Skip to content
Publicly Available Published by De Gruyter March 31, 2021

Mechanistic studies on the formation of 5-hydroxymethylfurfural from the sugars fructose and glucose

  • Yi Liu and Francesca M. Kerton ORCID logo EMAIL logo

Abstract

In recent years the transformations of fructose and glucose to the platform chemical 5-hydroxymethylfurfural (5-HMF) have been studied extensively, and a variety of mechanisms have been proposed. This review summarizes the varied mechanisms proposed and methods used to study the dehydration of biomass, such as fructose and glucose, to give 5-hydroxymethylfurfural. For fructose dehydration, two main mechanisms have been suggested including a cyclic and an acyclic pathway, of which the cyclic pathway dominates. The conversion of glucose to 5-HMF can proceed either through initial isomerization to fructose or a direct dehydration. For glucose to fructose isomerization, two main reaction pathways have been proposed (1,2-hydride shift and enolization). This review discusses the mechanisms that have been determined based on the evidence from experiments and/or calculations, and briefly introduces the techniques frequently used in such mechanistic studies. Mechanisms in this field are strongly dependent on the nature of the solvent and the catalyst used, so it is important that researchers have a general idea about the existing mechanisms, and the methods and techniques used for investigation, before pursuing their own mechanistic studies.

Introduction

With rapid industrial development in response to dramatic increases in human population, the world is currently experiencing a great demand for energy. Large emissions of carbon dioxide and other flue-gases (nitrogen oxides and sulfur dioxide) from fossil fuel combustion have caused environmental problems such as climate change and acid rain. The use of biomass as an alternative resource for both fuels and chemicals has been regarded as a promising and effective solution to the current situation [1], [2], [3], [4], [5], [6], [7], [8]. Some chemicals produced from biomass can play the role of platforms for the synthesis of a series of other compounds. 5-hydroxymethylfurfural (5-HMF) is a popular platform chemical which can be obtained from lignocellulosic biomass transformations [9].

Dehydration of carbohydrates to yield 5-HMF have mainly focused on the conversion of fructose, glucose and cellulose [10], [11], [12], [13], [14], [15], [16]. With both hydroxyl and aldehyde functional groups, 5-HMF can be an intermediate for a variety of value-added products and potential biofuels [17], [18], [19], [20]. For example, 2,5-furandicarboxylic acid (FDCA) can be obtained from the oxidation of 5-HMF [21], [22], [23]. FDCA has been listed among the 12 top platform chemicals from biomass by the US Department of Energy [2]. It can be used to produce polymers or pharmaceutical and photographic chemicals [24]. The hydrogenation of 5-HMF can produce 2,5-dimethylfuran (2,5-DMF), which is a potential transport fuel because of its good performance in ignition, emission and combustion compared with traditional gasoline [25].

Because of its valuable applications, numerous researches have been performed on the effective synthesis of 5-HMF from sugars, among these many involved mechanistic studies. In this present review, we summarize the mechanisms followed in conversion of fructose and glucose into 5-HMF. These have been proposed based on the basis of lab experiments or computational calculations, and we briefly introduce these techniques have been applied.

Techniques frequently used in mechanistic studies

NMR spectroscopy

This is a tool frequently used in mechanistic studies. Besides the commonly studied 1H and 13C NMR spectroscopies, NMR spectra of several other elements can also be obtained and used, such as 17O and 11B NMR spectroscopy. Isotopic labeling techniques are often used to trace locations of hydrogen and carbon atoms in compounds and intermediates.

Computational calculations

In recent years, computational chemistry has become increasingly important for investigating reaction mechanisms [26]. The advantages of computational calculations include their convenience to operate and that they can be performed at the same time as lab work. More and more researchers prefer to apply both experimental and computational work in their studies. The theoretical data can provide predictions and explanations for the experimental results, while the lab results provide factual support to the calculations. In the area of biomass transformation, computational studies are normally used to calculate the activation energies of every single step of a putative reaction pathway, analyze their endo- or exo-thermic properties and investigate the stabilities of proposed intermediates.

Besides these two main tools, other methods and tools such as kinetic studies, IR and UV spectroscopies, and mass spectrometry can also be used. In many cases, different techniques have been combined to ensure that the mechanisms proposed are more convincing.

Fructose

So far two main reaction pathways have been proposed for the dehydration of fructose to 5-HMF – a cyclic and an acyclic (open-chain) route [9, 2729]. Both mechanisms have been discussed in many studies since they were raised. However, evidence from lab experiments and computational calculations support the cyclic route over the acyclic one, so the discussion presented below will focus on the cyclic pathway.

Table 1 lists the 1H and 13C NMR data from the literature of the main intermediates and 5-HMF. The 1H and 13C NMR spectra of fructose in DMSO-d6 show that at room temperature the five isomers have chemical shifts ranging from 3.0–4.5 ppm and 50–220 ppm [34]. The NMR analysis of [13C-2] fructose further reveals that the cyclic isomers have resonances of their C2 atoms between 95 and 110 ppm and the open-chain fructose has its carbonyl resonance at 215 ppm [].

Table 1:

1H and 13C NMR data of the intermediates in Scheme 1 and 5-HMF.

Inter-mediateSolventδH (ppm)δC (ppm)Refs.
2DMSO-d65.57 (d, H1), 4.11 (dd, H3), 3.83 (H5), 3.74 (H4), 3.55 (H6), 3.44 (H6′)139.1 (C2), 118.9 (C1), 86.4 (C5), 76.3 (C4), 75.2 (C3), 61.5 (C6)[30]
3DMSO-d69.47 (s, H1), 6.25 (d, H3), 4.81 (dd, H4), 4.26 (td, H5), 3.46 (m, H6)184.3 (C1), 156.4 (C2), 122.3 (C3), 73.6 (C4), 90.3 (C5), 61.3 (C6)[30]
DMSO-d69.46 (s), 6.23 (d, H3)184.9 (C1), 156.9, 122.8[31]
DMSO-d6No proton data reported186 (C1), 157 (C2), 129 (C3), 85 (C4), 90 (C5), 62 (C6)[32]
DMSO-d69.47 (s, H1), 6.25 (d, H3), 4.81 (dd, H4), 4.26 (td, H5), 3.46 (m, H6)184.0 (C1), 156.4 (C2), 122.5 (C3), 90.1 (C5), 73.4 (C4), 61.1 (C6)[33]
5DMSO-d64.38 (dd, H5), 3.68 (dt, H3), 3.59 (m, H1), 3.53 (d, H6), 3.51 (d, H4), 3.41 (dd, H6′)57.6 (C1), 108.3 (C2), 82.6 (C5), 81.2 (C3), 78.3 (C4), 65.9 (C6)[30]
7DMSO-d6No proton data reported108.0 (C2), 81.0, 82.4, 78.1, 65.6 (C6), 57.4 (C1)[33]
8DMSO-d6No proton data reported109 (C2), 85 (C5), 78 (C3 & C4), 66 (C6), 63 (C1)[32]
5-HMFDMSO-d69.52 (s, H1), 7.48 (d, H3), 6.59 (dt, H4), 4.49 (s, H6)177.9 (C1), 162.0 (C5), 151.6 (C2), 124.5 (C3), 109.9 (C4), 56.0 (C6)[30]
DMSO-d66.57 (d), 7.46 (d), 9.51 (s)178.7, 162.8, 152.4, 125.0, 110.4, 58.6[31]
DMSO-d69.49 (s), 7.40 (d), 6.50 (d), 4.47 (s)178.3 (C1), 162.0, 151.8 (C2), 124.6, 110.0, 56.5 (C6)[33], [34]
D2ONo proton data reported183.2 (C1), 164.0, 154.5, 113.7, 58.2 (C6)[34]

To the best of our knowledge, the earliest research trying to determine the reaction mechanism for fructose dehydration through experiments was performed by Antal et al. [35]. Kinetic studies were performed for the reaction in water with and without acid catalysts. The detection of a very small amount of glucose and no mannose in the reaction mixture indicated that the Lobry de Bruyn–van Ekenstein transformation between fructose and glucose was very slow. Therefore, the enolization of fructose, a key step in this transformation, must be slow, and the enols (significant intermediates in the acyclic mechanism) were present in negligible amounts and had little influence in the conversion of fructose. Furthermore, if the 3-deoxyglycosulose intermediate in the acyclic route was generated during the reaction process, it should have undergone a benzylic acid rearrangement under the reaction conditions to produce glucometasaccharinic acids or lactones, which were not found in gas chromatography-mass spectrometry (GC-MS) analysis. Finally, the dehydration of sucrose was performed under similar conditions. The first step of this process is the hydrolysis of sucrose into a fructofuranosyl cation and glucose. The yield of 5-HMF from the fructose part (the fructofuranosyl cation) in the dehydration of sucrose was higher than from pure fructose, either with or without acid catalysts. This result was in accordance with the hypothesis that the fructofuranosyl cation was an important intermediate in the cyclic pathway of fructose dehydration to yield 5-HMF. This work is a milestone in the area of sugar transformations, and has been widely cited by others.

Scheme 1: Reaction pathways of fructose dehydration to 5-HMF proposed by Horváth [30] (red), Amarasekara [31] (black), Kimura [32] (blue) and Zhang [33], [34] (green).
Scheme 1:

Reaction pathways of fructose dehydration to 5-HMF proposed by Horváth [30] (red), Amarasekara [31] (black), Kimura [32] (blue) and Zhang [33], [34] (green).

Amarasekara et al. performed the dehydration of fructose at 150 °C in DMSO-d6 [31]. 1H and 13C NMR data showed that after 6 min an intermediate 3 (Scheme 1) appeared. DMSO was proposed to act as a Lewis base and interact with fructose to form intermediates 1 and 2. However, no NMR evidence was given to support their existence possibly because of their fast conversions to 3 during the reaction.

The Horváth group mapped the dehydration of fructose catalyzed by sulfuric acid (H2SO4) in different solvents through isotopic labeling experiments [30]. [13C-1] to [13C-6] fructose were respectively heated at 150 °C in DMSO-d6, and 13C NMR spectra were recorded to identify the intermediates 2, 3, and 5 (Scheme 1). In addition, deuterium oxide (D2O) was added to the starting materials and no deuterium incorporation into 5-HMF was seen. This indicates that the three steps from 4 towards 5-HMF were irreversible (otherwise 5-HMF-D1 would be produced). 5-HMF formation catalyzed by H2SO4 is not via fructoketose i.e. the acyclic pathway is not followed (otherwise 5-HMF-D3 would be produced). This study elegantly applied isotopic labeling techniques to analyze the different reaction pathways from the fructose isomers and the structures of possible intermediates. It is significant as it clearly showed for the first time that the often proposed fructoketose pathway is not followed.

In 2012 and 2016, Zhang et al. performed in situ 13C and 1H NMR studies of fructose conversion using different catalysts in both DMSO-d6 and water (10 wt% D2O added) [33], [34]. Two intermediates (3 and 7, Scheme 1) were identified in the 13C NMR spectra recorded during the reaction process, and their structures were determined through the analysis of distortionless enhancement by polarization transfer (DEPT) 135 13C NMR spectra, 13C NMR spectra of different 13C labeled fructose and high-resolution electrospray ionization mass spectra (HR ESI-MS). Besides 1H and 13C, 17O NMR analysis was also performed. During the reaction in DMSO-d6 containing H217O, the signal at 12.4 ppm from the naturally occurring 17O in DMSO-d6 was found to increase, indicating that 17O from H217O was incorporated into DMSO-d6. A control experiment of H217O (0.5 %) in DMSO-d6 under the same condition without fructose addition showed no change in the peak intensity at 12.4 ppm. Hence DMSO was believed to participate in the reaction process and consequently caused the incorporation of 17O atom. The Brønsted acid-catalyzed dehydration of fructose was simulated computationally. The interaction between DMSO and the protonated fructose was quite exergonic (−71.5 kJ/mol) and led to the formation of 7, which lost the DMSO molecule and one water molecule to produce 5, the intermediate found in the study by Horváth’s group. The process from 2 to 3 was exergonic (−70.3 kJ/mol), supporting the detection of the stable species 3 during experimental work. The geometry of 7 with one water molecule added was also optimized (7·H2O17adduct). This water molecule could provide hydrogen bonding and strengthen the bond between C2 from fructose and O from DMSO. This calculation supported the NMR observed incorporation of 17O atom into DMSO-d6. This study is interesting as it further confirms the involvement of DMSO in the reaction mechanism, and suggests that water can also influence the reaction process. It needs to be noted that in the mechanism proposed, the intermediate during the conversion of 7 to 3 has two possible forms, enol (2) or aldehyde (2′), but this study was not able to detect either one experimentally, probably because of its low concentration under the reaction conditions applied.

In the study by Kimura et al. [32], a new intermediate (8) in the dehydration of fructose in DMSO-d6 was proposed, but the participation of DMSO was not seen in the reaction mechanism. Two 13C NMR signals appeared at 109 and 157 ppm during the reaction and were assigned to two intermediates (3 and 8), and their structures were identified by using a range of 13C-labeled fructose for reactions and analyzing the corresponding 13C NMR spectra obtained. This study claimed to be the first one that discovered the existence of 8 in fructose dehydration. 8 is different from the DMSO-substituted (1) or the enol intermediate (2) in the mechanism previously proposed by Amarasekara et al. [31]. Therefore, a different reaction pathway was proposed (Scheme 1). However, it should be noted that 8 was possibly formed before or after the generation of other intermediates (1 or 2) and may be present in low (undetectable) concentrations in reactions studied by others.

In order to investigate the function of DMSO as a prominent solvent in fructose dehydration, the Vlachos group performed MD simulations of fructose and 5-HMF under ambient conditions [36]. Water or a water-DMSO mixture was used as the solvent system and the calculation results were compared to study the solvation effect of DMSO. According to the calculated radial distribution functions (RDF) between solute-solvent atom pairs and the volumetric maps of the local three-dimensional structural arrangement of the solvent around the solute, DMSO was more strongly coordinated with fructose than water. The coordination of fructose carbon atoms was related to the hydrogen bonding between fructose and water or DMSO. As both a hydrogen bond donor and acceptor, water molecules tended to stay around the oxygen atoms of the hydroxyl groups of fructose to lower the threshold of acid-catalyzed dehydration of fructose (the proton transfer from a hydronium ion to a hydroxyl group on C2, C3 or C4 of fructose). By contrast, DMSO is only a hydrogen bond acceptor, so it competed with water to form hydrogen bonds with the hydrogen atoms of fructose’s hydroxyl groups, especially around C2, C3 and C4. This helped to eliminate side reactions during fructose dehydration such as polymerization, by preventing the hydrogen atoms from removing the hydroxyl groups and forming glycosidic linkages with other fructose molecules. DMSO also had strong coordination with 5-HMF carbon atoms, especially with C1, inhibiting the cleavage of C1-C2 bond of 5-HMF that leads to the hydration of 5-HMF to formic and levulinic acid (LA). Therefore, this work provided evidence to support the use of hydrogen-bond acceptor solvents (especially DMSO) in the transformation of fructose.

During 2011 and 2012, four groups reported their computational studies of Brønsted acid-catalyzed dehydration of fructose in water [37], [38], [39], [40], [41]. The way to simulate the reaction proceeding via acid catalysis was to develop a protonated fructose model as the reactant. In all these studies, the comparison of all possible protonation sites indicated that protonation at the tertiary hydroxyl group (O2H) was the thermodynamically optimum one because (1) this model had the lowest gas-phase Gibbs energy; (2) the removal of the first water molecule had the lowest Gibbs energy of activation; (3) the reaction pathway would have led to side reactions if the protonation happened on other hydroxyl groups.

In the simulation of fructose dehydration in water at 90 °C performed by Caratzoulas and Vlachos [37], the participation of one water molecule was found to promote the proton transfer from O1 to O3 (11→12, Scheme 2) through decreasing the energy barrier. The C3 carbonium intermediate (13) obtained after the second dehydration was not stable and a hydride transfer from C4 to C3 immediately happened (14). The use of water as the solvent caused a higher energy barrier in hydride transfer via the reorganization of the polar solvent environment and the consequent solvation of the asymmetrical charge distribution. Later, they reported a microkinetic model for the acid-catalyzed dehydration of fructose and the accuracy of molecular simulations was tested [38]. The hydride transfer from C1 to C2 (4′11) was found to be the rate-limiting step. The dehydration of fructose deuterated at C1 (fructose-D1) was performed and kinetic isotopic effects (KIEs) were observed in fructose disappearance (kH/kD = 1.20) and 5-HMF appearance (kH/kD = 2.20), confirming that the hydride transfer from C1 to C2 was indeed rate-limiting [42]. The results were claimed to be in good agreement with experimental results [43].

Scheme 2: The reaction pathways of the dehydration of fructose to 5-HMF proposed by Caratzoulas and Vlachos and coworkers [37], [38] (black), Assary et al. [39], [40] (red: 2011; green: 2012) and the Pidko group [41] (blue).
Scheme 2:

The reaction pathways of the dehydration of fructose to 5-HMF proposed by Caratzoulas and Vlachos and coworkers [37], [38] (black), Assary et al. [39], [40] (red: 2011; green: 2012) and the Pidko group [41] (blue).

Assary et al. performed their calculations on fructose conversion in both neutral water and an acidic environment [39]. The high activation barriers in neutral water indicated that the dehydration of fructose at room temperature was not favored. In the acid-catalyzed simulation, a reaction route similar to Caratzoulas’ was given. The highest enthalpy of activation was 162.3 kJ/mol from the second water molecule loss (1213). This result claimed to be in agreement with reported experimental results using mineral or Lewis acids as catalysts, which were in a range of 129.7–142.2 kJ/mol [44], [45], [46], [47]. It was concluded that protonated intermediates in acid-catalyzed reaction systems would promote the dehydration of fructose. In the following year, the same group performed another computational study on the acid-catalyzed dehydration of fructose and glucose in water [40]. During the transformation of fructose to 5-HMF, three possible paths were proposed for the conversion of 4′ to 2 (Scheme 2). Compared with the generation of 19 through a hydride shift, the deprotonation of 4′ to form 2 directly or via 5 as an intermediate was more thermodynamically favored. It was believed that water promoted the conversion of 2 to 3 by acting as a proton mediator to form an eight-membered transition state (20). The direct generation of 2 from 4′ deprotonation was supported in the calculations performed by Arifin et al. [48].

In a study by the Pidko group [41], an astonishing 24 possible routes for fructose conversion were calculated, among which nine involved the formation of 5-HMF before the final product LA was obtained. On the most thermodynamically favored route, the dehydration of fructose to 5-HMF was −156 kJ/mol exergonic with 3 modeled as an intermediate (Scheme 2). According to the authors, the formation of a conjugated π system within the furan ring and the entropy increase during the dehydration were the main forces driving the conversion of fructose towards LA via 5-HMF formation.

In 2015, the Zhang group performed a DFT study of glucose dehydration in 1-butyl-3-methylimidazolium chloride (BmimCl) catalyzed by a Brønsted acid-functionalized ionic liquid ([BmimSO3H]Cl) [49]. Two mechanisms were proposed, and path I involved the dehydration of fructose to 5-HMF after the isomerization of glucose to fructose. The removal of the first and third water molecules was triggered through the interaction between the hydroxyl groups and the proton from the sulfonic acid group (-SO3H). The loss of the second water molecule was achieved through a ten-membered transition state, in which -SO3H acted as a proton shuttle to accept the proton at O1H and transfer its own proton to O3. The role of -SO3H here is similar to the water molecule in the conversion of 2 to 3 in the study by Assary (Scheme 2). In the study by Li et al. of fructose dehydration in 1-butyl-3-methylimidazolium bromide (BmimBr), a mechanism was proposed in which the removal of all three water molecules was triggered by the hydrogen bonding between fructose and Br, as indicated by FT-IR and 1H and 13C NMR spectra [50].

Most of the studies discussed above are fructose transformations catalyzed by a Brønsted acid. Guan et al. performed DFT studies of glucose and fructose dehydration catalyzed by trivalent metal chlorides in BmimCl [51]. The metal chlorides were proposed to form four-coordinate complexes after their dissolution in the ionic liquid, with three chloride and one imidazolium ligand, and catalyze the isomerization of glucose and the subsequent dehydration of fructose through the interaction between the hydroxyl groups and the metal centers.

As can be seen from the discussion above, 2, 3 and 4 are significant intermediates in cyclic mechanisms of fructose dehydration. These intermediates are therefore key species for others working in this field to detect and measure in order to understand their reaction pathways, and they may also act as clues towards possible pathways in transformations of other saccharide molecules.

Glucose

There have been two mechanisms generally proposed for the conversion of glucose to 5-HMF. As shown in Scheme 3, route A is the direct dehydration of glucose [52], [53]; while in route B the isomerization of glucose to fructose is a crucial step, and 5-HMF is obtained via dehydration of the generated fructose. Among the numerous studies on the transformation of glucose, most of them have focused on route B. Two mechanisms have been proposed for the isomerization of glucose to fructose (B1 and B2, Scheme 3). B1 is through an enolization and B2 is through a 1,2-hydride shift. Both mechanisms have been supported experimentally and computationally in various studies with a range of catalysts.

Scheme 3: Proposed mechanisms of glucose conversion to 5-HMF.
Scheme 3:

Proposed mechanisms of glucose conversion to 5-HMF.

Dehydration of glucose through its isomerization to fructose

In 2011 Matubayasi and co-workers [54] managed to detect all the straight chain and ring isomers of both glucose and fructose through in situ 13C NMR analysis of unlabeled and [13C-1] glucose in D2O, indicating that during the isomerization all isomeric forms of glucose and fructose were formed and therefore may play important roles in the mechanism for 5-HMF formation. The interconversion of glucose and fructose is realized through a Lobry de Bruyn–van Ekenstein transformation [55], [56], [57]. NMR data for said species are provided in Table 2.

Table 2:

1H and 13C NMR data of glucose, fructose, the intermediates in glucose conversion, and 5-HMF.

CompoundSolventδH (ppm)aδC (ppm)Refs.
GlucoseD2OOpen-chain: 205.0 (C1); α-glucopyranose: 92.5 (C1); β-glucopyranose: 97.4 (C1); α-glucofuranose: 97.7 (C1); β-glucofuranose: 102.8 (C1)[54]
DMSO-d6α-glucopyranose: 97.35 (C1), 75.25 (C2); β-glucopyranose: 92.62 (C1), 72.80 (C2); α-glucofuranose: 104.37 (C1), 80.35 (C2); β-glucofuranose: 102.73 (C1), 72.21 (C2)[58]
Solid-stateOpen-chain: 205 (C1)[59]
FructoseD2OOpen-chain: 213.1 (C1); α-fructopyranose: 66.2 (C1); β-fructopyranose: 65.6 (C1); α-fructofuranose: 64.4 (C1); β-fructofuranose: 64.8 (C1)[54]
DMSO-d6Open-chain: 71.17 (C1), 205.54 (C2)[58]
DMSO-d6103.6[60]
Solid-stateOpen-chain: 214 (C1)[59]
3-DGSolid-stateNot studied13C CP/MAS NMR: 110 (C2, absorbed on TiO2); 118 (C2, absorbed on phosphate/TiO2)[61]
22 (MClx = SnCl4)DMSO-d64.90 (s)Not reported[62]
22 (MClx = GeCl4)DMSO-d6Not reported76.4[60]
26DMSO-d65.33 (H1), 3.85 (H2), 3.77 (H3), 3.51 (H4), 3.49 (H5), 2.92 & 2.81 (H6)101.7 (C1), 83.6 (C2), 73.7 (C3), 70.6 (C4), 73.6 (C5)61.3 (C6)[63]
36DMSO-d6Not reportedTwo isomers: 72.58 & 74.15 (C1), 189.88 & 192.40 (C2)[58]
2DMSO-d6Not reported147.60 (C1), 136.30 (C2)[58]
5-HMFD2ONot reported180.9 (C1)[54]
DMSO-d6Not reported178.44 (C1), 152.16 (C2)[58]
DMSO-d69.49 (s, H1)Not reported[64]
DMSO-d6Not reported152.1 (C2), 124.3 (C3), 110.9 (C4), 56.1 (C6)[60]
  1. a1H NMR data not included for glucose and fructose due to presence of multiple isomers.

The study by Zhao et al. is a milestone because for the first time it revealed the interaction of metal halide catalysts with glucose [65]. Mutarotation of glucopyranose was observed in 1H and 13C NMR spectra after the addition of the catalysts. The competition studies using glycerol and glyceraldehyde indicated that it was the hemiacetal portion of glucopyranose instead of the polyalcohol portion that interacted with the catalyst. It was proposed that the metal center interacted with the hydroxyl groups to cause the mutarotation and the subsequent isomerization of glucose, but no experimental work was performed to verify this. A similar study was performed by Zhang et al. using hydroxyapatite supported chromium chloride (Cr-HAP) as the catalyst [66]. The 13C NMR analysis provided evidence for glucose mutarotation, and traces of fructose were detected, which came from the isomerization of glucose.

After Zhao’s work, several researchers used similar methods (NMR analysis and competition studies) to explore the mechanisms of glucose dehydration [60, 62, 64]. Different Lewis acids were used as catalysts. In the study by Han and co-workers using tin(IV) chloride (SnCl4) [62], broadening of the resonances of the hydroxyl groups of glucopyranose was observed in 1H NMR spectra after SnCl4 addition, due to the interaction of Cl ions and the protons of the hydroxyl groups (Scheme 3, 21, MClx = SnCl4). The glucose conversion and 5-HMF yield decreased drastically with the addition of ethylene glycol, indicating that a five-membered ring chelate complex was formed during the reaction. This intermediate was observed in 1H NMR spectra, coming from the interaction of the Sn center with the 1,2-dihydroxy group of glucose (22). 5-HMF was proposed to be generated either directly from an enediol intermediate (24) or through the fructose produced from 24. In a study by De et al., a 1H NMR study was carried out on glucose dehydration catalyzed by aluminum(III) chloride (AlCl3) [64]. A similar broadening of hydroxyl peaks caused by hydrogen bonding was observed. However, a 1,2-hydride shift mechanism was proposed for the subsequent isomerization of glucose, which was different from Han’s work. In a study by Zhang et al., [13C-2] glucose was used as the reactant with germanium(IV) chloride (GeCl4) as the catalyst [60]. The peak at 76.4 ppm in 13C NMR spectra during the reaction was assigned to 22 (Scheme 3, MClx = GeCl4). The addition of phenylfluorone decreased the glucose conversion and 5-HMF yield significantly because of its competitive complexation with GeCl4. An enolization pathway similar to Han’s was proposed for the subsequent conversion of 22 to fructose.

Compared with Zhao’s work, these studies took further steps to identify the intermediates of metal complexation in the isomerization process. However, all the studies discussed above were focused on confirming the complexation of the Lewis acid catalyst with glucose. None of them provided experimental evidence to support the proposed enolization or 1,2-hydride shift mechanisms of glucose isomerization. The isomerization pathways can be studied via 1H and 13C NMR spectroscopy using glucose deuterated at C2 (glucose-D2). If a 1,2-hydride shift occurs, the deuterium at the C2 will transfer to C1 of the fructose or 5-HMF obtained. Theoretically, the fructose should contain 100 % deuterium at C1 and the 5-HMF should be 50 % deuterated at C1. In 13C NMR spectra, because of disruption of the nuclear overhauser enhancement (NOE) caused by the deuterium atoms (2H), 13C resonance intensities of 13C-1H pairs will be doubly enhanced while the resonance intensities of 13C-2H pairs (appear as triplets) will remain unchanged [67]. If enolization happens, the deuterium will be expelled into the solvent so in theory no deuterium will be in the produced fructose and 5-HMF. However, it should be noted that a small amount of deuterium incorporation might be seen because of the proton/deuterium exchange. Conversely, if a deuterated solvent, such as D2O, is used, 50 % deuterium incorporation at C1 of fructose or 5-HMF is expected for an enolization pathway but not for a 1,2-hydride shift mechanism.

Raines and co-workers performed a mechanistic study on glucose dehydration through 1H and 2H NMR analysis [68]. Chromium(III) chloride (CrCl3) or CrCl2 was used as the catalyst. When one equivalent D2O was added in the starting materials, less than 5 % deuterium incorporation was detected at C1 of the 5-HMF produced. When glucose-D2 was used, approximately 33 % deuterium was incorporated at C1 of 5-HMF, and a signal of the aldehydic deuteron was observed in the 2H NMR spectrum. Therefore, the isomerization of glucose occurred via a 1,2-hydride shift. Similarly, in the study by Davis’ group of glucose-D2 isomerization using a tin-containing zeolite (Sn-Beta) [69], the 1H and 13C NMR spectra of the fructose product indicated that the deuterium at C2 of glucose-D2 transferred to C1 of the fructose. A KIE (kH/kD = 1.98) was observed in the initial reaction rate, further revealing that the 1,2-hydride shift was the rate-limiting step. In the study by Bermejo-Deval et al. [59], a DFT calculation was performed on the Sn-Beta-catalyzed isomerization of glucose through a 1,2-hydride shift. The isomerization catalyzed by a Sn-Beta open-site with one adjacent silanol group had an enthalpy of activation of 92.5 kJ/mol, which was in good agreement with the experimental value (88.7 ± 2.9 kJ/mol). A similar 1,2-hydride shift mechanism was proposed by Choudhary et al. [70] for the isomerization of glucose-D2 using homogeneous (CrCl3, AlCl3) and heterogeneous (Sn-Beta) Lewis acid catalysts, after the observation of KIEs and the analysis of 1H, 2H and 13C NMR spectra.

AlCl3-catalyzed dehydration of glucose was analyzed through a combination of ESI-MS/MS, ATR-IR and NMR spectroscopies by Tang et al. [71]. The ESI-MS/MS spectra indicated that glucose mainly coordinated with [Al(OH)2(H2O)4]+ in solution. The linear relationship between log k and log [Al] indicated that glucose isomerization was a first-order reaction with respect to the Al species, so the Al complexes involved in the rate-limiting step should be mononuclear. The addition of glyceraldehyde and 1,3-dihydroxyacetone significantly decreased the glucose conversion and fructose yield while the addition of glycerol made no difference. Therefore, the interaction between the carbonyl group of open-chain glucose and the Al complexes was crucial to initialize the isomerization. Both KIE and NOE effects were observed in the dehydration of glucose-D2, thus a 1,2-hydride shift mechanism was proposed. This conclusion was supported by the study by Shanks and co-workers [72]. When AlCl3 was used, a significant KIE was observed (kH/kD = 1.34) in glucose-D2 conversion.

Hensen’s group performed a series of studies of glucose complexation with chromium chlorides (CrCl2 or CrCl3) [73], [74], [75]. In their work about CrCl2 catalysis [73], extended X-ray absorption fine structure (EXAFS) spectra showed coordination of glucose to Cr2+ and the formation of a Cl- or O-bridged Cr dimer during the reaction. A binuclear route with two Cr2+ centers was proposed for the 1,2-hydride shift process after DFT calculations of Gibbs free energies of intermediates. Similar studies were performed using CrCl3·6H2O as the catalyst [75]. During the initial ring opening of glucose, the mononuclear path was slightly more favored than the binuclear path. Nevertheless, during the following deprotonation of O2H and the 1,2-hydride shift the coordination with two Cr centers effectively decreased the energy barrier. Therefore, similar to Cr2+, the binuclear Cr3+ complex was preferred for glucose isomerization via a 1,2-hydride shift. However, in the study by Guan et al. [51] of glucose isomerization catalyzed by trivalent metal chlorides (CrCl3, WCl3, MoCl3 and FeCl3), a mechanism through enolization was proposed after DFT calculations, which is different from Hensen’s studies.

In the study of CrCl3-catalyzed glucose dehydration in water by Vlachos and co-workers, it was found that the pH of the solution decreased from 2.9 to 1.6 during the process [76]. Therefore, it was assumed that the Lewis acid catalyst produced an intrinsic Brønsted acidity during its dissolution, solvation and further hydrolysis in the aqueous media, and the whole reaction was catalyzed by the combination of Lewis and Brønsted acidity. The speciation modeling of CrCl3 aqueous solutions with various concentrations was carried out. The calculated pH values at 22 and 140 °C were in good agreement with the experimental data, both of which showed that the intrinsic Brønsted acidity increased with an increase in CrCl3 concentration. The speciation analysis implied that the main species in the CrCl3 aqueous solution were [Cr(H2O)6]3+, [Cr(H2O)5Cl]2+ and [Cr(H2O)5OH]2+. With HCl addition, [Cr(H2O)5OH]2+ was the only species whose concentration kept decreasing. The kinetic experiments revealed that the rate of glucose consumption also decreased with HCl addition. In addition, the initial rate of glucose consumption had a linear relationship with the concentration of [Cr(H2O)5OH]2+. All these results indicated that [Cr(H2O)5OH]2+ was the species significantly influencing the isomerization of glucose in aqueous CrCl3. A 1,2-hydride shift mechanism was proposed and the coordination of glucose to the Cr complex was verified through EXAFS spectra analysis and CPMD (Car–Parrinello molecular dynamics) simulations. The EXAFS spectra showed that in solution the Cr center was originally coordinated to six water molecules, which were replaced then with glucose addition. In the CPMD simulations, initially two unhydrolyzed Cr3+ cations were created coordinated to a glucose molecule in water. After the calculation was started at 77 °C, only one Cr3+ remained coordinated to glucose at O1 and O2. Therefore, the subsequent simulations were performed with only one [Cr(H2O)5OH]2+ group, which contrasts with the binuclear Cr complex mechanism proposed by Hensen’s group [75]. It was claimed that the coordination of glucose to the Cr complex promoted the ring opening of glucose for the Lewis acid-catalyzed isomerization; while the intrinsic Brønsted acidity facilitated the subsequent dehydration of fructose to 5-HMF.

Wang and co-workers investigated the influence of organic solvents and Brønsted acid co-catalysts in the CrCl3-catalyzed dehydration of glucose [77]. The use of DMA led to the highest 5-HMF yield while DMSO resulted in the lowest one. Afterwards, the reaction was performed in DMA and co-catalyzed by several ionic liquids and their corresponding Brønsted acids. The 5-HMF yield decreased when H2SO4 was used. The reaction rate increased when HCl or HBr was used but the 5-HMF yield was not improved. Without CrCl3, the use of HCl or HBr alone resulted in 5-HMF production from fructose but not from glucose. Therefore, it was claimed that these co-catalysts functioned during the dehydration of fructose after glucose isomerization. DFT calculations were performed to probe the interactions among CrCl3, glucose and the co-catalysts. In the DMA-CrCl3-glucose system, similar Gibbs energy changes for CrCl3-DMA and CrCl3-glucose coordination indicated that they were both favored and the reaction was catalyzed by both CrCl3 and CrCl3-DMA. In the DMSO-CrCl3-glucose system, the extremely low Gibbs energy change of CrCl3-3DMSO coordination revealed the formation of a stable six-coordinated Cr complex, which inhibited the interaction of the Cr center with glucose for further reaction. Similarly, CrCl3-3HSO4 coordination was also quite energetically favored so glucose conversion was prevented. For the halide systems, five-coordinate Cr complexes (CrCl3-2Cl or CrCl3-2Br) were generated, having hemispheric structures so the Cr center was exposed for the interaction with glucose. The isomerization was proposed to occur via a 1,2-hydride shift. When the halides were involved, the activation barrier of the hydride shift process (the rate-limiting step) was slightly decreased but the overall barriers were increased.

Boric acid (B(OH)3) and its derivatives have been found to be able to form chelate complexes with carbohydrates [78], [79], [80], [81], [82], and catalyze the isomerization of aldohexoses to ketohexoses [83], [84]. The Riisager group achieved the first metal-free dehydration of glucose using B(OH)3 as the catalyst [85]. The 1H NMR analysis of glucose-D2 dehydration revealed less than 5 % deuterium incorporation, thus B(OH)3-catalyzed isomerization of glucose was proposed to proceed via an enolization mechanism (Scheme 4), which was different from the reactions catalyzed by metal Lewis acid catalysts. The isomerization of glucose with and without B(OH)3 catalysis were calculated through two mechanisms (1,2-hydride shift and enolization), and for both situations the enolization pathway was more favored with lower activation barriers. Caes et al. used a series of phenylboronic acids in the transformation of glucose and cellulose to 5-HMF [86]. Glucose-D2 and D2O were used, and the 1H NMR spectra showed that the deuterium from D2O instead of glucose-D2 was incorporated at C1 of the 5-HMF obtained, further indicating an enolization route similar to that proposed in Riisager’s work. This was also supported by Lukamto et al. [87] through a sophisticated mechanistic study including stereochemical considerations, competition reactions and NMR analyses. Ananikov and co-workers performed the conversion of carbohydrates to 5-HMF using boron trioxide (B2O3), and analyzed the reaction mixture through 1H, 13C and 11B NMR analyses [63]. During the reaction, a 1,2-borate complex was detected in both 1H and 13C NMR spectra, and was proposed to be 27. This may not be a true intermediate in the mechanism but a side-product formed from 26 (Scheme 4, R = OH). The function of B2O3 in this study was proposed to react with water, which could both facilitate the dehydration via Le Châtelier’s principle and generate B(OH)3 to promote the reaction.

Scheme 4: The dehydration of glucose to 5-HMF catalyzed by boronic acids (black) [63, 85–87] Brønsted acids (Qian: red [through isomerization]) [88]; orange (direct dehydration) [89], [90]; Pidko: (blue) [41] and Brønsted bases (green) [91].
Scheme 4:

The dehydration of glucose to 5-HMF catalyzed by boronic acids (black) [63, 8587] Brønsted acids (Qian: red [through isomerization]) [88]; orange (direct dehydration) [89], [90]; Pidko: (blue) [41] and Brønsted bases (green) [91].

As can be seen from the discussion above, in most studies on glucose isomerization the catalysts used were Lewis acids, mainly because of their excellent performance on the catalysis of glucose dehydration to get good 5-HMF yields. There have been some studies reported using Brønsted acid catalysts. Amarasekara and Razzaq performed a NMR study of glucose dehydration catalyzed by a functionalized Brønsted acidic ionic liquid, 1-(1-propylsulfonic)-3-methylimidazolium chloride ([C3SO3Hmim]Cl) [58]. [13C-1] and [13C-2] labeled glucose were used as reactants, and the resonances of C1 and C2 of the open-chain fructose, two intermediates (36 (Table 2) and 2) and 5-HMF were observed in 13C NMR spectra. The isomerization was proposed to proceed along a 1,2-hydride shift pathway because the authors claimed that no deuterium incorporation was detected. However, the solvent used herein was DMSO-d6, which does not release free deuterium into solution. Therefore, it would be more convincing if D2O was used as the solvent for the deuterium detection experiments.

Because of the harsh conditions (e.g. high temperature, long reaction time) required and the low yields of 5-HMF obtained in lab experiments, Brønsted acid-catalyzed glucose isomerization/dehydration has been studied extensively via computational calculations. In the work by Qian’s group, a cyclic route through a 1,2-hydride shift was proposed for glucose isomerization (Scheme 4) [88]. The CPMD-metadynamics (MTD) simulation results in the gas phase showed that the reaction barrier of the 1,2-hydride shift was as low as 20.9–29.3 kJ/mol because 4 was quite stable. Nevertheless, when the calculation was performed in water, 29 became less stable and the proton of O1H was easily removed by water to form 2′. In the DFT study carried out by Pidko and co-workers [41], the protonation of O5 led to the ring opening of glucose, and the protonated open-chain fructose was obtained via a 1,2-hydride shift (Scheme 4). This mechanism was supported by Daorattanachai et al. [91], but their calculations were performed in gas phase and there was no explanation about their choice of the O5 protonated glucose as the starting point. Therefore, further study is needed for more accurate and reasonable results. In a quantum mechanical study by Arifin et al. [48], both the cyclic and acyclic mechanisms of HCl-catalyzed isomerization of glucose were simulated in water and 1,3-dimethylimidazolium chloride ([Mmim]Cl). The cyclic mechanism was favored in water with an activation barrier of 99.6 kJ/mol, and the acyclic mechanism was favored in [Mmim]Cl with a barrier of 135.6 kJ/mol. In the DFT calculations by Li et al. of [BmimSO3H]Cl-catalyzed dehydration of glucose, two pathways were proposed [49]. Path I is through the isomerization and Path II is the direct conversion to 5-HMF through an enediol intermediate. It was claimed that both pathways were accessible because they had similar activation barriers. The catalysis by [BmimSO3H]Cl was achieved mainly through -SO3H acting as a proton shuttle, and the intermediates and transition states were stabilized via hydrogen bonding with Cl anions.

A poly-benzyl ammonium chloride resin was used as the catalyst in glucose dehydration by Zhang’s group [92]. An enolization route catalyzed by both the benzyl ammonium cation and the benzyl amine was proposed. The DFT calculations indicated that the whole isomerization process was exothermic, possibly because the intermediates were stabilized through hydrogen bonding. The dehydration of glucose was performed in DMSO/D2O system and deuterium incorporation at C1 of the 5-HMF product was observed in the 13C NMR spectra, hence the enolization path was verified.

The use of Brønsted bases as catalysts in glucose isomerization have also been investigated. The Davis group performed the isomerization of glucose-D2 using sodium hydroxide (NaOH), and no deuterium incorporation was observed [69]. Hence a proton transfer mechanism was proposed, in which the deuterium at C2 of glucose-D2 was transferred into the solution and a proton from the solution was incorporated into the fructose. A Brønsted-base catalyzed isomerization of glucose was simulated by Daorattanachai et al., starting from the deprotonation of O1 of glucose (32, Scheme 4) [91]. The deprotonated open-chain glucose (33) was isomerized to the deprotonated fructose (35) through the deprotonated 1,2-enediol intermediate (34) with an overall activation barrier of 173.7 kJ/mol. Carraher et al. performed the isomerization of glucose using aqueous triethylamine (TEA) [93]. A KIE was observed (kH/kD = 3.8) when using glucose-D2 as the reactant, indicating that the deprotonation at C2 was the rate-limiting step. A similar enolization mechanism was proposed.

Direct dehydration of glucose without isomerization to fructose

3-Deoxyglucosone (3-DG, Scheme 3, route A) has been proposed as an intermediate in the dehydration of carbohydrates to 5-HMF [28, 29, 52]. This mechanism has been supported by some early studies [94], [95], [96], [97], [98]. The study by Bols’ group verified this mechanism through a combination of experiments and calculations [53]. In the H2SO4-catalyzed dehydration of glucose, both 3-DG and fructose were observed in the ion chromatogram, but higher product yields and reaction rates were obtained from the conversion of 3-DG than fructose. The DFT calculations of glucose dehydration proceeding via 3-DG or fructose were performed. The Gibbs energies of all intermediates in the 3-DG route were lower than those in the fructose route, indicating the preference of the 3-DG route over the fructose one.

The Hara group performed glucose dehydration catalyzed by titanium oxide (TiO2) and phosphate-immobilized TiO2 (phosphate/TiO2) [61]. In the 2H and 13C NMR spectra, no deuterium incorporation was detected in the 5-HMF generated from glucose-D2, and incorporation much higher than 54 % was observed for the reaction of glucose-D1.Therefore, a mechanism through 3-DG was proposed. The 13C CP/MAS NMR spectra of [13C-2] glucose absorbed on TiO2 and phosphate/TiO2 at room temperature showed two resonances at 110 and 118 ppm respectively, which were assigned to C2 of 3-DG. The reaction was also performed using scandium triflate (Sc(OTf)3), and 55 and 54 % deuterium incorporation in the 5-HMF was seen from glucose-D2 and glucose-D1 respectively. This indicated that the Sc(OTf)3-catalyzed reaction proceeded through the isomerization route via a 1,2-hydride shift.

Qian’s group simulated the Brønsted acid-catalyzed conversion of glucose to 2′ both in vacuum and in water [89], [90]. As discussed above, compound 2 or 2′ was proposed to be an important intermediate in the dehydration of fructose to 5-HMF. Here it was also assumed to be significant in the direct dehydration of glucose to 5-HMF (Scheme 4, orange route). The pathway was initialized by the protonation at different hydroxyl groups or the O atom on the ring (O5), and in both situations (with and without solvent) only protonation at O2H could lead to C2-O2 cleavage and the subsequent formation of 29 (the precursor of 2′) via C1-O5 breaking and C2-O5 bonding. The NMR analyses from previous studies indicated that the downfield shifts of the hydroxyl groups after proton exchange with water decreased in the order O6H > O2H > O3H, O4H > O1H, corresponding to the decrease of their proton affinities [99], [100]. This further supports the calculation results that the O2H protonated glucose was the most likely starting point since O2H has a relatively high proton affinity. This cyclic mechanism of glucose direct dehydration was further verified using a more comprehensive CPMD-MTD method [101]. This reaction route is supported by Assary et al. [40] through Gibbs free energy calculations and Yang et al. [102] through the analysis of reaction fluxes in the kinetic models composed of several glucose reactions (isomerization, direct dehydration, etc.). Assary et al. also proposed an acyclic route, in which O1 of glucose was protonated after the ring opening. However, it should be noted that the protonation of O1H as the first step was rarely seen in computational studies of glucose dehydration.

Zhang et al. performed a DFT study of glucose transformation in supercritical water [103]. An acyclic mechanism without glucose isomerization was proposed. Supercritical water effectively decreased the activation barrier for the ring opening of glucose and the removal of the first two water molecules, but not the third dehydration (the rate-limiting step). Therefore, glucose dehydration was extremely challenging with a high energy demand in supercritical water.

For the transformation of glucose to 5-HMF, both the direct dehydration route and the indirect route through the isomerization to fructose have been investigated extensively. Among the isomerization studies, a 1,2-hydride shift is favored when metal Lewis acids are used as catalysts; while enolization is more dominant for B(OH)3 and boronic acid catalysts. The Brønsted acid-catalyzed dehydration/isomerization of glucose has been studied more computationally than experimentally, and a variety of mechanisms have been simulated. In the direct dehydration mechanisms proposed, 3-DG has been listed as a key intermediate.

Conclusions

Mechanisms of fructose and glucose dehydration to 5-HMF that were proposed from lab experiments or computational calculations have been reviewed. The mechanisms are closely related to the reaction environment, such as the solvent and catalyst used. NMR spectroscopy and computational calculations are the two most frequently used techniques in mechanistic studies, but other techniques such as kinetic studies, mass spectrometry, IR and UV spectroscopies can also be used to get a full picture of the reaction mechanisms.


Corresponding author: Francesca M. Kerton, Department of Chemistry, Memorial University of Newfoundland, St. John’s, A1B 3X7, NL, Canada, e-mail:

  1. Article note: A collection of peer-reviewed articles dedicated to Chemical Research Applied to World Needs (CHEMRAWN).

References

[1] Biomass Energy Basics, National Renewable Energy Laboratory, USA, https://www.nrel.gov/research/re-biomass.html (accessed Nov 2, 2020).Search in Google Scholar

[2] J. E. Holladay, J. J. Bozell, J. F. White, D. Johnson. reportTop value added chemicals from biomass, PNNL-16983 Technical Report, Pacific Northwest National Laboratory, Richland, Washington (2004), http://www.nrel.gov/docs/fy04osti/35523.pdf (accessed Nov 2, 2020).Search in Google Scholar

[3] F. Cherubini, A. H. Strømman. Biofuels Bioprod. Biorefin.5, 548 (2011).10.1002/bbb.297Search in Google Scholar

[4] J. N. Chheda, G. W. Huber, J. A. Dumesic. Angew. Chem. Int. Ed.46, 7164 (2007).10.1002/anie.200604274Search in Google Scholar PubMed

[5] F. M. Kerton, Y. Liu, K. W. Omari, K. Hawboldt. Green Chem.15, 860 (2013).10.1039/c3gc36994cSearch in Google Scholar

[6] J. Song, H. Fan, J. Ma, B. Han. Green Chem.15, 2619 (2013).10.1039/c3gc41141aSearch in Google Scholar

[7] S. G. Wettstein, D. M. Alonso, E. I. Gürbüz, J. A. Dumesic. Curr. Opin. Chem. Eng.1, 218 (2012).10.1016/j.coche.2012.04.002Search in Google Scholar

[8] A. Takagaki, S. Nishimura, K. Ebitani. Catal. Surv. Asia16, 164 (2012).10.1007/s10563-012-9142-3Search in Google Scholar

[9] R.-J. van Putten, J. C. van der Waal, E. De Jong, C. B. Rasrendra, H. J. Heeres, J. G. de Vries. Chem. Rev.113, 1499 (2013).10.1021/cr300182kSearch in Google Scholar PubMed

[10] A. Mukherjee, M.-J. Dumont, V. Raghavan. Biomass Bioenergy72, 143 (2015).10.1016/j.biombioe.2014.11.007Search in Google Scholar

[11] T. Wang, M. W. Nolte, B. H. Shanks. Green Chem.16, 548 (2014).10.1039/C3GC41365ASearch in Google Scholar

[12] W. Deng, Q. Zhang, Y. Wang. Sci. China Chem.58, 29 (2015).10.1007/s11426-014-5283-8Search in Google Scholar

[13] M. B. Fusaro, V. Chagnault, D. Postel. Carbohydr. Res.409, 9 (2015).10.1016/j.carres.2015.03.012Search in Google Scholar PubMed

[14] H. Rasmussen, H. R. Sørensen, A. S. Meyer. Carbohydr. Res.385, 45 (2014).10.1016/j.carres.2013.08.029Search in Google Scholar PubMed

[15] S. Dutta, S. Pal. Biomass Bioenergy62, 182 (2014).10.1016/j.biombioe.2013.12.019Search in Google Scholar

[16] S. Dutta, S. De, B. Saha, M. I. Alam. Catal. Sci. Technol.2, 2025 (2012).10.1039/c2cy20235bSearch in Google Scholar

[17] A. S. Nagpure, A. K. Venugopal, N. Lucas, M. Manikandan, R. Thirumalaiswamy, S. Chilukuri. Catal. Sci. Technol.5, 1463 (2015).10.1039/C4CY01376JSearch in Google Scholar

[18] O. Casanova, S. Iborra, A. Corma. J. Catal.275, 236 (2010).10.1016/j.jcat.2010.08.002Search in Google Scholar

[19] K. Mliki, M. Trabelsi. Res. Chem. Intermed.42, 8253 (2016).10.1007/s11164-016-2593-9Search in Google Scholar

[20] F. Yang, S. Zhang, Z. C. Zhang, J. Mao, S. Li, J. Yin, J. Zhou. Catal. Sci. Technol.5, 4602 (2015).10.1039/C5CY00750JSearch in Google Scholar

[21] E. Hayashi, T. Komanoya, K. Kamata, M. Hara. ChemSusChem10, 654 (2017).10.1002/cssc.201601443Search in Google Scholar PubMed

[22] X. Han, C. Li, X. Liu, Q. Xia, Y. Wang. Green Chem.19, 996 (2017).10.1039/C6GC03304KSearch in Google Scholar

[23] H. Choudhary, K. Ebitani. Chem. Lett.45, 613 (2016).10.1246/cl.160178Search in Google Scholar

[24] X. Tong, Y. Ma, Y. Li. Appl. Catal., A385, 1 (2010).10.1016/j.apcata.2010.06.049Search in Google Scholar

[25] S. Zhong, R. Daniel, H. Xu, J. Zhang, D. Turner, M. L. Wyszynski, P. Richards. Energy Fuels24, 2891 (2010).10.1021/ef901575aSearch in Google Scholar

[26] F. Jensen. in Introduction to Computational Chemistry, John Wiley & Sons Ltd, West Sussex, England, 2nd ed. (2017).Search in Google Scholar

[27] J. U. Nef. Ann. Chem.376, 1 (1910).10.1002/jlac.19103760102Search in Google Scholar

[28] E. F. L. J. Anet. J. Am. Chem. Soc.82, 1502 (1960).10.1021/ja01491a055Search in Google Scholar

[29] E. F. L. J. Anet. Aust. J. Chem.18, 240 (1965).10.1071/CH9650240Search in Google Scholar

[30] G. R. Akien, L. Qi, I. T. Horváth. Chem. Commun.48, 5850 (2012).10.1039/c2cc31689gSearch in Google Scholar PubMed

[31] A. S. Amarasekara, L. D. Williams, C. C. Ebede. Carbohydr. Res.343, 3021 (2008).10.1016/j.carres.2008.09.008Search in Google Scholar PubMed PubMed Central

[32] H. Kimura, M. Nakahara, N. Matubayasi. J. Phys. Chem. A117, 2102 (2013).10.1021/jp312002hSearch in Google Scholar PubMed

[33] J. Zhang, A. Das, R. S. Assary, L. A. Curtiss, E. Weitz. Appl. Catal. B181, 874 (2016).10.1016/j.apcatb.2014.10.056Search in Google Scholar

[34] J. Zhang, E. Weitz. ACS Catal.2, 1211 (2012).10.1021/cs300045rSearch in Google Scholar

[35] M. J. Antal, W. S. Mok, G. N. Richards. Carbohydr. Res.199, 91 (1990).10.1016/0008-6215(90)84096-DSearch in Google Scholar

[36] S. H. Mushrif, S. Caratzoulas, D. G. Vlachos. Phys. Chem. Chem. Phys.14, 2637 (2012).10.1039/c2cp22694dSearch in Google Scholar PubMed

[37] S. Caratzoulas, D. G. Vlachos. Carbohydr. Res.346, 664 (2011).10.1016/j.carres.2011.01.029Search in Google Scholar PubMed

[38] N. Nikbin, S. Caratzoulas, D. G. Vlachos. ChemCatChem4, 504 (2012).10.1002/cctc.201100444Search in Google Scholar

[39] R. S. Assary, P. C. Redfern, J. Greeley, L. A. Curtiss. J. Phys. Chem. B115, 4341 (2011).10.1021/jp1104278Search in Google Scholar PubMed

[40] R. S. Assary, T. Kim, J. J. Low, J. Greeley, L. A. Curtiss. Phys. Chem. Chem. Phys.14, 16603 (2012).10.1039/c2cp41842hSearch in Google Scholar PubMed

[41] G. Yang, E. A. Pidko, E. J. M. Hensen. J. Catal.295, 122 (2012).10.1016/j.jcat.2012.08.002Search in Google Scholar

[42] T. D. Swift, C. Bagia, V. Choudhary, G. Peklaris, V. Nikolakis, D. G. Vlachos. ACS Catal.4, 259 (2013).10.1021/cs4009495Search in Google Scholar

[43] F. S. Asghari, H. Yoshida. Ind. Eng. Chem. Res.46, 7703 (2007).10.1021/ie061673eSearch in Google Scholar

[44] C. Moreau, R. Durand, S. Razigade, J. Duhamet, P. Faugeras, P. Rivalier, P. Ros, G. Avignon. Appl. Catal., A145, 211 (1996).10.1016/0926-860X(96)00136-6Search in Google Scholar

[45] C. Moreau, A. Finiels, L. Vanoye. J. Mol. Catal. Chem.253, 165 (2006).10.1016/j.molcata.2006.03.046Search in Google Scholar

[46] Y. Li, X. Lu, L. Yuan, X. Liu. Biomass Bioenergy33, 1182 (2009).10.1016/j.biombioe.2009.05.003Search in Google Scholar

[47] C. Yao, Y. Shin, L.-Q. Wang, C. F. Windisch, W. D. Samuels, B. W. Arey, C. Wang, W. M. Risen, G. J. Exarhos. J. Phys. Chem. C111, 15141 (2007).10.1021/jp074188lSearch in Google Scholar

[48] M. Puripat, D. Yokogawa, V. Parasuk, S. Irle. J. Comput. Chem.37, 327 (2016).10.1002/jcc.24214Search in Google Scholar PubMed

[49] J. Li, J. Li, D. Zhang, C. Liu. J. Phys. Chem. B119, 13398 (2015).10.1021/acs.jpcb.5b07773Search in Google Scholar PubMed

[50] Y.-N. Li, J.-Q. Wang, L.-N. He, Z.-Z. Yang, A.-H. Liu, B. Yu, C.-R. Luan. Green Chem.14, 2752 (2012).10.1039/c2gc35845jSearch in Google Scholar

[51] J. Guan, Q. Cao, X. Guo, X. Mu. Comput. Theor. Chem.963, 453 (2011).10.1016/j.comptc.2010.11.012Search in Google Scholar

[52] E. F. L. J. Anet. Adv. Carbohydr. Chem.19, 181 (1964).Search in Google Scholar

[53] H. Jadhav, C. M. Pedersen, T. Sølling, M. Bols. ChemSusChem4, 1049 (2011).10.1002/cssc.201100249Search in Google Scholar PubMed

[54] H. Kimura, M. Nakahara, N. Matubayasi. J. Phys. Chem. A115, 14013 (2011).10.1021/jp206355eSearch in Google Scholar PubMed

[55] D. W. Harris, M. S. Feather. J. Org. Chem.39, 724 (1974).10.1021/jo00919a036Search in Google Scholar

[56] D. W. Harris, M. S. Feather. J. Am. Chem. Soc.97, 178 (1975).10.1021/ja00834a031Search in Google Scholar

[57] S. J. Angyal, in Glycoscience: Epimerisation, Isomerisation and Rearrangement Reactions of Carbohydrates, A. E. Stütz (Ed.), pp. 1–14, Springer Berlin Heidelberg, Berlin, Heidelberg, (2001).Search in Google Scholar

[58] A. S. Amarasekara, A. Razzaq. Carbohydr. Res.386, 86 (2014).10.1016/j.carres.2014.01.009Search in Google Scholar PubMed

[59] R. Bermejo-Deval, R. S. Assary, E. Nikolla, M. Moliner, Y. Román-Leshkov, S.-J. Hwang, A. Palsdottir, D. Silverman, R. F. Lobo, L. A. Curtiss. Proc. Natl. Acad. Sci. U.S.A.109, 9727 (2012).10.1073/pnas.1206708109Search in Google Scholar PubMed PubMed Central

[60] Z. Zhang, Q. Wang, H. Xie, W. Liu, Z. K. Zhao. ChemSusChem4, 131 (2011).10.1002/cssc.201000279Search in Google Scholar PubMed

[61] R. Noma, K. Nakajima, K. Kamata, M. Kitano, S. Hayashi, M. Hara. J. Phys. Chem. C119, 17117 (2015).10.1021/acs.jpcc.5b03290Search in Google Scholar

[62] S. Hu, Z. Zhang, J. Song, Y. Zhou, B. Han. Green Chem.11, 1746 (2009).10.1039/b914601fSearch in Google Scholar

[63] E. A. Khokhlova, V. V. Kachala, V. P. Ananikov. ChemSusChem5, 783 (2012).10.1002/cssc.201100670Search in Google Scholar PubMed

[64] S. De, S. Dutta, B. Saha. Green Chem.13, 2859 (2011).10.1039/c1gc15550dSearch in Google Scholar

[65] H. Zhao, J. E. Holladay, H. Brown, Z. C. Zhang. Science316, 1597 (2007).10.1126/science.1141199Search in Google Scholar PubMed

[66] Z. Zhang, Z. K. Zhao. Bioresour. Technol.102, 3970 (2011).10.1016/j.biortech.2010.11.098Search in Google Scholar PubMed

[67] D. Neuhaus, M. P. Williamson. in The Nuclear Overhauser Effect in Structural and Conformational Analysis, Wiley VCH, New York, 2nd ed. (1989).Search in Google Scholar

[68] J. B. Binder, A. V. Cefali, J. J. Blank, R. T. Raines. Energy Environ. Sci.3, 765 (2010).10.1039/b923961hSearch in Google Scholar

[69] Y. Román-Leshkov, M. Moliner, J. A. Labinger, M. E. Davis. Angew. Chem. Int. Ed.49, 8954 (2010).10.1002/anie.201004689Search in Google Scholar PubMed

[70] V. Choudhary, A. B. Pinar, R. F. Lobo, D. G. Vlachos, S. I. Sandler. ChemSusChem6, 2369 (2013).10.1002/cssc.201300328Search in Google Scholar PubMed

[71] J. Tang, X. Guo, L. Zhu, C. Hu. ACS Catal.5, 5097 (2015).10.1021/acscatal.5b01237Search in Google Scholar

[72] T. Wang, J. A. Glasper, B. H. Shanks. Appl. Catal., A498, 214 (2015).10.1016/j.apcata.2015.03.037Search in Google Scholar

[73] E. A. Pidko, V. Degirmenci, R. A. van Santen, E. J. Hensen. Angew. Chem. Int. Ed.49, 2530 (2010).10.1002/anie.201000250Search in Google Scholar PubMed

[74] E. A. Pidko, V. Degirmenci, R. A. van Santen, E. J. Hensen. Inorg. Chem.49, 10081 (2010).10.1021/ic101402rSearch in Google Scholar PubMed

[75] Y. Zhang, E. A. Pidko, E. J. Hensen. Chem. Eur. J.17, 5281 (2011).10.1002/chem.201003645Search in Google Scholar PubMed

[76] V. Choudhary, S. H. Mushrif, C. Ho, A. Anderko, V. Nikolakis, N. S. Marinkovic, A. I. Frenkel, S. I. Sandler, D. G. Vlachos. J. Am. Chem. Soc.135, 3997 (2013).10.1021/ja3122763Search in Google Scholar PubMed

[77] Y. Yang, W. Liu, N. Wang, H. Wang, Z. Song, W. Li. RSC Adv.5, 27805 (2015).10.1039/C5RA02057CSearch in Google Scholar

[78] R. van den Berg, J. A. Peters, H. van Bekkum. Carbohydr. Res.253, 1 (1994).10.1016/0008-6215(94)80050-2Search in Google Scholar

[79] S. Arimori, M. D. Phillips, T. D. James. Tetrahedron Lett.45, 1539 (2004).10.1016/j.tetlet.2003.12.013Search in Google Scholar

[80] D. Stones, S. Manku, X. Lu, D. G. Hall. Chem. Eur. J.10, 92 (2004).10.1002/chem.200305400Search in Google Scholar

[81] M. Dowlut, D. G. Hall. J. Am. Chem. Soc.128, 4226 (2006).10.1021/ja057798cSearch in Google Scholar

[82] M. Bielecki, H. Eggert, J. Norrild. J. Chem. Soc. Perkin Trans.2, 449 (1999).10.1039/a808896iSearch in Google Scholar

[83] J. F. Mendicino. J. Am. Chem. Soc.82, 4975 (1960).10.1021/ja01503a055Search in Google Scholar

[84] K. B. Hicks, E. V. Symanski, P. E. Pfeffer. Carbohydr. Res.112, 37 (1983).10.1016/0008-6215(83)88264-0Search in Google Scholar

[85] T. Ståhlberg, S. Rodriguez-Rodriguez, P. Fristrup, A. Riisager. Chem. Eur. J.17, 1456 (2011).10.1002/chem.201002171Search in Google Scholar PubMed

[86] B. R. Caes, M. J. Palte, R. T. Raines. Chem. Sci.4, 196 (2013).10.1039/C2SC21403BSearch in Google Scholar PubMed PubMed Central

[87] D. H. Lukamto, P. Wang, T. P. Loh. Asian J. Org. Chem.2, 947 (2013).10.1002/ajoc.201300185Search in Google Scholar

[88] X. Qian. Top. Catal.55, 218 (2012).10.1007/s11244-012-9790-6Search in Google Scholar

[89] X. Qian, M. R. Nimlos, D. K. Johnson, M. E. Himmel. Appl. Biochem. Biotechnol.121-124, 989 (2005).10.1385/ABAB:124:1-3:0989Search in Google Scholar

[90] X. Qian, M. R. Nimlos, M. Davis, D. K. Johnson, M. E. Himmel. Carbohydr. Res.340, 2319 (2005).10.1016/j.carres.2005.07.021Search in Google Scholar

[91] P. Daorattanachai, S. Namuangruk, N. Viriya-Empikul, N. Laosiripojana, K. Faungnawakij. J. Ind. Eng. Chem.18, 1893 (2012).10.1016/j.jiec.2012.04.019Search in Google Scholar

[92] X. Cao, S. P. Teong, D. Wu, G. Yi, H. Su, Y. Zhang. Green Chem.17, 2348 (2015).10.1039/C4GC02488ESearch in Google Scholar

[93] J. M. Carraher, C. N. Fleitman, J. -P. Tessonnier. ACS Catal.5, 3162 (2015).10.1021/acscatal.5b00316Search in Google Scholar

[94] M. Wolfrom, R. Schuetz, L. F. Cavalieri. J. Am. Chem. Soc.70, 514 (1948).10.1021/ja01182a025Search in Google Scholar

[95] L. Evans, A. Gillam. J. Chem. Soc.1943, 565 (1943).10.1039/JR9430000565Search in Google Scholar

[96] C. D. Hurd, L. L. Isenhour. J. Am. Chem. Soc.54, 317 (1932).10.1021/ja01340a048Search in Google Scholar

[97] M. Wolfrom, E. Wallace, E. Metcalf. J. Am. Chem. Soc.64, 265 (1942).10.1021/ja01254a017Search in Google Scholar

[98] K. Lourvanij, G. L. Rorrer. Ind. Eng. Chem. Res.32, 11 (1993).10.1021/ie00013a002Search in Google Scholar

[99] J. M. Harvey, M. C. Symons. J. Solut. Chem.7, 571 (1978).10.1007/BF00646035Search in Google Scholar

[100] J. M. Harvey, M. C. Symons, R. J. Naftalin. Nature261, 435 (1976).10.1038/261435a0Search in Google Scholar PubMed

[101] X. Qian. J. Phys. Chem. A115, 11740 (2011).10.1021/jp2041982Search in Google Scholar PubMed

[102] L. Yang, G. Tsilomelekis, S. Caratzoulas, D. G. Vlachos. ChemSusChem8, 1334 (2015).10.1002/cssc.201403264Search in Google Scholar PubMed

[103] Y. Zhang, C. Liu, X. Chen. J. Anal. Appl. Pyrolysis119, 199 (2016).10.1016/j.jaap.2016.02.018Search in Google Scholar

Published Online: 2021-03-31
Published in Print: 2021-04-27

© 2021 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/

Downloaded on 23.4.2024 from https://www.degruyter.com/document/doi/10.1515/pac-2020-1108/html
Scroll to top button