You are on page 1of 40

Chem Soc Rev

View Article Online


REVIEW ARTICLE View Journal | View Issue
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Catalytic oxidation of carbohydrates into organic


acids and furan chemicals
Cite this: Chem. Soc. Rev., 2018,
47, 1351
Zehui Zhang *a and George W. Huber *b

A wide variety of commodity chemicals can be produced from the catalytic oxidation of carbohydrates
or carbohydrate derived molecules including formic acid, acetic acid, glycolic acid, gluconic acid,
glucaric acid, malonic acid, oxalic acid, 2,5-diformylfuran (DFF), 5-hydroxymethyl-2-furancarboxylic acid
(HFCA), 5-formyl-2-furancarboxylic acid (FFCA), and 2,5-furandicarboxylic acid (FDCA). This review will
highlight the recent research progress in the development of new routes for the production of organic
acids and furan compounds via catalytic oxidation reactions. Particular attention will be paid to these
one-pot reactions with the requirements of an acidic site and a metal site. For the one-pot
transformation of cellobiose or lignocellulose into gluconic acid, these reactions were performed via a
one-step strategy using a single catalyst containing an acidic site and a metal site. However, a two-step
Received 23rd March 2017 strategy was adopted for the oxidative transformation of carbohydrates into DFF or FDCA in order
DOI: 10.1039/c7cs00213k to avoid the oxidation of the carbohydrates. The first step was performed for the dehydration of
carbohydrates into 5-hydroxylmethylfuran (HMF) in the presence of an acid catalyst, and the second
rsc.li/chem-soc-rev step was performed for the oxidation of HMF into DFF or FDCA with a metal catalyst.

1. Introduction
Biomass typically contains 40–45% weight percent of oxygen.1
a
Key Laboratory of Catalysis and Material Sciences of the State Ethnic Affairs In the petrochemical industry a wide range of commodity
Commission & Ministry of Education, College of Chemistry and Material Sciences,
chemicals are produced via selective oxidation of petroleum
South-Central University for Nationalities, Wuhan, 430074, China.
E-mail: zehuizh@mail.ustc.edu.cn
derived feedstocks. These molecules could potentially be more
b
Department of Chemical and Biological Engineering, University of inexpensively produced from biomass, because it is typically
Wisconsin-Madison, Madison, WI 53706, USA. E-mail: huber@engr.wisc.edu more economical to remove oxygen from a molecule than add

Professor Zehui Zhang obtained his George W. Huber is the Harvey


PhD degree at Dalian Institute of Spangler Professor at University of
Chemical Physics (DICP) in 2011, Wisconsin-Madison. His research
Chinese Academy of Science. After focus is on developing new catalytic
that, he joined the Key Laboratory processes for the production
of Catalysis and Materials Sciences of renewable liquid fuels and
of the Ministry of Education, South- chemicals. He is a co-founder of
Central University for Nationalities Anellotech (www.anellotech.com)
(SCUEC), and started his indepen- a biochemical company focused
dent research. His current research on commercializing catalytic fast
interest is the design of new pyrolysis, a technology to produce
catalytic materials for the con- renewable aromatics from biomass.
Zehui Zhang version of biomass into chemicals George W. Huber George did a post-doctoral stay with
and liquid fuels and organic Avelino Corma at the Polytechnical
transformations. He is the director of the catalysis and green University of Valencia, Spain. He obtained his PhD in Chemical
chemistry research group at SCUEC. He has published more than Engineering from University of Wisconsin-Madison (2005) studying
80 papers with a H-Index of 30 and has applied for 10 patents. under the direction of James Dumesic. George has published over 138
publications which have been cited over 24 000 times.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1351
View Article Online

Review Article Chem Soc Rev


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fig. 1 Chemicals produced by oxidation of glucose and glucose derived molecules.

oxygen via selective oxidization reactions. Many recent papers conversion of carbohydrates into these furan chemicals has
have been published on oxidation of carbohydrates or biomass also been reported using acid–metal bifunctional catalysts.4 As
derived molecules into value-added chemicals.2–5 Fig. 1 shows will be described in detail in this review the above mentioned
that the oxidative transformation of biomass or biomass-derived biomass-derived oxygenates can be alternatives to fossil resource
molecules can produce a wide range of products including formic derived chemicals as described in Table 1. For example, FDCA
acid, acetic acid, glycolic acid, malonic acid, oxalic acid, gluconic can replace the fossil resource ( p-xylene) for the production of
acid, glucaric acid and furan chemicals. For example, gluconic polyesters.7 Currently, 98% of p-xylene produced commercially
acid can be obtained by the direct oxidation of glucose, or the one- is used as a feedstock for the production of terephthalic acid, an
pot transformation of polysaccharides and lignocellulose over ingredient for polyethylene terephthalate (PET), which is then
acid–metal bifunctional catalysts. The catalytic oxidative cleavage converted to fibers, films, containers, packaging materials, molded
of carbon–carbon bonds generates smaller and molecular organic articles, and household consumable goods. However, some of the
acids such as formic acid, acetic acid, glycolic acid, malonic acid, biomass-derived chemicals have a high production cost and have
and oxalic acid. 5-Hydroxymethylfurfural (HMF) is a renewable limited annual production ability. For example, the production
platform chemical, which is produced from the dehydration of C6 cost of FDCA is high and the annual production ability is low
carbohydrates.6 Catalytic oxidation of HMF generates several furan (Table 1). There are very few companies that manufacture and sell
derivatives including 2,5-diformylfuran (DFF), 5-hydroxymethyl-2- FDCA on demand due to the lack of economic viability. The main
furancarboxylic acid (HFCA), 5-formyl-2-furancarboxylic acid current producers of FDCA are Synbias, V & V Pharma Industries,
(FFCA), and 2,5-furandicarboxylic acid (FDCA).4 The one-pot Carbone Scientific Tokyo Chemical Industry and Chemsky.

Table 1 Market analysis and applications of commodity chemicals produced by oxidation of carbohydrates

Market price Volume (metric


Chemicals (US $ per metric ton)8 tons per year) Main applications
Formic acid9 400 720 000 Silage and animal feed preservation, leather and tanning, textiles, formate salts,
pharmaceuticals/food chemicals, rubber chemicals, catalysts and plasticizers.
Acetic acid10 450–690 6 500 000 Production of cellulose. Acetate for photographic film and polyvinyl acetate
for wood glue, synthetic fibers and fabrics.
Glycolic acid11 100 000–300 000 15 000 USA12 As a dyeing and tanning agent in the textile industry; as a flavoring agent and
as a preservative in food processing; as a skin care agent in the pharmaceutical
industry; it is also used in adhesives and plastics; a useful intermediate for
organic synthesis, in a range of reactions.
Gluconic acid13 760–830 60 000 Food additive, an ingredient for concrete and raw materials for medicines
and biodegradable polymers.
FDCA14 690 000–720 000 40 Mainly used as a polymer building block as a substitute of terephthalic acid in the
production of polyesters and other current polymers containing an aromatic moiety.

1352 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

Aventium is currently running a pilot plant with a 40 ton per hydrolysis of formic acid methyl ester with water to produce
annum capacity in Geelen. However, Aventium has announced formic acid (this reaction is performed with excessive water to
a new technology involving a reactive separation technology shift the equilibrium at 373–413 K and atmospheric pressure).
and catalyst which would result in the economic production of Recently, there have been some papers on the synthesis
FDCA at USD 1000 per ton from 2016.14 of formic acid by aqueous catalytic partial oxidation of wet
The objective of this review is to document the production of biomass mainly over vanadium catalysts.
commodity chemicals that can be produced from carbo-
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

hydrates and carbohydrate derived furan compounds by chemical 2.1 Conversion of carbohydrates into formic acid with O2 as
oxidation routes. In this review, we will highlight the recent the oxidant
progress in the development of efficient catalytic systems Oxidation of carbohydrates into formic acid was first reported
for the production of formic acid, acetic acid, glycolic acid, in 1954.22 Schöpf and Wild carried out the oxidation of glucose
malonic acid, oxalic acid, gluconic acid, glucaric acid, HFCA, into formic acid using stoichiometric amounts of periodic acid
FFCA, DFF and FDCA. (H5IO6).22 Catalytic oxidation of carbohydrates to formic acid with
molecular oxygen (O2) would be preferred over using periodic
acid, as O2 is a clean oxidant. Fig. 2a shows the stoichiometric
2. Conversion of carbohydrates into reaction of one mole of glucose with three moles of O2 to produce
formic acid 6 moles of formic acid without any other product. However, a
stoichiometric amount of HI is also produced, when H5IO6 is
Formic acid is a colorless liquid with a pungent odor, which is used as the oxidant.
completely miscible with water and many polar solvents but The reaction pathway of the catalytic oxidation of cellulose
only partially miscible with hydrocarbons.9 Formic acid has into formic acid is illustrated in Fig. 2b. It contains two major
an annual production of 720 000 tons per year.9 Formic acid steps: the acid-catalyzed hydrolysis of carbohydrates (cellulose
is mainly used as a preservative and antibacterial agent in or lignocellulose) into monosaccharides (glucose or xylose) and
livestock feed.9 It can also be used in the production of leather the oxidative cleavage of C–C bonds in monosaccharides into
for tanning, dyeing and finishing, and in place of mineral acids formic acid. Over oxidation of monosaccharides will result in the
for various cleaning products, such as limescale removers and complete combustion of biomass to water and CO2, which are
toilet bowl cleaners.9 It can also be used in organic synthesis the thermodynamically favored products. Thus, it is of critical
reactions as a source of formyl groups15 and a hydrogen source importance that the selected catalyst systems do not oxidize
for transfer hydrogenation reactions.16 Recently, formic acid has formic acid to CO2 under the applied reaction conditions.
been proposed as a feedstock for H2 storage since it contains O2 is usually activated by metal catalysts. Vanadium(V)-
4.4 wt% mass fraction of hydrogen and can decompose into containing catalysts, especially polyoxometalates (POM) with a
H2 over metal catalysts under mild conditions (o100 1C).17–19 formula of H3+nPVnMo12 nO40 (HPA), are effective for formic
Formic acid has also been proposed to power fuel cells for acid production from biomass with O2 as an oxidant as shown
electricity generation and automobiles.20 This concept builds in Table 2.23–29 In 2010, Wasserscheid and co-workers first
on the 4.4 wt% share of hydrogen in formic acid resulting in reported the aqueous transformation of carbohydrates into formic
53 g hydrogen per liter of formic acid.17–19 The commercial acid in a batch reactor from 333 to 363 K using H5PV2Mo10O40
production of formic acid from fossil fuel derived methanol occurs (HPA-2) as the catalyst.23 Monosaccharides (glucose, xylose) and
by two steps:21 (1) the carboxylation of methanol with carbon disaccharides (cellobiose, sucrose) produced formic acid in yields
monoxide to generate formic acid methyl ester (this reaction is of around 50% at nearly full conversions (498%) at a substrate
performed in the liquid phase in the presence of 2.5 wt% sodium concentration of 30 mg mL 1 at 353 K and 30 bar O2 after 26 h
methoxide at 353 K and 45 bar CO); and (2) the subsequent (Table 2, entries 1–4). Fu and co-workers obtained similar results

Fig. 2 (a) Overall stoichiometric reaction for oxidation of glucose into formic acid; (b) oxidative conversion of carbohydrates (cellulose as an example)
into formic acid over H3+nPVnMo12 nO40 (HPA) catalysts. HPA-nox: the oxidized state of HPA; HPA-nred: the reduced state of HPA.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1353
View Article Online

Review Article Chem Soc Rev

Table 2 Oxidation of carbohydrates into formic acida

Feed con. Catalyst


Entry Substrate (mg mL 1) Catalyst ratiob T (K) O2 (bar) Time (h) Con. (%) Yieldc (%) Ref.
1 Glucose 30 H5PV2Mo10O40 22.2 353 30 26 498 47 23
2 Xylose 30 H5PV2Mo10O40 26.7 353 30 26 498 54 23
3 Cellobiose 30 H5PV2Mo10O40 11.7 353 30 26 498 47 23
4 Sucrose 30 H5PV2Mo10O40 11.7 353 30 26 498 48 23
5 Glucose 25 H5PV2Mo10O40 20 373 20 3 100 47 24
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

6 Cellulose 30 H5PV2Mo10O40 24.7 353 30 26 n.d. 1 23


7 Cellulose 10 H5PV2Mo10O40 & HCl 20 443 10 9 100 34 24
8 Cellulose 27 H5PV2Mo10O40 & H2SO4 17.5 453 30 0.08 100 61 25
9d Cellulose 27 H5PV2Mo10O40 & TSA 22.2 363 30 24 n.d. 19 26
10 Xylan 27 H5PV2Mo10O40 & TSA 27.3 363 30 24 n.d. 53 26
11 Lignin 27 H5PV2Mo10O40 & TSA — 363 30 24 n.d. 31 26
12 Pomace 33 H5PV2Mo10O40 & TSA — 363 30 24 n.d. 54.6 27
13 Cane trash 33 H5PV2Mo10O40 & TSA — 363 30 24 n.d. 48.9 27
14 Fruit pulp 33 H5PV2Mo10O40 & TSA — 363 30 24 n.d. 45.3 27
15 Spruce chips 33 H5PV2Mo10O40 & TSA — 363 30 24 n.d. 34.6 27
16 Glucose 50 H8PV5Mo7O40 34.8 363 30 8 100 60 28
17 Cellulose 50 H8PV5Mo7O40 & TSA 38.7 363 30 24 100 28 28
18 Cellulose 10 H4PVMo11O40 12.3 453 20 3 100 60.9 29
19 Cellulose 10 H5PV2Mo10O40 12.3 453 20 3 100 44.9 29
20 Cellulose 10 H6PV3Mo9O40 12.3 453 20 3 100 42.0 29
21 Cellulose 67 H4PVMo11O40 40 453 1 1 100 35.5 31
22 Cellulose 67 [MIMPS]3HPMo11VO40 40 453 1 1 93.3 51.3 31
23 Cellulose 67 [BMIM]3HPMo11VO40 40 453 1 1 15.2 5.2 31
24 Glucose 16.7 NaVO3–H2SO4 (0.7 wt%) 3.1 433 30 1 100 68.2 33
25 Xylose 16.7 NaVO3–H2SO4 (0.7 wt%) 3.7 433 30 1 min 100 68.2 33
26 Xylan 16.7 NaVO3–H2SO4 (0.7 wt%) 4.2 433 30 0.5 100 63.5 33
27 Sucrose 83.3 NaVO3–H2SO4 (0.7 wt%) 1.6 433 30 5 min 100 64.2 33
28 Wheat straw 83.3 NaVO3–H2SO4 (0.7 wt%) — 433 30 5 min 100 47 34
29 Cellulose 16.7 NaVO3–H2SO4 (2.0 wt%) 3.8 433 30 10 min — 57.7 37
30 Cellulose 16.7 NaVO3–H2SO4 (2.0 wt%) 3.8 453 30 10 min 37.5 37
31 Cellulose 8.1 VOSO4 10 453 20 1 — 53 38
32 Inulin 8.1 VOSO4 10 453 20 1 — 39 38
33 Starch 8.1 VOSO4 10 453 20 1 — 46 38
34e Glucose 18 H8PV5Mo7O40 20 363 20 48 100 85 39
35 Glucose 18 H8PV5Mo7O40 20 363 20 48 100 53 39
36 Beech wood 16.3 H8PV5Mo7O40 & TSA — 363 20 48 — 61 39
a
All reactions were carried out in water, except for entry 34, in a batch reactor. b The catalyst ratio is the molar ratio of substrates to the catalysts. For
cellulose and xylan as the substrates, the catalyst ratio is calculated as the molar ratio of glucose or xylose to the catalysts. c The yield is reported as a
carbon yield. d TSA is the abbreviation of p-toluenesulfonic acid. e This reaction was performed in a biphasic system of water/1-hexanol (v : v = 1).

(a 47% yield of formic acid) using the same HPA-2 catalyst p-toluenesulfonic acid (TSA) as the additive to promote the
(Table 2, entry 5).24 In each case, formic acid was the only transformation of water-insoluble biomass into formic acid
observed product in the liquid solution, and CO2 was the only over an HPA-2 catalyst.26 Formic acid was produced in a yield
other side product. However, the catalytic oxidation treatment of 19% after 24 h at 363 K at a cellulose concentration of
of cellulose over HPA-2 only produced a 1% yield of formic acid 27 mg mL 1 (Table 2, entry 9). The combined use of HPA-2 and
(Table 2, entry 6).23 Fu and co-workers improved the formic acid TSA was also effective at transforming xylan and lignin to formic
yield to 34% from cellulose at a concentration of 10 mg mL 1 acid with yields up to 53% and 31% (Table 2, entries 10 and 11),
after 9 h by the combined use of HPA-2 and 0.01 M HCl (Table 2, respectively. Wasserscheid and co-workers further applied this
entry 7), due to the higher efficiency of cellulose hydrolysis by method for the oxidative transformation of lignocellulose and
HCl.24 The use of HPA-2 and 1.6 wt% H2SO4 greatly improved the algae feedstocks into formic acid.27 Formic acid was produced
efficiency of the transformation of cellulose into formic acid, with yields of 34.6–54.6% from pomace, cane trash, fruit
which produced formic acid in a much higher yield of 61% at pulp and spruce chips (Table 2, entries 12–15). Subsequently,
100% cellulose conversion with a high substrate concentration Wasserscheid and co-workers prepared an improved heteropoly
of 30 mg mL 1 after 5 min at 453 K (Table 2, entry 8).25 The pH of acid H8PV5Mo7O40 (HPA-5), which produced formic acid yields of
the reaction system with HPA-2 and 0.01 M HCl was 1.87, while it 60% and 28% at 363 K from glucose (8 h) and cellulose (24 h) at a
was 0.56 for the HPA-2 and 1.6 wt% H2SO4 catalytic system. At substrate concentration of 50 mg mL 1, respectively (Table 2,
the lower pH of 0.56, the oxidation potential and electron affinity entries 16 and 17).28 The formic acid yield from cellulose over
increased due to the formation of protonated HPA-2 and VO2+ HPA-5 was 1.5 times higher than with HPA-2 at a higher cellulose
species, favoring the reduction of the catalyst and oxidation of concentration (Table 2, entries 9 vs. 17).26,28 The higher activity
the substrate, as well as the hydrolysis of cellulose into glucose. of the HPA-5 catalyst over the HPA-2 catalyst was due to the
Besides mineral acids, Wasserscheid and co-workers used 0.1 M presence of a much higher ratio of pervanadyl species (VO2+) for

1354 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

the reaction by the use of the HPA-5 catalyst. VO2+ has a higher a slight decrease of formic acid yield (51.3% for the first run vs.
redox potential (0.87 V vs. normal hydrogen electrode) than 49.5% for the third run).
HPA-2 (0.68 V vs. normal hydrogen electrode) at pH = 1, and In addition to the vanadium-containing heteropolyacids,23–29
hence HPA-5 has a higher catalytic activity. Formic acid was the vanadium salts have also been studied for the oxidative trans-
only product in the liquid solutions for all of the above cases, formation of carbohydrates into formic acid under the strong
and the only byproduct was CO2. acidic conditions.33,34 Sodium metavanadate (NaVO3) with H2SO4
In contrast to the results reported by Wasserscheid and has been reported to have catalytic activity for oxidation of
co-workers,28 Han and co-workers reported that the formic acid carbohydrates into formic acid.33,34 As shown in Fig. 3, cellu-
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

yield from cellulose increased with the decrease of vanadium lose conversion and formic acid selectivity increased on
content in the H3+nPVnMo12 nO40 catalysts at 453 K.29 The oxida- increasing the weight percentage of H2SO4 from 0.1 wt% to
tive conversion of cellulose over H3+nPVnMo12 nO40 catalysts at 0.7 wt%, which was due to the enhanced hydrolysis of cellulose
453 K did not require additional acid catalysts,29 while it required at higher weight percentage of H2SO4, and then kept almost
TSA to catalyze cellulose hydrolysis at 363 K.26,28 Under the stable beyond 0.7 wt%. The reaction was almost stopped
reaction conditions (453 K and 20 bar O2), cellulose was fully (cellulose conversion = 3%) without H2SO4. The highest formic
converted after 3 h, and the formic acid yields from cellulose acid yield (64.9%, Fig. 3) was close to that from glucose (68.2%,
followed the order: H4PVMo11O40 (60.9%) 4 H5PV2Mo10O40 Table 2, entry 24), suggesting that the first step of the hydrolysis of
(44.9%) 4 H6PV3Mo9O40 (42.0%) (Table 2, entries 18–20). As cellulose into glucose was highly efficient. This catalytic system
cellulose was completely converted after 3 h at 453 K, the yield was also effective at transforming xylose, xylan and sucrose into
of formic acid over H3+nPVnMo12 nO40 catalysts depended on the formic acid with yields up to 63% (Table 2, entries 25–27),33 and
selectivity of formic acid. The higher selectivity of H4PVMo11O40 even wheat straw (Table 2, entry 28).34 Similar to the vanadium-
with a single V atom than those of H3+nPVnMo12 nO40 (n = 2, 3) containing heteropolyacid catalytic systems,23–29 CO2 was also the
was possibly due to the synergetic effect of the multiple V atoms in sole byproduct over the NaVO3–H2SO4 catalytic system.33,34 The
H3+nPVnMo12 nO40 (n = 2, 3) causing over-oxidation, as implied selectivity of formic acid was dependent on the type of V–O
by their higher yields of CO2. Besides formic acid, acetic acid species in the reaction solution. As shown in Fig. 4, V4O124 is
was also obtained with yields around 5% for the three catalysts.29 predominant at pH around 7. V10O286 appears at pH o 6 and
The decomposition of formic acid can produce CO and CO2 at the becomes dominant in a pH range of 4.1–3.0. VO2+ appears at
temperature of 410 K. CO2 was the gas byproduct, which was pH o 2 and becomes dominant at pH o 1.35 Among them,
primarily formed from the intermediates and not formic acid, VO2+ showed the highest oxidative potential for selective oxida-
similar to what was observed by other authors.23–28 tion reactions.36 In fact, 51V-NMR and X-ray photoelectron
The H+ in the heteropolyacids can be exchanged with other spectra (XPS) analysis confirmed that VO2+ was the only V–O
metal cations or organic group.30 Some modified POM catalysts species by the use of 0.7 wt% H2SO4 with the reaction solution
were prepared by the exchange of the H+ in POM catalysts with pH lower than 1. Thus, it is proposed that in the catalytic cycle
organic sulfonic acid groups and demonstrated good catalytic VO2+ oxidizes carbohydrates by donating an oxygen atom first
performance for the oxidative transformation of cellulose to generate formic acid and VO2+, and VO2+ is then oxidized
into formic acid.31,32 For example, Liu and co-workers studied by molecular oxygen to generate VO2+. Furthermore, Wu and
the oxidative conversion of cellulose into formic acid over
modified POM catalyst-[MIMPS]3HPMo11VO40, which was pre-
pared by the exchange of three H+ in H4PMo11VO40 with -butyl
sulfonic acid group functionalized imidazole cations.31 Under
the same reaction conditions (453 K, 1 h, 10 bar O2), the parent
H4PMo11VO40 catalyst produced formic acid with a yield 35.5%
at a cellulose concentration of 67 mg mL 1, while the formic
acid yield increased to 51.3% over the [MIMPS]3HPMo11VO40
catalyst (Table 2, entries 21 and 22). Control experiments
indicated that the cations of the [MIMPS]3HPMo11VO40 catalyst
served two important roles. On the one hand, the sulfonic
acid group in [MIMPS]3HPMo11VO40 promoted the hydrolysis
of cellulose. [BMIM]3HPMo11VO40 (1-methyl-3-butyl-imidazole
cation) without the sulfonic acid group produced much lower
cellulose conversion (Table 2, entries 22 vs. 23). On the other
hand, [MIMPS]3HPMo11VO40 suppressed the over oxidation
of the intermediates to CO2 during the oxidation of glucose
Fig. 3 The effect of H2SO4 on the conversion of cellulose into formic
into formic acid, as the formic acid yield from glucose over
acid. Reaction conditions: initial O2 pressure (30 bar), cellulose (100 mg),
[MIMPS]3HPMo11VO40 was also higher than that over H4PMo11VO40 NaVO3 (22 mg), H2O (6 g), 433 K. Conversion was calculated based on the
(54.7% vs. 44.8%) after 1 h at 453 K and 10 bar O2. Furthermore, amount of unconverted carbohydrates. Adapted from ref. 34 with permission
the [MIMPS]3HPMo11VO40 catalyst showed a high stability with from the Royal Society of Chemistry, copyright 2015.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1355
View Article Online

Review Article Chem Soc Rev

yield of 9.2% (Table 2, entries 29 vs. 30). The yield of levulinic


acid was almost the same (0.8%) at 453 K. Besides temperature,
the oxidation step to formic acid was accelerated by the increase
of oxygen pressure.
VOSO4 was also used for the oxidative transformation of
cellulose into formic acid in water without acid.38 Formic acid
was produced in a yield of 53% from cellulose after 1 h at 413 K
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

and 20 bar O2 (Table 2, entry 31). Compared with the NaVO3–


H2SO4 catalytic system,37 it required a longer reaction time to
reach the maximum formic acid yield at lower cellulose concen-
tration (Table 2, entries 29 vs. 31), as this catalytic system did
not use H2SO4 to promote cellulose hydrolysis. This catalytic
system was also effective for the transformation of other
carbohydrates such as inulin into formic acid (Table 2, entries
32 and 33). Mechanism studies revealed that fructose, from the
isomerization of glucose by VOSO4, underwent a retrol-aldol
reaction to give two C3 intermediates (glyceraldehydes and
Fig. 4 Pourbaix diagram of vanadium at 25 1C and an ionic strength of
1,3-dihydroxyacetone, Fig. 6), which were oxidized into formic
1 M. Reproduced from ref. 35 with permission from Elsevier, copyright acid and CO2. Under a nitrogen atmosphere, lactic acid was
2003. produced with a yield of 54%, which was formed by the
isomerization of C3 intermediates. ESR and NMR spectra of
VOSO4 revealed that the oxidation state of vanadium did not
co-workers studied the relationship between oxidation and change during the conversion of cellulose into lactic acid under
hydrolysis in the oxidative conversion of cellulose into formic a nitrogen atmosphere, suggesting that these steps in Fig. 6 for
acid in NaVO3–H2SO4 aqueous solution.37 They observed that the formation of lactic acid were catalyzed by VIV. On the other
there were mainly two hydrolysis and two oxidations steps hand, the oxidation of VIV to VV by O2 and the subsequent
(Fig. 5): initial hydrolysis of cellulose to glucose and deep reduction of VV to VIV by reactants were observed, indicating
hydrolysis of glucose to levulinic acid; the oxidation of mono- that the redox between VIV and VV played key roles in the
saccharides to formic acid and the oxidation of levulinic acid to oxidative C–C cleavage reactions to form formic acid.
acetic acid. The reaction temperature, oxygen pressure and the The above examples for the oxidative conversion of biomass
concentration of H2SO4 all played crucial roles in the control of into formic acid were performed in monophasic systems with a
the reaction pathways. The increasing rate of deep hydrolysis maximum formic acid yield below 70%. Albert and co-workers
was faster than that of catalytic oxidation, when the reaction used a biphasic system (water/organic solvents) to improve
temperature was increased. For example, formic acid was produced formic acid yield via the in situ extraction of formic acid.39
in a yield of 57.7% from cellulose at 433 K after 10 min, and The water/1-hexanol (v : v = 1) biphasic system produced formic
levulinic acid and acetic acid were in low yields of 1.4 and 0.9%, acid with a high yield of 85% from glucose over an HPA-5
respectively (Table 2, entry 29). However, the formic acid yield catalyst at 363 K after 48 h (Table 2, entry 34), while the
greatly decreased to 37.5% at 453 K with an increased acetic acid oxidation of glucose in water only produced a 53% yield of
formic acid under identical conditions (Table 2, entry 35).
Glucose conversions were 100% either in water or water/
1-hexanol. As formic acid is stable both in water and 1-hexanol,
the higher selectivity of formic acid in the biphasic system is not
caused by the improvement of the stability of formic acid. The
authors claimed that the extraction of formic acid by 1-hexanol
reduced the pH in the aqueous phase during the reaction process
to increase the formic acid selectivity. An additional experiment
confirmed this assumption. Formic acid selectivity from
glucose oxidation over the HPA-5 catalyst decreased sharply from
53% in water to 26% by the addition of 20 wt% formic acid
to the initial reaction solution. Using this biphasic reaction
protocol, raw lignocellulosic biomass (beech wood) could also
be successfully converted into formic acid with a yield of 61%
(Table 2, entry 36).
Fig. 5 Proposed reaction pathways for the transformation of cellulose
It is generally accepted that the oxidation of carbohydrates
into formic acid in NaVO3–H2SO4 catalytic systems. Adapted from ref. 37 into formic acid over vanadium-containing POM catalysts
with permission from the Royal Society of Chemistry, copyright 2003. follows an electron-oxygen transfer mechanism (the oxidation

1356 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fig. 6 Reaction pathways for the conversion of glucose into formic acid over a VOSO4 catalyst. Adapted from ref. 38 with permission from Wiley,
copyright 2014.

Table 3 The oxidation of oxygenated molecules to formic acid using a HPA-2 catalysta

Entry Substrate (g mL 1) Time (h) Temperature (K) Conversion (%) Formic acid yieldb
1 Formic acid (0.05) 6 373 9 —
2 Methanol (0.025) 6 373 0 0
3 Formaldehyde (0.025) 6 373 13 0
4 Ethylene glycol (0.01) 3 373 10 0
5 1,3-Propylene glycol (0.05) 2 398 0 0
6 1,2-Propylene glycol (0.05) 3 398 50 17
7c 2,5-Dihydroxy-1,4-dioxane (0.01) 3 373 90 56
8 Glyoxal (0.01) 3 373 100 65
9 Glyceraldehyde (0.01) 3 373 100 55
a
Reaction conditions: 5 mol% of the HPA-2 catalyst and 20 bar O2 were used. b The yield of formic acid was calculated based on 1 mol of glucose
producing 6 mol of formic acid. c The substrate was the glycolaldehyde dimer. Adapted from ref. 24.

of carbohydrates by high valence (+5) vanadium catalysts, and possible intermediates and the pinacol oxidation mechanism
the oxidation of reduced vanadium species to the original high proposed by Neumann and Khenkin,40,41 Fu and co-workers
valence state vanadium catalysts, Fig. 6). As shown in Table 3, proposed that a possible mechanism for glucose oxidation
Fu and co-workers tried to study the possible mechanism of involved electron and oxygen transfer processes catalyzed by the
glucose oxidation into formic acid by the oxidation of some HPA-2 catalyst (Fig. 7) and that the reaction intermediates might
intermediates.24 Only formic acid and the feed molecules were be aldehydes. These possible intermediates were also claimed
detected in the product solution, for all the reactants used in by Marsh and co-workers for the oxidation of carbohydrates
Table 3. Firstly, 5 wt% formic acid solution was treated over the into formic acid over the NaVO3–H2SO4 catalytic system.35
HPA-2 catalyst at 373 K and 20 bar O2, and only 10% of formic As mentioned in this section, vanadium catalysts with V5+
acid was decomposed. As listed in Table 2, entry 5, 48% of CO2 are proved to have high rates of electron transfer and moderate
was produced during the oxidation of glucose into formic acid activity for the oxidation of carbohydrates into formic acid.
under the same reaction conditions. Therefore, it suggested However, formic acid yields from the current reported methods
that CO2 was not mainly from the decomposition of formic acid are not above 70% (Table 2), which is still far away from the
but the intermediates (Table 3, entry 1). No formic acid was theoretical value (100%). Selection of the right catalyst is critical
detected when methanol and formaldehyde were employed as to achieve high formic acid yields. Since complete combustion
the substrates (Table 3, entries 2 and 3), indicating that the of biomass to water and CO2 is always the thermodynamically
reaction mechanism could not be the a-cleavage of glucose to
form methanol or formaldehyde, which would be subsequently
oxidized into formic acid. The conversion of ethylene glycol,
glycolaldehyde dimer, and glyoxal over the same reaction time
showed that the active center of the substrates should be the
aldehyde group (Table 3, entries 4–9). The oxidation of oxalic acid,
glycolic acid, and glyoxylic acid produced very low conversion of
formic acid (less than 5%), suggesting that formic acid could
not be generated from these carboxylic acids.
In the presence of oxygen, the valence of the V and Mo atoms
in the catalysts was the same as that of the fresh catalyst.
However, the peak position of the Mo catalysts remained
basically unchanged in the absence of oxygen, whereas the Fig. 7 Possible pathway for glucose oxidation catalyzed by H5PV2Mo10O40
change in the peak position of V indicated that the valence of V (HPA-2) through an electron transfer and oxygen transfer reaction mechanism.
changed from five to four. Combining the reaction results for Adapted from ref. 24 with permission from Wiley, copyright 2012.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1357
View Article Online

Review Article Chem Soc Rev

favored reaction in biomass oxidation with molecular oxygen,


the applied catalyst should promote fast oxidation of all carbon
in biomass to formic acid, while showing no activity or negli-
gible activity in the further decomposition of formic acid into
CO2 and H2.

2.2 Conversion of carbohydrates into formic acid with H2O2


as the oxidant Fig. 9 Hydrothermal conversion of glucose into formic acid using H2O2
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

as the oxidant. Adapted from ref. 43 with permission from Royal Society of
H2O2 can also be used for the hydrothermal oxidation of Chemistry, copyright 2008.
carbohydrates into formic acid (Fig. 8). In 2002, Calvo and
Vallejo reported the oxidative conversion of carbohydrates into
organic acids in water with H2O2.42 A formic acid yield of 38% H2O2 as the oxidant over metal catalysts has been reported to be
was obtained in a continuous flow reactor from glucose performed under mild conditions. Jung and co-workers per-
at 523 K by the addition of NaOH (55.6 wt% of glucose), and formed the oxidation of glucose into formic acid over a homo-
the glucose conversion was 77%. In addition, acetic acid and genous N-heterocyclic carbenes-amidate Pd(II) complex at room
glycolic acid were also detected in yields of 17% and 22%, temperature.45 Formic acid was obtained in a high yield of 91%
respectively. For cellulose transformation, the hydrothermal after 16 h by the use of a stoichiometric amount of H2O2 and 6
oxidation was carried out at a higher temperature of 673 K equiv. NaOH. Although formic acid was achieved in a high
with 0.02 wt% H2SO4 and 12 equiv. H2O2 in a continuous flow yield, it is difficult to recycle the homogeneous Pd(II) complex
reactor, producing formic acid in a low yield of 19.4% with catalyst. Ebitani and co-workers prepared a MgO-supported
5.8% yield of acetic acid. Jin and Co-workers first reported the copper catalyst (denoted as CuCTAB/MgO) for the oxidation
hydrothermal oxidation of glucose into formic acid in a batch of glucose into formic acid by H2O2, where cetyltrimethyl-
reactor with a yield of 75% in 1.25 M KOH reaction solution ammonium bromide (CTAB) was used as the capping agent for
with 14.4 equiv. mol H2O2 at 523 K.43 The formic acid yield Cu nanoparticles.46 Formic acid was produced in a yield of 65%
increased from 24 to 75% with increasing KOH concentration from glucose after 12 h with 8 equiv. H2O2 at 393 K. In addition,
from 0 to 1.25 M within 1 min. The same authors then glyceric acid and glycolic acid were also produced in yields of
performed the oxidation of monosaccharides (glucose, fructose 29% and 16.6%, respectively. The CuCTAB/MgO catalyst was
and galactose) and disaccharides (cellobiose and maltose) at a stable without any significant loss of activity for three runs.
low temperature of 423 K,44 producing high formic acid yields The one advantage of using H2O2 and O2 as oxidants is that
(80–85%). Compared with previous work,43 a longer reaction the reduction product of the oxidants (H2O2 and O2) is H2O. In
time of 15 min and higher NaOH concentration of 2.0–2.5 M contrast to the use of O2 as the oxidant, the hydrothermal
were required. The alkali played two roles in enhancing the oxidation of carbohydrates into formic acid with H2O2 was
production of formic acid. One was the inhibition of the formic performed under harsh reaction conditions such as at a high
acid decomposition. The other role of a base is that the base reaction temperature up to 673 K. In addition, it is also difficult
was favorable for the selective oxidation at C-1 for aldoses, to get a high purity formic acid because other organic acids are
which leads to the formation of formic acid via the rupture of produced as byproducts. Taking all of the above cases for the
the C1–C2 bond. In addition to formic acid, a few low molecular oxidative transformation of carbohydrates into formic acid into
weight carboxylic acids including lactic acid, propanoic acid, consideration, catalytic oxidation with O2 as the oxidant can
acrylic acid and acetic acid were also identified by HPLC. As overcome the drawbacks caused by the hydrothermal oxidation
shown in Fig. 9, formic acid was proposed to be formed by the method, e.g., the facile availability of O2, wide feedstock scope and
rupture of C1–C2 or C5–C6 (a-scission) and/or the rupture high purity of formic acid. Moreover, hydrothermal treatment can
of C2–C3 (b-scission) or C4–C5 (b-scission) from glucose. The achieve a highly efficient, rapid conversion and simplify the reaction
former directly yields formic acid, but the latter yields formic procedure, which is key for industrial application. Also, the
acid via oxalic acid. Oxidation of oxalic acid under the same selectively of formic acid is high, rather than low. For acetic acid
conditions only produced a low formic acid yield (14%). This production, although its selectively was not high, the acetic acid
indicates that formic acid is formed from the direct oxidation purity was very high, which is important for separation. Further-
of glucose by a-scission. more, the hydrothermal conversion can achieve a continuous-
In contrast to the hydrothermal oxidation methods, the system, which is also key for industrial application.
catalytic oxidation of glucose into formic acid by the use of

3. Conversion of carbohydrates into


acetic acid
Acetic acid is the second simplest carboxylic acid next to formic
acid.10 Its principal use (40%) is in the manufacture of vinyl
Fig. 8 Hydrothermal oxidation of glucose into formic acid by H2O2. acetate.10 Acetic acid can also be converted to ethyl acetate or

1358 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

other esters by Fisher esterification.47 Dehydration of acetic an inert atmosphere. The first hydrothermal step produced
acid can be used to make acetic anhydride, which is used as an lactic acid, HMF and 2-FA, in yields of approximately 16%, 9%
acetylating agent, in the production of cellulose acetate. In the and 6% from cellulose at 573 K for 2 min, respectively. The
food industry, acetic acid is used as an acidity regulator and second step involves the oxidation of these intermediates into
as a condiment. The global annual demand of acetic acid is acetic acid. As shown in Table 4 (entries 1–3), the acetic acid
around 6.5 million tones.10 Currently, acetic acid is mainly yields using the two-step method from rice hulls, cellulose, and
produced by methanol carbonylation over rhodium or iridium- starch were nearly double compared with those using the one-
complex catalysts.47 However, this synthetic process requires
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

step method. Besides acetic acid, other products were also


relatively high temperatures (B448 K) and pressures (30 bar), produced including formic acid, maleic acid, malonic acid, and
exotic materials of construction, and extensive safety-related propionic acid. This process was scaled up to a continuous flow
equipment. Acetic acid can also be produced by fermentation of reaction system with a maximum capacity of 2 kg dry starch h 1.50
ethanol.48 Biological sources of acetic acid are of interest, but The highest acetic acid yield (17.5% carbon) in the continuous
currently uncompetitive with acetic acid produced from metha- flow reaction system was the same as the batch reactor system
nol carbonylation. Thus, chemical oxidation of renewable bio- (Table 4, entries 3 vs. 4).
mass would pave a new and effective way for the industrial The acetic acid yield can be increased by the addition of
production of acetic acid. mineral acids in the first step.51 In the presence of acid, levulinic
The oxidation of biomass into acetic acid typically occurs in acid is the main intermediate in the first step in Fig. 10b, which
two reactive steps: transformation of the carbohydrates into is formed via the dehydration of carbohydrates into HMF and
intermediates, and the oxidation of the intermediates into acetic the subsequent rehydration of HMF. Levulinic acid was then
acid. H2O2 is often used as an oxidant for the second step. The converted to acetic acid via hydrothermal oxidation. Under
intermediates include HMF, 2-furaldehyde (2-FA), lactic acid and optimal conditions (523 K, pH = 0.5, and a reaction time of
levulinic acid according to the reaction conditions. 300 s for the first step; 523 K, H2O2 = 12 equiv., and a reaction
Jin’s group has performed many interesting studies on the time of 60 s for the second step), the yields of acetic acid were as
conversion of carbohydrates into acetic acid. A two-step method high as 23% and 26% for glucose and fructose, respectively
was developed by Jin and co-workers to improve the yield of (Table 4, entries 5 and 6). The solution pH of the first step was
acetic acid.49 As depicted in Fig. 10a,49 carbohydrates are first important for the total yield of acetic acid. The yield of acetic
hydrothermally converted into HMF, 2-FA and lactic acid under acid was only approximately 10% for both glucose and fructose

Fig. 10 Proposed two-step processes for producing acetic acid via different intermediates.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1359
View Article Online

Review Article Chem Soc Rev

Table 4 Acetic acid production by oxidation of biomass

Reaction conditions
Substrate First step Second step Acetic acid yielda Purityb Ref.
c c
1 Rice hulls 573 K, 2 min 573 K, 1 min, H2O2 = 8.4 equiv. of glucose unit 21.7 (11.7) 75.5 (33.4) 44
2 Cellulose 573 K, 2 min 573 K, 1 min, H2O2 = 8.4 equiv. of glucose unit 16.3 (9.0)c 68.5 (26.1)c 44
3 Starch 573 K, 1 min 573 K, 1 min, H2O2 = 8.4 equiv. of glucose unit 17.5 (9.6)c 70.0 (28.2)c 44
4d Starch 573 K, 1 min 573 K, 1 min, H2O2 = 8.4 equiv. of glucose unit 17.5 — 45
47c
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

5 Glucose 523 K, 5 min, pH = 0.5 523 K, 1 min, H2O2 = 12 equiv. of glucose unit 23 46
6 Fructose 523 K, 5 min, pH = 0.5 523 K, 1 min, H2O2 = 12 equiv. of fructose 26 52c 46
7 Corn stover 1 M H2SO4, 463 K, 60 min 463 K, 540 min, 2.0 MPa O2 21.3 84.6 47
8 Glucose 573 K, 1 min, 0.64 M Ca(OH)2 573 K, 3 min, H2O2 = 8.4 equiv. of glucose unit 27 — 49
9 Glucose 548 K, 20 s, glucose/bentonite wt% = 1 548 K, 80 s, in air 27 — 54
a
Acetic acid yield was calculated based on carbon of the feedstock. b The purity of acetic acid is defined as the percentage of organic carbon of
acetic acid divided by the total organic carbon (TOC) of the liquid sample after the reaction. c The data are the results using the one-step method.
d
Results obtained by the continuous reactor. The flow rate of H2O2 was 6.5 L h 1.

without the control of acid. However, too strong acidic solution The discussed three kinds of two-step methods for the conver-
was also not suitable for the production of acetic acid. The sion of carbohydrates into acetic acid demonstrate that acetic
acetic acid yield from fructose decreased from 26% at pH = 0.5 acid was produced by the oxidation of the intermediates such
to 17% at pH = 0.24. Later, Fu and co-workers used this two-step as lactic acid, levulinic acid and furans. The transformation of
process for the transformation lignocellulosic biomass into lactic acid into acetic acid must cleave one carbon, thus the
acetic acid by the use of O2 in the second step.52 Acetic acid theoretical carbon yield of acetic acid should be 66.7%. The
was attained in a yield of 21.3% under the optimal conditions acetic acid yield from lactic acid was reported to be 40% by Jin
with a purity up to 84.6% of the total liquid product (Table 4, and co-workers.49 The relatively low yield of acetic acid also
entry 7). Other lignocellulosic feedstocks such as corncob, bagasse, accompanies another troublesome problem that the purity was
bamboo, poplar and pine could also be transformed into acetic not high because of the byproducts. According to the above
acid with yields of 15–20%. analysis, the current methods for the production of acetic acid
Jin and co-workers discovered that oxidation of lactic acid from biomass by catalytic oxidation only produce low yields
produced much more acetic acid (about 40%) than furans of acetic acid. Catalysts may be able to tune this selectively
(about 15% from HMF and 20% from 2-FA).49 A base promoted to produce acetic acid with higher selectivity under milder
the hydrothermal transformation of biomass into lactic acid.53 conditions.
Thus alkaline conditions were used to produce acetic acid from
glucose in 27% yield (Fig. 10c and Table 4, entry 8). Later, they
used bentonite as the solid base for the two-step transforma- 4. Conversion of carbohydrates into
tion of glucose into acetic acid.54 The second step just exposed glycolic acid
the reactor to the air for 3 min after the first step. This method
produced acetic acid with a yield of 27% at 548 K with 20 s for Glycolic acid is another important two-carbon carboxylic acid
the first step and 90 s for the second step (Table 4, entry 9). with a hydroxyl group, which is the smallest a-hydroxy acid. It
Being different from the above two-step methods, Yan and has been widely used in organic synthesis, biodegradable
co-workers performed one-step transformation of chitin and polymer synthesis, skin-care products, industrial rust removal,
waste shrimp shells into acetic acid in 2 M NaOH by the use of and food processing.11 The price for the glycolic acid market in
CuO with the molar ratio to carbon = 0.125.55 Acetic acid was 2011 was $2332 per metric ton, with an expected increase to
produced in a 32.3% yield from chitin at 573 K and 5 bar O2 $5375 per metric ton in 2018.56 The US annual market size of
after 35 min. The real oxidant was CuO, which was reduced to glycolic acid was 15 000 per ton in 2011.56 Glycolic acid is mainly
Cu2O during the reaction process. The as-formed Cu2O can be produced from the reaction of formaldehyde with synthesis gas
reoxidized to CuO by O2. The use of untreated tiger shrimp (carbonylation of formaldehyde). It can also be prepared by the
shell powder, consisting of 24.4% chitin, 17.5% protein, and reaction of chloroacetic acid with sodium hydroxide followed by
43.2% calcium carbonate (CaCO3), also produced an acetic acid re-acidification. The net reaction can be summarized as follows:
yield of 47.9% under the same conditions. The higher acetic ClCH2COOH + NaOH - HOCH2COOH + NaCl. Glycolic acid can
acid yield from raw shrimp shells over purified chitin suggested also be produced from cellulose as shown in Fig. 11.
that some carbon in proteins may contribute to acetic acid So far, the oxidative conversion of carbohydrates to glycolic
formation. Glucosamine was firstly formed by hydrolysis from acid has been less studied.57–59 Jin and co-workers performed
chitin, which then underwent deamination to produce glucose. the hydrothermal conversion of cellulose into glycolic acid using
Lactic acid was the main intermediate for the conversion of commercial CuO as the oxidant under alkaline conditions in a bath
glucose into acetic acid. reactor. They achieved a glycolic acid yield of 14.9% with acetic
There is no need for an oxidant for the stoichiometric acid (5.2% yield), formic acid (7.1% yield) and lactic acid (13.8%
conversion of glucose into acetic acid (C6H12O6 - 3CH3COOH). yield) as by products (reaction conditions: 573 K for 300 s with

1360 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

Fig. 11 Oxidative conversion of carbohydrates (cellulose as an example) into glycolic acid.


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

0.035 g of cellulose in 2 mL NaOH (1 M) and 2 mmol CuO).57 and H4SiMo12O40 were effective for the oxidation of these
Later Han and co-workers studied the oxidative conversion intermediates into glycolic acid.
of cellulose over four kinds of heteropoly acids at 453 K and The photocatalytic oxidation method was also reported for
6 bar O2.58 H3PMo12O40 and H4SiMo12O40 produced glycolic the conversion of cellulose into glycolic acid.59 This method
acid with yields of 49.3% and 46.5%, respectively, and other used a UV-Fenton reaction where strong oxidative species such
products were formic acid, acetic acid, levulinic acid, glucose as hydroxyl radicals ( OH) are produced by H2O2 catalyzed by
and fructose. Whereas, Mo-free heteropoly acids including FeCl2 under UV irradiation. Glycolic acid was produced in a very
H3PW12O40 and H3SiW12O40 produced only a small amount of low yield of 4%, and the major products were formic acid (31%)
glycolic acid in yields of less than 6%. These results suggested and CO2 (54%). The addition of ZnCl2 increased the glycolic
that the reaction pathway is determined by the type of metal acid yield to 33%, and decreased the formic acid and CO2 yields
atom in the heteropoly acids, and that Mo favors a selective to 14% and 27%, respectively. The authors proposed that the
oxidation of cellulose into glycolic acid. Han and co-workers Zn2+ enhanced the glycolic acid yield via complexation with
proposed a reaction pathway for the oxidation of cellulose into glyconic acid, thus preventing the further oxidative decomposi-
glycol acid (Fig. 12). Cellulose is firstly hydrolyzed to glucose, tion of glyconic acid.
which then undergoes successive retro-aldol condensations
to produce glycolaldehyde. In parallel, the isomerization of
glucose gives rise to fructose, which can also be converted by 5. Conversion of carbohydrates into
retro-aldol reactions through dihydroxyacetone and glyceralde- malonic acid and oxalic acid
hydes to glycolaldehyde and formaldehyde. The oxidation of
glycolaldehyde and formaldehyde produces glycolic acid and The oxidative conversion of carbohydrates can also generate two
formic acid, respectively. As glyceraldehyde and glycolaldehyde other dicarboxylic acids of malonic acid (propanedioic acid)
are also intermediates for the oxidation of glucose into formic and oxalic acid. Malonic acid with a structure of CH2(COOH)2 is
acid under similar reaction conditions by Fu and co-workers,24 a bulk chemical molecule for the preparation of malonates,
the difference of the oxidation products should be mainly barbituric acid, and meldrum’s acid, and it is also a feedstock to
due to the use of different catalysts. Vanadium-containing prepare 1,3-propanediol, which is one of the chemical precursors
heteropoly acids such as H5PV2Mo10O40 were effective for the of polytrimethylene terephthalate (PTT), an alternative to poly-
oxidation of these intermediates into formic acid, while ethylene terephthalate (PET) in the future.60 Malonic acid is
molybdenum-containing heteropoly acids such as H3PMo12O40 mainly prepared by the fermentation of glucose or starch.61

Fig. 12 Proposed pathways for the conversion of cellulose to glycolic acid. Adapted from ref. 58 with permission from American Chemical Society,
copyright 2012.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1361
View Article Online

Review Article Chem Soc Rev


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fig. 13 Reaction pathways for the conversion of cellulose to malonic acid. Adapted from ref. 62 with permission from American Chemical Society,
copyright 2014.

Yang and Qi reported a hydrothermal method for the oxidative As shown in Fig. 14, there are two main degradation pathways
transformation of cellulose into malonic acid at 573 K with a for glucose, namely C3–C4 scission, and C2–C3 scission.65
high concentration of NaOH.62 Cellulose was firstly dissolved in Glyceraldehyde, an important intermediate formed by the
an aqueous solution of NaOH and urea. Then the hydrothermal C3–C4 scission of glucose, has a good selectivity toward lactic
treatment of the 4 mg mL 1 cellulose solution at 573 K produced acid in alkali solution. Glycolaldehyde as another intermediate
malonic acid in a yield of 39.7% after 1 h at a NaOH concentration via C2–C3 scission could be oxidized to glycolic acid, which was
of 100 mg mL 1. At the same time, lactic acid, acetic acid and further oxidized to oxalic acid.
formic acid were produced in yields of 9.5%, 7.9% and 6.7%,
respectively. It was proposed that cellulose was firstly hydrolyzed to
glucose, which was subsequently degraded into dihydroxyacetone 6. Conversion of carbohydrates into
and glyceraldehyde, via isomerization and retro-aldol reactions. gluconic acid and glucaric acid
In the presence of oxygen, dihydroxyacetone was transformed
to malonic acid (HOOC–CH2–COOH) by a series of oxidation Gluconic acid is one of the oxidation products of glucose
and dehydration steps (Fig. 13). Meanwhile, lactic acid was without the cleavage of the C6 chain.66 Gluconic acid and its
formed from glyceraldehyde through consecutive dehydration salts are biocompatible and biodegradable chemicals, which
and 1,2-hydride shift reactions. have been widely used as a food additive, an ingredient for
Oxalic acid (OA) is an industrially important chemical with concrete and a raw material for medicines and biodegradable
various applications, such as the production of celluloid and polymers.13 The current annual worldwide production capacity
rayon, leather manufacture and dressing, and extraction of rare for gluconic acid and its derivatives is estimated to be about
earths from monazite.63 Oxalic acid was early manufactured by 60 000 tones per year.13 Current commercial production of
the oxidation of carbohydrates using nitric acid as the oxidant.63 sodium gluconate uses submerged fermentation with A. niger
Obviously, the use of molecular oxygen for the synthesis of oxalic and is based on the modified process developed by Blom et al.67
acid is environmentally-friendly. Han and co-workers reported a It involves fed-batch cultivation with intermittent glucose
new method for the oxidative conversion of cellulose into oxalic feedings and the use of sodium hydroxide as a neutralising
acid in 2 M NaOH solution catalyzed by CuO.64 Oxalic acid was agent. pH is held at 6.0–6.5 and the temperature at about 307 K.
produced in a yield of 41.5% with a catalytic amount of CuO at The productivity of this process is very high, since glucose
474 K and 3 bar O2. Formic acid, acetic acid, and lactic acid were is converted at a rate of 15 g (L h) 1. The biochemical method
also produced with yields of 20.1%, 6.6% and 3.0%, respectively. (fermentation) is competitive with chemical techniques

Fig. 14 Reaction pathway for the conversion of cellulose to oxalic acid. Adapted from ref. 64 with permission from American Chemical Society,
copyright 2016.

1362 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

(such as catalytic methods, and electrolytic methods), and is yield of 99.4% was obtained at a constant pH = 9 with air as the
at present the method of choice by industry. Although the oxidant after 2.6 h over the Pd–Bi/C catalyst (Table 5, entry 1).
biochemical method is used as the industrial method for the Whereas, a Pd/C catalyst produced gluconic acid with a 78.1%
production of gluconic acid, this biochemical process still has a yield at a glucose conversion of 82.6% after 24 h (Table 5,
few drawbacks such as the difficulties in using moulds and free entry 2). Bismuth (Bi) was proposed to prevent the oxygen
enzyme separation, waste-water removal, and dead microbes poisoning of the Pd surface by acting as a co-catalyst in the
and accumulated excretory substances of microbes that should oxidative dehydration mechanism (Fig. 16), and the Pd–Bi/C
be disposed of.68 In addition, the product of the biochemical
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

catalyst could be reused for five runs without any decrease of


process is sodium gluconate, as the fermentation process is gluconic acid yield. A similar phenomenon was also observed by
carried out at a relative high pH at 6.0–6.5. Recent developments Karski and co-workers, when they used Pd–Bi/C and Pd–TI/SiO2
have shown the potential of chemical catalytic procedures for catalysts for the oxidation of glucose into gluconic acid.72
oxidizing glucose with oxygen or air.69,70 The economic viability In contrast to the Pd–Bi/C and Pd–TI/SiO2 catalysts, Crossley
of the chemical catalytic processes depend largely on the and co-workers reported that the commercial Pd/C catalyst
activity, selectivity, lifetime and cost of the catalyst as well as (Sigma-Aldrich) was stable for the oxidation of glucose, which
on the measures required for product purification and the was obtained from the fast pyrolysis of bio-oils, followed by
energy demand. hydrolysis.73 Crossley obtained a gluconic acid yield of 83% after
24 h at 323 K at and a constant pH = 9.3 (Table 5, entry 3).73 After
6.1 Oxidation of glucose into gluconic acid with O2 as the the first run, fresh glucose was then added, and the yield of
oxidant gluconic acid after 3 h was the same as the first run.
Most researchers used supported noble metal catalysts (Pt, Pd Besides the addition of a second metal to Pd catalysts, the
and Au) for the oxidation of glucose into gluconic acid with O2 interaction between metal nanoparticles and a support also
as the oxidant (Fig. 15). A base is often required in this process. shows some effect on the catalyst activity and stability. Liu and
In 1995, Besson and co-workers studied the aerobic oxidation of co-workers studied two kinds of Pd/g-Al2O3 catalysts for glucose
glucose over the Pd/C and Pd–Bi/C catalysts.71 A gluconic acid oxidation. The 3.14% Pd/g-Al2O3-(P) catalyst was prepared by

Fig. 15 Catalytic oxidation of glucose into gluconic acid over metal catalysts.

Table 5 Catalytic oxidation of glucose into gluconic acid/salt with O2 as the oxidanta

Entry Catalyst Feed con. (M) Catalyst ratiob T (K) pHc Oxidant Time (h) Con. (%) Yield (%) Ref.
1
1 Pd–Bi/C 1.66 787 313 9 1.5 L min (1 bar air) 2.6 99.6 99.4 71
2 Pd/C 1.66 787 313 9 1.5 L min 1 (1 bar air) 24 82.6 78.1 71
3 Pd/C 0.02 — 323 9.3 60 mL min 1 (1 bar air) 24 83 83 73
4 Pd/g-Al2O3-(P) 0.4 139 323 9 400 mL min 1 (1 bar O2) 6 100 95 74
5 Pd/g-Al2O3-(C) 0.4 139 323 9 400 mL min 1 (1 bar O2) 8 95 90.3 74
6 Pd/HPS — — 353 6.0–7.5 100 mL min 1 (1 bar O2) 20 93.6 93.2 75
7d Pt/C 0.167 — 413 — 50 bar air 1 63 46 77
8d Pt/C 0.167 — 473 — 50 bar air 1 94 22.6 77
9 Pt/HT 0.13 68 323 — 1 bar O2 12 99 83 78
10 Au/C 0.22 1000 323 7 3 bar O2 2.5 100 99 85
11 Au/C 0.22 1000 323 9.5 3 bar O2 0.5 100 99 85
12 Au/C 0.22 1000 373 — 3 bar O2 6 100 99 85
13 Au/CMK-3 0.1 1000 383 — 3 bar O2 2 92 80.5 88
14 Au/TiO2 0.1 4378 313 11 500 mL min 1 (1 bar O2) 2 100 98 89
15 Au/TiO2-SIM 0.06 438 433 — 3 bar O2 1 71.1 67.3 90
16 Au/TiO2-CIM 0.06 438 433 — 3 bar O2 1 30.3 20.0 90
17 Au/TiO2-DP 0.06 438 433 — 3 bar O2 1 61.7 58.0 90
18 Au/TiO2-DP 0.167 140 338 — 2.3 bar O2 2 91 90 91
19 Au/ ZrO2-DP 0.167 140 338 — 2.3 bar O2 2 89 88 91
20 Au/CeO2-DP 0.167 140 338 — 2.3 bar O2 2 89 87 91
21 Au–Pd/MgO 0.5 1225 333 — 1 bar air 24 62 100 114
a
All of the reactions were carried out in water in a batch reactor, except for entries 7 and 8. b The catalyst ratio = the molar ratio of glucose to the
noble metal. c Some reactions were performed without pH control if there was no data in this column. d The reaction was performed in a
continuous flow reactor.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1363
View Article Online

Review Article Chem Soc Rev


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fig. 16 Glucose oxidation on the Pd–Bi/C catalyst. Adapted from ref. 71 with permission from Elsevier, copyright 1995.

the argon glow discharge plasma approach, including the over the Pt/C catalyst in a flow reactor without pH control at
impregnation of g-Al2O3 with PdCl2 and the subsequent treatment higher temperatures (413–473 K). Glucose conversion increased
with argon glow discharge plasma.74 The 3.14% Pd/g-Al2O3-(C) from 63% at 413 K to 94% at 473 K after 1 h, but gluconic acid
catalyst was also prepared by the conventional impregnation of selectivity decreased from 73% at 413 K to 24% at 473 K
g-Al2O3 with PdCl2 followed by calcination in air, and subsequent (Table 5, entries 7 vs. 8).77 The oxidation of glucose over the
reduction with H2. Both catalysts gave the same gluconic acid Pt/C catalyst without base required high reaction temperatures
selectivity (95%), while the Pd/g-Al2O3-(P) catalyst had higher (413–473 K),77 while it could be carried out at a lower tempera-
activity (Table 5, entries 4 vs. 5). Pd/g-Al2O3-(P) also showed ture of 323 K at pH = 9.76 Besides the addition of a base to
higher stability than Pd/g-Al2O3-(C), albeit both of them were control a constant pH, a Pt catalyst with a basic support can
not stable. The amount of Pd leaching from the Pd/g-Al2O3-(P) also promote glucose oxidation at low temperatures.78 A Pt/HT
catalyst was 4.4% in the first run, while it was 17.8% for the catalyst was prepared by adding Pt to hydrotalcite {HT, with a
Pd/g-Al2O3-(C) catalyst. The higher activity and stability of general formula of Mg6Al2CO3(OH)164(H2O)} and produced a
Pd/g-Al2O3-(P) over Pd/g-Al2O3-(C) was attributed to the stronger 99% conversion of glucose and a 83% yield of gluconic acid
interaction between Pd nanoparticles and the support. Diffuse after 12 h at 323 K (Table 5, entry 9).78 Fructose was also
reflectance Fourier transform infrared (DRIFT) spectra of CO detected in a yield of 8% by the isomerization of glucose, which
adsorption indicated that a lower percentage of CO was absorbed was catalyzed by the base support of HT. In addition, other small
on the surface of Pd/g-Al2O3-(P) than that on the surface of the molecular weight organic acids such as oxalic acid, malonic acid,
Pd/g-Al2O3-(C) catalyst. The lower amount of CO adsorbed over and lactic acid were also observed in minor amounts (4–5%).
the Pd/g-Al2O3-(P) catalyst was due to the stronger interaction Unlike the Pt/C catalyst,76 the Pt/HT catalyst did not suffer
between Pd nanoparticles and g-Al2O3. Another example to poisoning albeit 10% of the catalyst activity was lost in each
enhance the interaction between Pd nanoparticles and a support run, which was caused by the leaching of Pt. There was less
involved the deposition of Pd nanoparticles onto hypercrosslinked research on using Au catalysts for glucose oxidation than Pd and
polystyrene (HPS) (Pd/HPS).75 HPS contains micropores, meso- Pt catalysts for a long time. Since the discovery of the activity of
pores and macropores, which have strong interaction with Pd Au nanoclusters by Haruta in 2003,79 Au catalysts have received
nanoparticles. The glucose conversion and gluconic acid yield great attention in chemical reactions.80–82 Recently, supported
were 93.6% and 93.2% after 20 h at 353 K (Table 5, entry 6), Au catalysts were found to have catalytic activity towards the
respectively. After five runs, the glucose conversion (93.4%) was oxidation of glucose into gluconic acid.83,84 The particle size of
almost the same as the first run (93.6%), and the gluconic acid Au nanoparticles and the nature of the support were observed to
selectivity was also the same. have an important impact on the catalytic activity, selectivity and
Compared with Pd catalysts, Pt catalysts have also been stability.83,84 Supported Au catalysts mainly include Au/C,85–88
studied for the aerobic oxidation of glucose, which generally Au/TiO2,89–91 Au/g-Al2O3,92–99 Au/CeO2,91,100 Au/ZrO2,91,100,101
deactivate and have lower gluconic acid selectivity. In 1981, Au/SiO2102,103 and polymer supported Au catalysts.104,105 Biella
Kogelbauer and co-workers studied the oxidation of glucose in and co-workers studied a carbon supported Au catalyst (Au/C)
a weakly alkaline medium with a Pt/C catalyst.76 The authors for the oxidation of glucose at pH values from 7.0 to 9.5 and
proposed that the surface of the active Pt was initially covered without pH control (Table 5, entries 10–12).85 The gluconic acid
with chemisorbed oxygen (Pt–O), and glucose first reacted with yield at different pH (7.0 and 9.5) was as high as 99% at 323 K,
chemisorbed oxygen (Pt–O) to yield gluconic acid. Catalyst but a higher reaction rate was observed at a higher pH (Table 5,
deactivation occurs in this process, due to the formation of entries 10 and 11). The Au/C catalyst was also active without pH
platinum oxide (PtO2). Besides gluconic acid, uronic acid and control, but a higher reaction temperature of 373 K and a
glucaric acid were also detected in the reaction solution, and longer reaction time of 6 h were required to get similar results
the selectivity of gluconic acid was only 70% at 323 K at pH = 9. (Table 5, entry 12). The Au/C catalyst deactivated due to the
Later, Koklin and co-workers performed the oxidation of glucose leaching and sintering of Au nanoparticles at the pH below 9.5.

1364 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

There was no leaching of Au at pH = 9.5, but the size of the Au (CIM), deposition–precipitation (DP) and sol-immobilization
nanoparticles increased from 4.4 nm in the fresh catalyst to methods (SIM) were used for the preparation of the Au/TiO2-
5.9 nm after four runs. At pH = 7, 18% of Au was leached into CIM, Au/TiO2-DP and Au/TiO2-SIM catalysts, respectively. The
the solution after 4 runs and the Au size was 5.4 nm. At pH = 7, as-prepared catalysts were further calcined in static air at 250 1C
gluconic acid can chelate with Au nanoparticles, causing the for 3 h. The Au/TiO2-SIM catalyst showed better catalytic
leaching of Au from the Au/C catalyst. Later, ordered meso- performance than the other two kinds of catalysts at 433 K
porous carbon (OMC) supported Au catalysts were used for (Table 5, entries 15–17). The Au/TiO2-SIM catalyst not only had
the oxidation of glucose into gluconic acid.86,88 The Au/OMC
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

a small size of Au nanoparticles, but also showed the highest


catalyst with an average mesopore size of 5.4 nm and an Au dispersion of Au nanoparticles. Therefore, the Au/TiO2-SIM
nanoparticle size of 3.3 nm gave a high TOF of 4.308 mol catalyst showed the best catalytic activity. Again, the base-free
glucose molAu 1 s 1 at 313 K and pH = 9 with an O2 flow rate of oxidation of glucose over the Au/TiO2 catalyst was performed at
100 mL min 1 under atmospheric pressure,86 much higher than higher temperature than that with a constant pH = 11.89
the Au/C catalyst with a TOF of 0.023 mol glucose molAu 1 s 1 Román-Leshkov and co-workers compared the activities of the
at pH = 11 with the same temperature and O2 flow rate.87 The three kinds of metal oxide (TiO2, ZrO2, CeO2) supported Au
superior catalyst activity of the Au/OMC catalyst was ascribed to catalysts for glucose oxidation without pH control, which were
its mesopore structure that could not be blocked by Au nano- prepared using a deposition–precipitation (DP) method.91 The
particles, and the remaining space was enough for molecule three kinds of Au catalysts produced glucose conversion around
diffusion during the reaction process. Besides, abundant O 90% and gluconic acid selectivity up to 99% after 2 h at 338 K
and O2 species on the surface of the Au/OMC catalyst also and 2.3 bar O2 (Table 5, entries 18–20).91 The authors found
promoted the oxidation of glucose into gluconic acid. Later, that the deactivation of the Au catalysts was caused by the
Yuan and co-workers prepared another ordered mesoporous leaching and hydrothermal sintering of Au nanoparticles, as well
carbon (CMK-3)-supported Au catalyst (Au/CMK-3) for the base- as by the adsorption of reactive species. Lowering the surface
free oxidation of glucose.88 Au nanoparticles were uniformly density of Au nanoparticles on metal oxides decreased the
dispersed on the surface and in the channels of CMK-3 with an sintering rate of Au nanoparticles and enhanced the stability
average size of 3 nm. Glucose conversion reached 92% with a and activity of the catalysts. Alumina supported Au catalysts
gluconic acid selectivity of 87.5% after 2 h at 383 K and 3 bar O2 (Au/g-Al2O3) prepared using different methods were reported to
(Table 5, entry 13). Similar to the base-free oxidation of glucose show high catalytic activity and long-term stability towards the
over the Au/C catalyst,85 the oxidation of glucose over the oxidation of glucose in a continuous flow reactor.92–95,98,99 For
Au/CMK-3 catalyst without pH control was also performed at example, the Au/g-Al2O3 catalyst prepared using an incipient wetness
a relatively high temperature of 383 K.85,86 The Au/CMK-3 method showed a maximum activity of about 600 mmol g 1 min 1
catalyst lost its activity, due to the absorption of gluconic acid with 100% yield of sodium gluconate at 313 K and pH = 9 in a
molecules. XPS analysis revealed that a peak of O 1s with the batch reactor with an O2 flow rate of 500 mL min 1 at atmo-
binding energy at 534.1 eV was attributed to the –COOH group, spheric pressure.94 The Au/g-Al2O3 catalyst could be re-used
which increased from 5% in the fresh catalyst to 21% in the 20 times without any decrease of its activity. Similarly, Thielecke
spent catalyst. Two methods were applied to recover the activity of and co-workers reported that the Au/g-Al2O3 catalyst showed no
the spent catalyst. On the one hand, the treatment of Au/CMK-3 loss of its activity and selectivity during 70 days of continuous-
with NaOH at 363 K could produce glucose conversion up to 87% flow glucose oxidation.98,99
after 2 h at 383 K and 3 bar O2, which was close to the results of a Au containing bimetallic and trimetallic catalysts have also
fresh catalyst. However, the treatment of Au/CMK-3 by calcination been widely used for glucose oxidation either in the form of
at 773 K only produced glucose conversion of 55% after 2 h alloys or supported states, including Au–Ag,106–109 Au–Pt,110,111
under the same conditions. The authors ascribed the reason to Au–Pd,112–114 Au–Pt–Pd115 and Au–Pt–Ag,116–118 which have
the growth of Au nanoparticles to 3.5 nm during the calcina- been documented by Zhang and Toshima in 2013.69 The
tions process, while it was kept the same at 3.0 nm with the bimetallic catalysts generally showed better catalytic perfor-
treatment by NaOH. mance than the monometallic catalysts.69 For example, the
Metal oxide supported Au catalysts were found to be more bimetallic Au–Pt/C catalyst showed 6 times higher catalytic
stable than the Au/C catalyst under basic conditions. Prüße and activity (TOF) than the Au/C catalyst without pH control.111
co-workers prepared a TiO2 supported Au (Au/TiO2) catalyst In conclusion, supported Au catalysts typically show
using the deposition–precipitation method, which produced a higher activity and stability over supported Pd and Pt catalysts.
98% yield of gluconic acid within 2 h at 313 K and pH = 11 with Therefore, supported Au catalysts should be promising for the
an O2 flow rate at 500 mL min 1 under atmospheric pressure large-scale production of gluconic acid from glucose oxidation.
(Table 5, entry 14).89 The Au/TiO2 catalyst could be reused at High pH is generally required to achieve high activity and
least 17 times without a discernible loss of its activity. ICP selectivity with supported Au catalysts. Similar to the results
analysis indicated that there was no leaching of Au, and TEM reported by Thielecke and co-workers,99 the Au/g-Al2O3 catalyst
indicated that there was no change in the Au particle size. Later, could be used in a continuous reactor for the large scale
Hutchings and co-workers prepared three kinds of Au/TiO2 catalysts oxidation of glucose, but it was performed in an aqueous
for the base-free oxidation of glucose.90 Conventional impregnation solution of NaOH. The product is in the form of sodium

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1365
View Article Online

Review Article Chem Soc Rev

gluconate and would require acidification to be converted into reuse of the homogeneous FeSO4 catalyst, and the purification
gluconic acid. of the product from FeSO4 due to the high affinity of Fe3+ with
gluconic acid. In addition, this system also required the use
6.2 Oxidation of glucose into gluconic acid with H2O2 as the of a high amount of H2O2 (12 equiv.) compared with those with
oxidant Au catalysts.
The oxidation of glucose into gluconic acid is mostly performed
by the use oxygen as the oxidant rather than H2O2. However, 6.3 Conversion of polysaccharides and lignocellulose into
gluconic acid
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

there are a handful of studies where they use H2O2 as the


oxidant for glucose oxidation. In addition to O2, H2O2 was also Synthesis of gluconic acid from polysaccharides and ligno-
used for the oxidation of glucose because of its reasonable cellulose basically requires two successive steps: acid-catalyzed
price, safe storage, and low environmental impact. Prüße hydrolysis of lignocellulose or polysaccharides into glucose and
and co-workers performed the oxidation of glucose with H2O2 the subsequent oxidation of the produced glucose. Therefore,
using a commercial Au/g-Al2O3 catalyst.119 Under mild reaction bifunctional catalysts combining both metal and acidic sites are
conditions (313 K, pH = 9), quantitative gluconic yields (499%) essential to integrate the hydrolysis and oxidation steps into a
were obtained at glucose concentrations ranging from 2–50 wt% one-pot reaction for the conversion of cellulose or polysaccharides
using 1.1 equiv. H2O2. H2O2 was added stepwise in constant into gluconic acid.
portions during the course of the reaction until the total amount In 2009, Wang and co-workers first reported the transforma-
was reached to avoid the decomposition of H2O2. The apparent tion of cellobiose into gluconic acid over supported Au catalysts
activation energy was 48 kJ mol 1 in the range of 303–333 K, without pH control.123 Carbon nanotubes (CNTs) pretreated
which was similar to the apparent activation energy (47 kJ mol 1) with 68 wt% HNO3 supported Au catalyst (Au/CNT) produced
of glucose oxidation with O2.94 Later, Repo and co-workers the highest gluconic acid yield of 68% at a cellobiose conver-
performed microwave-assisted base-free oxidation of glucose sion of 81% after 3 h at 418 K and 5 bar O2 (Table 6, entry 1).
over the Au/g-Al2O3 catalyst with an average size of 2.4 nm.120 The treatment of CNTs with HNO3 created more acidic sites as
The microwave irradiation could accelerate the reaction rate, measured by NH3-TPD, which were used for the hydrolysis of
which produced a glucose conversion of 83% with a gluconic cellobiose. Using Al2O3 and MgO as supports also produced
acid selectivity of 87% in 10 min by the use of 2.2 equiv. H2O2 higher cellobiose conversions (93–100%), but the gluconic acid
at 393 K. Recycling experiments indicated that the Au/g-Al2O3 selectivity was lower (10–20%). The supports showed a great
catalyst retained constant activity during four consecutive runs, effect on the degradation of gluconic acid. For example, 14%,
even though the Au particle size was observed to increase 67% and 97% of gluconic acid degraded with CNT, Al2O3 and
from 2.4 nm to 3.6 nm. Omri and co-workers also reported a MgO, respectively. Similar to the Au/CNT catalyst, the acidic
microwave-assisted oxidation of glucose into gluconic acid over carbon xerogel (CX) supported Au catalyst (Au/CX) also showed
different metal-oxide supported Au catalysts with H2O2 as the activity for the one-pot conversion of cellobiose to gluconic
oxidant.121 Au/g-Al2O3 and Au/CeO2 showed higher catalytic acid.124 The carbon xerogels were prepared by the polyconden-
activity than Au/TiO2 with the same catalyst weight and a similar sation of resorcinol with formaldehyde (molar ration = 1 : 2) in
average size of around 2–3 nm. High conversions and high water, and the subsequent carbonation in a nitrogen atmo-
selectivity beyond 90% were produced over the Au/g-Al2O3 and sphere. Phenolic groups (acidic sites) were introduced on the
Au/CeO2 catalysts in 10 min with 1 equiv. of NaOH and 3 equiv. surface of the Au/CX catalyst for the hydrolysis of a glycosidic
of H2O2, while the glucose conversion was 46% and a gluconic bond. A selectivity of nearly 80% to gluconic acid and a glucose
acid selectivity of 69% was obtained from the Au/TiO2 catalyst. conversion of 75% were obtained after 1.25 h at 418 K and 5 bar
The peroxide radical (HO2 ) is supposed to be the active species O2 (Table 6, entry 2). The partial exchange of H+ in hetero-
for the oxidation of glucose into gluconate. A silica-supported Au polyacids (HPW) with large cations (e.g., Cs+, Ag+) not only
catalyst (Au/SiO2) suspended in 30% H2O2 was also reported for retained their acidity, but also made them not dissolve in water
the oxidation of glucose into gluconic acid.103 100% conversion or organic solvents.125 Zhang and co-workers prepared a bifunc-
of glucose and B80% yield of gluconic acid were attained within tional Au/Cs2HPW12O40 catalyst for the oxidative conversion of
1 h at room temperature and pH = 9.2 by the use of 6 equiv. cellobiose with a high gluconic acid yield of 96.5% after 3 h at
H2O2. Besides gluconic acid, glucaric acid was produced with a 418 K and 5 bar O2. (Table 6, entry 3).126 The Au/Cs2HPW12O40
selectivity of about 15%. Some catalyst stability problems were catalyst did not have a high rate of catalyzing gluconic acid
observed using this catalyst system. degradation as only 4.5% of gluconic acid was converted into
Besides supported Au catalysts, the inexpensive and abun- other byproducts by the use of pure gluconic acid as a feed. The
dant FeSO4 was also used for the oxidation of glucose into Au/Cs2HPW12O40 catalyst was more effective than the Au/CNT
gluconic acid with H2O2 as the oxidant.122 A high gluconic acid and Au/CX catalysts (Table 6, entries 1, 2 vs. 3). The better
yield up to 97% was obtained after 15 min with ultrasonic catalytic performance of Au/Cs2HPW12O40 was possibly due to
activation at 70 1C by the use of 12 equiv. H2O2. A hydroxyl its higher acidity, which catalyzed the hydrolysis of cellobiose
radical (HO ) was the oxidative species, which was produced faster. Recycling experiments of the Au/Cs2HPW12O40 catalyst
from the reaction of H2O2 with FeSO4. This catalytic system indicated that cellobiose conversion did not change significantly,
demonstrates many challenges including the recycling and but the selectivity of gluconic acid dropped by approximately 7.8%

1366 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

Table 6 One-pot conversion of cellobiose or lignocellulose into gluconic acid and in bath reactorsa

Feed con. Catalyst


Entry Substrates (mg mL 1) Catalyst ratiob T (K) O2 (bar) Time (h) Con. (%) Yield (%) Ref.
1 Cellobiose 5.13 Au/CNT 236 418 5 3 81 68 123
2 Cellobiose 4.1 Au/CXs 118 418 5 1.25 75 B80 124
3 Cellobiose 5.0 Au/Cs2HPW12O40 118 418 5 3 97.5 96.4 126
4 Cellobiose 0.4 Au/Cs1.2H1.8PW12O40 118 418 5 3 97 97 127
5 Cellulose 20 Au/Cs1.2H1.8PW12O40 118 418 10 11 — 60 127
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

6 Cellobiose 5 Au/Amberlyst-15 14 373 In air 24 — 44.9 128


7 Cellobiose 5.13 Au/TiO2 118 393 5 3 100 67.4 129
8 Cellobiose 10.3 0.5Cu–0.5Au/TiO2 58 418 10 3 100 88.5 130
9 Cellobiose 10.3 0.5Ru–0.5Au/TiO2 80 418 10 9 100 86.9 130
10 Cellobiose 10.3 0.5Co–0.5Au/TiO2 55 418 10 3 96.8 33.6 130
11 Cellobiose 10.3 0.5Pd–0.5Au/TiO2 83 418 10 3 96.8 35.6 130
12 Starch 9 Pt/AC-SO3H — 393 In air 24 — 40 132
13 Pullulan 9 Pt/AC-SO3H — 393 In air 24 — 35 132
14 Cellobiose 9 Pt/AC-SO3H 13 393 In air 24 — 46 132
a b
All of the reactions were carried out in water. The catalyst ratio = the molar ratio of cellobiose to the noble metals.

in the second run and then stabilized in the following runs. 0.5Ru–0.5Au/TiO2 catalyst to get similar results (Table 6, entry 9).
Similar work was later reported by An and co-workers by the use of Interestingly, the reaction pathways were affected by the com-
the Au/Cs1.2H1.8PW12O40 catalyst (Table 6, entry 4).127 Meanwhile, position of the bimetallic catalysts. For the reaction with the
the Au/Cs1.2H1.8PW12O40 catalyst was also effective for the oxidative 0.5Cu–0.5Au/TiO2 catalyst, it is suggested that cellobiose is
conversion of cellulose into gluconic acid with a yield of 60% after firstly converted to cellobionic acid, followed by the cleavage
11 h at 418 K and 10 bar O2 (Table 6, entry 5). The yield of gluconic of the b-1,4-glycosidic bonds to produce gluconic acid (Fig. 17,
acid from cellulose increased to 85% by co-feeding H3PO4 with Au/ path A). For the reaction over the 0.5Ru–0.5Au/TiO2 catalyst,
Cs1.2H1.8PW12O40. H3PO4 promoted the hydrolysis of cellulose into glucose was observed as the reaction intermediate and gluconic
glucose. A slight decrease of gluconic acid yield was observed over acid was formed as a result of glucose oxidation (Fig. 17, path B).
the H3PO4 and Au/Cs1.2H1.8PW12O40 catalytic system, due to the For the reactions over the 0.5Co–0.5Au/TiO2 and 0.5Pd–0.5Au/TiO2
growth of Au nanoparticles from 2.7 nm (fresh catalyst) to 3.9 nm catalysts, low gluconic acid yields were produced under the same
(the spent catalyst after 6 runs). Sakurai and co-workers deposited reaction conditions (Table 6, entries 10 and 11), and fructose was
Au nanoparticles with an average diameter of 7.3 nm on the surface observed as the intermediate from the isomerization of glucose
of Amberlyst-15 to generate a metal–acid bifunctional catalyst (Fig. 17, path C). Furthermore, 0.5Co–0.5Au/TiO2 and 0.5Pd–
(Au/Amberlyst-15) for cellobiose conversion to gluconic acid.128 0.5Au/TiO2 catalysts remarkably promoted the successive retro-aldol
The yield of gluconic acid was low at 44.9% together with a condensation reaction of fructose to produce glyceraldehyde and
glucose yield of 6.3% after 24 h at 373 K in air (Table 6, entry 6). glycolaldehyde, which were further oxidized to glycolic acid. For
Besides the use of supports with Brønsted acidity, metal- example, fructose, glyceraldehyde and glycolic acid were also
oxides with Lewis acidity were also used for the construction produced over the 0.5Co–0.5Au/TiO2 catalyst with the selectivity
of Au–acid bifunctional catalysts for the transformation of of 28.6%, 10.6% and 18.1%, respectively. The above discussed
cellobiose into gluconic acid. Yang and co-workers reported examples all used Au–acid bifunctional catalysts for the oxidative
that a TiO2 supported Au catalyst (Au/TiO2) could simultaneously conversion of cellobiose into gluconic acid. A sulfonated activated-
catalyze cellobiose hydrolysis and the subsequent oxidation of carbon supported Pt catalyst (Pt/AC-SO3H) was reported to show
glucose into gluconic acid.129 NH3-TPD profiles presented two- catalytic activity towards the oxidative conversion of cellobiose and
peaks with a total acidity up to 354.54 mmol g 1 for the Au/TiO2 even lignocelluloses such as starch and pullulan into gluconic acid.
catalyst. One weak peak at a low temperature of 534 K was Gluconic acid was produced with relatively low yields of 35–46%
attributed to the adsorption of NH3 with very weak Brønsted after 24 h in water at 393 K in air (Table 6, entries 12–14).132
acid, whereas the strong and conspicuous peak at 648 K was As listed in Table 6, cellobiose can be successfully converted
attributed to the desorption of NH3 on the strong Lewis acid into gluconic acid over bifunctional catalysts, but it is still not
sites of TiO2. The Au/TiO2 catalyst reduced at 700 1C under a H2 suitable for the large-scale production of gluconic acid due to the
flow produced 100% cellobiose conversion and 67.4% gluconic high cost of cellobiose. The use of cellulose or lignocellulose as the
acid yield within 3 h at 418 K and 5 bar O2 (Table 6, entry 7). The feedstocks should be encouraged for the production of gluconic
byproducts were glycolic acid (14.3% selectivity) and ethylene acids. Au catalysts with a strong acidic support are required to
glycol (13.2% selectivity). Later, the same authors improved this realize the hydrolysis of cellulose into glucose and the in situ
method by the use of TiO2 supported bimetallic catalysts.130,131 oxidation of glucose into gluconic acid simultaneously. A support
Complete conversion of cellobiose (100%) and a gluconic acid with strong Brønsted acids would be preferred to hydrolyze the
selectivity of 88.5% were produced over the 0.5Cu–0.5Au/TiO2 cellulose into glucose and stabilize Au nanoparticles with a
catalyst after 3 h at 418 K and 10 bar O2 (Table 6, entry 8). specific size and morphology for the subsequent oxidation of
Whereas, a longer reaction time of 9 h was required for the glucose into gluconic acid.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1367
View Article Online

Review Article Chem Soc Rev


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fig. 17 Proposed reaction pathways of cellobiose conversion to gluconic acid and its derivatives. Adapted from ref. 131 with permission from Elsevier,
copyright 2015.

6.4 Conversion of glucose to glucaric acid have been reported for the oxidation of carbohydrates into
Glucaric acid is another important oxidation product of glucose glucaric acid with molecular oxygen.139–145
with the formula of C6H10O8 (Fig. 18). It has been classified by Chaudhari and co-workers prepared a series of supported Pt
the United States Department of Energy as a versatile platform catalysts for the oxidation of glucose to glucaric acid with O2 as
chemical from biomass for making a number of useful products, the oxidant.139 Initially, the authors screened different kinds of
such as nylons, plastics, and food additives.133 It serves as an catalysts for the oxidation of sodium gluconate with bubbles
important precursor for one of the most widely used materials in of oxygen under atmospheric pressure. As shown in Table 7,
industry, adipic acid, which is conventionally obtained from TiO2 was a better support than CeO2 (Table 7, entries 1 vs. 2).
fossil-fuel-based feedstocks via energy-intensive processes using Pt catalysts outperformed other monometallic Pd, Cu and Co
expensive catalysts.134 catalysts in terms of conversion and glucaric acid selectivity
In the past, glucaric acid was produced by the oxidation of under the same conditions (Table 7, entries 1 vs. 5, and entries
glucose with nitric acid.135 Other methods of glucaric acid synthesis 2 vs. 3 and 4). Other products were tartronic acid, and oxalic
included either electrochemical or stoichiometric oxidation acid, along with monocarboxylic acids such as glyceric, lactic,
using undesirable reagents such as NaBr and NaOCl.136–138 glycolic, and formic acids. The PtCu/TiO2 bimetallic catalysts
High concentrations of these oxidants in the reaction medium exhibited much better performance than the Pt/TiO2 and Cu/TiO2
and the generation of significant amounts of toxic by-products catalysts, suggesting that there was a synergetic effect between
and inorganic salts are the major issues confronting the Cu and Pt (Table 7, entries 2, 4 vs. 6). The authors further
sustainability and environmental compatibility of these current studied the different ratios of Pt to Cu for the oxidation of
routes. Catalytic oxidation of carbohydrates into glucaric acid glucose to glucaric acid, and found that the Pt1Cu3/TiO2 catalyst
has become much more attractive in recent years. Unlike the gave the highest TOF of 3542 h 1. Kinetic data indicated that the
oxidation of glucose into gluconic acid, the oxidation of glucose addition of Cu to Pt altered the possible oxidation pathways,
is much more difficult, as it requires the simultaneous oxida- thus lowering the activation barriers for the formation of
tion of the hydroxyl group at the C6 position and the aldehyde glucaric acids and other acids. The catalytic oxidation of
group at the C1 position into carboxylic acid groups (Fig. 16). glucose over the Pt1Cu3/TiO2 catalyst produced glucaric acid
However, the oxidation of the hydroxyl group necessitates more with a yield of 25.4% at full conversion after 24 h at 318 K and
severe conditions than that of the aldehyde. Hence, side and 1 bar O2 by the addition of NaOH at a rate of 0.2 mL min 1 for
consecutive C–C cleavage reactions generally occur, resulting in the initial 2 h. The Pt1Cu3/TiO2 catalyst was stable and could be
a poor selectivity. Recently, a few kinds of supported Pt catalysts reused 3 times without significant loss of its catalytic activity.

Fig. 18 Catalytic oxidation of glucose into glucaric acid.

1368 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

Table 7 Comparison of mono- and bimetallic catalysts for sodium gluconate oxidation

Selectivity (%)
Entry Catalyst Conversion (%) Glucaric Tartronic Oxalic Others Carbon balance (%)
1 Pt/CeO2 28.0 47.0 31.1 6.4 15.6 100.1
2 Pt/TiO2 70.8 38.1 27.7 5.8 11.5 83.1
3 Pd/TiO2 41.1 44.2 28.3 8.3 11.5 92.3
4 Cu/TiO2 2.8 31.2 30.0 9.9 21.1 90.4
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

5 Co/CeO2 5.5 21.8 38.9 15.6 15.3 91.6


6 PtCu/TiO2 100 32.3 27.2 19.5 11.1 91.8
Notes: reaction conditions: 3 g sodium gluconate, 1.0 g NaOH, 0.1 g solid catalyst, metal loading 2 wt% or 2–2 wt%, T: 60 1C, others: glyceric,
glycolic, and formic acids. Reaction time: 6 h. Adapted from ref. 139.

Later, the same authors prepared another bimetallic PtPd/TiO2 Meanwhile, glucuronic acid easily isomerizes to 5-keto-gluconic
catalyst for the oxidation of glucose into glucaric acid.140 acid in the reaction medium. Direct C–C cleavage of 5-keto-
Similar to the Pt1Cu3/TiO2 catalyst, the PtPd/TiO2 catalyst also gluconic acid generated precursors for tartronic acid (Fig. 19).141
exhibited synergistic activity compared with the monometallic C–C cleavage of glucuronic acid generated C2 and C4 inter-
Pt and Pd ones. Compared with the Pt1Cu3/TiO2 catalyst, the mediates and quickly gave C2 intermediates, which will even-
PtPd/TiO2 catalyst showed higher catalytic performance. The tually form oxalic acid through further oxidation (Fig. 19).
oxidation of glucose produced glucaric acid with a selectivity Furthermore, direct C–C cleavage of gluconic acid can generate
of 44% at full conversion under the same reaction conditions. several monocarboxylic acid precursors such as formaldehyde,
Meanwhile, tartronic acid with a selectivity of 15% and oxalic glycolic aldehyde, and glyceraldehyde, which undergo further
acid with a selectivity of 6% were also produced. Concentration– reaction to form lactic, formic, glycolic, and glyceric acids
time profiles on the PtPd/TiO2 catalyst showed that no other (Fig. 19). Recent patents were issued using the base-free oxida-
products besides gluconic acid were formed before glucose was tion of glucose to glucaric acid over Pt and Pt–Au catalysts.142,143
completely consumed. These results indicated that glucose The glucaric acid yield was reported to be 66% in the presence
inhibited the second step oxidation of gluconic acid in the reaction of Pt/SiO2 in multiparallel millireactors after 8 h at 363 K and
medium, with the secondary oxidation and C–C cleavage occurring 5 bar O2. The yield reached 71% after 5 h at 393 K and 28 bar
after glucose was consumed. After there was no glucose in the O2 over the Au–Pt/TiO2 catalyst. However, these reactions were
reaction solution, the gluconic acid concentration gradually conducted on a micro-liter scale in a 96 well-reactor system,
decreased, which was firstly oxidized into glucuronic acid, followed which was neither reproduced nor conducted on a reasonable
by oxidation of glucuronic acid into glucaric acid (Fig. 17). laboratory scale.

Fig. 19 Reaction pathway for the oxdiation of glucose into glucaric acid over the PtPd/TiO2 catalyst. Adapted from ref. 140 with permission from
American Chemical Society, copyright 2016.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1369
View Article Online

Review Article Chem Soc Rev

Besides metal-oxide supported Pt catalysts, Pt/C catalysts also of chemicals. HMF, produced by the dehydration of C6
showed catalytic ability for the oxidative conversion of glucose carbohydrates, is considered to be one of the most important
into glucaric acid. In 1972, Dirkn and co-workers reported that platform chemicals derived from biomass.149,150 As shown in
the oxidation of gluconic acid over the Pt/C catalyst produced a Fig. 20, several kinds of furan derivatives can be obtained by the
55% glucaric acid yield (gluconic acid/Pt = 20) at controlled oxidation of HMF including DFF, HFCA, FFCA, and FDCA.
pH (8–11).144 Later, the oxidation of glucose over the Pt/C catalyst
was also performed at pH = 9. Gluconate was first formed with 7.1 Selective oxidation of HMF into HFCA and FFCA
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

an 80% yield at nearly total conversion of glucose. Then,


HFCA is produced by selective oxidation of the formyl group in
gluconate was further oxidized to glucarate.145 The final selec-
HMF. HFCA is currently only produced on a small scale for the
tivity of glucaric acid was 57% at a 97% glucose conversion.
fine chemical industry ($2.5 per mg, Sigma-Aldrich). It has been
Vlachos and co-workers studied the oxidation of glucose into
used as a monomer in the synthesis of various polyesters,151
glucaric acid over three kinds of commercially available Pt
and as a building block of an interleukin inhibitor.152 HMF can
catalysts (Pt/C, Pt/SiO2 and Pt/Al2O3) with 5 wt% Pt loading.146
be selectively oxidized into HFCA with stoichiometric amounts
Under the same reaction conditions (353 K and 6.2 bar O2),
of silver oxide153 or a mixture of silver and copper(II) oxides154
the catalytic activity of the Pt catalysts decreased in the order of
as oxidants at room temperature. These methods have little
Pt/C 4 Pt/SiO2 4 Pt/Al2O3. The TOFs at 20% conversion
practical application as the cost of the oxidants is high and
of glucose were 879, 195, and 102 h 1 for the Pt/C, Pt/SiO2 and
stoichiometric amounts of metal byproducts are produced. There
Pt/Al2O3 catalysts, respectively, while those for gluconic acid
have been fewer reports on the oxidation of HMF into HFCA with
oxidation, starting from gluconic acid as a substrate, are 285,
O2 as the oxidant. HFCA is an intermediate of the oxidation of
42, and 13 h 1, respectively. These data clearly indicated that the
HMF into FDCA. Gorbanev and co-workers obtained HFCA in a
oxidation rate of the aldehyde group was much faster in glucose
yield of 76% together with a FDCA yield of 20% after 14 h over the
than that of a –CH2OH group in gluconic acid. Gluconic acid is
Au/CeO2 catalyst at 403 K and 10 bar air with NaOH/HMF molar
the major product in an acidic solution, while the selectivity to
ratio = 1.155 Similar work was also reported by Davis and
glucaric acid is poor in a highly basic solution, due to the C–C
co-workers.156 Full HMF conversion with 92–93% selectivity
bond cleavage to low carbon chain carboxylic acids. Thus, the
towards HFCA was attained over the Au/C and Au/TiO2 catalysts
oxidation of glucose into glucaric acid was performed without
after 6 h and 6.9 bar O2 with NaOH/HMF molar ratio = 2.
the addition of a base. A high glucaric acid yield of 74% was
A method was developed for the oxidation of HMF into
produced after 10 h at 353 K and 13.8 bar O2 with glucose/Pt
HFCA using non-noble metal catalysts without a base.157 Bis(acetyl-
ratio = 54. Recently, Besson and co-workers studied the effect of
acetonato)dioxo-molybdenum(VI) [MoO2(acac)2] supported on
the possible impurities in the glucose residues on the oxidation
montmorillonite K-10 clay (K-10 clay-Mo) produced a 86.9%
of glucose.147 The oxidation of glucose produced glucaric acid in
HFCA yield after 3 h at 110 1C with flowing oxygen. DFF was
a 54% yield over the 4.7% Pt/C catalyst after 24 h at 333 K and
identified as the byproduct with a yield of 6.7%. The K-10 clay-Mo
pH = 9 with an air flow rate of 0.5 mL min 1. As regards the
catalyst had a HFCA yield of 86.9% in the first run and 83.9% in
residues in the hydrolysate, sodium salts (acetate and sulfate) or
the sixth run. This catalytic system had three distinct advantages
disaccharides (cellobiose and maltose) showed no significant
over the above examples with Au catalysts: (1) it did not contain
effect on the course of the reaction. The effect of HMF and
any non-precious metal, (2) it did the reaction without a base
furfural (the dehydration products of hexose and pentose sugars,
co-catalyst, and (3) downstream separation and purification of
respectively)148 were also studied, and furfural significantly
HFCA is easy. HFCA was precipitated out from toluene in the
slowed down the reaction rates of glucose oxidation into glucaric
form of needle-like crystals. However, the synthesis of HFCA in
acid over the Pt/C catalyst.
strong alkaline solution over Au catalysts produced HFCA in
its salt form, which makes it difficult to separate and purify
7. Synthesis of furan compounds via the product.
catalytic oxidation reactions The selective oxidation of HMF into FFCA is much more
difficult than the oxidation of HMF into HFCA. Recently,
HMF, furfural and other platform chemicals produced from Dibenedetto and co-workers reported a new method for the
biomass can also be catalytically oxidized into a wide variety oxidation of HMF into FFCA in water by the use of a mixed

Fig. 20 Synthesis of furan chemicals from HMF and carbohydrates.

1370 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

oxide CuOCeO2 as the catalyst.158 FFCA was produced in 90% oxidation of HMF into DFF using Cu(NO3)2/VOSO4 as the
yield at 383 K and 9 bar O2. The CuO by itself achieved a catalyst. A 98% DFF yield was produced after 1.5 h at 353 K
moderate selectivity of FFCA (40%) at a 33% conversion, while and 1 bar O2 (Table 8, entry 2).166 The oxidation of HMF over
CeO2 showed a higher selectivity towards FFCA (76.6%) at the Cu(NO3)2/VOSO4 catalyst can even produce DFF with a high
19.9% conversion of 19.9%. The formation of a CuOCeO2 yield of 98% at room temperature after 48 h (Table 8, entry 3).
mixed oxide increased the basicity of the catalyst. This is the highest reported yield for the oxidation of HMF
into DFF at room temperature. The molar ratio of Cu(NO3)2/
7.2 Selective oxidation of HMF to DFF with O2 [Cu(NO3)2 + VOSO4] influenced the reactivity and selectivity.167
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

DFF can be used in the synthesis of polymeric Schiff bases, as an The conversion of HMF goes through a maximum with increased
antifungal agent, in pharmaceuticals, in macro-cyclic ligands, in molar concentration of Cu(NO3)2. The highest DFF yield is
fluorescent materials and in resins.159–161 Several reports have obtained at a molar ratio of Cu(NO3)2 to VOSO4 = 1. The authors
used stoichiometric reagents such as BaMnO4, pyridinium claim that the high DFF selectivity is ascribed to the Cu2+ cation,
chlorochromate, NaOCl and 2,2,6,6-tetramethylpiperydine-1- which prohibits the oxidative C–C bond cleavage reaction
oxide (TEMPO) for oxidation of HMF into DFF.162–164 of HMF and prevents a radical reaction of DFF to humins.
As listed in Table 8, there are many reports that use O2 as the A mechanism was proposed by the authors where the oxidation
oxidant with both homogeneous and heterogeneous catalysts. of HMF into DFF was catalyzed by V5+ species (Fig. 21).
In 2001, Partenheimer and Grushin reported the catalytic The V4+ species was then regenerated to V5+ by the NOx,
oxidation of HMF into DFF over Co(OAc)2/Mn(OAc)2/Zr(OAc)4/ and NOy was reoxidized to NOx by O2. Tong and co-workers
NaBr catalysts (Table 8, entry 1).165 This method produced a reported that the combination of CuI and 1-hydroxybenzotriazole
57% DFF yield after 5.17 h at 348 K. HMF acetate (formed via synergistically promoted the oxidation of HMF into DFF in
the esterification of HMF with acetic acid) and COx were the 93% yield in dimethyl sulfoxide (DMSO, Table 8, entry 4).168
side products with a selectivity of 7.2% and 2.1%, respectively. 1-Hydroxybenzotriazole played a role in the abstraction of hydrogen
Xu and co-workers reported an effective method for the from HMF, promoting the formation of DFF. The use of Cu(NO3)2

Table 8 Catalytic oxidation of HMF into DFF in batch reactors

Entry Catalyst Catalyst ratioa Time (h) T (K) Solvent Oxidant Con. (%) Yield (%) Ref.
1 Co/Mn/Zr/Br HMF/Co = 200 5.17 348 Acetic acid 1 bar air 91 57 165
2 Cu(NO3)2/VOSO4 HMF/Cu = 100 1.5 353 Acetonitrile 1 bar O2 99 98 166
HMF/V = 100
3 Cu(NO3)2/VOSO4 HMF/Cu = 100 48 298 Acetonitrile 1 bar O2 99 98 166
HMF/V = 100
4 CuI HMF/Cu = 5 10 403 DMSO 3 bar O2 93.2 92.3 168
5 Cu(NO3)2 HMF/Cu = 50 7 323 Acetonitrile 1 bar O2 — 71 169
6 PVP–VO2+ complex HMF/V = 30 24 403 DMSO 10 bar air 82 81 170
7 SBA–Py–VO2+ HMF/V = 30 24 403 Triflurotoluene 10 bar air 50 49 170
8 SBA–NH2–VO2+ HMF/V = 81; 12 383 4-Chlorotoluene 1 bar O2 98.8 63.4 171
SBA–NH2–Cu2+ HMF/Cu = 154
9 V2O5 HMF/V = 2.8 6 443 Toluene 1 bar air 91 62.7 173
10 V2O5/TiO2 HMF/V = 2.8 4 363 Toluene 16 bar air 100 90 175
11 V2O5/H-beta HMF/V = 72 3 398 DMSO 10 bar O2 84 82 176
12 V2O5/Cu-MOR HMF/V = 3.6 7 373 DMSO 1 bar O2 499.9 91.5 177
13 V2O5/C HMF/V = 20 4 383 MIBK 2.8 bar O2 95.2 88.1 178
14 VO2–PANI/CNT HMF/V = 5.4 11 363 DMSO 10 bar O2 100 96 180
15 VOPO42H2O HMF/V = 4.0 6 353 MIBK 1 bar O2 98 49 181
16 VOPO42H2O HMF/V = 4.0 6 353 DMF 1 bar O2 56 55 181
17 C14VOPO4 HMF/V = 4.0 6 383 Toluene 1 bar O2 99 82 182
18 C14VOHPO4 HMF/V = 4.0 6 383 Toluene 1 bar O2 99 83 182
19 Ru/HT HMF/Ru = 23 6 393 DMF 1 bar O2 95 92 183
20 Ru/g-Al2O3 HMF/Ru = 56 4 403 Toluene 2.8 bar O2 100 97 184
21 Ru/C HMF/Ru = 80 — 383 Toluene 20 bar O2 100 96 185
22 Ru/CTF HMF/Ru = 40 3 353 MTBE 20 bar air 97.3 72.7 186
23 Ru/mPMF HMF/Ru = 96 12 393 Toluene 20 bar O2 99.6 85 188
24 RuCo(OH)2CeO2 HMF/Ru = 2.3 12 393 MIBK 1 bar O2 96.5 82.6 189
25 Ru3+–PVP/CNT HMF/Ru = 38.3 12 393 DMF 20 bar O2 100 94 191
26 Fe3O4@SiO2–NH2–Ru3+ HMF/Ru = 26.9 4 383 Toluene 1 bar O2 96.1 85.9 192
27 K-OMS-2 HMF/Mn = 0.16 6 383 DMSO 1 bar O2 100 99 193
28 K-OMS-2 HMF/Mn = 2.0 1 383 DMF 20 bar O2 100 99 194
29 Ag-OMS-2 HMF/Mn = 0.13 4 438 Isopropyl alcohol 15 bar air 100 99 195
30 K-OMS-2 HMF/Mn = 0.13 4 438 Isopropyl alcohol 15 bar air 62 51 195
31 Fe3O4/Mn3O4, DMF HMF/Mn = 0.96 4 393 DMF 1 bar O2 99.8 82.1 196
32 Mn0.70Cu0.05Al0.25 — 24 363 H2O 8 bar O2 90 78 197
33 FeCo/C-500 HMF/(Fe + Co) = 20 6 373 Toluene 10 bar O2 100 499 198
a
The catalyst ratio defined as the molar ratio of HMF to the active metals.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1371
View Article Online

Review Article Chem Soc Rev

catalyst produced a 91.5% DFF yield after 7 h at 373 K and 1 bar


O2 (Table 8, entry 12). The V2O5/Cu-MOR catalysts showed no
loss of catalytic activity after five runs. The V2O5/C catalyst also
had a high stability and high DFF yield (Table 8, entry 13).178
The V2O5/C catalyst showed no activity loss after 4 recycle runs.
XPS and TEM analysis of the fresh and spent catalysts showed
no change in the catalyst structure. Liu and co-workers system-
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

atically studied VOx on different supports and tested them for


Fig. 21 Promoted role of Cu(NO3)2 in aerobic oxidation of HMF to DFF the aerobic oxidation of HMF into DFF.179 VOx deposited on
catalyzed by VOSO4. Adapted from ref. 116 with permission from Elsevier,
acidic supports showed a much higher activity. The structure of
copyright 2011.
VOx changed from isolated monovanadate to 2D polyvanadate
domains and finally to crystalline V2O5 clusters with an
and the additive N-hydroxyphthalimide produced DFF with a increased V loading. The activity of the supported VOx catalysts
yield of 71% in acetonitrile after 7 h (Table 8, entry 5).169 first increased with the increase of the surface density of VOx,
Heterogeneous catalysts have also been used for the oxidation and then decreased, while the selectivity of DFF continuously
of HMF into DFF including supported vanadium (V), ruthenium increased with the surface density of VOx and then kept stable.
(Ru) and manganese (Mn). Navarro and co-workers performed the Vanadium dioxide (VO2) immobilized on polyaniline (PINI)-
oxidation of HMF over poly(4-vinylpyridine) supported vanadyl functionalized carbon nanotubes (CNTs) (VO2–PANI/CNT) was
acetylacetonate [VO(acac)2] (abbreviated as PVP–VO2+).170 They also used to produce a DFF yield of 96% (Table 8, entry 14).180
used pyridine to enhance the catalytic activity. Pyridine acted as However, the catalyst was not stable and the HMF conversion
a base to abstract the proton of the hydroxyl group of HMF. An decreased from 91% in the initial run to 81% and 62% for the
82% HMF conversion and a DFF selectivity of 99% were obtained second and third runs, respectively.
in DMSO after 24 h over the PVP–VO2+ catalyst with pyridine Vanadium phosphate oxide based catalysts showed catalytic
(0.8 equiv. to HMF) as an additive (Table 8, entry 6). However, 28% activity towards the oxidation of HMF into DFF. Carlini and
of the PVP–VO2+ catalyst leached after the first run. The catalyst co-workers studied the oxidation of HMF into DFF over a
was then supported on a functionalized mesoporous-silica support VOPO42H2O catalyst.181 Both the catalytic activity of VOPO42H2O
containing trimethyl and g-aminopropyl groups (SBA–Py–VO2+). and the selectivity of DFF were greatly affected by the solvents.
When the SBA–Py–VO2+ catalyst was used, only 50% conversion of A 50% yield of DFF was obtained with this catalyst at a 98%
HMF was achieved in triflurotoluene but with a 98% DFF selectiv- HMF conversion in MIBK (Table 8, entry 15). The selectivity
ity after 24 h at 403 K and 10 bar air (Table 8, entry 7). No leaching increased to 99% using an N,N-dimethylformamide (DMF) solvent,
of the SBA–Py–VO2+ catalyst was observed. The lower catalytic but the HMF conversion sharply decreased to 56% (Table 8,
activity of SBA–Py–VO2+ was attributed to a possible inaccessibility entry 16). No reaction was observed in water. Dumeignil and
of the reactants to the active sites owing to the surface capping co-workers improved the catalytic performance of vanadium
with trimethyl groups. Inspired by the results reported by Xu and phosphate oxides by intercalating long alkyl chain trimethyl-
co-workers,166 Liu and co-workers used g-aminopropyl group ammonium into the catalyst structure (C14VOPO4 and C14VOHPO4
modified SBA-15 as the carrier to support Cu2+ and VO2+ to catalysts).182 DFF was produced in yields of 82% and 83% after 6 h
prepare SBA-15–NH2–Cu2+ and SBA-15–NH2–VO2+ catalysts.171,172 over the C14VOPO4 and C14VOHPO4 catalysts, respectively (Table 8,
However, a lower DFF selectivity of 63.4% was produced at a high entries 17 and 18). The authors proposed that this improvement
HMF conversion of 98.8% (Table 8, entry 8) compared with the was ascribed to the reduced oxidative properties of the vanadium
homogeneous catalytic system.166 The other oxidation product was phosphorus oxides, which prevented the deeper oxidation of HMF
FDCA. The combined SBA-15–NH2–Cu2+ and SBA-15–NH2–VO2+ into organic acids by intercalating long alkyl chain trimethyl-
catalysts were stable for the oxidation of HMF into DFF with ammonium. However, the C14VOPO4 catalyst deactivated with
high DFF selectivity. a 91% yield in the first run and 62% yield in the third run.
V2O5 and supported V2O5 were also used for the oxidation As discussed above, the vanadium catalysts were not effective
of HMF into DFF.173–178 HMF oxidation catalyzed by V2O5 in at effectively converting the HMF into DFF.170,173,181 Some of
toluene produced a 91% HMF conversion and 69% of DFF the vanadium catalysts were not stable.170,176,181,182
selectivity at 443 K and 1 bar air (Table 8, entry 9).173 Moreau Ru catalysts were also reported to be active for the oxidation
and co-workers performed the oxidation of HMF into DFF in of HMF into DFF. A hydrotalcite-supported ruthenium catalyst
toluene over the V2O5/TiO2 catalyst with a 90% DFF yield (Ru/HT) produced a 92% DFF yield after 6 h at 393 K (Table 8,
(Table 8, entry 10).175 V2O5/H-beta produced an 82% yield of entry 19).183 No leaching of Ru was observed. A Ru/g-Al2O3
DFF from HMF in DMSO (Table 8, entry 11).176 However, the catalyst produced a 97% DFF yield after 4 h at 403 K (Table 8,
V2O5 leached and 45% of the total activity was a result of entry 20).184 However, the Ru/g-Al2O3 catalyst lost its activity
the catalytic species dissolved from the solid catalyst. Wang decreasing to 70% and 30% conversions after the second and
and co-workers improved the activity and stability of the zeolite third runs, respectively. The decreased activity of the Ru/g-Al2O3
supported V2O5 catalyst using a Cu-containing mordenite catalyst was due to the deposition of insoluble polymeric furan
(MOR) zeolite (Cu-MOR) as the support.177 The V2O5/Cu-MOR compounds, which blocked the Ru sites. Calcination of the

1372 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

catalyst could remove the carbon deposits but produced a magnetic supports to facilitate catalyst recycling.192 A magnetic
volatile ruthenium oxide phase. Liu and co-workers studied Fe3O4@SiO2–NH2–Ru3+ catalyst was prepared by using g-amino-
the oxidation of HMF into DFF over different kinds of Ru propyl group functionalized core–shell structured Fe3O4@
catalysts at 383 K.185 For example, the TOF of Ru/C was 61.2 h 1 SiO2.192 A 99.3% HMF conversion and a 86.4% DFF yield were
at 383 K and 20 bar O2, and 30.1, 22.7, 11.2, 6.9, 2.7 and 7.4 h 1 for produced over the Fe3O4@SiO2–NH2–Ru3+ catalyst after 4 h at
Ru/Al2O3, Ru/ZSM-5, Ru/TiO2, Ru/ZrO2, Ru/CeO2 and Ru/MgO 393 K (Table 8, entry 26). After the reaction, the Fe3O4@SiO2–
catalysts, respectively. The DFF selectivity over the Ru/C catalyst NH2–Ru3+ catalyst could be readily recovered from the reaction
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

was 96.2%, and 82.4, 12.4, 57.0, 65.9, 58.4 and 21.4% for Ru/Al2O3, mixture by a permanent magnet. However, a decrease of the
Ru/ZSM-5, Ru/TiO2, Ru/ZrO2, Ru/CeO2 and Ru/MgO, respectively. catalytic activity was observed with DFF yields of 86.4% and
These results suggested that the inert carbon supported Ru catalyst 80.8% in the first and the sixth runs, respectively.
favored the oxidation of HMF into DFF by avoiding the possible Besides V and Ru catalysts, Mn catalysts are also often used for
degradation and polymerization of HMF, which were observed the oxidation of HMF into DFF. A cryptomelane-type manganese
over other Ru catalysts on the support with different acidity– oxide octahedral molecular sieve (KMn8O16nH2O; K-OMS-2), which
basicity and reducibility. A 96% DFF yield was produced over the possesses a 2  2 hollandite structure with one-dimensional pores,
Ru/C catalyst (Table 8, entry 21). Ru supported on organic porous showed a high activity towards the oxidation of HMF.193 A 99% DFF
polymers was also used for the oxidation of HMF into DFF.186–188 yield was produced in DMSO after 6 h at 383 K (Table 8, entry 27).
Palkovits et al. prepared a covalent triazine framework (CTF) The K-OMS-2 catalyst could be reused five times without a decrease
supported Ru catalyst (Ru/CTF) for the oxidation of HMF.186 CTFs in DFF yield. However, recycling a catalyst at high conversion does
are a class of highly stable polymers formed by the trimeriza- not demonstrate that the catalyst is stable. As shown in Fig. 22,
tion of aromatic dinitriles in molten ZnCl2 that have a surface HMF is oxidized by Mn4+O2 , and Mn4+O2 is simultaneously
area of 2349 m2 g 1. The Ru/CTF catalyst with a high surface reduced to a Mn3+ ion. The Mn3+ ion is then reoxidized by O2
area and a large pore volume showed a high catalytic activity. to produce Mn4+O2 to finish the catalytic cycle. Similar results
A DFF yield of 72.7% was obtained in a methyl t-butyl ether were also observed in DMF over the K-OMS-2 catalyst (Table 8,
(MTBE) solvent over the Ru/CTF catalyst (Table 8, entry 22). The entry 28).194 Although high DFF yields were obtained in DMSO
selectivity of the oxidation products was also affected by the and DMF over the K-OMS-2 catalyst, product purification was
reaction solvents. The reaction in water over the Ru/CTF difficult due to the high boiling points of DMF and DMSO.
catalyst produced FFCA, FDCA and DFF with yields of 35.3%, Yadav and Sharma observed that a silver exchanged K-OMS-2
37.7% and 8% after 1 h at 413 K and 20 bar air.187 Islam and catalyst (Ag-OMS-2) was active for the oxidation of HMF in
co-workers prepared a mesoporous poly-melamine-formaldehyde isopropyl alcohol.195 A 99% DFF yield was produced after 4 h at
material (surface area: 536 m2 g 1) supported Ru catalyst 483 K (Table 8, entry 29). The isopropyl alcohol could be easily
(Ru@mPMF) for the oxidation of HMF.188 The Ru@mPMF removed under reduced pressure, and the purity of DFF was
catalyst gave a 99.6% conversion of HMF and a 85% yield of up to 99%. Compared with Ag-OMS-2, the K-OMS-2 catalyst
DFF after 12 h at 393 K (Table 8, entry 23). The Ru@mPMF produced a lower HMF conversion of 62% and a lower DFF
catalyst could be reused without leaching of Ru. selectivity of 82% (Table 8, entries 29 vs. 30). NH3-TPD results
Besides supported metallic Ru catalysts, other kinds of Ru indicated that the surface acidity of the Ag-OMS-2 catalyst
catalysts were also studied for the oxidation of HMF into DFF. decreased, which inhibited the side reactions such as polymeriza-
The trimetallic oxide of RuCo(OH)2CeO2, prepared from RuCl3, tion of HMF into humins. HMF conversion and DFF selectivity
Co(NO3)2, and Ce(NO3)2 under basic conditions, was also were stable during five recycling experiments for the Ag-OMS-2
reported to be active for the oxidation of HMF.189 A 96.5% catalyst. A magnetic Fe3O4 supported Mn3O4 (Fe3O4/Mn3O4)
HMF conversion and a 82.6% DFF yield were produced after catalyst was used for the oxidation of HMF.196 This catalyst was
12 h at 393 K (Table 8, entry 24). The HMF conversion was 5.4% prepared by the solvent-thermal treatment of Mn(OAc)2 in the
with Co–Ce oxide, suggesting that Ru is the active species for presence of Fe3O4 nanoparticles. An 82.1% DFF yield and a full
the oxidation of HMF. Full HMF conversion with low DFF HMF conversion were produced in DMF after 12 h at 393 K
selectivity (58%) was observed with Ce–Ru oxide, suggesting (Table 8, entry 31). The Fe3O4/Mn3O4 catalyst could be readily
that cobalt plays a role in the enhancement of DFF selectivity. collected by a permanent magnet without a loss of activity. As
Although there was no leaching of Ru, HMF conversion decreased listed in Table 8, the oxidation of HMF into DFF was almost
from 96.5% in the first run to 88.6% in the fifth run. Similarly, carried out in organic solvents. Recently, Florea and co-workers
magnetic ZnFe1.65Ru0.35O4 was also prepared via the alkali- reported that a manganese–copper layered double hydroxide
coprecipitation of the metal salts and used for the oxidation of (Mn0.70Cu0.05Al0.25) could promote the oxidation of HMF into
HMF.190 A high DFF yield of 93.5% was obtained for the DFF in water,197 which was prepared by the co-precipitation of
reaction at 383 K for 4 h in DMSO with an oxygen flow, whereas these metal salts under basic conditions, followed by calcination
the maximum FDCA yield of 91.2% was achieved for the reaction in air. The optimal results were produced at 363 K and 8 bar O2
at 403 K for 16 h in H2O/DMSO. Chen and co-workers anchored with a DFF yield of 78% after 24 h (Table 8, entry 32). The catalyst
Ru3+ to poly(4-vinylpyridine)-functionalized carbon-nanotubes was stable for 5 runs. The oxidation reaction is initiated by the
(PVP/CNTs), which produced a DFF yield of 94% after 12 h at reduction of Cu2+ to Cu+ species that is accompanied by a release
393 K (Table 8, entry 25).191 Ru has also been added onto of an electron (Fig. 23, step 1). Then a cooperative effect between

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1373
View Article Online

Review Article Chem Soc Rev


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fig. 22 The mechanism for K-OMS-2-catalyzed conversion of HMF to DFF. Adapted from ref. 193 with permission from Royal Society of Chemistry,
copyright 2012.

Fig. 23 Proposed mechanism for the oxidation of HMF over the Mn0.70Cu0.05Al0.25 catalyst. Adapted from ref. 197 with permission from Elsevier,
copyright 2016.

the two cations results from the oxidation of Mn3+ to Mn4+ by the oxygen of FeCo oxides facilitated HMF oxidation to DFF, which
released electron (step 2). Furthermore, the oxidation of HMF to was later replenished by O2 molecules. The catalyst could be
DFF occurs with reduction of Mn4+ to Mn3+. The role of the reused in up to six runs without any significant loss of reactivity.
increased basicity is to induce the weakening of the hydroxyl In summary, Mn catalysts generally showed better catalytic
group of HMF, which makes the molecule easier to oxidize. performance over V and Ru catalysts towards the oxidation of
Other non-noble catalysts have also been reported to be effective HMF into DFF in terms of the DFF yield and the catalyst stability.
for the oxidation of HMF into DFF. Luque and co-workers reported Some Mn catalysts such as K-OMS-2193 and Ag-OMS-2194 not only
that a magnetic Fe–Co/C catalyst was active for the oxidation of produced quantitative DFF yield, but also showed high stability.
HMF into DFF.198 This catalyst was prepared by the pyrolysis This is because the active species of Mn in K-OMS-2193 and
of bimetallic MOF (MIL-45b) containing Co and Fe with 1,3,5- Ag-OMS-2194 are presented in the form of compounds, which was
benzenetricarboxylate (BTC) as the organic ligand. A quantitative unlike the supported vanadium oxides176 or Ru nanoparticles,184
DFF yield (499%) was produced after 6 h at 373 K (Table 8, suffering from leaching and the growth of metal nanoparticles.
entry 33). The catalytic performance of the FeCo/C catalyst was In addition, Mn based catalysts are cheap. Thus, Mn catalysts
attributed to the presence of well-ordered Fe and Co oxides show great potential in the practical production of DFF
derived from MIL-45b. The authors claimed that the lattice from HMF.

1374 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

7.3 TEMPO mediated oxidation of HMF into DFF


2,2,6,6-Tetramethylpiperidine-N-oxide (TEMPO) is a stable radical
with the formula of (CH2)3(CMe2)2NO, which is mainly used as
a co-catalyst for a variety of oxidation reactions.199 There were
some reports on the TEMPO mediated oxidation of HMF into
DFF.200–207 Chung and co-workers studied the oxidation of
HMF into DFF by the use of 4-hydroxy-2,2,6,6-tetramethyl-
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

piperidine-1-oxyl (4-hydroxy-TEMPO radical) as the oxidant


with [bis(acetoxy)-iodo]benzene (BAIB) as the co-oxidant and
acetic acid as an additive.200 A 66% DFF yield was obtained
in ethyl acetate after 0.75 h at 303 K over a 4-hydroxy-TEMPO
(0.2 equiv.)/BAIB (1.5 equiv.)/CH3COOH (0.1 equiv.) catalytic
system. As shown in Fig. 24, acetic acid aids in the initial
formation of the oxoammonium form of 4-hydroxy-TEMPO
(compound 2). HMF is then oxidized to DFF, and compound Fig. 25 Proposed reaction pathway for the aerobic oxidation of HMF into
2 is simultaneously reduced to hydroxylamine (compound 3). DFF using a Fe3O4@SiO2-TEMPO catalyst. Adapted from ref. 203 with
Then, compound 3 is oxidized by BAIB to regenerate the 4-hydroxy- permission from American Chemical Society, copyright 2015.

TEMPO radical to finish the catalytic cycle. Furthermore, TEMPO


was chemically bonded on the surface of SBA-15 (SBA-15-TEMPO).201
SBA-15-TEMPO provided a similar DFF yield (73%) with homo- grafted onto the magnetic support by the reaction of 4-hydroxyl-
geneous TEMPO (78%), which was due to its large surface area TEMPO with the g-aminopropyl functionalized Fe3O4@SiO2–NH2.
(526 m2 g 1) and large pore size (5.5 nm), facilitating the Quantitative yield of DFF (499%) was produced in toluene
diffusive transport of reactants and products. Although SBA-15- after 18 h at 323 K and 1 bar O2 using the Fe3O4@SiO2-TEMPO
TEMPO is recyclable, this method is still less attractive than the catalyst (2 mol% TEMPO) with tert-butyl nitrite (5 mol%)
methods by the use of oxygen as the oxidant because this as a co-catalyst and acetic acid (25 mol%) as an additive. As
approach requires the use of stoichiometric amounts of BAIB shown in Fig. 25, the mediation of electron transfer from HMF
and it produces iodobenzene as a byproduct. Hence, it is desirable to O2 by NO/NO2 couples and the immobilized nitroxyl radical
to perform the TEMPO mediated oxidation of HMF by the use of (Fe3O4@SiO2-TEMPO) through a combination of two redox cycles
oxygen as the terminal oxidant. is required. The catalyst was successfully recovered by an
Riisager and co-workers reported that HMF could be oxidized external magnet.
to DFF at room temperature in air under atmospheric pressure Hou and co-workers developed a more sustainable method
over a TEMPO (0.1 equiv.)/2,2-bipyridine (0.3 equiv.)/CuCl for the TEMPO mediated oxidation of HMF into DFF with
(0.3 equiv.) catalytic system.202 A 92% DFF yield was produced graphene oxide (GO) as co-catalysts.204 A 99.6% DFF yield
in acetonitrile after 24 h. Cu(I) was coordinated by 2,2-bipyridin, was produced in acetonitrile after 18 h at 373 K in air in
which was oxidized to Cu(II) by TEMPO. HMF was then oxidized the presence of GO and TEMPO. Low HMF conversions were
by Cu(II) species to give DFF and Cu(I) species. The reduction obtained under these same reaction conditions with only
state of TEMPO was then reoxidized by O2 to finish the catalytic TEMPO or GO, suggesting that GO and TEMPO played syner-
cycle. Karimi and co-workers used magnetically separable getic roles in the oxidation of HMF into DFF. Electron spin
TEMPO for the oxidation of HMF.203 TEMPO was chemically resonance (ESR) measurements indicated that the edge sites
with unpaired electrons and carboxylic acid groups were the
active catalytic sites on GO. As shown in Fig. 26, TEMPO binds
to the carboxylic acid groups in GO via a hydrogen bond and
forms an electron-donor complex. The unpaired electrons of
porous GO reduce molecular oxygen to form adsorbed super-
oxide radicals ( O2 ), which stabilize the positive charge of the
holes. The anchored TEMPO is oxidized by the positive charge
of the hole to form the oxoammonium cation, and the active
site on GO is regenerated. HMF is oxidized by the oxoammo-
nium cation, producing DFF, and the produced hydroxylamine
is reoxidized by  O2 . The TEMPO molecule is regenerated. The
electronic structure of graphene, and hence catalytic properties,
can be further modified by the introduction of certain hetero-
atoms, such as N, B, and P.205,206 The same authors enhanced
Fig. 24 Proposed reaction pathway for the aerobic oxidation of HMF into
the efficiency of TEMPO mediated oxidation of HMF by the use
DFF using TEMPO/BAIB/CH3COOH. Adapted from ref. 200 with permis- of nitrogen-doped graphene.207 A quantitative DFF yield was
sion from springer, copyright 2014. reached after 6 h at 383 K under atmospheric air pressure,

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1375
View Article Online

Review Article Chem Soc Rev


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fig. 26 Proposed mechanism for the aerobic oxidation of HMF into DFF using TEMPO and graphene oxide catalysts. Adapted from ref. 204 with
permission from American Chemical Society, copyright 2015.

while 18 h residence time was required to get 99.6% DFF yield HMF oxidation was started by the addition of V catalysts at
under the same reaction conditions using unmodified GO.204 423 K with a bubble of air. The DFF yields were 52% and 67%
based on the starting fructose over VHPO40.5H2O and V2O5
7.4 One-pot conversion of carbohydrates into DFF catalysts, respectively (Table 9, entries 1 and 2). However, the
As shown in Fig. 27 the one-pot conversion of glucose requires high reaction temperature at 423 K caused the decomposition of
three parallel steps including the isomerization of glucose into DMSO into Me2SO2 and Me2S, which contaminated the product
fructose,208 the dehydration of fructose into HMF, and the of DFF. Therefore, further purification by vacuum sublimation
oxidation of HMF into DFF. As listed in Table 9, fructose is and Soxhlet extraction was required to obtain pure crystalline
mainly used as the feedstock for the one-pot synthesis of DFF DFF. A polyaniline grafted vanadyl acetylacetonate [polyaniline-
over metal–acid catalysts. In 2002, Halliday and co-workers first VO(acac)2] catalyst could produce an HMF conversion of 99.2%
studied the conversion of fructose into DFF using a two-step and DFF yield of 86.2% after 12 h at 383 K and 1 bar O2.211
method.209,210 Fructose initially dehydrated to HMF in DMSO at The one-pot method by the combined use of Amberlyst 15
383 K to produce a 85% HMF yield over an acidic ion-exchange and polyaniline-VO(acac)2 catalysts produced DFF with a yield
BioRad AG 50W-X8 resin. After the filtration of the acid catalyst, of 42% after 12 h at 383 K and 1 bar O2 (Table 9, entry 3).

Fig. 27 Synthesis of DFF from glucose and fructose via a one-pot reaction.

Table 9 Catalytic conversion of carbohydrates into DFF in batch reactors

Entry Catalyst Substrates T (K) Solvent Method Con. (%) Yield (%) Ref.
+
1 VOPO4, H -exchange resin Fructose 423 DMSO Two-step — 52 209
2 V2O5, H+-exchange resin Fructose 423 DMSO Two-step — 77 209
3 Amberlyst-15, polyaniline-VO(acac)2 Fructose 383 DMSO One-step 100 42 211
4-Chlorotoluene
4 Amberlyst-15, polyaniline-VO(acac)2 Fructose 383 DMSO Two-step 100 71 211
4-Chlorotoluene
5 V-g-C3N4(H+) Fructose 403 DMSO One-step 99 45 212
6 g-C3N4(H+), V-g-C3N4(H+) Fructose 403 DMSO Two-step 100 63 212
7 Cs2HPMo11VO40 Fructose 383 DMSO Two-step 99 60 213
8 Cs0.5H2.5PMo12 Fructose 433 DMSO One-step — 69 214
9 WO3HO-VO-@Fe3O4 Fructose 353, 298 Isopropanol Two-step 71 215
10 Fe3O4–SBA–SO3H, K-OMS-2 Fructose 373 DMSO Two-step 499 80 192
11 Amberlyst-15, Ru/HT Fructose 373 DMF Two-step 499 49 193
12 Amberlyst-15, HT, Ru/HT Glucose 373 DMF Three-step 98 25 183
13 Amberlyst-15, HT, Ru/HT Raffinose 393 DMF Three-step 100 27 183
14 CrCl3/NaBr//NaVO3 Glucose 383 DMF Two-step — 55 216
15 g-Fe2O3@HAP-Ru; Fe3O4@SiO2–SO3H Fructose 383 DMSO Two-step 100 79 217
4-Chlorotoluene
16 GO Fructose 413 DMSO One-step — 53 220
17 GO Fructose 413 DMSO Two-step — 73 220
18 CC–SO3H–NH2 Glucose 413 DMSO Two-step — 44 222

1376 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

A two-step method was then used to further improve the yield. The the one-step method produced both HMF and DFF in less than
dehydration of fructose over the Amberlyst 15 catalyst produced a 5% yields at full fructose conversion, which was possibly
91.8% HMF yield after 1.5 h at 383 K. After the removal of the caused by the oxidation of fructose. Ebitani and co-workers
Amberlyst 15 catalyst, the oxidation of HMF over the [polyaniline- performed the conversion of fructose into DFF by the combined
VO(acac)2] catalyst produced DFF in an overall yield of 71% use of Amberlyst-15 and Ru/HT catalysts.183 The one-step
after 12 h at 383 K and 1 atm O2 (Table 9, entries 3 vs. 4). method produced a DFF yield of 13% and HMF yield of 44%
The above two examples required the use of one acid catalyst at fructose conversion 499% after 9 h at 373 K and 1 bar O2.
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

and one metal catalyst. Chen and co-workers prepared proto- The DFF yield increased to 49% using a two-step method
nated vanadium-doped graphitic carbon nitride [V-g-C3N4(H+)] (Table 9, entry 11), including fructose dehydration at 373 K
as a bifunctional catalyst for the oxidative conversion of fructose for 3 h and the subsequent HMF oxidation at 373 K for 6 h. The
into DFF in DMSO.212 The V-g-C3N4(H+) catalyst was prepared by one-pot conversion of glucose into DFF was performed by the
the thermal pyrolysis of dicyandiamide with ammonium meta- use of HT, Amberlyst-15 and Ru/HT catalysts, in which HT
vanadate (NH4VO3) at 813 K for 4 h, followed by treatment with catalyzed the isomerization of glucose into fructose. As shown
HCl. A 99% fructose conversion and a 45% DFF yield were in Fig. 28, the three-step method produced an overall DFF yield
produced after 8 h at 403 K and 1 bar O2 in one-step (Table 9, of 25% (Table 9, entry 12). Furthermore, DFF was produced in a
entry 5). However, the two-step method produced an overall DFF yield of 27% from naturally occurring indigestible trisaccharide
yield of 63%, including the first dehydration step at 393 K for raffinose (galactose, glucose, and fructose units) by the same
2 h and the subsequent oxidation step at 403 K for 6 h (Table 9, method (Table 9, entry 13).216 Hu and co-workers improved the
entry 6). The lower DFF yield by the one-step method was due to DFF yield to 55% from glucose using a two-step method over
the possible oxidation of fructose to undesired byproducts.150 the CrCl3/NaBr/NaVO3 catalysts (Table 9, entry 14).217 Glucose
V-g-C3N4(H+) was not stable due to the leaching of H+, but the was first converted to HMF in DMF with a yield of 65% at 383 K
V-g-C3N4(H+) catalyst could be regenerated by the treatment for 6 h by CrCl3/NaBr, and then the oxidation of HMF was
with HCl. The vanadium-containing heteropolyacids were also performed at 383 K and 1 bar O2 for 10 h over the NaVO32H2O
used as bifunctional catalysts for the transformation of fructose catalyst. Similarly, Lin and co-workers also performed the one-
into DFF.213,214 The one-step conversion of fructose over the pot conversion of glucose into DFF in DMF over the AlCl3/NaBr/
Cs2HPMo11VO40 catalyst produced DFF in a yield of 26% in vanadium catalysts, which produced DFF yields of 34–48% with
DMSO after 6 h at 393 K and 1 bar O2.213 The two-step approach
increased the DFF yield to 60% by the first dehydration
of fructose at 383 K for 2 h, followed by oxidation of HMF at
393 K and 1 bar O2 for 6 h (Table 9, entry 7).213 The catalyst
deactivated (60% in the first run vs. 53% in the fourth run),
due to the leaching of vanadium, and the deposition of
insoluble polymeric furan compounds on the catalyst surface.
Cs0.5H2.5PMo12O40 produced a higher DFF yield of 69% from
fructose than Cs2HPMo11VO40 (Table 9, entry 8).214 The DFF
yield declined slightly in the first two runs, then remained
stable after the third cycle. The XRD patterns of the spent
catalyst were the same as the fresh catalyst. To facilitate the
recycling of the bifunctional catalysts, a magnetic bifunctional
WO3HO-VO@Fe3O4 nanocatalyst was prepared and tested for
the synthesis of DFF from fructose.215 An 82% HMF yield was
obtained with the tungstic acid-mediated fructose dehydration
at 353 K within 1 h. After cooling to room temperature, the
oxidation of HMF was performed by the addition of H2O2,
affording 71% DFF yield after 15 h (Table 9, entry 9). The
magnetic WO3HO-VO@Fe3O4 catalyst demonstrated a high
stability with about 5% decrease of DFF yield after the fifth run.
Besides V catalysts, Mn and Ru catalysts were also used
for the oxidative transformation of fructose into DFF. Fu and
co-workers performed the two-step synthesis of DFF from
fructose by the combined use of Fe3O4–SBA–SO3H and K-OMS-2 Fig. 28 Glucose conversion into fructose (’), HMF (n) and DFF (K). HT
catalysts.192 The first step of fructose dehydration of HMF in was initially added into the reactor. The Amberlyst-15 catalyst was then
added after 3 h. The Ru/HT catalyst was further added after 6 h under
DMSO produced HMF with a yield of 80% after 2 h at 383 K over
oxygen flow (20 mL min 1). Reaction conditions: glucose (0.1 g), HT (0.2 g),
the Fe3O4-SBA-SO3H catalyst. The HMF was quantitatively con- Amberlyst-15 (0.1 g), Ru/HT (0.2 g), DMF (3 mL), 373 K. Reproduced with
verted into DFF over the K-OMS-2 catalyst within 6 h at 383 K and permission from ref. 183 with permission from American Chemical Society,
1 bar O2 (Table 9, entry 10). In contrast to the two-step method, copyright 2011.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1377
View Article Online

Review Article Chem Soc Rev

almost complete glucose conversion.218 In order to recycle the catalytic systems are mainly based on the conventional metal
catalysts, two magnetic g-Fe2O3@HAP-Ru and Fe3O4@SiO2-SO3H bromide catalysts (Co/Mn/Br) that are used for the oxidation of
catalysts were used for the synthesis of DFF from fructose.219 para-xylene to terephthalic acid.230 In 2001, Partenheimer and
g-Fe2O3 was firstly encapsulated by hydroxyapatite (HAP), and Grushin performed the oxidation of HMF in acetic acid over the
then Ca2+ in HAP was exchanged with Ru3+ to generate the Co(OAc)2/HBr/Mn(OAc)2 catalysts, producing 60.9% yield of
g-Fe2O3@HAP-Ru catalyst. Firstly, the dehydration of fructose FDCA after 3 h at 398 K and 70 bar air pressure.165 Other
produced HMF with a yield of 90.1% after 2.5 h at 383 K over the byproducts were HMF acetate and the decarbonylation of HMF.
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

Fe3O4@SiO2–SO3H catalyst. After the removal of the Fe3O4@ Later, Subramaniam and co-workers improved the FDCA yield
SiO2–SO3H catalyst, the oxidation of HMF produced DFF with to 90% at a 1/0.015/0.5 molar ratio of Co, Mn, and Br, 7% (v/v)
an overall yield of 79% after 24 h at 383 K and 1 bar O2 over the water, 30 bar (CO2/O2 = 1/1, mol/mol), and 453 K.231 Similar
g-Fe2O3@HAP-Ru catalyst (Table 9, entry 15). work was also reported for the oxidation of HMF over the
Hou and co-workers reported a metal-free transformation of Co(OAc)2/Zn(OAc)2/NaBr catalysts.232 This method produced
fructose into FDCA.220 Graphene oxide (GO) was used as a FDCA with a yield of 60% in acetic acid at an HMF conversion
bifunctional and metal-free catalyst for the transformation of of 90% after 3 h at 363 K and 1 bar O2. Unlike the Co(OAc)2/
fructose into DFF. The oxidation of HMF produced a 90% HBr/Mn(OAc)2 catalytic system,165 FFCA was also produced in a
DFF yield in DMSO after 24 h at 413 K with an O2 flow rate of yield of 29% over the Co(OAc)2/Zn(OAc)2/NaBr catalysts. Due
20 mL min 1. DMSO was very crucial for this catalytic system, to the difficulty in the recycling of homogeneous catalysts,
as the same reaction in DMF could not produce any oxidation extensive attention has been paid by researchers to developing
products. The oxidation of HMF was attributed to the in situ heterogeneous catalytic systems.
formation of the active component from DMSO at high As discussed in Section 7.2, the oxidation of HMF into DFF
temperature, and the reaction pathway involved the proton- was mainly catalyzed by V, Ru and Mn catalysts in organic
ation of the alcohol group of HMF and a nucleophilic addition solvents without a base. In contrast to the oxidation of HMF
of DMSO, as reported in the literature.221 The reaction under a into DFF, the oxidation of HMF into FDCA is mainly performed
N2 atmosphere produced less than 2% HMF conversion indi- over noble-metal (Pd, Pt and Au) catalysts in an alkaline
cating that the real oxidant was O2 not DMSO in this catalytic solution (Table 10). In 1983, Verdeguer et al. studied the
system. The one-step method produced DFF in a yield of 53% oxidation of HMF into FDCA over a Pt/C catalyst.233 FDCA
after 24 h, whereas the two consecutive methods (the first step was produced in a yield of 81% at full HMF conversion after
under a N2 atmosphere for 2 h and the second step under an O2 2 h at 298 K (Table 10, entry 1). The FDCA yield increased to
atmosphere for 22 h) improved the DFF yield to 73%, based on 99% over a Pt–Pb/C catalyst (Table 10, entry 2). Kinetic studies
fructose (Table 9, entries 16 and 17). Jadhav et al. further indicated that the formyl group was oxidized first, followed by
developed a similar catalytic system for the one-pot conversion the hydroxymethyl group. The HMF conversions were 100% after
of glucose to DFF in DMSO by the use of a carbon material 1 h at 373 K and 40 bar air with Na2CO3/HMF molar ratio = 2 over
functionalized with acid–base sites (CC–SO3H–NH2).222 The both the Pt/C and Pt–Bi/C catalysts, but the FDCA yields were
base sites acted as the site for the isomerization of glucose 51% and 83% for the Pt/C and Pt–Bi/C catalysts, respectively. Bi
into fructose. The two-step method (the first step under N2 was claimed to increase the resistance of the Pt/C catalyst to
atmosphere for 15 h and the second step under an O2 atmo- oxygen poisoning. After 4 h, FDCA was produced in a nearly
sphere for 27 h) produced a 44% DFF yield from glucose at quantitative yield (499%) over the Pt–Bi/C catalyst (Table 10,
413 K (Table 9, entry 18). The CC–SO3H–NH2 catalyst could be entry 3). Observed initial intermediates were HFCA and DFF,
reused five times, maintaining its activity. which were rapidly oxidized to FFCA (the main intermediate).
The FDCA yield slowly decreased from 100% to 90% in the fifth
7.5 Catalytic oxidation of HMF into FDCA with O2 as the run with FFCA as the byproduct over the Pt–Bi/C catalyst. Davis
oxidant and co-workers studied the catalytic performance of the Pt/C,
FDCA is another important HMF oxidation product, which is Pd/C and Au/C catalysts for HMF oxidation in 0.3 M NaOH
produced by the simultaneous oxidation of the methylhydroxyl solution at 296 K and 6.90 bar O2 pressure.156 The HMF conver-
and formyl group in HMF. It has been identified by the US sion reached 100% after 6 h for the three catalysts with FDCA
Department of Energy as one of 12 priority chemicals for and HFCA as the oxidation products. The FDCA yields were 79%
establishing the ‘‘bridge’’ between fossil fuel resources and and 71% for the Pt/C and Pd/C catalysts, respectively (Table 10,
biomass resources.223 FDCA can be used as a replacement for entries 4 and 5), while 92% yield of HFCA was produced for the
terephthalic acid, in various polyesters.224,225 Avantium announced Au/C catalyst (Table 10, entry 6), Strasser et al. also observed that
plans to produce 40 ton per year of FDCA monomer at a pilot plant the Pt/C catalyst showed higher catalytic activity than the Pd/C
in 2011 and 30 000 to 50 000 ton per year in 2016.226 catalyst.235 Besides the Pt/C catalyst, reduced graphene oxide
Oxidation of HMF into FDCA was first carried out with (RGO) was also used to stabilize Pt nanoparticles (Pt/RGO) for
stoichiometric oxidants such as KMnO4, N2O4 and HNO3.227–229 the oxidation of HMF at 298 K.236 FDCA was produced in a yield
The catalytic oxidation of HMF into FDCA with green oxidants of 84% with an HFCA yield of 16% over the Pt/RGO catalyst
particularly molecular oxygen has been extensively examined with after 24 h with NaOH/HMF molar ratio = 5 and 1 bar O2
both homogeneous and heterogeneous catalysts. The homogeneous (Table 10, entry 7). The Pt/RGO catalyst could be reused with

1378 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

Table 10 Catalytic oxidation of HMF into FDCA in batch reactorsa

Entry Catalyst Catalyst ratioa Oxidant and base T (K) Time (h) Con. (%) Yield (%) Ref.
1 Pt/C 100 1 bar O2, 1.25 M NaOH 298 2 100 81 233
2 Pt–Pb/C 100 1 bar O2, 1.25 M NaOH 298 2 100 99 233
3 Pt–Bi/C 100 40 bar air, 2 equiv. Na2CO3 373 4 100 499 234
4 Pt/C 150 6.9 bar O2, 2 equiv. NaOH 296 6 100 79 156
5 Pd/C 150 6.9 bar O2, 2 equiv. NaOH 296 6 100 71 156
6 Au/C 150 6.9 bar O2, 2 equiv. NaOH 296 6 100 7 156
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

7 Pt/RGO 39 1 bar O2, 5 equiv. NaOH 298 24 100 84 236


8 Fe3O4@C/Pt 27 1 bar O2, 1.67 equiv. Na2CO3 363 4 100 84 237
9 Pt/C-EDA-4.1 49 10 bar O2 373 12 100 96 238
10 Pt/ZrO2 100 10 bar air, 8.3 equiv. Na2CO3 373 12 100 98 239
11 Pt–Bi/TiO2 100 40 bar air, 2 equiv. Na2CO3 373 10 100 499 240
12 Pt/Al2O3 100 1 bar O2, 1 equiv. Na2CO3 413 12 96 96 243
13 Pt/PVP 19 1 bar O2 353 24 100 94 244
14 Pt/IP 19 1 bar O2 353 24 100 99 244
15 Pt/C–O–Mg 50 10 bar O2 383 12 100 97 245
16 Pd/PVP 100 1 bar O2, 35 mL min 1, 1.25 equiv. NaOH 363 6 499 90 246
17 Pd/ZrO2/La2O3 100 1 bar O2, 35 mL min 1, 1.25 equiv. NaOH 363 8 499 90 247
18 Pd/Al2O3 100 1 bar O2, 35 mL min 1, 1.25 equiv. NaOH 363 8 499 78 247
19 Pd/TiO2 100 1 bar O2, 35 mL min 1, 1.25 equiv. NaOH 363 8 499 53 247
20 Pd/HT 21.2 1 bar O2, 35 mL min 1 373 7 499 499 248
21 g-Fe2O3@HAP-Pd 43.1 1 bar O2, 0.5 equiv. K2CO3 373 6 97 92.9 249
22 C–Fe3O4–Pd 55 1 bar O2, 0.5 equiv. K2CO3 353 4 98.1 91.8 250
23 Pd/C@Fe3O4 28 1 bar O2, 0.5 equiv. K2CO3 353 6 98.4 86.7 251
24 Au/TiO2 100 20 bar O2, 20 equiv. NaOH 303 18 100 71 155
25 Au/CeO2 150 10 bar Air, 4 equiv. NaOH 403 5 100 96 255
26 Au/TiO2 150 10 bar Air, 4 equiv. NaOH 403 5 100 84 255
27 Au/m-CeO2 100 10 bar O2, 4 equiv. NaOH 343 4 100 92 257
28 Au/Ce0.85Bi0.15O2 d 150 10 bar O2, 4 equiv. NaOH 338 2 100 100 258
29 Au/HT 40 1 bar O2 368 7 499 499 259
30 Au/HY 100 3 bar O2, 5 equiv. NaOH 333 6 499 499 261
31 Au–Cu/TiO2 100 10 bar O2, 4 equiv. NaOH 368 4 100 99 262
32 Au8–Pd2/C 200 10 bar O2, 2 equiv. NaOH 333 6 499 499 263
33 Au/C 200 10 bar O2, 2 equiv. NaOH 333 6 499 80 263
34 Pd/C 200 10 bar O2, 2 equiv. NaOH 333 6 499 9 263
35 Au–Pd/CNT 100 5 bar O2 373 12 100 94 265
36 Pt/CNT 100 5 bar O2 368 14 100 98 267
37 Ru/AC 100 40 bar air, 4 equiv. NaHCO3 373 2 100 75 268
38 Ru/AC-NaOCl 100 40 bar air, 4 equiv. NaHCO3 373 4 100 55 268
39 Ru/MnCo2O4 33.6 24 bar air 393 24 100 99.1 269
40b Mn0.75Fe0.25Ox 50 8 bar O2, 4 equiv. NaOH 363 24 100 90 271
41b NNC-873 317.5 1 bar O2, 100 mL min 1 353 48 100 11 274
3 equiv. K2CO3
42b NNC-973 317.5 1 bar O2, 100 mL min 1 353 48 100 35 274
3 equiv. K2CO3
b
43 NNC-1073 317.5 1 bar O2, 100 mL min 1 353 48 100 72 274
3 equiv. K2CO3
44b NNC-1173 317.5 1 bar O2, 100 mL min 1 353 48 100 80 274
3 equiv. K2CO3
a b
For entries 1 to 39, the catalyst ratio was defined as the molar ratio of HMF to the metals. For entries 40–44, it was defined as the weight ratio of
HMF to the catalyst.

full HMF conversion in each run. However, the FDCA yield 4.1 wt% nitrogen showed the highest activity, producing FDCA
showed a slight decrease with a slight increase in the HFCA yield. with a yield of 96.0% after 12 h (Table 10, entry 9). XPS
To facilitate catalyst recycling, a magnetic core–shell structured indicated that graphene-type nitrogen, pyrrole-type nitrogen
Fe3O4@C supported Pt catalyst was prepared for HMF oxidation, and pyridine-type nitrogen were present in the catalysts. The
producing a quantitative FDCA yield after 4 h at 363 K (Table 10, FDCA yields were proportional to the amount of graphene-type
entry 8).237 The Fe3O4@C/Pt catalyst could be reused at least three nitrogen. Thus, graphene-type nitrogen played a key role in the
times without significant performance loss. The above carbon selective oxidation of HMF. However, the Pt/C-EDA-4.1 catalyst
based Pt catalysts were used in alkaline solution for the oxidation was not stable due to growth of Pt nanoparticles.
of HMF into FDCA. A catalyst consisting of Pt supported on a Metal oxide supported Pt catalysts were also studied for the
carbon material with the introduction of nitrogen atoms (basic oxidation of HMF into FDCA. The Pt/ZrO2 catalyst produced a
sites) (Pt/C-EDA) could promote the base-free oxidation of HMF 98% FDCA yield at 373 K and 10 bar air (Table 10, entry 10).239
into FDCA.238 The Pt/C-EDA catalyst was prepared by first hydro- Besson and co-workers compared the catalytic activity of
thermal treatment of glucose, resorcinol, and ethylenediamine Pt/TiO2 with Pt/ZrO2.240 Pt/TiO2 showed a higher catalytic activity
(EDA), and second pyrolysis. The Pt/C-EDA-4.1 catalyst with than Pt/ZrO2. Moreover, the addition of bismuth (Bi) to Pt/TiO2

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1379
View Article Online

Review Article Chem Soc Rev

improved the catalyst activity and stability, similar to the conversion over the Pd/HT catalyst. Several kinds of magnetic
Pt–Bi/C catalyst.234 Quantitative oxidation of HMF into FDCA Pd catalysts were studied for the oxidation of HMF into FDCA
was obtained after 10 h at 373 K and 40 bar air over the Pt–Bi/ to facilitate catalyst recycling.249–251 Similar to the g-Fe2O3@
TiO2 catalyst (Bi/Pt = 0.22) (Table 10, entry 11). The Pt–Bi/TiO2 HAP-Ru3+ catalyst,219 a magnetic g-Fe2O3@HAP-Pd catalyst was
catalyst was stable for four runs. The Pt/Al2O3 catalyst also showed prepared.249 An HMF conversion of 97% and FDCA yield of
good catalytic performance for the oxidation of HMF into 92.9% were produced after 6 h at 373 K and 1 bar O2 over the
FDCA.241–243 For example, Dhepe and co-workers reported that g-Fe2O3@HAP-Pd catalyst (Table 10, entry 21). This catalytic
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

FDCA was produced in a yield of 96% after 12 h at 413 K and 1 bar system did not require the use of excessive base (with 0.5 equiv.
O2, but at a high Pt loading (Table 10, entry 12).243 The HMF of K2CO3 to neutralize FDCA). The catalyst could be reused (a
conversion remained at 100% for each run by the reduction of the 92.9% FDCA yield in the first run vs. 90.7% in the fifth run)
spent catalyst with H2. However, a small decrease in the FDCA without the growth of Pd nanoparticles. Graphene oxide with
yield was still observed. abundant oxygen functional groups has received great interest
PVP and poly-1,3-bis(4-vinylbenzyl)imidazolium chloride to immobilize nanoparticles.252,253 A magnetically separable
(abbreviated as IP) stabilized Pt nanoparticles (abbreviated as graphene oxide supported Pd catalyst (C-Fe3O4-Pd) was also
Pt/PVP, Pt/IP) were active for the base-free oxidation of HMF used for the oxidation of HMF into FDCA, which was prepared
into FDCA.244 DFF and HFCA were the intermediates, and FFCA by the simultaneous deposition of Fe3O4 and Pd nanoparticles
was the main intermediate, suggesting that the oxidation of on graphene oxide.250 A 98.1% HMF conversion and a 91.8%
the hydroxymethyl group was easier than the oxidation of the FDCA yield were produced after 4 h at 353 K and 1 bar O2 with
formyl group. FDCA was produced in yields of 94% and 99% 0.5 equiv. K2CO3 (Table 10, entry 22). Similarly, a core–shell
after 24 h over the Pt/PVP and Pt/IP catalysts, respectively structured C@Fe3O4 supported Pd catalyst was prepared for the
(Table 10, entries 13 and 14). The Pt/IP catalyst could be oxidation of HMF into FDCA.251 An HMF conversion of 100%
recycled with a full HMF conversion in each run, but the FDCA and FDCA yield of 86.7% were produced over the Pd/C@Fe3O4
yield decreased to nearly 87% with an FFCA yield of 13% in catalyst after 6 h at 353 K (Table 10, entry 23). The Pd/C@Fe3O4
the sixth run. The Pt/C–O–Mg catalyst with a basic support catalyst could be reused with a very slight decrease in FDCA
also promoted the base-free oxidation of HMF into FDCA. The yield (86.7% in the first run vs. 83.7% in the fifth run). XPS
C–O–Mg support was prepared by coating MgO particles with confirmed that metallic Pd(0) was retained, and no aggregation
carbon and then removing the free MgO before loading Pt onto of Pd nanoparticles was observed in the spent catalyst.
the carbon surface.245 A 97% FDCA yield was produced after 12 h Compared with Pd and Pt catalysts, Au catalysts have
at 383 K and 10 bar O2 (Table 10, entry 15). More importantly, the received much more attention for the oxidation of HMF into
Pt/C–O–Mg catalyst could be reused ten times. The excellent FDCA. Riisager and co-workers studied the oxidation of HMF
activity and stability may come from the formation of a C–O–Mg into FDCA over a commercial Au/TiO2 catalyst.155 HFCA was the
bond, which created new, strong and stable basic sites. Further intermediate. A 71% FDCA yield was produced at full HMF
scaling up the HMF amount to 10 mmol, the isolated FDCA yield conversion after 18 h at 303 K and 20 bar O2 with 20 equiv. of
reached 74.9% with very high purity (99.5%). NaOH (Table 10, entry 24). Without a base, the deactivation of
Several kinds of Pd catalysts were also active for the oxida- the Au/TiO2 catalyst was caused by the absorption of FDCA on
tion of HMF into FDCA. PVP stabilized Pd nanoparticles the gold surface. NaOH reacted with FDCA to form a salt in the
(Pd/PVP) with a smaller particle size produced a higher FDCA reaction solution. Similar phenomena were also observed by
yield.246 The Pd/PVP catalyst with a Pd diameter of 1.8 nm pro- Davis et al. using the same catalyst.254 The activity of the Au
duced a 90% FDCA yield after 6 h at 363 K (Table 10, entry 16), catalysts was greatly affected by the support.255 Under identical
while it was 81% with a Pd diameter of 2.0 nm. Furthermore, Pd/ conditions (10 bar air, 403 K, 4 equiv. NaOH), FDCA was produced
PVP deposited on metal oxides (TiO2, g-Al2O3 and ZrO2/La2O3) was in a yield of 84% at full HMF conversion after 8 h over the Au/TiO2
preprared.247 Full HMF conversions (499%) were achieved for catalyst, while the yield was 100% after 5 h over the Au/CeO2
different catalysts after 8 h at 363 K, but Pd/ZrO2/La2O3 produced catalyst (Table 10, entries 25 and 26). The lower selectivity of FDCA
the highest FDCA yield of 90% (Table 10, entries 17–19). No over the Au/TiO2 catalyst was caused by the formation of decar-
significant aggregation of Pd nanoparticles was observed in the boxylated and ring-opening byproducts from HMF. Albonetti et al.
spent Pd/ZrO2/La2O3 catalyst, while serious aggregation of Pd also observed that the Au/CeO2 catalyst showed a better catalytic
nanoparticles was noted in other catalysts. The Pd/ZrO2/La2O3 performance than the Au/TiO2 catalyst.256 Furthermore, they
was recycled for three runs with a full HMF conversion in prepared an ordered mesoporous CeO2 supported Au catalyst
each run, but the FDCA yield decrease from 90% in the first run (Au/mCeO2) for the oxidation of HMF into FDCA.257 HFCA was
to 86% in the third run. Similar to the Pt/C–O–Mg catalyst the main product within 30 min, and FDCA was produced in a yield
with a basic support,245 the Pd/HT catalyst with the basic of 92% after 4 h at full HMF conversion at 343 K and 10 bar O2
Mg–Al–hydrotalcite (HT) support was also active for the base- (Table 10, entry 27). Yang and co-workers found that the catalytic
free oxidation of HMF into FDCA.248 A quantitative FDCA yield activity of Au/CeO2 could be enhanced by the addition of Bi into
was produced in water after 7 h at 373 K with HFCA and FFCA CeO2.258 Full HMF conversions were attained after 1 h at 338 K and
as the intermediates (Table 10, entry 20). A slight decrease in 10 bar O2 over the Au/CeO2 and Au/Ce1 xBixO2 d (0.08 r x r 0.2)
FDCA yield was observed after the 5th run with full HMF catalysts. The FDCA and HFCA yields were 39% and 57% over

1380 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

the Au/CeO2 catalyst, respectively. While FDCA yields around catalyst produced an 80% FDCA yield together with a 20%
75% were produced over the Au/Ce1 xBixO2 d catalysts. 100% HFCA yield after 6 h (Table 10, entry 33), while HFCA was
yield of FDCA was obtained after 2 h over the Au/Ce0.85Bi0.15O2 d produced in a 91% yield over the Pd/C catalyst after 6 h
catalyst (Table 10, entry 28). The higher activity of the Au/ (Table 10, entry 34). These results indicated that Pd and Au
Ce1 xBixO2 d catalysts was attributed to the oxygen activation synergetically promoted the oxidation of HMF into FDCA. The
and hydride transfer enhanced by Bi doping and the large Au8–Pd2/C catalyst could be reused for five runs with the FDCA
amount of oxygen vacancies. The Au/Ce0.85Bi0.15O2 d catalyst yield remaining 499%. The synergetic effects of Pd and Au
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

could be reused with full HMF conversions for four runs. were present in the Au–Pd/TiO2 for the oxidation of HMF into
However, the FDCA yield showed a slight decrease in conjunc- FDCA.264 For most cases, the oxidation of HMF into FDCA over
tion with a slight increase of the HFCA yield. Similar to the Au catalysts was performed by the use of an excessive base.
Pd/HT catalyst,248 the 1.92 wt%Au/HT catalyst could also promote A carbon nanotube (CNT) supported Au–Pd (Au–Pd/CNT) catalyst
the base-free oxidation of HMF into FDCA.259 A quantitative FDCA was active for the base-free oxidation of HMF to FDCA.265 A 94%
yield was produced with HFCA as the main intermediate after 7 h FDCA yield was produced over the Au–Pd/CNT catalyst after 12 h
at 368 K and 1 bar O2 (Table 10, entry 29). Although HMF at 373 K and 5 bar O2 (Table 10, entry 35). The treatment of CNT
conversion remained 100% for three runs, the FDCA yield showed with H2O2 generated carbonyl/quinone and phenol groups on
a slight decrease in conjunction with a slight increase of the HFCA its surfaces, facilitating the adsorption of HMF and the inter-
and FFCA yields. The spent Au/HT catalyst showed no change in mediates on the catalyst surface. The CNT supported Pt catalyst
the oxidation state, morphology and particle size distribution of (Pt/CNT) was also active for the base-free oxidation of HMF into
Au nanoparticles. Wilson and co-workers260 reported that the FDCA.267 A 98% FDCA yield was produced after 14 h at 368 K and
alkali-free HT supported Au catalyst (2.0 wt% Au-HT) produced 5 bar O2 (Table 10, entry 36). The Pt/CNT catalyst was very stable
80% HMF conversion with HFCA as the product in the absence without a decrease in FDCA yield.
of a base. The competitive adsorption between strongly-bound As listed in Table 8, Ru catalysts were mainly used for
HMF and reactively-formed oxidation intermediates site-blocks the oxidation of HMF into DFF in organic solvents. Kerdi and
gold. Aqueous NaOH dramatically promotes solution phase co-workers reported that the active carbon supported Ru
HMF activation, liberating free gold sites able to activate the catalyst (Ru/AC) could catalyze the oxidation of HMF into FDCA
alcohol function within the meta-stable HFCA reactive inter- in alkaline solution via DFF as the intermediate.268 The intro-
mediate. Thus the authors claimed that the base-free oxidation duction of oxygen-functional groups on the surface of active
of HMF into FDCA over the Au/HT catalyst reported by Ebitani carbon by NaOCl resulted in a decrease of the catalytic activity.
and co-workers259 was caused either by alkali contaminants Full HMF conversion was attained after 2 h over the Ru/AC
leached from Na2CO3/NaOH precipitated hydrotalcites, or par- catalyst with a 75% FDCA yield, while 4 h was required for the
tially soluble brucite co-existing with the hydrotalcite. The con- Ru/AC-NaOCl catalyst to get a full HMF conversion and a 55%
finement of Au nanoparticles in a rigid cage was an effective way FDCA yield (Table 10, entries 37 vs. 38). The oxygen functional
to prevent Au nanoparticle aggregation. Xu et al. confined Au groups made the catalyst surface more hydrophilic and more
nanoparticles of 1 nm into a supercage of Y zeolite (Au/HY) for polar, facilitating the preferential adsorption of water compara-
HMF oxidation.261 An FDCA yield 499% was produced after 6 h at tively to the substrates. A MnCo2O4 spinel supported Ru catalyst
333 K (Table 10, entry 28). The HY supercage serves as a natural (Ru/MnCo2O4) was reported to be active for the base-free
nanocage to encapsulate the extremely small Au nanoclusters to oxidation of HMF into FDCA, which was due to the presence
prevent them from aggregation. The particle size of Au was of Lewis and Brønsted acids in the support.269 Although a high
essentially unaltered in the spent catalyst. However, the FDCA FDCA yield of 99.1% was attained in water at 393 K under 24 bar
yield decreased slightly from 99% in the first run to 88% in the air, the use of high Ru loading (HMF/Ru molar ratio = 33.6)
fourth run. makes this method high cost for practical applications (Table 10,
Some Au containing bimetallic catalysts were also studied entry 39).269 Some recent progress has been made for the non-
for HMF oxidation. Pasini et al. studied the oxidation of HMF noble metal or metal free oxidation of HMF into FDCA. Amos
into FDCA over Cu–Au/TiO2 catalysts.262 Au–Cu/TiO2 catalysts Mugweru et al. reported a mixed oxide of Li2CoMn3O8 with NaBr
with different Au/Cu ratio showed higher activity by at least a as the co-catalyst for the oxidation of HMF into FDCA.270 An
factor of 2 in each case as compared to the Au/TiO2 catalyst at the isolated FDCA yield of 80% was produced on a gram scale, but
same weight percentage. The enhanced catalytic activity was the reaction was performed in acetic acid under harsh condi-
attributed to the isolation of Au sites by Cu. The Au–Cu/TiO2 tions (423 K, 55 bar O2). Neatu et al. performed the oxidation of
catalyst with the molar ratio Au/Cu = 1 showed the highest HMF into FDCA by the use of MnxFey mixed oxides.271 MnxFey
catalytic activity, which afforded a 99% FDCA yield at 368 K and mixed oxides were prepared by the co-precipitation of Fe3+ and
10 bar O2 (Table 10, entry 31). This catalyst was stable without the Mn2+ at pH = 9–10, followed by calcination at 773 K in air. The
leaching and aggregation of the metal nanoparticles. Supported Mn0.75Fe0.25 mixed oxide (Mn0.75Fe0.25Ox) showed the highest
bimetallic Pd–Au catalysts also showed better catalytic perfor- catalytic activity. The authors found that the oxidation of HMF
mance than the corresponding Au catalysts.263–266 For example, into FFCA could be performed without a catalyst. A 79% HMF
an FDCA yield 499% was produced after 4 h at 333 K over the conversion and an 88% FFCA selectivity were produced after
Au8–Pd2/C catalyst (Table 10, entry 32).263 However, the Au/C 12 h at 363 K and 8 bar O2 with NaOH/HMF molar ratio = 1 : 4.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1381
View Article Online

Review Article Chem Soc Rev

However, the humins from this step caused the deactivation of


the catalyst. Therefore, a two-step method was adopted. After
the first step, humin species were removed using a simple
precipitation/filtration method by adjusting the pH to 1. Then
the oxidation of the purified FFCA was carried out over the
Mn0.75Fe0.25Ox catalyst, which gave a 90% FDCA yield after 24 h
(Table 10, entry 40). In addition, other kinds of Mn oxides such
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

as MnOx–CeO2 and MnO2 were also used for the oxidation of


HMF into FDCA under basic conditions.272,273 More excitingly,
the metal-free oxidation of HMF into FDCA was recently
reported.274 In recent years, nitrogen-doped carbon catalysts
have been discovered to be active for several types of chemical
reactions.275 Wu and co-workers reported a first example on the Fig. 29 Overall reaction scheme and proposed mechanism for the oxi-
metal-free oxidation of HMF into FDCA by the use of nitrogen- dation of HMF in aqueous solution. Dioxygen (not shown) serves as a
doped carbon catalysts (abbreviated as NNC-X, where X repre- scavenger of electrons that are deposited into the metal particles during
the catalytic cycle. Adapted from ref. 276 with permission from Royal
sents the pyrolysis temperature). NNC catalysts were prepared
Society of Chemistry copyright 2012.
by the pyrolysis of ZIF-8 (one-kind of zeolitic imidazole frame-
work), followed by treatment with HCl to remove the zinc
species.274 Full HMF conversions were all observed for the 7.6 Catalytic oxidation of HMF into FDCA using t-BuOOH and
NNC-X catalysts. FDCA yields of 11%, 35%, 72%, and 80% were H2O2 as the oxidant
observed in water after 48 h at 353 K over NNC-873, NNC-973,
NNC-1073, and NNC-1173 as the catalysts (Table 10, entries 41–44), Besides O2, there are also some reports on the oxidation of HMF
respectively. The FDCA yields were proportional to the amount into FDCA with t-BuOOH or H2O2 as the oxidant. Riisager et al.
of graphene-type nitrogen in the NNC-X catalysts, suggesting performed the oxidation of HMF into FDCA in acetonitrile with
that the graphene-type nitrogen atoms were the active sites for t-BuOOH as the oxidant and copper salts as the catalysts.202
the activation of O2 to give oxidative species of oxygen radicals. A 43% FDCA yield was produced after 48 h at 25 1C with the
However, the NNC-1173 catalyst was not stable with a decrease CuCl–LiBr catalysts, and that was 45% with CuCl2. Not only are
in FDCA yield from 80% to 70%, due to a decrease in the the FDCA yields lower, but it is also difficult to recycle the
amount of graphene-type nitrogen. homogeneous catalysts. The merrifield resin supported Co(II)-
Little attention has been paid to the mechanism of oxidation meso-tetra(4-pyridyl)-porphyrin catalyst could also promote
of HMF into FDCA over supported metal catalysts. Davis et al. the oxidation of HMF into FDCA in acetonitrile by the use of
studied the kinetics and mechanism of the oxidation of HMF into t-BuOOH, affording a 90.4% FDCA yield after 24 h at 373 K.278
FDCA over Pt/C and Au/C catalysts in alkaline solutions.276,277 This catalyst could be reused without loss of its catalytic activity.
According to the isotope labeling technology, a plausible mecha- Although the oxidation of HMF into FDCA with t-BuOOH can be
nism was proposed for the Pt/C and Au/C catalysts (Fig. 27). The performed without a base, the large scale production of FDCA is
incorporation of two 18O atoms was observed in HFCA when the limited, because of the explosive peroxide. H2O2 is another
reaction is performed in H218O. The aldehyde side chain is environmentally-friendly and green oxidant. Tetra-ethylpyridinium
believed to undergo rapid reversible hydration to a geminal diol octamolybdate ([EPy]4Mo8O26) was active for the oxidation of
via nucleophilic addition of a hydroxide ion to the carbonyl and HMF into FDCA with H2O2 as the oxidant.279 A 99.5% FDCA
subsequent proton transfer from water to the alkoxy ion inter- yield was produced after 2 h at 373 K with H2O2/NaOH/HMF
mediate (Fig. 27, step 1). The second step is the dehydrogenation molar ratio = 18 : 2 : 1.
of the geminal diol intermediate, facilitated by the hydroxide ions
adsorbed on the metal surface, to produce the carboxylic acid 7.7 Electrocatalytic oxidation of HMF
(Fig. 27, step 2). An oxidation experiment performed in the The advantage of electrocatalytic oxidation is that it is due to
absence of a base over both Au and Pt catalysts resulted in no electron transfer and does not require chemical oxidants.280,281
conversion of HMF, indicating the important role of OH in the The first report of electrochemical oxidation of HMF into FDCA
oxidation of HMF into FDCA at 295 K. Hydroxide ions on the was in 1995 in an H-shaped cell.282 Carbon-black supported
catalyst surface then facilitate the activation of the C–H bond noble metal catalysts were studied for the electrocatalytic oxida-
in the alcohol side-chain to form the aldehyde intermediate, tion of HMF in alkaline solution.283 In electrocatalytic oxidation
5-formyl-2-furancarboxylic acid (FFCA). Then, FFCA was oxi- of HMF, a competitive oxidation is the oxidation of water to O2.
dized to FDCA followed by steps 4 and 5, which were the same Choi and co-workers employed a redox mediator (TEMPO) to
as steps 1 and step 2, respectively. Therefore, this mechanism inhibit the oxidation of water over a Pt electrode.284 The use of
in Fig. 29 can explain the incorporation of all 4 18O atoms in TEMPO significantly reduced the necessary overpotential to
FDCA when the reaction is performed in labeled water. The role initiate HMF oxidation. As shown in Fig. 30, TEMPO was firstly
of O2 is used to scavenge electrons from the catalyst surface, oxidized to TEMPO+ near the electrode, and TEMPO+ promoted
thereby closing the catalytic cycle. the HMF oxidation. High FDCA yield (Z99%) and Faradaic

1382 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

which made these processes less green and less cost-effective.


Therefore, much effort is still needed to develop new catalytic
methods for the base-free oxidation of HMF into FDCA with
O2 as the oxidant with high activity, selectivity and stability.
Carefully choosing a suitable support and the active metals
seems to be important to realize this aim.

7.8 Conversion of carbohydrates into FDCA


Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

There are a few examples on the oxidative transformation of


carbohydrates into FDCA (Fig. 31). In 2000, Kröger and co-workers
first reported the transformation of fructose into FDCA via a
two-step process using a membrane as shown in Fig. 32.289
Fig. 30 Scheme of electrocatalytic oxidation of HMF into FDCA with
TEMPO as the mediator. Adapted from ref. 284 with permission from Fructose dehydration to HMF was first conducted in water at
Nature Publishing Group, copyright 2015. 353 K with Lewatit SPC 108 (a microporous cationic ion
exchanger in H+ form) as the solid acid catalyst. The HMF
transferred into the MIBK phase through the membrane, followed
efficiency (Z93%) were obtained in water at pH = 9.2. Kinetic by the oxidation of HMF into FDCA over a PtBi/C catalyst at 353 K.
studies and cyclic voltammetry indicated that DFF was the reaction The reaction rate was controlled by the diffusion of HMF through
intermediate. Similar TEMPO mediated electrocatalytic oxidation the membrane. After 7 days, FDCA was produced in an overall
of HMF was performed in a CH2Cl2/H2O biphasic system, produ- yield of 25% with a selectivity of 50%. Levulinic acid was also
cing an HMF conversion of 92% and DFF yield of 64%.285 produced in a similar yield of 25%. Ribeiro and Schuchardt
Sun and co-workers made great progress in the electrocata- studied the one-pot transformation of fructose into FDCA over a
lytic oxidation of HMF over a noble-metal-free bifunctional bifunctional catalyst.290 Cobalt acetylacetonate encapsulated in
electrocatalyst.286 Ni2P nanoparticles arrayed on nickel foam sol–gel silica was used as the acid–metal bifunctional catalyst for
(Ni2P/NF) were used as a bifunctional electrocatalyst to couple the one-step conversion of fructose into FDCA. Fructose conver-
HMF oxidation and H2 evolution in alkaline solution. Linear sion of 72% with an FCDA selectivity of 99% was produced at
sweep voltammetry indicated that the oxidation of HMF was 433 K and 20 bar air in water. Fructose was dehydrated into HMF
significantly easier than the oxidation of water catalyzed by on the silica catalyst. The HMF was then oxidized on the metal
Ni2P/NF. The Ni2P/NF catalyst could be employed both as a sites. No other intermediates were observed. Zhang et al. reported
cathode and an anode for the HMF oxidation and H2 produc- a two-step method for the synthesis of FDCA from fructose that
tion, respectively, which produced a high current density
(50 mA cm 2) with a voltage at least 200 mV less than that of
pure water splitting electrolysis. In addition, near unity Faradaic
efficiencies were achieved for both H2 (100%) and FDCA (98%)
generation, together with robust stability. HMF oxidation into
FDCA catalyzed by Ni2P/NF likely followed the HMFCA route.
Meanwhile, the same authors also reported similar work by the
use of other kinds of electrocatalysts such as Ni3S2/NF287 and
CoP/copper foam288 for the coupling of HMF oxidation and
H2 evolution. The electrocatalytic methods with the superior
performance for H2 and FDCA production, render the current
catalytic systems very appealing for sustainable energy conver-
sion technologies.
In summary, the oxidation of HMF into FDCA was mainly
carried out over noble metal catalysts with O2 as the oxidant.
Fig. 32 Scheme of the processes in the membrane reactor: 1 – HMF
Au catalysts have shown encouraging performance for this formation in a water phase, 2 – diffusion of HMF in an MIBK phase and
reaction over Pt and Pd catalysts. In most cases, the oxidation 3 – HMF oxidation. Reproduced with permission from ref. 289 with
reactions were carried out with an excessive amount of base, permission from springer, copyright 2000.

Fig. 31 Synthesis of FDCA from glucose via a one-pot reaction.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1383
View Article Online

Review Article Chem Soc Rev

involved conversion of fructose into HMF in isopropanol with stable for glucose oxidation than Pt and Pd catalysts which
HCl followed by oxidation of HMF into FDCA over an Au/HT undergo oxidation. Au supported on acidic materials can con-
catalyst.291 Humin impurities decreased the catalyst activity vert cellobiose into gluconic acid, but is ineffective for cellulose
and stability. The humin impurities could be removed by transformation into gluconic acid. Compared with the oxida-
evaporating isopropanol from the original solution of HMF at tion of glucose into gluconic acid, the selective oxidation of
40 1C under reduced pressure, and then extracting the HMF glucose into glucaric acid is much more difficult, which is
from the solids with water. The overall FDCA yield from fructose currently performed over supported Pt catalysts.
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

was 83%. Zhang and co-workers then designed a triphasic Oxidation of HMF produces DFF and FDCA. Oxidation
approach for the conversion of fructose and glucose into of HMF into DFF was mainly carried out in organic solvents
FDCA.292 HMF was produced in phase I (TEAB or water) from without a base with transition metal (vanadium or manganese)
fructose with Amberlyst-15 or glucose by CrCl3. The HMF in or ruthenium catalysts. Vanadium catalysts showed low catalytic
phase I was extracted into MIBK (phase II) and transferred into performance in terms of activity, DFF selectivity and catalyst
water (phase III), where FDCA was produced with an Au–Pd/HT stability, while OMS-2 (KMn8O16nH2O) based catalysts had
catalyst with O2. Overall FDCA yields of 78% and 50% were higher activity and stability in the oxidation of HMF into DFF.
obtained from fructose and glucose, respectively. The same The oxidation of HMF into FDCA was mainly performed in water
group then used a commercial Ru/C catalyst and FDCA as the with an excess of base over supported noble metal catalysts (Pt,
acid catalyst for the two-step transformation of fructose into Pd, and Au). Au catalysts had higher selectivity and stability than
FDCA.293 Ru/C catalyzed the base-free oxidation of HMF to FDCA Pt and Pd catalysts in the aerobic oxidation of HMF into FDCA.
with a yield of 88% (after 10 h at 293 K and 2 bar O2). The FDCA The one-pot conversion of carbohydrates into DFF or FDCA is
could then be used as an acid catalyst for the dehydration of much slower than using HMF as the starting material.
fructose to HMF in an isopropanol/H2O system, where 64% HMF Although great progress has recently been achieved for the
yield was produced after 30 min at 443 K. Then the second step catalytic oxidation of carbohydrates or their derivate HMF into
of HMF oxidation to FDCA was demonstrated with an overall important organic acids or furanic compounds, further improve-
yield of 53% after 15 h (83% yield in the second step). The rate of ments are still necessary in many cases with the aim to realize the
oxidation of synthesized HMF (from sugar) was slower than the commercial application of these processes for the synthesis of
rate of oxidation of pure commercial HMF (15 h vs.10 h), which valuable chemicals through biorefineries. (1) Catalytic oxidation
may be due to some minor water soluble impurities in the in a biomass refinery should not be limited to the currently
freshly synthesized HMF solution. To facilitate the separation reported routes, but also should pay special attention to devel-
of the acid–metal catalysts, two magnetic catalysts were used for oping new routes, or even exploring novel strategies to produce
the oxidative conversion of fructose into FDCA.294 Fe3O4@SiO2– new products; (2) efforts should be devoted to the design of active
SO3H was used as the acid catalyst for the dehydration of and effective catalysts for the transformation of carbohydrates
fructose HMF to produce HMF with a yield of 93.1% at 373 K into the target products under mild reaction conditions with high
after 2 h. After the removal of the Fe3O4@SiO2–SO3H catalyst, conversion and selectivity especially for the production of formic
HMF was then oxidized into FDCA with t-BuOOH over a nano- acid and acetic acid. (3) In order to lower the production cost,
Fe3O4–CoOx catalyst at 353 K. FDCA was obtained in an overall much attention should be directed to the development of none
yield of 59.8% after 15 h based on the starting fructose. noble based catalysts for the oxidative transformation of biomass
with high activity, stability and facile catalyst recycling. (4) It is
expected that the design novel multi-functional catalysts will
8. Conclusions and perspectives allow more substrates as the starting materials to produce
chemicals via several steps in a one-pot reaction, thus avoiding
As listed in Fig. 1, a wide variety of commodity chemicals the costly intermediate separation process. (5) It is necessary to
including formic acid, acetic acid, glycolic acid, gluconic acid, get more insights into the mechanism of the transformation
glucaric acid, malonic acid, oxalic acid, and furan chemicals reactions, and the structure–property relationship of the cata-
(DFF and FDCA) can be produced from the catalytic oxidation lysts, which is of great significance from both scientific and
of carbohydrates and carbohydrate derivative HMF. Polyoxo- practical viewpoints. (6) It is extremely important to perform
metalates or polyoxometalate based catalysts are mainly used the large-scale production of chemicals via catalytic oxidation
for the conversion of cellulose or glucose into formic acid with reactions in a biorefinery. Many current methods are technically
molecular oxygen as the oxidant. Carbon dioxide is the typical feasible, but not economical. Although these processes still need
undesired byproduct from this reaction and is produced to be improved for practical use, we believe that further progress
by oxidation of an intermediate. Hydrothermal oxidation by will realize a biorefinery by heterogeneous catalysis to ensure
the use of H2O2 as the oxidant can convert the carbohydrates sustainable development.
into acetic acid or formic acid in low yields as the harsh
hydrothermal reaction conditions (high temperatures and high
pressure) cause other organic acids to be formed. Gluconic acid Conflicts of interest
can be produced by the oxidation of glucose over supported
metal catalysts under high pH conditions. Au catalysts are more There are no conflicts to declare.

1384 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

Acknowledgements 27 J. Albert and P. Wasserscheid, Green Chem., 2015, 17,


5164–5171.
This work was supported by the funding offered by the China 28 J. Albert, D. Luders, A. Bosmann, D. M. Guldi and
scholarship council (201408420018). P. Wasserscheid, Green Chem., 2014, 16, 226–237.
29 J. Z. Zhang, M. Sun, X. Liu and Y. Han, Catal. Today, 2014,
233, 77–82.
30 Y. Leng, J. Wang, D. R. Zhu, X. Q. Ren, H. Q. Ge and
References
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

L. Shen, Angew. Chem., Int. Ed., 2009, 48, 168–171.


1 G. W. Huber, S. Iborra and A. Corma, Chem. Rev., 2006, 31 J. L. Xu, H. Y. Zhang, Y. F. Zhao, Z. Z. Yang, B. Yu, H. J. Xu
106, 4044–4098. and Z. M. Liu, Green Chem., 2014, 16, 4931–4935.
2 W. P. Deng, Q. H. Zhang and Y. Wang, Catal. Today, 2014, 32 K. X. Li, L. L. Bai, P. N. Amaniampong, X. L. Jia, J. M. Lee
234, 31–41. and Y. H. Yang, ChemSusChem, 2014, 7, 2670–2677.
3 T. F. Wang, M. W. Nolte and B. H. Shanks, Green Chem., 33 W. Wang, M. Niu, Y. Hou, W. Wu, Z. Liu, Q. Liu, S. Ren and
2014, 16, 548–572. K. N. Marsh, Green Chem., 2014, 16, 2614–2618.
4 X. L. Tong, Y. Ma and Y. D. Li, Appl. Catal., A, 2010, 385, 34 M. G. Niu, Y. C. Hou, S. H. Ren, W. Z. Wu and K. N. Marsh,
1–13. Green Chem., 2015, 17, 453–459.
5 Y. Kwon, K. J. P. Schouten, J. C. van der Waal, E. de Jong 35 B. M. Weckhuysen and D. E. Keller, Catal. Today, 2003, 78,
and M. T. M. Koper, ACS Catal., 2016, 6, 6704–6717. 25–46.
6 M. E. Zakrzewska, E. Bogel-Lukasik and R. Bogel-Lukasik, 36 M. Tani, T. Sakamoto, S. Mita, S. Sakaguchi and Y. Ishii,
Chem. Rev., 2011, 111, 397–417. Angew. Chem., Int. Ed., 2005, 44, 2586–2588.
7 G. Z. Papageorgiou, D. G. Papageorgiou, Z. Terzopoulou 37 M. G. Niu, Y. C. Hou, S. H. Ren, W. H. Wang, Q. T. Zheng
and D. N. Bikiaris, Eur. Polym. J., 2016, 83, 202–229. and W. Z. Wu, Green Chem., 2015, 17, 335–342.
8 http://www.alibaba.com/?spm=a2700.7699653.a271qf.40. 38 Z. C. Tang, W. P. Deng, Y. L. Wang, E. Z. Zhu, X. Y. Wan,
APWJep. Q. H. Zhang and Y. Wang, ChemSusChem, 2014, 7, 1557–1567.
9 https://en.wikipedia.org/wiki/Formic_acid. 39 J. Reichert, B. Bruner, A. Jess, P. Wasserscheid and
10 https://en.wikipedia.org/wiki/Acetic_acid. J. Albert, Energy Environ. Sci., 2015, 8, 2985–2990.
11 https://en.wikipedia.org/wiki/Glycolic_acid. 40 A. M. Khenkin and R. Neumann, J. Am. Chem. Soc., 2008,
12 http://www.transparencymarketresearch.com/glycolic-acid- 130, 14474–14476.
market.html. 41 A. M. Khenkin, G. Leitus and R. Neumann, J. Am. Chem.
13 https://en.wikipedia.org/wiki/Gluconic_acid. Soc., 2010, 132, 11446–11448.
14 http://www.grandviewresearch.com/industry-analysis/fdca- 42 L. Calvo and D. Vallejo, Ind. Eng. Chem. Res., 2002, 41,
industry. 6503–6509.
15 L. F. Fieser and J. E. Jones, Org. Synth., 1955, 3, 590. 43 F. M. Jin, J. Yun, G. M. Li, A. Kishita, K. Tohji and
16 R. V. Jagadeesh, D. Banerjee, P. B. Arockiam, H. Junge, H. Enomoto, Green Chem., 2008, 10, 612–615.
K. Junge, M. M. Pohl, J. Radnik, A. Bruckner and M. Beller, 44 J. Yun, G. D. Yao, F. M. Jin, H. Zhong, A. Kishita, K. Tohji,
Green Chem., 2015, 17, 898–902. H. Enomoto and L. Wang, AIChE J., 2016, 62, 3657–3663.
17 S. Zhang, O. Metin, D. Su and S. H. Sun, Angew. Chem., Int. 45 N. Zargari, Y. Kim and K. W. Jung, Green Chem., 2015, 17,
Ed., 2013, 52, 3681–3684. 2736–2740.
18 Y. Himeda, Green Chem., 2009, 11, 2018–2022. 46 H. Choudhary, S. Nishimura and K. Ebitani, Appl. Catal., B,
19 A. Boddien, D. Mellmann, F. Gartner, R. Jackstell, H. Junge, 2015, 162, 1–10.
P. J. Dyson, G. Laurenczy, R. Ludwig and M. Beller, Science, 47 B. C. Gates, H. Knozinger and F. C. Jentoft, Adv. Catal.,
2011, 333, 1733–1736. 2010, 53, 1–45.
20 L. G. Feng, J. F. Chang, K. Jiang, H. G. Xue, C. P. Liu, 48 N. Saichana, K. Matsushita, O. Adachi, I. Frebort and
W. B. Cai, W. Xing and J. J. Zhang, Nano Energy, 2016, 30, J. Frebortova, Biotechnol. Adv., 2015, 33, 1260–1271.
355–361. 49 F. M. Jin, Z. Y. Zhou, T. Moriya, H. Kishida, H. Higashijima
21 W. Reutemann and H. Kieczka, Formic Acid, Ullmann’s and H. Enomoto, Environ. Sci. Technol., 2005, 39, 1893–1902.
Encyclopedia of Industrial Chemistry, 2011. 50 F. M. Jin, Z. Y. Zhou, A. Kishita, H. Enomoto, H. Kishida
22 C. Schopf and H. Wild, Chem. Ber., 1954, 87, 1571–1575. and T. Moriya, Chem. Eng. Res. Des., 2007, 85, 201–206.
23 R. Wolfel, N. Taccardi, A. Bosmann and P. Wasserscheid, 51 Y. Fang, X. Zeng, P. Yan, Z. Z. Jing and F. M. Jin, Ind.
Green Chem., 2011, 13, 2759–2763. Eng. Chem. Res., 2012, 51, 4759–4763.
24 J. Li, D. J. Ding, L. Deng, Q. X. Guo and Y. Fu, ChemSusChem, 52 J. Li, Y. B. Huang, Q. X. Guo and Y. Fu, Acta Chim. Sin.,
2012, 5, 1313–1318. 2014, 72, 1223–1227.
25 T. Lu, M. G. Niu, Y. C. Hou, W. Z. Wu, S. H. Ren and 53 L. Y. Li, F. Shen, R. L. Smith and X. H. Qi, Green Chem.,
F. Yang, Green Chem., 2016, 18, 4725–4732. 2017, 19, 76–81.
26 J. Albert, R. Wolfel, A. Bosmann and P. Wasserscheid, 54 X. Y. Gao, H. Zhong, G. D. Yao, W. M. Guo and F. M. Jin,
Energy Environ. Sci., 2012, 5, 7956–7962. Catal. Today, 2016, 274, 49–54.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1385
View Article Online

Review Article Chem Soc Rev

55 X. Y. Gao, X. Chen, J. G. Zhang, W. M. Guo, F. M. Jin and 78 A. Tathod, T. Kane, E. S. Sanil and P. L. Dhepe, J. Mol.
N. Yan, ACS Sustainable Chem. Eng., 2016, 4, 3912–3920. Catal. A: Chem., 2014, 388, 90–99.
56 http://www.transparencymarketresearch.com/glycolic- 79 A. Haruta, Chem. Rec., 2003, 3, 75–87.
acid-market.html. 80 M. Stratakis and H. Garcia, Chem. Rev., 2012, 112, 4469–4506.
57 F. W. Wang, Y. Q. Wang, F. M. Jin, G. D. Yao, Z. B. Huo, 81 M. Pan, A. J. Brush, Z. D. Pozun, H. C. Ham, W. Y. Yu,
X. Zeng and Z. Z. Jing, Ind. Eng. Chem. Res., 2014, 53, G. Henkelman, G. S. Hwang and C. B. Mullins, Chem. Soc.
7939–7946. Rev., 2013, 42, 5002–5013.
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

58 J. Z. Zhang, X. Liu, M. Sun, X. H. Ma and Y. Han, ACS 82 M. D. Hughes, Y. J. Xu, P. Jenkins, P. McMorn, P. Landon,
Catal., 2012, 2, 1698–1702. D. I. Enache, A. F. Carley, G. A. Attard, G. J. Hutchings,
59 Z. She, J. G. Wang, J. P. Ni, X. Q. Liu, R. Y. Zhang, H. N. Na F. King, E. H. Stitt, P. Johnston, K. Griffin and C. J. Kiely,
and J. Zhu, RSC Adv., 2015, 5, 5741–5744. Nature, 2005, 437, 1132–1135.
60 A. Corma, S. Iborra and A. Velty, Chem. Rev., 2007, 107, 83 B. T. Kusema and D. Y. Murzin, Catal. Sci. Technol., 2013,
2411–2502. 3, 297–307.
61 C. S. K. Lin, L. A. Pfaltzgraff, L. Herrero-Davila, 84 C. D. Pina, E. Falletta and M. Rossi, Chem. Soc. Rev., 2012,
E. B. Mubofu, S. Abderrahim, J. H. Clark, A. A. Koutinas, 41, 350–369.
N. Kopsahelis, K. Stamatelatou, F. Dickson, S. Thankappan, 85 S. Biella, L. Prati and M. Rossi, J. Catal., 2002, 206, 242–247.
Z. Mohamed, R. Brocklesby and R. Luque, Energy Environ. 86 C. Y. Ma, W. J. Xue, J. J. Li, W. Xing and Z. P. Hao, Green
Sci., 2013, 6, 426–464. Chem., 2013, 15, 1035–1041.
62 L. Y. Yan and X. Y. Xi, ACS Sustainable Chem. Eng., 2014, 2, 87 M. Comotti, C. D. Pina, E. Falletta and M. Rossi, J. Catal.,
897–901. 2006, 244, 122–125.
63 W. Riemenschneider and M. Tanifuji, Oxalic acid, Ullmann’s 88 P. Y. Qi, S. S. Chen, J. Chen, J. W. Zheng, X. L. Zheng and
Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag Y. Z. Yuan, ACS Catal., 2015, 5, 2659–2670.
GmbH & Co. KGaA, 2000. 89 A. Mirescu, H. Berndt, A. Martin and U. Prusse, Appl.
64 Z. W. Jiang, Z. R. Zhang, J. L. Song, Q. L. Meng, H. C. Zhou, Catal., A, 2007, 317, 204–209.
Z. H. He and B. X. Han, ACS Sustainable Chem. Eng., 2016, 90 Y. L. Cao, X. Liu, S. Iqbal, P. J. Miedziak, J. K. Edwards,
4, 305–311. R. D. Armstrong, D. J. Morgan, J. W. Wang and
65 A. Costine, J. S. C. Loh, F. Busetti, C. A. Joll and A. Heitz, G. J. Hutchings, Catal. Sci. Technol., 2016, 6, 107–117.
Ind. Eng. Chem. Res., 2013, 52, 5572–5581. 91 Y. R. Wang, S. Van de Vyver, K. K. Sharma and Y. Roman-
66 P. N. Amaniampong, Q. T. Trinh, B. Wang, A. Borgna, Leshkov, Green Chem., 2014, 16, 719–726.
Y. H. Yang and S. H. Mushrif, Angew. Chem., Int. Ed., 2015, 92 U. Prusse, M. Herrmann, C. Baatz and N. Decker, Appl.
54, 8928–8933. Catal., A, 2011, 406, 89–93.
67 R. H. Blom, V. F. Pfeifer, A. J. Moyer, D. H. Traufler and 93 A. Mirescu and U. Prusse, Appl. Catal., B, 2007, 70, 644–652.
H. F. Conway, Ind. Eng. Chem., 1952, 44, 435–440. 94 C. Baatz and U. Prusse, J. Catal., 2007, 249, 34–40.
68 J. Bao, K. Furumoto, K. Fukunaga and K. Nakao, Biochem. 95 C. Baatz, N. Thielecke and U. Prusse, Appl. Catal., B, 2007,
Eng. J., 2001, 8, 91–102. 70, 653–660.
69 H. J. Zhang and N. Toshima, Catal. Sci. Technol., 2013, 3, 96 I. V. Delidovich, B. L. Moroz, O. P. Taran, N. V. Gromov,
268–278. P. A. Pyrjaev, I. P. Prosvirin, V. I. Bukhtiyarov and V. N. Parmon,
70 M. J. Climent, A. Corma and S. Iborra, Green Chem., 2011, Chem. Eng. J., 2013, 223, 921–931.
13, 520–540. 97 R. Saliger, N. Decker and U. Prusse, Appl. Catal., B, 2011,
71 M. Besson, F. Lahmer, P. Gallezot, P. Fuertes and G. Fleche, 102, 584–589.
J. Catal., 1995, 152, 116–121. 98 N. Thielecke, K. D. Vorlop and U. Prusse, Catal. Today,
72 S. Karski, T. Paryjczak and I. Witonska, Kinet. Catal., 2003, 2007, 122, 266–269.
44, 618–622. 99 N. Thielecke, M. Ayternir and U. Prusse, Catal. Today, 2007,
73 D. Santhanaraj, M. R. Rover, D. E. Resasco, R. C. Brown 121, 115–120.
and S. Crossley, ChemSusChem, 2014, 7, 3132–3137. 100 T. Benko, A. Beck, O. Geszti, R. Katona, A. Tungler, K. Frey,
74 X. Liang, C. J. Liu and P. Kuai, Green Chem., 2008, 12, L. Guczi and Z. Schay, Appl. Catal., A, 2010, 388, 31–36.
1318–1322. 101 T. Ishida, N. Kinoshita, H. Okatsu, T. Akita, T. Takei and
75 V. Y. Doluda, I. B. Tsvetkova, A. V. Bykov, V. G. Matveeva, M. Haruta, Angew. Chem., Int. Ed., 2008, 47, 9265–9268.
A. I. Sidorov, M. G. Sulman, P. M. Valetsky, B. D. Stein, 102 K. Odrozek, K. Maresz, A. Koreniuk, K. Prusik and
E. M. Sulman and L. M. Bronstein, Polymer-Protected and J. Mrowiec-Bialon, Appl. Catal., A, 2014, 475, 203–210.
Au-Containing Bi- and Trimetallic Nanoparticles as Novel 103 P. Bujak, P. Bartczak and J. Polanski, J. Catal., 2012, 295,
Catalysts for Glucose Oxidation, Green Process Synth., 2013, 15–21.
2, 25–34. 104 T. Ishida, S. Okamoto, R. Makiyama and M. Haruta, Appl.
76 J. M. H. Dirkx and H. S. Vanderbaan, J. Catal., 1981, 67, 1–13. Catal., A, 2009, 353, 243–248.
77 A. E. Koklin, T. A. Klimenko, A. V. Kondratyuk, V. V. Lunin 105 T. Ishida, H. Watanabe, T. Bebeko, T. Akita and M. Haruta,
and V. I. Bogdan, Kinet. Catal., 2015, 56, 84–88. Appl. Catal., A, 2010, 377, 42–46.

1386 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

106 H. J. Zhang and N. Toshima, Appl. Catal., A, 2012, 447, 81–88. 132 A. Onda, T. Ochi and K. Yanagisawa, Catal. Commun.,
107 H. J. Zhang, N. Toshima, K. Takasaki and M. Okumura, 2011, 12, 421–425.
J. Alloys Compd., 2014, 586, 462–468. 133 K. Hasegawa, K. Shimidzu, S. Morimoto, Y. Hachino,
108 H. J. Zhang, M. Haba, M. Okumura, T. Akita, S. Hashimoto K. Hasegawa, Y. Shono, T. Horiuchi, K. Tabuse, Y. Masui,
and N. Toshima, Langmuir, 2013, 29, 10330–10339. T. Yamaguchi and Y. Fujita, Yakugaku Zasshi, 2008, 128,
109 H. M. Yin, C. Q. Zhou, C. X. Xu, P. P. Liu, X. H. Xu and 135–140.
Y. Ding, J. Phys. Chem. C, 2008, 112, 9673–9678. 134 S. V. de Vyver and Y. Román-Leshkov, Catal. Sci. Technol.,
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

110 H. J. Zhang and N. Toshima, J. Colloid Interface Sci., 2013, 2013, 3, 1465–1479.
394, 166–176. 135 I. A. Colon, R. Fernandezgarcia, L. Amoros and H. Blay,
111 M. Comotti, C. D. Pina and M. Rossi, J. Mol. Catal. A: J. Am. Chem. Soc., 1949, 71, 4131–4133.
Chem., 2006, 251, 89–92. 136 D. S. Bin, H. Wang, J. X. Li, H. Wang, Z. Yin, J. L. Kang,
112 E. Smolentseva, B. T. Kusema, S. Beloshapkin, M. Estrada, B. Q. He and Z. H. Li, Electrochim. Acta, 2014, 130, 170–178.
E. Vargas, D. Y. Murzin, F. Castillon, S. Fuentes and 137 M. Ibert, P. Fuertes, N. Merbouh, C. Feasson and
A. Simakov, Appl. Catal., A, 2011, 392, 69–79. F. Marsais, Carbohydr. Res., 2011, 346, 512–518.
113 H. J. Zhang, K. Kawashima, M. Okumura and N. Toshima, 138 M. Ibert, P. Fuertes, N. Merbouh, C. Fiol-Petit, C. Feasson
J. Mater. Chem. A, 2014, 2, 13498–13508. and F. Marsais, Electrochim. Acta, 2010, 55, 3589–3594.
114 P. J. Miedziak, H. Alshammari, S. A. Kondrat, T. J. Clarke, 139 X. Jin, M. Zhao, J. Shen, W. J. Yan, L. M. He, P. S. Thapa,
T. E. Davies, M. Morad, D. J. Morgan, D. J. Willock, S. Q. Ren, B. Subramaniam and R. V. Chaudhari, J. Catal.,
D. W. Knight, S. H. Taylor and G. J. Hutchings, Green 2015, 330, 323–329.
Chem., 2014, 16, 3132–3141. 140 X. Jin, M. Zhao, M. Vora, J. Shen, C. Zeng, W. J. Yan,
115 H. J. Zhang, L. L. Lu, Y. N. Cao, S. Du, Z. Cheng and P. S. Thapa, B. Subramaniam and R. V. Chaudhari, Ind.
S. W. Zhang, Mater. Res. Bull., 2014, 49, 393–398. Eng. Chem. Res., 2016, 55, 2932–2945.
116 H. J. Zhang and N. Toshima, Appl. Catal., A, 2011, 400, 9–13. 141 I. Matzerath, W. Klaui, R. Klasen and H. Sahm, Inorg.
117 H. J. Zhang, M. Okumura and N. Toshima, J. Phys. Chem. C, Chim. Acta, 1995, 237, 203–205.
2011, 115, 14883–14891. 142 V. J. Murphy, J. Shoemaker, G. Zhu, R. Archer, G. F. Salem
118 N. Toshima and H. J. Zhang, Macromol. Symp., 2012, 317, and E. L. Dias, US2011/0306790, 2011.
149–159. 143 J. Boucher, C. Chirat and D. Lachenal, Cellul. Chem.
119 R. Saliger, N. Decker and U. Prusse, Appl. Catal., B, 2011, Technol., 2015, 49, 303–308.
102, 584–589. 144 J. M. H. Dirkn, H. S. van der Baan and J. M. A. J. J. van den
120 S. Rautiainen, P. Lehtinen, M. Vehkamaki, K. Niemela, Broen, Carbohydr. Res., 1977, 59, 63–72.
M. Kemell, M. Heikkila and T. Repo, Catal. Commun., 2016, 145 M. Besson, G. Fleche, P. Fuertes, P. Gallezot and
74, 115–118. F. Lahmer, Recl. Trav. Chim. Pays-Bas, 1996, 115, 217–221.
121 M. Omri, G. Pourceau, M. Becuwe and A. Wadouachi, ACS 146 J. Lee, B. Saha and D. G. Vlachos, Green Chem., 2016, 18,
Sustainable Chem. Eng., 2016, 4, 2432–2438. 3815–3822.
122 D. Rinsant, G. Chatel and F. Jerome, ChemCatChem, 2014, 147 E. Derrien, P. Marion, C. Pinel and M. Besson, Org. Process
6, 3355–3359. Res. Dev., 2016, 20, 1265–1275.
123 X. S. Tan, W. P. Deng, M. Liu, Q. H. Zhang and Y. Wang, 148 P. Maki-Arvela, T. Salmi, B. Holmbom, S. Willfor and
Chem. Commun., 2009, 7179–7181. D. Y. Murzin, Chem. Rev., 2011, 111, 5638–5666.
124 K. M. Eblagon, M. F. R. Pereira and J. L. Figueiredo, Appl. 149 T. F. Wang, M. W. Nolte and B. H. Shanks, Green Chem.,
Catal., B, 2016, 184, 381–396. 2014, 16, 548–572.
125 O. Poole, K. Alharbi, D. Belic, E. F. Kozhevnikova and 150 S. P. Teong, G. S. Yi and Y. G. Zhang, Green Chem., 2014,
I. V. Kozhevnikov, Appl. Catal., B, 2017, 202, 446–453. 16, 2015–2026.
126 J. Zhang, X. Liu, M. N. Hedhili, Y. Zhu and Y. Han, 151 H. Hirai, J. Macromol. Sci., Chem., 1984, 21, 1165–1179.
ChemCatChem, 2011, 3, 1294–1298. 152 A. C. Braisted, J. D. Oslob, W. L. Delano, J. Hyde,
127 D. L. An, A. H. Ye, W. P. Deng, Q. H. Zhang and Y. Wang, R. S. McDowell, N. Waal, C. Yu, M. R. Arkin and
Chem. – Eur. J., 2012, 18, 2938–2947. B. C. Raimundo, J. Am. Chem. Soc., 2003, 125, 3714–3715.
128 H. Sakurai, K. Koga and M. Kiuchi, Catal. Today, 2015, 251, 153 T. Elhajj, A. Masroua, J. C. Martin and G. Descotes, Bull.
96–102. Soc. Chim. Fr., 1987, 855–860.
129 P. N. Amaniampong, K. X. Li, X. L. Jia, B. Wang, A. Borgna 154 B. W. Lew, US Pat. 3326944, 1967.
and Y. H. Yang, ChemCatChem, 2014, 6, 2105–2114. 155 Y. Y. Gorbanev, S. K. Klitgaard, J. M. Woodley, C. H.
130 P. N. Amaniampong, X. L. Jia, B. Wang, S. H. Mushrif, Christensen and A. Riisager, ChemSusChem, 2009, 2,
A. Borgna and Y. H. Yang, Catal. Sci. Technol., 2015, 5, 672–675.
2393–2405. 156 S. E. Davis, L. R. Houk, E. C. Tamargo, A. K. Datye and
131 P. N. Amaniampong, A. Y. Booshehri, X. L. Jia, Y. Dai, R. J. Davis, Catal. Today, 2011, 160, 55–60.
B. Wang, S. H. Mushrif, A. Borgna and Y. H. Yang, Appl. 157 Z. H. Zhang, B. Liu, K. L. Lv, J. Sun and K. J. Deng, Green
Catal., A, 2015, 505, 16–27. Chem., 2014, 16, 2762–2770.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1387
View Article Online

Review Article Chem Soc Rev

158 M. Ventura, M. Aresta and A. Dibenedetto, ChemSusChem, 184 C. A. Antonyraj, J. Jeong, B. Kim, S. Shin, S. Kim, K. Y. Lee
2016, 9, 1096–1100. and J. K. Cho, J. Ind. Eng. Chem., 2013, 19, 1056–1059.
159 P. Vinke, D. de Wit, T. J. W. De Goede and H. Van Bekkum, 185 J. F. Nie, J. H. Xie and H. C. Liu, J. Catal., 2013, 301, 83–91.
New developments in selective oxidation by heterogeneous 186 J. Artz, S. Mallmann and R. Palkovits, ChemSusChem, 2014,
catalysis, Elsevier, New York, 1989. 8, 672–679.
160 A. Gandini and N. M. Belgacem, Polym. Int., 1998, 47, 187 J. Artz, S. Mallmann and R. Palkovits, ChemSusChem, 2015,
267–276. 8, 3832–3838.
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

161 M. Baumgarten and N. Tyutyulkov, Chem. – Eur. J., 1998, 4, 188 K. Ghosh, R. A. Molla, M. A. Iqubal, S. S. Islam and
987–989. S. M. Islam, Appl. Catal., A, 2016, 520, 44–52.
162 F. W. Lichtenthaler and S. Mondel, Pure Appl. Chem., 1997, 189 Y. Wang, B. Liu, K. Huang and Z. Zhang, Ind. Eng. Chem.
69, 1853–1866. Res., 2014, 53, 1313–1319.
163 L. Cottier, G. Descotes, E. Viollet, J. Lewkowski and 190 Z. Z. Yang, W. Qi, R. X. Su and Z. M. He, Energy Fuels, 2017,
R. Skowronski, J. Heterocycl. Chem., 1995, 32, 927–930. 31, 533–541.
164 A. S. Amarasekata, D. Green and E. McMillan, Catal. 191 J. Z. Chen, J. W. Zhong, Y. Y. Guo and L. M. Chen, RSC Adv.,
Commun., 2008, 9, 286–288. 2015, 5, 5933–5940.
165 W. Partenheimer and V. V. Grushin, Adv. Synth. Catal., 192 S. G. Wang, Z. H. Zhang, B. Liu and J. L. Li, Ind. Eng. Chem.
2001, 343, 102–111. Res., 2014, 53, 5820–5827.
166 J. P. Ma, Z. T. Du, J. Xu, Q. H. Chu and Y. Pang, Chem- 193 Z. Z. Yang, J. Deng, T. Pan, Q. X. Guo and Y. Fu, Green
SusChem, 2011, 4, 51–54. Chem., 2012, 14, 2986–2989.
167 X. Q. Jia, J. P. Ma, M. Wang, Z. T. Du, F. Lu, F. Wang and 194 J. F. Nie and H. C. Liu, J. Catal., 2014, 316, 57–66.
J. Xu, Appl. Catal., A, 2014, 482, 231–236. 195 G. D. Yadav and R. V. Sharma, Appl. Catal., B, 2014, 147,
168 X. L. Tong, Y. F. Sun, X. Q. Bai and Y. D. Li, RSC Adv., 2014, 293–301.
4, 44307–44311. 196 B. Liu, Z. H. Zhang, K. L. Lv, K. J. Deng and H. M. Duan,
169 M. O. Kornpanets, O. V. Kushch, Y. E. Litvinov, O. L. Pliekhov, Appl. Catal., A, 2014, 472, 64–71.
K. V. Novikova, A. O. Novokhatko, A. N. Shendrik, A. V. 197 F. Neatu, N. Petrea, R. Petre, V. Somoghi, M. Florea and
Vasilyev and I. O. Opeida, Catal. Commun., 2014, 57, 60–63. V. I. Parvulescu, Catal. Today, 2016, 278, 66–73.
170 O. C. Navarro, A. C. Canos and S. I. Chornet, Top. Catal., 198 R. Q. Fang, R. Luque and Y. W. Li, Green Chem., 2016, 18,
2009, 52, 304–314. 3152–3157.
171 X. X. Liu, J. F. Xiao, H. Ding, W. Z. Zhong, Q. Xu, S. P. Su 199 R. A. Sheldon, Catal. Today, 2015, 247, 4–13.
and D. L. Yin, Chem. Eng. J., 2016, 283, 1315–1321. 200 N. Mittal, G. M. Nisola, L. B. Malihan, J. G. Seo, S. P. Lee and
172 L. F. Liao, Y. Liu, Z. Y. Li, J. P. Zhuang, Y. B. Zhou and W. J. Chung, Korean J. Chem. Eng., 2014, 31, 1362–1367.
S. Z. Chen, RSC Adv., 2016, 6, 94976–94988. 201 N. Mittal, G. M. Nisola, J. G. Seo, S. P. Lee and W. J. Chung,
173 G. Durand, P. Faugeras, F. Laporte, C. Moreau, M. C. Neau, J. Mol. Catal. A: Chem., 2015, 404, 106–114.
G. Roux, D. Tichit and C. Toutremepuich, Agrichimie, 202 T. S. Hansen, I. Sadaba, E. J. Garcia-Suarez and A. Riisager,
WO9617836, 1995. Appl. Catal., A, 2013, 456, 44–50.
174 R. Durand, P. Faugeras, F. Laporte, C. Moreau, M. C. Neau, 203 B. Karimi, H. M. Mirzaei and E. Farhangi, ChemCatChem,
C. Doutremepuich, G. Roux and D. Tichit, EP0796254, 2014, 6, 758–762.
1997. 204 G. Q. Lv, H. L. Wang, Y. X. Yang, T. S. Deng, C. M. Chen,
175 C. Moreau, R. Durand, C. Pourcheron and D. Tichit, Stud. Y. L. Zhu and X. L. Hou, ACS Catal., 2015, 5, 5636–5646.
Surf. Sci. Catal., 1997, 108, 399–406. 205 X. Wang, X. Li, L. Zhang, Y. Yoon, P. K. Weber, H. Wang,
176 I. Sadaba, Y. Y. Gorbanev, S. Kegnaes, S. S. R. Putluru, R. W. J. Guo and H. Dai, Science, 2009, 324, 768–771.
Berg and A. Riisager, ChemCatChem, 2013, 5, 284–293. 206 C. Su and K. P. Lou, Acc. Chem. Res., 2013, 46, 2275–2285.
177 W. Zhang, J. Y. Xie, W. Hou, Y. Q. Liu, Y. Zhou and J. Wang, 207 G. Q. Lv, H. L. Wang, Y. X. Yang, X. Li, T. S. Deng,
ACS Appl. Mater. Interfaces, 2016, 8, 23122–23132. C. M. Chen, Y. L. Zhu and X. L. Hou, Catal. Sci. Technol.,
178 C. A. Antonyraj, B. Kim, Y. Kim, S. Shin, K. Y. Lee, I. Kim 2016, 6, 2377–2386.
and J. K. Cho, Catal. Commun., 2014, 57, 64–68. 208 C. Liu, J. M. Carraher, J. L. Swedberg, C. R. Herndon,
179 J. F. Nie and H. C. Liu, Pure Appl. Chem., 2012, 84, 765–777. C. N. Fleitman and J. P. Tessonnier, ACS Catal., 2014, 4,
180 Y. Y. Guo and J. Z. Chen, ChemPlusChem, 2015, 80, 4295–4298.
1760–1768. 209 V. V. Grushin, N. Herron and G. A. Halliday, WO2003024947,
181 C. Carlini, P. Patrono, A. M. R. Galletti, G. Sbrana and 2003.
V. Zima, Appl. Catal., A, 2005, 289, 197–204. 210 G. A. Halliday, R. J. Young and V. V. Grushin, Org. Lett.,
182 F. L. Grasset, B. Katryniok, S. Paul, V. Nardello-Rataj, 2003, 5, 2003–2005.
M. Pera-Titus, J. M. Clacens, F. De Campo and F. Dumeignil, 211 F. H. Xu and Z. H. Zhang, ChemCatChem, 2015, 7,
RSC Adv., 2013, 3, 9942–9948. 1470–1477.
183 A. Takagaki, M. Takahashi, S. Nishimura and K. Ebitani, 212 R. L. Liu, J. Z. Chen, L. M. Chen, Y. Y. Guo and J. W. Zhong,
ACS Catal., 2011, 1, 1562–1565. ChemCatChem, 2014, 6, 3174–3181.

1388 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018
View Article Online

Chem Soc Rev Review Article

213 R. L. Liu, J. Z. Chen, L. M. Chen, Y. Y. Guo and J. W. Zhong, 240 H. A. Rass, N. Essayem and M. Besson, ChemSusChem,
ChemPlusChem, 2014, 79, 1448–1454. 2015, 8, 1206–1217.
214 Y. Liu, L. F. Zhu, J. Q. Tang, M. Y. Liu, R. Q. Cheng and 241 P. Vinke, H. H. van Dam and H. van Bekkum, Stud. Surf.
C. W. Hu, ChemSusChem, 2014, 7, 3541–3547. Sci. Catal., 1990, 55, 147–157.
215 N. Mittal, G. M. Nisola, L. B. Malihan, J. G. Seo, H. Kim, 242 P. Vinke, W. van der Poel and H. van Bekkum, Stud. Surf.
S. P. Lee and W. J. Chung, RSC Adv., 2016, 6, 25678–25688. Sci. Catal., 1991, 59, 385–394.
216 S. Dabral, S. Nishimura and K. Ebitani, ChemSusChem, 243 R. Sahu and P. L. Dhepe, React. Kinet., Mech. Catal., 2014,
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

2014, 7, 260–267. 112, 173–187.


217 X. Xiang, L. He, Y. Yang, B. Guo, D. Tong and C. Hu, Catal. 244 S. Siankevich, G. Savoglidis, Z. F. Fei, G. Laurenczy, D. T. L.
Lett., 2011, 141, 735–741. Alexander, N. Yan and P. J. Dyson, J. Catal., 2014, 315, 67–74.
218 S. Q. Zhang, W. F. Li, X. H. Zeng, Y. Sun and L. Lin, 245 X. W. Han, L. Geng, Y. Guo, R. Jia, X. H. Liu, Y. G. Zhang
BioResources, 2014, 9, 4568–4580. and Y. Q. Wang, Green Chem., 2016, 18, 1597–1604.
219 Z. H. Zhang, Z. L. Yuan, D. G. Tang, Y. S. Ren, K. L. Lv and 246 B. Siyo, M. Schneider, M. M. Pohl, P. Langer and
B. Liu, ChemSusChem, 2014, 7, 3496–3504. N. Steinfeldt, Catal. Lett., 2014, 144, 498–506.
220 G. Q. Lv, H. L. Wang, Y. X. Yang, T. S. Deng, C. M. Chen, 247 B. Siyo, M. Schneider, J. Radnik, M. M. Pohl, P. Langer and
Y. L. Zhu and X. L. Hou, Green Chem., 2016, 18, 2302–2307. N. Steinfeldt, Appl. Catal., A, 2014, 478, 107–116.
221 Q. Girka, B. Estrine, N. Hoffmann, J. Le Bras, S. Marinkovic 248 Y. B. Wang, K. Yu, D. Lei, W. Si, Y. J. Feng, L. L. Lou and
and J. Muzart, React. Chem. Eng., 2016, 1, 176–182. S. X. Liu, ACS Sustainable Chem. Eng., 2016, 4, 4752–4761.
222 P. V. Rathod, S. D. Nale and V. H. Jadhav, ACS Sustainable 249 Z. H. Zhang, J. D. Zhen, B. Liu, K. L. Lv and K. J. Deng,
Chem. Eng., 2017, 5, 701–707. Green Chem., 2015, 17, 1308–1317.
223 J. J. Bozell and G. R. Petersen, Green Chem., 2010, 12, 250 N. Mei, B. Liu and Z. H. Zhang, Catal. Sci. Technol., 2015, 5,
539–554. 3194–3202.
224 S. Rajendran, R. Raghunathan, I. Hevus, R. Krishnan, 251 B. Liu, Y. S. Ren and Z. H. Zhang, Green Chem., 2015, 17,
A. Ugrinov, M. P. Sibi, D. C. Webster and J. Sivaguru, 1610–1617.
Angew. Chem., Int. Ed., 2015, 54, 1159–1163. 252 I. Lee, J. B. Joo and M. Shokouhimehr, Chin. J. Catal., 2015,
225 J. Deng, X. Q. Liu, C. Li, Y. H. Jiang and J. Zhu, RSC Adv., 36, 1799–1810.
2015, 5, 15930–15939. 253 W. J. Wang, Y. C. Li, Z. W. Kang, F. Wang and J. C. Yu,
226 https://www.avantium.com/press-releases/avantium- Appl. Catal., B, 2016, 182, 184–192.
named-2015-global-cleantech-100/. 254 B. N. Zope, S. E. Davis and R. J. Davis, Top. Catal., 2012, 55,
227 N. Jacquel, R. Saint-Loup, J. P. Pascault, A. Rousseau and 24–32.
F. Fenouillot, Polymer, 2015, 59, 234–242. 255 O. Casanova, S. Iborra and A. Corma, ChemSusChem, 2009,
228 T. Elhajj, A. Masroua, J. C. Martin and G. Descotes, Bull. 2, 1138–1144.
Soc. Chim. Fr., 1987, 855–860. 256 S. Albonetti, A. Lolli, V. Morandi, A. Migliori, C. Lucarelli
229 M. Toshinari, K. Hirokazu, K. Takenobu and M. Hirohide, and F. Cavani, Appl. Catal., B, 2015, 163, 520–530.
US232815, 2007. 257 A. Lolli, R. Amadori, C. Lucarelli, M. G. Cutrufello, E. Rombi,
230 Y. W. Cheng, X. Li, L. J. Wang and Q. B. Wang, Ind. Eng. F. Cavani and S. Albonetti, Microporous Mesoporous Mater.,
Chem. Res., 2006, 45, 4156–4162. 2016, 226, 466–475.
231 X. B. Zuo, P. Venkitasubramanian, D. H. Busch and 258 Z. Z. Miao, Y. B. Zhang, X. Q. Pan, T. X. Wu, B. Zhang,
B. Subramaniam, ACS Sustainable Chem. Eng., 2016, 4, J. W. Li, T. Yi, Z. D. Zhang and X. G. Yang, Catal. Sci.
3659–3668. Technol., 2015, 5, 1314–1322.
232 B. Saha, S. Dutta and M. M. Abu-Omar, Catal. Sci. Technol., 259 N. K. Gupta, S. Nishimura, A. Takagaki and K. Ebitani,
2012, 2, 79–81. Green Chem., 2011, 13, 824–827.
233 P. Verdeguer, N. Merat and A. Gaset, J. Mol. Catal., 1993, 260 L. Ardemani, G. Cibin, A. J. Dent, M. A. Isaacs, G. Kyriakou,
85, 327–344. A. F. Lee, C. M. A. Parlett, S. A. Parry and K. Wilson, Chem.
234 H. A. Rass, N. Essayem and M. Besson, Green Chem., 2013, Sci., 2015, 6, 4940–4945.
15, 2240–2251. 261 J. Y. Cai, H. Ma, J. J. Zhang, Q. Song, Z. T. Du, Y. Z. Huang
235 K. R. Vuyyuru and P. Strasser, Catal. Today, 2012, 195, 144–154. and J. Xu, Chem. – Eur. J., 2013, 19, 14215–14223.
236 W. Q. Niu, D. Wang, G. H. Yang, J. Sun, M. B. Wu, 262 T. Pasini, M. Piccinini, M. Blosi, R. Bonelli, S. Albonetti,
Y. Yoneyama and N. Tsubaki, Bull. Chem. Soc. Jpn., 2014, N. Dimitratos, J. A. Lopez-Sanchez, M. Sankar, Q. He,
87, 1124–1129. C. J. Kiely, G. J. Hutchings and F. Cavani, Green Chem.,
237 Y. W. Zhang, Z. M. Xue, J. F. Wang, X. H. Zhao, Y. H. Deng, 2011, 13, 2091–2099.
W. C. Zhao and T. C. Mu, RSC Adv., 2016, 6, 51229–51237. 263 A. Villa, M. Schiavoni, S. Campisi, G. M. Veith and L. Prati,
238 X. W. Han, C. Q. Li, Y. Guo, X. H. Liu, Y. G. Zhang and ChemSusChem, 2013, 6, 609–612.
Y. Q. Wang, Appl. Catal., A, 2016, 526, 1–8. 264 A. Lolli, S. Albonetti, L. Utili, R. Amadori, F. Ospitali,
239 M. A. Lilga, R. T. Hallen, J. Hu, J. F. White and M. J. Gray, C. Lucarelli and F. Cavani, Appl. Catal., A, 2015, 504,
US2008103318-A1, 2007. 408–419.

This journal is © The Royal Society of Chemistry 2018 Chem. Soc. Rev., 2018, 47, 1351--1390 | 1389
View Article Online

Review Article Chem Soc Rev

265 X. Y. Wan, C. M. Zhou, J. S. Chen, W. P. Deng, Q. H. Zhang, 279 S. Li, K. M. Su, Z. H. Li and B. W. Cheng, Green Chem.,
Y. H. Yang and Y. Wang, ACS Catal., 2014, 4, 2175–2185. 2016, 18, 2122–2128.
266 Z. Y. Gui, W. R. Cao, S. Saravanamurugan, A. Riisager, 280 B. B. Sarma, I. Efremenko and R. Neumann, J. Am. Chem.
L. F. Chen and Z. W. Qi, ChemCatChem, 2016, 8, 3636–3643. Soc., 2015, 137, 5916–5922.
267 C. M. Zhou, W. P. Deng, X. Y. Wan, Q. H. Zhang, Y. H. Yang 281 B. Karimi, M. Rafiee, S. Alizadeh and H. Vali, Green Chem.,
and Y. Wang, ChemCatChem, 2015, 18, 2853–2863. 2015, 17, 991–1000.
268 F. Kerdi, H. A. Rass, C. Pinel, M. Besson, G. Peru, B. Leger, 282 G. Grabowski, J. Lewkowski and R. Skowroński, Electro-
Published on 03 January 2018. Downloaded by National University of Kaohsiung on 22/02/2018 10:09:12.

S. Rio, E. Monflier and A. Ponchel, Appl. Catal., A, 2015, chim. Acta, 1991, 36, 1995.
506, 206–219. 283 D. J. Chadderdon, L. Xin, J. Qi, Y. Qiu, P. Krishna,
269 D. K. Mishra, H. J. Lee, J. Kim, H. S. Lee, J. K. Cho, Y. W. Suh, K. L. More and W. Z. Li, Green Chem., 2014, 16, 3778–3786.
Y. Yi and Y. J. Kim, Green Chem., 2017, 19, 1619–1623. 284 H. G. Cha and K. S. Choi, Nat. Chem., 2015, 7, 328–333.
270 A. Jain, S. C. Jonnalagadda, K. V. Ramanujachary and 285 V. P. Kashparova, V. A. Klushin, D. V. Leontyeva,
A. Mugweru, Catal. Commun., 2015, 58, 179–182. N. V. Smirnova, V. M. Chernyshev and V. P. Ananikov,
271 F. Neatu, R. S. Marin, M. Florea, N. Petrea, O. D. Pavel and Chem. – Asian J., 2016, 11, 2578–2585.
V. I. Parvulescu, Appl. Catal., B, 2016, 180, 751–757. 286 B. You, N. Jiang, X. Liu and Y. J. Sun, Angew. Chem., Int. Ed.,
272 E. Hayashi, T. Komanoya, K. Kamata and M. Hara, Chem- 2016, 55, 9913–9917.
SusChem, 2017, 10, 654–658. 287 B. You, X. Liu, N. Jiang and Y. J. Sun, J. Am. Chem. Soc.,
273 X. W. Han, C. Q. Li, X. H. Liu, Q. N. Xia and Y. Q. Wang, 2016, 138, 13639–13646.
Green Chem., 2017, 19, 996–1004. 288 N. Jiang, B. You, R. Boonstra, I. M. T. Rodriguez and
274 C. V. Nguyen, Y. T. Liao, T. C. Kang, J. E. Chen, Y. J. Sun, ACS Energy Lett., 2016, 1, 386–390.
T. Yoshikawa, Y. Nakasaka, T. Masuda and K. C. W. Wu, 289 M. Kroger, U. Prusse and K. D. Vorlop, Top. Catal., 2000,
Green Chem., 2016, 18, 5957–5961. 13, 237–242.
275 K. S. Lakhi, D. H. Park, K. Al-Bahily, W. Cha, 290 M. L. Ribeiro and U. Schuchardt, Catal. Commun., 2003, 4,
B. Viswanathan, J. H. Choy and A. Vinu, Chem. Soc. Rev., 83–86.
2017, 46, 72–101. 291 G. S. Yi, S. P. Teong, X. K. Li and Y. G. Zhang, ChemSusChem,
276 S. E. Davis, B. N. Zope and R. J. Davis, Green Chem., 2012, 2014, 7, 2131–2137.
14, 143–147. 292 G. S. Yi, S. P. Teong and Y. G. Zhang, ChemSusChem, 2015,
277 S. E. Davis, A. D. Benavidez, R. W. Gosselink, J. H. Bitter, 8, 1151–1155.
K. P. de Jong, A. K. Datye and R. J. Davis, J. Mol. Catal. A: 293 G. S. Yi, S. P. Teong and Y. G. Zhang, Green Chem., 2016,
Chem., 2014, 388, 123–132. 18, 979–983.
278 L. C. Gao, K. J. Deng, J. D. Zheng, B. Liu and Z. H. Zhang, 294 S. G. Wang, Z. H. Zhang and B. Liu, ACS Sustainable Chem.
Chem. Eng. J., 2015, 270, 444–449. Eng., 2015, 3, 406–412.

1390 | Chem. Soc. Rev., 2018, 47, 1351--1390 This journal is © The Royal Society of Chemistry 2018

You might also like