Production of 4-Hydroxymethylfurfural from ... - ACS Publications

Feb 1, 2016 - Production of 4‑Hydroxymethylfurfural from Derivatives of Biomass-. Derived Glycerol for Chemicals and Polymers. Min-Shu Cui,. †,‡...
0 downloads 0 Views 1MB Size
Subscriber access provided by MAHIDOL UNIVERSITY (UniNet)

Article

Production of 4-Hydroxymethylfurfural from Derivatives of Biomass-derived Glycerol for Chemicals and Polymers Min-Shu Cui, Jin Deng, Xing-Long Li, and Yao Fu ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.5b01657 • Publication Date (Web): 01 Feb 2016 Downloaded from http://pubs.acs.org on February 7, 2016

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.

ACS Sustainable Chemistry & Engineering 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 10

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

Production of 4-Hydroxymethylfurfural from Derivatives Biomass-derived Glycerol for Chemicals and Polymers

of

Min-Shu Cui,†,‡Jin Deng,*,† Xing-Long Li,†,§ Yao Fu*,† †

iChEM, CAS Key Laboratory of Urban Pollutant Conversion, Anhui Province Key Laboratory of Biomass Clean Energy, Department

of Chemistry, University of Science and Technology of China, Hefei 230026, P. R. China ‡

Nano Science and Technology Institute, University of Science and Technology of China, Suzhou 215123, P. R. China

§

School of Medical Engineering, and Key Laboratory of Advanced Functional Materials and Devices, Hefei University of Technology,

Hefei 230009, P. R. China * Corresponding Author: Jin Deng [email protected] Abstract:Replacing petroleum feedstocks by sustainable resources requires new and efficient methods to convert simple biomassderived compounds to functionally useful and versatile chemicals. We discovered a new path for potential industrial-scale conversion of glycerol, a voluminously available byproduct from worldwide biodiesel production, to 5- and 4-hydroxymethylfurfurals through a cost effective process.Staring from glycerol derivatives, glyceraldehyde (GLYD) or dihydroxyacetone (DHA), 5-HMF or 4-HMF could be obtained through base-catalyzed condensation and acid-catalyzed dehydration steps in batch process. Effects of reaction temperature, base and acid well demonstrate the product distribution. To promote this process into commercial-scale application, we also performed the two-step conversion in continuous process.Under optimal conditions, 4-HMF could be obtained in 80% isolated yield. To interpret the potential utilization of 4-HMF, we also explored the application of 4-HMF and synthesized several functional molecules for both pharmaceutical and material sciences. Keywords:Biomass, Hydroxymethylfurfural, Glycerol, Sustainable chemistry, Glyceraldehyde, Dihydroxyacetone

Introduction In an era of diminishing fossil fuel reserves, a sustainable future for the chemical industry requires renewable feedstocks. In this context, recent efforts have been focused on converting various biomass resources to biofuels and value-added chemicals.1, 2 One such process that already produces significant social impact is the worldwide production and utilization of biodiesel made through the transesterification of vegetable oil or animal fat. A recent statistics showed that the global biodiesel production increased by 34% in 2008 and reached over 11 million tons per year.3 Hydroxymethyl furfural (HMF), including 5-HMF and 4-HMF, is biomass-derived molecule which has potentials to be upgraded into numerous valuable chemicals.4-10Previously, 5-HMF produced from biomass-derived carbohydrates11-14was proven to be a key precursor in the synthesis of liquid alkanes15-18 and 2,5-dimethylfuran19-21 which can be used as transportation fuels. 5-HMF and its derivatives may also replace petroleum-derived building blocks to produce polymers22-25 and fine chemicals26-30. Currently, reports on 4-HMF production and its utilization for fuels has been published by our group.31 The utilization of 4-HMF can also be demonstrated by its application in the synthesis of various 2,4-substituted furans. Substituted furans are important organic intermediates in both pharmaceutical and material sciences.32-35 The rapid growth of biodiesel production raises a challenging problem on how to consume the massive by-product of the transesterification process, i.e. glycerol. Intensive studies have thus been conducted to invent new transformations of glycerol and its derivatives to valued fine chemicals for future bio-refineries.36-38 Recent progresses made the conversion of glycerol to simple compounds possible such as acrolein39, 40, dihydroxyacetone41, 42,1,2-propanediol43and 1,3-propanediol44, 45. Further research is needed to convert glycerol to more functionally useful and economically feasible chemicals.Efficient conversion of derivatives of biomass-derived glycerol to HMF can provide additional support for the development of HMF-based economy. Herein we developed a novel method for the selective conversion of glycerol derivatives glyceraldehyde (GLYD) and dihydroxyacetone (DHA), to 4- and 5-hydroxymethylfurfural (HMF), thereby providing a cost effective route for the synthesis of these valuable chemical intermediates. A schematic representation of our Scheme 1. A schematic representation of the process for the conversion process is outlined in Scheme 1. The first step is a selective of bio-derived glycerol to 4- and 5-hydroxymethylfurfurals and its potential oxidation of glycerol to glyceraldehyde (GLYD) or dihydroxyacetone application. (DHA). This step has already been industrialized by using either

1

ACS Paragon Plus Environment

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 10

bio-catalyzed (for GLYD46) or metal-catalyzed oxidations (for DHA,over Pt-Bi47-49, over Au50-52, over Pd53, 54, over Ir55, 56;also see the supporting information Part 1 for details). Our main invention is the second step that involves: a) base catalyzed condensation of GLYD and/or DHA to ketohexoses, and followed by b) acid catalyzed dehydration of the ketohexoses to 5- and, more interestingly, 4HMF without separating any intermediate.

Experimental Section Procedure for model batch systems: A mixture of 0.4 g GLYD or DHA and 0.4 g Amberlite® IRA-900 basic ion exchange resin in 4 ml water is stirred at 0°C for 24 hours. After filtration to remove the resin, the solution is condensed under vacuum to remove most water. The residue is re-dissolved in 2 ml DMSO, to which is added 0.2 g Amberlyst®-15 acidic ion exchange resin or 5mol% lanthanide chloride. This mixture is stirred at 110°C for 5 hours and then cooled to room temperature and filtered. The product mixture was diluted with a known mass of deionized water, centrifuged to sediment insoluble products, and analyzed withHPLC using external standard method. Soluble bases and acids (such as NaOH and HCl) can also be employed to mediate the transformation without sacrificing yield. Procedure for model continuous systems: Aqueous solution of 10% DHA first passes through a fixed bed reactor (R1) packed with Amberlite® IRA-900 basic resin (temperature = -2°C, LHSV=1.0h-1) to produce dendroketose. The reaction solution (S1) was then condensed under reduced pressure continuously. DMSO solution of 10% dendroketose passes through the second fixed bed reactor (R2) packed with Amberlyst®-15 acidic resin (temperature = 110°C, LHSV=2.0 h-1) and give 4- and 5-HMF in yield of 84.2%. The reaction solution was condensed then added into NaHCO3 aqueous solution. 4- and 5-HMF was extracted continuously using CH2Cl2. The product was readily obtained and purified from the CH2Cl2 solution in 80% isolated yield (4-HMF:5-HMF = 99.5:0.5). Please See supporting information Part 2 for HPLC analytical methods and Part 5 for detailed procedure of 4-HMF utilization to synthesize various 2,4-substituted furans.

Results and Discussion Batch Process Starting from DHA, the reaction process produced 4-HMF in 77.7% HPLC yield. More careful analysis revealed that about 4.5% 5-HMF was also produced along with 4-HMF. On the other hand, GLYD was used as the starting material, the above process produced 5-HMF in 71.4% HPLC yield. To understand the origin of the product distribution, different ratios of GLYD and DHA were used as the starting material. It was found that when the GLYD:DHA ratio were 4:0, 3:1, 2:2, 1:3, and 0:4 in the starting material at 5°C, the 5- vs. 4-HMF ratio in the product were 71.4%:0, 69.2%:3.2%, 61.5%:12.5%, 36.1%:41.2%, and 4.5%:77.6%, respectively(Figure 1). To interpret the observations, we propose that GLYD and DHA can isomerize to each other under basic conditions (Scheme 2). The conversion of DHA to GLYD is relatively slow and the condensation of two DHA molecules only produces a branchedchain ketohexose, DL-dendroketose. This branched-chain ketohexose is proposed to give 4-HMF through acidic dehydration. On the other hand, the conversion of GLYD to DHA is a very facile process.57 Subsequent condensation of GLYD with DHA can produce straight-chain ketohexoses including DL-fructose and DLsorbose which can be further dehydrated to generate 5-HMF. Thus, DHA tends to give 4HMF as the major product, whereas 5-HMF is the major product from GLYD. In addition, GLYD and its condensation product 5-HMF could further react under acid conditions which presents a decline in HMF yield as the ratio of

Figure 1. The ratio of 4- and 5-HMF from different ratios of GLYD and DHA (Determined by HPLC). Conditions: IRA-900, 5°C, 12h in condensation step, followed by Amberlyst-15, 110°C, 5h in dehydration step.

Scheme 2. Proposed mechanism for selective 4- and 5-HMF formation from DHA and GLYD.

2

ACS Paragon Plus Environment

Page 3 of 10

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

Figure 3a.Effect of temperatures in the condensation step. Conditions: IRA-900, various temperatures, 12h in condensation step, followed by Amberlyst-15, 110°C, 5h in dehydration step.

Figure 2.13C-NMR Spectrum of DHA and GLYD in D2O.

GLYD increases. Also, we noticed that no product of GLYD self-condensation has been detected. One possible reason for this phenomenon would Figure 3b.Effect of bases in the condensation step. Conditions: bases, be the formation of GLYD hydrate in aqueous solution. Normally, 25°C, 12h in condensation step, followed by Amberlyst-15, 110°C, 5h in the reaction between substrate and its hydrate is reversible. But dehydration step. for GLYD, it totally presents in the form of hydrate in aqueous phase, hence brings about the lack of α-hydrogen which leads to a termination for self-condensation reaction. To verify this demonstration, we did 13C-NMR test for both DHA and GLYD in D2O (Figure 2). As shown in the figure, in aqueous solution, partial DHA exists as the form of its hydrate. In contrast, GLYD has been converted to its hydrate entirely hence no product of selfcondensation has been detected. Further experimental results are consistent with the above mechanistic proposal (Figure 3). Under basic conditions, DHA could form ketohexoses with bothisomer GLYD and itself.And the active energy of self-condensation is lower than that of isomerization. Therefore, we deduced that the reaction temperature would affect the product distribution during the condensation process. A series experiments were performed to explore the product distribution and ketohexoses selectivityunder different temperatures (Figure 3a). Using DHA as substrate, reactions were performed for 12h over IRA-900. When the reaction temperature increased from 0 to 90°C for the condensation step, the 4- vs. 5-HMF ratio dropped dramatically from ca. 19:1 to 2:1. This observation is explained by that the rate of DHA to GLYD conversion is accelerated under higher temperatures. And in our previous work31, we designed a series in-situ NMR tests to verify the validation of this ratio 19:1. In addition, it should be noted that DHA can also isomerize to lactic acid under higher temperatures.58-60Thus, increase of temperature also reduces the overall yield of HMFs (see supporting information Part 2 for details). Later, we explore the effect of base under 25°C in 12h (Figure 3b). When different bases were used in the ketohexose formation step to convert DHA, we observed a small but noticeable variation of 4- vs. 5-HMF ratio in the final product. This result can be explained by the varied reaction rate of DHA to GLYD conversion in the presence of base. On overall yield of HMF, catalytic activity of inorganic bases, except Ca(OH)2, is better than that of strong basic ion exchange resins. And efficiency of macroporous ion exchange 3

ACS Paragon Plus Environment

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

resin (IRA-900) is better than that of ordinary ion exchange resins (IRA-400 and IRA-402). Also, we found that over strong basic ion exchange resins, except IRA-410, the results showed higher selectivity of 4-HMF than over inorganic bases. Since the functional group of IRA-410, -N+Me2CH2CH2OH, contains hydroxyl with a lone pair, it shows more activity on the catalysis of DHA isomerization to GLYD, to increase the formation of 5-HMF. In addition, among such inorganic bases, catalytic result of Ca(OH)2 presents little consistence with that of others. A higher selectivity on 4-HMF and a lower yield of overall HMF was observed. The complexation of Ca2+ and fructose in straight-chain ketohexoses would be the main reason for this unique phenomenon. Also we noted that the 4- vs. 5-HMF ratio is much less sensitive to acid used in the dehydration step (Figure 3c).Since Amberlyst-15 is dry resin with characteristic of macroporous and excellent thermal stability, that shows more convenience for usage and less incline of coking, we decided to utilize Amberlyst-15 as acid catalyst in the following detailed studies.

Page 4 of 10

Figure 3c. Effect of bases in the condensation step and acids in the dehydration step. Conditions: bases, 25°C, 12h in condensation step, followed by acids, 110°C, 5h in dehydration step.

Continuous Process The above synthetic process optimized the detailed conditions for the conversion from DHA and GLYD to 4- and 5-HMF. To increase the application potential of the conversion, we next developed a continuous process that may be further scaled up to industrial manufacture. In the process (Figure 4). An aqueous solution of 10% DHA first passes through a fixed bed reactor (R1) packed with basic resin. Dendroketose was produced in the fixed bed reactor R1, and the solution S1was condensed continuously under reduced pressure. The catalytic activity of the Amberlite® IRA-900 basic resin was found to be fully maintained after 36 hours (Figure S8).Subsequently a DMSO solution of 10% dendroketose passes through the second fixed bed reactor (R2) packed with acidic resin.The catalytic activity of the acidic Amberlyst®-15 resin was also found to be fully maintained after 36 hours in the continuous process. 4- and 5-HMF were produced in the fixed bed reactor R2 (HMF total yield is 84.2%, products selectivity 4-HMF:5-HMF = 95.2:4.8, determined by HPLC), and the solution S2 was condensed continuously under reduced pressure. The final condensed solution of 4- and 5-HMF was added into NaHCO3 aqueous solution and subjected to a continuous extraction process with CH2Cl2. Comparing batch procedure with continuous process, we found the latter process is much more efficient especiallyin space-time yield. Detailed comparing Figure 4. Schematic diagram of the process for conversion of DHA to HMF. Diagram includes aldol experiments were performed as condensation of DHA to form dendroketose in a fixed bed reactor over a basic resin (R1); evaporation of water follows (Table 1). Initially, different from the liquid solvent containing dendroketose (S1); dehydration of dendroketose to 4-HMF over a H+ resin condensation time and temperature (R2); evaporation of DMSO from the solvent containing 4-HMF (S2); and separation of HMF from the extracting were studied followed by dehydration step in batch procedure. We found that the maximum selectivity and yield of 4-HMF was obtained as we increased the reaction time to 24h and decreased the reaction temperature to 0°C (freezing point of the solvent). Later, we performed the condensation step using continuous process at -2°C while keep dehydration step using batch procedure at 110°C for 5h. A continuous decrease of overall HMF yield was observed as the liquid hourly space velocity (LHSV) in condensation step increased. But considering both the efficiency of continuous process and 4-HMF yield, we choose 1.0h-1 as the optimal LHSV for condensation step. Then we explored the product distribution in a complete two-step continuous procedure. A series LHSV from 1.0h-1 to 6.0h-1 in 4

ACS Paragon Plus Environment

Page 5 of 10

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

dehydration step were tested following the optimal condensation step. And ultimately we achieved the best selectivity and yield of 4HMF when LHSV was 2.0h-1. The space-time yield in the condensation step was found to be 48 times higher compared to the batch process and ca. 10 times higher in the dehydration step. Also, it is worth noting that when the liquid hourly space velocity (LHSV) in the condensation step is increased, the total HMF yield decreased remarkably but the 4-HMF selectivity increased. This observation is consistent with our above explanation that the conversion of DHA to dendroketose is faster than the isomerization of DHA to GLYD. In contrast, when the LHSV in the dehydration step is increased, the total HMF yield slightly increases at first and then gradually decreases but the 4-HMF selectivity increases a bit. This observation demonstrates that the dehydration of branched-chain ketohexose to 4-HMF is easier than that of straight-chain ketohexose to 5-HMF. It is proposed that straight-chain ketohexose could form the pyranose and furanose, whereas the branched-chain ketohexose could only form the furanose which could convert to furan derivative more easily than pyranose. Table 1. Comparison of Batch Technics and Continuous Technics. Condensation step Tec

Batch

Reaction time/LHSV

Reaction temp

12h

5℃

12h

0℃

24h

Batch

HMF yield a

4-HMF: :5HMF b

4-HMF yield a

5-HMF yield a

110℃

82.2%

94.5:5.6

77.7%

4.5%

5h

110℃

82.3%

95.4:4.6

78.5 %

3.8%

Reaction time/LHSV

Reaction temp

5h

0℃

5h

110℃

85.1%

95.1:4.9

81.0%

4.1%

-2℃

5h

110℃

86.2%

93.3:6.7

80.4%

5.8%

-1

-2℃

5h

110℃

83.6%

94.1:5.9

78.7%

4.9%

-1

-2℃

5h

110℃

76.0%

94.8:5.2

72.0%

4.0%

-1

-2℃

5h

110℃

55.2%

95.8:4.2

52.9%

2.3%

-1

-2℃

5h

1.0h 2.0h 3.0h 5.0h

-1

1.0h

-1

1.0h

-1

1.0h Continuous

Tec

-1

0.5h

Continuous

Dehydration step

-1

1.0h

-1

1.0h

-1

1.0h

Batch

-2℃

33.4%

1.3%

110℃

83.6%

95.2:4.8

79.6%

4.0%

-1

110℃

84.2%

95.2:4.8

80.2%

4.0%

-1

110℃

83.9%

95.3:4.7

80.0%

3.9%

-1

110℃

81.9%

95.4:4.6

78.1%

3.8%

-1

110℃

81.6%

95.6:4.4

78.0%

3.6%

-1

110℃

79.8%

95.7:4.3

76.4%

3.4%

3.0h Continuous

-2℃

96.1:3.9

2.0h

-2℃

-2℃

34.8%

1.0h

-2℃

-2℃

110℃ -1

4.0h 5.0h 6.0h

a. HPLC yield, b. determined with HPLC.

Application of 4-HMF As mentioned before, 4-HMF could be upgraded to valuable 2,4-substituted furans which are important organic intermediates in both pharmaceutical and material sciences. Previous synthesis of 2,4-substituted furans usually required multiple steps and therefore 2,4substituted furans are fairly expensive compounds.61, 62 In this study, we have successfully synthesized 4-HMF from 1,3dihydroxyacetone in 80% isolated yield (eq. 1). From 4-HMF we can readily synthesize other 2,4-disubstituted furans as functional molecules including furan-2,4-diyldimethanol (eq. 2, which could convert to inhibitor of calcineurin32), 4-(hydroxymethyl)furan-2carboxylic acid (eq. 3, which could convert to prostaglandin receptor EP433), furan-2,4-dicarboxylic acid (eq. 4, which could convert to liquid crystal materials34), and (E)-5-(3-ethoxy-3-oxoprop-1-enyl)furan-3-carboxylic acid (eq. 5, which could convert to photo crosslinking of copolymers35) in high yields compared with former methods (Scheme 3). Thus the current study demonstrates new methods for the economical synthesis of 2,4-disubstituted furans.Moreover, with these intermediates on hand, we studied the new synthesis pathway for several functional 2,4-disubstituted furans molecules, including 1,5-dibenzoyloxymethyl-7-oxabicyclo[2.2.1]hept-5-ene2,3-dicarboxylic (cantharidin compound), 2,4-furandicarboxylic acid bis(4-methoxyphenyl) (nematic liquid crystal molecules), and 4hydroxymethyl-2-furanacrylic acid ethyl ester and ethyl-6-hydroxyhexonoate copolymer (S9-15). Cantharidin analogues can act on a series of tumor cells since it is capable of inhibiting the activity of calcineurin and affecting the cell cycle. Sodeoka et al.32 synthesized cantharidin compounds which showed a highly targeted inhibition of calcineurin in a total yield of 10% in 2003 for the first time (S10). The main reason would be explained as follows: 3-substituted furans were used to build the structure of 2,4- disubstituted furans in 89% yield, while about 75% products is 2,3- disubstituted furans. This phenomenon also makes Keay et al. 61, 62believe that it is difficult to synthesize 2,4-disubstituted furans.Since we have already obtained 2,4-disubstituted

5

ACS Paragon Plus Environment

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 10

furan compound 4-HMF in high yield, we tried to apply 4-HMF as starting material in the synthesis of cantharidin and achieved a yield of 71% (eq. 6). Nematic liquid crystal molecules are generally utilized in LCD TVs, notebook computers, and various types of display element. Dewar et al.34 demonstrated the key factor that influence the stability of nematic liquid crystal molecules is the bond-angle between the double carboxyl groups linking with the central ring. As the first discovered liquid crystal, terephthalic acid diester molecules, the bond-angle between two carboxyl is 180°. Comparing with the common replacement of petroleum-derived terephthalic acid, 2,5-furan dicarboxylic acid, which bond-angle between two carboxyl is 137°, 2,4- furan dicarboxylic acid is more suitable for synthesis of nematic liquid crystal molecules since the bond-angle is 160°. In this work (eq. 7), we obtained the nematic liquid crystal molecule 2,4-furandicarboxylic acid bis(4-methoxyphenyl) in 82% isolated yield (S12). Also, report has been published35 that esters of 5-HMF derived from renewable sources could be used to synthesize photosensitive polyesters for negative offset plates (S13). In order to expand the application of 4-HMF, we tried to synthesis photosensitive copolymer using 4-hydroxymethyl-2-furanacrylic acid ethyl ester (4-HMFAE) as starting material. Referred to published literature, 4-HMFAE and ethyl-6-hydroxyhexonoate were stirred in ratio 1:9 under optimal conditions and obtained 4HMFAE and ethyl-6-hydroxyhexonoate copolymer which had similar photochemistry characterization with 5-HMF based copolymer (eq. 8).

Conclusion In summary, we discovered a novel chemical path for the conversion of GLYD and DHA to 5- and, more interestingly, 4HMF through an operationally simple and cost effective process in ca. 80% yields. This transformation is not only new in organic chemistry, but also important because large quantities of Scheme 3. Synthesis of 4-HMF from 1,3-dihydroxyacetone and synthesis biomass-derived glycerol may be efficiently converted through of various 2,4-substituted furans from 4-HMF. this approach to 5- and 4-HMF which are industrially useful and versatile chemicals. Previous studies have already showed that 5-HMF can serve as a vital intermediate for future bio-refineries to produce liquid fuels, pharmaceutics, and polymers. The chemical analogy between 4- and 5-HMF implies that 4-HMF from bio-derived glycerol may become a new type of voluminously available platform molecule for the development of sustainable HMF-based economy in industrial scale.

ASSOCIATED CONTENT Supporting Information Oxidation of glycerol, data of figure 1-3,and detailed preparation and NMR spectrum of compound 1-8are available in SI.

Author Information CorrespondingAuthor *Jin Deng. E-mail: [email protected];* Yao Fu. E-mail: [email protected]. Fax: +86-551-6360-6689.

Acknowledgements This work was supported by the 973 Program (2012CB215305), NSFC (21402181, 21325208, 21172209), IPDFHCPST (2014FXCX006), CAS (KJCX2-EW-J02), SRFDP (20123402130008), CPSF (2014M561835), FRFCU (WK2060190025, WK2060190033) and PCSIRT. 6

ACS Paragon Plus Environment

Page 7 of 10

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

References (1) Gandini, A.; Lacerda, T. M.; Carvalho, A. J. F.; Trovatti, E. Progress of Polymers from Renewable Resources: Furans, Vegetable Oils, and Polysaccharides. Chem. Rev. 2015, DOI: 10.1021/acs.chemrev.5b00264. (2) Caes, B. R.; Teixeira,R. E.; Knapp,K. G.; Raines, R. T. Biomass to Furanics: Renewable Routes to Chemicals and Fuels. ACS Sustainable Chem. Eng.2015, 3, 2591–2605. (3) Emerging markets online. Biodiesel 2020: a global markets survey, available at {http://www.emerging-markets.com/biodiesel/}. (4) van Putten,R.-J.;van der Waal,J. C.;de Jong, E.; Rasrendra, C. B.; Heeres,H. J.; Vries, J. G.Hydroxymethylfurfural, A Versatile Platform Chemical Made from Renewable Resources. Chem. Rev.2013, 113, 1499–1597. (5) Nagpure, A. S.; Lucas, N.; and Chilukuri, S. V. Efficient Preparation of Liquid Fuel 2,5-Dimethylfuran from Biomass-Derived 5Hydroxymethylfurfural over Ru–NaY Catalyst. ACS Sustainable Chem. Eng.2015, 3, 2909–2916. (6) Bohre, A.; Dutta, S.; Saha, B.; and Abu-Omar, M. M. Upgrading Furfurals to Drop-in Biofuels: An Overview. ACS Sustainable Chem. Eng.2015, 3, 1263–1277. (7) Wang, S.; Zhang, Z.; Liu, B. Catalytic Conversion of Fructose and 5-Hydroxymethylfurfural into 2,5-Furandicarboxylic Acid over a Recyclable Fe3O4−CoOx Magnetite Nanocatalyst. ACS Sustainable Chem. Eng.2015, 3, 406−412. (8) Cao, X.; Teong, S. P.; Wu, D.; Yi, G.; Su, H.; Zhang, Y. An enzyme mimic ammonium polymer as a single catalyst for glucose dehydration to 5-hydroxymethylfurfural. Green Chem.2015, 17, 2348-2352. (9) Teong, S. P.; Yi, G.; Cao, X.; Zhang, Y. Poly-benzylic Ammonium Chloride Resins as Solid Catalysts for Fructose Dehydration. ChemSusChem2014, 7, 2120- 2126. (10) Lai, L.; Zhang Y. The Production of 5-Hydroxymethylfurfu ral from Fructose in Isopropyl Alcohol : A Green and Efficient System. ChemSusChem2011, 4, 1745-1748. (11) Rosatella, A. A.; Simeonov, S. P.; Fradea, R. F. M.; Afonso, C. A. M. 5-Hydroxymethylfurfural (HMF) as a building block platform: Biological properties, synthesis and synthetic applications. Green. Chem.2011, 13, 754-793; (12) Zakrzewska, M. E.; Bogel-Łukasik, E.; Bogel-Łukasik, R. Ionic Liquid-Mediated Formation of 5-Hydroxymethylfurfural—A Promising Biomass-Derived Building Block. Chem. Rev.2011, 111, 397-417. (13) Lewkowski,J. Synthesis, chemistry and applications of 5-hydroxymethylfurfural and its derivatives. Arkivoc2001, Part 1, 17-54. (14) Schön, M.; Schnürch, M.; Mihovilovic, M. D.Application of continuous flow and alternative energy devices for 5hydroxymethylfurfural production. Mol Divers2011, 15, 639-643. (15) Huber, G. W.; Chheda, J. N.; Barrett, C. J.; Dumesic, J. A. Production of Liquid Alkanes by Aqueous-Phase Processing of Biomass-Derived Carbohydrates. Science2005, 308, 1446-1450. (16) Chatterjee, M.; Matsushima, K.; Ikushima, Y.; Sato, M.; Yokoyama, T.; Kawanami, H.; Suzukia, T.; Production of linear alkane via hydrogenative ring opening of a furfural-derived compound in supercritical carbon dioxide. Green. Chem.2010, 12, 779-782. (17) Julis, J.; Hölschera, M.; Leitner, W. Selective hydrogenation of biomass derived substrates using ionic liquid-stabilized ruthenium nanoparticles. Green. Chem.2010, 12, 1634-1639. (18) West, R. M.; Liu, Z. Y.; Peter, M.; Dumesic,J. A. Liquid Alkanes with Targeted Molecular Weights from Biomass-Derived Carbohydrates. ChemSusChem2008, 1, 417-424. (19) Román-Leshkov, Y.; Barrett, C. J.; Liu, Z. Y.; Dumesic, J. A. Production of dimethylfuran for liquid fuels from biomass-derived carbohydrates. Nature2007, 447, 982-986. (20) Thananatthanachon, T.; Rauchfuss, T. B. Efficient Production of the Liquid Fuel 2,5-Dimethylfuran from Fructose Using Formic Acid as a Reagent. Angew. Chem. Int. Ed.2010, 49, 6616-6618. (21) Chidambaram, M.; Bell, A. T. A two-step approach for the catalytic conversion of glucose to 2,5-dimethylfuran in ionic liquids. Green Chem.2010, 12, 1253-1262. (22) Gandini, A.; Silvestre, A. J. D.; Neto, C. P.; Sousa, A. F.; Gomes, M. The furan counterpart of poly (ethylene terephthalate): An alternative material based on renewable resources. J Polym. Sci. Pol. Chem. 2009, 47, 295-298. (23) Eerhart, A. J. J. E.; Faaij, A. P. C.; Patel, M. K.Replacing fossil based PET with biobased PEF; process analysis, energy and GHG balance. Energy Environ. Sci.2012, 5, 6407-6422. (24) Jiang, M.; Liu, Q.; Zhang, Q.; Ye, C.; Zhou, G. Y. A series of furan-aromatic polyesters synthesized via direct esterification method based on renewable resources. J. Polym. Sci. A2012, 50, 1026-1036. (25) Ma, J. P.; Yu, X. F.; Xu, J.; Pang, Y. Synthesis and crystallinity of poly (butylene 2,5-furandicarboxylate). Polymer2012, 53, 4145-4151. (26) Buntara, T.; Noel, S.; Phua, P. H.; Melián-Cabrera, I.; Vries, J. G.; Heeres,H. J. Caprolactam from Renewable Resources: Catalytic Conversion of 5-Hydroxymethylfurfural into Caprolactone. Angew. Chem. Int. Ed. 2011, 50, 7083-7087. (27) Du, Z. T.;Ma, J.P.; Wang, F.; Liu, J. X.; Xu, J. Oxidation of 5-hydroxymethylfurfural to maleic anhydride with molecular oxygen. Green Chem.2011, 13, 554-557.

7

ACS Paragon Plus Environment

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 8 of 10

(28) Yang, Z. Z.; Deng, J.; Pan, T.; Guo, Q. X.; Fu, Y. A one-pot approach for conversion of fructose to 2,5-diformylfuran by combination of Fe3O4-SBA-SO3H and K-OMS-2. Green Chem.2012, 14, 2986-2989. (29) Alamillo, R.; Tucker, M.; Chia, M.; Pagán-Torres, Y.; Dumesic, J. The selective hydrogenation of biomass-derived 5hydroxymethylfurfural using heterogeneous catalysts. Green Chem.2012, 14, 1413-1419. (30) Gupta, N. K.; Nishimura, S.; Takagaki, A.; Ebitani, K. Hydrotalcite-supported gold-nanoparticle-catalyzed highly efficient basefree aqueous oxidation of 5-hydroxymethylfurfural into 2,5-furandicarboxylic acid under atmospheric oxygen pressure. Green Chem.2011, 13, 824-827. (31) Deng, J.; Pan T.; Xu, Q.; Chen, M.; Zhang, Y.; Guo, Q.; Fu Y. Linked strategy for the production of fuels via formose reaction. Scientific Reports 2013, 3. (32) Baba, Y.; Hirukawa, N.; Tanohira, N.; Sodeoka, M. Structure-Based Design of a Highly Selective Catalytic Site-Directed Inhibitor of Ser/Thr Protein Phosphatase 2B (Calcineurin). J. Am. Chem. Soc.2003, 125, 9740-9749. (33) Clark, D. E.; Clark, K. L.; Coleman, R. A.; Davis, R. J. WO2004067524, 2004. (34) Dewar, M. J. S.; Riddle, R. M.; Factors influencing the stabilities of nematic liquid crystals. J. Am. Chem. Soc.1975, 97, 66586662. (35) E. Lasseuguettea, A. Gandini, M. N. Belgacema, H-J. Timpe, Synthesis, characterization and photocross-linking of copolymers of furan and aliphatic hydroxyethylesters prepared by transesterification. Polymer2005, 46, 5476-5483. (36) Lopes, J. M.; Petrovski, Z.; Bogel-Łukasik, R.; Bogel-Łukasik, E.; Heterogeneous palladium-catalyzed telomerization of myrcene with glycerol derivatives in supercritical carbon dioxide: a facile route to new building blocks. Green. Chem.2011, 13, 20132016. (37) Conceiçao, L.; Bogel-Łukasik, R.; Bogel-Łukasik, E. Supercritical CO2 as an effective medium for a novel conversion of glycerol and alcohols in the heterogeneous telomerisation of butadiene. Green. Chem.2012, 14, 673-681. (38) Melo, C. I.; Rodrigues, A. I.; Bogel-Łukasik, R.; Bogel-Łukasik, E. Outlook on the Phase Equilibria of the Innovative System of “Protected Glycerol”: 1,4-Dioxaspiro[4.5]decane-2-methanol and Alternative Solvents. J. Phys. Chem.A2012, 116, 1765-1773. (39) Liu, L.; Ye, X. P.; Bozell, J. J. A Comparative Review of Petroleum-Based and Bio-Based Acrolein Production. ChemSusChem2012, 5, 1162 – 1180. (40) Martin. A.; Armbruster, U.; Atia, H. Recent developments in dehydration of glycerol toward acrolein over heteropolyacids. Eur. J. Lipid Sci. Technol.2012, 114, 10–23. (41) Lari, G. M.; Mondelli, C.; Pérez-Ramı́rez, J. Gas-Phase Oxidation of Glycerol to Dihydroxyacetone over Tailored Iron Zeolites. ACS Catal.2015, 5, 1453–1461. (42) Zhang, Y.; Zhang, N.; Tong, Z. -R.; Xu, Y. –J. Identification of Bi2WO6 as a highly selective visible-light photocatalyst toward oxidation of glycerol to dihydroxyacetone in water. Chem. Sci.2013, 4, 1820-1824. (43) Wang, Y.; Zhou, J.; Guo, X. Catalytic hydrogenolysis of glycerol to propanediols: a review. RSC Adv.2015, 5, 74611-74628. (44) Priya, S. S.; Kumar, V. P.; Kantam, M. L.; Bhargava, S. K.; Srikanth, A.; Chary, K. V. R. High Efficiency Conversion of Glycerol to 1,3-Propanediol Using a Novel Platinum–Tungsten Catalyst Supported on SBA-15. Ind. Eng. Chem. Res.2015, 54, 9104 – 9115. (45) Chen, S.; Qi, P.; Chen, J.; Yuan, Y. Platinum nanoparticles supported on N-doped carbon nanotubes for the selective oxidation of glycerol to glyceric acid in a base-free aqueous solution. RSC Adv.2015, 5, 31566–31574. (46) Wolf, H. J. US Patent 4353987, 1982. (47) Kimura, H.; Tsuto, K.; Wakisaka, T.; Kazumi, Y.; Inaya, Y. Selective oxidation of glycerol on a platinum-bismuth catalyst. Appl. Catal. A: General1993, 96, 217-228. (48) H. Kimura, K. Tsuto, US Patent 5274187, 1993. (49) Hu, W.; Knight, D.; Lowry, B.; Varma, A.; Selective Oxidation of Glycerol to Dihydroxyacetone over Pt−Bi/C Catalyst: Optimization of Catalyst and Reaction Conditions. Ind. Eng. Chem. Res.2010, 49, 10876-10882. (50) Demirel, S.; Lehnert, K.; Lucas, M.; Claus, P. Use of renewables for the production of chemicals: Glycerol oxidation over carbon supported gold catalysts. Appl. Catal. B: Environmental2007, 70, 637-643. (51) Dimitratos, N.; Wagland, A.; Hutchings, G. J.; Stitt,E. H. Enhanced selective glycerol oxidation in multiphase structured reactors. Catal. Today2009, 145, 169-175. (52) Rodrigues, E. G.; Pereira, M. F. R.; Chen, X.; Delgado, J. J.; Orfao, J. J. M. Influence of activated carbon surface chemistry on the activity of Au/AC catalysts in glycerol oxidation. J. Catal.2011, 281, 119-127. (53) Painter, R. M.; Pearson, D. M.; Waymouth,R. M. Selective Catalytic Oxidation of Glycerol to Dihydroxyacetone. Angew. Chem. Int. Ed.2010, 49, 9456-9459. (54) Hirasawa, S.; Nakagawa, Y.; Tomishige, K. Selective oxidation of glycerol to dihydroxyacetone over a Pd-Ag catalyst. Catal. Sci. Technol.2012, 2, 1150-1152. (55) Farnetti, E.; Kašpara, J.; Crotti,C. A novel glycerol valorization route: chemoselective dehydrogenation catalyzed by iridium derivatives. Green Chem.2009, 11, 704-709.

8

ACS Paragon Plus Environment

Page 9 of 10

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

(56) Crotti, C.; Kašparb, J.; Farnetti, E. Dehydrogenation of glycerol to dihydroxyacetone catalyzed by iridium complexes with P–N ligands. Green Chem.2010, 12, 1295-1300. (57) Nagorski, R. W.; Richard, J. P. Mechanistic Imperatives for Aldose−Ketose Isomerization in Water:  Specific, General Baseand Metal Ion-Catalyzed Isomerization of Glyceraldehyde with Proton and Hydride Transfer. J. Am. Chem. Soc.2001, 123, 794-802. (58) Yang, G.-Y.; Ke, Y.-H.; Ren H.-F.; Liu, C.-L.; Yang, R.-Z.; Dong, W.-S. The conversion of glycerol to lactic acid catalyzed by ZrO2-supported CuO catalysts. Chem. Eng.J. 2016, 283, 759-767. (59) Wang, Y.; Deng, W.; Wang, Y.; Wan, H. Chemical synthesis of lactic acid from cellulose catalysed by lead (II) ions in water. Nat. Commun. 2013, 4. (60) Sharninghausen, L. S.; Campos, J.; Manas, M. G.; Crabtree, R. H. Efficient selective and atom economic catalytic conversion of glycerol to lactic acid. Nat. Commun. 2014, 5. (61) Bures, E.; Spinazz, P. G.; Beese, G.; Hunt, I. R.; Rogers, C.; Keay, B. A. Regioselective Preparation of 2,4-, 3,4-, and 2,3,4Substituted Furan Rings. 1. [1,4] O f C and [1,4] C f O Silyl Migrations of Silyl Ethers and Esters Attached to Furan and Thiophene Rings. J. Org. Chem.1997, 62, 8741-8749. (62) Bures, E.; Nieman, J. A.; Yu, S. Y.; Spinazz, P. G.; Bontront, J. J.; Hunt, I. R.; Rauk, A.; Keay, B. A. Regioselective Preparation of 2,4-, 3,4-, and 2,3,4-Substituted Furan Rings. 2.1 Regioselective Lithiation of 2-Silylated-3-substituted Furan Rings. J. Org. Chem.1997, 62, 8750-8759.

9

ACS Paragon Plus Environment

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

For Table of Contents Use Only:

Production of 4-Hydroxymethylfurfural from Derivatives of Biomass-derived Glycerol for Chemicals and Polymers Min-Shu Cui, Jin Deng, Xing-Long Li, Yao Fu A new path for potential industrial-scale conversion of glycerol and potential application for the products.

10

ACS Paragon Plus Environment

Page 10 of 10