Production of 5-hydroxymethylfurfural and levulinic acid from lignocellulosic biomass and catalytic upgradation

Production of 5-hydroxymethylfurfural and levulinic acid from lignocellulosic biomass and catalytic upgradation

Industrial Crops & Products 130 (2019) 184–197 Contents lists available at ScienceDirect Industrial Crops & Products journal homepage: www.elsevier...

2MB Sizes 0 Downloads 83 Views

Industrial Crops & Products 130 (2019) 184–197

Contents lists available at ScienceDirect

Industrial Crops & Products journal homepage: www.elsevier.com/locate/indcrop

Production of 5-hydroxymethylfurfural and levulinic acid from lignocellulosic biomass and catalytic upgradation Xiaoyun Lia,b, Rui Xua, Jiaxin Yanga, Shuangxi Nieb, Dan Liue, Ying Liue, Chuanling Sia,b,c,d,e,

T ⁎

a

Tianjin Key Laboratory of Pulp and Paper, Tianjin University of Science and Technology, Tianjin 300457, China Guangxi Key Laboratory of Clean Pulp & Papermaking and Pollution Control, College of Light Industry and Food Engineering, Guangxi University, Nanning 530004, China c State Key Laboratory of Tree Genetics and Breeding, Northeast Forestry University, Harbin 150040, China d State Key Laboratory of Pulp & Paper Engineering, South China University of Technology, Guangzhou 510640, China e Tianjin Jianfeng Natural Product R&D Co., Ltd., Tianjin 300457, China b

A R T I C LE I N FO

A B S T R A C T

Keywords: 5-Hydroxymethylfurfural Levulinic acid Lignocellulosic biomass Pretreatment Catalytic Upgrading application

5-Hydroxymethylfurfural (HMF) and levulinic acid (LA), produced from lignocellulosic biomass, can be catalytically upgraded to valuable chemicals and fuels. Lignocellulosic biomass has been considered as an alternative carbon resource for fossil. This work reviews the recent development of HMF and LA production from lignocellulosic biomass, which focuses on pretreatment of lignocellulosic biomass, preparation principle for HMF and LA and the upgrading application. Traditionally, the preparation of HMF and LA is catalyzed by mineral acids in aqueous solutions. Recently, using organic solvents as reaction medium have been investigated in many reports. We carefully summary the production process of HMF and LA with homogeneous or heterogeneous catalysts in organic solvents; meanwhile, the advantages and challenges in different reaction systems are also pointed out. Furthermore, the applications of HMF and LA are described in details.

1. Introduction Lignocellulosic biomass is a long-term alternative carbon source to fossil, and it could be used as raw material to prepare liquid fuels and valuable chemicals (Dai et al., 2018; Delidovich et al., 2014; Mariscal et al., 2016; Rebeiz et al., 1984). Lignocellulosic biomass is cheap, abundant and widely available, and is considered as a sustainable carbon source (Lin et al., 2018). Plant biomass accounts for half of the total lignocellulosic biomass. The annual output of terrestrial plants in the world is about 1.7–2.0 × 1011 tons. Lignocellulosic biomass, rich in reserves and widely distributed, can be used for the sustainable production of biofuels and chemicals (Nie et al., 2018). Sources of lignocellulosic biomass mainly include agricultural crop residues (such as wheat straw, bagasse, rice husk, corn cob), forestry by-products (such as pine wood, aspen, sawdust) and algal biomass (Zhang et al., 2010). In China, a large number of lignocellulosic biomass is derived from agricultural solid waste. However, these agricultural solid wastes have not been effectively utilized, causing waste of resource and environmental pollution. Over the recent years, the catalytic conversion of lignocellulosic biomass has attracted increasing attentions. The use of lignocellulosic biomass as feedstock for the production of biofuels and chemicals not only reduces production costs but also



decreases carbon dioxide emissions. Among these chemicals, 5-hydroxymethylfurfural (HMF) and levulinic acid (LA) are valuable platform compounds, which are intermediates for preparing many valuable chemicals. HMF made from hexose (glucose or fructose) and lignocellulosic biomass is one of the most appealing platform chemicals. Many catalytic systems have been reported for the efficient preparation of 5-HMF (Alam et al., 2018; Galkin et al., 2016; Widsten et al., 2018; Zuo et al., 2017). HMF can be transformed to a number of useful compounds, such as, 5-arylaminomethyl-2-furanmethanol, 5-hydroxymethylfuroic acid, furfuryl alcohol, LA, levulinate esters (Hu et al., 2017) and 5-ethoxymethyl furfural (Kumari et al., 2018). LA is one of the 12 most valuable platform compounds identified by the U.S. Department of Energy. Nowadays, LA has received extensive attention in the preparation of pharmaceuticals, dyes, coatings, pesticides, plastic additives, resins and lubricant additives. In addition, LA can be upgraded to liquid fuels. LA and methanol (or ethanol) produce methyl and ethyl esters under the action of acidic catalysts. The proportion of these esters can reach 20% in mixed diesel fuel. LA can also be used in the preparation of 2-methyltetrahydrofuran, which could be mixed with gasoline up to 70% without the need of engine modification (Obregón et al., 2016). LA is further derivatized to synthesize pyridinyl levulinate, 2-

Corresponding author at: Tianjin Key Laboratory of Pulp and Paper, Tianjin University of Science and Technology, Tianjin 300457, China.

https://doi.org/10.1016/j.indcrop.2018.12.082 Received 6 September 2018; Received in revised form 9 December 2018; Accepted 26 December 2018 0926-6690/ © 2018 Elsevier B.V. All rights reserved.

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

made up of phenylpropane units through carbon-carbon and ether bonds. It is difficult for microbes to decompose the structure of this connection. The most inert component in the cell wall of plants is lignin. Cell wall is difficult to digest because of its high lignin content. Besides the three principal polymers of cellulose (30–50%), hemicellulose (20–35%) and lignin (15–20%), lignocellulosic biomass also includes pectin, organic solvent extracts, and ash (Wyman et al., 2005).

methyl-3-indole acetic acid, other pesticide intermediates and 2-mercapto-4-methyl-5-thiazole acetic acid, indomethacin pharmaceutical intermediates. The calcium decanoate produced by LA and calcium carbonate can be used for intravenous injection in medical field, which is conducive to bone formation and can maintain the excitability of muscles and nerves. LA and its derivatives are added to skin care products or cosmetics for sebum secretion, acne, ntibacterial anti-inflammatory treatment. In addition, LA can offer raw material to jasmine flavors. There are two reaction paths for the preparation of LA. The first one is catalyzed hydrolysis of furfuryl alcohol. In this path, furfural is initially hydrogenated to furfuryl alcohol. After acid catalysis, LA is formed through hydrolysis, ring opening and rearrangement reactions. This process is used by chemical companies such as French Organic Synthesis and U.S. Fitch. The second path to produce LA is direct hydrolysis of lignocellulosic biomass. It has been reported that strong acids can catalyze the conversion of carbo hexaose (eg, glucose, fructose) or cellulose, lignocellulosic biomass to LA and formic acid in aqueous solution in the temperature range of 140–225 °C (Mariscal et al., 2016; O’Neill et al., 2013; Sitthisa et al., 2011b; D. Sun et al., 2016; Z. Sun et al., 2016). The most effective acidic catalysts are sulfuric acid, hydrochloric acid or trifluoroacetic acid, with acid concentrations ranging from 0.1 to 2 mol/L. Carbo hexaose is firstly dehydrated to form HMF and then rapidly converted to LA and formic acid. In the Biofine process, biomass feedstock was hydrolyzed and dehydrated to 5-hydroxyfurfural after mixing with 1.5–3% sulfuric acid, 210–220 °C in a plug flow reactor for 12 s, by reducing the formation of degradation products. In the second reactor, 5-hydroxyfurfural was converted to LA and formic acid with residence time at 190–200 °C for 20 min. The reaction conditions were optimized to reduce the formation of formic acid and furfural, and the maximum yield of LA observed was 70%. In order to reduce the formation of formic acid and furfural, the reaction conditions were optimized to obtain 70% yield of LA. Because of this process (Hayes et al., 2006), Biofine company won the small business award-"Presidential Green Chemistry Challenge Award".

3. Lignocellulosic biomass pretreatment technology As shown in Fig. 3, the purpose of the lignocellulosic biomass pretreatment technique is to disrupt the complex network structure among hemicellulose, cellulose and lignin in the cellulosic biomass, providing favorable conditions for the conversion of lignocellulosic biomass into HMF and LA. The pretreatment methods for lignocellulosic biomass are mainly divided into physical, chemical and biological methods. The principles of choosing a pretreatment method are to minimize extra energy consumption and have good compatibility with the next process. 3.1. Physical pretreatment The physical pretreatment methods of lignocellulosic biomass include mechanical crushing, ultrasonic and microwave pretreatments. Mechanical crushing is the smashing of cellulosic biomass into granules by mechanical milling, reducing the crystallinity of cellulose, increasing the contact area between the catalyst and the raw materials in the reaction process, and increasing the hydrolysis efficiency. Ball-milling, which is one of milling methods, has been investigated in several studies (Liao et al., 2014; Liu et al., 2016). This mechanical technique leads to the decrease of the particles size and crystallinity of cellulose by cleaving hydrogen bonds in cellulose, therefore the yield of glucose is improved (Boissou et al., 2015; Nemoto et al., 2017). The particle size of cellulose was decreased by jet-milling treatment without reducing the crystallinity index, while the crystalline of cellulose was converted into amorphous within 1 h by methods of ball-milling and rob-milling (Avolio et al., 2012; Suzuki and Nakagami, 1999). The disadvantage of mechanical crushing pretreatment is that the smaller the particle size of the raw material, the higher the energy consumption. The principle of ultrasonic and microwave pretreatments is that the molecular structures of hemicellulose, cellulose and lignin in cellulosic biomass are changed. After the ultrasonic and microwave pretreatments, the crystalline structure of cellulose is destroyed, the crystallinity is reduced, and the raw material hydrolysis efficiency is improved (Ha et al., 2011). Chen and co-workers observed that the diameter of cellulose was decreased into 5–20 nm with ultrasonic treatment at a higher output power (> 1000 w) (Peleteiro et al., 2015). The cost of this pretreatment method is high and cannot meet the needs of industrialized large-scale applications. Steam explosion, discovered by Mason in 1925, has been considered as a suitable pretreatment method of lignocellulosic biomass (Biermann et al., 1984), which has been applied to the industry of fiberboard originally. The biomass structure was destroyed by sudden decompression in “Mansonite Wet-Form Process” (Spalt, 1977). Steam explosion is a relatively inexpensive because there is no need of external catalyst addition (Schwald et al., 1989).

2. The structure of lignocellulosic biomass Lignocellulosic biomass mainly consists of cellulose, hemicellulose and lignin, which is a fibrous material in the cell walls of plants (Fig. 1). Cellulose is a linear crystalline polymer composed of glucose (Xie et al., 2018). Hemicellulose is an amorphous polymer, which composes of carbo pentaose and carbo hexose. And lignin has the skeleton that supports the plant cell wall, surrounded by cellulose and hemicellulose. Cellulose is a glucose polymer with β-glycosidic bonds, which is connected by a network of inter- and intra-molecular hydrogen bonding, and the crystal structure is stabilized by van der Waals forces as presented in Fig. 2. The properties of the hemicellulose and lignin differ with the type of biomass. Xylan is the major component of hemicellulose, though arabinan, hemicellulosic glucan, galactan, and mannan are also present. The key structure of hemicellulose is composed of a linear backbone of β(1→4) linked D-xylopyranosyl residues. Glycosyl groups of hemicelluloses in plant species are different. The hemicellulose in the plants of the family Gramineae is the poly (arabinose-4-O-methyl glucuronide (Mamman et al., 2008)). The structural formula of hemicellulose is characterized as the D-xylose base is connected to the main chain through β-1,4-glycosides, while the branched chain is formed by the L-furan and 4-O-methyl-D-furan glucuronides on the C2 or C3 of the main chain. In addition, O-acetylgroups are also occasionally connected to chains. The main component of hemicellulose in cork is glucomannan (Hendriks and Zeeman, 2009). Lignin is a structural skeleton in plant cell walls. The degree of lignification is involved with the compressive strength of stems and the dehydration of plant cell walls. Lignin is an amorphous aromatic network polymer (Wang et al., 2018a), which is

3.2. Chemical pretreatment Chemical pretreatment includes hydrothermal, dilute acid and alkali pretreatment. Hydrothermal pretreatment is generally carried out at the temperature range of 150–260 °C, which resolves lignocellulosic biomass into soluble part and solid residue. After liquid hot water (LHW)pretreatment, 95.8 mol% of the C6-fractions in the raw bagasse were recovered (Schmidt et al., 2017). The high-temperature liquid water pretreatment offers several advantages: simple operation, high recovery rate of monosaccharide, no need for separation of catalyst. 185

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Fig. 1. Lignocellulosic biomass composition and representative structures of constituent components in plants. Adapted with permission from Ref. (Ratanakhanokchai et al., 2013).

Fig. 2. Cellulose from intra-molecular condensation of β-glucose at the1,4-position. Adapted with permission from Ref. (Alonso et al., 2013b).

Additionally, hydrothermal pretreatment is a method with low cost and high process compatibility. The purpose of dilute acid pretreatment is to produce sugar monomers. The acids used in the pretreatment include oxalic acid, sulfuric acid, hydrochloric acid and phosphoric acid. Dilute acid pretreatment has been employed to obtain fermentable sugars in bioethanol production process with the commercial application scale. The intermediate product (mainly glucose) in bioethanol production is the same as the production of HMF and LA. Therefore, a modification of the hydrolysis processes could be used in the preparation of HMF and LA. The red maple wood was soaked in 0.5% dilute sulfuric acid, heated for 27.5 min at 170 °C in the autoclave, and 84.4% of the hemicellulose in the raw material was recovered, and 78.8% of the sugar was the xylose monomer (Zhang et al., 2013a). Although this method is effective for the hydrolysis of oligomers, sulfuric acid used in the process was corrosive to the reaction equipment. In order to reduce the corrosiveness, oxalic acid was used as catalyst instead of inorganic acid, and a similar yield of carbohydrates was observed comparing with dilute sulfuric acid pretreatment (Zhang et al., 2013b). The principle of alkali pretreatment is that lignin in lignocellulose is dissolved in alkaline solution (sodium hydroxide, potassium hydroxide

Fig. 3. Schematic representation of lignocellulosic biomass pretreatment. Adapted with permission from Ref. (Gu et al., 2018).

and calcium hydroxide), which destroys the network structure of the wood cellulose. The hydroxyl ion in alkali solution weakens the hydrogen bond between cellulose and hemicelluloses, and saponification between hemicelluloses and other components. Consequently, alkali pretreatment causes the increase of lignocellulose porosity, the decrease of ester bonds, partial dissolution of hemicellulose, and the decrease of cellulose crystallinity. The enzymatic products of energy cane bagasse pretreated with dilute ammonia contains sugars (xylose and glucose) and non-sugar compounds (acetic acid, formic acid, furfural, HMF, levulinic acid and phenolic compounds). Compared with acid hydrolysis pretreatment, the enzymatic hydrolysis after dilute ammonia 186

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Fig. 4. Production of 5-hydroxymethylfurfural and levulinic acid from lignocellulose. Adapted with permission from Ref. (Amarasekara et al., 2008; Antal et al., 1990a, b; Brown et al., 1982; Rackemann and Doherty, 2011; Wolfrom et al., 1948; Zhang and Weitz, 2012.).

lignocellulosic derivatives such as furfural, HMF and LA. Compared with physical/chemical pretreatments, there are several advantages associated with biological pretreatment such as low energy consumption, high yield of targeted products and no formation of toxicants (RiaMillati et al., 2011; Saritha et al., 2012). However, biological pretreatment needs strict control of growth conditions and large space, and it takes a long time to carry out (Chandra et al., 2007).

pretreatment generats fewer amounts of non-sugar compounds, because less hemicellulose dissolved under the milder treatment conditions (Deng and Aita, 2018a). The flocculants polyethylenimine (PEI) was also applied to remove the non-sugar compounds from enzymatic hydrolysate of energy cane bagasse with dilute ammonia pretreated (Deng et al., 2018b). The drawbacks of this pretreatment method is that alkali used is costly, and the waste liquid is difficult to recover, which result in environmental problems. The pretreatment of lignocellulosic biomass led to 18–20% of total expense in the whole process because of the energy requirement (Eggeman and Elander, 2005; Mosier et al., 2005). Therefore, it is necessary to choose an energy-saving pretreatment method to decrease the expense of pretreatment. Furthermore, the feasibility of large-scale production needs to be took into consideration to make it a commercially productive process.

4. Synthesis of HMF and LA Originally, glucose and fructose were used as raw materials for the synthesis of HMF and LA. The dehydration reaction of fructose to HMF easily takes place over acid catalyst. Glucose need to be isomerized into fructose firstly then converted to HMF. Glucose was proved to be a good feedstock for extensive preparation of HMF and LA because of the low cost. Additionally, lignocellulosic biomass as a renewable carbon source, is also used as raw material for the preparation of HMF and LA. The conversion of lignocellulosic biomass into HMF and LA is very important to produce fuels and value-added chemicals, for lignocellulosic biomass is easily obtained in nature and non-edible.

3.3. Biological pretreatment The biological pretreatment uses natural microorganisms. They secret hydrolytic enzyme and ligninolytic enzyme, which result in depolymerization of lignin, and the lignocellulose are altered or degraded (Pérez et al., 2002). In biological pretreatment, cell wall structure of biomass is destroyed, which is beneficial to the hydrolysis of cellulose and hemicellulose (Sharma et al., 2017). Monomeric sugars generated from biopolymers in biological pretreatment are used for preparation of

4.1. Mechanistic studies of HMF production Currently, the preparation of HMF is mainly started from lignocellulosic biomass. The conversion of lignocellulose into HMF 187

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

due to the complex structure of lignocellulosic biomass.

generally includes three steps: (1) hydrolysis of cellulose into glucose, (2) isomerization of glucose into fructose, (3) dehydration of fructose into HMF and by-products such as humins, 2-hydroxy acetyl (Jr et al., 1990) and 5-methyl furfural (Brown et al., 2010). The conversion of C6 sugars into HMF was investigated in many reports, and the mechanisms of acid-catalyzed production process has been carefully studied, in which nuclear magnetic resonance spectroscopy was applied to elucidate intermediate products and explore the possible reaction pathways. Fructose is initially produced from glucose through mutarotation and isomerization. To explain the dehydration process of fructose into HMF, there are two types of mechanisms proposed by researchers: a cyclic route through the fructofuranosyl intermediate and an acyclic route through the 1,2-enediol intermediate (Kuster, 1990). The cyclic route begins with cyclic fructofuranose. A tertiary carbenium cation is produced from the dehydration of the hemiacetal at C2. Then HMF is formed through two consecutive β-dehydrations (van Putten et al., 2013). In acyclic pathway, fructose was dehydrated into 1,2-enediol (an intermediate of the reaction), which is assumed as the rate-limiting step. HMF is produced via two consecutive β-dehydrations and a ring closure reaction (van Putten et al., 2013).

5. Solvents 5.1. Water as solvent Application of water as solvent in the production of HMF and LA usually caused low yields. Cai reported that yield of HMF was only 2.4% from maple wood using 1 wt% H2SO4 at 170 °C for 40 min (Cai et al., 2014). The low yield of HMF (3.4%) was also obtained when HCl (pH 0.21) solution was applied as catalyst in water (Yemiş and Mazza, 2012). 5-HMF was observed in the preparation process of hydrothermal carbonization (HTC) and the production of 5-HMF varied with the HTC duration (Borrero-López et al., 2018). Recently, hydrothermal carbonbased catalyst (HTC-SO3H) was used in the HMF production from cellulose in water, and 46% of HMF yield was observed at 180 °C for 5 min (Wataniyakul et al., 2018). Algal biomass was used as raw materials in Kang’s study. Firstly, algal biomass was soaked in sulfuric acid and the galactose in the soaking solution is used to prepare LA. The solid residue after soaking is used for fermentation to produce ethanol. According to the literature, the yield of LA was 45.88% (Kang et al., 2013). The maximum yield of nearly 60% of the LA is obtained, when the reaction condition is 99.6 mM of pure cellulose as feedstock, 0.927 mol/L of hydrochloric acid as catalyst at temperature of 180–200 °C (Shen and Wyman, 2012). LA is easily degraded into humins in strong acid condition, resulting in low yield of LA. Yi et al. investigated the reaction of corn straw in the hydrothermal system with AlCl3 at 140 °C and the conversion rate of the hemicellulose component after 1 h was 85.1%, and the conversion rate of cellulose and lignin was 10.7% and 23.9% (Yi et al., 2013), respectively. Li et al. Studied the preparation of acetylpropionic acid by catalyzing the hydrolysis of corn cob slag in AlCl3-NaCl system, using AlCl3 as the catalyst and sodium chloride as a hydrothermal method. At a lower temperature, the yield of acetylpropionic acid was up to 46.8% (Li et al., 2015). Zheng et al. (2017) investigated the preparation of LA from corn stalk with FeCl3 as catalyst, and the observed maximum yield of LA was 48.89% at 180 °C for 10 min catalyzed by 0.5 mol/L FeCl3 solution. Wang et al. (2018b) reported that a peak yield of LA was 68.0 mol % over 20 g/L FeCl3 in 40 wt% NaCl solutions, which was higher than the non-NaCl systems (48.5 mol %). The results indicated that NaCl improved the dissolution rate and the hydrolysis rate of cellulose in xylose residues. Although the homogeneous inorganic acid and metal chloride catalyst have the characteristics of high catalytic efficiency and simple to operate, there exists some problems, such as serious corrosion of equipment, high cost of acid recovery and difficulty in purification of LA. In order to alleviate the above problems, some researchers considered that the solid acid was a good choice to replace inorganic acid and metal chloride catalyst in the hydrolysis reaction. Alkaline-treated zeolite was used as catalyst in the conversion of xylose into LA, and the peak of LA yield was 30% with the modified zeolite in 0.25 M NaOH

4.2. Mechanistic studies of LA production As shown in Fig. 4, the furan cycles of HMF molecule are opened through dehydration steps, then 2,5-dioxo-3-hexenal is produced. 5,5Dihydroxypent-3-en-2-one and formic acid are produced after the CeC bond between C-1 and C-2 of 2,5-dioxo-3-hexenal was broken. Eventually, LA is generated from 2,5-dioxo-3-hexena through several steps. The mechanism for the production of LA was presented in details by Horvat et al. (1985). Production of valuable chemicals from lignocellulosic biomass is an important and practical process, because lignocellulosic biomass is widely distributed, abundant in nature and renewable (Huber et al., 2006). As shown in Fig. 5, LA can be prepared from the cellulose in lignocellulosic biomass, which includes hydrolysis of cellulose into glucose, production of HMF from glucose, and formation of LA from HMF. However, there are many difficulties in this process: (1) The crystalline structure of cellulose packed densely with the inter- and intermolecular hydrogen bond, (2) The cellulose is insoluble in water and other solvents, (3) the chemical characteristics of cellulose is stable under many conditions. The production of LA can also begin with hemicellulose (C5 sugars), as presented in Fig. 5. The reaction process consists of three steps. The first step is formation of furfural from xylose catalyzed by acid. The second step is reduction of furfural into furfuryl alcohol catalyzed by metal. The last step is hydrolysis of furfuryl alcohol into LA catalyzed by Brønsted acid. The conversion of furfuryl alcohol into LA is one of the integration reactions of glucose and xylose reaction pathways (Mellmer et al., 2015). However, there are some challenges for the preparation of HMF and LA from real lignocellulosic biomass: stringent reaction conditions (such as higher reaction temperature and catalyst concentration) and low yield of target chemicals,

Fig. 5. Production of HMF and LA from lignocellulosic biomass. Adapted with permission from Ref. (Mellmer et al., 2015). 188

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Table 1 Production of HMF from sugars and biomass waste by homogeneous catalysts in organic solvent and biphasic. Feed

Catalyst

Solvent

Temperature (°C)

Time (min)

HMF yield

References

Fructose Wood ear mushroom Fructose Barley husk

– [BBIM-SO3][NTf2] HCl Sulphanilic acid

130 140 100 150

60 15 240 60

67.2 wt% 47 wt% 90.3 mol% 41 mol%

(Jia et al., 2014) (Alam et al., 2018) (Zuo et al., 2018) (Karimi and Mohamad-Mirzaei, 2016)

Mixed saccharide Mixed saccharide Vegetable waste Bread waste Eucalyptus

H3PO4 NaHSO4/ZnSO4, SnCl4 SnCl4 Levulinic acid

DMSO DMA-LiCl DES/acetonitrile Water-DMSO/2butanol-MIBK Acetone/water THF/water DMSO/water DMSO/water MTHF/water

180 160 140 160 180

80 90 20 20 15

36 mol% 40 mol% 4.9 wt% 21.4 wt% 93 mg/L

(Widsten et al., 2018) (Widsten et al., 2018) (Yu et al., 2016) (Yu et al., 2017) (Seemala et al., 2016)

into the organic phase, which was separated from the acidic aqueous phase (Luterbacher et al., 2014). As shown in Table 1, the DES/acetonitrile biphasic system resulted in 90.3 mol% of HMF from fructose catalyzed by HCl (Zuo et al., 2018). HMF at 40 mol% was obtained using mixed saccharide as feedbacks with H3PO4 in acetone/water (Widsten et al., 2018). Biphasic systems were also used in the hydrolysis process of lignocellulosic biomass such as vegetable waste (Yu et al., 2017), bread waste (Yu et al., 2017), Eucalyptus (Seemala et al., 2016) and straw (Karimi and Mohamad-Mirzaei, 2016). As shown in Table 2, a series of functionalized ionic liquids (SFILs) were prepared by microwave-assisted method. The maximum yield of LA was 55.0%, which was obtained with SO3H-functionalized ionic liquids as catalysts (Ren et al., 2013). HPA ionic liquids were applied in the conversion of cellulose into LA in a water/methyl isobutyl ketone (MIBK) system, and the peak yield of LA (63.1%) was obtained at 140 °C for 12 h. Moreover, separation of products and catalysts was easy, and the activity of the recycled HPA ionic liquids was not obviously decreased (Sun et al., 2012). The acidic ionic liquids were also used as catalyst for preparation of LA from cellulose. The peak of LA from cellulose was 86.1% under the reaction condition of 1 g of IL and 6 g of water at 170 °C for 5 h (Ren et al., 2015). The application of dicationic ionic liquids as a catalyst in LA preparation was studied by Khan et al, and the maximum LA yield was 55% over [C4(Mim)2] [(2HSO)(H2SO4)2] at 100 °C for 3 h (Khan et al., 2018). Biphasic reactor systems were also applied for the production of LA. A 57 wt% of LA yield from rice straw was obtained in the hydrochloric acid/dichloromethane system (Kumar et al., 2018). Water/paraffin oil biphasic process was proposed for production of levulinic acid, and the maximum yield of LA reached to 45 mol% from corn grain (Licursi et al., 2018).

(Chamnankid et al., 2014). Weingarten et al. (2012) used Amberlyst 70 as a solid acid catalyst to convert the cellulose into LA in aqueous phase, while the maximum of LA yield was only 28%. The solid acid in aqueous phase generally reveal several drawbacks such as low yield, long reaction time and high price. It is necessary to improve the catalytic activity of solid acids to alleviate these drawbacks. 5.2. Organic solvents and biphasic systems Recently many liquid solvents have been used in the hydrolysis process, which are substitute for water and organic solvents (Mushrif and Vasudevan, 2015), ionic liquid (Aylak et al., 2017; Xiao and Song, 2014) and biphasic systems (Dumesic et al., 2013; Kumar et al., 2018; Rackemann and Doherty, 2015). The types of catalysts are divided into homogeneous and heterogeneous catalysts. 5.2.1. Homogeneous catalyst in organic solvents and biphasic systems The moderate polarity solvents have been applied in the HMF production process (Table 1). An HMF yield of 67% was observed at 130 °C using fructose as crude material with no catalyst in dimethyl sulfoxide (DMSO) (Jia et al., 2014). DMSO is a favorable solvent for the fructose dehydration, because DMSO decreased the production of byproducts and possess hermodynamic advantages (Sim et al., 2012; Yu and Tsang, 2017). Additionally, the yield of HMF reached to 47 wt% from wood ear mushroom using [BBIM-SO3][NTf2] as the catalyst in DMA-LiCl systems (Alam et al., 2018). As shown in Fig. 6, biphasic systems were applied in the conversion of biomass into furans, for target chemical were continuously extracted

5.2.2. Heterogeneous catalyst in organic solvents and biphasic systems The homogeneous catalyst used to produce HMF and LA is difficult to be separated from the products and reaction system. Additionally, the waste liquid formed in the reaction process causes environment pollution. Aiming to solve these problems, solid acid catalysts attract researchers’ attentions. Common solid acid catalysts include carbonbased catalysts, molecular sieves, ion-exchanged resins and heteropoly acids. Carbon materials own a variety of surface structures, such as amorphous surface and non-aromatic ring, which are made up of surface −OH groups. Furthermore, carbon materials have many advantages: high porosity, large specific surface area, perfect electron conductivity, etc. Carbon materials could be converted to be Brønsted acid catalyst after modified with –SO3H group. Molecular sieves act as solid acid catalysts, which are applied in the preparation process of HMF and LA. Commonly used molecular sieve types are H-ZSM-5, Hmordenite, HY molecular sieve, H-Beta, SAPO, etc. Jeong et al. used NaOH modified-zeolite Y as catalysts for production of LA from C5 sugar, and 42.7% of LA yield was obtained at 190 °C for 180 min (Jeong et al., 2018). The kinetics of HMF production from glucose over HBEA25 zeolite in water was reported by Swift et al., and the peak yield of

Fig. 6. Biphasic systems for the dehydration of carbohydrates into valuable chemicals. 189

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Table 2 LA production from fructose, glucose and polymer using homogeneous catalyst in organic solvents and biphasic. Feed

Catalyst

Solvent

Temperature (°C)

Time (min)

LA yield (wt%)

References

Fructose Glucose Cellulose Cellulose Cellulose Cellulose Starch Rice straw Corn grain

[C3SO3Hmim]HSO4 [C3SO3Hmim]HSO4 SO3H-functionalized ILs HPA ionic liquids [C3SO3Hmim]HSO4 [C4(Mim)2][(2HSO4)(H2SO4)2 HPA ionic liquids HCl H2SO4

Water/[C3SO3Hmim]HSO4 Water/[C3SO3Hmim]HSO4 SO3H-functionalized ILs MIBK Water/[C3SO3Hmim]HSO4 Water/[C4(Mim)2] [(2HSO4)(H2SO4)2 MIBK Water/dichloromethane Water/paraffin oil

170 170 160 140 170 100 140 200 190

300 300 30 720 300 180 300 180 60

76.7 wt% 86.1 wt% 55 wt% 63.1 wt% 65.6 wt% 55 wt% 48.7 wt% 57 wt% 45 mol%,

(Ren et al., 2015) (Ren et al., 2015) (Ren et al., 2013) (Sun et al., 2012) (Ren et al., 2015) (Khan et al., 2018) (Sun et al., 2012) Kumar et al., 2018) Licursi et al.,2018

reaction temperature and a large number of strong acid sites, it usually leads to the formation of the degradation product, humins. Combination of ultrasonic irradiation and alumina loading on SBA-15 was applied in the conversion of fructose into HMF, and the HMF yield was 47% observed at room temperature under sonication treatment (Babaei et al., 2018). Sn20/γ-Al2O3 and Fe3O4@SiO2-SO3H catalysts were synthesized for the dehydration of glucose into HMF, and caused 27.5% and 70% HMF yield, respectively (Marianou et al., 2018; Elsayed et al., 2018). Additionally, several metal-containing silicoaluminophosphate molecular sieves (MeSAPO-11, MeSAPO-5, MeSAPO-34, MeSAPO-44) were prepared for the HMF production, and MeSAPO-11 showed the highest catalytic activity, which was attributed to more acid sites (Sun et al., 2018). Therefore, when the solid acid catalyst is designed, the number of strong acid sites needs to be optimized to improve the catalytic activity, which is beneficial to the production of HMF and LA. In addition, the type of acid sites in solid acid catalysts is also an important factor determining the reaction pathway. Ionic liquids are green solvents, which have the characteristics of high boiling point, low volatility and high polarity, can be used in the production process of platform compounds (HMF, LA and furfural). Ionic liquids are liquid at room temperature, composed of organic macromolecules and small molecules of anions. Their melting point is lower than 100 °C. Ionic liquids are mainly sorted into alkyl imidazolium, quaternary ammonium salt, alkyl pyridine and quaternary phosphonium salts according to their cations. The carbonaceous solid acid derived from corn stalk was applied in the conversion of corn stalk into HMF in [BMIM][Cl] ionic liquid, and the HMF yield reached to 44.1% at 150 °C for 30 min. The application of corn stalk-derived carbon catalyst offers a green and efficient conversion process of lignocellulosic biomass into HMF (Yan et al., 2014). Zhang et al. investigated the conversion of lignocellulosic biomass (pine, straw and corn straw) to HMF and furfural over CrCl3 in imidazole [C4MIm]Cl ionic liquids with microwave heating, and 45–52% of HMF yield was obtained (Zhang and Zhao, 2010). Application of ionic liquids as solvents could decrease the reaction temperature. Due to the high cost of ionic liquids synthesis, ionic liquids as solvents are difficult to be applied in the large-scale production. The solid acid conjugated ionic liquid is a green way for biomass conversion. γ-Valerolactone (GVL), derived from lignocellulosic biomass, is an ideal alternative to the sustainable non-polar solvent in the process of biomass conversion to platform compounds. GVL can improve the catalytic activity and selectivity of target chemicals (Alonso et al., 2013a; Li et al., 2018; Qi and Horváth, 2012). The application of GVL as a solvent in the production of HMF and LA from lignocellulosic biomass offers several advantages. (1) GVL is derived from lignocellulose and is an excellent green solvent that can be used for the preparation of valuable chemicals (Li et al., 2017). (2) Wet biomass can be employed in the conversion process due to GVL is miscible with water. (3) There is no need for any neutralization/purification steps in upgrading process of LA to GVL. (4) GVL results in a high cellulose deconstruction rate, and the regeneration of solvents is easy, which supply an ideal process for formation of LA from lignocellulose (Alonso et al., 2013a).

Fig. 7. Brønsted and Lewis acid sites in solid acid catalysts: “BA” represents Brönsted acid sites and “LA” represents Lewis acid sites. Adapted with permission from Ref.(Bhaumik and Dhepe, 2016).

HMF was observed when the ratio of Lewis to Brønsted acid sites was about 0.3 (Swift et al., 2016). Molecular sieves have both Brønsted and Lewis acid sites. Ion-exchanged resins bearing –SO3H groups, such as Amberlyst-15 and Nafion-117, are also Brønsted acid. Heteropoly acids (HPA) as typical Brønsted acid are composed of various oxoacids (transition metal-oxygen anion clusters). Heteropoly acids, which is originally homogeneous, become heterogeneous when the protons are partially replaced by bigger ions (such as Cs+, Ag+). The Brønsted and Lewis acid sites in solid acid catalysts are presented in Fig. 7. As shown in Table 3, the preparation of HMF from fructose was catalyzed by mesoporous silica heterogenized perfluorosulfonic acid resin in DMSO. 85% Yield of HMF was observed at 190 °C for 120 min (Dou et al., 2018). The conversion of aquatic microalgae into HMF with one-pot method was investigated by Wang et al. (Wang et al., 2016). The commercial H-ZSM-5 zeolites were used as catalysts in NaCl/THF system in the conversion process. The yield of HMF was 48.0% at 200 °C for 2 h. There are few reports of solid acid applied in the LA preparation, which may be due to the reaction of LA production using solid acid was difficult. Alonso et al. investigated the conversion of cellulose into LA using Amberlyst 70 as the catalyst, 69% of LA yield was obtained after 16 h at 190 °C (Alonso et al., 2013a). Under the conditions of high 190

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Table 3 HMF production from fructose, glucose and polymers using solid catalyst in organic solvents and biphasic. Feed

Catalyst

Solvent

Temperature (°C)

Time (min)

HMF yield (wt%)

References

Fructose Fructose Fructose Fructose Fructose Glucose Glucose Glucose Glucose Glucose Glucose Inulin Corn stalk Corn stalk Starch Cassava waste Microalgae

Silica heterogenized perfluorosulfonic acid resin Sulfonated carbon sphere from glucose. Sulfonated polyaniline. HMOR Alumina loading on SBA-15 Sn-SBA-15 SAPO-34 SAPO-44 Silica-supported phosphotungstic acid Sn20/γ-Al2O3 Fe3O4@SiO2-SO3H K10 clay-Cr3+ Carbonaceous solid acid catalyst Biochar-Mg-Sn SAPO-44 Sulfonated carbon-based catalyst H-ZSM-5

DMSO DMSO DMSO Water/MIBK DMSO Water/THF GVL Water/MIBK Water/Acetone Water/DMSO water/MIBK [BMIM]Cl [BMIM]Cl AMIMCl/isopropanol Water/MIBK Acetone/DMSO NaCl/THF

90 160 140 165 Room temperature 130 170 175 160 150 140 120 150 100 175 250 200

120 90 180 90 150 20 40 83 140 60 1440 120 30 180 360 60 120

85 90 71 74 47 36 93.6 67 78.3 27.5 70 83 44.1 63.57 68 30 48

(Dou et al., 2018) (Zhao et al., 2016) (Dai et al., 2017) (Ordomsky et al., 2012) (Babaei et al., 2018) (Gallo et al., 2013) (Zhang et al., 2017) (Bhaumik and Dhepe, 2013) (Huang et al., 2018) (Marianou et al., 2018) (Elsayed et al., 2018) (Román-Leshkov et al., 2010) (Yan et al., 2014) (Liu et al., 2018) (Bhaumik and Dhepe, 2013) (Daengprasert et al., 2011) (Wang et al., 2016)

(Morone et al., 2015; Yan et al., 2015). LA can also be used as lubricants or plasticizers, as well as used in ketals production. In addition, LA can be utilized to produce thermosets, polyols and thermoplastics.

SAPO-34 was prepared by Zhang et al. with hydrothermal synthesis and was utilized in the HMF production from hexoses using GVL as solvents (Table 3). The maximum yield of HMF was 93.6% at 170 °C. The SAPO-34 as catalyst can be recycled and reused for at least five runs without obviously loss of catalytic activity (Zhang et al., 2017). Wettstein et al. studied the preparation of LA and γ-valerolactone in a biphasic reaction system with cellulose as the raw materials (Wettstein et al., 2012). The yield of LA reached 70%, and 75% of LA was extracted into GVL. Because the solvent GVL was also a product, this method not only saved the step of separating the product from the solvent, but also reduced the accumulation of the solid humins species in the chemical reactor by using the renewable solvent. Therefore, application of GVL as solvent gives a great advantage in the process optimization and economic benefits in the conversion process of lignocellulosic into valuable materials.

6.1. 5-Aminolevulinic acid 5-Aminolevulinic acid (ALA) is a naturally-occurring substance existing in the cells of biological organism. ALA is not only a kind of nontoxic and pollution-free green agrochemicals (act as herbicides) (Rebeiz et al., 1984), but also a photodynamic therapy medicine used in cancer treatment (Bedwell et al., 1992). The 5-aminolevulinic acid and its derivatives are natural precursors of the photosensitizer in the photodynamic therapy (PDT), which attract many attentions in treatment for different cancers. The protoporphyrin IX (PpIX), which is a strong photosensitizer, are biosynthetically generated from 5-aminolevulinic acid and appropriate derivatives in the photodynamic therapy of 5aminolevulinic acid (Juzeniene et al., 2010). 5-Aminolevulinic acidphotodynamic therapy involves administration of 5-aminolevulinic acid, suitable derivatives, and protoporphyrin IX (PpIX) induced by 5aminolevulinic acid. The protoporphyrin IX (PpIX) is a powerful photosensitizer in photodynamic therapy. Singlet oxygen are generated from protoporphyrin IX in present of light, which breach tumor and tumor-associated cells. Additionally, protoporphyrin IX are used to detect the cancer cells (Hyun et al., 2009; Zenzen and Zankl, 2003). The amino group selectively introduced at the C5-position of LA in the ALA synthesis pathway as presented in Fig. 10. The LA molecule was brominated in alcohol solvent to activate the C5 position, and the mixtures of 5-bromoesters and 3-bromoesters were prepared after the amination process, which could be separated by distillation (MacDonald, 1974). The nucleophilic nitrogen species were used in the amination process of 5-bromolevulinate (Hyun-JoonHa et al., 1994). Many references have reported various methods of the ALA preparation (Dabrowski et al., 2003; Hayes et al., 2006; MacDonald, 1974). A typical method of ALA production contains three steps. The first chain is bromination of LA in methanol. The reaction between methyl 5-bromolevulinic acid and potassium phthalimide serves as the second step. The last step comes to the hydrolysis of the phthalimide derivative (Hyun et al., 1994). Dabrowski et al. proposed another reaction process of the 5-bromolevulinic acid synthesis, in which the 5-bromo ester was converted into ALA hydrochloride using an azide derivative (Dabrowski et al., 2003).

6. Transformation of HMF and LA to valuable chemical and fuels HMF and LA have been considered as building blocks for the production of petrochemicals. Such molecules would be proposed as green and sustainable substitutes to chemicals made from petroleum. HMF, a platform compound obtained from the conversion of biomass, can be used to prepare 5-arylaminomethyl-2-furanmethanol (AAMFM), LA (D. Sun et al., 2016; Z. Sun et al., 2016), levulinate esters (Hu et al., 2017; Zhang et al., 2019), 1,2,6-hexanetriol (HTO), 1-hydroxyhexane-2,5dione (HHD), 1,6-hexanediol (HDO) 5-chloromethylfurfural (CMF) and 5-ethoxymethylfurfural (5-EMF) (Fig. 8). 5-Arylaminomethyl-2-furanmethanol is a natural precursor of the photosensitizer in the photodynamic therapy used in the cancer treatment. 1,2,6-Hexanetriol is used as a humidity regulation and viscosity-controlling agent in the production of drugs and cosmetics. In addition, furandicarboxylic acid (FDCA) can be produced from HMF. FDCA was applied to make polyethylene furandicarboxylate (PEF), an alternative for polyethylene terephthalate (PET) (Rosatella et al., 2011). 2,5-Diformylfuran (DFF) and 2,5-bishydroxymethyl furan (BHMF) used as polymer precursors could be prepared from HMF (Rosatella et al., 2011; Hao et al., 2016). 5-Ethoxymethylfurfural (5-EMF) as a potential biofuel alternative compound could be yielded from HMF, and amberlyst-15 improved the production of 5-EMF from 5-HMF (Zuo et al., 2018). LA easily prepared from lignocellulosic biomass can be converted into a wide range of biofuels, including γ-valerolactone (GVL), 1,4pentanediol (1,4-PeD), 2-methyltetrahydrofuran (MTHF) and levulinate esters (Cui et al., 2018; Yan et al., 2015) (Fig. 9). Many high-valued chemicals are produced from LA, such as 5-aminolevulinic acid (ALA), 2-butanone, acetylacrylic acid, acrylic acid and diphenolic acid (DPA)

6.2. 1,2,6-Hexanetriol 1,2,6-Hexanetriol (HTO) is regarded as a humidity regulation and 191

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Fig. 8. Valuable chemicals and biofuels derive from HMF.

cleavage of carbonoxygen bond to produce a ring-opening intermediate (B2). The last step is the HTO formation from the hydrogenation of B2. Yao et al. reported that a series of mixed oxide such as Ni-Al, Co-Al and Ni-Co-Al were utilized as catalysts for the HTO production from HMF (Yao et al., 2013). Ni-Co-Al had higher catalytic activity than other mixed oxide catalysts due to the synergetic effect between Ni and Co. Ni was conducive to hydrogenation reaction of aldehyde group and furan ring of HMF, while Co results in the cleavage of carbon-oxygen bond of HMF (Nakagawa et al., 2012; Xu et al., 2011; Yang et al., 2015). The cooperation of Ni and Co species enhance the catalytic activity in the production of HTO from HMF through the hydrogenation-ringopening reaction.

viscosity-controlling agent as applying in the production of drugs and cosmetics (Endo and Sawada, 2001; Miura et al., 1999; Nolan et al., 2016, 2014). Recently, HTO was prepared through a hydrogenationring opening reaction using HMF as feedstock. A reaction path of the HTO production via hydrogenation-ring-opening reaction of HMF was showed in Fig. 11 (Chia et al., 2011; Sitthisa et al., 2011a; Xu et al., 2011). The first step was the formation of DHMF through the hydrogenation reaction of the aldehyde group in HMF. The second step is the adsorption of the furan ring of DHMF on 0.5Ni2.5CoAl with two modes: parallel (a) and tilted (b). The parallel mode could cause the total hydrogenation DHMF into DHMTHF. The tilted modes generate an important semi-hydrogenation intermediate (B1), which lead to the

Fig. 9. Valuable chemicals and biofuels derive from LA. 192

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Fig. 10. The conversion of LA into 5-aminolevulinic acid. Adapted with permission from reference. Adapted with permission from Ref. (Werpy et al., 2004).

Fig. 11. The HTO formation via hydrogenation-ring-opening of HMF. Adapted with permission from Ref. (Yao et al., 2013).

aromatic amines with electron donating groups > aromatic. Aromatic amines with halogen groups led to low yield of AAMFM, which was less than 30.0% in 24 h (Cukalovic and Stevens, 2010). To increase the yield of AAMFM, Xu et al. (2014) developed an efficient strategy, in which the AAMFM was produced from the direct reductive amination with dichloro-bis (2,9-dimethyl-1,10-phenanthroline) ruthenium (II) (Ru (DMP)2Cl2) in ethanol solvent (Xu et al., 2014). The yields of AAMFM reached 94.0%, 95.0% and 95.0% at 60 °C and 12 bar H2 for 20 h, when substrates for the reductive amination HMF were parafluoroaniline, parachloroaniline and parabromoaniline. Those results suggested that aromatic amines containing electron withdrawing groups were extremely reactive in this reaction process. Furthermore, the direct reductive amination HMF into AAMFM could extended to various secondary amines.

6.3. 5-Arylaminomethyl-2-furanmethanol 5-Arylaminomethyl-2-furanmethanol (AAMFM) presents an aminomethyl moiety and a hydroxymethyl moiety on the furan ring. AAMFM is an original material to prepare the pharmaceuticals such as calcium antagonists, muscarinic agonists, antifilarial agents and cholinergic agents (Agarwal et al., 2011; Chieffi et al., 2015; Cukalovic and Stevens, 2010; Kegnaes et al., 2012; Xu et al., 2014). The production of AAMFM generally starts from furfural or furfuryl alcohol (Chatterjee et al., 2016; Nishimura et al., 2016; Zhang et al., 2016). Increasing the C5 position of the furan ring is very difficult because of the inactivation of the C2 aldehyde group (Cukalovic and Stevens, 2010). Many studies reported that reaction conditions in this process were tough: the reaction time was delayed, reaction temperature and pressure were increased (Cukalovic and Stevens, 2010). Cukalovic and Stevens reported that AAMFM could be prepared from HMF through reductive amination reaction at room temperature and atmospheric pressure (Cukalovic and Stevens, 2010). This reaction process consisted of two key steps: the production of aldimine and in-situ reduction of aldimine with NaBH4 (Fig. 12). The decreasing order of amines reactivity in the reductive amination of HMF into AAMFM was as following: aliphatic amines >

6.4. γ-Valerolactone γ-Valerolactone (GVL) is a sustainable platform chemical to produce liquid alkenes, and it also acts as a green solvent for the conversion of lignocellulosic biomass. The mechanism for the production of GVL from LA has been proposed in many reports (Alonso et al., 2013b; Yan et al., 193

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Fig. 12. Reaction path of AAMFM from HMF through reductive amination. Adapted with permission from Ref. (Cukalovic and Stevens, 2010).

Fig. 13. Reaction pathways for production of γ-valeroclatone from LA.

(Chia and Dumesic, 2011). The maximum yield of GVL was 92% obtained in 2-butanol with ZrO2 as catalyst. Additionally, the Ni/NiO catalyst showed extremely high catalytic activity in the hydrogenation reaction of levulinic acid because Ni/NiO heterojunctions can improve the hydrogenation reactions (Song et al., 2017).

2015). There are several reaction routes to prepare GVL that begin with LA (Fig. 13). γ-Hydroxyvaleric acid is formed through the hydrogenation reaction of LA, and GVL is generated when the γ-hydroxyvaleric acid ring-closes by intramolecular esterification (Alonso et al., 2013b). A second possible reaction route begin with LA dehydration to product angelica lactone, then angelica lactone is hydrogenated to form GVL, with ring-closing by intramolecular esterification and loses a water molecule spontaneously to form GVL. This route needs an acidic condition, and generally results in lower yields because the coke produced from the angelica lactone with acidic catalyst (Serrano-Ruiz et al., 2010). When levulinic acid esters present, the pathway is similar to that starting with LA. The hydroxy levulinic ester is produced from levulinic acid esters in hydrogenation reaction, then ring closing reactions happen in intramolecular transesterification to form GVL and alcohol (Gürbüb et al., 2011; Starodubtseva et al., 2005). Hengne et al. (Hengne and Rode, 2012) reported that Cu-ZrO2 and Cu-Al2O3 nanocomposites catalyzed the hydrogenation of LA and its ester. GVL selectivity was more than 90% using Cu–ZrO2 and Cu–Al2O3 as catalyst. In presence of methanol, the corresponding ester was firstly produced, then hydrogenated to generate GVL and methanol. The catalyst is a good substitute for the noble metal catalyst, although it would lose selectivity when recycled, probably owing to metal sintering. Recently, CuAg/Al2O3 was used to catalyze the formation of GVL from LA, and the Cu leaching in the reaction was obviously restrained, due to the addition of Ag which lead to suppress Cu leaching (Zhang et al., 2018). The metal oxides, such as ZrO2 and γ-Al2O3, were applied to catalyze the production of GVL employing secondary alcohols as the solvent and hydrogen donor

6.5. 2-Methyltetrahydrofuran 2-Methyltetrahydrofuran (MTHF) is a fuel additive with high flammability, which could be added up to 30% by volume in fuels with no decrease of performance, and engine modifications are not needed. Noble-metal (Mizugaki et al., 2016) and nonnoble-metal catalysts (Upare et al., 2011) have been studied to produce MTHF from LA. Mizugaki et al. (2016) envisioned that the production of MTHF from LA via the intermediate 1,4-pentanedio (1,4-PeD), which was dehydrative cyclized to MTHF catalyzed by the solid acid (Fig. 14). The peak yield of MTHF (86%) was observed using Pt − Mo/H-β as catalyst at 130 °C for 24 h. This was caused by the facts that hydrogenation of LA to 1,4-PeD was enhanced by the synergistic effect between Pt NPs and MoOx, and the cyclodehydration of 1,4-PeD was catalyzed by H-β. 7. Conclusions HMF and LA, as valuable platform chemicals, could be upgraded to fuel and other important chemicals. In this work, we give a comprehensive review of the development in the production of HMF and LA from lignocellulosic biomass and their upgrading applications. An 194

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Fig. 14. Reaction path of MTHF from LA through 1,4-PeD. Adapted with permission from Ref. (Mizugaki et al., 2016).

energy-efficient pretreatment method is vital to reduce the total cost in the whole process. The number of strong acid sites and the type of acid sites need to be considered in preparation process of the solid acid catalyst, which are crucial factors for the formation of target chemicals. Ionic liquids and γ-valerolactone are green solvents applied in the production of HMF and LA. The upgradation of HMF and LA to valuable chemicals and fuels are also discussed. The synthetic process of 5-arylaminomethyl-2-furanmethanol, 1,2,6-hexanetriol, γ-valerolactone, 5aminolevulinic acid and 2-methyltetrahydrofuran (MTHF) as the derivatives of HMF and LA are described in details.

Boissou, F., Sayoud, N., Karine, D.O.V., Barakat, A., Marinkovic, S., Estrine, B., J, rôme, F., 2015. Acid‐assisted ball milling of cellulose as an efficient pretreatment process for the production of butyl glycosides. ChemSusChem 8, 3263–3269. Borrero-López, A.M., Masson, E., Celzard, A., Fierro, V., 2018. Modelling the reactions of cellulose, hemicellulose and lignin submitted to hydrothermal treatment. Ind. Crops Prod. 124, 919–930. Brown, D.W., Floyd, A.J., Kinsman, R.G., Ali, Y., 1982. Dehydration reactions of fructose in non-aqueous media. J. Chem. Technol. Biotechnol. 32, 920–924. Brown, D.W., Floyd, A.J., Kinsman, R.G., Roshanhyphen, Ali,Y., 2010. Dehydration reactions of fructose in non-aqueous media. J. Chem. Technol. Biol. 32, 920–924. Cai, C.M., Nagane, N., Kumar, R., Wyman, C.E., 2014. Coupling metal halides with a cosolvent to produce furfural and 5-HMF at high yields directly from lignocellulosic biomass as an integrated biofuels strategy. Green Chem. 16, 3819–3829. Chamnankid, B., Ratanatawanate, C., Faungnawakij, K., 2014. Conversion of xylose to levulinic acid over modified acid functions of alkaline-treated zeolite Y in hot-compressed water. Chem. Eng. J. 258, 341–347. Chandra, R., Bura, R., Mabee, W., Berlin, A., Pan, X., 2007. Saddler J. Substrate pretreatment: the key to effective enzymatic hydrolysis of lignocellulosics. Adv. Biochem. Eng. Biotechnol. 108, 67–93. Chatterjee, M., Ishizaka, T., Kawanami, H., 2016. Reductive amination of furfural to furfurylamine using aqueous ammonia solution and molecular hydrogen: an environmentally friendly approach. Green Chem. 18, 487–496. Chia, M., Dumesic, J.A., 2011. Liquid-phase catalytic transfer hydrogenation and cyclization of levulinic acid and its esters to γ-valerolactone over metal oxide catalysts. Chem. Commun. (Camb.) 47, 12233–12235. Chia, M., Pagán-Torres, Y.J., Hibbitts, D., Tan, Q., Pham, H.N., Datye, A.K., Neurock, M., Davis, R.J., Dumesic, J.A., 2011. Selective hydrogenolysis of polyols and cyclic ethers over bifunctional surface sites on rhodium-rhenium catalysts. J. Am. Chem. Soc. 133, 12675–12689. Chieffi, G., Braun, M., Esposito, D., 2015. Continuous reductive amination of biomassderived molecules over carbonized filter paper‐supported FeNi alloy. ChemSusChem 8, 3590–3594. Cui, J., Tan, J., Zhu, Y., Chen, F., 2018. Aqueous hydrogenation of levulinic acid to 1,4‐pentanediol over Mo‐modified Ru/activated carbon catalyst. Chemsuschem. 11, 1316–1320. Cukalovic, A., Stevens, C.V., 2010. Production of biobased HMF derivatives by reductive amination. Green Chem. 12, 1201–1206. Sun, D., Sato, S., Ueda, W., Primo, A., Garcia, H., Corma, A., 2016. Production of C4 and C5 alcohols from biomass-derived materials. Green Chem. 18, 2579–2597. Dabrowski, Z., Kwaśny, M., Kamiński, J., Bełdowicz, M., Lewicka, L., Obukowicz, B., Kaliszewski, M., Pirozyńska, E., 2003. The synthesis and applications of 5-aminolevulinic acid (ALA) derivatives in photodynamic therapy and photodiagnosis. Acta Pol. Pharm. 60, 219–224. Daengprasert, W., Boonnoun, P., Laosiripojana, N., Goto, M., Shotipruk, A., 2011. Application of sulfonated carbon-based catalyst for solvothermal conversion of cassava waste to hydroxymethylfurfural and furfural. Ind. Eng. Chem. Res. 50, 327–330. Dai, J.H., Zhu, L.F., Tang, D.Y., Fu, X., Tang, J.Q., Guo, X.W., Hu, C.W., 2017. Sulfonated polyaniline as a solid organocatalyst for dehydration of fructose into 5-hydroxymethylfurfural. Green Chem. 19, 1932–1939. Dai, L., Liu, R., Si, C., 2018. A novel functional lignin-based filler for pyrolysis and feedstock recycling of poly (L-lactide). Green Chem. 20, 1777–1783. Delidovich, I., Leonhard, K., Palkovits, R., 2014. Cellulose and hemicellulose valorisation: an integrated challenge of catalysis and reaction engineering. Synth. Lect. Energy Environ. Technol. Sci. Soc. 7, 2803–2830. Deng, F., Aita, G.M., 2018a. Detoxification of dilute ammonia pretreated energy cane bagasse enzymatic hydrolysate by soluble polyelectrolyte flocculants. Ind. Crops Prod. 112, 681–690. Deng, F., Cheong, D.Y., Aita, G.M., 2018b. Optimization of activated carbon detoxification of dilute ammonia pretreated energy cane bagasse enzymatic hydrolysate by response surface methodology. Ind. Crops Prod. 115, 166–173. Dou, Y., Zhou, S., Oldani, C., Fang, W., Cao, Q., 2018. 5-Hydroxymethylfurfural production from dehydration of fructose catalyzed by solid acid. Fuel 214, 45–54. Dumesic, J.A., Alonso, D.M., Gürbüz, E.I., Wettstein, S.G., 2013. Production of levulinic acid, furfural, and gamma valerolactone from C5 and C6 carbohydrates in mono- and biphasic systems using gamma-valerolactone as a solvent. J. Polym. Sci. Polym. Chem. Edi. 14, 1311–1316. Eggeman, T., Elander, R.T., 2005. Process and economic analysis of pretreatment technologies. Bioresour. Technol. 96, 2019–2025. Elsayed, I., Mashaly, M., Eltaweel, F., Jackson, M.A., Hassan, E.B., 2018. Dehydration of

Acknowledgements This work was financially supported by the State Key Laboratory of Tree Genetics and Breeding (K2017101), Guangxi Key Laboratory of Clean Pulp & Papermaking and Pollution Control, College of Light Industry and Food Engineering, Guangxi University (KF201801-5), China Postdoctoral Science Foundation (2018M641661), State Key Laboratory of Pulp and Paper Engineering (201822), and Foundation of Tianjin Key Laboratory of Pulp & Paper (Tianjin University of Science & Technology), China (201810). References Agarwal, A., Awasthi, S.K., Murthy, P.K., 2011. In vivo antifilarial activity of some cyclic and acylic alcohols. Med. Chem. Res. 430–434. Alam, M.I., De, S., Khan, T.S., Haider, M.A., Saha, B., 2018. Acid functionalized ionic liquid catalyzed transformation of non-food biomass into platform chemical and fuel additive. Ind. Crops Prod. 123, 629–637. Alonso, D.M., Gallo, J.M.R., Mellmer, M.A., Wettstein, S.G., Dumesic, J.A., 2013a. Direct conversion of cellulose to levulinic acid and gamma-valerolactone using solid acid catalysts. Catal. Sci. Technol. 3, 927–931. Alonso, D.M., Wettstein, S.G., Dumesic, J.A., 2013b. Gamma-valerolactone, a sustainable platform molecule derived from lignocellulosic biomass. Green Chem. 15, 584–595. Amarasekara, A.S., Williams, L.D., Ebede, C.C., 2008. Mechanism of the dehydration of dfructose to 5-hydroxymethylfurfural in dimethyl sulfoxide at 150 °C: an NMR study. Carbohyd. Res. 343, 3021–3024. Antal Jr, M.J., Mok, W.S., Richards, G.N., 1990a. Mechanism of formation of 5-(hydroxymethyl)-2-furaldehyde from D-fructose and sucrose. Carbohyd. Res. 199, 91–109. Antal Jr., Michael Jerry, Mok, W.S.L., Richards, G.N., 1990b. Mechanism of formation of 5-(hydroxymethyl)-2-furaldehyde from d-fructose and sucrose. Carbohyd. Res. 199, 91–109. Avolio, R., Bonadies, I., Capitani, D., Errico, M.E., Gentile, G., Avella, M., 2012. A multitechnique approach to assess the effect of ball milling on cellulose. Carbohyd. Polym. 87, 265–273. Aylak, A.R., Akmaz, S., Koc, S.N., 2017. An efficient heterogeneous CrOx-Y zeolite catalyst for glucose to HMF conversion in ionic liquids. Part. Sci. Technol. 35, 490–493. Babaei, Z., Najafi Chermahini, A., Dinari, M., Saraji, M., Shahvar, A., 2018. Cleaner production of 5-hydroxymethylfurfural from fructose using ultrasonic propagation. J. Clean. Prod. 198, 381–388. Bedwell, J., Macrobert, A.J., Phillips, D., Bown, S.G., 1992. Fluorescence distribution and photodynamic effect of ALA-induced PP IX in the DMH rat colonic tumour model. Brit. J. Canser. 65, 818–824. Bhaumik, P., Dhepe, P.L., 2013. Influence of properties of SAPO’s on the one-pot conversion of mono-, di- and poly-saccharides into 5-hydroxymethylfurfural. RSC Adv. 3, 17156–17165. Bhaumik, P., Dhepe, P.L., 2016. Solid acid catalyzed synthesis of furans from carbohydrates. Catal. Rev. 58, 36–112. Biermann, C., Schultz, T., Mcginnia, G., 1984. Rapid steam hydrolysis/extraction of mixed hardwoods as a biomass pretreatment. J. Wood Chem. Technol. 4, 111–128.

195

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

cellulose and solid acid catalyst. Energ. Fuel 28, 5778–5784. Licursi, D., Antonetti, C., Parton, R., Raspolli Galletti, A.M., 2018. A novel approach to biphasic strategy for intensification of the hydrothermal process to give levulinic acid: use of an organic non-solvent. Bioresour. Technol. 264, 180–189. Lin, X., Wu, Z., Zhang, C., Liu, S., Nie, S., 2018. Enzymatic pulping of lignocellulosic biomass. Ind. Crops Prod. 120, 16–24. Liu, Q.Y., Jin, T., Cai, C.L., Ma, L.L., Wang, T.J., 2016. Enhanced sugar alcohol production from cellulose by pretreatment with mixed ball-milling and solid acids. Bioresources 11, 1843–1854. Liu, L., Yang, X., Hou, Q., Zhang, S., Ju, M., 2018. Corn stalk conversion into 5-hydroxymethylfurfural by modified biochar catalysis in a multi-functional solvent. J. Clean. Prod. 187, 380–389. Luterbacher, J.S., Martin Alonso, D., Dumesic, J.A., 2014. Targeted chemical upgrading of lignocellulosic biomass to platform molecules. Green Chem. 16, 4816–4838. MacDonald, S., 1974. Methyl 5-bromolevulinate. Can. J. Chem. 52, 3257–3258. Mamman, A.S., Jongmin, L., Yeongcheol, K., Intaek, H., Nojoong, P., Youngkyu, H., Chang, J.S., Jinsoo, H., 2008. Furfural: hemicellulose/xylose-derived biochemical. Biofuel Bioprod. Bioresour. 2, 438–454. Marianou, A.A., Michailof, C.M., Pineda, A., Iliopoulou, E.F., Triantafyllidis, K.S., Lappas, A.A., 2018. Effect of Lewis and Brønsted acidity on glucose conversion to 5-HMF and lactic acid in aqueous and organic media. Appl. Catal. A-gen. 555, 75–87. Mariscal, R., Mairelestorres, P., Ojeda, M., Sádaba, I., Granados, M.L., 2016. Furfural: a renewable and versatile platform molecule for the synthesis of chemicals and fuels. Synth. Lect. Energy Environ. Technol. Sci. Soc. 9, 1144–1189. Mellmer, M.A., Gallo, J.M.R., Alonso, D.M., Dumesic, J.A., 2015. Selective production of levulinic acid from furfuryl alcohol in THF solvent systems over H-ZSM-5. ACS Catal. 5, 3354–3359. Miura, Y., Hata, M., Yuge, M., Numano, K., Iwakiri, K., 1999. Allergic contact dermatitis from 1, 2, 6-hexanetriol in fluocinonide cream. Contact Derm. 41, 118–119. Mizugaki, T., Togo, K., Maeno, Z., Mitsudome, T., Jitsukawa, K., Kaneda, K., 2016. Onepot transformation of levulinic acid to 2-methyltetrahydrofuran catalyzed by Pt-Mo/ H-β in water. ACS Sustain. Chem. Eng. 4, 682–685. Morone, A., Apte, M., Pandey, R.A., 2015. Levulinic acid production from renewable waste resources: bottlenecks, potential remedies, advancements and applications. Renew. Sust. Energ. Rev. 51, 548–565. Mosier, N., Wyman, C., Dale, B., Elander, R., Lee, Y.Y., Holtzapple, M., Ladisch, M., 2005. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour. Technol. 96, 673–686. Mushrif, S.H., Vasudevan, V., 2015. Insights into the solvation of glucose in water, dimethyl sulfoxide (DMSO), tetrahydrofuran (THF) and N,N-dimethylformamide (DMF) and its possible implications on the conversion of glucose to platform chemicals. RSC Adv. 5, 20756–20763. Nakagawa, Y., Nakazawa, H., Watanabe, H., Tomishige, K., 2012. Total hydrogenation of Furfural over a silica-supported nickel catalyst prepared by the reduction of a nickel nitrate precursor. ChemCatChem 4, 1791–1797. Nemoto, S., Ueno, T., Watthanaphanit, A., Hieda, J., Saito, N., 2017. Crystallinity and surface state of cellulose in wet ball-milling process. J. Appl. Polym. Sci. 134, 7. Nie, S., Zhang, C., Zhang, Q., Zhang, K., Zhang, Y., Tao, P., Wang, S., 2018. Enzymatic and cold alkaline pretreatments of sugarcane bagasse pulp to produce cellulose nanofibrils using a mechanical method. Ind. Crops Prod. 124, 435–441. Nishimura, S., Mizuhori, K., Ebitani, K., 2016. Reductive amination of furfural toward furfurylamine with aqueous ammonia under hydrogen over Ru-supported catalyst. Res. Chem. Intermediat. 42, 1–12. Nolan, M.R., Sun, G., Shanks, B.H., 2014. On the selective acid-catalysed dehydration of 1, 2, 6-hexanetriol. Catal. Sci. Technol. 4, 2260–2266. Nolan, M.R., Bejile, A., Enombo, S.-L., Shanks, B.H., 2016. Directing polyol dehydration via modification of acid catalysts with metals. Top. Catal. 59, 29–36. O’Neill, B.J., Jackson, D.H., Crisci, A.J., Farberow, C.A., Shi, F., Alba-Rubio, A.C., Lu, J., Dietrich, P.J., Gu, X., Marshall, C.L., 2013. Stabilization of copper catalysts for liquidphase reactions by atomic layer deposition. Angew. Chemie 52, 13808–13812. Obregón, I., Gandarias, I., Alshaal, M.G., Mevissen, C., Arias, P.L., Palkovits, R., 2016. The role of the hydrogen source on the selective production of γ-valerolactone and 2methyltetrahydrofuran from levulinic acid. ChemSusChem 9, 2488–2495. Ordomsky, V.V., Van, d.S, J., Schouten, J.C., Nijhuis, T.A., 2012. Effect of solvent addition on fructose dehydration to 5-hydroxymethylfurfural in biphasic system over zeolites. J. Catal. 287, 68–75. Peleteiro, S., Rivas, S., Alonso, J.L., Santos, V., Parajo, J.C., 2015. Utilization of ionic liquids in lignocellulose biorefineries as agents for separation, derivatization, fractionation, or pretreatment. J. Agr. Food. Chem. 63, 8093–8102. Pérez, J., Muñoz-Dorado, J., de la Rubia, T., Martínez, J., 2002. Biodegradation and biological treatments of cellulose, hemicellulose and lignin: an overview. Int. Microbiol. 5, 53–63. Qi, L., Horváth, I.T., 2012. Catalytic conversion of fructose to γ-valerolactone in γ-valerolactone. ACS Catal. 2, 2247–2249. Rackemann, D.W., Doherty, W.O., 2015. The conversion of lignocellulosics to levulinic acid. Biofuel Bioprod. Bior. 5, 198–214. Ratanakhanokchai, K., Waeonukul, R., Pason, P., Tachaapaikoon, C., Kyu, K.L., Sakka, K., Kosugi, A., Mori, Y., 2013. Paenibacillus curdlanolyticus strain B-6 multienzyme complex: a novel system for biomass utilization. Biomass Now. 369–394. Rebeiz, C., Montazer-Zouhoor, A., Hopen, H., Wu, S., 1984. Photodynamic herbicides: 1. Concept and phenomenology. Enzyme. Microb. Tech. 6, 390–396. Ren, H.F., Zhou, Y.G., Liu, L., 2013. Selective conversion of cellulose to levulinic acid via microwave-assisted synthesis in ionic liquids. Bioresour. Technol. 129, 616–619. Ren, H., Girisuta, B., Zhou, Y., Liu, L., 2015. Selective and recyclable depolymerization of cellulose to levulinic acid catalyzed by acidic ionic liquid. Carbohyd. Polym. 117, 569–576. Ria-Millati, I., Syamsiah, S., Niklasson, C., Cahyanto, M.N., Lundquist, K., Taherzadeh, M.J., 2011. Biological pretreatment of lignocelluloses with white-rot fungi and its applications: a review. Bioresources 6, 1–36. Román-Leshkov, Y., Moliner, M., Labinger, J.A., Davis, M.E., 2010. Mechanism of glucose

glucose to 5-hydroxymethylfurfural by a core-shell Fe3O4@SiO2-SO3H magnetic nanoparticle catalyst. Fuel 221, 407–416. Endo, K., Sawada, T., 2001. Control of polymer structure by a chain-transfer reaction in the radical polymerization of acrylamide by β-mercaptopropionic acid and 1, 2, 6hexanetriol trithioglycolate. Collloid. Polym. Sci. 279, 1058–1063. Galkin, K.I., Krivodaeva, E.A., Romashov, L.V., Zalesskiy, S.S., Kachala, V.V., Burykina, J.V., Ananikov, V.P., 2016. Critical influence of 5-hydroxymethylfurfural aging and decomposition on the utility of biomass conversion in organic synthesis. Angew. Chem. Int. Ed. 55, 1–6. Gallo, J.M.R., Alonso, D.M., Mellmer, M.A., Dumesic, J.A., 2013. Production and upgrading of 5-hydroxymethylfurfural using heterogeneous catalysts and biomass-derived solvents. Green Chem. 15, 85–90. Gu, H., An, R., Bao, J., 2018. Pretreatment refining leads to constant particle size distribution of lignocellulose biomass in enzymatic hydrolysis. Chem. Eng. J. 352, 198–205. Gürbüb, E.I., Alonso, D.M., Bond, J.Q., Dumesic, J.A., 2011. Reactive extraction of levulinate esters and conversion to γ-valerolactone for production of liquid fuels. ChemSusChem 4, 357–361. Ha, S.H., Mai, N.L., An, G., Koo, Y.M., 2011. Microwave-assisted pretreatment of cellulose in ionic liquid for accelerated enzymatic hydrolysis. Bioresour. Technol. 102, 1214–1219. Hao, W., Li, W., Tang, X., Lin, L., 2016. Catalytic transfer hydrogenation of biomassderived 5-hydroxymethyl furfural to the building block 2,5-bishydroxymethyl furan. Green Chem. 18, 1080–1088. Hayes, D.J., Fitzpatrick, S., Hayes, M.H., Ross, J.R., 2006. The biofine process-production of levulinic acid, furfural, and formic acid from lignocellulosic feedstocks. Biore. Industr. Proc. Prod. 1, 139–164. Hendriks, A., Zeeman, G., 2009. Pretreatments to enhance the digestibility of lignocellulosic biomass. Bioresour. Technol. 100, 10–18. Hengne, A.M., Rode, C.V., 2012. Cu-ZrO2 nanocomposite catalyst for selective hydrogenation of levulinic acid and its ester to γ-valerolactone. Green Chem. 14, 1064–1072. Horvat, J., Klaić, B., Metelko, B., Šunjić, V., 1985. Mechanism of levulinic acid formation. Tetrahedron Lett. 26, 2111–2114. Hu, X., Jiang, S., Wu, L., Wang, S., Li, C.Z., 2017. One-pot conversion of the biomassderived xylose and furfural into levulinate esters via acid catalysis. Chem. Commun. (Camb.) 53, 2938–2941. Huang, F., Su, Y., Tao, Y., Sun, W., Wang, W., 2018. Preparation of 5-hydroxymethylfurfural from glucose catalyzed by silica-supported phosphotungstic acid heterogeneous catalyst. Fuel 226, 417–422. Huber, G.W., Sara Iborra, A., Corma, A., 2006. Synthesis of transportation fuels from biomass: chemistry, catalysts, and engineering. Chem. Rev. 106, 4044–4098. Hyun, J.H., Sung, K.L., Young, J.H., Jun, W.P., 1994. Selective bromination of ketones. A convenient synthesis of 5-aminolevulinic acid. Synthetic. Commun. 24, 2557–2562. Hyun, D.S., Kim, H.T., Jheon, S.H., Park, S.I., Kim, J.K., 2009. A preliminary study of protoporphyrin-IX as a potential candidate for identification of lung cancer cells using fluorescence microscopy. Photodiagnosis Photodyn. Ther. 6, 221–226. Jeong, H., Park, S.Y., Ryu, G.H., Choi, J.H., Kim, J.H., Choi, W.S., Lee, S.M., Choi, Choi, J.W, I.G., 2018. Catalytic conversion of hemicellulosic sugars derived from biomass to levulinic acid. Catal. Commun. 117, 19–25. https://doi.org/10.1016/j.catcom. 2018.04.016. Jia, S., Xu, Z., Zhang, Z.C., 2014. Catalytic conversion of glucose in dimethylsulfoxide/ water binary mix with chromium trichloride: role of water on the product distribution. Chem. Eng. J. 254, 333–339. Juzeniene, A., Juzenas, P., Kaalhus, O., Iani, V., Moan, J., 2010. Temperature effect on accumulation of protoporphyrin IX after topical application of 5-aminolevulinic acid and its methylester and hexylester derivatives in normal mouse skin. Photochem. Photobiol. 76, 452–456. Kang, M., Kim, S.W., Kim, J.W., Kim, T.H., Kim, J.S., 2013. Optimization of levulinic acid production from Gelidium amansii. Renew. Energy 54, 173–179. Karimi, B., Mohamad-Mirzaei, H., 2016. Sulphanilic acid as a recyclable bifunctional organocatalyst in selective conversion of lignocellulosic biomass to 5-HMF. Green Chem. 18, 2282–2286. Kegnaes, S., Mielby, J., Mentzel, U.V., Jensen, T., Fristrup, P., Riisager, A., 2012. One-pot synthesis of amides by aerobic oxidative coupling of alcohols or aldehydes with amines using supported gold and base as catalysts. Chem. Commun. (Camb.) 48, 2427–2429. Khan, A.S., Man, Z., Bustam, M.A., Kait, C.F., Nasrullah, A., Ullah, Z., Sarwono, A., Ahamd, P., Muhammad, N., 2018. Dicationic ionic liquids as sustainable approach for direct conversion of cellulose to levulinic acid. J. Clean. Prod. 170, 591–600. Kumar, S., Ahluwalia, V., Kundu, P., Sangwan, R.S., Kansal, S.K., Runge, T.M., Elumalai, S., 2018. Improved levulinic acid production from agri-residue biomass in biphasic solvent system through synergistic catalytic effect of acid and products. Bioresour. Technol. 251, 143–150. Kumari, P.K., Rao, B.S., Padmakar, D., Pasha, N., Lingaiah, N., 2018. Lewis acidity induced heteropoly tungustate catalysts for the synthesis of 5-ethoxymethyl furfural from fructose and 5-hydroxymethylfurfural. Mol. Catal. 448, 108–115. Kuster, B., 1990. 5-Hydroxymethylfurfural (HMF). A review focussing on its manufacture. StarchStrke 42, 314–321. Li, J., Jiang, Z., Hu, L., Hu, C., 2015. Selective conversion of cellulose in corncob residue to levulinic acid in an aluminum trichloride-sodium chloride system. ChemSusChem 7, 2482–2488. Li, X., Liu, Q., Luo, C., Gu, X., Lu, L., Lu, X., 2017. Kinetics of furfural production from corn cob in γ-valerolactone using dilute sulfuric acid as catalyst. ACS Sustain. Chem. Eng. 5, 8587–8593. Li, X., Liu, Q., Si, C., Lu, L., Luo, C., Gu, X., Liu, W., Lu, X., 2018. Green and efficient production of furfural from corn cob over H-ZSM-5 using γ-valerolactone as solvent. Ind. Crops Prod. 120, 343–350. Liao, Y., Liu, Q., Wang, T., Long, J., Qi, Z., Ma, L., Yong, L., Li, Y., 2014. Promoting hydrolytic hydrogenation of cellulose to sugar alcohols by mixed ball milling of

196

Industrial Crops & Products 130 (2019) 184–197

X. Li et al.

Coordinated development of leading biomass pretreatment technologies. Bioresour. Technol. 96, 1959–1966. Xiao, Y., Song, Y.F., 2014. Efficient catalytic conversion of the fructose into 5-hydroxymethylfurfural by heteropolyacids in the ionic liquid of 1-butyl-3-methyl imidazolium chloride. Appl. Catal. A-gen. 484, 74–78. Xie, H., Du, H., Yang, X., Si, C., 2018. Recent strategies in preparation of cellulose nanocrystals and cellulose nanofibrils derived from raw cellulose materials. Int. J. Polym. Sci. 5, 1–25. Xu, W., Wang, H., Liu, X., Ren, J., Wang, Y., Lu, G., 2011. Direct catalytic conversion of furfural to 1, 5-pentanediol by hydrogenolysis of the furan ring under mild conditions over Pt/Co2AlO4 catalyst. Chem. Commun. (Camb.) 47, 3924–3926. Xu, Z., Yan, P., Xu, W., Jia, S., Xia, Z., Chung, B., Zhang, Z.C., 2014. Direct reductive amination of 5-hydroxymethylfurfural with primary/secondary amines via Ru-complex catalyzed hydrogenation. RSC Adv. 4, 59083–59087. Yan, L., Liu, N., Wang, Y., Machida, H., Qi, X., 2014. Production of 5-hydroxymethylfurfural from corn stalk catalyzed by corn stalk-derived carbonaceous solid acid catalyst. Bioresour. Technol. 173, 462–466. Yan, K., Jarvis, C., Gu, J., Yan, Y., 2015. Production and catalytic transformation of levulinic acid: a platform for speciality chemicals and fuels. Renew. Sust. Energ. Rev. 51, 986–997. Yang, P., Cui, Q., Zu, Y., Liu, X., Lu, G., Wang, Y., 2015. Catalytic production of 2, 5dimethylfuran from 5-hydroxymethylfurfural over Ni/Co3O4 catalyst. Catal. Commun. 66, 55–59. Yao, S., Wang, X., Jiang, Y., Wu, F., Chen, X., Mu, X., 2013. One-step conversion of biomass-derived 5-hydroxymethylfurfural to 1, 2, 6-hexanetriol over Ni-Co-Al mixed oxide catalysts under mild conditions. ACS Sustain. Chem. Eng. 2, 173–180. Yemiş, O., Mazza, G., 2012. Optimization of furfural and 5-hydroxymethylfurfural production from wheat straw by a microwave-assisted process. Bioresour. Technol. 109, 215–223. Yi, J., He, T., Jiang, Z., Li, J., Hu, C., 2013. AlCl3 catalyzed conversion of hemicellulose in corn stover. Chin. J. Catal. 34, 2146–2152. Yu, I.K.M., Tsang, D.C.W., 2017. Conversion of biomass to hydroxymethylfurfural: a review of catalytic systems and underlying mechanisms. Bioresour. Technol. Rep. 238, 716–732. Yu, I., Dcw, T., Ack, Y., Chen, S.S., Ok, Y.S., Poon, C.S., 2016. Valorization of food waste into hydroxymethylfurfural: dual role of metal ions in successive conversion steps. Bioresour.Technol. 219, 338–347. Yu, I.K.M., Tsang, D.C.W., Yip, A.C.K., Chen, S.S., Wang, L., Ok, Y.S., Poon, C.S., 2017. Catalytic valorization of starch-rich food waste into hydroxymethylfurfural (HMF): controlling relative kinetics for high productivity. Bioresour. Technol. 237, 222–230. Sun, Z., Xue, L., Wang, S., Shi, J., 2016. Single step conversion of cellulose to levulinic acid using temperature-responsive dodeca-aluminotungstic acid catalysts. Green Chem. 18, 742–752. Zenzen, V., Zankl, H., 2003. Protoporphyrin IX-accumulation in human tumor cells following topical ALA-and h-ALA-application in vivo. Cancer Lett. 202, 35–42. Zhang, J., Weitz, E., 2012. An in situ NMR study of the mechanism for the catalytic conversion of fructose to 5-hydroxymethylfurfural and then to levulinic acid using 13C labeled D-fructose. ACS Catal. 2, 1211–1218. Zhang, J., Zhang, B., Zhang, J., Lin, L., Liu, S., Ouyang, P., 2010. Effect of phosphoric acid pretreatment on enzymatic hydrolysis of microcrystalline cellulose. Biotechnol. Adv. 28, 613–619. Zhang, T., Kumar, R., Wyman, C.E., 2013a. Enhanced yields of furfural and other products by simultaneous solvent extraction during thermochemical treatment of cellulosic biomass. RSC Adv. 3, 9809–9819. Zhang, T., Kumar, R., Wyman, C.E., 2013b. Sugar yields from dilute oxalic acid pretreatment of maple wood compared to those with other dilute acids and hot water. Carbohyd. Polym. 92, 334–344. Zhang, Q., Li, S.S., Zhu, M.M., Liu, Y.M., He, H.Y., Cao, Y., 2016. Direct reductive amination of aldehydes with nitroarenes using bio-renewable formic acid as a hydrogen source. Green Chem. 18, 2507–2513. Zhang, L., Xi, G., Chen, Z., Qi, Z., Wang, X., 2017. Enhanced formation of 5-HMF from glucose using a highly selective and stable SAPO-34 catalyst. Chem. Eng. J. 307, 877–883. Zhang, L., Mao, J., Li, S., Yin, J., Sun, X., Guo, X., Song, C., Zhou, J., 2018. Hydrogenation of levulinic acid into gamma-valerolactone over in situ reduced CuAg bimetallic catalyst: strategy and mechanism of preventing Cu leaching. Appl. Catal. B-Environ. 232, 1–10. Zhang, Z., Hu, X., Zhang, S., Liu, Q., Hu, S., Xiang, J., Wang, Y., Lu, Y., 2019. Direct conversion of furan into levulinate esters via acid catalysis. Fuel 237, 263–275. Zhang, Z., Zhao, Z.K., 2010. Microwave-assisted conversion of lignocellulosic biomass into furans in ionic liquid. Bioresour. Technol. 101, 1111–1114. Zhao, J., Zhou, C., He, C., Dai, Y., Jia, X., Yang, Y., 2016. Efficient dehydration of fructose to 5-hydroxymethylfurfural over sulfonated carbon sphere solid acid catalysts. Catal. Today 264, 123–130. Zheng, X., Zhi, Z., Gu, X., Li, X., Zhang, R., Lu, X., 2017. Kinetic study of levulinic acid production from corn stalk at mild temperature using FeCl3 as catalyst. Fuel 187, 261–267. Zuo, M., Le, K., Li, Z., Jiang, Y., Zeng, X.H., Sun, T.X., Lu, Y.L., 2017. Green process for production of 5-hydroxymethylfurfural from carbohydrates with high purity in deep eutectic solvents. Ind. Crops Prod. 99, 1–6. Zuo, M., Le, K., Feng, Y., Li, Z., Zeng, Xh., Tang, X., Sun, Y., Lin, Lu., 2018. An effective pathway for converting carbohydrates to biofuel 5-ethoxymethylfurfural via, 5-hydroxymethylfurfural with deep eutectic solvents (DESs). Ind. Crops Prod. 112, 18–23.

isomerization using a solid Lewis acid catalyst in water. Angew. Chem. 49, 8954–8957. Rosatella, A.A., Simeonov, S.P., Frade, R.F.M., Afonso, C.A.M., 2011. 5Hydroxymethylfurfural (HMF) as a building block platform: biological properties, synthesis and synthetic applications. Green Chem. 13, 754–793. Saritha, M., Arora, A., Lata, 2012. Biological pretreatment of lignocellulosic substrate for enhanced delignification and enzymatic digestibility. Ind. J. Microbiol. 52, 122–130. Schmidt, L.M., Mthembu, L.D., Reddy, Deenadayalu, N, Kaltschmitt, Smirnova Irina, M., 2017. Levulinic acid production integrated into a sugarcane bagasse based biorefinery using thermal-enzymatic pretreatment. Ind. Crops Prod. 99, 172–178. Schwald, W., Breuil, C., Brownell, H.H., Chan, M., Saddler, J.M., 1989. Assessment of pretreatment conditions to obtain fast complete hydrolysis on high substrate concentrations. Appl. Biochem. Biotech. 20-21, 29–44. Seemala, B., Haritos, V., Tanksale, A., 2016. Levulinic acid as a catalyst for the production of 5‐hydroxymethylfurfural and furfural from lignocellulose biomass. Chemcatchem 8, 640–647. Serrano-Ruiz, J.C., West, R.M., Dumesic, J.A., 2010. Catalytic conversion of renewable biomass resources to fuels and chemicals. Annu. Rev. Chem. Biomol. Eng. 1, 79–100. Sharma, H.K., Xu, C., Qin, W., 2017. Biological pretreatment of lignocellulosic biomass for biofuels and bioproducts: an overview. Waste Biomass Valori. 1–17. Shen, J., Wyman, C.E., 2012. Hydrochloric acid-catalyzed levulinic acid formation from cellulose: data and kinetic model to maximize yields. AIChE J. 58, 236–246. Sim, S.E., Kwon, S., Koo, S., 2012. Bis-sulfonic acid ionic liquids for the conversion of fructose to 5-hydroxymethyl-2-furfural. Molecules 17, 12804–12811. Sitthisa, S., Pham, T., Prasomsri, T., Sooknoi, T., Mallinson, R.G., Resasco, D.E., 2011a. Conversion of furfural and 2-methylpentanal on Pd/SiO2 and Pd–Cu/SiO2 catalysts. J. Catal. 280, 17–27. Sitthisa, S., Sooknoi, T., Ma, Y., Balbuena, P.B., Resasco, D.E., 2011b. Kinetics and mechanism of hydrogenation of furfural on Cu/SiO2 catalysts. J. Catal. 277, 1–13. Song, S., Yao, S., Cao, J., Di, L., Wu, G., Guan, N., Li, L., 2017. Heterostructured Ni/NiO composite as a robust catalyst for the hydrogenation of levulinic acid to γ-valerolactone. Appl. Catal. B-Environ. 217, 115–124. Spalt, H., 1977. Chemical changes in wood associated with wood fiberboard manufacture. ACS Symp. Seri. Am. Chem. Soc. 12, 193–219. Starodubtseva, E., Turova, O., Vinogradov, M., Gorshkova, L., Ferapontov, V., 2005. Enantioselective hydrogenation of levulinic acid esters in the presence of the Ru IIBINAP-HCl catalytic system. Russ. Chem. Bull. 54, 2374–2378. Sun, Z., Cheng, M., Li, H., Shi, T., Yuan, M., Wang, X., Jiang, Z., 2012. One-pot depolymerization of cellulose into glucose and levulinic acid by heteropolyacid ionic liquid catalysis. RSC Adv. 2, 9058–9065. Sun, X., Wang, J., Chen, J., Zheng, J., Shao, H., Huang, C., 2018. Dehydration of fructose to 5-hydroxymethylfurfural over MeSAPOs synthesized from bauxite. Microporous Mesoporous Mater. 259, 238–243. Suzuki, T., Nakagami, H., 1999. Effect of crystallinity of microcrystalline cellulose on the compactability and dissolution of tablets. Eur. J. Pharm. Biopharm. 47, 225–230. Swift, T.D., Nguyen, H., Erdman, Z., Kruger, J.S., Nikolakis, V., Vlachos, D.G., 2016. Tandem Lewis acid/Brønsted acid-catalyzed conversion of carbohydrates to 5-hydroxymethylfurfural using zeolite beta. J. Catal. 333, 149–161. Upare, P.P., Lee, J.M., Hwang, Y.K., Hwang, D.W., Lee, J.H., Halligudi, S.B., Hwang, J.S., Chang, J.S., 2011. Direct hydrocyclization of biomass-derived levulinic acid to 2methyltetrahydrofuran over nanocomposite copper/silica catalysts. ChemSusChem 4, 1749–1752. Van Putten, R.J., Jc, V.D.W., De, J.E., Rasrendra, C.B., Heeres, H.J., de Vries, J.G., 2013. Hydroxymethylfurfural, a versatile platform chemical made from renewable resources. Chem. Rev. 113, 1499–1597. Wang, J.J., Tan, Z.C., Zhu, C.C., Miao, G., Kong, L.Z., Sun, Y.H., 2016. One-pot catalytic conversion of microalgae (Chlorococcum sp.) into 5-hydroxymethylfurfural over the commercial H-ZSM-5 zeolite. Green Chem. 18, 452–460. Wang, C., Zhang, Q., Chen, Y., Zhang, X., Xu, F., 2018a. Highly efficient conversion of xylose residues to levulinic acid over FeCl3 catalyst in green salt solutions. ACS Sustain. Chem. Eng. 6, 3154–3161. Wang, G., Xia, Y., Sui, W., Si, C., 2018b. Lignin as a novel tyrosinase inhibitor: effects of sources and isolation processes. ACS Sustain. Chem. Eng. 6, 9510–9518. Wataniyakul, P., Boonnoun, P., Quitain, A.T., Kida, T., Laosiripojana, N., Shotipruk, A., 2018. Preparation of hydrothermal carbon acid catalyst from defatted rice bran. Ind. Crops Prod. 117, 286–294. Weingarten, R., Conner, W.C., Huber, G.W., 2012. Production of levulinic acid from cellulose by hydrothermal decomposition combined with aqueous phase dehydration with a solid acid catalyst. Synth. Lect. Energy Environ. Technol. Sci. Soc. 5, 7559–7574. Werpy, T.A., Holladay, J.E., White, J.F., 2004. Top value added chemicals from biomass: I. Results of screening for potential candidates from sugars and synthesis gas. US Depar. Energ. 45–49. Wettstein, S.G., Alonso, D.M., Chong, Y., Dumesic, J.A., 2012. Production of levulinic acid and gamma-valerolactone (GVL) from cellulose using GVL as a solvent in biphasic systems. Synth. Lect. Energy Environ. Technol. Sci. Soc. 5, 8199–8203. Widsten, P., Murton, K., West, M., 2018. Production of 5-hydroxymethylfurfural and furfural from a mixed saccharide feedstock in biphasic solvent systems. Ind. Crops Prod. 119, 237–242. Wolfrom, M., Schuetz, R.D., Cavalieri, L.F., 1948. Chemical interactions of amino compounds and sugars. III. 1 the conversion of D-glucose to 5-(hydroxymethyl)-2-furaldehyde. J. Am. Chem. Sci. 70, 514–517. Wyman, C.E., Dale, B.E., Elander, R.T., Holtzapple, M., Ladisch, M.R., Lee, Y.Y., 2005.

197