Efficient Synthesis of 2,5-Furandicarboxylic Acid from Furfural Based

Aug 20, 2018 - ... to p-phthalic acid, from furfural based platform molecule was explored. ... Acids onto Multifunctional-Based Fe3O4@SiO2 Magnetic Ca...
0 downloads 0 Views 448KB Size
Subscriber access provided by University of South Dakota

Article

Efficient Synthesis of 2,5-Furandicarboxylic Acid from Furfural Based Platform through Aqueous-phase Carbonylation Sicheng Zhang, Guanfei Shen, Yuyan Deng, Yu Lei, Jing-Wen Xue, Zhuqi Chen, and Guochuan Yin ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b02780 • Publication Date (Web): 20 Aug 2018 Downloaded from http://pubs.acs.org on August 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 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 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.

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 25 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

ACS Sustainable Chemistry & Engineering

Efficient Synthesis of 2,5-Furandicarboxylic Acid from Furfural Based Platform through Aqueousphase Carbonylation Sicheng Zhang, Guanfei Shen, Yuyan Deng, Yu Lei, Jing-Wen Xue, Zhuqi Chen and Guochuan Yin* School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, 1037 Luoyu Rd, Wuhan 430074, Hubei Province, PR China; Key laboratory of Material Chemistry for Energy Conversion and Storage (Huazhong University of Science and Technology), Ministry of Education, PR China. ∗

Corresponding author:

E-mail address: [email protected] (G. Yin)

ACS Paragon Plus Environment

1

ACS Sustainable Chemistry & Engineering 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 2 of 25

ABSTRACT

Deriving bulk chemicals from sustainable biomass for chemical industries has become a worldwide consensus due to the rapid depletion of fossil feedstock. Here, synthesis of 2,5furandicarboxylic acid (FDCA), an alternative to p-phthalic acid, from furfural based platform molecule was explored. Using 5-bromo-furoic acid as the starting chemical which has been industrially available through bromination of furoic acid, 98% yield of FDCA can be achieved by catalytic carbonylation in aqueous solution with palladium catalyst, and FDCA separation can be easily performed through simple acidification of reaction media in large scale synthesis. This new route to FDCA has offered furfural a promising sub-chemical with large market. Meanwhile, it also provides an alternative source to FDCA originating from C5 based bulk biomass, that is, hemicellulose, thus offering an opportunity to relieve the current stress in FDCA synthesis from 5-hydroxymethylfurfural originally from cellulose to replace p-phthalic acid in polymer industry.

KEYWORDS: carbonylation, renewable resources, furfural, 2,5-furandicarboxylic acid, palladium.

ACS Paragon Plus Environment

2

Page 3 of 25 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

ACS Sustainable Chemistry & Engineering

INTRODUCTION

As the only bulk, sustainable resource of organic carbon, biomass is a unique and promising candidate to, at least, partly replace the fossil feedstock as the carbon source of chemical industry.1-3 The transformations of versatile carbohydrates, in particular cellulose and hemicellulose which represent up to 70% of global biomass resources, to versatile chemical platform molecules have attracted increasing attentions in recent years.5-7 Among recommended platform molecules,8-10 furfural is one of the most attractive, because it has been industrially produced for several decades, and its bulk raw materials, that is, C5 based carbohydrates from hemicellulose in corncobs, oat, wheat bran, and sawdust, etc., do not compete with the food of human supply.11 However, the challenge is that the markets of its sub-chemicals like furfuryl alcohol, furoic acid and furan are very limited, which seriously blocks its utilizations as the carbon source of chemical industry. As a result, its industrial production is approximately 0.25 M tons/y.12 In view of its promising potentiality as a sustainable platform molecule with its ongoing industrial process, transformations of furfural to fuels and chemicals have attracted much attention than ever,13-15 including our and other endeavors to synthesis maleic acid and anhydride from furfural by catalytic oxidation.16-24 Like furfural originating from hemicellulose, 5-hydroxymethylfurfural (HMF) can be also potentially, but not yet, produced from cellulose in large scale, although Avantium even announced a pilot plant for FDCA production with the capacity of 50,000 tons/y12. Since it contains two functional substituents on the furan frame, its sub-chemicals are more flexible than those of furfural, thus demonstrating more applications in industry. For example, 2,5furandicarboxylic acid (FDCA) has been recognized as an alternative to p-phthalic acid in polymer industry.25 However, even though great endeavours have been paid on exploring new

ACS Paragon Plus Environment

3

ACS Sustainable Chemistry & Engineering 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 4 of 25

catalysts for HMF synthesis,26 the technical bottlenecks in transforming bulk cellulose to HMF in large scale are still existing, which apparently block utilizations of its sub-chemicals like FDCA as the bulk chemicals in industry. In exploring alternative route for FDCA synthesis, Fu and co-workers even reported the synthesis of FDCA through disproportionation of furoic acid with the formation of furan, while Fischer and Kanan independently explored the C-H carboxylations of furoic acid with CO2 to FDCA under strong alkaline conditions or with carbonate promoter.27-30 We recently developed a four-step synthesis of FDCA from furoic acid through bromination, esterification, carbonylation and hydrolysis, and an one-step synthesis of another HMF derivative through oxidative carbonylation of furfuryl acetate.31,32 Although these new processes are still lengthy or low efficiency, the new developments have convinced that transforming furfural based platforms to HMF derivatives is accessible. Here, we present an efficient one-step synthesis of FDCA from 5bromo-furoic acid through aqueous-phase carbonylation. Regarding 5-bromo-furoic acid can be feasibly synthesized from furfural based derivatives,33,34 and its production has been successfully industrialized through bromination of furioc acid with bromine, this novel catalytic route has offered an alternative solution to FDCA synthesis originally from bulk hemicellulose with high efficiency. EXPERIMENTAL SECTION Materials. Unless otherwise noted, all of the reagents are analytic purity grade and were used without further purification. 5-Bromo-furoic acid was purchased from Shandong Youbang Biochemical Technology Co., Ltd. Pd(OAc)2 was purchased from Strem Chemicals Inc. Pd(TFA)2 was from Energy Chemical. Pd(CH3CN)2Cl2 came from Shanghai Boka-chem Tech Inc. Pd(dba)2 was from Nanjing JinruiJiuAn Biotechnology Co., Ltd. PdCl2 and different bases

ACS Paragon Plus Environment

4

Page 5 of 25 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

ACS Sustainable Chemistry & Engineering

were all purchased from Sinopharm Chemical Reagent Co., Ltd. 1H and 13C NMR spectra were recorded on a Bruker AV-400 using TMS as an internal reference. Analytical methods After a typical catalytic carbonylation reaction, the product mixture was diluted with a known mass of the mobile phase, and then filtered and analyzed by HPLC. The HPLC instrument was equipped with a UV detector and a C18 column (250 mm × 4.6 mm), the ratio of mobile phase was water : MeOH : CH3CN = (80% : 10% : 10%, v/v/v) containing concentrated H2SO4 (0.1%) and the flowing rate was fixed at 1 mL min−1. The temperature of the column was 303 K. The wavelength for analysis of FDCA, FA and 5-bromo-furoic acid was 247 nm. In the gram scale synthesis, after the reaction, FDCA was directly isolated from the reaction medium through acidification with HCl, and dried under air for weight and NMR analysis. Synthesis of FDCA through catalytic carbonylation of 5-bromo-furoic acid with CO balloon. Pd(dba)2 (5.3 mg, 9.2 µmol) and TPPTS (10.5 mg, 18.4 µmol) were successively added in water (2 mL) in a glass tube, and the mixtures were next stirred for 30 min under Ar atmosphere. Then, NaHCO3 (60.5 mg, 0.72 mmol) and 5-bromo-furoic acid (34.4 mg, 0.18 mmol) were added in, and the reaction tube was evacuated and back-filled with CO (five times, balloon), then heated up to 90 oC for 8 h under stirring with CO balloon. After the reaction, the reaction mixture was cooled to room temperature and vented to discharge the excess CO carefully in the fume hood. Yield of FDCA and conversion of 5-bromo-furoic acid were quantitatively analyzed by HPLC. Synthesis of FDCA through catalytic carbonylation of 5-bromo-furoic acid under pressured CO. Pd(dba)2 (5.3 mg, 9.2 µmol) and TPPTS (10.5 mg, 18.4 µmol) were successively

ACS Paragon Plus Environment

5

ACS Sustainable Chemistry & Engineering 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 25

added in water (2 mL) in a glass tube, and the mixtures were stirred for 30 min under Ar atmosphere. NaHCO3 (60.5 mg, 0.72 mmol) and 5-bromo-furoic acid (34.4 mg, 0.18 mmol) were then added in. After the reaction tube was sealed inside the stainless steel reactor (50 mL), the system was purged with CO, pressurizing to 10 atm and venting for four times to clean up the air residues. Then, the autoclave was pressurized to 2.5 atm with CO and stirred in an oil bath to keep the reaction temperature at 90 °C for 8 h. After the reaction, the reactor was allowed to cool down to room temperature and carefully depressurized to normal pressure in the fume hood. Yield of FDCA and conversion of 5-bromo-furoic acid were quantitatively analyzed by HPLC. A gram scale synthesis of FDCA by aqueous carbonylation. Pd(dba)2 (80.5 mg, 0.14 mmol) and TPPTS (160.0 mg, 0.28 mmol) were successively added in water (45 mL) in a glass reactor, and the mixture was stirred for 30 min under Ar atmosphere. After that, the generated suspension was filtrated, and NaHCO3 (1.62 g, 18.9 mmol) and 5-bromo-furoic acid (1.03 g, 5.4 mmol) were then added to the resulting pale green solution. The pH value of this aqueous solution was determined to be 7.4. Next, the carbonylation reaction was conducted following the procedure in a 100 mL stainless steel reactor under the pressurized CO (5 atm) as described above for 14 h. After the reaction, the reactor was allowed to cool down to room temperature and carefully depressurized to normal pressure in the fume hood. After filtration of trace palladium black, the reaction solution was acidified by HCl, which leads to the immediate formation of a white precipitate. After filtration, the precipitates were washed with water, and dried in vacuum, giving 93% isolated yield of FDCA (0.78 g, 5.0 mmol). Anal. for FDCA, found: C, 46.40%;

H, 2.75%; calculated: C 46.17%; H, 2.58%. The HPLC and 31P NMR analysis disclosed that there is no TPPTS ligand detectable in the FDCA product, and ICP analysis of Pd displayed that only very limited Pd (11 ppm) existing in the FDCA product.

ACS Paragon Plus Environment

6

Page 7 of 25 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

ACS Sustainable Chemistry & Engineering

RESULTS AND DISCUSSION

Table 1. The palladium soruce screening for carbonylation of 5-bromo-furoic acid.a

Br

O 1

COOH

5 mol% [Pd] 10 mol% TPPTS CO, NaHCO3, H2O, 90 oC, 8 h

HOOC

O

COOH

+

H

COOH

O

2

3

Yield (%)b Entry

a

[Pd] catalyst

Conv. (%) 2

3

1

Pd(OAc)2

>99

87

7

2

Pd(CH3CN)2Cl2

>99

84

5

3

PdCl2

>99

86

6

4

Pd(TFA)2

>99

84

6

5

Pd(dba)2

>99

93

2

6c

Pd(dba)2

8

6

10 (Table 2, entries 1-7), and the formation of FA could be sharply prohibited when the pH value was below 8. Generally, the FDCA yield increased, meanwhile the FA yield decreased with the pH decline following the order of Cs2CO3 < K2CO3 < Na2CO3 < KHCO3 < NaHCO3. Table 2. The influence of different bases on carbonylation of 5-bromo-furoic acid with Pd(dba)2/TPPTS catalyst. a

Entry

a

Base

pH

Yield (%)b

pH (with substrate)

Conv. (%) 2

3

1

KOH

13.2

13.1

>99

43

44

2

NaOH

13.2

13.1

>99

69

25

3

DBU

13.1

13.0

64

22

31

4

K3PO4

12.5

12.0

>99

62

27

5

Cs2CO3

11.6

10.5

>99

49

48

6

K2CO3

11.6

10.3

>99

68

24

7

Na2CO3

11.5

10.2

>99

74

23

8

KHCO3

8.4

7.6

>99

90

6

9

NaHCO3

8.1

7.3

>99

93

2

10

Na2HPO4

9.0

7.0

>99

91

4

11

AcONa

7.9

5.1

>99

98

99

91

2

13

AcONH4

6.8

5.1

>99

83

2

Conditions: Pd(dba)2 (5 mol%), TPPTS (10 mol%), 1 (0.18 mmol), base (0.72 mmol), CO (10

atm), H2O (2 mL), 90 oC, 8 h. DBU = 1,8-Diazabicyclo[5.4.0]undec-7-ene. b HPLC yields.

ACS Paragon Plus Environment

10

Page 11 of 25 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

ACS Sustainable Chemistry & Engineering

Inspired by these results, other weak bases like Na2HPO4, NaOAc, KOAc and AcONH4 was subsequently tested (entries 10-13), and adding NaOAc exhibited the highest carbonylation efficiency, giving 98% yield of FDCA with only a trace of FA formation under current conditions. Interestingly, adding three types of acetate salts, which generated the identical pH environment (pH value of 5.1) for carbonylation, demonstrated the different carbonylation efficiency, giving the FDCA yield following the order of NaOAc> KOAc>AcONH4. A plausible explanation of this phenomenon could be that these univalent cations demonstrated the different influence on the coagulation of Pd(0) in the order of NH4+>K+>Na+ which obeys the Hofmeister sequence,38 which led to the different carbonylation efficiency, even that the added TPPTS ligand is able to stabilize Pd(0) in the aqueous medium as reported.39,40 Next, the influence of CO pressure on carbonylation was investigated using NaHCO3 and NaOAc as base, respectively (Figure 2). In the case of NaOAc, the FDCA yield increased with the increase of CO pressure, and a 98% yield was achieved under 10 atm of CO. Remarkably, using NaHCO3 as the base, the same yield (98%) of FDCA was achieved under only 2.5 atm of CO, while a slight decrease of FDCA yield was observed with further increase of CO pressure. Accordingly, NaHCO3 was employed as the base in following studies because of its relatively low CO pressure requirement for efficient carbonylation. The influence of increased CO pressure can be attributed to the competitive coordination of CO on the palladium center, which slightly disturbed the carbonylation proceeding. Similar retarding effects of increased CO pressure were also observed in other Pd-catalyzed carbonylation reactions.41,42

ACS Paragon Plus Environment

11

ACS Sustainable Chemistry & Engineering

100

NaOAc NaHCO3

80

Yield %

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 25

60

40 0.0

2.5

5.0

7.5

10.0

CO pressure / atm

Figure 2. The influence of the CO pressure on the carbonylation reaction. Conditions: Pd(dba)2 (5 mol%), TPPTS (10 mol%), 1 (0.18 mmol), NaHCO3 (0.72 mmol), CO, H2O (2 mL), 90 oC, 8 h. Next, the influence of NaHCO3 amount along with the pH value of solution on carbonylation was investigated, and the results are summarized in Table 3. In the absence of NaHCO3, 5bromo-furoic acid was not completely soluble in water at room temperature, and the pH value of the reaction medium was around 2.5. Due to the low solubility of 5-bromo-furoic acid, its conversion was quite low (33%) with only 12% yield of FDCA formation. The FDCA yield could be improved up to 49% with 84% conversion of 5-bromo-furoic acid by adding 1 equiv. of NaHCO3. In this case, the pH value of reaction medium was 5.1. Further increasing the NaHCO3 loading from 1 to 4 equiv. made the yield and selectivity of FDCA continuously increasing until achieving 98% yield with 98% selectivity, and the pH value also gradually increased from 5.1 to 7.3. However, increasing the NaHCO3 loading to 5 equiv. made the pH value of reaction medium rising to pH 7.5, which led to the increased dehalogenation product (FA) formation with the

ACS Paragon Plus Environment

12

Page 13 of 25 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

ACS Sustainable Chemistry & Engineering

decline of FDCA selectivity, similar to that of adding 4 equiv. of KHCO3 having pH value of 7.6 (Table 2, entry 8). Further over-loading of NaHCO3 to 6 equiv. resulted in the conversion decreased as well, which was plausibly attributed to that the high electrolyte concentration of the reaction medium may have slightly disturbed the catalysis for carbonylation. However, it is worth mentioning that using 4 equiv. of NaOAc as base, which generated a pH 5.1 value of reaction medium, also provided 98% yield of FDCA (Table 2, entry 11). Apparently, not only the pH value of reaction medium, but the nature of added base may also affect the carbonylation selectivity. Table 3. The influence of the NaHCO3 amount on the carbonylation reaction.a

NaHCO3 (equiv.)

pH (with substrate)

Conv. (%)

pH

1

0

6.1

2.5

2

1

7.7

3

2

4

2

3

Selectivity of 2 (%)

33

12

5

36

5.1

84

49

3

58

7.8

7.0

90

70

3

78

3

8.0

7.2

95

92

2

97

5

4

8.1

7.3

>99

98

1

98

6

5

8.3

7.5

>99

89

4

89

7

6

8.4

7.5

92

78

9

85

Entry

a

Yield (%)b

Conditions: Pd(dba)2 (5 mol%), TPPTS (10 mol%), 5-bromo-furoic acid (0.18 mmol), NaHCO3,

CO (2.5 atm), H2O (2 mL), 90 oC, 8 h. b HPLC yields. In optimizing the Pd/TPPTS ratio, it was found that the ratio of 1:2 offered the best carbonylation efficiency, affording 98% yield of FDCA (Figure 3). However, the ratio of 1:1 also provided a very nice carbonylation results, giving 82% yield of FDCA with complete conversion of 5-bromo-furoic acid, implicating that the active palladium catalyst for

ACS Paragon Plus Environment

13

ACS Sustainable Chemistry & Engineering

carbonylation is either a monomeric or multiple phosphine ligated Pd(0) species as well as those in the literature.43,44 The presence of extra TPPTS ligand, for example, ratio of 1:2, may benefit preventing the formation of palladium black, thus improve its carbonylation efficiency. Conversion / % Yield of FDCA / % Yield of FA / %

100

Conv. % / Yield %

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 14 of 25

80 60 40 20 0 1

2

3

The ratio of TPPTS/Pd(0)

Figure 3. The influence of the TPPTS/Pd(0) ratio on the carbonylation reaction. Conditions: Pd(dba)2 (5 mol%), 5-bromo-furoic acid (0.18 mmol), NaHCO3 (0.72 mmol), CO (10 atm), H2O (2 mL), 90 oC, 8 h. The time course of reaction under 2.5 atm of CO pressure further highlighted the high efficiency of Pd/TPPTS catalyzed carbonylation of 5-bromo-furoic acid in aqueous solution (Figure 4). At the first half hour, it has achieved 61% yield of FDCA with 65% conversion. With the reaction proceeding, it could achieve complete carbonylation in 8 h, and further extending the carbonylation time did not cause the yield of FDCA decreasing, supporting that FA as a byproduct was generated from dehalogenation of 5-bromo-furoic acid rather than monodecarboxylation of FDCA. Remarkably, the selectivity of FDCA consistently remained at the high level in the whole process.

ACS Paragon Plus Environment

14

Page 15 of 25

100

Conv. % / Yield %

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

ACS Sustainable Chemistry & Engineering

80

Conv. % Yield %

60 40 20 0 0

2

4

6

8

10

12

Time / h

Figure 4. The time-course of carbonylation under 2.5 atm of CO. Conditions: Pd(dba)2 (5 mol%), TPPTS (10 mol%), 1 (0.18 mmol), NaHCO3 (0.72 mmol), CO (2.5 atm), H2O (2 mL), 90 oC. Inspired by the high activity of Pd/TPPTS catalyst in this carbonylation of 5-bromo-furoic acid, a gram-scale synthesis was next conducted in aqueous solution. For this gram scale test, the free ligand dba as a suspended solid was firstly recycled through filtration before the carbonylation, and the initial pH of the reaction medium was determined to be 7.4. The carbonylation was firstly carried out under 2.5 atm of CO, which gave a relatively lower yield (72%) of FDCA. However, slightly elevating the CO pressure to 5 atm made the carbonylation proceeding smoothly. After the reaction, a white solid product of FDCA was feasibly precipitated through simple acidification of the reaction medium with HCl, affording FDCA in 93% isolated yield, thus highlighting the merit of this aqueous carbonylation in FDCA separation in large scale synthesis (Scheme 1).

ACS Paragon Plus Environment

15

ACS Sustainable Chemistry & Engineering 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 25

Scheme 1. A gram scale synthesis of FDCA by aqueous carbonylation. Conditions: Pd(dba)2 (2.5 mol%), TPPTS (5 mol%), 1 (5.4 mmol), NaHCO3 (18.9 mmol), CO (5 atm), H2O (45 mL), 90 o

C, 14 h. In the literature, several mechanisms have been proposed for carbonylation of C-X bonds in

different systems. 45-48 Based on our results of FA formation as a dehalogenation byproduct, we proposed two competitive pathways for this Pd/TPPTS catalysed 5-bromo-furoic acid carbonylation as shown in Scheme 2. At a relatively low pH conditions, that is, pH 8, otherwise, it was, at least, competitive with path A at pH > 8. In the case of pH > 8, the nucleophilic attack of hydroxide on the intermediate II, rather than the CO coordination, may happen immediately with the release of bromide, which generated the intermediate PdII-OH (V). Although this kind of ligand exchange prior to CO coordination was not so popularly realized in Pd-catalysed carbonylation, it was also reported in the presence of strong bases such as NaOH or KOH.49-51 In the next step, CO insertion of the PdII-OH bond, rather than Pd(II)-C bond in the intermediate V happened to generate the intermediate VI which is different from that of the intermediate IV. The intermediate VI proceeded two competitive pathways in the next step: one to generate FDCA as the expected final product through reductive elimination, the other releasing CO2 to generate the

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering 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 25

intermediate VII having a PdII-hydride moiety. The reductive elimination of this intermediate gave FA as the dehalogenation product, which was observed in our system. In the literature, both CO insertion of the PdII-OH moiety to generate the PdII-CO2H moiety and next CO2 release from the PdII-CO2H moiety to generate the PdII-hydride have been reported; 52,53 thereof, similar CO insertion and CO2 release can also happen here. Although path A and B may competitively exist in this reaction, as the carbonylation efficiency shown in Table 2, path A is dominant at the pH values of reaction medium below 8, while path B makes significant, if not dominant, contribution at pH above 8. CONCLUSION

In summary, a new route to synthesize FDCA has been explored from furfural based platform molecule. Through catalytic carbonylation of 5-bromo-furoic acid in aqueous solution, above 98% HPLC yield, or 93% isolated yield of FDCA can be achieved under low CO pressure. Since the carbonylation was carried out in aqueous solution, separation of FDCA product can be performed by simple acidification of reaction medium in large scale synthesis, highlighting the merit of this process in its potential applications. In view of the limited market of sustainable furfural resources as well as the current technical bottlenecks in FDCA production starting from cellulose in large scale, this novel process has provided an opportunity to transform C5 based furfural platform to C6 derivatives which offered an alternative source to FDCA originally from hemicellulose. ASSOCIATED CONTENT

Supporting Information. This material is available free of charge on the ACS Publications website at DOI: Carbonylation condition optimizations, 1H NMR,

13

C NMR and HPLC characterizations of the

products.

ACS Paragon Plus Environment

18

Page 19 of 25 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

ACS Sustainable Chemistry & Engineering

ACKNOWLEDGMENTS

This work was financially supported by the National Natural Science Foundation of China (No. 21573082 and 21872059). The product identifications by NMR were performed in Analytical and Testing Center, Huazhong University of Science and Technology. REFERENCES

(1) Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41, DOI 10.1039/C1CS15147A. (2) Sheldon, R. A. Green and sustainable manufacture of chemicals from biomass: state of the art. Green Chem. 2014, 16, DOI 10.1039/C3GC41935E. (3) Besson, M.; Gallezot, P.; Pinel, C. Conversion of biomass into chemicals over metal catalysts. Chem. Rev. 2014, 114, DOI 10.1021/cr4002269. (4) Xia, Q.; Chen, Z.; Yi, S.; Gong, X.; Wang, H.; Liu, X.; Parker, S. F.; Han, X.; Yang, S.; Wang, Y. Direct hydrodeoxygenation of raw woody biomass into liquid alkanes. Nat. Commun. 2016, 7, DOI 10.1038/ncomms11162. (5) Zhao, S.; Cheng, M.; Li, J.; Tian, J.; Wang, X. One pot production of 5hydroxymethylfurfural with high yield from cellulose by a Brønsted–Lewis–surfactant-combined heteropolyacid catalyst. Chem. Commun. 2011, 47, DOI 10.1039/C0CC04444J. (6) Guerbuez, E. I.; Wettstein, S. G.; Dumesic, J. A. Conversion of hemicellulose to furfural and levulinic acid using biphasic reactors with alkylphenol solvents. ChemSusChem 2012, 5, DOI 10.1002/cssc.201100608. (7) Yang, L.; Su, J.; Carl, S.; Lynam, J. G.; Yang, X.; Lin, H. Catalytic conversion of hemicellulosic biomass to lactic acid in pH neutral aqueous phase media. Appl. Catal. B-Environ. 2015, 162, DOI 10.1016/j.apcatb.2014.06.025. (8) Werpy, T.; Petersen, G. Top value added chemicals from biomass, U.S. Department of Energy, Oak Ridge, TN, USA, 2004, DOI 10.2172/15008859. (9) Choi, S.; Song, C. W.; Shin, J. H.; Lee, S. Y. Biorefineries for the production of top building block chemicals and their derivatives. Metab. Eng. 2015, 28, DOI 10.1016/j.ymben.2014.12.007.

ACS Paragon Plus Environment

19

ACS Sustainable Chemistry & Engineering 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 20 of 25

(10) Bozell, J. J.; Petersen, G. R. Technology development for the production of biobased products from biorefinery carbohydrates-the US Department of Energy's “Top 10” revisited. Green Chem. 2010, 12, DOI 10.1039/B922014C. (11) Zeitsch, K. J. The chemistry and technology of furfural and its many by-products, Elsevier: London, 2000. (12) Delidovich, I.; Hausoul, P. J.; Deng, L.; Pfützenreuter, R.; Rose, M.; Palkovits, R. Alternative monomers based on lignocellulose and their use for polymer production. Chem. Rev. 2016, 116, DOI 10.1021/acs.chemrev.5b00354. (13) Bohre, A.; Dutta, S.; Saha, B.; Abu-Omar, M. M. Upgrading furfurals to drop-in biofuels: An overview. ACS Sustain. Chem. Eng. 2015, 3, DOI 10.1021/acssuschemeng.5b00271. (14) Li, X.; Jia, P.; Wang, T. Furfural: A promising platform compound for sustainable production of C4 and C5 chemicals. ACS Catal. 2016, 6, DOI 10.1021/acscatal.6b01838. (15) Ouyang, W.; Yepez, A.; Romero, A. A.; Luque, R. Towards industrial furfural conversion: Selectivity and stability of palladium and platinum catalysts under continuous flow regime. Catal. Today 2018, 308, DOI 10.1016/j.cattod.2017.07.011. (16) Shi, S.; Guo, H.; Yin, G. Synthesis of maleic acid from renewable resources: catalytic oxidation of furfural in liquid media with dioxygen. Catal. Commun. 2011, 12, DOI 10.1016/j.catcom.2010.12.033. (17) Guo, H.; Yin, G. Catalytic aerobic oxidation of renewable furfural with phosphomolybdic acid catalyst: an alternative route to maleic acid. J. Phys. Chem. C, 2011, 115, DOI 10.1021/jp2054712. (18) Lan, J.; Chen, Z.; Lin, J.; Yin, G. Catalytic aerobic oxidation of renewable furfural to maleic anhydride and furanone derivatives with their mechanistic studies. Green Chem. 2014, 16, DOI 10.1039/C4GC00829D. (19) Li, X.; Ho, B.; Lim, D. S. W.; Zhang, Y. Highly efficient formic acid-mediated oxidation of renewable furfural to maleic acid with H2O2. Green Chem. 2017, 19, DOI 10.1039/C6GC03020C. (20) Araji, N.; Madjinza, D. D.; Chatel, G.; Moores, A.; Jerome, F.; De Oliveira Vigier, K. Synthesis of maleic and fumaric acids from furfural in the presence of betaine hydrochloride and hydrogen peroxide. Green Chem. 2017, 19, DOI 10.1039/C6GC02620F. (21) Alba-Rubio, A. C.; Fierro, J. L. G.; Leon-Reina, L.; Mariscal, R.; Dumesic, J. A.; Lopez Granados, M. Oxidation of furfural in aqueous H2O2 catalysed by titanium silicalite: deactivation

ACS Paragon Plus Environment

20

Page 21 of 25 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

ACS Sustainable Chemistry & Engineering

processes and role of extraframework Ti oxides. Appl. Catal. B-Environ 2017, 202, DOI 10.1016/j.apcatb.2016.09.025. (22) Alonso-Fagundez, N.; Agirrezabal-Telleria, I.; Arias, P. L.; Fierro, J. L. G.; Mariscal, R.; Granados, M. L. Aqueous-phase catalytic oxidation of furfural with H2O2: high yield of maleic acid by using titanium silicalite-1. RSC Adv. 2014, 4, DOI 10.1039/C4RA11563E. (23) Du, Z.; Ma, J.; Wang, F.; Liu, J.; Xu, J. Oxidation of 5-hydroxymethylfurfural to maleic anhydride with molecular oxygen. Green Chem. 2011, 13, DOI 10.1039/C0GC00837K. (24) Alonso-Fagúndez, N.; Ojeda, M.; Mariscal, R.; Fierro, J. L.G.; López Granados, M. Gas phase oxidation of furfural to maleic anhydride on V2O5/γ-Al2O3 catalysts: Reaction conditions to slow down the deactivation. J. Catal. 2017, 348, DOI 10.1016/j.jcat.2016.12.005. (25) Deng, J.; Liu, X.; Li, C.; Jiang Y.; Zhu, J. Synthesis and properties of a bio-based epoxy resin from 2,5-furandicarboxylic acid (FDCA), RSC Adv., 2015, 5, DOI 10.1039/C5RA00242G. (26) Iris, K. M.; Tsang, D. C. W. Conversion of biomass to hydroxymethylfurfural: a review of catalytic systems and underlying mechanisms. Bioresour. Technol. 2017, 238, DOI 10.1016/j.biortech.2017.04.026. (27) Pan, T.; Deng, J.; Xu, Q.; Zuo, Y.; Guo, Q.; Fu, Y. Catalytic conversion of furfural into a 2, 5-furandicarboxylic acid-based polyester with total carbon utilization. ChemSusChem 2013, 6, DOI 10.1002/cssc.201200652. (28) Fischer, R.; Fišerová, M., One-step synthesis of furan-2, 5-dicarboxylic acid from furan-2carboxylic

acid

using

carbon

dioxide.

ARK.

J.

Org.

Chem.

2013,

4,

DOI

10.3998/ark.5550190.p008.285. (29) Dick, G. R.; Frankhouser, A. D.; Banerjee, A.; Kanan, M. W. A scalable carboxylation route to furan-2, 5-dicarboxylic acid. Green Chem. 2017, 19, DOI 10.1039/C7GC01059A. (30) Banerjee, A.; Dick, G. R.; Yoshino, T.; Kanan, M. W. Carbon dioxide utilization via carbonate-promoted C–H carboxylation. Nature, 2016, 531, DOI 10.1038/nature17185. (31) Zhang, S.; Lan, J.; Chen, Z.; Yin, G.; Li, G. Catalytic synthesis of 2,5-furandicarboxylic acid from furoic acid: transformation from C5 platform to C6 derivatives in biomass utilizations. ACS Sustain. Chem. Eng. 2017, 5, DOI 10.1021/acssuschemeng.7b02396. (32) Zhang, S.; Shen, G.; Chen, Z.; Yin, G. Accessing the HMF derivatives from furfural acetate through oxidative carbonylation. ChemistrySelect 2017, 2, DOI 10.1002/slct.201701263.

ACS Paragon Plus Environment

21

ACS Sustainable Chemistry & Engineering 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 22 of 25

(33) Chaubey, G. S.; Kharsyntiew, B.; Mahanti, M. K. Oxidation of substituted 2-furaldehydes by quinolinium dichromate: a kinetic study. J. Phys. Org. Chem. 2004, 17, DOI 10.1002/poc.698. (34) Fu, L.; Fu, W.; Cheng, P.; Xie, Z.; Fan, C.; Shi, M.; Ling, J.; Hou, J.; Zhan, X.; Chen, H. A diketopyrrolopyrrole molecule end-capped with a furan-2-carboxylate moiety: the planarity of molecular geometry and photovoltaic properties. J. Mater. Chem. A, 2014, 2, DOI 10.1039/C3TA13534A. (35) Xin, Z.; Gøgsig, T. M.; Lindhardt, A. T.; Skrydstrup, T. An efficient method for the preparation of tertiary esters by Palladium-catalyzed alkoxycarbonylation of aryl bromides. Org. Lett. 2012, 14, DOI 10.1021/ol203057w. (36) Min, B.; Kim, D.; Park, H.; Jun, C. Pd/C-catalyzed carbonylative esterification of aryl halides with alcohols by using oxiranes as CO sources. Chem.-Eur. J. 2016, 22, DOI 10.1002/chem.201600570. (37) Carpentier, J.; Castanet, Y.; Brocard, J.; Mortreux, A.; Petit, F. A novel and convenient method for palladium-catalysed alkoxycarbonylation of aryl and vinyl halides using HCO2R/NaOR system. Tetrahedron Lett. 1991, 32, DOI 10.1016/S0040-4039(00)92287-1. (38) Parsons, D. F.; Bostrom, M.; Nostro, P. L.; Ninham, B. W. Hofmeister effects: interplay of hydration, nonelectrostatic potentials, and ion size. Phys. Chem. Chem. Phys. 2011, 13, DOI 10.1039/C1CP20538B. (39) Kohlpaintner, C. W.; Beller, M. Palladium-catalyzed carbonylation of benzyl chlorides to phenylacetic acids—a new two-phase process. J. Mol. Catal. A-Chem. 1997, 116, DOI 10.1016/S1381-1169(96)00119-7. (40) Bertoux, F.; Monflier, E.; Castanet, Y.; Mortreux, A. Palladium catalyzed hydroxycarbonylation of olefins in biphasic system: beneficial effect of alkali metal salt and protective-colloid agents on the stability of the catalytic system. J. Mol. Catal. A-Chem. 1999, 143, DOI 10.1016/S1381-1169(98)00359-8. (41) Klaus, S.; Neumann, H.; Zapf, A.; Strübing, D.; Hübner, S.; Almena, J.; Riermeier, T.; Groß, P.; Sarich, M.; Krahnert, W.; Rossen, K.; Beller, M. A general and efficient method for the formylation of aryl and heteroaryl bromides. Angew. Chem. Int. Ed., 2006, 45, DOI 10.1002/anie.200502697.

ACS Paragon Plus Environment

22

Page 23 of 25 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

ACS Sustainable Chemistry & Engineering

(42) Ito, T.; Mori, K.; Mizoroki, T.; Ozaki, A. Effect of base on palladium-black catalyzed carbonylation of iodobenzene. B. Chem. Soc. Jpn. 1975, 48, DOI 10.1246/bcsj.48.2091. (43) Zheng, Q.; Liu, Y.; Chen, Q.; Hu, M. H.; Helmy, R.; Sherer, E. C.; Welch, C. J.; Chen, H. Capture of reactive monophosphine-ligated palladium(0) intermediates by mass spectrometry. J. Am. Chem. Soc. 2015, 137, DOI 10.1021/jacs.5b08905. (44) Liegault, B.; Renaud, J.; Bruneau, C. Activation and functionalization of benzylic derivatives by palladium catalysts. Chem. Soc. Rev. 2008, 37, DOI 10.1039/B704255H. (45) Bryndza, H. E. Mechanism of carbon monoxide insertion into metal-oxygen bonds. Reaction of (DPPE)Pt(OCH3) R (R= CH3 or OCH3) with CO. Organometallics 1985, 4, DOI 10.1021/om00128a037. (46) Komiya, S.; Akai, Y.; Tanaka, K.; Yamamoto, T.; Yamamoto, A. Reductive elimination of aryl carboxylates from acyl(aryloxy) nickel (II) and-palladium(II) complexes. Organometallics 1985, 4, DOI 10.1021/om00125a033. (47) Garrou, P. E.; Heck, R. F. The mechanism of carbonylation of halo (bis ligand) organoplatinum(II), -palladium(II), and -nickel(II) complexes. J. Am. Chem. Soc. 1976, 98, DOI 10.1021/ja00430a018. (48) Moser, W. R.; Wang, A. W.; Kildahl, N. K. Mechanistic studies of the palladium-catalyzed reaction of methanol with bromobenzene and carbon monoxide to produce methyl benzoate. 1. Stoichiometric study. J. Am. Chem. Soc. 1988, 110, DOI 10.1021/ja00217a020. (49) Grushin, V. V.; Alper, H. Alkali-induced disproportionation of palladium(II) tertiary phosphine complexes, [L2PdCl2], to LO and palladium(0). Key intermediates in the biphasic carbonylation of ArX catalyzed by [L2PdCl2]. Organometallics, 1993, 12, DOI 10.1021/om00029a052. (50) Ando, K.; Nakazawa, J.; Hikichi, S. Synthesis, characterization and aerobic alcohol oxidation catalysis of palladium (II) complexes with a bis(imidazolyl)borate ligand. Eur. J. Inorg. Chem. 2016, DOI 10.1002/ejic.201600206. (51) Bryndza, H. E.; Tam, W. Monomeric metal hydroxides, alkoxides, and amides of the late transition metals: synthesis, reactions, and thermochemistry. Chem. Rev. 1988, 88, DOI 10.1021/cr00089a009.

ACS Paragon Plus Environment

23

ACS Sustainable Chemistry & Engineering 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 24 of 25

(52) Cámpora, J.; Palma, P.; del Río, D.; Álvarez, E. CO insertion reactions into the M-OH bonds of monomeric nickel and palladium hydroxides. Reversible decarbonylation of a hydroxycarbonyl palladium complex. Organometallics 2004, 23, DOI 10.1021/om040002i. (53) Johansson, R.; Wendt, O. F. Synthesis and reactivity of (PCP) palladium hydroxy carbonyl and related complexes toward CO2 and phenylacetylene. Organometallics 2007, 26, DOI 10.1021/om0611709.

ACS Paragon Plus Environment

24

Page 25 of 25 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

ACS Sustainable Chemistry & Engineering

For Table of Contents Use Only

Synopsis: An alternative route to synthesize FDCA was developed by using 5-bromofuroic acid, a C5 based furfural derivative through aqueous carbonylation.

ACS Paragon Plus Environment

25