One-Pot Synthesis of 2,5-Furandicarboxylic Acid ... - ACS Publications

*Corresponding author: Jiayu Xin, Associate professor, Ph.D., E-mail: [email protected],. 11. Tel./fax: +86-10-62558174. 12. *Corresponding author: Suoj...
0 downloads 0 Views 655KB Size
Subscriber access provided by READING UNIV

Article

One-Pot Synthesis of 2,5-Furandicarboxylic Acid from Fructose in Ionic Liquids Dongxia Yan, Gongying Wang, Kai Gao, Xingmei Lu, Jiayu Xin, and Suojiang Zhang Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.7b04947 • Publication Date (Web): 16 Jan 2018 Downloaded from http://pubs.acs.org on January 20, 2018

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

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

Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

One-Pot Synthesis of 2,5-Furandicarboxylic Acid from

1

Fructose in Ionic Liquids

2

Dongxia Yan,†,‡,§ Gongying Wang,† Kai Gao,‡,§ Xingmei Lu,‡,§ Jiayu Xin,*,‡ and Suojiang

3

Zhang*,‡

4 5



6

China

7



8

Engineering, State Key Laboratory of Multiphase Complex Systems, Institute of Process

9

Engineering, Chinese Academy of Sciences, Beijing, 100190, P. R. China.

Chengdu Institute of Organic Chemistry, Chinese Academy of Sciences, Chengdu, 610041, P. R.

Beijing Key Laboratory of Ionic Liquids Clean Process, Key Laboratory of Green Process

10

§

11

*

12

Tel./fax: +86-10-62558174

13

*

14

ABSTRACT

15

2,5-Furandicarboxylic acid (FDCA), which is usually produced from HMF catalyzed

16

by noble metal catalysts, is an important bio-based monomer for the degradable

17

polymer polyethylene furandicarboxylate (PEF). In order to reduce the high costs of

18

starting material and catalysts, a novel approach for the direct conversion of fructose

19

into FDCA was developed by employing [Bmim]Cl as a solvent with non-noble metal

20

(Fe-Zr-O) as a catalyst. Relatively high FDCA yield was obtained at full fructose

21

conversion under optimal conditions. The kinetic study revealed that the oxidation of

University of Chinese Academy of Sciences, Beijing 100049, P. R. China.

Corresponding author: Jiayu Xin, Associate professor, Ph.D., E-mail: [email protected],

Corresponding author: Suojiang Zhang, Professor, Ph.D., E-mail: [email protected]

1

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

intermediate FFCA to FDCA possessed the highest activation energy, indicating this

2

step is most affected by reaction temperature. Additionally, in the IL-promoted

3

reaction system, other biomass sources, such as glucose, galactose, mannose, starch

4

and cellulose also can be directly converted, with lower FDCA yield compared with

5

that of fructose due to the ineffective isomerization of aldohexoses into fructose.

6

KEYWORDS: Fructose, 5-Hydroxymethylfurfural, 2,5-Furandicarboxylic acid,

7

Non-noble metal catalyst, Ionic liquids

8

1. INTRODUCTION

9

With growing concerns on environmental pollution and the depletion of nonrenewable

10

fossil resources, the interests in fuels and chemicals synthesized from biomass

11

resources have been growing.1-3 2,5-Furandicarboxylic acid (FDCA), as one of the

12

most important biomass-based platform molecules produced from biomass-derived

13

5-hydroxymethylfurfural (HMF) or C6-based carbohydrates,4 has a similar structure

14

with petroleum-derived terephthalic acid (TPA), is considered as a potential substitute

15

for TPA in the production of polyethylene terephthalate (PET).5, 6 More importantly,

16

FDCA-based polyethylene furandicarboxylate (PEF) is a renewable and degradable

17

bio-based polymer with better performance in terms of thermal stability, mechanical

18

and barrier properties than its analogue, TPA-based PET.7-9 Therefore, the exploration

19

of FDCA production has drawn tremendous attention in recent years.4, 10

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

Traditionally, FDCA is produced from the oxidation of HMF. In early reports, the

2

oxidation of HMF to FDCA was performed by using high cost and highly toxic

3

stoichiometric oxidants, such as chromate (CrO42−), dichromate (Cr2O72−) and

4

permanganate (MnO4−) salts.11 Substantial progresses have been achieved in FDCA

5

production from HMF using air or molecular oxygen as oxidants catalyzed by various

6

supported noble metal catalysts in (basic) aqueous media, such as Au,12 Pt,13 Pd,14

7

Ru15 or their alloys.16 Recently, the oxidation of HMF to FDCA has also been

8

accomplished using earth-abundant metals, including nano-Fe3O4-CoOx catalyst,

9

Co(II)-meso-tetra(4-pyridyl)-porphyrin catalyst,17,

18

MnxFey mixed oxide,19 MnO2

10

and its mixed oxide MnOx-CeO2.20, 21 Generally, these processes require not abundant

11

and very expensive HMF as a starting material, which limits the large-scale and

12

sustainable production of FDCA, thus becomes a barrier for its commercial

13

applications. Therefore, it is more attractive to produce FDCA from largely available

14

and inexpensive source.

15

In recent years, some researchers have tried to synthesize FDCA directly from

16

fructose. For example, Kröger et al. used a biphasic water/methyl isobutyl ketone

17

(MIBK) system separated by a polytetrafluoroethylene (PTFE) membrane to convert

18

fructose into FDCA, resulting in a moderate yield of 25%.22 Later, Ribeiro and

19

Schuchardt synthesized a Co(acac)3-gel bifunctional catalyst for the one-pot

20

conversion of fructose into FDCA, affording a 72% fructose conversion.23 Recently,

21

Zhang et al. designed a two step method for the conversion of fructose into FDCA.24, 3

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

25

2

chloride (PBnNH3Cl) resin as catalysts, and then oxidized into FDCA over Au/HT

3

catalyst after purification with water-extraction. Zhang and co-workers have

4

accomplished a one-pot conversion of sugars into FDCA in a triphasic system, which

5

is composed of tetraethylammonium bromide (TEAB) or water-methyl isobutyl

6

ketone (MIBK)-water.26 In the reaction system, HMF was firstly formed in TEAB or

7

water phase using Amberlyst-15 solid acid catalyst, and then transferred to water

8

phase through the bridge phase of MIBK, where HMF is further oxidized to FDCA

9

over Au8Pd2/HT catalyst in the presence of Na2CO3. In order to reduce the cost of the

10

catalyst and facilitate the catalyst recycle, Zhang et al. employed Fe3O4@SiO2-SO3H

11

and Nano-Fe3O4-CoOx magnetic catalysts for the conversion of fructose into FDCA

12

through a two-step process.17 Fructose is firstly converted to HMF over the

13

Fe3O4@SiO2-SO3H acid catalyst in DMSO, subsequently, the formed HMF was

14

oxidized to FDCA over Nano-Fe3O4-CoOx catalyst by using t-BuOOH as an oxidant

15

after the removal of Fe3O4@SiO2-SO3H catalyst. Similarly, Yang and co-workers also

16

used magnetic solid acid (Fe3O4-RGO-SO3H) and ZnFe1.65Ru0.35O4 catalysts for a

17

one-pot, two-step conversion of fructose into FDCA.27 HMF was firstly produced in

18

dry DMSO using Fe3O4-RGO-SO3H catalyst, and then oxidized into FDCA over

19

ZnFe1.65Ru0.35O4 catalyst in the mixture of DMSO and H2O after separation of

20

magnetic solid acid from the reaction solution.

HMF was firstly produced in isopropanol using HCl or poly-benzyl ammonium

4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

As discussed above, FDCA is usually produced by two-step processes from

2

fructose: the dehydration of fructose to HMF was firstly catalyzed by acid catalyst,

3

then the formed HMF was oxidized to FDCA usually conducted in basic environment.

4

Importantly, prior to the second reaction step, separation of the acid catalyst and

5

purification of HMF are usually required. The multistep process not only leads to high

6

costs, but also leads to a great consumption of energy, solvent, and time. In addition,

7

the second step for the oxidation of HMF to FDCA usually employs noble metal

8

catalyst, making the price of FDCA less competitive than TPA. Therefore, a non-noble

9

metal catalyzed one-pot process for the direct conversion of fructose into FDCA

10

without additional separation step is highly desired.

11

In this work, as ionic liquids (ILs) usually used as an excellent solvent and

12

co-catalyst for biomass pretreatment and conversion,28-30 a systematic study on the

13

direct synthesis of FDCA from fructose using non-noble metal catalyst in ILs was

14

presented. The effects of reaction parameters including acid catalyst, reaction

15

temperature, reaction time, O2 pressure and catalyst dosage on the catalytic

16

conversion of fructose were investigated. A plausible reaction mechanism and

17

pathway were proposed based on experimental data and previous studies, and HMFA

18

instead of DFF as the key intermediate for the formation of FDCA was confirmed. To

19

the best of our knowledge, there is no report about the direct synthesis of FDCA from

20

fructose in ILs, let alone using non-noble metal as an oxidation catalyst. It is believed

5

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 32

1

that the novel IL-promoted non-noble metal catalytic reaction system for the direct

2

conversion of fructose into FDCA reported here is an attractive process.

3

2. EXPERIMENTAL SECTION

4

2.1. Materials. HMF (98%), 5-hydroxymethyl-2-furancarboxylic acid (HMFA) (95%),

5

2,5-diformylfuran (DFF) (98%), 5-formyl-2-furancarboxylic acid (FFCA) (98%) and

6

FDCA (97%) were purchased from J&K Chemical Co. Ltd. (Beijing, China).

7

Zr(NO 3 ) 4 ·5H 2 O, CrCl3·6H 2 O, fructose (99%), glucose, galactose, mannose and

8

cellulose were supplied by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).

9

Fe(NO 3 ) 3 ·9H 2 O and starch were purchased from Xilong Chemical Co., Ltd.

10

(Guangdong, China). Sulfuric acid (guaranteed reagent) and ammonia (wt% 25%~

11

28%) were purchased from Beijing Chemicals Co. Ltd. (Beijing, China). O2 (99.9%)

12

was supplied by Beijing Beiwen Gas Factory. Nafion-NR50 and Amberlyst-15 were

13

supplied

14

1-Butyl-3-methylimidazolium chloride ([Bmim]Cl) was commercially available

15

(purchased from Linzhou Keneng Material Technology Co., Ltd.) and dried under

16

vacuum at 60 oC for 24 h before use. The commercial chemicals were of analytical

17

grade and therefore were used as received.

by

Alfa

Aesar

Chemical

Co.,

Ltd.

(Shanghai,

China).

18

2.2. Catalyst Preparation and Characterization. The Fe0.6Zr0.4O2 catalyst was

19

prepared by a hydrothermal method as reported previously.31, 32 The required amounts

20

of Fe(NO3)3·9H2O and Zr(NO3)4·5H2O (atomic ratio of Fe/Zr=6:4) were dissolved in

6

ACS Paragon Plus Environment

Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

deionized water and well mixed. The total concentration of Fe and Zr cation was 0.25

2

mol/L. A brown slurry was precipitated by gradually dripping the mixed nitrate

3

solution into an ammonia solution (5 mol/L) while maintaining the pH at about 10

4

with stirring. After stirring and settling for 2 h, respectively, 65 ml of slurry was

5

transferred to 100 ml Teflon-lined stainless steel autoclaves and reacted at 220 oC for

6

48 h. After cooling, the sample was washed with distilled water and ethanol until pH

7

of the decanted water was ~7 and then air-dried at 110 oC for 12 h. Experimental

8

details for catalyst characterization, including X-ray diffraction analysis (XRD), BET

9

surface area analysis (BET) and transmission electron microscope (TEM) were

10

described in Supporting Information.

11

2.3. Catalytic Reactions and Products Analysis. The preparation of FDCA from

12

fructose was carried out in a 50 ml batch-type Teflon-lined stainless-steel autoclave.

13

Typically, 1.0 g [Bmim]Cl ILs, 0.1 mmol fructose, 0.01 g Fe0.6Zr0.4O2 catalyst, as well

14

as 0.01 g Amberlyst-15 were added into the autoclave, and purged 3 times with O2

15

then pressurized to 2 MPa at room temperature. The reaction was carried out at a

16

given reaction time and temperature by vigorous stirring with a magnetic stirrer. After

17

reaction, the products were diluted with deionized H2O and filtered using PTFE 0.2

18

µm filters, subsequently analyzed by HPLC. Fructose, HMF, HMFA, DFF, FFCA and

19

FDCA were analyzed by a Shimazu LC-20A HPLC instrument equipped with

20

refractive index (RI) and photo-diode array (PDA) detectors. The HPLC utilized a

21

Bio-Rad Aminex HPX-87H column (300 mm × 7.8 mm ) and 5 mM H2SO4 aqueous 7

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 32

1

solution as a mobile phase flowing at 0.6 ml min−1 and 35 oC to perform the

2

separation. The retention times and calibrations for observed products were

3

determined by injecting known concentrations of standard products, and quantified

4

using an external standard calibration curve method.

5

3. RESULTS AND DISCUSSION

6

3.1. Catalyst Preparation and Characterization. The Fe0.6Zr0.4O2 catalyst was

7

prepared by a hydrothermal method as reported previously.31,

8

as-synthesized Fe0.6Zr0.4O2 catalyst has high degrees of acidity and basicity, as well as

9

good reducibility and oxygen mobility, which are beneficial to the redox properties

10

and activity of the catalyst, thus favor the oxidation of HMF to FDCA. As discussed

11

in our previous reports,31, 32 the powder XRD pattern revealed that the Fe0.6Zr0.4O2

12

was successfully synthesized, which was consisted of ZrO2, Fe2O3 and their solid

13

solutions (Figure S1a). TEM image revealed the particle sizes of the catalyst powder

14

was about 6.9 nm in diameter (Figure S1b). The BET surface area of the synthesized

15

catalyst was found to be 96 m2/g.

32

And the

16

3.2. Catalytic Conversion of Fructose into FDCA in ILs by Various Acid

17

Catalysts. Several acid catalysts combined with Fe0.6Zr0.4O2 oxidation catalyst were

18

used for the synthesis of FDCA from fructose, the results are summarized in Table 1.

19

It can be seen that the FDCA yield is strongly influenced by the acid catalyst. The use

20

of Amberlyst-15 solid acid together with Fe0.6Zr0.4O2 catalyst gave a 46.4% yield of

8

ACS Paragon Plus Environment

Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

FDCA, while Nafion-NR50 and CrCl3·6H2O catalysts exhibited moderate to low

2

yields of FDCA (29.6 and 24.4%, respectively), although they also showed complete

3

conversions of fructose (Table 1, entries 2 and 3). These results indicated that the

4

Amberlyst-15 is the most suitable acid catalyst for the reaction. Importantly, relatively

5

low FDCA yield (15.1 and 29.2%, respectively) were observed for Amberlyst-15 and

6

Fe0.6Zr0.4O2 catalyst used alone (Table 1, entries 4 and 5), revealing that the catalytic

7

conversion of fructose into FDCA needs the cooperation of the dehydration catalyst

8

and oxidation catalyst. Additionally, according to the blank experiment, the reaction

9

occurred even in the absence of any catalyst, an 11.4% FDCA yield (Table 1, entry 6)

10

was obtained in ILs at 160 oC after a reaction time of 24 h. This implies that the ILs

11

exhibited an excellent synergistic catalytic effect for the catalytic conversion of

12

fructose.

13 14 15 16 17 18 19

Table 1. Synthesis of FDCA from fructose by using different acids and Fe0.6Zr0.4O2 catalysts.

9

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Catalyst

Page 10 of 32

Yield (%)

Fructose

Entry Acid

1 2

Oxidation

FDCA

FFCA

HMFA

HMF

conversion (%)

1

Amberlyst-15

Fe0.6Zr0.4O2

46.4±1.7

3.3±0.2

0.2±0.0

2.9±0.2

100.0±0.0

2

Nafion-NR50

Fe0.6Zr0.4O2

29.6±1.4

1.7±0.3

0.4±0.1

3.2±0.4

100.0±0.0

3

CrCl3·6H2O

Fe0.6Zr0.4O2

24.4±1.1

1.9±0.3

0.2±0.0

1.1±0.1

100.0±0.0

4

Amberlyst-15

15.1±0.8

2.0±0.3

0.1±0.0

1.2±0.2

100.0±0.0

5



29.2±0.9

2.4±0.4

0.3±0.0

1.2±0.2

100.0±0.0

6



11.4±0.8

0.9±0.1

0.1±0.0

0.9±0.1

100.0±0.0

— Fe0.6Zr0.4O2 —

Reaction conditions: fructose (0.1 mmol), acid catalyst (0.01 g), Fe0.6Zr0.4O2 catalyst (0.01 g), [Bmim]Cl (1 g), O2 (2 MPa), 160 oC, 24 h.

3

Based on the above results we may assume that the acid catalyst played an

4

important role to enhance the yield of FDCA. In order to reveal the reason of catalytic

5

performance of various acid catalysts, Amberlyst-15, Nafion-NR50 and CrCl3·6H2O,

6

as well as [Bmim]Cl itself were used for the catalytic dehydration of fructose into

7

HMF. Their catalytic performances are shown in Figure S2. Among these results,

8

>84% HMF yield was obtained for Amberlyst-15 and Nafion-NR50, which was much

9

better than that of CrCl3·6H2O/[Bmim]Cl under the same conditions. Especially for

10

Amberlyst-15 catalyst, >83% HMF yield was attained as short as 1 min, which was

11

consistent with the previous report,33 that a HMF yield of 82.2% was obtained from

12

fructose as short as 1 min in [Bmim]Cl ILs when the reaction was catalyzed by

13

Amberlyst-15 catalyst. The rapid conversion of fructose to HMF can avoid the

14

oxidation or degradation of fructose into byproducts, resulting in higher FDCA yield

15

when the reaction was carried out in the presence of Amberlyst-15 and Fe0.6Zr0.4O2 10

ACS Paragon Plus Environment

Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

catalysts. Compared with the use of [Bmim]Cl alone, the combined use of

2

CrCl3·6H2O provided a high FDCA yield in a shorter time. The results indicate that

3

both Brønsted and Lewis acid sites are active in fructose conversion into HMF.

4

However, as reported, the Lewis acid sites are also able to catalyze the formation of

5

humins from sugar,34, 35 which should probably be responsible for the relatively low

6

FDCA yield catalyzed by CrCl3·6H2O and Fe0.6Zr0.4O2 catalysts in this work. Since

7

Amberlyst-15 resin showed the best catalytic activity for HMF and FDCA production

8

from fructose under our experimental conditions, this solid acid catalyst was chosen to

9

use together with Fe0.6Zr0.4O2 catalyst for the catalytic conversion of fructose in the

10

followed experiments.

11

3.3. Effects of Reaction Parameters on Catalytic Conversion of Fructose into

12

FDCA. FDCA was prepared by the catalytic conversion of fructose over

13

Amberlyst-15 and Fe0.6Zr0.4O2 catalysts using O2 as the oxidant. The effects of

14

reaction time, reaction temperature, oxygen pressure, and catalyst amount on the

15

fructose conversion and FDCA yield were systematically investigated, as shown in

16

Figure 1.

11

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

100

100

a

Yield/Conversion (%)

Yield/Conversion (%)

50 40 30

FDCA yield FFCA yield HMFA yield HMF yield Fructose conversion

20 10

b

FDCA yield FFCA yield HMFA yield HMF yield Fructose conversion

80 60 40 20 0

0 4

8

12

16

20

24

100

28

120

Time (h)

140

160

180

Temperature (℃)

100

100

c

40 30

FDCA yield FFCA yield HMFA yield HMF yield Fructose conversion

20 10

d

50

Yield/Conversion (%)

50

Yield/Conversion (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 32

40 30

FDCA yield FFCA yield HMFA yield HMF yield Fructose conversion

20 10 0

0 0

1

2

3

0

5

10

15

20

Catalyst amount (mg)

O2 pressure (MPa)

1 2

Figure 1. Effects of different reaction conditions on the preparation of FDCA by catalytic

3

conversion of fructose using Fe0.6Zr0.4O2 and Amberlyst-15 catalysts: (a) reaction time, (b)

4

reaction temperature, (c) oxygen pressure, and (d) catalyst loading. If not specified, the default

5

reaction conditions were as follows: fructose (0.1 mmol), Amberlyst-15 (0.01 g), Fe0.6Zr0.4O2

6

catalyst (0.01 g), [Bmim]Cl (1 g), O2 (2 MPa), 160 oC, 24 h.

7

Figure 1a shows the evolutions of fructose conversion and products yields with

8

the reaction time in the presence of Amberlyst-15 and Fe0.6Zr0.4O2 catalysts at 160 oC

9

and 2 MPa O2. It is noted that at the initial stage of the reaction (4 h), the formation of

10

HMF was clearly observed (22% yield) with a complete conversion of fructose, which

11

indicated that fructose was first dehydrated into HMF over protic acid sites of

12

Amberlyst-15. Thereafter, with the increase of reaction time, the formed HMF was 12

ACS Paragon Plus Environment

Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

further converted to FDCA, thus, the yield of HMF gradually decreased with the

2

increase of FDCA yield. The highest FDCA yield of 46.4% was obtained after 24 h,

3

and then decreased. This may be attributed to the transformation of FDCA to other

4

byproducts with the increase of reaction time. In addition, the HMFA intermediate

5

rather than the DFF intermediate can be detected with little content, while the FFCA

6

content firstly increased then decreased with the increase of reaction time. That is

7

because FFCA was further oxidized into FDCA.

8

The effect of reaction temperature on FDCA yield is presented in Figure 1b. The

9

results show that the reaction temperature had a remarkable effect on the conversion

10

of fructose into FDCA. Higher reaction temperature clearly produced higher FDCA

11

yield. The conversion of fructose at 100 oC almost produced no FDCA after 24 h,

12

while could proceed slowly at 120 oC with a low FDCA yield of 2.9% after 24 h.

13

However, the dehydration of fructose at 120 oC could produce HMF yield of 82.2%.33

14

In our study, we also obtained 84.6% HMF yield as short as 0.5 min at 120 oC (see

15

Table S1). Thus the main reason for the low FDCA yield from fructose at 120 oC was

16

that oxidation of the formed HMF into FDCA at the temperature was not effective,

17

which was consistent with the previous report.31 In order to improve the yield of

18

FDCA, the reaction temperature was further increased to 140 oC. As can be seen from

19

Figure 1b, FDCA yield greatly increased. With a further increase in the temperature to

20

160 oC and 180 oC, FDCA yield further increased to 46.4%, and then sharply

21

decreased to 28.1%. This is most likely due to the poor stability of fructose and HMF 13

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

and their favored condensation and degradation reactions to the formation of humins

2

at higher temperatures36-38 (see Figure S3). This result was consistent with the

3

previous report on the effect of reaction temperature on synthesis of FDCA from

4

HMF.31 Therefore, 160 oC was an appropriate reaction temperature for the conversion

5

of fructose into FDCA in our catalytic system.

6

The influence of O2 pressure on the catalytic conversion of fructose was

7

investigated and is plotted in Figure 1c. Initially, the reaction was performed with 2

8

MPa oxygen at 160 oC for 24 h over Fe0.6Zr0.4O2 catalyst in [Bmim]Cl. Under these

9

conditions, 46.4% FDCA yield was obtained (Figure 1c). By decreasing the oxygen

10

pressure to 1 MPa and atmospheric pressure under the same reaction conditions, the

11

FDCA yield decreased to 31.7% and 22.1%, respectively. It is reported that the

12

solubility of oxygen increases with the increase of oxygen pressure.12 It can be

13

assumed that a higher amount of oxygen is present on the active metal sites and hence

14

lead to a better performance for FDCA production if there is a higher concentration of

15

oxygen in the reaction system. This is consistent with previous report that higher

16

oxygen pressure favors higher FDCA yield.39 With a further increase in oxygen

17

pressure from 2 MPa to 3 MPa, the FDCA yield slightly decreased from 46.4% to

18

45.3%, without significant change. It was because the oxygen pressure of 2 MPa was

19

sufficient for the reaction.

20

Catalytic conversion of fructose into FDCA was also carried out with various

21

amounts of Fe0.6Zr0.4O2 catalyst in order to study the effect of catalyst loading. As 14

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

shown Figure 1d, in all cases, fructose was fully converted, which indicated that the

2

oxidation catalyst loading had little influence on the conversion of fructose, which

3

probably mainly depended on other reaction conditions, such as the dehydration

4

catalyst, reaction time, reaction temperature, and reaction solvent. However, the

5

catalyst loading showed a significant effect on the yield of FDCA, which increased

6

with the increase of Fe0.6Zr0.4O2 catalyst amount from 0 mg (blank experiment) to 10

7

mg. A FDCA yield of 15.1% was obtained without oxidation catalyst and sharply

8

increased to 20.7% and 46.4% by the increase of catalyst loading to 5 mg and 10 mg,

9

respectively. However, FDCA yield was not further increased with further increased

10

the catalyst amount to 15 and 20 mg. The possible reason might be that the active

11

catalytic sites of 10 mg of Fe0.6Zr0.4O2 catalyst were sufficient for the oxidation of

12

formed HMF from fructose.

13

3.4. Mechanism and Kinetics Studies. In the catalytic process for the production

14

of FDCA from fructose, other byproducts and intermediates, including HMF, HMFA

15

and FFCA were also detected, while DFF was not detected (see the time course of

16

fructose conversion and product yield in Figure 1a). Taken together with the

17

investigated effect of reaction temperature, O2 pressure, catalyst amount (Figure 1b,

18

1c and 1d), and previous studies,40-44 the plausible reaction mechanism for the

19

catalytic conversion of fructose into FDCA over Amberlyst-15 and Fe0.6Zr0.4O2

20

catalysts was proposed. As shown in Scheme S1, the conversion of fructose into

21

FDCA is a multistep process which involves various important intermediates, as 15

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 32

1

discussed above. Firstly, the fructose dehydrate to form HMF through loss of three

2

water molecules in the presence of Amberlyst-15 catalyst, then the formed HMF is

3

further oxidized to generate HMFA, rather than DFF, which finally undergoes

4

sequential oxidation to form FFCA and FDCA. In order to confirm the existence of

5

intermediate products (particularly HMF, HMFA and FFCA), the HPLC

6

chromatograms were recorded at various interval of times (Figure S4). The proposed

7

mechanism is in line with the previous studies about the mechanism for the

8

dehydration of fructose to HMF,40-42 and the mechanism for the oxidation of HMF to

9

FDCA.32, 43, 44 It is worth mention that ILs can help a lot to convert fructose to FDCA.

10

Because [Bmim]Cl is an ideal solvent for fructose and FDCA, and has relatively high

11

catalytic activity for HMF formation and conversion.31

O

OH

OH

k1

O

HO HO

Fructose

12

OH

k2

O HO

O HO

OH

k3

O

O

k4

O

O

HO

OH O O

O

OH

HMF

HMFA

FFCA

FDCA

Figure 2. A reaction pathway for fructose conversion to FDCA.

13

Based on the experimental results, a reaction scheme used for the development of

14

the kinetic model of fructose conversion was proposed, as shown in Figure 2, where ki

15

represents various rate constants. The fructose conversion mainly consists of four

16

steps: (1) fructose dehydration to form HMF, (2) formed HMF oxidation to produce

17

HMFA, (3) HMFA further oxidation to produce FFCA, and (4) finally FFCA

18

oxidation to form FDCA. Recent literatures reported that the orders of reactions for

19

the conversion of fructose to HMF and HMF to FDCA, were revealed as first orders.27, 16

ACS Paragon Plus Environment

Page 17 of 32

1

33

2

was performed based on first order initially.

According to this, the kinetic analysis of the conversion of fructose to FDCA in ILs

-2.0

a

-2.5 -3.0 -3.5 80℃ 100℃ 120℃ 140℃

-4.0 -4.5

b

-2.4

ln( nHMF/mmol)

ln( nFructose/mmol)

-2.0

-2.8 -3.2 -3.6 120℃ 140℃ 160℃ 180℃

-4.0 -4.4

-5.0 -4.8

0

10

20

30

40

50

60

70

80

90

0

60

Time (s)

180

240

-2

c

-2.5

ln( nFFCA/mmol)

-3.5 120℃ 140℃ 160℃ 180℃

-4.0 -4.5

0

60

120

d

-3

-3.0

-5.0

120

Time (min)

-2.0

ln( nHMFA/mmol)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

-4 -5 120℃ 140℃ 160℃ 180℃

-6 -7

180

240

300

360

420

0

480

60

120

180

240

300

360

Time (min)

Time (min)

3 4

Figure 3. Ln (n) versus reaction time for (a) fructose, (b) HMF, (c) HMFA, and (d) FFCA

5

conversion.

6

To study the reaction kinetics, the effect of temperature and time on production of

7

FDCA from fructose was investigated. Plots of ln (n) (n is the various moles of

8

material) versus reaction time (t) were made to obtain first order kinetic constants.

9

The linearity of Figure 3 supported the first order assumption for conversion of

10

fructose to FDCA, therefore the validity of the selected first order kinetic model was

11

confirmed, and the reaction rate constants calculated from the slopes of the line at

17

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1

different reaction temperature (see Table S2-Table S5) were credible. As expected, all

2

rate constants increased by increasing reaction temperature. The activation energy (Ea)

3

and pre-exponential factor (A) were obtained by applying the following Arrhenius

4

equation based on the experimental data:

5

(1)

6

where k is the rate constant, A is the pre-exponential factor, R is the ideal gas constant,

7

and Ea is the activation energy.

6 Fructose conversion HMF conversion HMFA conversion FFCA conversion

4 2

Ea = 59.7 kJ/mol lnA = 23.8

-1

ln k (h )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 32

Ea = 110.2 KJ/mol

0

lnA = 30.3 Ea = 82.7 KJ/mol

-2

lnA = 22.3

Ea = 86.4 KJ/mol

-4 2.1

lnA = 22.5

2.2

2.3

2.4

2.5

3 -1

2.6

2.7

2.8

2.9

-1

10 T (K ) 8 9

Figure 4. Arrhenius plots of ln k based on different temperature.

10

According to the reaction rate constants (see Table S2-Table S5 and Figure 3),

11

Arrhenius plots were generated, as shown in Figure 4. The values of A were obtained

12

from the intercept of the plots, while the values of Ea were obtained from the slope,

18

ACS Paragon Plus Environment

Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

respectively. An activation energy of 59.7 kJ/mol was determined for fructose

2

conversion, which was comparatively lower than previous value reported in the

3

literature (ca. 65.2 kJ/mol).33 As reported, the variation in the activation energy of a

4

reaction is possibly due to several factors such as the catalyst types, reaction solvents

5

and heating method.45 Compared with the other two oxidation steps of HMF to HMFA

6

(82.7 kJ/mol) and HMFA to FFCA (86.4 kJ/mol), the oxidation of FFCA to FDCA

7

possessed a higher activation energy (110.2 kJ/mol), which is in line with the previous

8

study.46 This indicates that this reaction is most affected by temperature, thus, high

9

temperature enhances the formation of FDCA from FFCA over Fe0.6Zr0.4O2 catalyst in

10

ionic liquids.

11

3.5. Investigation on the Catalyst Recycling. To test the reusability of the

12

catalyst, Amberlyst-15 and Fe0.6Zr0.4O2 catalysts were successively reused for four

13

times under the optimal conditions for the catalytic conversion of fructose into FDCA.

14

After the reaction, the catalysts were thoroughly washed with ethanol for three times

15

before vacuum-drying at 70 oC overnight, and then the catalysts were reused. It is

16

important to note that, during the recycle experiments, some insoluble humins were

17

also formed, and mixed together with the catalysts. However, the unique dissolving

18

ability of [Bmim]Cl ILs can make it dissolved completely under the reaction

19

conditions, thus, the adsorption of humins on the active sites of the catalysts could be

20

avoid. As shown in Figure 5, there was slight loss of activity in the yield of FDCA

21

duo to the accumulation of humus during the recycle experiments. In the first run, 19

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1

fructose conversion and FDCA yield were 100% and 46.4%, respectively, while in the

2

fourth run, they were still maintained at 100% and 38.2%. This phenomenon indicates

3

that Amberlyst-15 and Fe0.6Zr0.4O2 catalysts are relatively stable during the recycle

4

experiments.

100

Yield/Conversion (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fructose conversion FDCA yield FFCA yield HMFA yield HMF yield

80 60 40 20 0

1

2

3

4

Run No.

5 6

Figure 5. Recycle experiments of Amberlyst-15 and Fe0.6Zr0.4O2 catalysts. Reaction conditions:

7

fructose (0.1 mmol), Amberlyst-15 (0.01 g), Fe0.6Zr0.4O2 catalyst (0.01 g), [Bmim]Cl (1 g), O2 (2

8

MPa), 160 oC, 24h.

9

3.6 Preparation of FDCA by the One-Pot Catalytic Conversion of Other

10

Biomass Source. As reported,47, 48 besides fructose, HMF can be prepared by other

11

biomass source, such as glucose, galactose, mannose, starch and cellulose. In this

12

paper, the biomass sources mentioned above were also used as starting materials to

13

prepare FDCA by a one-pot catalytic conversion method using Amberlyst-15 and

14

Fe0.6Zr0.4O2 catalysts. The results are summarized in Table S6. Unfortunately, except

15

fructose, other biomass sources exhibited very low yields of FDCA, even though they 20

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

showed high conversions. The possible reason might be that the Brønsted acidic resin,

2

Amberlyst-15, is an effective catalyst for the dehydration of fructose to HMF, which

3

can further be oxidized into FDCA. However, it is not a suitable catalyst for the

4

isomerization of glucose to fructose, which is necessary before fructose dehydration

5

to HMF, and can greatly promote by Lewis acid.49 In addition, the synthesis of FDCA

6

from starch and cellulose needs to go through hydrolysis, isomerization, dehydration

7

and oxidation four procedures. In the ILs solvent, hydrolysis of cellulosic

8

polysaccharides cannot carry out effectively in the absence of water, which imposed

9

restriction for the following reaction, resulting in low FDCA yield.

10

3.7 Comparisons among various FDCA production methods. In order to

11

evaluate the advantages of our reaction system for the synthesis of FDCA from

12

fructose, the developed method in this study was compared with others as shown in

13

Table 3. The supported noble metal catalysts showed excellent performances for

14

FDCA production from HMF (Table 3, Entries 1 and 2). However, these methods not

15

only required to use additional base, but also required to use expensive HMF as a

16

starting material, and expensive noble metals as catalysts, which limit the industrial

17

scale production of FDCA. Widely available and inexpensive fructose has also been

18

reported for the synthesis of FDCA (Table 3, Entries 3-5) with moderate to high

19

FDCA yields. Similar to the methods for the synthesis of FDCA from HMF, high cost

20

noble metals and high concentrations of alkaline were usually employed in these

21

reaction system. Additionally, these methods usually employed complicated biphasic 21

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

or triphasic reaction systems, or employed two-step processes: (1) the dehydration of

2

fructose to HMF, and (2) the oxidation of formed HMF to FDCA after purification or

3

after the removal of the acid catalysts. There was also report about the FDCA

4

synthesis from fructose (Table 3, Entry 6) without additional base or noble metal

5

catalyst. However, the reaction system employed t-BuOOH as the oxidant, which was

6

high cost and not friendly to the environment. Although, our method showed lower

7

FDCA yield as compared with some of the above mentioned methods, especially

8

when use HMF as a starting material, it demonstrated some advantages. The

9

non-noble metal catalyzed direct conversion of fructose into FDCA in [Bmim]Cl ILs

10

used a earth-abundant and inexpensive starting material, catalyst and oxidant, which

11

makes the FDCA production much more economical. More importantly, the use of

12

[Bmim]Cl ILs as a unique reaction solvent not only avoids the use of additional base

13

in the reaction, but also makes the synthesis of FDCA from fructose use a one-step

14

process in one-pot, which leads to a great saving in energy, solvent, and time because

15

of the avoidance of the complicated separation steps.

16 17 18 19 20 22

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1 2

Table 2. Comparison our method for the synthesis of FDCA with other methods

Starting

FDCA

Entry

Catalyst

Oxidant

Base

Process

material

1

2

Refs. yield (%)

HMF

Au/CeO2

HMF

γ-Fe2O3@HAP-Pd(0)

4 Equiv of

One-pot,

NaOH

one-step

0.5 Equiv of

One-pot,

K2CO3

one-step

Air

O2

99

12

92.9

14

25

22

74

25

78

26

59.8

17

Biphasic 3

Fructose

Lewatit SPC 108/PtBi/C

O2

Base free system

4

5

Fructose

Fructose

PBnNH3Cl/Au/HT

Amberlyst-15/Au8Pd2/HT

1 Equiv of

Two-step

Na2CO3

process

1 Equiv of

Triphasic

Na2CO3

system

O2

O2

Fe3O4@SiO2-SO3H/nano6

Fructose

Two-step t-BuOOH

Base free

Fe3O4-CoOx

process One-pot,

7

Fructose

Amberlyst-15/Fe0.6Zr0.4O2

O2

Base free

This 46

one-step

3

4. CONCLUSIONS

4

In summary, an economical method for the direct conversion of fructose to FDCA in

5

[Bmim]Cl ILs was reported with Amberlyst-15 and Fe-Zr-O as catalysts for the first

6

time. The dehydration of fructose into HMF was efficiently accomplished with the

7

help of Amberlyst-15 and [Bmim]Cl, and subsequently followed by the further

8

oxidation of the formed HMF using Fe-Zr-O in the same reactor without separation of 23

ACS Paragon Plus Environment

work

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

acid catalyst from the reaction system. Under the base-free conditions, a FDCA yield

2

of 46.4% was obtained with the fructose conversion of 100%. The possible reaction

3

mechanism was proposed based on experimental data and previous studies, and a

4

plausible reaction pathway and corresponding kinetics studies were also investigated.

5

More importantly, FDCA can also be produced from other biomass sources, including

6

glucose, galactose, mannose, starch and cellulose. This one-pot conversion method for

7

the synthesis of FDCA from fructose exhibits promising potential in sustainable and

8

industrial production of FDCA since the widely available and cheap renewable

9

resource can be utilized as a starting material. Furthermore, this novel reaction

10

process avoids the complicated separation of intermediates and catalyst, thereby

11

reducing the corresponding costs, energy and solvent.

12



13

Supporting Information

14

The details for catalyst characterization; Figure S1, powder X-ray diffraction pattern and TEM

15

image of catalyst; Figure S2, effect of acid catalysts on the dehydration of fructose to HMF; Table

16

S1, effect of reaction time on the conversion of fructose; Figure S3, the insoluble humins formed

17

after the reaction; Scheme S1, putative mechanism for the one-pot synthesis of FDCA from

18

fructose; Figure S4, HPLC chromatograms for conversion of fructose to FDCA; Table S2, kinetic

19

parameters of fructose conversion in ionic liquids; Table S3, kinetic parameters of HMF

20

conversion in ionic liquids; Table S4, kinetic parameters of HMFA conversion in ionic liquids;

ASSOCIATED CONTENT

24

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

Table S5, kinetic parameters of FFCA conversion in ionic liquids; Table S6, synthesis of FDCA

2

from other biomass source by using Amberlyst-15 and Fe0.6Zr0.4O2 catalysts in ionic liquids. The

3

Supporting Information is available free of charge on the ACS Publications website at DOI:

4

5



6

Corresponding Authors

7

*Tel./fax: +86-10-62558174. E-mail: [email protected] (J. Xin), [email protected] (S. Zhang).

8

Notes

9

The authors declare no competing financial interest.

10



AUTHOR INFORMATION

ACKNOWLEDGEMENTS

11

This work was supported by the National Key Technology R&D Program of

12

China (2016YFB0301900), the National Natural Science Foundation of China (No.

13

21576269, 21210006, 21506231), and the Science & Technology Pillar Program of

14

Sichuan Province (2015GZ0065).

15



REFERENCES

16

(1) Ragauskas, A. J.; Williams, C. K.; Davison, B. H.; Britovsek, G.; Cairney, J.;

17

Eckert, C. A.; Frederick, W. J.; Hallett, J. P.; Leak, D. J.; Liotta, C. L.; Mielenz, J. R.;

18

Murphy, R.; Templer, R.; Tschaplinski, T. The path forward for biofuels and

19

biomaterials. Science 2006, 311 (5760), 484-489.

20

(2) Chheda, J. N.; Huber, G. W.; Dumesic, J. A. Liquid-phase catalytic processing of

21

biomass-derived oxygenated hydrocarbons to fuels and chemicals. Angew. Chem. Int. 25

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

Page 26 of 32

Ed. 2007, 46 (38), 7164-83.

2

(3) Ruppert, A. M.; Weinberg, K.; Palkovits, R. Hydrogenolysis goes bio: from

3

carbohydrates and sugar alcohols to platform chemicals. Angew. Chem. Int. Ed. 2012,

4

51 (11), 2564-2601.

5 6

(4) Zhang, Z. H.; Deng, K. J. Recent advances in the catalytic synthesis of 2,5-furandicarboxylic acid and its derivatives. ACS Catal. 2015, 5 (11), 6529-6544.

7

(5) Gandini, A.; Coelho, D.; Gomes, M.; Reis, B.; Silvestre, A. Materials from

8

renewable resources based on furan monomers and furan chemistry: work in progress.

9

J. Mater. Chem. 2009, 19 (45), 8656-8664.

10

(6) Pan, T.; Deng, J.; Xu, Q.; Zuo, Y.; Guo, Q. X.; Fu, Y. Catalytic conversion of

11

furfural into a 2,5-furandicarboxylic acid-based polyester with total carbon utilization.

12

ChemSusChem 2013, 6 (1), 47-50.

13

(7) Zhang, D. H.; Dumont, M. J. Advances in polymer precursors and bio-based

14

polymers synthesized from 5-hydroxymethylfurfural. J. Polym. Sci., Part A: Polym.

15

Chem. 2017, 55 (9) 1478-1492.

16

(8) Burgess, S. K.; Leisen, J. E.; Kraftschik, B. E.; Mubarak, C. R.; Kriegel, R. M.;

17

Koros, W. J. Chain mobility, thermal, and mechanical properties of poly(ethylene

18

furanoate) compared to poly(ethylene terephthalate). Macromolecules 2014, 47 (4),

19

1383-1391.

20

(9) Gandini, A.; Silvestre, A. J. D.; Neto, C. P.; Sousa, A. F.; Gomes, M. The furan

21

counterpart of poly(ethylene terephthalate): an alternative material based on

22

renewable resources. J. Polym. Sci., Part A: Polym. Chem. 2009, 47 (1), 295-298.

23

(10) Liu, B.; Zhang, Z. H. One-pot conversion of carbohydrates into furan

24

derivatives

via

furfural

and

5-hydroxylmethylfurfural

25

ChemSusChem 2016, 9 (16), 2015-2036.

as

intermediates.

26

(11) Hansen, T. S.; Sádaba, I.; García-Suárez, E. J.; Riisager, A. Cu catalyzed

27

oxidation of 5-hydroxymethylfurfural to 2,5-diformylfuran and 2,5-furandicarboxylic

28

acid under benign reaction conditions. Appl. Catal. A: Gen. 2013, 456, 44-50. 26

ACS Paragon Plus Environment

Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

(12) Casanova, O.; Iborra, S.; Corma, A. Biomass into chemicals: aerobic oxidation

2

of 5-hydroxymethyl-2-furfural into 2,5-furandicarboxylic acid with gold nanoparticle

3

catalysts. ChemsusChem 2009, 2 (12), 1138-1144.

4

(13) Siankevich, S.; Savoglidis, G.; Fei, Z. F.; Laurenczy, G.; Alexander, D. T. L.;

5

Yan, N.; Dyson, P. J. A novel platinum nanocatalyst for the oxidation of

6

5-hydroxymethylfurfural into 2,5-furandicarboxylic acid under mild conditions. J.

7

Catal. 2014, 315, 67-74.

8 9

(14) Zhang, Z. H.; Zhen, J. D.; Liu, B.; Lv, K. L.; Deng, K. J. Selective aerobic oxidation

of

the

biomass-derived

precursor

5-hydroxymethylfurfural

to

10

2,5-furandicarboxylic acid under mild conditions over a magnetic palladium

11

nanocatalyst. Green Chem. 2015, 17 (2), 1308-1317.

12

(15) Gorbanev, Y. Y.; Kegnaes, S.; Riisager, A. Selective aerobic oxidation of

13

5-hydroxymethylfurfural in water over solid ruthenium hydroxide catalysts with

14

magnesium-based supports. Catal. Lett. 2011, 141 (12), 1752-1760.

15

(16) Nishimura, S.; Yakita, Y.; Katayama, M.; Higashimine, K.; Ebitani, K. The role

16

of negatively charged Au states in aerobic oxidation of alcohols over hydrotalcite

17

supported AuPd nanoclusters. Catal. Sci. Technol. 2013, 3 (2), 351-359.

18

(17) Wang, S. G.; Zhang, Z. H.; Liu, B. Catalytic conversion of fructose and

19

5-hydroxymethylfurfural

into

2,5-furandicarboxylic

acid

over

a

recyclable

20

Fe3O4-CoOx magnetite nanocatalyst. ACS Sustainable Chem. Eng. 2015, 3 (3),

21

406-412.

22

(18) Gao, L. C.; Deng, K. J.; Zheng, J. D.; Liu, B.; Zhang, Z. H. Efficient oxidation

23

of biomass derived 5-hydroxymethylfurfural into 2,5-furandicarboxylic acid catalyzed

24

by Merrifield resin supported cobalt porphyrin. Chem. Eng. J. 2015, 270, 444-449.

25

(19) Neațu, F.; Marin, R. S.; Florea, M.; Petrea, N.; Pavel, O. D.; Pârvulescu, V. I.

26

Selective

27

heterogeneous catalysts. Appl. Catal. B: Environ. 2016, 180, 751-757.

28

oxidation

of

5-hydroxymethyl

furfural

over

non-precious

metal

(20) Hayashi, E.; Komanoya, T.; Kamata, K.; Hara, M. Heterogeneously-catalyzed 27

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

aerobic oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid with

2

MnO2. ChemSusChem 2017, 10 (4), 654-658.

3

(21) Han, X. W.; Li, C. Q.; Liu, X. H.; Xia, Q. N.; Wang, Y. Q. Selective oxidation of

4

5-hydroxymethylfurfural to 2,5-furandicarboxylic acid over MnOx-CeO2 composite

5

catalysts. Green Chem. 2017, 19 (4), 996-1004.

6

(22) Kröger, M.; Prüße, U.; Vorlop, K.-D. A new approach for the production of

7

2,5-furandicarboxylic acid by in situ oxidation of 5-hydroxymethylfurfural starting

8

from fructose. Topics in Catalysis 2000, 13 (3), 237-242.

9

(23) Ribeiro, M. L.; Schuchardt, U. Cooperative effect of cobalt acetylacetonate and

10

silica in the catalytic cyclization and oxidation of fructose to 2,5-furandicarboxylic

11

acid. Catal. Commun. 2003, 4 (2), 83-86.

12

(24) Yi, G. S.; Teong, S. P.; Li, X. K.; Zhang, Y. G. Purification of biomass-derived

13

5-hydroxymethylfurfural and its catalytic conversion to 2,5-furandicarboxylic acid.

14

ChemsusChem 2014, 7 (8), 2131-2135.

15

(25) Teong, S. P.; Yi, G. S.; Cao, X. Q.; Zhang, Y. G. Poly-benzylic ammonium

16

chloride resins as solid catalysts for fructose dehydration. ChemSusChem 2014, 7 (8),

17

2120-2126.

18

(26) Yi, G. H.; Teong, S. P.; Zhang, Y. G. The direct conversion of sugars into

19

2,5-furandicarboxylic acid in a triphasic system. ChemsusChem 2015, 8 (7),

20

1151-1155.

21

(27) Yang, Z. Z.; Qi, W.; Su, R. X.; He, Z. M. Selective synthesis of

22

2,5-diformylfuran and 2,5-furandicarboxylic acid from 5-hydroxymethylfurfural and

23

fructose catalyzed by magnetically separable catalysts. Energy Fuels 2017, 31 (1),

24

533-541.

25

(28) Xin, J. Y.; Yan, D. X.; Cao, R. M.; Lu, X. M.; Dong, H. X.; Zhang, S. J.

26

Sub/supercritical carbon dioxide induced phase switching for the reaction and

27

separation in ILs/methanol. Green Energy Environ. 2016, 1 (2), 144-148.

28

(29) Liu, M.; Jia, S Y..; Gong, Y. Y.; Song, C. S.; Guo, X. W. Effective hydrolysis of 28

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

cellulose into glucose over sulfonated sugar-derived carbon in an ionic liquid. Ind.

2

Eng. Chem. Res. 2013, 52 (24), 8167-8173.

3

(30) Zhou, J. X.; Huang, T. Y.; Zhao, Y. L.; Xia, Z.; Xu, Z. W.; Jia, S. Y.; Wang, J. J.;

4

Zhang, Z. C. Solvent mediation for enhanced separation of 5-hydroxymethylfurfural

5

from 1-butyl-3-methylimidazolium chloride. Ind. Eng. Chem. Res. 2015, 54 (32),

6

7977-7983.

7

(31) Yan, D. X.; Xin, J. Y.; Shi, C. Y.; Lu, X. M.; Ni, L. L.; Wang, G. Y.; Zhang, S. J.

8

Base-free conversion of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid in

9

ionic liquids. Chem. Eng. J. 2017, 323, 473-482.

10

(32) Yan, D. X.; Xin, J. Y.; Zhao, Q.; Gao, K.; Lu, X.M.; Wang, G.Y.; Zhang, S.J.

11

Fe-Zr-O catalyzed base-free aerobic oxidation of 5-HMF to 2,5-FDCA as a bio-based

12

polyester monomer. Catal. Sci. Technol. 2017, DOI: 10.1039/C7CY01704A.

13

(33) Qi, X. H.; Watanabe, M.; Aida, T. M.; Smith, J. R. L. Efficient process for

14

conversion of fructose to 5-hydroxymethylfurfural with ionic liquids. Green Chem.

15

2009, 11 (9), 1327-1331.

16

(34) Ramli, N. A. S.; Amin, N. A. S. A new functionalized ionic liquid for efficient

17

glucose conversion to 5-hydroxymethyl furfural and levulinic acid. J. Mol. Catal. A:

18

Chem. 2015, 407, 113-121.

19

(35) Weingarten, R.; Tompsett, G. A.; Conner, W. C.; Huber, G. W. Design of solid

20

acid catalysts for aqueous-phase dehydration of carbohydrates: the role of Lewis and

21

Brønsted acid sites. J. Catal. 2011, 279 (1), 174-182.

22

(36) Wang, S. G.; Zhang, Z. H.; Liu, B.; Li, J. L. Environmentally friendly oxidation

23

of biomass derived 5-hydroxymethylfurfural into 2,5-diformylfuran catalyzed by

24

magnetic separation of ruthenium catalyst. Ind. Eng. Chem. Res. 2014, 53 (14),

25

5820-5827.

26

(37) Roy Goswami, S.; Dumont, M.-J.; Raghavan, V. Microwave assisted synthesis

27

of 5-hydroxymethylfurfural from starch in AlCl3·6H2O/DMSO/[BMIM]Cl system.

28

Ind. Eng. Chem. Res. 2016, 55 (16), 4473-4481. 29

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

(38)

Hu,

L.;

Sun,

Y.;

Lin,

L.

Efficient conversion

Page 30 of 32

of

glucose

into

2

5-hydroxymethylfurfural by chromium(III) chloride in inexpensive ionic liquid. Ind.

3

Eng. Chem. Res. 2012, 51 (3), 1099-1104.

4

(39)

Vuyyuru,

K.

R.;

Strasser,

P.

Oxidation

of

biomass

derived

5

5-hydroxymethylfurfural using heterogeneous and electrochemical catalysis. Catal.

6

Today. 2012, 195 (1), 144-154.

7 8

(40) Akien, G. R.; Qi, L.; Horváth, I. T. Molecular mapping of the acid catalysed dehydration of fructose. Chem. Commun. 2012, 48 (47), 5850-5852.

9

(41) Nikbin, N.; Caratzoulas, S.; Vlachos, D. G. A first principles-based microkinetic

10

model for the conversion of fructose to 5-hydroxymethylfurfural. ChemcatChem 2012,

11

4 (4), 504-511.

12

(42) Zhang, J.; Yu, X. X.; Zou, F. X.; Zhong, Y. H.; Du, N.; Huang, X. R.

13

Room-temperature

14

5-hydroxymethylfurfural in high efficiency. ACS Sustainable Chem. Eng. 2015, 3 (12),

15

3338-3345.

16

ionic

liquid

system

converting

fructose

into

(43) Gupta, N. K.; Nishimura, S.; Takagaki, A.; Ebitani, K. Hydrotalcite-supported

17

gold-nanoparticle-catalyzed

highly

efficient

base-free

aqueous

oxidation

of

18

5-hydroxymethylfurfural into 2,5-furandicarboxylic acid under atmospheric oxygen

19

pressure. Green Chem. 2011, 13 (4), 824-827.

20

(44) Wang, Y. B.; Yu, K.; Lei, D.; Si, W.; Feng, Y. J. Lou, L.-L.; Liu, S. X.

21

Basicity-tuned hydrotalcite-supported Pd catalysts for aerobic oxidation of

22

5-hydroxymethyl-2-furfural under mild conditions. ACS Sustainable Chem. Eng. 2016,

23

4 (9), 4752-4761.

24

(45) van Putten, R.-J.; van der Waal, J. C.; de Jong, E.; Rasrendra, C. B.; Heeres, H.

25

J.; de Vries, J. G. Hydroxymethylfurfural, a versatile platform chemical made from

26

renewable resources. Chem. Rev. 2013, 113 (3), 1499-1597.

27

(46) Zheng, L. F.; Zhao, J. Q.; Du, Z. X.; Zong, B. N.; Liu, H. C. Efficient aerobic

28

oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid on Ru/C catalysts. 30

ACS Paragon Plus Environment

Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

Sci. China Chem. 2017, 60 (7), 950-957.

2

(47) Caes, B. R.; Teixeira, R. E.; Knapp, K. G.; Raines, R. T. Biomass to furanics:

3

renewable routes to chemicals and fuels. ACS Sustainable Chem. Eng. 2015, 3 (11),

4

2591-2605.

5

(48) Rosatella, A. A.; Simeonov, S. P.; Frade, R. F. M.; Afonso, C. A. M.

6

5-Hydroxymethylfurfural (HMF) as a building block platform: biological properties,

7

synthesis and synthetic applications. Green Chem. 2011, 13 (4), 754-793.

8

(49) Choudhary, V.; Mushrif, S. H.; Ho, C.; Anderko, A.; Nikolakis, V.; Marinkovic,

9

N. S.; Frenkel, A. I.; Sandler, S. I.; Vlachos, D. G. Insights into the interplay of lewis

10

and

brønsted

acid

catalysts

in

glucose

and

11

5-(hydroxymethyl)furfural and levulinic acid in aqueous media. J. Am. Chem. Soc.

12

2013, 135 (10), 3997-4006.

31

ACS Paragon Plus Environment

fructose

conversion

to

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2

Table of contents graphic:

3

32

ACS Paragon Plus Environment

Page 32 of 32