Utilization of Ionic Liquids in Lignocellulose Biorefineries as Agents for

Sep 3, 2015 - CITI (Centro de Investigación, Transferencia e Innovación), University of Vigo, Tecnopole, San Cibrao das Viñas, 32900 Ourense, Spain...
0 downloads 0 Views 581KB Size
Subscriber access provided by The University Library | University of Sheffield

Perspective

Utilization of ionic liquids in lignocellulose biorefineries as agents for separation, derivatization, fractionation or pretreatment Susana Peleteiro, Sandra Rivas, Jose L. Alonso, Valentín Santos, and Juan C. Parajo J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.5b03461 • Publication Date (Web): 03 Sep 2015 Downloaded from http://pubs.acs.org on September 8, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 35

Journal of Agricultural and Food Chemistry

1

Utilization of ionic liquids in lignocellulose biorefineries as agents for

2

separation, derivatization, fractionation or pretreatment

3 4

Susana Peleteiro,†,‡ Sandra Rivas,

†,‡

5

Parajó*,†,‡

6



7

Science. Polytechnical Building, As Lagoas, 32004 Ourense, Spain.

8



9

San Cibrao das Viñas, Ourense, Spain.

José L. Alonso,

†,‡

Valentín Santos†,‡ and Juan C.

Chemical Engineering Department. University of Vigo (Campus Ourense). Faculty of

CITI (Centro de Investigación, Transferencia e Innovación), University of Vigo, Tecnopole,

10 11

*Author to whom correspondence should be addressed (Phone: +34988387033, Fax.:

12

+34988387001; e-mail: [email protected])

13

14

ABSTRACT: Ionic liquids (ILs) can play multiple roles in lignocellulose biorefineries,

15

including utilization as agents for the separation of selected compounds, or as reaction media

16

for processing lignocellulosic materials (LCM). Imidazolium-based ILs have been proposed

17

for separating target components from LCM biorefinery streams, for example the dehydration

18

of ethanol-water mixtures, or the extractive separation of biofuels (ethanol, butanol) or lactic

19

acid from the respective fermentation broths. As in other industries, ILs are potentially

20

suitable for removing Volatile Organic Compounds or carbon dioxide from gaseous

21

biorefinery effluents. On the other hand, cellulose dissolution in ILs allows to carry out

22

homogeneous derivatization reactions, opening new ways for product design and/or for

23

improving the quality of the products. Imidazolium-based ILs are also suitable for processing

24

native LCM, allowing the integral benefit of the feedstocks via separation of polysaccharides 1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

25

and lignin. Even strongly lignified materials can yield celullose-enriched substrates highly

26

susceptible to enzymatic hydrolysis upon ILs processing. Recent developments in enzymatic

27

hydrolysis include the identification of ILs causing limited enzyme inhibition and the

28

utilization of enzymes with improved performance in the presence of ILs.

29 30

Keywords: Ionic liquids, lignocellulosic materials, biorefineries, pretreatment, fractionation

31

32

Introduction: lignocellulosic materials as feedstocks for the industry

33

The mankind is facing major challenges such as the increased pressure on supplies

34

caused by the growing global population, the anthropogenic climate change, the fast depletion

35

of fossil resources, the volatility of oil prices, and the geopolitical risks affecting the safe

36

supply of raw materials. In this context, the development of cost- and energy-efficient

37

processes for manufacturing renewable transportation fuels and chemicals to supplement or

38

replace those derived from petroleum is imperative.1 It has been suggested that the future

39

chemical industry should be based on novel routes based on renewable raw materials and

40

providing chemicals with similar or more advanced properties than the ones currently

41

produced.2

42

Environmental sustainability would be favoured by using widespread, renewable

43

resources as raw materials, following one of the guiding principles of Green Chemistry.3 In

44

quantitative terms, biomass offers the only sustainable alternative to fossil fuels as a source of

45

carbon for our chemical and material needs.4 Lignocellulosic materials (denoted LCM) stand

46

for the many types of vegetal biomass mainly made up of carbohydrate polymers (cellulose

47

and hemicelluloses) and polyphenol-based lignin.5 The plant cell walls found in 2

ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35

Journal of Agricultural and Food Chemistry

48

lignocellulosic biomass are complex structures difficult to break down, or deconstruct, into

49

their component polymers and monomers.6

50

Typical LCM suitable as feedstocks for chemical processing include softwoods,

51

hardwoods, dedicated energy crops, industrial byproducts (such as bran or bagasse),

52

agricultural residues and byproducts (such as straw, grasses, husks or shells), and some

53

municipal solid wastes (such as cardboard, waste paper and gardening residues). LCM present

54

favourable characteristics as sustainable and environmentally friendly sources of chemicals

55

and fuels,7 including:

56



low purchase cost,8

57



large availability and huge generation rate. Nature produces about 180 billion metric

58

tons of biomass/year, of which about 75% is in the form of carbohydrates.9 Based on

59

the global annual production of biomass (1 × 1011 tons), and on the specific energy

60

contents of biomass and crude oil, it has been concluded that in only one decade,

61

Earth’s plants can renew in the form of cellulose, hemicellulose, and lignin all of the

62

energy stored as conventional crude oil,10

63



secure supply, as LCM can be produced locally,11

64



low nitrogen and sulfur contents,

65



CO2- neutrality, since vegetal biomass comes from photosynthesis, and the CO2 from

66

biomass does not represent a net input to the amount of carbon already making part of

67

the carbon cycle,

68



no competition with the food chain.12

69

However, the use of sustainable feedstocks is not enough for sustainability, and must be

70

completed with the protection of the environment using greener methodologies.4

71

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

72

Strategies for the utilization of lignocellulosic materials: biorefineries

73

In general terms, the benefit of LCM can be accomplished according to two different

74

philosophies:

75



gasification), or

76 77

direct utilization as a whole (for example, by combustion, pyrolysis, liquefaction or



LCM “fractionation” (in which the major components of the feedstock are separated

78

into “fractions” made up of compounds with related properties) followed by separate

79

processing of each fraction for specific purposes.

80

LCM fractionation is the conceptual basis of the “lignocellulose biorefinery”, which has

81

been defined by the National Renewable Energy Laboratory (NREL) as ‘‘a facility that

82

integrates conversion processes and equipment to produce fuels, power and chemicals from

83

biomass’’. According to this definition, biorefineries are expected to work under a general

84

idea analogous to today’s petroleum refineries.13 The IEA Bioenergy Task 42 broadened the

85

biorefinery concept, proposing that biorefinery stands for “the sustainable processing of

86

biomass into a spectrum of bio-based products (food, feed, chemicals and/or materials) and

87

bioenergy (biofuels, power and/or heat)”, in a way that a biorefinery can be a concept, a

88

facility, a process, a plant, or even a cluster of facilities.14 Integrated biorefineries,

89

characterized by a reduced carbon footprint of the final products, can provide a sustainable

90

approach to valuable products that can also improve biomass processing economics as well as

91

environmental issues,14 and have been considered crucial for extracting the maximum value

92

from biomass.15 In the medium term, relatively small-scale biorefineries making use of local

93

or regional resources have been considered as the most favored method to introduce more

94

advanced (green, whole-crop, and lignocellulosic) biorefinery processes into the market.14

95

The main issue in LCM processing lies within its complex structure and chemical

96

composition, an important issue when envisioning new processes.16 4

ACS Paragon Plus Environment

Page 4 of 35

Page 5 of 35

Journal of Agricultural and Food Chemistry

97

Usually, LCM fractionation intends the selective separation of the structural

98

components (lignin, cellulose and hemicelluloses), which have to be recovered and further

99

processed to yield an array of tailored commercial products. A recent analysis of the potential

100

demand, the biomass availability and the energy efficiency led to the conclusion that the

101

production of chemicals from biomass is more beneficial than the production of transportation

102

fuels or electricity.17

103

Desirably, operation in biorefineries should follow the basic principles of Green

104

Chemistry, for example regarding an efficient utilization of raw materials, and avoiding the

105

utilization of wastes instead of performing end-of-pipe waste remediation,3 enabling the

106

production of end-products fulfilling the societal needs, and shortening the dependence on

107

fossil resources.18 In this way, a sustainable industrial and societal development can be

108

achieved by meeting the needs of the present generation without compromising the needs of

109

future generations to meet their own needs.19

110

On the other hand, the implementation of biorefineries has been identified as the most

111

promising route to the creation of a new domestic biobased industry (NREL, 2014), paving

112

the way for a future low carbon bioeconomy,20,21 which would be driven by clean, sustainable

113

environmental development, economic growth and green politics,22 and result in a highly

114

efficient and cost-effective processing of biological feedstocks to a range of bio-based

115

products.23 On this basis, the implementation of biorefineries could also contribute to the

116

revitalization of rural areas.24 There is a long way to go, since just about 5% of the European

117

economy was bio-based in 2010; whereas in the USA, about 12% of all products (excluding

118

energy use) in 2010 was originated from biomass.25

119 120

Ionic liquids in lignocellulose biorefineries

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

121

Increasing attention is being paid to the utilization of ionic liquids (ILs) in a variety of fields.

122

ILs are salts composed of large organic cations and inorganic or organic anions, with melting

123

points below 100 ºC, which can be designed to be liquid at room temperature.26 Their low

124

volatility (many ILs can be distilled at 200–300 ºC just under significantly reduced pressure

125

and at very low distillation rate)27 limits the losses in vapour phase. This is an important

126

advantage over the conventional solvents employed in industries, which can be responsible for

127

emissions of Volatile Organic Compounds (VOC), identified as a major source of waste,28,29

128

and show hazards related to inhalation and explosion.30 Additionally, some ILs present

129

catalytic activity. In this case, they show many of the advantages of both homogenous and

130

heterogeneous catalysts, such as high acid density, uniform catalytic active centers, easy

131

separation and recyclability.31

132

The most common ILs can be classified in four groups according to their cations:

133

quaternary ammonium ILs, N-alkylpyridinium ILs, N-alkyl-isoquinolinium ILs, and

134

imidazolium-based ILs.32 Among them, the latter group has received special attention, and is

135

the focus of this study. Figure 1 presents the general formula of 3-methylimidazolium-based

136

ILs ([mim]) employed for biomass processing, as well as the nomenclature employed in this

137

work for the various ILs.

138

ILs are considered as green chemicals,33,34 and offer a unique environment for

139

chemistry, biocatalysts, separation, material synthesis, and electrochemistry.35 However, ILs

140

can be used as more than just alternative green solvents, since they differ from molecular

141

solvents by their unique ionic character and their structure and organization, which can lead to

142

specific effects.27 Non-flammability, lack of toxicity and thermal and chemical stability are

143

usual ILs properties of special importance for application in biorefineries. An ideal ionic

144

liquid should be inexpensive, non-toxic, biodegradable and recyclable.3

6

ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

Journal of Agricultural and Food Chemistry

145

Thermally stable ILs can be employed as reaction media or catalysts to carry out

146

reactions at high temperatures using conventional equipment, in a way that their negligible

147

vapour pressure would facilitate the recovery of volatile reaction products. Due to the unique

148

solvent properties of ILs, these solvents promise advantages with regard to conversion and

149

selectivity, leading to energy savings when compared to conventional solvents.26 Additionally,

150

their properties can be tuned for specific chemical tasks36 by selecting the type of the

151

constituent cation and anion:37 for example, a number of ionic liquids show unique solvating

152

properties.38 The increasing knowledge on the chemical, physical and technical properties of

153

ILs is expected to allow a wider and more efficient utilization.

154 155

Potential applications of ILs in biorefineries

156

Based on the above information on the nature of LCM and properties of ILs, a number of

157

potential applications of ILs in lignocellulose biorefineries have been identified, including

158

their utilization as agents for separation, cellulose dissolution and derivatization, LCM

159

fractionation or LCM pretreatment (see next sections). The utilization of ILs as reaction media

160

and/or catalyst for the hydrolysis-dehydration of polysaccharides into furans is also an

161

interesting research subject, but it falls out of the scope of this article.

162 163

ILs as separation agents for compounds present in biorefinery streams

164

ILs find a number of applications in biorefineries, for example in processes dealing with the

165

production of second-generation biofuels (such as bioethanol or biobutanol), or short-chain

166

organic acids. In this kind of technologies, LCM are processed to hydrolyze polysaccharides

167

(cellulose and/or hemicelluloses) into sugars, which are fermented to yield the target products.

168

One of the major techno-economic challenges in biorefineries manufacturing second-

169

generation ethanol from LCM is the comparatively low ethanol content of the fermentation 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

170

broth. The ethanol concentration can be increased by ordinary distillation up to concentrations

171

near the azeotrope (95.5 wt% ethanol), but the separation becomes increasingly difficult when

172

approaching this threshold. On the other hand, fuel applications require almost anhydrous

173

ethanol, making the utilization of alternative separation technologies (typically, azeotropic or

174

extractive distillation) necessary. Both azeotropic and extractive distillations involve the

175

addition of a third component (entrainer) able to break the azeotrope. Imidazolium-based ionic

176

liquids have been successfully employed as separation agents for the dehydration of ethanol-

177

water mixtures,26,39,40 as they can greatly enhance the relative volatility of ethanol over water,

178

showing advantages derived from low viscosity, thermal stability, good solubility and lower

179

corrosiveness respect to the high melting salts suitable for the same purpose.40 In a

180

comparative evaluation of [bmim]BF4, [emim]BF4 and [bmim]Cl (see Figure 1 for

181

nomenclature), the ability to improve the separation was found to follow the order [bmim]Cl >

182

[emim][BF4] > [bmim][BF4], in a way that the two best ILs performed better than the

183

reference compound.39,41 Figueroa et al.41 simulated the ethanol recovery using a number of

184

ILs ([bmim]Cl, [emim]Cl, [emim]BF4, [bmim]Ac, [bmim]BF4, [bmim]N(CN)2, [hmim]Cl and

185

[bmim]mSO4), and concluded that, in the best cases, high purity ethanol could be obtained at

186

96 wt% recovery. In a related study, Meinsderma et al.42 proposed an energy-efficient process

187

for ethanol dehydration based on extractive distillation with [emim]N(CN)2 enabling the

188

production of 99.91% ethanol in facilities at the pilot scale. Other reported advantages of ILs

189

as separation agents include:40

190



ILs do not pollute the distillate due to their non-volatile character,

191



reduced heat duties are required because of their non-volatility, high selectivities and heat capacities,

192 193 194



ILs properties such as solubility, capacity, selectivity, viscosity and thermal stability) can be tailored, and 8

ACS Paragon Plus Environment

Page 8 of 35

Page 9 of 35

Journal of Agricultural and Food Chemistry

only one distillation column is required.

195



196

Alternatively, ILs can be used for ethanol extraction from aqueous solutions. Arlt et al.43

197

proposed a number of ILs (including imidazolium- and phosphonium- based ones) as potential

198

solvents for the extractive separation of close-boiling or azeotropic mixtures, including water–

199

ethanol. This approach was claimed to be superior to conventional extractive rectification in

200

terms of cost-effectiveness and exergetic aspects, as a result of the selectivity and the unusual

201

characteristic profile of the ionic liquids. Solubility data for ternary mixtures of ethanol, water

202

and a hydrophobic IL ([bmim]PF6, [hmim]PF6 or [omim]PF6) were reported by Swatloski et

203

al.,44 whereas [hmim]Tf2N was found to be suitable for separating ethanol from water, but

204

acceptable recovery rates could be only achieved using unreasonably high solvent/feed

205

ratios.45

206

tetradecyltrihexylphosphonium-based ILs obtained by combination with seven different

207

anions have been recently proposed as candidates for ethanol recovery from aqueous solutions

208

by liquid–liquid extraction.46 On the basis of equilibrium calculations, the authors suggested

209

the use of a single liquid–liquid extraction stage coupled to extractive fermentation as a

210

favourable solution to recover ethanol from dilute aqueous solutions, a possibility favored by

211

the comparatively low toxicity of some of the ILs assayed.

Based

on

ternary

liquid–liquid

equilibrium

studies,

a

number

of

212

Biobutanol (n-butanol of biological origin) is an attractive biofuel,47,48 with higher energy

213

density and lower volatility as compared to ethanol. Biobutanol production from LCM

214

presents a number of challenging technological problems, some of them related to the

215

complex composition of culture media, which contain large amounts of water, microbial cells

216

(usually, bacteria belonging to the Clostridium genus), non-polysaccharide byproducts

217

resulting from LCM hydrolysis, residual sugars and volatile fermentation co-products (such as

218

acetone and ethanol). Additionally, the fermentation products are toxic to cells, a fact limiting

219

the maximum achievable butanol concentration (which is typically within the 1-2 wt% range). 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

220

In a screening study, Knoshaug and Zhang49 found just two Lactobacillus strains able to

221

tolerate and grow in up to 3 wt% butanol, the same level achieved by a mutant Clostridium

222

biejerinckii strain.30 Distillation, the technology traditionally employed to separate the volatile

223

components present in this type of fermentation broths,50 involves a high steam

224

consumption.51 Alternative, energy-efficient separation methods have been proposed,

225

including adsorption-desorption, extraction and membrane-based technologies. ILs may find

226

application in these two latter technologies.

227

Very high boiling extractants have been recommended as extraction solvents for butanol

228

fermentation products, a purpose for which ILs are suitable,30 either by direct contact with the

229

broth inside the fermenter52 or by downstream processing. ILs containing anions such as PF6−

230

or Tf2N− are water immiscible and enable the formation of biphasic systems suitable for

231

extraction applications.53 In particular, ILs such as [hmim]PF6 or [bmim]Tf2N, together with

232

others bearing a tetracyanoborate anion (such as [dmim]TCB) have been considered for n-

233

butanol extraction;30,51,54 whereas the same purpose has been achieved using imidazolium-

234

based ILs with alkyl chains of varying length in combination with tetrafluoroborate or

235

trifluoromethanesulfonate anions.30 The hydrophobicity was correlated to the butanol

236

distribution coefficient between ILs and water, whereas the extraction efficiency and

237

selectivity were directly related to the polarity.30 Ha et al.55 studied eleven imidazolium-based

238

ionic liquids as extraction agents to recover butanol from aqueous media, and also found an

239

interrelationship between extraction efficiency, selectivity and polarity. The best results were

240

achieved with [omim]Tf2N, which showed high extraction efficiency and allowed more than

241

74% butanol recovery in a single extraction stage. Fadeev and Meagher56 employed

242

[bmim]PF6 and [omim]PF6 to extract butanol from aqueous solutions, and subjected the ionic

243

liquid-rich phase (simulating the composition in equilibrium with 1 wt% butanol – water

244

solution) to pervaporation through a commercial polydimethylsiloxane membrane, obtaining 10

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35

Journal of Agricultural and Food Chemistry

245

an IL - free permeate. Pervaporation with IL-supported membranes 51,57 has been employed to

246

reduce the IL demand and to avoid its direct contact with the production organisms, avoiding

247

possible toxic effects.58

248

Imidazolium-based ILs have also been proposed as separation agents for short-chain

249

organic acids, including lactic acid.59–62 The biotechnological production of lactic acid in

250

biorefineries by hydrolysis-fermentation of LCM is an interesting possibility, since this

251

compound is an important commodity chemical for the production of specialty chemicals such

252

as 2,3-pentanedione, acrylic acid, propionic acid, pyruvic acid and polylactic acid. Lactic acid

253

can be obtained at high yield by fermentation of LCM-derived pentoses or hexoses.63

254

Depending on the operational conditions, the lactic fermentation can be stereospecific, or it

255

may lead mixtures of the two isomers. Interestingly, polylactic acid stereocomplexes with

256

improved properties can be produced from pure optical isomers. Several 1-alkyl-3-

257

methylimidazolium hexafluorophosphates (including [bmim]PF6, [hmim]PF6 and [omim]PF6)

258

have been employed for the extractive fermentation of lactic acid, acting as a water-

259

immiscible phase (instead of conventional organic solvents) to which lactic acid can be

260

transferred in situ from the fermentation broth, avoiding cell inhibition.59,60 Although the

261

extractability of the organic acid in the pure ILs was poor, the addition of tri-n-butylphosphate

262

increased the extraction ability to a level similar to those of conventional organic solvents.

263

The relatively low toxicity of the considered ILs allowed the growth of the fermenting cells

264

(Lactobacillus), which were able to consume glucose and to produce lactate. A comparative

265

evaluation of the various ILs showed that the length of the alkyl substituent in the

266

imidazolium cation had little influence on cell survival.60

267

In biorefineries, ILs can play a number of general roles common to other industries, for

268

example in the removal of Volatile Organic Compounds (VOC) and hazardous pollutants

269

from effluents.64–66 ILs such as [bmim]PF6, [bmim]Tf2N, or [bmim]BF4 can be directly 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

270

employed in bioreactors, either for performing enzyme-mediated biocatalytic reactions67 or in

271

the presence of whole-cells, for example in VOC absorption-biodegradation systems,67,68

272

enabling the biotransformation of toxic substrates in the presence of an immiscible liquid

273

phase acting as a substrate reservoir.

274

The ability of ILs to selectively dissolve gases, together with their non-volatility, makes

275

them potentially for gas separations69 (for example, CO2), acting as conventional liquid phases

276

in absorbers or as components of supported liquid membranes.53,70,71 CO2 has relatively high

277

solubility in a number of imidazolium ILs, particularly in those containing the

278

bis(trifluoromethylsulfonyl)imide anion, whereas there is little difference in CO2 solubility

279

between ionic liquids possessing the tetrafluoroborate or hexafluorophosphate anion.69

280

Amine-functionalized imidazolium ionic liquids have been specifically designed for CO2

281

capture.72

282 283

ILs as agents for cellulose dissolution - regeneration and homogeneous derivatization

284

Cellulose is hardly soluble in many conventional solvents because of its strong intermolecular

285

hydrogen bonding. Traditional cellulose solvents, such as carbon disulfide, N-

286

methylmorpholine-N-oxide (NMMO), and mixtures of N, N- dimethylacetamide and lithium

287

chloride (DMAC/LiCl), show a number of practical drawbacks, including limited dissolving

288

capability, toxicity, high cost, difficult solvent recovery, uncontrollable side reaction, and

289

instability during cellulose processing,73 and/or the need for multi-step pretreatments followed

290

by prolonged stirring.74

291

The ability of some molten ILs (N-alkylpyridinium or N- arylpyridinium chloride

292

salts) for dissolving cellulose (in the presence of a nitrogen-containing base) is known since

293

the mid 1930s,75 but the interest in this topic was renewed when Swatloski et al.76 claimed the

294

ability of ILs made of a number cations (including alkyl-imidazolium-based ones) and anions 12

ACS Paragon Plus Environment

Page 12 of 35

Page 13 of 35

Journal of Agricultural and Food Chemistry

295

(including Cl−, I−, PF6−, BF4−, acetate and trifluoroacetate) for achieving cellulose dissolution.

296

Cellulose dissolution in ILs involves the interaction of hydroxyl groups in cellulose with both

297

the cation and anion in IL.77 The oxygen atoms of OH groups and IL anions act as electron

298

donors along the dissolution process, in a way that anions acting as strong electron donors (for

299

example, halogen and pseudohalogen ions) gave good cellulose dissolution results; whereas

300

hydrogen atoms of hydroxyl groups and IL cations act as electron acceptors. The review by

301

Wang et al.78 summarizes extensive literature information on ILs able to dissolve cellulose,

302

the respective cellulose solubilities and the suitable operational conditions.

303

After dissolution, cellulose can be regenerated by adding a miscible anti-solvent such

304

as water, alcohols, ethers or ketones. When an anti-solvent such as water is added to the

305

homogeneous IL-cellulose system, the IL ions form hydrogen bonds with water molecules,

306

and are displaced into the aqueous phase; whereas cellulose (which previously interacted with

307

IL) is expelled and rebuilt its intra- and inter- molecular hydrogen bonds, and is then

308

precipitated.79

309

Cellulose dissolution/ regeneration in ILs is affected by many factors, including the

310

type of anion and cation, the basicity and H-bonding capability of the anion, and the position

311

and length of side chain in cation.80 As representative examples, [bmim]Cl,76 [amim]Cl;73,81

312

and [emim]Ac82can dissolve cellulose at comparatively high concentrations. Based on their

313

physicochemical properties, ILs with strongly basic anions (such as formate, acetate, or

314

phosphate) have been cited as favorable candidates for cellulose dissolution under mild

315

operational conditions.73,74,83 The dissolved cellulose could be readily regenerated by anti-

316

solvent addition: for example, water, ethanol, or acetone have been employed to regenerate

317

cellulose from [bmim]Cl;84 whereas Zhang et al.81 employed water to regenerate cellulose

318

from [amim]Cl.

13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

319

The regenerated cellulose can be different from the native cellulose in terms of macro

320

and microstructure, crystallinity and even degree of polymerization, depending on the

321

processing conditions.79,85 In fact, cellulose regenerated from ILs was found to be essentially

322

amorphous and porous.38 Additionally, it can be noted that complete cellulose dissolution in

323

ILs provides the opportunity of working with new functional groups, opening new ways for

324

product design or enabling a better control of the reaction (for example, regarding the degree

325

of substitution).86 The manufacture of cellulose derivatives in homogenous systems by

326

esterification, etherification,73,86 acylation, carbanilation,87 carboxymethylation, succination,

327

phthalation,88 alkylation, silylation, and halogenation86 has been reported. Other interesting

328

examples of cellulose derivatization in ILs have been reported in related studies.89–94 On the

329

other hand, cellulose dissolution -regeneration enables the recovery of cellulose nanofibers, or

330

its utilization as films, powder, gels, or capsules.95

331 332

ILs as agents for LCM fractionation

333

The complicated chemical composition and morphology of native LCM, characterized by the

334

presence of both non-structural components and a crosslinked, tridimensional matrix of

335

polysaccharides and lignin, together with the crystalline nature of cellulose, are major

336

hindrances for processing. Considered as feedstocks for chemical and/or biotechnological

337

processess, LCM are recalcitrant substrates that must be deconstructed (or fractionated) to

338

achieve the separation of the their major components. This is a key step in the conversion of

339

lignocellulosic biomass into fuels and valuable chemicals and materials.96 LCM dissolution

340

may imply the hydrolysis of carbohydrate polymers, or the chemical disruption of the

341

lignocellulose composite.97 ILs or ILs mixtures (eventually in the presence of catalysts, water

342

and/or other components) have demonstrated great promise as efficient solvents for biomass

343

fractionation, via total or partial dissolution.98 The interaction of ILs with lignocellulose may 14

ACS Paragon Plus Environment

Page 14 of 35

Page 15 of 35

Journal of Agricultural and Food Chemistry

344

involve ionic, π–π, and hydrogen bonding interactions. LCM dissolution in ILs depends on a

345

number of factors, including the type of constituent cation and anion, the nature and previous

346

processing of biomass (for example, milling or drying), the characteristics of the substrate

347

(including composition, particle size distribution, moisture), the relative amount of LCM,

348

other operational conditions (including temperature heating profile, maximum temperature

349

and processing time), and the possible presence of additional components in the medium (for

350

example, water, catalysts or co-solvents). For example, DMSO has been used in combination

351

with [bmim]Cl or [amim]Cl to achieve the partial dissolution of lignified raw materials,96,99

352

whereas the utilization of two ILs ([emim]Ac in combination with [amim]Cl or [bmim]Cl) has

353

been proposed for the same purpose.100

354

Complete dissolution has been reported for LCM of low lignin content (such as

355

bagasse, straws, grasses, or husks), but also for recalcitrant substrates such as hardwoods or

356

softwoods. Concerning wood dissolution, [bmim]Cl and [emim]Ac have been identified as the

357

two most promising ionic liquids.101 Representative studies and reviews on LCM dissolution

358

using a number of imidazolium-based ionic liquids have been reported in the past few

359

years.37,38,88,96,97,99,102–112 Once biomass is dissolved, the various structural components can be

360

recovered using a number of operational strategies, many of them based on the precipitation of

361

cellulose (or cellulose-enriched solids) upon anti-solvent addition.97

362

An alternative to completely dissolving biomass is to selectively dissolve just a single

363

biomass component.12 This approach may involve hemicellulose hydrolysis or lignin extraction.

364

Delignification-based processes usually enable the recovery of polysaccharides (at least,

365

cellulose) as insoluble solids.26,113,114 Representative information on the delignification of

366

multiple LCM in media containing imidazolium-based ionic liquids (including [emim]Ac,

367

[emim]Ace, [emim]ABS, [emim][Cl], [bmim]Ac, [bmim]Ace, [bmim]mSO4, [bmim]Br,

368

[bmim]BF4, [bmim]PF6, [amim]Cl, and [hmim]Cl) has been summarized in recent 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

369

studies.37,97,100,113,115-118

370

The ideal technology for LCM biorefineries would be the one where the three major

371

components of the biomass were obtained at high yields and purities.119 However, the

372

recovery of hemicelluloses (the most easily hydrolyzable of the three main lignocellulosic

373

polymers) is challenging,37 as this fraction can be considerably depolymerized and dissolved

374

in ILs, and eventually decomposed (as it has been found in the processing of pine wood with

375

[Hmim]Cl),113 resulting in losses that increase the cost of downstream separation.110 While

376

delignification increases with high temperatures, such conditions promote hemicellulose

377

losses.120 On the other hand, treatments allowing delignification and hemicellulose

378

solubilization simultaneously may entail losses of cellulose as soluble hydrolysis products.114

379

In some dissolution-regeneration approaches, hemicellulose and cellulose can be obtained as

380

separated fractions from the regenerated material, using specific solvents to maximize the

381

fractionation process.37,121 The selective dissolution of hemicelluloses (up to 39% of the total)

382

from extracted or native spruce wood has been addressed in literature using switchable ionic

383

liquids.101 Starting from bleached paper-grade pulps, separate fractions of hemicelluloses and

384

cellulose (both of high purity) were recovered upon processing with [emim]Ac containing 15-

385

20% water.122 Under severe conditions (for example, in high temperature treatments with

386

strongly acidic ILs), the formation of carbohydrate degradation products able to react with

387

lignin (leading to the formation of pseudolignin) has been reported.114 Oppositely, in

388

treatments with ILs possessing strongly basic anions, hemicelluloses are converted into

389

oligosaccharides, generating only trace amounts of monomeric sugars (xylose and glucose).123

390

In some operational schemes, cellulose and lignin are precipitated sequentially, and the

391

hemicellulose-derived saccharides left in the ILs phase are subjected to further chemical

392

modification.124

16

ACS Paragon Plus Environment

Page 16 of 35

Page 17 of 35

Journal of Agricultural and Food Chemistry

393

The integral benefit of the various LCM components can be achieved by coupling

394

conventional separation treatments with IL processing. Following this idea, dissolution in

395

[bmim]Cl followed by NaOH extraction has been reported for bagasse utilization,121 whereas

396

the implementation of a first stage of hemicellulose removal (for example, hydrothermolysis

397

or prehydrolysis) before treatment with ILs (aiming at the separation of cellulose and lignin)

398

has also been proposed.97 This latter approach follows the philosophy underlying the

399

commercial prehydrolysis-kraft process, and is being studied as an alternative to increase the

400

added-value from hemicelluloses in the conventional kraft pulping technology.

401 402

ILs as pretreatment agents for the enzymatic hydrolysis of cellulose

403

The enzymatic hydrolysis of cellulose leads to glucose solutions suitable as fermentation

404

media for manufacturing a wide range of biofuels, solvents, polymers and other marketable

405

compounds. However, native LCM is recalcitrant to the action of cellulases, a fact attributed

406

to a number of factors, such as substrate accessibility, cellulose degree of polymerization,

407

crystallinity, particle size, and porosity, as well as to the presence of hemicellulose and

408

lignin.111 In particular, cellulose is embedded in a matrix of lignin and hemicelluloses, which

409

are crosslinked by ester and ether linkages in the plant cell wall, hindering the access of

410

enzymes to the cellulose glycosidic bonds. To disrupt the cell wall structure, some type of

411

pretreatment (by mechanical, biological, physical and/or chemical methods, individually or in

412

combination) is necessary. This type of processing is referred in literature as “pretreatment”,

413

and is been extensively reviewed owing to its economic and technological importance.

414

Chemical pretreatments modify the composition of substrates, and some of them can be

415

considered as true fractionation treatments. This dual behavior (as pretreatment and/or

416

fractionation agents) defines well the potential of ILs in the field: in fact, a number of

417

references included in this work (particularly, some of the ones employed to discuss the LCM 17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

418

fractionation by ILs) are directly oriented to the production of substrates suitable for

419

enzymatic hydrolysis, since ILs show ability for removing lignin, disrupting the crystalline

420

structure of cellulose, and increasing the accessibility of enzymes to cellulose.32 The ILs most

421

frequently utilized for LCM pretreatment are the ones suitable for cellulose dissolution, for

422

example [emim]Ac, [amim]Cl, [bmim]Cl, and dialkylimidazolium dialkylphosphates, which

423

have been applied to a number of substrates.104,112,125–129 Depending on the IL employed, the

424

effects caused by processing may resemble those of aqueous stages performed under alkaline

425

or acidic conditions, as it happens for acetate- and chloride- based ILs, respectively.130 Data

426

on the comparative saccharification efficiency achievable with selected ILs have been

427

reported recently.129

428

In typical processing schemes, LCM are treated with ILs to cause dissolution-

429

regeneration or delignification, yielding cellulose-enriched substrates of decreased

430

crystallinity and improved enzyme accessibility, which are extensively washed to remove the

431

residual IL and further hydrolyzed with enzymes at high yield with fast kinetics. It can be

432

noted that ILs are comparatively expensive, and that efficient and cost-effective recovery

433

methods have to be implemented for economic feasibility. The best IL depend on the substrate

434

considered;

435

different substrates.109 In terms of saccharification yields, material recovery and

436

delignification, [emim]Ac ranks as the most suitable IL for biomass pretreatment;130 but

437

[bmim]Cl performed very well for simple cellulosic substrates, and [bmim]HSO4 has been

438

recommended for processing hybrid aspen.112 Acetate-based ILs present advantages over

439

chloride-based ILs derived from their lower melting point, lower viscosity, lower corrosive

440

character and ability for working with higher loadings.131 Strongly acidic ionic liquids may

441

cause partial polysaccharide depolymerization, hydrolysis and dehydration, leading to the

442

formation of furans as reaction byproducts.

112

and for a given IL, the optimal processing conditions must be optimized for

18

ACS Paragon Plus Environment

Page 18 of 35

Page 19 of 35

Journal of Agricultural and Food Chemistry

443

It must be considered that, in integrated processes, ILs may be present in the aqueous

444

hydrolysis media as a result of recycling streams coming from IL recovery stages. This is

445

important since the presence of ILs in the hydrolysis media (even in trace amounts) can inhibit

446

or denature enzymes, depending on the type and concentration of the considered IL.

447

According to Turner et al.132, Trichoderma reesei cellulases are subjected to ionic strength-

448

induced inactivation and unfolding in media containing as little as 22 mM [bmim]Cl.

449

Refolding of denatured cellulase was possible only when the enzyme was contacted with

450

solutions containing 0–5% [bmim]Cl. The Cl− ion was, in part, responsible for the inactivation

451

of the cellulase, since the IL produces a dehydrating and denaturing environment.

452

In most cases, the enzyme activity drops quickly when increasing the ionic liquid

453

concentration, but no general conclusions on this subject can be drawn owing to the diversity

454

of conditions assayed. In integrated biorefinery processes, keeping a very low IL

455

concentration in the hydrolysis media would require extensive processing and clean up of the

456

cellulosic substrates. Because of this, the cellulases should be preferably able to perform

457

optimally in solutions containing up to 15 wt% IL, which correspond with the amount retained

458

by the hydrolysis substrates upon processing.133 To deal with this problem, two possible

459

approaches can be followed (individually or simultaneously): utilization of ILs causing

460

minimal inhibition and/or using cellulases with increased tolerance to the IL.

461

The IL [emim]Ac shows minimal inhibitory effects on enzymes compared to ionic

462

liquids containing Cl− or Br− anions;96,133,134 whereas favorable compatibility has been also

463

reported for [emim]DEP,135 [mmim]DEP and [mmim]DMP.136 Concerning the role of

464

enzymes, Wahlström and Suurnäkki125 summarized literature on the action of cellulases and

465

other glycosyl hydrolases in IL solutions. Interesting possibilities lay in the utilization of

466

enzymes from extremophilic strains able to perform at high temperature and salt

467

concentration, as the utilization of enzymes capable of withstanding higher concentrations of 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

468

salt and ionic liquids and higher pH would make the washing steps less cumbersome,

469

facilitating the industrial biomass conversion process.133,137 Datta et al.133 investigated the

470

stability of recombinant, extremophilic enzymes, as a function of IL concentration, and

471

reported that (unlike the reference Trichoderma viride cellulase), the tested enzymes retained

472

between 44% and 79% of their activity in the presence of 15% [emim]Ac. In this field, a GH5

473

family cellulase able to keep activity in the presence 30% (v/v) IL has been identified;138

474

whereas an hyperthermostable archaeal endo-glucanase was reported to retain a considerable

475

activity (30–50%) in the presence of 25% (v/v) [mmim]DMP.139

476

In conclusion, ILs are potentially suitable for a number of duties in LCM biorefineries.

477

Considered as agents for physicochemical processing of liquid process streams, ILs are

478

suitable as entrainers for the dehydration of ethanol-water mixtures, as well as agents for the

479

in situ separation of microbial metabolites (including selected alcohols and acids) from the

480

fermentation media. ILs are also suitable for preventing pollution from gaseous waste streams,

481

for example by removing Volatile Organic Compounds or carbon dioxide from them.

482

Concerning the chemical modification of native LCM, operational schemes based on

483

dissolution-regeneration or delignification provide interesting alternatives for developing

484

alternative processing technologies. One of the most suited schemes for LCM biorefineries

485

(pretreatment/fractionation followed by enzymatic hydrolysis and fermentation) could be

486

conveniently perfomed by using selected ILS (causing limited losses of enzymatic activity

487

and/or showing enhanced compatibility with microorganisms), as well as by using tolerant

488

enzymes produced by extremophilic microorganisms.

489 490

Acknowledgement

20

ACS Paragon Plus Environment

Page 20 of 35

Page 21 of 35

Journal of Agricultural and Food Chemistry

491

The authors are grateful to the Spanish “Ministry of Economy and Competitivity” for

492

supporting this study, in the framework of the research Project “Advanced processing

493

technonologies for biorefineries” (reference CTQ2014-53461-R), partially funded by the

494

FEDER program of the European Union. Ms. Sandra Rivas thanks the Ministry for her

495

predoctoral grant.

496 497

References

498 499

(1) Yu, W.; Tang, Y.; Mo, L.; Chen, P.; Lou, H.; Zheng, X. One-step hydrogenation-

500

esterification of furfural and acetic acid over bifunctional Pd catalysts for bio-oil upgrading.

501

Bioresour. Technol. 2011, 102, 8241–8246.

502 503 504 505 506 507

(2) Blanco-Rosete, S.; Webb, C. Emerging biorefinery markets: global context and prospects for Latin America. Biofuels, Bioprod. Biorefining 2008, 2, 331–342. (3) Sheldon, R. Green and sustainable manufacture of chemicals from biomass: state of the art. Green Chem. 2014, 16, 950–963. (4) Clark, J. H.; Deswarte, F. E. I.; Farmer, T. J. The integration of green chemistry into future biorefineries. Biofuels, Bioprod. Biorefining 2009, 3, 72–90.

508

(5) Mittal, A.; Chatterjee, S. G.; Scott, G. M.; Amidon, T. E. Modeling xylan

509

solubilization during autohydrolysis of sugar maple and aspen wood chips: Reaction kinetics

510

and mass transfer. Chem. Eng. Sci. 2009, 64, 3031–3041.

511

(6) Marcuschamer, D. K.; Simmons, B. A.; Blanch, H. W. Techno-economic analysis of a

512

lignocellulosic ethanol biorefinery with ionic liquid pre-treatment. Biofuels, Bioprod.

513

Biorefining 2011, 5, 562–569.

514

(7) Sen, S. M.; Binder, J. B.; Raines, R. T.; Maravelias, C. T. Conversion of biomass to

515

sugars via ionic liquid hydrolysis: process synthesis and economic evaluation. Biofuels

516

Bioprod. Biorefining 2012, 6, 444–452.

517 518 519 520

(8) Shao, X.; Lynd, L. Kinetic modeling of xylan hydrolysis in co- and countercurrent liquid hot water flow-through pretreatments. Bioresour. Technol. 2013, 130, 117–124. (9) Corma, A.; Iborra, S.; Velty, A. Chemical routes for the transformation of biomass into chemicals. Chem. Rev. 2007, 107, 2411–2502. 21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

521 522 523

(10)

Page 22 of 35

Binder, J. B.; Raines, R. T. Fermentable sugars by chemical hydrolysis of

biomass. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 4516–4521. (11)

Vargas-Hernández, D.; Rubio-Caballero, J. M.; Santamaría-González, J.;

524

Moreno-Tost, R.; Mérida-Robles, J. M.; Pérez-Cruz, M.; Jiménez-López, A.; Hernández-

525

Huesca, R.; Maireles-Torres, P. Furfuryl alcohol from furfural hydrogenation over copper

526

supported on SBA-15 silica catalysts. J. Mol. Catal. A Chem. 2014, 383-384, 106–113.

527 528 529

(12)

Rackemann, D. W.; Doherty, W. O. The conversion of lignocellulosics to

levulinic acid. Biofuels Bioprod. Bioref. 2011, 5, 198–214. (13)

Budarin, V. L.; Shuttleworth, P. S.; Dodson, J. R.; Hunt, A. J.; Lanigan, B.;

530

Marriott, R.; Milkowski, K. J.; Wilson, A. J.; Breeden, S. W.; Fan, J.; Sin, E. H. K.; Clark, J.

531

H. Use of green chemical technologies in an integrated biorefinery. Energ. Environ. Sci. 2011,

532

4, 471–479.

533 534 535 536 537 538 539

(14)

Clark, J. H.; Luque, R.; Matharu, A. S. Green Chemistry, biofuels, and

biorefinery. Annu. Rev. Chem. Biomol. Eng. 2012, 3, 183–207. (15)

Beach, E. S.; Cui, Z.; Anastas, P. T. Green Chemistry: A design framework for

sustainability. Energ. Environ. Sci. 2009, 2, 1038–1049. (16)

Luque, R. Catalytic biomass processing : prospects in future biorefineries.

Curr. Green Chem. 2015, 2, 90–95. (17)

Bos, H. L.; Sanders, J. P. M. Raw material demand and sourcing options for the

540

development of a bio-based chemical industry in Europe. Biofuels Bioprod. Bioref. 2013, 7,

541

246–259.

542

(18)

Rosatella, A.; Simeonov, S. P.; Frade, R. F. M.; Afonso, C. A. M. 5-

543

Hydroxymethylfurfural (HMF) as a building block platform: Biological properties, synthesis

544

and synthetic applications. Green Chem. 2011, 13, 754–793.

545 546 547

(19)

World Commission on Environment and Development. Report: Our Common

Future (The Brundtland Report). Med. Confl. Surviv. 1988, 4, 17-25. (20)

German

Federal

Government.

Biorefineries

Roadmap.

Retrieved

at:

548

http://www.bmbf.de/pub/BMBF_Roadmap-Bioraffinerien_en_bf.pdf. Last accessed, July

549

2015.

550

(21)

Höltinger, S.; Schmidt, J.; Schönhart, M.; Schmid, E. A spatially explicit

551

techno-economic assessment of green biorefinery concepts. Biofuels Bioprod. Bioref. 2014, 8,

552

325–341. 22

ACS Paragon Plus Environment

Page 23 of 35

553

Journal of Agricultural and Food Chemistry

(22)

De Jong, W.; Marcotullio, G. Overview of biorefineries based on bo-

554

production of furfural, existing concepts and novel developments. Int. J. Chem. React. Eng.

555

2010, 8, 1-27.

556 557 558 559 560

(23)

De Jong, E.; Higson, A.; Walsh, P.; Wellisch, M. Product developments in the

bio-based chemicals arena. Biofuels Bioprod. Bioref.2012, 6, 606–624. (24)

Cherubini, F. The biorefinery concept: Using biomass instead of oil for

producing energy and chemicals. Energy Convers. Manag. 2010, 51, 1412–1421. (25)

Vandermeulen, V.; Van der Steen, M.; Stevens, C. V.; Van Huylenbroeck, G.

561

Industry expectations regarding the transition toward a biobased economy. Biofuels Bioprod.

562

Bioref. 2012, 6, 453–464.

563 564 565 566 567 568 569 570 571

(26)

Stark, A. Ionic liquids in the biorefinery: a critical assessment of their potential.

Energy Environ. Sci. 2011, 4, 19-32 (27)

Olivier-Bourbigou, H.; Magna, L.; Morvan, D. Ionic liquids and catalysis:

Recent progress from knowledge to applications. Appl. Catal. A Gen. 2010, 373, 1–56. (28)

Sheldon, R. A. Green solvents for sustainable organic synthesis: state of the art.

Green Chem. 2005, 7 (5), 267-278. (29)

Sheldon, R. A. Fundamentals of green chemistry: efficiency in reaction design.

Chem. Soc. Rev. 2012, 41 (4), 1437-1451. (30)

Simoni, L. D.; Chapeaux, A.; Brennecke, J. F.; Stadtherr, M. A. Extraction of

572

biofuels and biofeedstocks from aqueous solutions using ionic liquids. Comput. Chem. Eng.

573

2010, 34, 1406–1412.

574

(31)

Long, J.; Guo, B.; Teng, J.; Yu, Y.; Wang, L.; Li, X. SO3H-functionalized ionic

575

liquid: Efficient catalyst for bagasse liquefaction. Bioresour. Technol. 2011, 102, 10114–

576

10123.

577

(32)

578 579 580 581

Liu, C.-Z.; Wang, F.; Stiles, A. R.; Guo, C. Ionic liquids for biofuel production:

Opportunities and challenges. Appl. Energy 2012, 92, 406–414. (33)

Earle, M. J.; Seddon, K. R. Ionic liquids. Green solvents for the future. Pure

Appl. Chem. 2000, 72 (7), 1391–1398. (34)

Huddleston, J. G.; Visser, A. E.; Reichert, W. M.; Willauer, H. D.; Broker, G.

582

A.; Rogers, R. D. Characterization and comparison of hydrophilic and hydrophobic room

583

temperature ionic liquids incorporating the imidazolium cation. Green Chem. 2001, 3, 156–

584

164. 23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

585

(35)

Niknam, K.; Damya, M. 1-butyl-3-methylimidazolium hydrogen sulfate

586

[bmim] HSO4 : An efficient reusable acidic ionic liquid for the synthesis of 1,8-dioxo-

587

octahydroxanthenes. J. Chin. Chem: Soc. 2009, 56, 659–665.

588

(36)

Fraga-Dubreuil, J.; Bourahla, K.; Rahmouni, M.; Bazureau, J. P.; Hamelin, J.

589

Catalysed esterifications in room temperature ionic liquids with acidic counteranion as

590

recyclable reaction media. Catal. Commun. 2002, 3 (5), 185–190.

591

(37)

Mäki-Arvela, P.; Anugwom, I.; Virtanen, P.; Sjöholm, R.; Mikkola, J. P.

592

Dissolution of lignocellulosic materials and its constituents using ionic liquids-A review. Ind.

593

Crops Prod. 2010, 32, 175–201.

594 595 596

(38)

Vancov, T.; Alston, A. S.; Brown, T.; McIntosh, S. Use of ionic liquids in

converting lignocellulosic material to biofuels. Renew. Energy 2012, 45, 1–6. (39)

Seiler, M.; Jork, C.; Kavarnou, A.; Arlt, W.; Hirsch, R. Separation of

597

azeotropic mixtures using hyperbranched polymers or ionic liquids. AIChE J. 2004, 50, 2439–

598

2454.

599

(40)

Huang, H. J.; Ramaswamy, S.; Tschirner, U. W.; Ramarao, B. V. A review of

600

separation technologies in current and future biorefineries. Sep. Purif. Technol. 2008, 62, 1–

601

21.

602

(41)

Figueroa, J. J.; Hoss Lunelli, B.; Maciel Filho, R.; Wolf Maciel, M. R.

603

Improvements on anhydrous ethanol production by extractive distillation using ionic liquid as

604

solvent. Procedia Eng. 2012, 42, 1016–1026.

605

(42)

Meindersma, G. W. ; Quijada-Maldonado, E.; Aelmans, T. A. M.; Gutiérrez

606

Hernéndez, J. P.; de Haan, A. B. Ionic liquids in extractive distillation of ethanol/water: from

607

laboratory to pilot plant. ACS Symp. Ser. 2012, 117, 239–257.

608 609 610

(43)

Arlt, W.; Seiler, M.; Jork, C.; T. S. Ionic liquids as selective additives for

separation of close-boiling or azeotropic mixtures. US Pat. 20040133058 A1, 2004. (44)

Swatloski, R. P.; Visser, A. E.; Reichert, W. M.; Broker, G. A.; Farina, L. M.;

611

Holbrey, J. D.; Rogers, R. D. On the solubilization of water with ethanol in hydrophobic

612

hexafluorophosphate ionic liquids. Green Chem. 2001, 4 (2), 81–87.

613

(45)

Chapeaux, A.; Simoni, L. D.; Ronan, T. S.; Stadtherr, M. A.; Brennecke, J. F.

614

Extraction

615

bis(trifluoromethylsulfonyl)imide. Green Chem. 2008, 10, 1301-1306.

of

alcohols

from

water

with

24

ACS Paragon Plus Environment

1-hexyl-3-methylimidazolium

Page 24 of 35

Page 25 of 35

616

Journal of Agricultural and Food Chemistry

(46)

Neves, C. M. S. S.; Granjo, J. F. O.; Freire, M. G.; Robertson, A.; Oliveira, N.

617

M. C.; Coutinho, J. A. P. Separation of ethanol–water mixtures by liquid–liquid extraction

618

using phosphonium-based ionic liquids. Green Chem. 2011, 13, 1517-1526.

619 620 621 622 623 624 625 626 627

(47)

Dürre, P. Biobutanol: An attractive biofuel. Biotechnol. J. 2007, 2 (12), 1525–

1534. (48)

Bankar, S. B.; Survase, S. A.; Ojamo, H.; Granström, T. Biobutanol: the

outlook of an academic and industrialist. RSC Adv. 2013, 3, 24734-24757. (49)

Knoshaug, E. P.; Zhang, M. Butanol tolerance in a selection of

microorganisms. Appl. Biochem. Biotechnol. 2009, 153), 13–20. (50)

Green, E. M. Fermentative production of butanol-the industrial perspective.

Curr. Opin. Biotechnol. 2011, 22, 337–343. (51)

Heitmann, S.; Krings, J.; Kreis, P.; Lennert, A.; Pitner, W. R.; Górak, A.;

628

Schulte, M. M. Recovery of n-butanol using ionic liquid-based pervaporation membranes.

629

Sep. Purif. Technol. 2012, 97, 108–114.

630 631 632 633 634 635 636 637 638 639 640

(52)

Kaminski, W.; Tomczak, E.; Gorak, A. bioutanol - production and purifciaton

methods. Ecol. Chem. Eng. S-Chemia I Inz. Ekol. S 2011, 18, 31–37. (53)

Zhao, H.; Xia, S.; Ma, P. Use of ionic liquids as “green” solvents for

extractions. J. Chem. Technol. Biotechnol. 2005, 80, 1089–1096. (54)

Kubiczek, A.; Jonowe, C. Ionic liquids for the extraction of n-butanol from

aqueous solutions. Proc. ECOPole 2013, 7, 125-131. (55)

Ha, S. H.; Mai, N. L.; Koo, Y. M. Butanol recovery from aqueous solution into

ionic liquids by liquid-liquid extraction. Process Biochem. 2010, 45, 1899–1903. (56)

Fadeev, A. G.; Meagher, M. M. Opportunities for ionic liquids in recovery of

biofuels. Chem. Commun. 2001, 3, 295–296. (57)

Plaza, A.; Merlet, G.; Hasanoglu, A.; Isaacs, M.; Sánchez, J.; Romero, J.

641

Separation of butanol from ABE mixtures by sweep gas pervaporation using a supported

642

gelled ionic liquid membrane: Analysis of transport phenomena and selectivity. J. Memb. Sci.

643

2013, 444, 201–212.

644

(58)

Cascon, H. R.; Choudhari, S. K.; Nisola, G. M.; Vivas, E. L.; Lee, D. J.;

645

Chung, W. J. Partitioning of butanol and other fermentation broth components in

646

phosphonium and ammonium-based ionic liquids and their toxicity to solventogenic

647

clostridia. Sep. Purif. Technol. 2011, 78, 164–174. 25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

648

(59)

Matsumoto, M.; Mochiduki, K.; Fukunishi, K.; Kondo, K. Extraction of

649

organic acids using imidazolium-based ionic liquids and their toxicity to Lactobacillus

650

rhamnosus. Sep. Purif. Technol. 2004, 40, 97–101.

651 652 653

(60)

Matsumoto, M.; Mochiduki, K.; Kondo, K. Toxicity of ionic liquids and

organic solvents to lactic acid-producing bacteria. J. Biosci. Bioeng. 2004, 98, 344–347. (61)

Li, Q.; Jiang, X.; Zou, H.; Cao, Z.; Zhang, H.; Xian, M. Extraction of short-

654

chain organic acids using imidazolium-based ionic liquids from aqueous media. J. Chem.

655

Pharmac. Res. 2014, 6, 374–381.

656

(62)

Tonova, K.; Svinyarov, I.; Bogdanov, M. G. Biocompatible ionic liquids in

657

liquid – liquid extraction of lactic acid : A comparative study. Int. J. Chem. Mol. Nuclear Mat.

658

Metallurgical Eng. 2015, 9, 554–558.

659

(63)

Gullón, P.; Romaní, A.; Vila, C.; Garrote, G.; Parajó, J. C. Potential of

660

hydrothermal treatments in lignocellulose biorefineries. Biofuels Bioprod. Bioref. 2012, 6,

661

219–232.

662

(64)

663 664 665 666

Zhu, S. Use of ionic liquids for the efficient utilization of lignocellulosic

materials. J. Chem. Technol. Biotechnol. 2008, 83, 777–779. (65)

Milota, M. R.; Li, K. VOC and HAP recovery using ionic liquids. Retrieved

from: http://www.osti.gov/scitech/servlets/purl/909870. Last accessed, July 2015. (66)

Quijano, G.; Couvert, A.; Amrane, A.; Darracq, G.; Couriol, C.; Le Cloirec, P.;

667

Paquin, L.; Carrié, D. Absorption and biodegradation of hydrophobic Volatile Organic

668

Compounds in ionic liquids. Water Air Soil Pollut. 2013, 224, 1-9.

669 670 671

(67)

Quijano, G.; Couvert, A.; Amrane, A. Ionic liquids: Applications and future

trends in bioreactor technology. Bioresour. Technol. 2010, 101, 8923–8930. (68)

Quijano, G.; Couvert, A.; Amrane, A.; Darracq, G.; Couriol, C.; Le Cloirec, P.;

672

Paquin, L.; Carrié, D. Potential of ionic liquids for VOC absorption and biodegradation in

673

multiphase systems. Chem. Eng. Sci. 2011, 66, 2707–2712.

674

(69)

Cadena, C.; Anthony, J. L.; Shah, J. K.; Morrow, T. I.; Brennecke, J. F.;

675

Maginn, E. J. Why is CO2 so soluble in imidazolium-based ionic liquids?. J. Am. Chem. Soc.

676

2004, 126, 5300–5308.

677 678 679 680

(70)

Galán Sánchez, L. M.; Meindersma, G. W.; de Haan, A. B. Solvent properties

of functionalized ionic liquids for CO2 absorption. Chem. Eng. Res. Des. 2007, 85, 31–39. (71)

Hasib-ur-Rahman, M.; Siaj, M.; Larachi, F. Ionic liquids for CO2 capture-

Development and progress. Chem. Eng. Process. Process Intensif. 2010, 49, 313–322. 26

ACS Paragon Plus Environment

Page 26 of 35

Page 27 of 35

681 682 683

Journal of Agricultural and Food Chemistry

(72)

Bates, E. D.; Mayton, R. D.; Ntai, I.; Davis Jr, J. H. CO2 capture by a task

specific ionic liquid. J. Am. Chem. Soc. 2002, 124, 926–927. (73)

Cao, Y.; Wu, J.; Zhang, J.; Li, H.; Zhang, Y.; He, J. Room temperature ionic

684

liquids (RTILs): A new and versatile platform for cellulose processing and derivatization.

685

Chem. Eng. J. 2009, 147, 13–21.

686 687

(74)

Fukaya, Y.; Hayashi, K.; Wada, M.; Ohno, H. Cellulose dissolution with polar

ionic liquids under mild conditions: required factors for anions. Green Chem. 2008, 10, 44-46.

688

(75)

Graenacher, C. Cellulose solution. Pat. US 1943176 A. 1934

689

(76)

Swatloski, R. P.; Rogers, R. D.; Holbrey, J. D. Dissolution and processing of

690 691 692 693 694 695 696 697 698 699

cellulose using ionic liquids Pat WO 2003029329 A2 2003 .(77)

Feng, L.; Chen, Z. L. Research progress on dissolution and functional

modification of cellulose in ionic liquids. J. Mol. Liq. 2008, 142, 1–5. (78)

Wang, H.; Gurau, G.; Rogers, R. D. Ionic liquid processing of cellulose. Chem.

Soc. Rev. 2012, 41, 1519-1537. (79)

Tan, H. T.; Lee, K. T. Understanding the impact of ionic liquid pretreatment on

biomass and enzymatic hydrolysis. Chem. Eng. J. 2012, 183, 448–458. (80)

Gupta, K. M.; Jiang, J. Cellulose dissolution and regeneration in ionic liquids:

A computational perspective. Chem. Eng. Sci. 2015, 121, 180–189. (81)

Zhang, H.; Wu, J.; Zhang, J.; He, J. 1-allyl-3-methylimidazolium chloride room

700

temperature ionic liquid: A new and powerful nonderivatizing solvent for cellulose.

701

Macromolecules 2005, 38, 8272–8277.

702 703 704

(82)

Ignatyev, I.; Van Doorslaer, C.; Mertens, P. G. N.; Binnemans, K.; de Vos, D.

E. Synthesis of glucose esters from cellulose in ionic liquids. Holzfors.2011, 66, 417–425. (83)

Fukaya, Y.; Sugimoto, A.; Ohno, H. Superior solubility of polysaccharides in

705

low viscosity, polar and halogen-free 1,3-dialkylimidazolium formates. Biomacromolecules

706

2006, 7, 3295–3297.

707 708 709

(84)

Swatloski, R. P.; Spear, S. K.; Holbrey, J. D.; Rogers, R. D. Dissolution of

cellulose with ionic liquids. J. Am. Chem. Soc. 2002, 124, 4974–4975. (85)

Zhao, H.; Jones, C. L.; Baker, G. A.; Xia, S.; Olubajo, O.; Person, V. N.

710

Regenerating cellulose from ionic liquids for an accelerated enzymatic hydrolysis. J.

711

Biotechnol. 2009, 139, 47–54.

712 713

(86)

Wu, J.; Zhang, J.; Zhang, H.; He, J.; Ren, Q.; Guo, M. Homogeneous

acetylation of cellulose in a new ionic liquid. Biomacromolecules 2004, 5, 266–268. 27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

714 715 716 717 718

(87)

Barthel, S.; Heinze, T. Acylation and carbanilation of cellulose in ionic liquids.

Green Chem. 2006, 8, 301-306. (88)

Ogawa, S.; Miyafuji, H. Reaction behavior of milled wood lignin in an ionic

liquid, 1-ethyl-3-methylimidazolium chloride. J. Wood Sci. 2015, 61, 285–291. (89)

Gericke, M.; Liebert, T.; Heinze, T. Interaction of ionic liquids with

719

polysaccharides, 8-synthesis of cellulose sulfates suitable for polyelectrolyte complex

720

formation. Macromol. Biosci. 2009, 9, 343–353.

721 722 723

(90)

Mormann, W.; Wezstein, M. Trimethylsilylation of cellulose in ionic liquids.

Macromol. Biosci. 2009, 9, 369–375. (91)

Köhler, S.; Liebert, T.; Schöbitz, M.; Schaller, J.; Meister, F.; Günther, W.;

724

Heinze, T. Interactions of ionic liquids with polysaccharides 1. Unexpected acetylation of

725

cellulose with 1-ethyl-3-methylimidazolium acetate. Macromol. Rapid Commun. 2007, 28,

726

2311–2317.

727 728 729

(92)

Köhler, S.; Liebert, T.; Heinze, T. Ammonium-based cellulose solvents suitable

for homogeneous etherification. Macromol. Biosci. 2009, 9, 836–841. (93)

Crépy, L.; Chaveriat, L.; Banoub, J.; Martin, P.; Joly, N. Synthesis of cellulose

730

fatty esters as plastics-influence of the degree of substitution and the fatty chain length on

731

mechanical properties. ChemSusChem 2009, 2, 165–170.

732

(94)

Li, W. Y.; Jin, A. X.; Liu, C. F.; Sun, R. C.; Zhang, A. P.; Kennedy, J. F.

733

Homogeneous modification of cellulose with succinic anhydride in ionic liquid using 4-

734

dimethylaminopyridine as a catalyst. Carbohydr. Polym. 2009, 78, 389–395.

735 736 737

(95)

Ohno, H.; Fukaya, Y. Task specific ionic liquids for cellulose technology.

Chem. Lett. 2009, 38, 2–7. (96)

Wang, X.; Li, H.; Cao, Y.; Tang, Q. Cellulose extraction from wood chip in an

738

ionic liquid 1-allyl-3-methylimidazolium chloride (AmimCl). Bioresour. Technol. 2011, 102,

739

7959–7965.

740 741 742

(97)

Brandt, A.; Gräsvik, J.; Hallett, J. P.; Welton, T. Deconstruction of

lignocellulosic biomass with ionic liquids. Green Chem. 2013, 15, 550–583. (98)

Zhu, S.; Yu, P.; Wang, Q.; Cheng, B.; Chen, J.; Wu, Y. Breaking the barriers of

743

lignocellulosic ethanol production using ionic liquid technology. Bioresources 2013, 8, 1510–

744

1512.

28

ACS Paragon Plus Environment

Page 28 of 35

Page 29 of 35

745

Journal of Agricultural and Food Chemistry

(99)

Fort, D. A.; Remsing, R. C.; Swatloski, R. P.; Moyna, P.; Moyna, G.; Rogers,

746

R. D. Can ionic liquids dissolve wood? Processing and analysis of lignocellulosic materials

747

with 1-n-butyl-3-methylimidazolium chloride. Green Chem. 2007, 9, 63-69.

748

(100)

Espinoza-Acosta, J. L:; Torres-Chávez, P. I:; Carvajal-Millán, E.; Ramírez-

749

Wong, B.; Bello-Pérez, L. A.; Montaño-Leyva, B. Ionic liquids and organic solvents for

750

recovering lignin from lignocellulosic miomass. 2014, 9, 3660–3687.

751

(101)

Anugwom, I.; Mäki-Arvela, P.; Virtanen, P.; Willför, S.; Sjöholm, R.; Mikkola,

752

J. P. Selective extraction of hemicelluloses from spruce using switchable ionic liquids.

753

Carbohydr. Polym. 2012, 87, 2005–2011.

754 755 756

(102)

Kilpelainen, I.; Xie, H.; King, A; Granstrom, M.; Heikkinen, S.; Argyropoulos,

D. S. Dissolution of Wood in Ionic Liquids. J. Agric. Food Chem. 2007, 55, 9142–9148. (103)

Li, C.; Knierim, B.; Manisseri, C.; Arora, R.; Scheller, H. V.; Auer, M.; Vogel,

757

K. P.; Simmons, B. A.; Singh, S. Comparison of dilute acid and ionic liquid pretreatment of

758

switchgrass: Biomass recalcitrance, delignification and enzymatic saccharification. Bioresour.

759

Technol. 2010, 101, 4900–4906.

760

(104)

Singh, S.; Simmons, B. A.; Vogel, K. P. Visualization of biomass

761

solubilization and cellulose regeneration during ionic liquid pretreatment of switchgrass.

762

Biotechnol. Bioeng. 2009, 104, 68–75.

763 764 765

(105)

Samayam, I. P.; Schall, C. A. Saccharification of ionic liquid pretreated

biomass with commercial enzyme mixtures. Bioresour. Technol. 2010, 101, 3561–3566. (106)

Sun, N.; Rahman, M.; Qin, Y.; Maxim, M. L.; Rodríguez, H.; Rogers, R. D.

766

Complete dissolution and partial delignification of wood in the ionic liquid 1-ethyl-3-

767

methylimidazolium acetate. Green Chem. 2009, 11, 646.

768

(107)

Lee, S. H.; Doherty, T. V.; Linhardt, R. J.; Dordick, J. S. Ionic liquid-mediated

769

selective extraction of lignin from wood leading to enhanced enzymatic cellulose hydrolysis.

770

Biotechnol. Bioeng. 2009, 102, 1368–1376.

771

(108)

Da Costa Lopes, A. M.; João, K. G.; Morais, A. R. C.; Bogel-Łukasik, E.;

772

Bogel-Łukasik, R. Ionic liquids as a tool for lignocellulosic biomass fractionation. Sustain.

773

Chem. Process. 2013, 1, 1-31.

774

(109)

Shill, K.; Padmanabhan, S.; Xin, Q.; Prausnitz, J. M.; Clark, D. S.; Blanch, H.

775

W. Ionic liquid pretreatment of cellulosic biomass: Enzymatic hydrolysis and ionic liquid

776

recycle. Biotechnol. Bioeng. 2011, 108, 511–520. 29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

777

(110)

Xu, F.; Shi, Y. C.; Wang, D. Enhanced production of glucose and xylose with

778

partial dissolution of corn stover in ionic liquid, 1-ethyl-3-methylimidazolium acetate.

779

Bioresour. Technol. 2012, 114, 720–724.

780

(111)

Sathitsuksanoh, N.; George, A.; Zhang, Y. H. P. New lignocellulose

781

pretreatments using cellulose solvents: A review. J. Chem. Technol. Biotechnol. 2013, 88,

782

169–180.

783

(112)

Gräsvik, J.; Winestrand, S.; Normark, M.; Jönsson, L. J.; Mikkola, J. P.

784

Evaluation of four ionic liquids for pretreatment of lignocellulosic biomass. BMC Biotechnol.

785

2014, 14, 1-11.

786

(113)

Cox, B. J.; Ekerdt, J. G. Pretreatment of yellow pine in an acidic ionic liquid:

787

Extraction of hemicellulose and lignin to facilitate enzymatic digestion. Bioresour. Technol.

788

2013, 134, 59–65.

789

(114)

Brandt, A.; Hallett, J. P.; Leak, D. J.; Murphy, R. J.; Welton, T. Ionic liquid

790

pretreatment of lignocellulosic biomass with ionic liquid-water mixtures Green Chem. 2011,

791

13, 2489–2499.

792

(115)

Tan, S. S. Y.; MacFarlane, D. R.; Upfal, J.; Edye, L. A; Doherty, W. O. S.;

793

Patti, A. F.; Pringle, J. M.; Scott, J. L. Extraction of lignin from lignocellulose at atmospheric

794

pressure using alkylbenzenesulfonate ionic liquid. Green Chem. 2009, 11, 339-345.

795

(116)

Fu, D.; Mazza, G.; Tamaki, Y. Lignin extraction from straw by ionic liquids

796

and enzymatic hydrolysis of the cellulosic residues. J. Agric. Food Chem. 2010, 58, 2915–

797

2922.

798 799 800

(117)

Fu, D.; Mazza, G. Optimization of processing conditions for the pretreatment

of wheat straw using aqueous ionic liquid. Bioresour. Technol. 2011, 102, 8003–8010. (118)

Pinkert, A.; Goeke, D. F.; Marsh, K. N.; Pang, S. Extracting wood lignin

801

without dissolving or degrading cellulose: investigations on the use of food additive-derived

802

ionic liquids. Green Chem. 2011, 13, 3124-3136.

803

(119)

Eerhart, A. J. J. E.; Huijgen, W. J. J.; Grisel, R. J. H.; van der Waal, J. C.; de

804

Jong, E.; de Sousa Dias, A.; Faaij, A. P. C.; Patel, M. K. Fuels and plastics from

805

lignocellulosic biomass via the furan pathway; a technical analysis. RSC Adv. 2014, 4, 3536-

806

3549.

807 808

(120)

Bensah, E. C.; Mensah, M. Chemical pretreatment methods for the production

of cellulosic ethanol: Technologies and innovations. Int. J. Chem. Eng. 2013, 1, 1-21. 30

ACS Paragon Plus Environment

Page 30 of 35

Page 31 of 35

809

Journal of Agricultural and Food Chemistry

(121)

Lan, W.; Liu, C. F.; Sun, R. C. Fractionation of bagasse into cellulose,

810

hemicelluloses, and lignin with ionic liquid treatment followed by alkaline extraction. J.

811

Agric. Food Chem. 2011, 59, 8691–8701.

812

(122)

Froschauer, C.; Hummel, M.; Iakovlev, M.; Roselli, A.; Schottenberger, H.;

813

Sixta, H. Separation of hemicellulose and cellulose from wood pulp by means of ionic

814

liquid/cosolvent systems. Biomacromolecules 2013, 14, 1741–1750.

815

(123)

Arora, R.; Manisseri, C.; Li, C.; Ong, M. D.; Scheller, H. V.; Vogel, K.;

816

Simmons, B. A.; Singh, S. Monitoring and analyzing process streams towards understanding

817

ionic liquid pretreatment of switchgrass (Panicum virgatum L.). Bioenergy Res. 2010, 3, 134–

818

145.

819

(124)

Van Spronsen, J.; Cardoso, M. A. T.; Witkamp, G. J.; de Jong, W.; Kroon, M.

820

C. Separation and recovery of the constituents from lignocellulosic biomass by using ionic

821

liquids and acetic acid as co-solvents for mild hydrolysis. Chem. Eng. Process. Process

822

Intensif. 2011, 50, 196–199.

823 824 825

(125)

Wahlström, R. M.; Suurnäkki, a. Enzymatic hydrolysis of lignocellulosic

polysaccharides in the presence of ionic liquids. Green Chem. 2015, 17, 694–714. (126)

Bose, S.; Armstrong, D. W.; Petrich, J. W. Enzyme-catalyzed hydrolysis of

826

cellulose in ionic liquids: A green approach toward the production of biofuels. J. Phys. Chem.

827

B 2010, 114, 8221–8227.

828

(127)

Yáñez, R.; Gómez, B.; Martínez, M.; Gullón, B.; Alonso, J. L. Valorization of

829

an invasive woody species, Acacia dealbata, by means of ionic liquid pretreatment and

830

enzymatic hydrolysis. J. Chem. Technol. Biotechnol. 2014, 89, 1337–1343.

831

(128)

Kamiya, N.; Matsushita, Y.; Hanaki, M.; Nakashima, K.; Narita, M.; Goto, M.;

832

Takahashi, H. Enzymatic in situ saccharification of cellulose in aqueous-ionic liquid media.

833

Biotechnol. Lett. 2008, 30, 1037–1040.

834

(129)

Khare, S. K.; Pandey, A.; Larroche, C. Current perspectives in enzymatic

835

saccharification of lignocellulosic biomass. Biochem. Eng. J. 2015 (In press) DOI:

836

10.1016/j.bej.2015.02.033.

837

(130)

Karatzos, S. K.; Edye, L. A; Doherty, W. O. Sugarcane bagasse pretreatment

838

using three imidazolium-based ionic liquids; mass balances and enzyme kinetics. Biotechnol.

839

Biofuels 2012, 5, 62.

31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

840

(131)

Francisco, M.; Mlinar, A. N.; Yoo, B.; Bell, A. T.; Prausnitz, J. M. Recovery of

841

glucose from an aqueous ionic liquid by adsorption onto a zeolite-based solid. Chem. Eng. J.

842

2011, 172, 184–190.

843

(132)

Turner, M. B.; Spear, S. K.; Huddleston, J. G.; Holbrey, J. D.; Rogers, R. D.

844

Ionic liquid salt-induced inactivation and unfolding of cellulase from Trichoderma reesei.

845

Green Chem. 2003, 5, 443-447.

846

(133)

Datta, S.; Holmes, B.; Park, J. I.; Chen, Z.; Dibble, D. C.; Hadi, M.; Blanch, H.

847

W.; Simmons, B. A.; Sapra, R. Ionic liquid tolerant hyperthermophilic cellulases for biomass

848

pretreatment and hydrolysis. Green Chem. 2010, 12, 338-345.

849 850 851

(134)

Li, L.; Yu, S. T.; Liu, F. S.; Xie, C. X.; Xu, C. Z. Cellulose in aqueous-ionic

liquid media by microwave pretreatment. BioResources 2011, 6, 4494–4504. (135)

Li, Q.; He, Y. C.; Xian, M.; Jun, G.; Xu, X.; Yang, J. M.; Li, L. Z. Improving

852

enzymatic hydrolysis of wheat straw using ionic liquid 1-ethyl-3-methyl imidazolium diethyl

853

phosphate pretreatment. Bioresour. Technol. 2009, 100, 3570–3575.

854

(136)

Engel, P.; Mladenov, R.; Wulfhorst, H.; Jäger, G.; Spiess, A. C. Point by point

855

analysis: how ionic liquid affects the enzymatic hydrolysis of native and modified cellulose.

856

Green Chem. 2010, 12, 1959-1966.

857

(137)

Guerriero, G.; Hausman, J.-F.; Strauss, J.; Ertan, H.; Siddiqui, K. S.

858

Destructuring plant biomass: Focus on fungal and extremophilic cell wall hydrolases. Plant

859

Sci. 2015, 234, 180–193

860

(138) Pöttkamper, J.; Barthen, P.; Limberger, N.; Schwaneberg, U.; Schenk, A.; Schulte,

861

M.; Ignatiev N.; Streit, W. Applying metagenomics for the identification of bacterial

862

cellulases that are stable in ionic liquids. Green Chem., 2009, 11, 957–965.

863

(139)

Graham, J. E.; Clark, M. E.; Nadler, D. C.; Huffer, S.; Chokhawala, H. A.;

864

Rowland, S. E.; Blanch, H. W.; Clark, D. S.; Robb, F. T. Identification and characterization of

865

a multidomain hyperthermophilic cellulase from an archaeal enrichment. Nat. Commun. 2011,

866

2, 1-10.

867

32

ACS Paragon Plus Environment

Page 32 of 35

Page 33 of 35

868

Journal of Agricultural and Food Chemistry

FIGURE LEGEND

869 870 871

Figure 1. General nomenclature employed for imidazolium-based ILs and representative examples of ILs employed for biomass processing.

872

33

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

873

Page 34 of 35

TOC GRAPHIC

874 875 876 877 878

Ionic liquids R1

+

N

N

CH3



A

Products

879 880 881

LIGNOCELLULOSE BIOREFINERY .Physical separation .Chemical modification

34

ACS Paragon Plus Environment

Page 35 of 35

Journal of Agricultural and Food Chemistry

GENERAL FORMULA

R1

N

N+

CH3

GENERAL NOMENCLATURE

Anion Substituent Methyl imidazolium cation (mim)

Representative ILs cited in this study [Hmim]Cl [mmim]DEP [mmim]DMP [emim]ABS [emim]Ac [emim]Ace [emim]BF4 [emim]Cl [emim]DEP

[emim]N(CN)2 [bmim]Ac [bmim]Ace [bmim]BF4 [bmim]Br [bmim]Cl [bmim]HSO4 [bmim]mSO4 [bmim]N(CN)2

[R1mim]A

A─

[bmim]N(CN)2 [bmim]PF6 [bmim]Tf2N [hmim]Cl [hmim]PF6 [hmim]Tf2N [omim]PF6 [omim]Tf2N [dmim]TCB [amim]Cl

Typical R1 and nomenclature Hydrogen, H Methyl, m Ethyl, e Butyl, b Hexyl, h Octyl, o Decyl, d Allyl, a

Typical A− and nomenclature Chloride, ClBromide, BrTetrafluoroborate, BF4− Hexafluorophosphate, PF6 − Dicyanamide, N(CN)2 − Hydrogen Sulfate, HSO4 − Acetate, Ac − Tetracyanoborate, TCB − Methylsulfate, mSO4 − Dimethylphosphate, DMP − Diethylphosphate, DEP − Acesulfamate, Ace − Bis(trifluoromethylsulfonyl)imide,Tf2N − Alkylbenzenesulfonate, ABS −

Figure 1

ACS Paragon Plus Environment