 K, 30 min 6 MPa (H2) 463 K, 24 h 5 MPa (H2) 423 K, 48 h 3.5 MPa (H2) + C4MIMCl 518 K, 30 min 6 MPa (H2) 463 K, 24 h 6 MPa (H2) Main products Sorbitol Sorbitol Glucose Hexitols EG 1,2-PD Polyol Sorbitol Sorbitol Mannitol Sorbitol Glucose Sorbitol Mannitol EG Sorbitol Glucose Glucose Sorbitol Sorbitol Mannitol EG Sorbitol Mannitol Erythritol EG Yield (%) 0.7 23.18 1.4 39.3 15% 14% 65 27 25 6 18.5 42.6 9.5 6.0 14.2 15 o1 26 74 3.9 1.9 61.0 50.0 6.2 12.8 2.5 Researchers/Year Fukuoka et al./2006 Yan et al./2006 Luo et al./2007 Deng et al./2010 Geboers et al./2010 Fukuoka et al./2006 Yan et al./2006 Ji et al./2008 Jolle et al./2009 Ignatyev et al./2010 Ji et al./2008 de Vyver et al./2010 Ref. 205 202 206 211 209 205 202 214 204 208 214 212

Ru/HUSY (2.5 wt%) Ru/PVP (1 wt%) Ru/C (4 wt%) Ru/C H4SiW12O40 and Ru/C catalysts Pt/g-Al2O3 (2.5 wt%)

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Pt/PVPb (1 wt%) Pt/Al2O3 (2.5 wt%) Pt/g-Al2O3 HRuCl(CO)(PPh3)3, Rh/C 2%Ni30%W2C/AC Ni/carbon ber

HUSY: H form of ultrastable Y zeolite. PVP: poly(Nvinyl-2-pyrrolidone). EG: ethylene glycol. PD: propanediol.

Scheme 10 Catalytic conversion of cellulose into polyols by integrated hydrolysis and hydrogenation.206,214 Adapted and reprinted with permission from ref. 206. Copyright 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim; Adapted and reprinted with permission from ref. 214. Copyright 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

to exchange with protons of the hydroxy groups on the surface of support Al2O3.203 These resultant H+ species can reversibly spill over onto the surface of the catalyst support.203,204 Consequently, more Brnsted acid sites formed on the support
5608 Chem. Soc. Rev., 2011, 40, 55885617

which can then catalyze the hydrolysis of cellulose to glucose, followed by metal-catalyzed hydrogenation. This is part of the reasons that the presence of Pt/g-Al2O3 was found to signicantly increase the initial rate of dissolution-conversion as well
This journal is
c

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 11 Cellulose structure and the potential monomers formed following cleavage of the COC bonds at positions a or b.202 Reprinted with permission from ref. 202. Copyright 2006 American Chemical Society.

as the selectivity to polyols, especially sorbitol.203,204 Moreover, in most cases, the hydrolysis reaction could be a rate-determining step.203207 Thus an ecient conversion of cellulose into polyols can be realized through a combination of hydrolysis and hydrogenation over such a class of bifunctional supported metal catalysts. In addition to Pt/g-Al2O3, on the surface of a carbonsupported Ru cluster (Ru/C) catalyst, H+ ions could also be reversibly formed in situ in hot water, thereby resulting in both catalytic hydrolysis and instantaneous hydrogenation.205,206 Moreover, integrated depolymerization of cellulose and hydrogenation can be carried out using ionic liquid C4MIMCl as a solvent in the presence of hydrogen.207 For example, Ignatyev et al.208 recently discovered that the combination of a heterogeneous metal catalyst and a homogeneous Ru catalyst proved to be eective in such a process in the ionic liquid, possibly because the Ru compound can enhance the transfer of hydrogen to the metallic surface. In addition, heteropoly acids together with supported noble metal catalysts exhibit good performances in the conversion of lignocellulosic biomass into valuable platform chemicals via combined hydrolysis and hydrogenation.209 More recently, Palkovits et al.210 showed that heteropoly acids together with supported Ru catalysts show not only high activity but also remarkable selectivity for sugar alcohols reaching up to 81% yield of C4 to C6 sugar alcohols at 433 K for 7 h. Noticeably, over supported Ru catalysts in the presence of basic solution, an integrated one-pot conversion of cellulose could be developed to produce valuable specic polyols such as ethylene glycol(EG), 1,2-propanediol, and 1,2,5-pentanetriol.211 Ni supported onto carbon is also an alternative catalyst for a one-pot process of catalytic cellulose to polyols.212 Interestingly, a series of bimetallic catalysts including Ru-W/ AC, Ir-W/AC, Pd-W/AC, Pt-W/AC, Ni-W/SiC, and Ni-W/ TiO2 proved to be highly active and selective for the formation of EG from cellulose.213 W was considered as a key component for degradation of cellulose, namely, the CC cracking reactions, while the transition metals were mainly responsible for the hydrogenation reactions of unsaturated intermediates. Recently, Ji et al.214 demonstrated that supported tungsten carbide catalysts can replace supported noble metal-based catalysts to catalyze the degradation of cellulose. Moreover, the use of Ni-W2C/AC catalysts showed remarkable selectivity towards EG with a yield up to 61 wt% after 30 min at 518 K and 6 MPa H2. In addition to integrated hydrolysis and
This journal is
c

hydrogenation, integrated aldol-condensation and hydrogenation of lignocellulosic biomass (corncobs) to produce water-soluble C5C15 compounds was recently developed in a single reactor system over Pd/WO3ZrO2 catalysts.215 The promising progress means that there is still plenty of scope to design more ecient catalysts for the one-pot catalytic conversion of cellulose into value-added polyols and other ne chemicals. 7.2 Integrated hydroprocessing to produce fuels Another major objective of integrating hydroprocessing is to convert lignocellulosic biomass to liquid hydrocarbons so that they can directly replace the present gasoline or diesel fuels in transportation applications. In this context, the processes of hydrogenating hydrolytic compounds and hydrogenating pyrolysis oil have been intensively investigated. Moreover, lignocellulosic biomass is used more often as a starting material, rather than pure cellulose model samples. Studies have shown the possibilities for yielding diesel and gasoline from lignocellulosic biomass by integrated hydrogenation with acid-catalyzed hydrolysis. Recently it was revealed, for example, that an integrated process can produce gasoline with high octane numbers from maple wood.216 Such a process involves dilute acid-catalyzed (H2SO4, oxalic acid) hydrolysis of maple wood into aqueous carbohydrate solutions at 433 K, followed by aqueous phase hydrodeoxygenation of the resulting sugar solutions. Thus it led to a product similar to gasoline with a carbon yield of up to 57% and an estimated octane number of 96.5. Because of the complicated compositions of hydrolytic products, hydrogenation deoxygenation of aqueous carbohydrate stream was realized by a two-stage selectively catalytic process: rst the catalyst bed contained a Ru/C catalyst at 393 K and the second catalyst bed contained a Pt/zirconium phosphate catalyst at 518 K. As discussed previously, pretreatment by organic solvent can be signicantly helpful to dissolve or solvolyze lignin, cellulose or hemicellulose in lignocellulosic biomass. Clearly, it is also conducive to the integrated process of hydrolysis and hydrogenation. It was demonstrated that the pretreatment of cellulose in 1-hexanol at 623 K, followed by hydrocracking catalyzed by Pt/HZSM-5(23) at 673 K in the presence of hydrogen, produced C2C9 alkanes with a remarkable yield up to 89%.217 Compounds other than hydrocarbons from lignocellulosic materials also have potential for use as fuels. For example, Lange and co-workers218 reported that over bifunctional
Chem. Soc. Rev., 2011, 40, 55885617 5609

The Royal Society of Chemistry 2011

View Online

catalysts Pt/TiO2 and Pt/ZSM-5 valeric esters as biofuels were produced from lignocellulosic materials. The synthesis of valeric biofuels consisted of acid hydrolysis of lignocellulosic materials to levulinic acid, hydrogenation of the levulinic acid to g-valerolactone (gVL) and valeric acid (VA), and its subsequent esterication to valeric biofuels (Scheme 12A). According to the proposed mechanism for the conversion of gVL to VA over bifunctional catalysts (Scheme 12B), the optimal production of VA requires a balancing of the acidic and hydrogenation functionalities of the bifunctional metal/ zeolite catalyst. Therefore, changing the ratio of metal to zeolite led to dierent intermediates and products. Low metal loading onto zeolite resulted in an increase in the coproduction of pentenoic acid. High metal loading favored the formation of methyl tetrahydrofuran (MTHF), pentanal/pentanol, and/or pentane/butane. In particular, it is noteworthy that valeric biofuels were found to be fully compatible for blending with gasoline or diesel. More recently, Serrano-Ruiz et al. described219,220 that the deconstruction of cellulose in an aqueous solution of H2SO4 yielded an equimolar mixture of levulinic acid and formic acid. The formic acid could then be decomposed into H2 and CO2 and thus H2 was consumed to reduce levulinic acid to GVL in the H2SO4 solution over a Ru/C catalyst. Moreover, the formation of GVL brought about easy separation and recycling of the used H2SO4 in the cellulose deconstruction reactor. Remarkably, in a single reactor by means of a dual catalyst bed of Pd/Nb2O5 and ceria-zirconia; levulinic acid and GVL could be catalytically upgraded to 5-nonanone and then liquid hydrocarbon fuels similar to diesel and gasoline. In addition to hydrotreating hydrolytic products, hydroprocessing of the pyrolysis oil of lignocellulosic biomass is also a promising option for the production of CH-rich fuels from lignocellulosic biomass. One example is the extraordinary nding by Vispute and co-workers documenting that the further hydroprocessing of the pyrolysis oil greatly increased its intrinsic hydrogen content, thereby leading to light olens and aromatic hydrocarbons.221 Over zeolite catalyst, the yield of aromatic hydrocarbons and light olens could be tuned through the amount of hydrogen added to the biomass feedstock during hydroprocessing. As discussed, fast pyrolysis of lignocellulosic biomass is a relatively economical way to produce pyrolysis oil. As such, an ecient integrated hydrotreating denitely oers remarkable potential for production of renewable liquid bio-based diesel and gasoline from lignocellulosic biomass.

Prospects and concluding remarks

The limitation in the availability and sustainability of fossil fuels such as coal, petroleum and natural gas, along with the negative ecological and environmental impacts of their uses compel many researchers to develop new environmentally friendly processes on the basis of renewable feedstocks.222 Lignocellulosic biomass, in particular plant cell wall-based cellulose, is the most abundant and bio-renewable resource on earth. It has been considered as one of the most attractive alternatives to replace fossil resources for the production of fuels, fuel additives and ne chemicals.
5610 Chem. Soc. Rev., 2011, 40, 55885617

Over the past decades, several catalytic systems, namely hydrolysis, liquefaction, gasication, pyrolysis and deoxygenation/hydrogenation for the catalytic conversion of lignocellulosic biomass have increasingly received attention. Generally, acid catalysts are primarily used in the catalytic hydrolysis of lignocellulosic biomass to produce glucose and xylose; metal supported catalysts and zeolites are primarily evaluated in the gasication of lignocellulosic biomass to produce H2 or syngas; and alkaline catalysts are dominantly investigated in the fast pyrolysis. Over multifunctional catalysts, the integrated hydroprocessing of lignocellulosic biomass specically lead to sorbitol and mannitol, EG or hydrocarbon-based diesel and gasoline. Theoretically, the scale-up of a chemocatalytic process of lignocellulosic biomass can produce fuels and chemicals in large quantities in a short reaction period. However, no matter what type of catalytic processes employed, present techniques have not yet met the required level for commercialization in an economical and environmentally stable fashion. The present liquefaction, gasication and pyrolysis of lignocellulosic biomass yield very complicated products in most cases. The upgrading and separation of products requires high energy consumption. The invention of a heterogeneously solid catalyst with extraordinary activity and selectivity remains an essential task of paramount importance. A paradoxical, challenging issue is that the heterogeneous reaction over solid catalysts commonly suers from low catalytic activity due to the solidsolid contact between the insoluble lignocellulosic biomass and solid catalyst. The utilization of ionic liquids as solvents and/or as catalysts provides an alternative solution. However, the use of the ionic liquids might bring about new problems such as separation and extraction of the targeted products and increased cost. Fast pyrolysis of lignocellulosic biomass has proven to be more ecient than liquefaction and gasication. It is expected that upgraded pyrolysis oils can act as fuels directly. However, provided that gasication processes can selectively produce syngas with cost-eective cleaning technology, it is also certainly of great commercial value because syngas can be converted to clean fuels and ne chemicals by currently available catalytic technologies, for example syngas to methanol and ethanol and then to hydrocarbons and olens.223 Also, it is much desired to directly gasify and pyrolyze wet lignocellulosic biomass because drying lignocellulosic biomass is an energy-intensive and time-consuming operation.224 In the short term, fossil crude oils are still available in large quantities. Therefore, the overall costs for production of commodity fuels and chemicals from lignocellulosic biomass processes by the present catalytic techniques are indeed less competitive in comparison with large-scales petroleum industries. In the long run, integrated production of fuels and ne commodity chemicals from lignocellulosic biomass feedstocks could be the most economical and renewable alternative in a sustainable society. To put the catalytic processes of lignocellulosic biomass into use, some key points to be addressed in the years ahead include: (1) In-depth understanding of the mechanism of the heterogeneously catalytic conversion of lignocellulosic biomass. It is necessary to identify the specic features of each component
This journal is
c

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

The Royal Society of Chemistry 2011

View Online

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

Scheme 12 (A) Reaction scheme and key performance factors for the individual process steps in the formation of valeric biofuels (S: selectivity, [mol%]; P: productivity, [tproductm3reactorh1]; and C: concentration, [wt%]). LA: levulinic acid; gVL: g-valerolactone; VA: valeric acid; EV: ethyl valerate; EG: ethylene glycol; PG: propylene glycol; IER: acidic ion-exchange resin. (B) Probable reaction mechanism for the conversion of gVL to VA over bifunctional catalysts. MTHF: methyl tetrahydrofuran; Pe: 1-pentanol; PV: Pentyl valerate; PeOH: 1-pentanol.218 Adapted and reprinted with permission from ref. 218. Copyright 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

such as cellulose, lignocelluloses and lignin in the reaction networks. It includes the interrelationship among the feedstock, the catalyst, the reaction conditions and the product distribution. In addition, the specic sources of lignocellulosic biomass to the desired products should also be established more clearly in each process.225 In-depth insights into the structure of cellulose and lignin biopolymers from the interdisciplinary sciences of plant and chemistry are surely useful to develop a strategy to convert it.226,227 (2) Strategic design and preparation of multifunctional catalysts for highly active and selective conversion of lignocellulosic biomass. This involves acidity, basicity, metal-support interactions, the shape of catalyst particles, mechanical and chemical stability, deactivation and reusability. Because of the peculiar features of lignocellulosic biomass and its degraded products, novel reactors and separation systems, including process intensication, for continuous one-pass or one-pot operation should be developed so as to reduce the cost, minimize waste disposal and energy consumption, and maximize the productivity.
This journal is
c

(3) Ecient production of liquid hydrocarbons from lignocellulosic biomass. Firstly, because liquid hydrocarbons can be directly used in current engines, there is no need for building new infrastructure for distribution. Secondly, if lignocellulosic biomass is to replace fossil oils mostly or even completely, similar to current petroleum industry, the large-scale process of lignocellulosic biomass should, for the most part, produce hydrocarbon-based liquid fuels. In addition to platform chemicals and ne chemicals, the identication of certain specic chemicals in the conversion of lignocellulosic biomass as a new type of fuels or fuel additives is also of much signicance.228,229 All in all, to make the catalytic conversion of lignocellulosic biomass into valuable products practicable is one of the most important topics in chemistry. It involves many challenges and opportunities. Though the studies on the chemocatalytic conversion of cellulose have received intensive attention over the past few decades and many breakthroughs have been achieved therein. A noteworthy thing is that besides cellulose, lignocellulosic biomass contains hemicellulose and lignin which are
Chem. Soc. Rev., 2011, 40, 55885617 5611

The Royal Society of Chemistry 2011

View Online

still less investigated. Moreover, the most workable and economical way could be to use lignocellulosic biomass directly, rather than pure cellulose. It might be right to say that the technology of catalytic conversion of lignocellulosic biomass to fuels and chemicals remains in its infancy. Recent advances have shown an exciting prospect of putting catalytic conversion of lignocellulosic biomass to good use. The next decades will certainly witness the use and growth of catalytic production of fuels and chemicals from lignocellulosic biomass with continual technological innovations.

Acknowledgements
Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

We thank anonymous reviewers for thoroughly reading the manuscript, providing thoughtful, valuable suggestions and comments on the early version of this manuscript, which almost lead to a new version. We would like to thank the Publishers for granting permissions to reproduce copyright materials cited or adapted in this manuscript. The results and illustrations are credited to the original researchers to whom we would like to express our gratitude. The authors wish to acknowledge the nancial support from the Distinguished Young Scholar Grants from the Natural Scientic Foundation of Zhejiang Province (R4100436), the National Natural Scientic Foundation of China (20773110, 20541002), Zhejiang 151 talents project, and Project (2009C14G2020021) from Science and Technology Department of Zhejiang Provincial Government for the related research and development. The authors wish to acknowledge the support of the School of Chemical Engineering and ARC Centre of Excellence for Functional Nanomaterials, AIBN in the University of Queensland during the preparation of this work.

References
1 M. A. Rubio Rodriguez, J. De Ruyck, P. Roque Diaz, V. K. Verma and S. Bram, An LCA based indicator for evaluation of alternative energy routes, Appl. Energy, 2011, 88(3), 630635. 2 W. P. Nel and C. J. Cooper, Implications of fossil fuel constraints on economic growth and global warming, Energy Policy, 2009, 37(1), 166180. 3 S. Shaee and E. Topal, When will fossil fuel reserves be diminished?, Energy Policy, 2009, 37(1), 181189. 4 G. W. Huber, S. Iborra and A. Corma, Synthesis of transportation fuels from biomass: Chemistry, catalysts, and engineering, Chem. Rev., 2006, 106(9), 40444098. 5 C. H. C. Zhou, J. N. Beltramini, Y. X. Fan and G. Q. (Max) Lu, Chemoselective catalytic conversion of glycerol as a biorenewable source to valuable commodity chemicals, Chem. Soc. Rev., 2008, 37(3), 527549. 6 Y. X. Fan, C. H. Zhou and X. H. Zhu, Selective catalysis of lactic acid to produce commodity chemicals, Catal. Rev. Sci. Eng., 2009, 51(3), 293324. 7 C. H. Zhou, J. N. Beltramini, C. X. Lin, Z. P. Xu, G. Q. M. Lu and A. Tanksale, Selective oxidation of biorenewable glycerol with molecular oxygen over Cu-containing layered double hydroxide-based catalysts, Catal. Sci. Technol., 2011, 1(1), 111122. 8 Denition of Biomass, in Biomass Gasication and Pyrolysis, ed. P. Basu, Elsevier, 2011, pp. 325326. 9 C. Somerville, S. Bauer, G. Brininstool, M. Facette, T. Hamann, J. Milne, E. Osborne, A. Paredez, S. Persson, T. Raab, S. Vorwerk and H. Youngs, Toward a system approach to understanding plant cell walls, Science, 2004, 306(5705), 22062211.

10 M. Carrier, A. Loppinet-Serani, D. Denux, J.-M. Lasnier, F. Ham-Pichavant, F. Cansell and C. Aymonier, Thermogravimetric analysis as a new method to determine the lignocellulosic composition of biomass, Biomass Bioenergy, 2011, 35(1), 298307. 11 C. Somerville, H. Youngs, C. Taylor, S. C. Davis and S. P. Long, Feedstocks for lignocellulosic biofuels, Science, 2010, 329(5993), 790791. 12 E. Taarning, C. M. Osmundsen, X. B. Yang, B. Voss, S. I. Andersen and C. H. Christensen, Zeolite-catalyzed biomass conversion to fuels and chemicals, Energy Environ. Sci., 2011, 4(3), 793804. 13 W. Boerjan, J. Ralph and M. Baucher, Lignin biosynthesis, Annu. Rev. Plant Biol., 2003, 54, 519546. 14 L. Shuai and X. J. Pan, Biohydrocarbons: next generation liquid fuels from lignocellulosic biomass, Research Progress In Paper Industry And Biorenery (4th Isetpp), 2010, 13, 12931297. 15 L. Jimenez, A. Rodriguez, A. Perez, A. Moral and L. Serrano, Alternative raw materials and pulping process using clean technologies, Ind. Crops Prod., 2008, 28(1), 1116. 16 Wheat straw as a paper ber source. p2-1, Clean Washington Center, Washington, 1997. http://www.cwc.org/paper/pa971rpt. pdf. 17 L. Jimenez, F. Lopez and C. Martnez, Paper from sorghum stalks, Holzforschung, 1993, 47(6), 529533. 18 E. Maekawa, Studies on Hemicellulose of Bamboo, Wood Research: bulletin of the Wood Research Institute Kyoto University, 1976, 59/60, 153179. 19 N. Cordeiro, M. N. Belgacem, I. C. Torres and J. C. V. P. Mourad, Chemical composition and pulping of banana pseudostems, Ind. Crops Prod., 2004, 19(2), 147154. 20 Y. S. Sun, X. B. Lu, S. T. Zhang, R. Zhang and X. Y. Wang, Kinetic study for Fe(NO3)3 catalyzed hemicellulose hydrolysis of dierent corn stover silages, Bioresour. Technol., 2011, 102(3), 29362942. 21 M. P. Pandey and C. S. Kim, Lignin depolymerization and conversion: a review of thermochemical methods, Chem. Eng. Technol., 2011, 34(1), 2941. 22 H. D. Willauer, J. G. Huddleston, M. Li and R. D. Rogers, Investigation of aqueous biphasic systems for the separation of lignins from cellulose in the paper pulping process, J. Chromatogr., Biomed. Appl., 2000, 743(12), 127135. 23 S. R. Collinson and W. Thielemans, The catalytic oxidation of biomass to new materials focusing on starch, cellulose and lignin, Coord. Chem. Rev., 2010, 254(1516), 18541870. 24 D. Klemm, B. Heublein, H. Fink and A. Bohn, Cellulose: fascinating biopolymer and sustainable raw material, Angew. Chem., Int. Ed., 2005, 44(22), 33583393. 25 S. Van De Vyver, J. Geboers, P. A. Jacobs and B. F. Sels, Recent advances in the catalytic conversion of cellulose, ChemCatChem, 2011, 3(1), 8294. 26 Y. Nishiyama, J. Sugiyama, H. Chanzy and P. Langan, Crystal structure and hydrogen bonding system in cellulose 1(alpha), from synchrotron X-ray and neutron ber diraction, J. Am. Chem. Soc., 2003, 125(47), 1430014306. 27 D. Mohan, C. U. Pittman and P. H. Steele, Pyrolysis of wood/ biomass for bio-oil: A critical review, Energy Fuels, 2006, 20(3), 848889. 28 R. F. Tester, J. Karkalas and X. Qi, Starchcomposition, ne structure and architecture, J. Cereal Sci., 2004, 39(2), 151165. 29 A. E. Farrell, R. J. Plevin, B. T. Turner, A. D. Jones, M. OHare and D. M. Kammen, Ethanol can contribute to energy and environmental goals, Science, 2006, 311(5760), 506508. 30 R. F. Service, Is there a road ahead for cellulosic ethanol?, Science, 2010, 329(5993), 784785. 31 M. R. Schmer, K. P. Vogel, R. B. Mitchell and R. K. Perrin, Net energy of cellulosic ethanol from switchgrass, Proc. Natl. Acad. Sci. U. S. A., 2008, 105(2), 464469. 32 H. Danner and R. Braun, Biotechnology for the production of commodity chemicals from biomass, Chem. Soc. Rev., 1999, 28(6), 395405. 33 J. J. Bozell and G. R. Petersen, Technology development for the production of biobased products from biorenery carbohydrates the US Department of Energys Top 10 revisited, Green Chem., 2010, 12(4), 539554.

5612

Chem. Soc. Rev., 2011, 40, 55885617

This journal is

The Royal Society of Chemistry 2011

View Online

34 K. C. Kwon, H. Mayeld, T. Marolla, B. Nichols and M. Mashburn, Catalytic deoxygenation of liquid biomass for hydrocarbon fuels, Renewable Energy, 2011, 36(3), 907915. 35 L. Ge, P. Wang and H. Mou, Study on saccharication techniques of seaweed wastes for the transformation of ethanol, Renewable Energy, 2011, 36(1), 8489. 36 C. D. Risbrudt, R. J. Ross, J. J. Blankenburg and C. A. Nelson, Forest products laboratory-Supporting the nations armed forces with valuable wood research for 90 years, Forest Products Journal, 2007, 57(12), 614. 37 E. Princi, S. Vicini, E. Pedemonte, G. Gentile, M. Cocca and E. Martuscelli, Synthesis and mechanical characterisation of cellulose based textiles grafted with acrylic monomers, European Polymer Journal, 2006, 42(1), 5160. 38 Y. P. Zhang and L. R. Lynd, Toward an aggregated understanding of enzymatic hydrolysis of cellulose: Noncomplexed cellulase systems, Biotechnol. Bioeng., 2004, 88(7), 797824. 39 M. Pedersen and A. S. Meyer, Lignocellulose pretreatment severity-relating pH to biomatrix opening, New Biotechnol., 2010, 27(6), 739750. 40 R. Bharadwaj, A. Wong, B. Knierim, S. Singh, B. M. Holmes, M. Auer, B. A. Simmons, P. D. Adams and A. K. Singh, High-throughput enzymatic hydrolysis of lignocellulosic biomass viain situ regeneration, Bioresour. Technol., 2011, 102(2), 13291337. 41 M. Sto cker, Biofuels and biomass-to-liquid fuels in the biorenery: catalytic conversion of lignocellulosic biomass using porous materials, Angew. Chem., Int. Ed., 2008, 47(48), 92009211. 42 A. J. Ragauskas, C. K. Williams, B. H. Davison, G. Britovsek, J. Cairney, C. A. Eckert, W. J. Frederick, J. P. Hallett, D. J. Leak, C. L. Liotta, J. R. Mielenz, R. Murphy, R. Templer and T. Tschaplinski, The path forward for biofuels and biomaterials, Science, 2006, 311(5760), 484489. 43 R. R. Davda and J. A. Dumesic, Renewable hydrogen by aqueous-phase reforming of glucose, Chem. Commun., 2004, 3637. 44 R. R. Davda, J. W. Shabaker, G. W. Huber, R. D. Cortright and J. A. Dumesic, A review of catalytic issues and process conditions for renewable hydrogen and alkanes by aqueous-phase reforming of oxygenated hydrocarbons over supported metal catalysts, Appl. Catal., B, 2005, 56(12), 171186. 45 N. Yan, C. Zhao, W. J. Gan and Y. Kou, Polyols: A new generation of energy platform?, Chin. J. Catal., 2006, 27(12), 11591163. 46 B. Girisuta, L. Janssen and H. J. Heeres, A kinetic study on the conversion of glucose to levulinic acid, Chem. Eng. Res. Des., 2006, 84(A5), 339349. 47 M. Sasaki, T. Adschiri and K. Arai, Production of cellulose II from native cellulose by near- and supercritical water solubilization, J. Agric. Food Chem., 2003, 51(18), 53765381. 48 K. Grodzka, A. Maciejec and K. Krygier, Attempts to apply microcrystalline cellulose as a fat replacer in low fat mayonnaise emulsions, Zywnosc, Poland, 2005, 12(1), 5261. 49 H. Zhao, J. H. Kwak, Y. Wang, J. A. Franz, J. M. White and J. E. Holladay, Eects of crystallinity on dilute acid hydrolysis of cellulose by cellulose ballmilling study, Energy Fuels, 2006, 20(2), 807811. 50 R. W. Torget, J. S. Kim and Y. Y. Lee, Fundamental aspects of dilute acid hydrolysis/fractionation kinetics of hardwood carbohydrates. 1. Cellulose hydrolysis, Ind. Eng. Chem. Res., 2000, 39(8), 28172825. 51 Q. Xiang, J. S. Kim and Y. Y. Lee, A comprehensive kinetic model for diluteacid hydrolysis of cellulose, Appl. Biochem. Biotechnol., 2003, 106(13), 337352. 52 Z. Jia and D. Shen, Eect of reaction temperature and reaction time on the preparation of low-molecular-weight chitosan using phosphoric acid, Carbohydr. Polym., 2002, 49(4), 393396. 53 X. S. Zhuang, S. R. Wang, H. An, Z. Y. Luo and K. F. Cen, Cellulose hydrolysis research for liquid fuel production under low concentration acids, J. Zhejiang Univ., 2006, 40(6), 9971001. 54 C. N. Hamelinck, G. V. Hooijdonk and A. P. C. Faaij, Ethanol from lignocellulosic biomass: techno-economic performance in short-, middle- and long-term, Biomass Bioenergy, 2005, 28(4), 384410. 55 B. Rivas, J. M. Dominguez, H. Dominguez and J. C. Parajo, Bioconversion of posthydrolysed autohydrolysis liquors: an

56 57 58

59

60 61 62 63 64 65 66 67

68 69

70 71

72 73

74 75

76 77

alternative for xylitol production from corn cobs, Enzyme Microb. Technol., 2002, 31(4), 431438. R. F. Hu, L. Lin, T. J. Liu and S. J. Liu, Dilute sulfuric acid hydrolysis of sugar maple wood extract at atmospheric pressure, Bioresour. Technol., 2010, 101(10), 35863594. A. Emmel, A. L. Mathias, F. Wypych and L. P. Ramos, Fractionation of Eucalyptus grandis chips by dilute acid-catalysed steam explosion, Bioresour. Technol., 2003, 86(2), 105115. I. De Bari, F. Nanna and G. Braccio, SO2-catalyzed steam Fractionation of aspen chips for bioethanol production: Optimization of the catalyst impregnation, Ind. Eng. Chem. Res., 2007, 46(23), 77117720. A. Liu, Y. Park, Z. L. Huang, B. W. Wang, R. O. Ankumah and P. K. Biswas, Product identication and distribution from hydrothermal conversion of walnut shells, Energy Fuels, 2006, 20(2), 446454. P. Lenihan, A. Orozco, E. ONeill, M. N. M. Ahmad, D. W. Rooney and G. M. Walker, Dilute acid hydrolysis of lignocellulosic biomass, Chem. Eng. J., 2010, 156(2), 395403. F. S. Asghari and H. Yoshida, Conversion of Japanese red pine wood (Pinus densiora) into valuable chemicals under subcritical water conditions, Carbohydr. Res., 2010, 345(1), 124131. M. Matsunaga and H. Matsui, Super-rapid chemical conversion of Sugi wood by Supercritical and subcritical water treatment, Mokuzai Gakkaishi, 2004, 50(5), 325332. C. Schacht, C. Zetzl and G. Brunner, From plant materials to ethanol by means of supercritical uid technology, J. Supercrit. Fluids, 2008, 46(3), 299321. T. Rogalinski, K. Liu, T. Albrecht and G. Brunner, Hydrolysis kinetics of biopolymers in subcritical water, J. Supercrit. Fluids, 2008, 46(3), 335341. J. H. Lin, Y. H. Chang and Y. H. Hsu, Degradation of cotton cellulose treated with hydrochloric acid either in water or in ethanol, Food Hydrocolloids, 2009, 23(6), 15481553. Y. Sun, J. P. Zhuang, L. Lin and P. K. Ouyang, Clean conversion of cellulose into fermentable glucose, Biotechnol. Adv., 2009, 27(5), 625632. T. vom Stein, P. Grande, F. Sibilla, U. Commandeur, R. Fischer, W. Leitner and P. D. de Maria, Salt-assisted organic-acidcatalyzed depolymerization of cellulose, Green Chem., 2010, 12(10), 18441849. J. B. Binder and R. T. Raines, Fermentable sugars by chemical hydrolysis of biomass, Proc. Natl. Acad. Sci. U. S. A., 2010, 107(10), 45164521. Y. Yu and H. W. Wu, Signicant Dierences in the Hydrolysis Behavior of Amorphous and Crystalline Portions within Microcrystalline Cellulose in Hot-Compressed Water, Ind. Eng. Chem. Res., 2010, 49(8), 39023909. M. Misono, Acid catalysts for clean production. Green aspects of heteropolyacid catalysts, Comptes Rendus De Lacademie Des Sciences Serie Fascicule C-Chimie, 2000, 3(6), 471475. (a) J. H. Clark, Solid acids for green chemistry, Acc. Chem. Res., 2002, 35(9), 791797; (b) C. H. Zhou, An overview on strategies towards clay-based designer catalysts for green and sustainable catalysis, Appl. Clay Sci., 2011, 53(2), 8796. T. Okuhara, Water-tolerant solid acid catalysts, Chem. Rev., 2002, 102(10), 36413665. S. Van de Vyver, L. Peng, J. Geboers, H. Schepers, F. de Clippel, C. J. Gommes, B. Goderis, P. A. Jacobs and B. F. Sels, Sulfonated silica/carbon nanocomposites as novel catalysts for hydrolysis of cellulose to glucose, Green Chem., 2010, 12(9), 15601563. A. Abbadi, K. F. Gotlieb and H. van Bekkum, Study on solid acid catalyzed hydrolysis of maltose and related polysaccharides, Starch/Staerke, 1998, 50(1), 2328. P. L. Dhepe, M. Ohashi, S. Inagaki, M. Ichikawa and A. Fukuoka, Hydrolysisof sugars catalyzed by water-tolerant sulfonated mesoporous silicas, Catal. Lett., 2005, 102(34), 163169. J. Hegner, K. C. Pereira, B. De Boef and B. L. Lucht, Conversion of cellulose to glucose and levulinic acid via solid-supported acid catalysis, Tetrahedron Lett., 2010, 51(17), 23562358. Q. Xiang, Y. Y. Lee, P. O. Petterson and R. Torget, Heterogeneous aspects of acid hydrolysis of a-celluloses, Appl. Biochem. Biotechnol., 2003, 105, 505514.

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

This journal is

The Royal Society of Chemistry 2011

Chem. Soc. Rev., 2011, 40, 55885617

5613

View Online

78 D. Yamaguchi, M. Kitano, S. Suganuma, K. Nakajima, H. Kato and M. Hara, Hydrolysis of cellulose by a solid acid catalyst under optimal reaction conditions, J. Phys. Chem. C, 2009, 113(8), 31813188. 79 A. Onda, T. Ochi and K. Yanagisawa, Selective hydrolysis of cellulose into glucose over solid acid catalysts, Green Chem., 2008, 10(10), 10331037. 80 J. Hayashi, A. Kazehaya, K. Muroyama and A. P. Watkinson, Preparation of activated carbon from lignin by chemical activation, Carbon, 2000, 38(13), 18731878. 81 Y. Y. Wu, Z. H. Fu, D. L. Yin, Q. Xu, F. L. Liu, C. L. Lu and L. Q. Mao, Microwave-assisted hydrolysis of crystalline cellulose catalyzed by biomass char sulfonic acids, Green Chem., 2010, 12(4), 696700. 82 S. Suganuma, K. Nakajima, M. Kitano, D. Yamaguchi, H. Kato, S. Hayashi and M. Hara, Hydrolysis of cellulose by amorphous carbon bearing SO3H, COOH, and OH groups, J. Am. Chem. Soc., 2008, 130(38), 1278712793. 83 J. Pang, A. Wang, M. Zheng and T. Zhang, Hydrolysis of cellulose into glucose over carbons sulfonated at elevated temperatures, Chem. Commun., 2010, 46(37), 69356937. 84 A. Chareonlimkun, V. Champreda, A. Shotipruk and N. Laosiripojana, Reactions of C5 and C6-sugars, cellulose, and lignocellulose under hot compressed water (HCW) in the presence of heterogeneous acid catalysts, Fuel, 2010, 89(10), 28732880. 85 K. Shimizu, H. Furukawa, N. Kobayashi, Y. Itaya and A. Satsuma, Eects of Brnsted and Lewis acidities on activity and selectivity of heteropolyacid-based catalysts for hydrolysis of cellobiose and cellulose, Green Chem., 2009, 11(10), 16271632. 86 J. Tian, J. H. Wang, S. Zhao, C. Y. Jiang, X. Zhang and X. H. Wang, Hydrolysis of cellulose by the heteropoly acid H3PW12O40, Cellulose, 2010, 17(3), 587594. 87 C. Tagusagawa, A. Takagaki, S. Hayashi and K. Domen, Ecient Utilization of nanospace of layered transition metal oxide HNbMoO6 as a strong, water-tolerant solid acid catalyst, J. Am. Chem. Soc., 2008, 130(23), 72307231. 88 A. Takagaki, C. Tagusagawa and K. Domen, Glucose production from saccharides using layered transition metal oxide and exfoliated nanosheets as a water-tolerant solid acid catalyst, Chem. Commun., 2008, 42, 53635365. 89 P. Wang, H. B. Yu, S. H. Zhan and S. Q. Wang, Catalytic hydrolysis of lignocellulosic biomass into 5-hydroxymethylfurfural in ionic liquid, Bioresour. Technol., 2011, 102(5), 41794183. 90 R. Rinaldi, R. Palkovits and F. Schuth, Depolymerization of cellulose using solid catalysts in ionic liquids, Angew. Chem., Int. Ed., 2008, 47(42), 80478050. 91 C. H. Zhou, Emerging trends and challenges in synthetic claybased materials and layered double hydroxides, Appl. Clay Sci., 2010, 48(12), 14. 92 N. Villandier and A. Corma, One pot catalytic conversion of cellulose into biodegradable surfactants, Chem. Commun., 2010, 46(24), 44084410. 93 I. A. Ignatyev, P. G. N. Mertens, C. Van Doorslaer, K. Binnemans and D. E. de Vos, Cellulose conversion into alkylglycosides in the ionic liquid 1-butyl-3-methylimidazolium chloride, Green Chem., 2010, 12(10), 17901795. 94 W. P. Deng, M. Liu, Q. H. Zhang, X. S. Tan and Y. Wang, Acidcatalysed direct transformation of cellulose into methyl glucosides in methanol at moderate temperatures, Chem. Commun., 2010, 46(15), 26682670. 95 S. Cheng, I. Dcruz, M. Wang, M. Leitch and C. Xu, Highly ecient liquefaction of woody biomass in hot-compressed alcohol-water co-solvents, Energy Fuels, 2010, 24, 46594667. 96 T. Yamada, M. Aratani, S. Kubo and H. Ono, Chemical analysis of the product in acid-catalyzed solvolysis of cellulose using polyethylene glycol and ethylene carbonate, J. Wood Sci., 2007, 53(6), 487493. 97 S. P. Fan, S. Zakaria, C. H. Chia, F. Jamaluddin, S. Nabihah, T.-K. Liew and F.-L. Pua, Comparative studies of products obtained from solvolysis liquefaction of oil palm empty fruit bunch bres using dierent solvents, Bioresour. Technol., 2011, 102(3), 35213526. 98 M. Mascal and E. B. Nikitin, High-yield conversion of plant biomass into the key value-added feedstocks 5-(hydroxymethyl)furfural, levulinic acid, and levulinic esters via 5-(chloromethyl)furfural, Green Chem., 2010, 12(3), 370373.

99 J. Liu, R. Takada, S. Karita, T. Watanabe, Y. Honda and T. Watanabe, Microwave-assisted pretreatment of recalcitrant softwood in aqueous glycerol, Bioresour. Technol., 2010, 101(23), 93559360. 100 M. Kunaver, E. Jasiukaityte and V. Tisler, Liquefaction of lignocellulosic biomass and its potential applications, J. Biotechnol., 2007, 131S, S23S23. 101 D. W. Rackemann and W. O. S. Doherty, The conversion of lignocellulosics to levulinic acid, Biofuels, Bioprod. Bioren., 2011, 5(2), 198214. 102 X. H. Qi, M. Watanabe, T. M. Aida and R. L. Smith, Catalytical conversion of fructose and glucose into 5-hydroxymethylfurfural in hot compressed water by microwave heating, Catal. Commun., 2008, 9(13), 22442249. 103 B. D. Schutt and M. A. Abraham, Evaluation of a monolith reactor for the catalytic wet oxidation of cellulose, Chem. Eng. J., 2004, 103(13), 7788. 104 L. Peng, L. Lin, J. Zhang, J. Zhuang, B. Zhang and Y. Gong, Catalytic Conversion of cellulose to levulinic acid by metal chlorides, Molecules, 2010, 15, 52585272. 105 Z. H. Zhang and Z. B. K. Zhao, Microwave-assisted conversion of lignocellulosic biomass into furans in ionic liquid, Bioresour. Technol., 2010, 101, 11111114. 106 C. Z. Li, Z. H. Zhang and Z. B. Zhao, Direct conversion of glucose and cellulose to 5-hydroxymethylfurfural in ionic liquid under microwave irradiation, Tetrahedron Lett., 2009, 50(38), 54035405. 107 Y. Su, H. M. Brown, X. W. Huang, X. D. Zhou, J. E. Amonette and Z. C. Zhang, Single-step conversion of cellulose to 5-hydroxymethylfurfural (HMF), a versatile platform chemical, Appl. Catal., A, 2009, 361(12), 117122. 108 K. Seri, T. Sakaki, M. Shibata, Y. Inoue and H. Ishida, Lanthanum(III)-catalyzed degradation of cellulose at 250 1C, Bioresour. Technol., 2002, 81(3), 257260. 109 L. Z. Kong, G. M. Li, H. Wang, W. Z. He and F. Ling, Hydrothermal catalytic conversion of biomass for lactic acid production, J. Chem. Technol. Biotechnol., 2008, 83(3), 383388. 110 E. Meszaros, E. Jakab, G. Varhegyi, P. Szepesvary and B. Marosvolgyi, Comparative study of the thermal behavior of wood and bark of young shoots obtained from an energy plantation, J. Anal. Appl. Pyrolysis, 2004, 72(2), 317328. 111 C. J. Knill and J. F. Kennedy, Degradation of cellulose under alkaline conditions, Carbohydr. Polym., 2003, 51(3), 281300. 112 D. L. Geatches, S. J. Clarkand and H. C. Greenwell, Role of clay minerals in oil-forming reactions, J. Phys. Chem. A, 2010, 114(10), 35693575. 113 P. D. Jesus, N. Boukis, B. Kraushaar-Czarnetzki and E. Dinjus, Gasication of corn and clover grass in supercritical water, Fuel, 2006, 85(78), 10321038. 114 T. Minowa, F. Zhen, T. Ogi and G. Varhegyi, Liquefaction of cellulose in hot compressed water using sodium carbonate: Products distribution at dierent reaction temperatures, J. Chem. Eng. Jpn., 1997, 30(1), 186190. 115 S. Karagoz, T. Bhaskar, A. Muto, Y. Sakata and M. A. Uddin, Low-temperature hydrothermal treatment of biomass: eect of reaction parameters on products and boiling point distributions, Energy Fuels, 2004, 18(1), 234241. 116 P. Szabo, T. Minowa and T. Ogi, Liquefaction of cellulose in hot compressed water using sodium carbonate-part 2: Thermogravimetric mass spectrometric study of the solid residues, J. Chem. Eng. Jpn., 1998, 31(4), 571576. 117 S. Karagoz, T. Bhaskar, A. Muto, Y. Sakata, T. Oshiki and T. Kishimoto, Low-temperature catalytic hydrothermal treatment of wood biomass: analysis of liquid products, Chem. Eng. J., 2005, 108(12), 127137. 118 S. Karagoz, T. Bhaskar, A. Muto and Y. Sakata, Hydrothermal upgrading of biomass: Eect of K2CO3 concentration and biomass/water ratio on products distribution, Bioresour. Technol., 2006, 97(1), 9098. 119 J. Akhtar, S. K. Kuang and N. S. Amin, Liquefaction of empty palm fruit bunch (EPFB) in alkaline hot compressed water, Renewable Energy, 2010, 35(6), 12201227. 120 A. Demirbas, Yields of oil products from thermochemical biomass conversion processes, Energy Convers. Manage., 1998, 39(7), 685690.

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

5614

Chem. Soc. Rev., 2011, 40, 55885617

This journal is

The Royal Society of Chemistry 2011

View Online

121 Z. Fang, T. Minowa, R. L. Smith, T. Ogi and J. A. Kozinski, Liquefaction and gasication of cellulose with Na2CO3 and Ni in subcritical water at 350 1C, Ind. Eng. Chem. Res., 2004, 43(10), 24542463. 122 Y. J. Qian, C. J. Zuo, H. Tan and J. H. He, Structural analysis of bio-oils from sub-and supercritical water liquefaction of woody biomass, Energy, 2007, 32(3), 196202. 123 C. C. Song, H. Q. Hu, S. W. Zhu, G. Wang and G. H. Chen, Nonisothermal catalytic liquefaction of corn stalk in subcritical and supercritical water, Energy Fuels, 2004, 18(1), 9096. 124 S. Kumar and R. B. Gupta, Hydrolysis of microcrystalline cellulose in subcritical and supercritical water in a continuous ow reactor, Ind. Eng. Chem. Res., 2008, 47(23), 93219329. 125 A. Hammerschmidt, N. Boukis, E. Hauer, U. Galla, E. Dinjus, B. Hitzmann, T. Larsen and S. D. Nygaard, Catalytic conversion of waste biomass by hydrothermal treatment, Fuel, 2011, 90(2), 555562. 126 M. Tymchyshyn and C. B. Xu, Liquefaction of bio-mass in hotcompressed water for the production of phenolic compounds, Bioresour. Technol., 2010, 101(7), 24832490. 127 S. M. Kadangode, Lignin conversion into reformulated hydrocarbon and partially oxygenated gasoline compositions, PhD Dissertation, Department of Chemical and Fuels Engineering, The University of Utah, 2001. 128 J. E. Miller, L. Evans, A. Littlewolf and D. E. Trudell, Batch microreactor studies of lignin and lignin model compound depolymerization by bases in alcohol solvents, Fuel, 1999, 78(11), 13631366. 129 A. E. Putun, N. Ozbay, E. A. Varol, B. B. Uzun and F. Ates, Rapid and slow pyrolysis of pistachio shell:eect of pyrolysis conditions on the product yields and characterization of the liquid product, Int. J. Energy Res., 2007, 31(5), 506514. 130 C. Di Blasi, C. Branca and A. Galgano, Biomass screening for the production of furfural via thermal decomposition, Ind. Eng. Chem. Res., 2010, 49(6), 26582671. 131 M. M. Wright, D. E. Daugaard, J. A. Satrio and R. C. Brown, Techno-economic analysis of biomass fast pyrolysis to transportation fuels, Fuel, 2010, 89, S11S19. 132 M. W. Nolte and M. W. Liberatore, Viscosity of Biomass Pyrolysis Oils from Various Feedstocks, Energy Fuels, 2010, 24, 66016608. 133 R. P. Anex, A. Aden, F. K. Kazi, J. Fortman, R. M. Swanson, M. M. Wright, J. A. Satrio, R. C. Brown, D. E. Daugaard, A. Platon, G. Kothandaraman, D. D. Hsu and A. Dutta, Technoeconomic comparison of biomass-to-transportation fuels via pyrolysis, gasication, and biochemical pathways, Fuel, 2010, 89, S29S35. 134 M. Garcia-Perez, A. Chaala, H. Pakdel, D. Kretschmer and C. Roy, Vacuum pyrolysis of softwood and hardwood biomassComparison between product yields and bio-oil properties, J. Anal. Appl. Pyrolysis, 2007, 78(1), 104116. 135 J. Piskorz, P. Majerski, D. Radlein, A. Vladars-Usas and D. S. Scott, Flash pyrolysis of cellulose for production of anhydro-oligomers, J. Anal. Appl. Pyrolysis, 2000, 56(2), 145166. 136 S. Karagoez, Energy production from the pyrolysis of waste biomasses, Int. J. Energy Res., 2009, 33(6), 576581. 137 Z. Y. Luo, S. R. Wang, Y. F. Liao and K. F. Cen, Mechanism Study of Cellulose Rapid Pyrolysis, Ind. Eng. Chem. Res., 2004, 43(18), 56055610. 138 R. Lanza, D. D. Nogare and P. Canu, Gas phase chemistry in cellulose fast pyrolysis, Ind. Eng. Chem. Res., 2009, 48(3), 13911399. 139 X. L. Zhuang, H. X. Zhang, J. Z. Yang and H. Y. Qi, Preparation of levoglucosan by pyrolysis of cellulose and its citric acid fermentation, Bioresour. Technol., 2001, 79(1), 6366. 140 D. Lv, M. Xu, X. Liu, Z. Zhan, Z. Li and H. Yao, Eect of cellulose, lignin, alkali and alkaline earth metallic species on biomass pyrolysis and gasication, Fuel Process. Technol., 2010, 91(8), 903909. 141 Q. A. Lu, C. Q. Dong, X. M. Zhang, H. Y. Tian, Y. P. Yang and X. F. Zhu, Selective fast pyrolysis of biomass impregnated with ZnCl2 to produce furfural: Analytical Py-GC/MS study, J. Anal. Appl. Pyrolysis, 2011, 90(2), 204212. 142 G. Dobele, G. Rossinskaja, T. Diazbite, G. Telysheva, D. Meier and O. Faix, Application of catalysts for obtaining

143

144 145

146

147

148 149 150 151 152 153 154 155 156

157

158

159 160

161 162 163 164

1,6-anhydrosaccharides from cellulose and wood by fast pyrolysis, J. Anal. Appl. Pyrolysis, 2005, 74(12), 401405. Q. Lu, W. M. Xiong, W. Z. Li, Q. X. Guo and X. F. Zhu, Catalytic pyrolysis of cellulose with sulfated metal oxides: A promising method for obtaining high yield of light furan compounds, Bioresour. Technol., 2009, 100(20), 48714876. H. Kawamoto, S. Saito, W. Hatanaka and S. Saka, Catalytic pyrolysis of cellulose in sulfolane with some acidic catalysts, J. Wood Sci., 2007, 53(2), 127133. H. Kawamoto, W. Hatanaka and S. Saka, Thermochemical conversion of cellulose in polar solvent (sulfolane) into levoglucosan and other low molecular-weight substances, J. Anal. Appl. Pyrolysis, 2003, 70(2), 303313. M. Q. Chen, J. Wang, M. X. Zhang, M. G. Chen, X. F. Zhu, F. F. Min and Z. C. Tan, Catalytic eects of eight inorganic additives on pyrolysis of pine wood sawdust by microwave heating, J. Anal. Appl. Pyrolysis, 2008, 82(1), 145150. C. Di Blasi, C. Branca and A. Galgano, Thermal and catalytic decomposition of wood impregnated with sulfur- and phosphoruscontaining ammonium salts, Polym. Degrad. Stab., 2008, 93(2), 335346. N. Shimada, H. Kawamoto and S. Saka, Dierent action of alkali/alkaline earth metal chlorides on cellulose pyrolysis, J. Anal. Appl. Pyrolysis, 2008, 81(1), 8087. A. Pattiya, J. O. Titiloye and A. V. Bridgwater, Fast pyrolysis of cassava rhizome in the presence of catalysts, J. Anal. Appl. Pyrolysis, 2008, 81(1), 7279. M. Olazar, R. Aguado, J. Bilbao and A. Barona, Pyrolysis of sawdust in a conical spouted-bed reactor with a HZSM-5 catalyst, AIChE J., 2000, 46(5), 10251033. R. French and S. Czernik, Catalytic pyrolysis of biomass for biofuels production, Fuel Process. Technol., 2010, 91(1), 2532. T. R. Carlson, T. P. Vispute and G. W. Huber, Green gasoline by catalytic fast pyrolysis of solid biomass derived compounds, ChemSusChem, 2008, 1(5), 397400. P. T. Williams and J. Onwudili, Composition of products from the supercritical water gasication of glucose:a model biomass compound, Ind. Eng. Chem. Res., 2005, 44(23), 87398749. J. J. Saastamoinen, R. Taipale, M. Horttanainen and P. Sarkomaa, Propagation of the ignition front in beds of wood particles, Combust. Flame, 2000, 123(12), 214226. P. Lahijani and Z. A Zainal, Gasication of palm empty fruit bunch in a bubbling uidized bed: A performance and agglomeration study, Bioresour. Technol., 2011, 102(2), 20682076. A. A. Lappas, D. K. Iatridis and I. A. Vasalos, Production of Liquid Biofuels in a Fluid Catalytic cracking pilot-plant unit using waxes produced from a biomass-to-liquid (btl) process, Ind. Eng. Chem. Res., 2011, 50(2), 531538. M. Watanabe, H. Inomata, M. Osada, T. Sato, T. Adschiri and K. Arai, Catalytic eects of NaOH and ZrO2 for partial oxidative gasication of n-hexadecane and lignin in supercritical water, Fuel, 2003, 82(5), 545552. N. Nipattummakul, I. I. Ahmed, S. Kerdsuwan and A. K. Gupta, Hydrogen and syngas production from sewage sludge via steam gasication, Int. J. Hydrogen Energy, 2010, 35(21), 1173811745. C. Di Blasi, Combustion and gasication rates of lignocellulosic chars, Prog. Energy Combust. Sci., 2009, 35(2), 121140. X. H. Hao, L. J. Guo, X. Mao, X. M. Zhang and X. J. Chen, Hydrogen production from glucose used as a model compound of biomass gasied in supercritical water, Int. J. Hydrogen Energy, 2003, 28(1), 5564. A. J. Byrd, K. K. Pant and R. B. Gupta, Hydrogen production from glucose using Ru/Al2O3 catalyst in supercritical water, Ind. Eng. Chem. Res., 2007, 46(11), 35743579. A. J. Byrd, K. K. Pant and R. B. Gupta, Hydrogen production from ethanol by reforming in Supercritical water using Ru/Al2O3 catalyst, Energy Fuels, 2007, 21(6), 35413547. S. Su, W. Li, Z. Bai, H. Xiang and J. Bai, Eects of ionic catalysis on hydrogen production by the steam gasication of cellulose, Int. J. Hydrogen Energy, 2010, 35(10), 44594465. A. Tanksale, C. H. Zhou, J. N. Beltramini and G. Q. Lu, Hydrogen production by aqueous phase reforming of sorbitol using bimetallic NiPt catalysts: metal support interaction, J. Incl. Phenom. Macrocycl. Chem., 2009, 65(12), 8388.

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

This journal is

The Royal Society of Chemistry 2011

Chem. Soc. Rev., 2011, 40, 55885617

5615

View Online

165 Y. Izumizaki, K. C. Park, T. Yamamura, H. Tomiyasu, B. Goda and Y. Fujii, Exothermic hydrogen production system in supercritical water from biomass and usual domestic wastes with an exploitation of RuO2 catalyst, Prog. Nucl. Energy, 2008, 50(26), 438442. 166 A. Baiker, Supercritical uids in heterogeneous catalysis, Chem. Rev., 1999, 99(2), 453473. 167 C. Fushimi, S. Katayama and A. Tsutsumi, Elucidation of interaction among cellulose, lignin and xylan during tar and gas evolution in steam gasication, J. Anal. Appl. Pyrolysis, 2009, 86(1), 8289. 168 Y. Richardson, J. Blin, G. Vollea, J. Motuzas and A. Julbe, In situ generation of Ni metal nanoparticles as catalyst for H2-rich syngas production from biomass gasication, Appl. Catal., A, 2010, 382(2), 220230. 169 K. C. Park and H. Tomiyasu, Gasication reaction of organic compounds catalyzed by RuO2 in supercritical water, Chem. Commun., 2003, 6, 694695. 170 Y. Izumizaki, K. C. Park, Y. Tachibana, H. Tomiyasu and Y. Fujii, Organic decomposition in supercritical water by an aid of ruthenium(IV) oxide as a catalyst-Exploitation of biomass resources for hydrogen production, Prog. Nucl. Energy, 2005, 47(14), 544552. 171 M. Osada, T. Sato, M. Watanabe, T. Adschiri and K. Arai, Lowtemperature catalytic gasication of lignin and cellulose with a ruthenium catalyst in supercritical water, Energy Fuels, 2004, 18(2), 327333. 172 M. Zhao, N. H. Florin and A. T. Harrisa, Mesoporous supported cobalt catalysts for enhanced hydrogen production during cellulose decomposition, Appl. Catal., B, 2010, 97(12), 142150. 173 X. H. Hao, L. J. Guo, X. M. Zhang and Y. Guan, Hydrogen production from catalytic gasication of cellulose in supercritical water, Chem. Eng. J., 2005, 110(13), 5765. 174 R. Hashaikeh, Z. Fang, I. S. Butler and J. A. Kozinski, Sequential hydrothermal gasication of biomass to hydrogen, Proc. Combust. Inst., 2004, 30(2), 22312237. 175 R. Dolan, S. Yin and Z. Tan, Eects of headspace fraction and aqueous alkalinity on subcritical hydrothermal gasication of cellulose, Int. J. Hydrogen Energy, 2010, 35(13), 66006610. 176 N. Taccardi, D. Assenbaum, M. E. M. Berger, A. Boesmann, F. Enzenberger, R. Woelfel, S. Neuendorf, V. Goeke, N. Schoedel, H. J. Maass, H. Kistenmacher and P. Wasserscheid, Catalytic production of hydrogen from glucose and other carbohydrates under exceptionally mild reaction conditions, Green Chem., 2010, 12(7), 11501156. 177 B. D. Schutt, B. Serano, R. L. Cerro and M. A. Abraham, Production of chemicals from cellulose and biomass-derived compounds through catalytic sub-critical water oxidation in a monolith reactor, Biomass Bioenergy, 2002, 22(5), 365375. 178 Y. Usui, T. Minowa, S. Inoue and T. Ogi, Selective hydrogen production from cellulose at low temperature catalyzed by supported group 10 metal, Chem. Lett., 2000, 10, 11661167. 179 T. Yoshida and Y. Matsumura, Gasication of cellulose, xylan, and lignin mixtures in supercritical water, Ind. Eng. Chem. Res., 2001, 40(23), 54695474. 180 T. Miyazawa, T. Kimura, J. Nishikawa, K. Kunimori and K. Tomishige, Catalytic properties of Rh/CeO2/SiO2 for synthesis gas production from biomass by catalytic partial oxidation of tar, Sci. Technol. Adv. Mater., 2005, 6(6), 604614. 181 M. Asadullah, K. Fujimoto and K. Tomishige, Catalytic performance of Rh/CeO2 in the gasication of cellulose to synthesis gas at low temperature, Ind. Eng. Chem. Res., 2001, 40(25), 58945900. 182 Z. A. B. Z. Alauddin, P. Lahijani, M. Mohammadi and A. R. Mohamed, Gasication of lignocellulosic biomass in uidized beds for renewable energy development: A review, Renewable Sustainable Energy Rev., 2010, 14(9), 2852286. 183 M. Asadullah, S. Ito, K. Kunimori, M. Yamada and K. Tomishige, Energy ecient production of hydrogen and syngas from biomass: Development of low-temperature catalytic process for cellulose gasication, Environ. Sci. Technol., 2002, 36(20), 44764481. 184 K. Tomishige, M. Asadullah, S. I. Ito and K. Kunimori, Highly ecient production of synthesis gas by catalytic gasication of

185 186

187 188 189

190 191 192 193 194 195 196

197 198 199 200

201

202

203

204

biomass at low reaction temperature, Kagaku Kogaku Ronbunshu, 2002, 28(6), 666672. K. Tomishige, M. Asadullah and K. Kunimori, Novel catalysts for gasication of biomass with high conversion eciency, Catal. Surv. Asia, 2003, 7(4), 219233. M. Asadullah, T. Miyazawa, S. Ito, K. Kunimori and K. Tomishige, Demonstration of real biomass gasication drastically promoted by eective catalyst, Appl. Catal., A, 2003, 246(1), 103116. S. Kaewluan and S. Pipatmanomai, Potential of synthesis gas production from rubber wood chip gasication in a bubbling uidised bed gasier, Energy Convers. Manage., 2011, 52, 7584. P. J. Dauenhauer, B. J. Dreyer, N. J. Degenstein and L. D. Schmidt, Millisecond reforming of solid biomass for sustainable fuels, Angew. Chem., Int. Ed., 2007, 46(31), 58645867. N. Schumacher, A. Boisen, S. Dahl, A. A. Gokhale, S. Kandoi, L. C. Grabow, J. A. Dumesic, M. Mavrikakis and I. Chorkendor, Trends in low-temperature water-gas shift reactivity on transition metals, J. Catal., 2005, 229(2), 265275. T. Minowa and T. Ogi, Hydrogen production from cellulose using a reduced nickel catalyst, Catal. Today, 1998, 45(14), 411416. A. Sinag, T. Yumak, V. Balci and A. Kruse, Catalytic hydrothermal conversion of cellulose over SnO2 and ZnO nanoparticle catalysts, J. Supercrit. Fluids, 2011, 56(2), 179185. T. Minowa and S. Inoue, Hydrogen production from biomass by catalytic gasication in hot compressed water, Renewable Energy, 1999, 16(14), 11141117. F. L. P. Resende and P. E. Savage, Eect of metals on supercritical water gasication of cellulose and lignin, Ind. Eng. Chem. Res., 2010, 49(6), 26942700. K. Tasaka, T. Furusawa and A. Tsutsumi, Steam gasication of cellulose with cobalt catalysts in a uidized bed reactor, Energy Fuels, 2007, 21(2), 590595. K. Tasaka, T. Furusawa and A. Tsutsumi, Biomass gasication in uidized bed reactor with Co catalyst, Chem. Eng. Sci., 2007, 62(1820), 55585563. A. Pinheiro, N. Dupassieux, D. Hudebine and C. Geantet, Impact of the presence of carbon monoxide and carbon dioxide on gas oil hydrotreatment: investigation on liquids from biomass cotreatment with petroleum cuts, Energy Fuels, 2011, 25, 804812. E. F. Iliopoulou, Review of CC coupling reactions in biomass exploitation processes, Curr. Org. Synth., 2010, 7(6), 587598. A. G. Sergeev and J. F. Hartwig, Selective, nickel-catalyzed hydrogenolysis of aryl ethers, Science, 2011, 332(6028), 439443. S. A. Rezzoug and R. Capart, Liquefaction of wood in two successive steps: solvolysis in ethylene-glycol and catalytic hydrotreatment, Appl. Energy, 2002, 72(34), 631644. H. Li, D. Yu, Y. Hu, P. Sun, J. Xia and H. Huang, Eect of preparation method on the structure and catalytic property of activated carbon supportednickel oxide catalysts, Carbon, 2010, 48(15), 45474555. R. Palkovits, K. Tajvidi, J. Procelewska, R. Rinaldia and A. Ruppert, Hydrogenolysis of cellulose combining mineral acids and hydrogenation Catalysts, Green Chem., 2010, 12(6), 972978. N. Yan, C. Zhao, C. Luo, P. J. Dyson, H. C. Liu and Y. Kou, One-step conversion of cellobiose to C6-alcohols using a ruthenium nanocluster catalyst, J. Am. Chem. Soc., 2006, 128(27), 87148715. (a) P. L. Dhepe and A. Fukuoka, Cracking of cellulose over supported metal catalysts, Catal. Surv. Asia, 2007, 11(4), 186191; (b) P. F. Aparicio, A. G. Ruiz and I. R. Ramos, Hydrogen adsorbed species at the metal/support interface on a Pt/Al2O3 catalyst, J. Chem. Soc., Faraday Trans., 1997, 93(19), 35633567; (c) F. Ahmed, Md. K. Alam, A. Suzuki, M. Koyama, H. Tsuboi, N. Hatakeyama, A. Endou, H. Takaba, C. A. Del Carpio, M. Kubo and Akira Miyamoto, Dynamics of Hydrogen Spillover on Pt/g-Al2O3 Catalyst Surface: A Quantum Chemical Molecular Dynamics Study, J. Phys. Chem. C, 2009, 113(35), 1567615683. (a) V. Jolle, F. Chambon, F. Rataboul, A. Cabiac, C. Pinel, E. Guillon and N. Essayem, Non-catalyzed and Pt/gammaAl2O3-catalyzed hydrothermal cellulose dissolution-conversion: inuence of the reaction parameters and analysis of the unreacted

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

5616

Chem. Soc. Rev., 2011, 40, 55885617

This journal is

The Royal Society of Chemistry 2011

View Online

205

206

207

208

209

210

211 212

213

214

215

cellulose, Green Chem., 2009, 11(12), 20522060; (b) V. Jollet, F. Chambon, F. Rataboul, A. Cabiac, C. Pinel, E. Guillon and N. Essayem, Non-catalyzed and Pt/g-Al2O3 catalyzed hydrothermal cellulose dissolution-conversion: inuence of the reaction parameters, Top. Catal., 2010, 53(1518), 12541257. (a) A. Fukuoka and P. L. Dhepe, Catalytic conversion of cellulose into sugar alcohols, Angew. Chem., Int. Ed., 2006, 45(31), 51615163; (b) X. W. Sha, M. T. Knippenberg, A. C. Cooper, G. P. Pez and H. S. Cheng, Dynamics of hydrogen spillover on carbon-based materials, J. Phys. Chem. C, 2008, 112(44), 1746517470. C. Luo, S. A. Wang and H. C. Liu, Cellulose conversion into polyols catalyzed by reversibly formed acids and supported ruthenium clusters in hot water, Angew. Chem., Int. Ed., 2007, 46(40), 76367639. R. P. Swatloski, S. K. Spear, J. D. Holbrey and R. D. Rogers, Dissolution of cellose with ionic liquids, J. Am. Chem. Soc., 2002, 124(18), 49744975. I. A. Ignatyev, C. Van Doorslaer, P. G. N. Mertens, K. Binnemans and D. E. De Vos, Reductive splitting of cellulose in the ionic liquid 1-butyl-3-methylimidazolium chloride, ChemSusChem, 2010, 3(1), 9196. J. Geboers, S. Van de Vyver, K. Carpentier, K. de Blochouse, P. A. Jacobs and B. F. Sels, Ecient catalytic conversion of concentrated cellulose feeds to hexitols with heteropoly acids and Ru on carbon, Chem. Commun., 2010, 46(20), 35773579. R. Palkovits, K. Tajvidi, A. M. Ruppert and J. Procelewska, Heteropoly acids as ecient acid catalysts in the one-step conversion of cellulose to sugar alcohols, Chem. Commun., 2011, 47(1), 576578. T. Y. Deng, J. Y Sun and H. C. Liu, Cellulose conversion to polyols on supported Ru catalysts in aqueous basic solution, Sci. China Chem., 2010, 53(7), 14761480. S. Van de Vyver, J. Geboers, M. Dusselier, H. Schepers, T. Vosch, L. Zhang, G. Van Tendeloo, P. A. Jacobs and B. F. Sels, Selective bifunctional catalytic conversion of cellulose over reshaped ni particles at the tip of carbon nanobers, ChemSusChem, 2010, 3(6), 698701. M. Y. Zheng, A. Q. Wang, N. Ji, J. F. Pang, X. D. Wang and T. Zhang, Transition metal-tungsten bimetallic catalysts for the conversion of cellulose into ethylene glycol, ChemSusChem, 2010, 3(1), 6366. N. Ji, T. Zhang, M. Y. Zheng, A. Q. Wang, H. Wang, X. D. Wang and J. G. Chen, Direct catalytic conversion of cellulose into ethylene glycol using nickel-promoted tungsten carbide catalysts, Angew. Chem., Int. Ed., 2008, 47(44), 85108513. W. Dedsuksophon, K. Faungnawakij, V. Champreda and N. Laosiripojana, Hydrolysis/dehydration/aldol-condensation/ hydrogenation of lignocellulosic biomass and biomass-derived

216

217 218

219 220 221

222 223 224 225 226

227 228 229

carbohydrates in the presence of Pd/WO3ZrO2 in a single reactor, Bioresour. Technol., 2011, 102(2), 20402046. N. Li, G. A. Tompsett, T. Y. Zhang, J. Shi, C. E. Wyman and G. W. Huber, Renewable gasoline from aqueous phase hydrodeoxygenation of aqueous sugar solutions prepared by hydrolysis of maple wood, Green Chem., 2011, 13(1), 91101. K. Murata, Y. Liu and M. Inaba, Hydrocracking of biomassderived materials into alkanes in the presence of platinum-based catalyst and hydrogen, Catal. Lett., 2010, 140(12), 813. J. P. Lange, R. Price, P. M. Ayoub, J. Louis, L. Petrus, L. Clarke and H. Gosselink, Valeric biofuels: a platform of cellulosic transportation fuels, Angew. Chem., Int. Ed., 2010, 49(26), 44794483. J. C. Serrano-Ruiz, D. J. Braden, R. M. West and J. A. Dumesic, Conversion of cellulose to hydrocarbon fuels by progressive removal of oxygen, Appl. Catal., B, 2010, 100(12), 184189. J. C. Serrano-Ruiz, D. Wang and J. A. Dumesic, Catalytic upgrading of levulinic acid to 5-nonanone, Green Chem., 2010, 12(4), 574577. T. P. Vispute, H. Zhang, A. Sanna, R. Xiao and G. W. Huber, Renewable chemical commodity feedstocks from integrated catalytic processing of pyrolysis oils, Science, 2010, 330(6008), 12221227. A. Cho, Infographic energy tricky tradeos, Science, 2010, 329(5993), 786787. J. He and W. N. Zhang, Techno-economic evaluation of thermo-chemical biomass-to-ethanol, Appl. Energy, 2011, 88(4), 12241232. n, M. Lauer and F. Verhoe, L. Fagerna s, J. Brammer, C. Wile Drying of biomass for second generation synfuel production, Biomass Bioenergy, 2010, 34(9), 12671277. C. Somerville, H. Youngs, C. Taylor, S. C. Davis and S. P. Long, Feedstocks for lignocellulosic biofuels, Science, 2010, 329(5993), 790792. Y. P. Zhou, H. Stuart-Williams, G. D. Farquhar and C. H. Hocart, The use of natural abundance stable isotopic ratios to indicate the presence of oxygen-containing chemical linkages between cellulose and lignin in plant cell walls, Phytochemistry, 2010, 7(89), 982993. S. Zabler, O. Paris, I. Burgert and P. Fratzl, Moisture changes in the plant cell wall force cellulose crystallites to deform, J. Struct. Biol., 2010, 171(2), 133141. J. R. Regalbuto, Cellulosic biofuels-got gasoline?, Science, 2009, 325(5942), 822824. E. L. Kunkes, D. A. Simonetti, R. M. West, J. C. Serrano-Ruiz, C. A. Gartner and J. A. Dumesic, Catalytic conversion of biomass to monofunctional hydrocarbons and targeted liquidfuel classes, Science, 2008, 322(5900), 417421.

Downloaded by University of Oxford on 21 October 2011 Published on 24 August 2011 on http://pubs.rsc.org | doi:10.1039/C1CS15124J

This journal is

The Royal Society of Chemistry 2011

Chem. Soc. Rev., 2011, 40, 55885617

5617

Anda mungkin juga menyukai