Route for Conversion of Furfural to Ethylcyclopentane - ACS Omega

2 days ago - Institute of Technology, University of Tartu, Nooruse 1, Tartu 50411 , ... ACS AuthorChoice - This is an open access article published un...
0 downloads 0 Views 428KB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2018, 3, 10211−10215

http://pubs.acs.org/journal/acsodf

Route for Conversion of Furfural to Ethylcyclopentane Aleksei Bredihhin,† Siim Salmar,‡ and Lauri Vares*,† †

Institute of Technology, University of Tartu, Nooruse 1, Tartu 50411, Estonia Institute of Chemistry, University of Tartu, Ravila 14a, Tartu 50411, Estonia



ACS Omega 2018.3:10211-10215. Downloaded from pubs.acs.org by 185.250.43.21 on 08/31/18. For personal use only.

S Supporting Information *

ABSTRACT: A method for conversion of furfural, widely available platform chemical from biomass, to ethylcyclopentane, is reported. Ethylcyclopentane is a cyclic alkane with a relatively high octane number (RON 67, bp 103 °C) and could potentially serve as a drop-in biofuel. The synthetic route includes a transformation of furfural to 1-(furan-2yl)propan-1-ol that is further subjected to Piancatelli rearrangement to give 5-ethyl-4-hydroxycyclopent-2-en-1-one. The subsequent hydrodeoxygenation affords ethylcyclopentane in 48% isolated yield.



INTRODUCTION To achieve liquid transportation, fuels from biomass have been an intense research object in recent years.1−6 Among the various routes studied, conversion of bio-derived furanes to alkanes has received significant attention.7−13 One of the key steps for obtaining alkanes from furanes is total hydrodeoxygenation (HDO), and a number of different catalysts, including lanthanide triflates,10−13 have recently been developed14 for this transformation. Most of the reported methods provide linear or lowly branched alkanes that have low research octane numbers (RON). Synthesis of cyclic alkanes, which have higher RON numbers, requires different approach. One possibility is to apply the Piancatelli rearrangement,15,16 which allows the transformation of 2-furylcarbinols (including furfuryl alcohol) into substituted cyclopentenones. Cyclopentenones are an important scaffold for the synthesis of a variety of biologically active target molecules17,18 and therefore the Piancatelli rearrangement has been extensively studied. Recent advances in this field include a use of new catalytic systems,19−21 microwave irradiation,22 and an aza version of this rearrangement.23−28 In addition, this transformation has been applied to achieve one-step conversion of furfural to cyclopentanone29,30 in high yield. Cyclopentanone is an important intermediate in biofuel production and it can be reduced to cyclopentane (Ru/ZSM-5, 4 MPa H2, 180 °C).30 Furthermore, the product of self-condensation of cyclopentanone can be hydrodeoxygenated to bicyclopentane (Pd−SiO2 or Ni−SiO2-DP, 6 MPa H2, 230 °C),30−33 which is a perspective jet range biofuel (Scheme 1a). In contrast to furfuryl alcohol, HMF does not undergo Piancatelli rearrangement under normal conditions. However, recently, Ohyama has reported a direct transformation of HMF to 3-hydroxymethyl-cyclopentanone using hydrogenative conditions (3−8 MPa H2, 140 °C) in combination with lanthanide metal oxide catalysts (Scheme 1b).34−36 In addition, 3-methyl2-cyclopenten-1-one was formed in a low yield during the reduction of HMF with Zn.37 HDO of these cyclopentane skeleton containing compounds, obtained from HMF, could © 2018 American Chemical Society

afford methylcyclopentane (MCP), which can also serve as a bio-based drop-in gasoline substituent. A different approach to access MCP was recently described by Bell et al.38 In this procedure MCP was obtained by selfcondensation of 2,5-hexanedione and subsequent HDO (Pt/ NbOPO4, 3.0 MPa H2, 150 °C) of resulting 3-methylcyclopent-2-enone. In addition, MCP is forming as a low-yield byproduct during HDO of cellulose,39 sorbitol,40 and guaiacol41 to alkanes. However, to the best of our knowledge ethylcyclopentane is a completely new compound for application as biofuel. It has the boiling point of 103 °C and relatively high octane number (RON 67),42 making it a potential drop-in biofuel. In this study, we report a route for obtaining ethylcyclopentane from furfural, which includes addition of ethyl group to furfural, Piancatelli rearrangement, and subsequent HDO of obtained cyclopentenone 3 (Scheme 1c).



RESULTS AND DISCUSSION

Furfural (1) was easily converted to 1-(furan-2-yl)propan-1-ol (2) in 96% yield using ethylmagnesium bromide (Scheme 2). Addition of ethylmagnesium halide to furfural is a known reaction,43 and this step is somewhat problematic in large scale because of the stoichiometric amount of Grignard reagent. However, our route not only introduces the ethyl group but also the OH group, which is necessary for the subsequent Piancatelli rearrangement. Also, ethylmagnesium bromide is a widely available reagent potentially obtainable directly from ethanol or via bioethylene. To minimize the amount of used solvents, furfural was added neat to the EtMgBr in MTHF and the product was isolated without any chromatography. In addition, magnesium halides, which are formed in the work-up Received: March 28, 2018 Accepted: August 7, 2018 Published: August 30, 2018 10211

DOI: 10.1021/acsomega.8b00588 ACS Omega 2018, 3, 10211−10215

ACS Omega

Article

Scheme 1. Various Synthetic Strategies toward Cyclopentanic Alkanes from Furanes

point for the optimization of conditions for our furfuryl alcohol 2. In our hands, the use of 5 mol % trifluoroacetic acid (TFA) together with 10 mol % of Dy(OTf)3 in t-BuOH/water (5:1) system resulted in relatively low yield and incomplete conversion (Table 1, entry 1). The use of La(OTf)3 instead of Dy(OTf)3 gave no improvement (entry 2). Increasing the reaction time to 48 h afforded cyclopentenone 3 in 51% yield (entry 3). Further increase of the temperature and catalytic loading resulted in nearly the same result in 24 h (50%, entry 4). The use of other solvent systems (AcOH/water 5:1 and AcCN/water 5:1) mainly led to a complex mixture of byproducts (data not shown). In addition, the reaction in the t-BuOH/water system in the absence of the lanthanide catalyst using correspondingly 5, 10, and 20 mol % of TFA was studied. Generally, longer reaction time and higher catalyst loading afforded higher product yields. The best result was achieved using 20 mol % of TFA (47% yield, entry 5). Amberlyst-15 as a catalyst was also tested; however, it led to the formation of byproducts and only a minor amount of target cyclopentenone was observed. The byproducts are formed mainly through the elimination reaction and Friedel−Crafts alkylation. Because the use of lanthanide or acid catalyst furnished the yields not exceeding 51%, we applied a modified procedure that was previously reported by Frelek et al.44 for obtaining the compound 3 (entry 6). This reaction proceeded in pure boiling water in relatively concentrated fashion (6.8 g of 2 in 170 mL H2O) without any added catalyst. Moreover, our modification allowed separating the product in pure form, without using any chromatography in 51% yield. Considering the simplicity and avoidance of relatively expensive lanthanide catalysts makes this procedure (entry 6) superior to the other methods described in Table 1. Next, HDO of cyclopentenone 3 using the La(OTf)3 catalyst in combination with Pd catalyst and H2 gas was explored (Table 2). At 20 bar H2 (34 bar at 200 °C), the use of 0.5 mol % Pd catalyst resulted in ethylcyclopenanone as the main product (entry 1). At the same time, the use of 1.6 mol % Pd afforded a complex mixture of products, where only traces of desired ethylcyclopentane (4) was detected (entry 2). Increasing the hydrogen pressure to 40 bar (64 bar at 200 °C) afforded 19% of desired product (entry 3). An exchange of the solvent to EtOAc increased the yield of 4 to 56% (entry 4).

Scheme 2. Synthesis of 1-(Furan-2-yl)propan-1-ol

process, could be used as de-icing reagents on roads (MgCl2) or for the production of magnesium. Next, the obtained alcohol 2 was subjected to Piancatelli rearrangement and the reaction conditions were optimized (Table 1). Recently, methods for performing Piancatelli Table 1. Optimization of Piancatelli Rearrangement of 2a

entry

catalyst

1 2 3 4

10 mol % Dy(OTf)3 5 mol % TFA 10 mol % La(OTf)3 5 mol % TFA 10 mol % Dy(OTf)3 5 mol % TFA 10 mol % Dy(OTf)3 10 mol % TFA 20 mol % TFA none

5 6

temp (°C)

time (h)

yielda (%)

80 80 80 90

24 24 48 24

44 43 51 50

90 100

48 4

47 51b

a

Isolated yield. bReaction was performed in pure distilled water.

rearrangement of alkyl-substituted furyl carbinols in water were described.22,44,45 However, in these reactions, the yield of target compounds was relatively low and 4-hydroxy-2methylcyclopent-2-en-1-one was obtained in pure boiling water44 in 40% yield. Addition of MgSO4 as the catalyst at 160 °C increased to yield to 56%.45 In addition, Reiser has described the method for Piancatelli rearrangement of compound 2 in water using microwave irradiation in 54% yield. On the other hand, a recent work of Read de Alaniz suggests that alkyl-substituted furyl carbinols can be transformed to cyclopentenones in up to 90% yield.19 In this study, however, the yield of the product varied greatly depending of the alkyl substituent used (90% for butyl, but 56% for isopropyl). Thus, we have decided to use this work as a starting 10212

DOI: 10.1021/acsomega.8b00588 ACS Omega 2018, 3, 10211−10215

ACS Omega

Article

Table 2. Optimization of HDO Conditionsa

entry

H2 (bar)

co-catalyst

solvent

yieldb (%)

1 2 3 4 5 6 7 8 9 10

20 20 40 40 40 40 40 40 40 40

La(OTf)3 La(OTf)3 La(OTf)3 La(OTf)3 La(OTf)3 Dy(OTf)3 La(OTf)3 La(OTf)3 Dy(OTf)3 La(OTf)3

AcOH AcOH AcOH EtOAc EtOAc EtOAc AcOH EtOAc EtOAc neat

tracesc traces 19 56 85d 94d 23e 29e 36e,f 48g

reagents were purged with argon prior to use. All reagents and solvents including anhydrous THF were obtained from commercial sources and used as received, without additional purification. HDO reactions were performed in a Parr Instrument Company pressure vessel. Petroleum ether (PE) used refers to the fraction boiling in the 40−65 °C range. Nuclear magnetic resonance (NMR) spectroscopy was performed on a Bruker Avance II 400 MHz spectrometer using the residual solvent peak (CDCl3, 7.26 ppm for 1H and 77.16 ppm for 13C spectra) as an internal standard. The infrared spectra were measured on Shimadzu IRAffinity-1 FTIR spectrometer using ATR module with a ZnSe crystal. High-resolution mass spectrometry (HRMS) measurements were performed on a Thermo Electron LTQ Orbitrap HR-MS spectrometer. GC−MS measurements were performed on an Agilent Technologies 7890A gas chromatograph equipped with a 5975C inert MSD system with Triple-Axis Detector and HP5MS (30 m, 0.25 mm, 0.25 μ) column. The reactions were monitored using thin-layer chromatography and visualized with UV light or KMnO4 solution. The products were purified using flash chromatography on silica gel 60 (0.040−0.063 mm, 230−400 mesh ASTM). Synthetic Procedures. 1-(Furan-2-yl)propan-1-ol (2). A 100 mL round-bottom flask was charged with EtMgBr solution in MeTHF (3.056 M, 33 mL, 100 mmol), cooled to 0 °C, and then neat furfural (9.608 g, 100 mmol) was added dropwise in 15−30 min. The resulting thick suspension was diluted with dry THF (10 mL). Then, the cooling bath was removed and the reaction mixture was allowed to stir for 2 h at rt. After this, the reaction mixture was quenched with water (3 × 10 mL). NB. The mixture heats up strongly. After cooling, the resulting thick suspension was transferred to a separation funnel and partitioned between Et2O (100 mL) and hydrochloric acid (aq) (1 M, 100 mL). The organic fraction was separated and the water fraction was additionally extracted with Et2O (2 × 100 mL). Organic fractions were combined, dried with MgSO4, and concentrated under reduced pressure. The crude product was diluted with a mixture of PE and diethyl ether (1:1, 30 mL), filtered through a short plug of silica gel, and evaporated again to give a product (12.107 g, 96%) as yellowish oil. 1H NMR (400 MHz, CDCl3): δ 7.36 (dd, J = 1.8, 0.7 Hz, 1H), 6.32 (dd, J = 3.2, 1.8 Hz, 1H), 6.22 (dt, J = 3.2, 0.7 Hz, 1H), 2.07 (br s, 1H), 4.59 (t, J = 6.8 Hz, 1H), 1.79−1.93 (m, 2H), 0.94 (t, J = 7.4 Hz, 3H). 13C NMR (100 MHz, CDCl3): δ 156.9, 142.0, 110.2, 106.0, 69.3, 28.8, 10.0. IR (ATR-FTIR, neat): ν 3352, 2967, 2936, 2878, 1458, 1226, 1150, 937, 880, 810, 733 cm−1. HRMS (ESI) m/z: [M + Na]+ calcd for C7H10O2Na, 149.0573; found, 149.0567. 5-Ethyl-4-hydroxycyclopent-2-en-1-one (3). Procedure A. A round-bottom flask (250 mL) was charged with 1-(furan-2yl)propan-1-ol (2) (1.262 g, 10 mmol), and t-BuOH (70 mL), water (14 mL), Dy(OTf)3 (600 mg, 10 mol %), and TFA (80 μL, 10 mol %) were added. The reaction mixture was stirred at rt for 5 min to obtain homogeneous solution. After this, the flask was equipped with a condenser, placed in a preheated (90 °C) oil bath, and stirred for 24 h. Then, the reaction mixture was cooled down and transferred to a separation funnel and partitioned between Et2O (70 mL) and a mixture of aqueous solutions of NaCl (sat., 50 mL) and NaHCO3 (sat., 20 mL). The organic layer was separated and aqueous fraction was additionally extracted with Et2O (1 × 70 mL). Organic fractions were combined, concentrated, and purified with column chromatography (PE/Et2O 1:3) to give the desired

a

Reaction conditions: compound 3 (2 mmol), Pd/C (1.6 mol %), La(OTf)3 (15 mol %), solvent (5 mL), H2 (20−40 bar), 200 °C, 24 h. b Detected by GC. c0.5 mol % Pd/C used. d6 mmol scale. eCrude rearrangement product is used as starting material; yield is calculated over two steps from 2 (in 10 mmol scale). f12 mmol scale. gIsolated yield; 19 mmol scale, Pd/C (1.6 mol %), La(OTf)3 (5 mol %).

Applying the same conditions on a larger scale resulted in the formation of ethylcyclopentane in a high yield (85%, entry 5). The use of Dy(OTf)3 afforded the product in even higer yield (94%, entry 6). The attempt to reuse the catalytic system (after evaporation of solvents) gave roughly 15% lower yield compared with the first use; however, no special procedures for restoring the catalytic activity were applied. Taking into account the high efficiency achieved in the HDO reaction, an attempt to combine Piancatelli rearrangement and the HDO steps was made. Thus, a crude reaction mixture from Piancatelli rearrangement (10 mol % La(OTf)3, 10 mol % TFA, 90 °C, 24 h) was concentrated, dissolved in an EtOAc or AcOH, and after the addition of another 5 mol % of La(OTf)3 and Pd/C, was subjected to HDO. However, assuming that efficiency of Piancatelli rearrangement step was about 50%, in both cases (Table 2, entries 7 and 8), the yields over two steps were nearly half of expected. An analogous twostep reaction was performed using Dy(OTf)3 (entry 9). After the first step (10 mol % Dy(OTf)3, 10 mol % TFA, 90 °C, 24 h), the reaction mixture was evaporated, and ethyl acetate (15 mL) followed by Pd/C was added. This time the yield was slightly better (entry 8 vs entry 9), but still remained below the expectations. In addition, to isolate the target compound 4 in pure form, HDO of 3 was performed in solvent-free conditions (Table 2, entry 10). As a result, the pure 4 was isolated in 48% yield.



CONCLUSION A three-step route for obtaining ethylcyclopentane from biomass-derived furfural is described. Transformation of furfural to 1-(furan-2-yl)propan-1-ol (2) and HDO of 3 proceeds in excellent yield (96 and 94% correspondingly), thus making the Piancatelli rearrangement a limiting step. Scope and limitations of the route are discussed.



EXPERIMENTAL SECTION General Remarks. Reactions with organometallic reagents were carried out under argon atmosphere in dried glassware. Syringes that were used to transfer anhydrous solvents or 10213

DOI: 10.1021/acsomega.8b00588 ACS Omega 2018, 3, 10211−10215

ACS Omega

Article

compound (628 mg, 50%) as yellowish oil. 1H NMR (400 MHz, CDCl3): δ 7.49 (dd, J = 5.8, 2.2 Hz, 1H), 6.15 (dd, J = 5.8, 1.3 Hz, 1H), 4.64−4.68 (m, 1H), 3.15 (br s, 1H), 2.17 (ddd, J = 8.6, 4.8, 2.4 Hz, 1H) 1.80−1.90 (m, 1H), 1.46−1.57 (m, 1H), 0.99 (t, J = 7.5 Hz, 3H). 13C NMR (100 MHz, CDCl3): δ 208.7, 162.4, 134.3, 76.2, 56.6, 21.5, 11.5. IR (ATRFTIR, neat): ν 3410, 2967, 2937, 2878, 1697, 1462, 1342, 1172, 1092, 1026, 934, 799 cm−1. HRMS (ESI) m/z: [M + Na]+ calcd for C7H10O2Na, 149.0573; found, 149.0566. Procedure B. A round-bottom flask (250 mL) was charged with distilled water (170 mL) and equipped with a reflux condenser and a magnetic stirring bar. Then, the flask was placed on a preheated oil bath (120 °C), water brought to boiling, and then 1-(furan-2-yl)propan-1-ol (2) (1.7 g, 13.5 mmol) was added. The addition was repeated three times more every 20 min (6.8 g, 54 mmol in total). The mixture was stirred with heating for additional 3 h and then allowed to cool down to rt. The resulting heterogenic mixture was transferred to the separation funnel and washed first with PE (4 × 20 mL) and then with the mixture of PE and Et2O (1:1, 2 × 15 mL) to remove byproducts. These organic extracts contained mostly byproducts and therefore were discarded. Subsequent addition of solid NaCl (51 g) to the aqueous phase has triggered an additional separation of yellow organic phase (∼3 mL). The organic phase was separated and remaining completely clear and colorless water solution was extracted with EtOAc (4 × 60 mL). Ethyl acetate fractions were combined, dried with MgSO4, and evaporated to give compound 3 (3.082 g) as a colorless oil. The organic fraction obtained upon the addition of NaCl was evaporated, dissolved in a mixture of PE and Et2O (1:1, 6 mL), and washed with brine (6 × 30 mL) until compound 3 is removed from the organic phase. The remaining organic phase was discarded. Then, the aqueous fractions were combined and extracted again with EtOAc (4 × 20 mL). Ethyl acetate fractions were combined, dried with MgSO4 and evaporated to give an additional portion of compound 3 (400 mg) as colorless oil. Both fractions of compound 3 had the same appearance and purity by NMR and therefore were combined resulting totally 3.482 g (51%) of the target compound. Ethylcyclopentane (4). Typical Procedure. A 5-ethyl-4hydroxycyclopent-2-en-1-one (3) (756 mg, 6 mmol) was dissolved in EtOAc (15 mL). La(OTf)3 (510 mg, 15 mol %) was added, followed by 10 wt % Pd/C (102 mg, 1.6 mol % Pd). The reaction vessel was equipped with a magnetic stirrer and placed into the pressure reactor. The reactor was washed three times with H2, then pressurized to 40 bar H2 (rt), and then heated to 200 °C for 24 h. After this, the reactor was cooled down to rt and then a pressure was released. Obtained liquid was filtered, washed with aqueous solution of NaCl (sat., 10 mL) and dried with MgSO4. The obtained liquid was analyzed by gas chromatography, which confirmed the formation of ethylcyclopentane in 85% yield. Solvent-Free Procedure. A flask was charged with 5-ethyl-4hydroxycyclopent-2-en-1-one (3) (2400 mg, 19 mmol), La(OTf)3 (567 mg, 5 mol %) and 10 wt % Pd/C (340 mg, 1.6 mol % Pd). The reaction vessel was equipped with a magnetic stirrer and placed into the pressure reactor. The reactor was washed three times with H2, then pressurized to 40 bar H2 (rt), and then heated to 200 °C for 24 h. After this, the reactor was cooled down to rt, and then, a pressure was released. All nonvolatile components have remained in the reaction flask, and the colorless liquid that has condensed into

the reactor was collected and dried with MgSO4 to give pure ethylcyclopentane (897 mg) in a 48% yield. 1H NMR (400 MHz, CDCl3): δ 1.63−1.78 (m, 3H), 1.53−1.61 (m, 2H), 1.46−1.52 (m, 2H), 1.31 (quin, J = 7.2 Hz, 2H), 1.03−1.11 (m, 2H), 0.89 (t, J = 7.2 Hz, 3H). 13C NMR (100 MHz, CDCl3): δ 42.2, 32.5, 29.1, 25.4, 13.3.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b00588. Copies of 1H and 13C NMR spectra, GC measurement details, and chromatograms (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (L.V.). ORCID

Aleksei Bredihhin: 0000-0002-0823-2549 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was supported by the Estonian Environmental Investment Centre (project no 11064), State Forest Management Centre (Estonia) and the European Regional Development Fund (MOBTT21). The GC−MS analyses were supported by institutional research funding (IUT20-15) of the Estonian Ministry of Education and Research. We would like to thank Eliise Tammekivi from Institute of Chemistry, Tartu University, Estonia, for performing some of GS−MS measurements.



REFERENCES

(1) Alonso, D. M.; Bond, J. Q.; Dumesic, J. A. Catalytic conversion of biomass to biofuels. Green Chem. 2010, 12, 1493−1513. (2) Hu, L.; Zhao, G.; Hao, W.; Tang, X.; Sun, Y.; Lin, L.; Liu, S. Catalytic conversion of biomass-derived carbohydrates into fuels and chemicals via furanic aldehydes. RSC Adv. 2012, 2, 11184−11206. (3) Climent, M. J.; Corma, A.; Iborra, S. Conversion of biomass platform molecules into fuel additives and liquid hydrocarbon fuels. Green Chem. 2014, 16, 516−547. (4) Bohre, A.; Dutta, S.; Saha, B.; Abu-Omar, M. M. Upgrading Furfurals to Drop-in Biofuels: An Overview. ACS Sustainable Chem. Eng. 2015, 3, 1263−1277. (5) Bredihhin, A.; Vares, L. Conversion of Biomass to Biofuels. Encyclopedia of Physical Organic Chemistry; John Wiley & Sons, Inc., 2016; pp 3653−3697. (6) Hu, L.; Lin, L.; Wu, Z.; Zhou, S.; Liu, S. Recent advances in catalytic transformation of biomass-derived 5-hydroxymethylfurfural into the innovative fuels and chemicals. Renewable Sustainable Energy Rev. 2017, 74, 230−257. (7) Yang, J.; Li, S.; Zhang, L.; Liu, X.; Wang, J.; Pan, X.; Li, N.; Wang, A.; Cong, Y.; Wang, X.; et al. Hydrodeoxygenation of furans over Pd-FeO x /SiO 2 catalyst under atmospheric pressure. Appl. Catal., B 2017, 201, 266−277. (8) Xu, J.; Li, N.; Yang, X.; Li, G.; Wang, A.; Cong, Y.; Wang, X.; Zhang, T. Synthesis of Diesel and Jet Fuel Range Alkanes with Furfural and Angelica Lactone. ACS Catal. 2017, 7, 5880−5886. (9) Deng, Q.; Xu, J.; Han, P.; Pan, L.; Wang, L.; Zhang, X.; Zou, J.-J. Efficient synthesis of high-density aviation biofuel via solvent-free aldol condensation of cyclic ketones and furanic aldehydes. Fuel Process. Technol. 2016, 148, 361−366. 10214

DOI: 10.1021/acsomega.8b00588 ACS Omega 2018, 3, 10211−10215

ACS Omega

Article

(10) Sutton, A. D.; Waldie, F. D.; Wu, R.; Schlaf, M.; “Pete” Silks, L. A.; Gordon, J. C. The hydrodeoxygenation of bioderived furans into alkanes. Nat. Chem. 2013, 5, 428−432. (11) Song, H.-J.; Deng, J.; Cui, M.-S.; Li, X.-L.; Liu, X.-X.; Zhu, R.; Wu, W.-P.; Fu, Y. Alkanes from Bioderived Furans by using Metal Triflates and Palladium-Catalyzed Hydrodeoxygenation of Cyclic Ethers. ChemSusChem 2015, 8, 4250−4255. (12) Dutta, S.; Saha, B. Hydrodeoxygenation of Furylmethane Oxygenates to Jet and Diesel Range Fuels: Probing the Reaction Network with Supported Palladium Catalyst and Hafnium Triflate Promoter. ACS Catal. 2017, 7, 5491−5499. (13) Keskiväli, J.; Wrigstedt, P.; Lagerblom, K.; Repo, T. One-step Pd/C and Eu(OTf) 3 catalyzed hydrodeoxygenation of branched C 11 and C 12 biomass-based furans to the corresponding alkanes. Appl. Catal., A 2017, 534, 40−45. (14) For review see: Nakagawa, Y.; Liu, S.; Tamura, M.; Tomishige, K. Catalytic Total Hydrodeoxygenation of Biomass-Derived Polyfunctionalized Substrates to Alkanes. ChemSusChem 2015, 8, 1114− 1132. (15) Piancatelli, G.; Scettri, A.; Barbadoro, S. A useful preparation of 4-substituted 5-hydroxy-3-oxocyclopentene. Tetrahedron Lett. 1976, 17, 3555−3558. (16) Piancatelli, G.; D’Auria, M.; D’Onofrio, F. Synthesis of 1,4Dicarbonyl Compounds and Cyclopentenones from Furans. Synthesis 1994, 1994, 867−889. (17) Simeonov, S. P.; Nunes, J. P. M.; Guerra, K.; Kurteva, V. B.; Afonso, C. A. M. Synthesis of Chiral Cyclopentenones. Chem. Rev. 2016, 116, 5744−5893. (18) Roche, S. P.; Aitken, D. J. Chemistry of 4-Hydroxy-2cyclopentenone Derivatives. Eur. J. Org. Chem. 2010, 2010, 5339− 5358. (19) Fisher, D.; Palmer, L. I.; Cook, J. E.; Davis, J. E.; Read de Alaniz, J. Efficient synthesis of 4-hydroxycyclopentenones: dysprosium(III) triflate catalyzed Piancatelli rearrangement. Tetrahedron 2014, 70, 4105−4110. (20) de Alaniz, J. R.; Palmer, L. Lewis Acid Catalyzed Rearrangement of Furylcarbinols: The Aza- and Oxa-Piancatelli Cascade Reaction. Synlett 2014, 25, 08−11. (21) Liu, J.; Shen, Q.; Yu, J.; Zhu, M.; Han, J.; Wang, L. A Concise Domino Synthesis of Benzo-1,4-heterocycle Compounds via a Piancatelli/C-N Coupling/Michael Addition Process Promoted by La(OTf)3. Eur. J. Org. Chem. 2012, 2012, 6933−6939. (22) Reiser, O.; Ulbrich, K.; Kreitmeier, P. Microwave- or Microreactor-Assisted Conversion of Furfuryl Alcohols into 4Hydroxy-2-cyclopentenones. Synlett 2010, 2010, 2037−2040. (23) Li, H.; Tong, R.; Sun, J. Catalytic Enantioselective AzaPiancatelli Rearrangement. Angew. Chem., Int. Ed. 2016, 55, 15125− 15128. (24) Cai, Y.; Tang, Y.; Atodiresei, I.; Rueping, M. Catalytic Asymmetric Piancatelli Rearrangement: Brønsted Acid Catalyzed 4π Electrocyclization for the Synthesis of Multisubstituted Cyclopentenones. Angew. Chem., Int. Ed. 2016, 55, 14126−14130. (25) Xu, Z.-L.; Xing, P.; Jiang, B. Intramolecular Aza-Piancatelli Rearrangement of Alkyl- or Arylamines Promoted by PPh3/Diethyl Azodicarboxylate. Org. Lett. 2017, 19, 1028−1031. (26) Lebœuf, D.; Schulz, E.; Gandon, V. Calcium(II)-Catalyzed AzaPiancatelli Reaction. Org. Lett. 2014, 16, 6464−6467. (27) Reddy, B. V. S.; Reddy, Y. V.; Lakshumma, P. S.; Narasimhulu, G.; Yadav, J. S.; Sridhar, B.; Reddy, P. P.; Kunwar, A. C. In(OTf)3Catalyzed Tandem Aza-Piancatelli Rearrangement/Michael Reaction for the Synthesis of 3,4-Dihydro-2H-Benzo[b][1,4]Thiazine and Oxazine Derivatives. RSC Adv. 2012, 2, 10661−10666. (28) Veits, G. K.; Wenz, D. R.; Read de Alaniz, J. Versatile Method for the Synthesis of 4-Aminocyclopentenones: Dysprosium(III) Triflate Catalyzed Aza-Piancatelli Rearrangement. Angew. Chem., Int. Ed. 2010, 49, 9484−9487. (29) Hronec, M.; Fulajtárová, K.; Vávra, I.; Soták, T.; Dobročka, E.; Mičušík, M. Carbon supported Pd-Cu catalysts for highly selective

rearrangement of furfural to cyclopentanone. Appl. Catal., B 2016, 181, 210−219. (30) Yang, Y.; Du, Z.; Huang, Y.; Lu, F.; Wang, F.; Gao, J.; Xu, J. Conversion of furfural into cyclopentanone over Ni-Cu bimetallic catalysts. Green Chem. 2013, 15, 1932−1940. (31) Deng, Q.; Nie, G.; Pan, L.; Zou, J.-J.; Zhang, X.; Wang, L. Highly selective self-condensation of cyclic ketones using MOFencapsulating phosphotungstic acid for renewable high-density fuel. Green Chem. 2015, 17, 4473−4481. (32) Sheng, X.; Li, N.; Li, G.; Wang, W.; Yang, J.; Cong, Y.; Wang, A.; Wang, X.; Zhang, T. Synthesis of high density aviation fuel with cyclopentanol derived from lignocellulose. Sci. Rep. 2015, 5, 9565. (33) Yang, J.; Li, N.; Li, G.; Wang, W.; Wang, A.; Wang, X.; Cong, Y.; Zhang, T. Synthesis of renewable high-density fuels using cyclopentanone derived from lignocellulose. Chem. Commun. 2014, 50, 2572−2574. (34) Ohyama, J.; Kanao, R.; Ohira, Y.; Satsuma, A. The effect of heterogeneous acid-base catalysis on conversion of 5-hydroxymethylfurfural into a cyclopentanone derivative. Green Chem. 2016, 18, 676− 680. (35) Ohyama, J.; Kanao, R.; Esaki, A.; Satsuma, A. Conversion of 5hydroxymethylfurfural to a cyclopentanone derivative by ring rearrangement over supported Au nanoparticles. Chem. Commun. 2014, 50, 5633−5636. (36) Ohyama, J.; Ohira, Y.; Satsuma, A. Hydrogenative ringrearrangement of biomass derived 5-(hydroxymethyl)furfural to 3(hydroxymethyl)cyclopentanol using combination catalyst systems of Pt/SiO2 and lanthanoid oxides. Catal. Sci. Technol. 2017, 7, 2947− 2953. (37) Ren, D.; Song, Z.; Li, L.; Liu, Y.; Jin, F.; Huo, Z. Production of 2,5-hexanedione and 3-methyl-2-cyclopenten-1-one from 5-hydroxymethylfurfural. Green Chem. 2016, 18, 3075−3081. (38) Sacia, E. R.; Deaner, M. H.; Louie, Y. L.; Bell, A. T. Synthesis of biomass-derived methylcyclopentane as a gasoline additive via aldol condensation/hydrodeoxygenation of 2,5-hexanedione. Green Chem. 2015, 17, 2393−2397. (39) Op de Beeck, B.; Dusselier, M.; Geboers, J.; Holsbeek, J.; Morré, E.; Oswald, S.; Giebeler, L.; Sels, B. F. Direct catalytic conversion of cellulose to liquid straight-chain alkanes. Energy Environ. Sci. 2015, 8, 230−240. (40) Weng, Y.; Qiu, S.; Ma, L.; Liu, Q.; Ding, M.; Zhang, Q.; Zhang, Q.; Wang, T. Jet-Fuel Range Hydrocarbons from Biomass-Derived Sorbitol over Ni-HZSM-5/SBA-15 Catalyst. Catalysts 2015, 5, 2147− 2160. (41) Wang, Y.; He, T.; Liu, K.; Wu, J.; Fang, Y. From biomass to advanced bio-fuel by catalytic pyrolysis/hydro-processing: Hydrodeoxygenation of bio-oil derived from biomass catalytic pyrolysis. Bioresour. Technol. 2012, 108, 280−284. (42) Gjervan, T.; Prestvik, R.; Holmen, A. Catalytic Reforming. Basic Principles in Applied Catalysis; Baerns, M., Ed.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2004; pp 125−158. (43) Camp, J. E.; Craig, D. Reactivity of Dearomatised Furans Synthesised via the Decarboxylative Claisen Rearrangement. Tetrahedron Lett. 2009, 50, 3503−3508. (44) Frelek, J.; Karchier, M.; Madej, D.; Michalak, K.; Rózȧ ński, P.; Wicha, J. Chemoenzymatic Approach to Optically Active 4-Hydroxy5-alkylcyclopent-2-en-1-one Derivatives: An Application of a Combined Circular Dichroism Spectroscopy and DFT Calculations to Assignment of Absolute Configuration. Chirality 2014, 26, 300− 306. (45) Michalak, K.; Wicha, J. A convenient preparation of (S)-(−)-4hydroxy-2-methylcyclopent-2-en-1-one and its application as a chiral synthetic equivalent of 2-methylcyclopent-2-en-1-one in the terpenoid synthesis. Tetrahedron 2014, 70, 5073−5081.

10215

DOI: 10.1021/acsomega.8b00588 ACS Omega 2018, 3, 10211−10215