Next Article in Journal
Custom G4 Microarrays Reveal Selective G-Quadruplex Recognition of Small Molecule BMVC: A Large-Scale Assessment of Ligand Binding Selectivity
Previous Article in Journal
Variability of Major Phenyletanes and Phenylpropanoids in 16-Year-Old Rhodiola rosea L. Clones in Norway
Previous Article in Special Issue
Recent Progress of Cu-Catalyzed Azide-Alkyne Cycloaddition Reactions (CuAAC) in Sustainable Solvents: Glycerol, Deep Eutectic Solvents, and Aqueous Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

HFIP-Promoted Synthesis of Substituted Tetrahydrofurans by Reaction of Epoxides with Electron-Rich Alkenes

Departamento de Química Orgánica and Instituto de Síntesis Orgánica (ISO), Facultad de Ciencias, Universidad de Alicante. Apdo. 99, E-03080 Alicante, Spain
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(15), 3464; https://doi.org/10.3390/molecules25153464
Submission received: 3 July 2020 / Revised: 26 July 2020 / Accepted: 28 July 2020 / Published: 30 July 2020
(This article belongs to the Special Issue Solvent-Dependent Organic Transformations)

Abstract

:
In the present work, the employment of fluorinated alcohols, specifically 1,1,1,3,3,3-hexafluoroisopropanol (HFIP), as solvent and promoter of the catalyst-free synthesis of substituted tetrahydrofuranes through the addition of electron-rich alkenes to epoxydes is described. The unique properties of this fluorinated alcohol, which is very different from their non-fluorinated analogs, allows carrying out this new straightforward protocol under smooth reaction conditions affording the corresponding adducts in moderate yields in the majority of cases. Remarkably, this methodology has allowed the synthesis of new tetrahydrofuran-based spiro compounds as well as tetrahydrofurobenzofuran derivatives. The scope and limitations of the process are also discussed. Mechanistic studies were also performed pointing towards a purely ionic or a SN2-type process depending on the nucleophilicity of the alkene employed.

1. Introduction

The substituted tetrahydrofuran structure is present in a wide variety of bioactive natural compounds and has gained considerable interest in pharmaceutical research. In general, natural compounds containing tetrahydrofuran ring derivatives have been found in different classes of terrestrial and marine organisms [1,2,3]. One of these representative examples are Caloxylanes A and B, both isolated from the Caribbean marine sponge Calyx podatypa [4,5], or Corsifuran A, isolated from the heartwood of the tree Thespesia populnea [6]. Other examples are the lignans Fragransin C1 (among other compounds from the Fragransin family) [1,7] and Conocarpan [1,8], both having demonstrated to exhibit biological activity (Figure 1).
It is not surprising then that a considerable amount of strategies have been already described in order to gain access to these interesting molecules. Among them, probably the most straightforward method is based on the reaction between alkenes and epoxides, which are commercially available and highly abundant in bulk. This perfect atom-economy route provides direct access to different substituted tetrahydrofurans allowing a wide range of substitution patterns on the structure. However, to the best of our knowledge, only a limited number of publications following this protocol have been reported, being those mainly radical [9] or metal-catalyzed processes (Scheme 1) [10,11,12,13].
In the last years, our research group has become interested in the use of fluorinated alcohols as solvents and promoters of organic reactions [14,15]. The unique chemical and physical properties that fluoroalkyl alcohols have in comparison with their non-fluorinated analogues, such as their high hydrogen bond donor ability, high polarity and ionizing power, and low nucleophilicity values together with the slightly acidic character, make them perfect candidates as promoters of reactions involving ionic processes [16,17,18,19,20,21,22]. On the other hand, fluorinated alcohols have already proven to be efficient promoters in the ring-opening reaction of epoxides with different nucleophiles [23,24,25,26].
With all these precedents in mind, we envisioned a new strategy based on the use of fluorinated alcohols as solvents and reaction promoters in a metal and radical-free ring-opening reaction of epoxides with different electron-rich alkenes as nucleophiles in order to obtain the corresponding substituted tetrahydrofurans in an efficient, cost-effective, and environmentally friendly chemical manner. The results of this investigation are herein described.

2. Results and Discussion

Firstly, the reaction between styrene oxide (1a) and α-methylstyrene (1b) was selected as a model in order to obtain the optimal reaction conditions. Different solvents were selected to evaluate their performance as promoters at 45 °C (Table 1, entries 1–4). When water and 2-propanol, which possesses quite high polarity and hydrogen bond ability, were used, the reaction produced the diol 4 as major product and failed (Table 1, entries 1 and 2). Next, 1,1,1,3,3,3-hexafluoroisopropanol (HFIP) and 2,2,2-trifluoroethanol (TFE), as readily available and inexpensive fluorinated alcohols, were tested. As observed in the table, whereas the reaction with HFIP afforded the desired product in high conversion, TFE barely produced tetrahydrofuran 3aa (Table 1, entries 3 and 4, respectively). This sharp contrast in the performance of both fluorinated alcohols in this transformation can be explained by their different properties. Thus, HFIP has higher acidity (pKa(TFE) = 12.37, pKa(HFIP) = 9.30); higher hydrogen bond ability (αTFE = 1.51, αHFIP = 1.96), which can facilitate the activation of epoxide ring; and much lower nucleophilicity (NTFE = −2.78, NHFIP = −4.23) [16,17,18,19,20,21,22]. This last parameter would explain the obtention of fluoroalkyl ether 5 as major product when TFE was essayed. On the contrary, the corresponding fluorinated ether 6, derived from HFIP (along with phenylacetaldehyde and acetophenone) was obtained only as by-product. The absence of any solvent was also checked and, as was expected, the reaction did not take place (Table 1, entries 5). Then, efforts to improve the conversion of 3aa by using a series of HFIP/CH2Cl2 [16,17,18,19,20,21,22] mixtures were implemented (Table 1, entries 6–8), but turned out to be unsuccessful in all the cases. Lowering the reaction temperature to 25 °C also resulted in a drop in the conversion towards the desired product (Table 1, entry 9). Other changes in reaction stoichiometry were also essayed but did not produce any amelioration. After the search for the best conditions, those described in entry 3, involving the use of HFIP at 45 °C, were selected as optimal, realizing that reaction was complete in less than 10 h.
With these conditions in hand, the scope of the reaction was next investigated. Different electron-rich alkenes were employed as nucleophiles for the ring opening reaction of styrene oxide (Scheme 2). It is important to mention that in the majority of cases, the ring-opening reaction was highly regioselective, but when a mixture of diastereoisomers was obtained, low diastereoselective ratios were observed. First, a selection of substituted styrenes was chosen. As mentioned above, α-methylstyrene (2a) produced the corresponding tetrahydrofuran 3aa in moderate isolated yield. Better results were observed when a more electron-rich alkene, 2b, was employed; reaching up to 67% yield for 3ab. The more sterically crowded 1,1-diphenylethylene (2c) gave the corresponding product in only modest yield. In this case, some amount (25%) of the other regioisomer was also obtained, probably due to the mentioned steric hindrance. Next, styrene was essayed obtaining the corresponding caloxylane 3ad (as 65:35 mixture of diastereosiomers) in 39% yield. Similar results were obtained when 4-chlorostyrene was employed. Surprisingly, the more electron-rich alkene, 4-methoxystyrene (2f), gave rise to the corresponding product in low yields. At this point, it is worth mentioning that in the majority of the styrenes employed the presence of dimers or trimers of the styrene were detected by GC-MS, thus lowering the yield of the process. Methylenecyclohexane (2g) was next tested, obtaining spiro-compound 3ag in modest yield. Stilbenes were also submitted to the reaction conditions but failed and only a low conversion was observed when the cis-isomer was employed. Trisubstituted alkenes such as 1-phenylcyclohexene (2i) were also taken into account, obtaining the interesting octahydrobenzofuran derivative 3ai in 34% yield. Finally, benzocondensed alkenes were also essayed. Thus, whereas indene produced the corresponding product 3aj in modest yield, the reaction with 1,2-dihydronaphthalene barely worked. Finally, when benzofuran (2l) was employed as alkene, the corresponding tetrahydrofuro[3,2-b]benzofuran derivative 3al, arising from the attack of benzofuran through its 2-position onto the epoxide, obtained with moderate yield and quite good diastereoselectivity. It is important to remark that in all the cases ether 6 along with the products coming from the Meinwald rearrangement [27] of the epoxide (benzaldehyde and acetophenone, the first one with higher proportion) were obtained as by-products.
In order to further expand the scope of the reaction, other epoxides were tested with those alkenes that provided the best results. First, α-methylstyrene oxide (1b) was evaluated. Good yields were achieved when 2a and 2b were the alkenes employed. However, modest yields were only achieved when ethylenecyclohexane (2g) and benzofuran (2l) were used. Next, when 1-phenylcyclohexene oxide (1c) was the substrate submitted to the reaction with the same alkenes, modest yields were obtained for adducts 3ca and 3cb. Unfortunately, the reaction with 2g did not work. Although, in a modest 38% yield, benzofuran (2l) rendered the interesting tetracyclic compound 3cl. Finally, commercial available ethyl 3-methyl-3-phenylglycidate (1d) was also tested. It is remarkable that, contrary to the normal trend observed concerning the low diastereoselectivity achieved in previous cases, moderate to good diastereoselectivities were achieved when 1d was the substrate employed. Modest yield was achieved when α-methylstyrene (2a) was employed, rendering the densely substituted 3da in 43% yield. However, the more electron-rich alkene 2b did not produce satisfactory results, being the dimerization and trimerization product of the alkene the major products observed by GC-MS. To our surprise, alkene 2c gave rise to the corresponding tetrahydrofuran 3dc in 60% yield and a 90:10 diastereomeric ratio. Encouraged by this result, styrene was also employed obtaining 3dd in modest yield. Unfortunately, the reaction with alkenes 2g and 2j turned out to be unsuccessful and low conversions towards the desired products were obtained. Finally, benzofuran (2l) rendered the corresponding tricyclic compound in a modest 44% yield. It is worth mentioning that other epoxides such as cyclohexene oxide, 1-octene oxide, indene oxide, and cis- and trans-stilbene oxide were also submitted to the reaction with the alkenes depicted in Table 2; however, to our regret the reaction failed.
At this point, as epoxides are synthesized from alkenes, and fluorinated alcohols have proven to be efficient mediators in the oxidation of alkenes using H2O2 [28], we decided to explore the possibility of performing the HFIP-promoted alkene oxidation/ring opening of epoxides in a one-pot reaction (Scheme 3). For such purpose, an excess of α-methylstyrene (2a) was treated with 1 equivalent of H2O2 (30%) for 24 h. After this time among a myriad of products detected by GC-MS coming from the ring opening of the epoxide with H2O or HFIP, 2-phenylpropanaldehyde from Meinwald rearrangement and acetophenone from oxidative cleavage of the alkene, tetrahydrofuran 3aa was observed in 27% conv. Although the product was obtained in low amount, it can be seen as a proof of concept that substituted tetrahydrofurans can be easily obtained from readily available materials as styrenes.
Concerning the reaction mechanism, three possible scenarios were taken into account (Scheme 4). In route 1, direct nucleophilic attack of the alkene onto activated epoxide would occur rendering intermediate A, which cyclizes to afford the corresponding tetrahydrofuran. In route 2, intermediate A is obtained as consequence of a double nucleophilic attack, the first one carried out by the more abundant HFIP followed by a nucleophilic substitution. In the route 3, carbocationic intermediate C is formed, which could be stabilized by the formation of ionic pair or by other electrostatic interactions with HFIP. Route 2 was soon discarded due to the fact that intermediate B has been observed in the reaction and if this would have been the operating route, longer reaction times would render higher conversions, which did not happen. Nevertheless, ether 6 was synthesized by reacting 1a with HFIP by 8 h at room temperature, and after a quick purification, was allowed to react with α-methylstyrene (2a) for 24 h under the optimized conditions. After this time, no reaction was observed. In order to find out whether route 1 or 3 was operating in the process, we decided to carry out the reaction using enantiopure (R)-styrene oxide and α-methylstyrene (2a) and styrene (2d) as alkenes (Scheme 5). Thus, if route 1 is the one taking place, the configuration of the stereocenter will be somehow preserved. As depicted in Scheme 4, corroborated by chiral HPLC analysis (see Supplementary Materials for further details), when styrene (2d) was the nucleophile, the stereochemistry of the chiral center was lost giving a racemic mixture in each diastereoisomer of caloxylane (3ad). However, the better nucleophile α-methylstyrene (2a) gave rise to a mixture of diastereoisomers both presenting a loss of enantiopurity in the chiral centre (46% ee and 49% ee, respectively), which was determined by chiral HPLC analysis (see Supplementary Materials for further details). Therefore, these experimental evidences point that the mechanism of the reaction seemingly is highly dependent on the nucleophilicity of the alkene employed. Thus, whereas styrene (2d) apparently follows a purely ionic route (SN1-type mechanism (route 3, Scheme 4)) in the α-methylstyrene (2a) case, predominantly a SN2-type pathway (route 1, Scheme 3) is operating.

3. Materials and Methods

All reagents and solvents were obtained commercially and used without further purification. Substrates that were not commercially available were synthesized according to known literature procedures. NMR spectra were performed on a Bruker AV-300 or Bruker AV-400 (Bruker Corporation, Billerica, MA, USA) using CDCl3 as solvent and TMS as internal standard unless otherwise stated. Conversions and low-resolution mass spectra (MS) of the tetrahydrofurans 3 were recorded in the electron impact mode (EI, 70 eV, He as carrier phase) using an Agilent GC/MS 5973 Network Mass Selective Detector spectrometer apparatus equipped with a HP-5MS column (30 m × 0.25 mm) (Agilent technologies, Bilbao, Spain) and giving fragment ions in m/z with relative intensities (%) in parentheses. High-resolution mass spectra (HRMS) were obtained on an Agilent 7200 Quadrupole-Time of Flight apparatus (Q-TOF) (Agilent Technologies), with the ionization employed being electron impact (EI). Chiral HPLC analysis was performed in an Agilent 1100 Series HPLC equipped with a G1315B diode array detector and a Quat Pump G1311A (Agilent Technologies) equipped with the corresponding Daicel chiral column. Analytical TLC was performed on Merck silica gel plates and the spots visualized with UV light at 254 nm (Merck millipore, Darmstadt, Germany). Flash chromatography employed Merck silica gel 60 (0.040–0.063 mm). Silica gel 60 F254 containing gypsum was employed for preparative layer chromatography (Merck millipore).

General Procedure for the HFIP-Promoted Synthesis of Substituted Tetrahydrofurans

In a capped tube, onto a mixture of the corresponding epoxide (0.15 mmol) and alkene (0.25 mmol), HFIP (150 µL) was added in one portion. The reaction was then stirred at 45 °C for 6–15 h, until the reaction was judged to be completed (no starting epoxide remaining) by GC-MS. After this time, solvent was evaporated and the crude material was directly purified by flash chromatography or preparative TLC.
2-Methyl-2,4-diphenyltetrahydrofuran (3aa) [11]: yellow oil; purification by flash chromatography (hexane/EtOAc), 54% yield; (cis/trans) = 55:45; 1H NMR (300 MHz, CDCl3): cis isomer: δH = 7.52–7.45 (m, 4H), 7.43–7.28 (m, 10H), 7.27–7.16 (m, 6H), 4.42 (t, J = 7.6 Hz, 1H), 3.84 (dd, J = 10.0, 8.2 Hz, 1H), 3.70 (tt, J = 10.3, 7.7 Hz, 1H), 2.41–2.15 (m, 2H), 1.63 (s, 3H) ppm; further signals for the trans isomer: δH = 4.35 (t, J = 8.4 Hz, 1H), 4.00 (t, J = 8.7 Hz, 1H), 3.31 (ddd, J = 15.9, 11.3, 8.6 Hz, 1H), 2.81–2.60 (m, 2H), 1.68 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): cis isomer: δC = 148.9, 140.8, 128.5, 128.3, 127.4, 126.6, 126.5, 124.5, 74.4, 47.9, 45.8, 30.6, ppm; further signals for trans isomer: 147.6, 141.6, 128.5, 128.2, 127.3, 126.6, 126.4, 124.7, 73.9, 48.3, 44.6, 30.2 ppm; MS (EI): m/z 238 (M+, 0.31%), 224 (19), 223 (100), 193 (12), 117 (27), 115 (18), 105 (90), 91 (16), 77 (18). Chiral HPLC analysis: Chiralpak IA column, Hexane/iPrOH 99:1, flow rate = 0.2 mL/min, λ = 210 nm, retention times: = 25.5 and 26.5 min. (major diastereosiomer) and 27.4 and 28.4 min. (minor diastereoisomer).
2-(4-Methoxyphenyl)-2-methyl-4-phenyl tetrahydrofuran (3ab): orange oil; purification by flash chromatography (hexane/EtOAc), 67% yield; (cis/trans) = 55:45; 1H NMR (300 MHz, CDCl3): cis isomer: δH = 7.43–7.37 (m, 3H), 7.33–7.28 (m, 3H), 7.27–7.16 (m, 6H), 6.95–6.93 (m, 2H), 6.92–6.90 (m, 2H), 4.40 (t, J = 7.9 Hz, 1H), 3.84 (s, 3H), 3.82 (dd, J = 3.5, 1.9 Hz, 1H), 3.76–3.64 (m, 1H), 2.63 (dd, J = 12.4, 8.0 Hz, 1H), 2.32 (dd, J = 12.3, 10.5 Hz, 1H), 2.24–2.17 (m, 1H), 1.61 (s, 3H) ppm; further signals for the trans isomer: δH = 4.33 (t, J = 8.4 Hz, 1H), 3.98 (t, J = 8.6 Hz, 1H), 3.85 (s, 3H), 3.39–3.25 (m, 1H), 2.71 (dd, J = 12.1, 7.1 Hz, 1H), 2.32 (dd, J = 12.3, 10.5 Hz, 1H), 1.66 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): mixture of isomers, δC = 158.3, 158.2, 141.7, 141.1, 141.0, 139.7, 128.5, 128.5, 127.4, 127.3, 126.6, 126.6, 125.8, 125.7, 113.6, 113.5, 85.2, 84.7, 74.4, 73.9, 55.3, 55.2, 48.3, 48.1, 45.8, 44.6, 30.6, 30.3 ppm; MS (EI): m/z 268 (M+, 6%), 254 (17), 253 (93), 135 (100), 117 (14), 91 (11); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C18H20O2 268.1463, found 268.1463.
2,2,4-Triphenyltetrahydrofuran (3ac): yellow oil; purification by flash chromatography (hexane/EtOAc), 38% estimated yield (not purely isolated); 1H NMR (300 MHz, CDCl3): δH = 7.54–7.49 (m, 4H), 7.37–7.34 (m, 4H), 7.34–7.30 (m, 4H), 7.27–7.24 (m, J = 1.6 Hz, 2H), 7.24–7.22 (m, 1H), 4.48 (t, J = 8.4 Hz, 1H), 4.07 (t, J = 8.7 Hz, 1H), 3.61–3.43 (m, J = 16.0, 11.0, 5.2 Hz, 1H), 3.23 (dd, J = 12.3, 7.1 Hz, 1H) ppm; MS (EI): m/z 300 (M+, 55%), 270 (15), 224 (71), 223 (100), 192 (21), 191 (13), 179 (13), 178 (15), 165 (21), 118 (34), 117 (42), 115 (12), 105 (96), 91 (14), 77 (27); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C22H20O 300.1514, found 300.1509.
2,4-Diphenyltetrahydrofuran (Caloxylane A and B) (3ad) [11]: yellow oil; purification by flash chromatography (hexane/EtOAc), 39% yield; (cis/trans) = 65:35; 1H NMR (300 MHz, CDCl3): cis isomer: δH = 7.49–7.29 (m, 20H), 5.11 (dd, J = 10.2, 5.7 Hz, 1H), 4.39 (t, J = 8.2 Hz, 1H), 4.05 (t, J = 8.5 Hz, 1H), 3.74–3.63 (m, 1H), 2.85–2.72 (m, 1H), 2.05 (q, J = 10.5, 1.9 Hz, 1H) ppm; further signals for the trans isomer: δH = 5.26 (dd, J = 7.7, 5.8 Hz, 1H), 4.50 (t, J = 8.5, 7.4 Hz, 1H), 3.98 (t, J = 8.2 Hz, 1H), 3.63–3.51 (m, 1H), 2.57–2.45 (m, 1H), 2.36 (q, J = 12.5, 8.3, 5.8 Hz, 1H) ppm; 13C NMR (75 MHz, CDCl3): cis isomer: δC = 142.6, 141.7, 128.6, 128.4, 127.4, 127.2, 125.7, 81.8, 75.1, 46.0, 43.7, pm; further signals for the trans isomer: δC = 143.6, 142.0, 128.6, 128.3, 127.3, 127.1, 125.5, 80.6, 75.1, 44.4, 42.7 ppm; MS (EI): m/z 224 (M+, 34%), 195 (14), 194 (93), 193 (100), 179 (58), 178 (89), 165 (13), 146 (27), 133 (34), 120 (27), 117 (90), 115 (57), 105 (45), 91 (48), 77 (30). Chiral HPLC analysis: Chiralcel OD-H column, Hexane/iPrOH 99:1, flow rate = 0.7 mL/min, λ = 210 nm, retention times: = 23.1 and 23.2 min. (major diastereosiomer) and 28.0 and 35.4 min. (minor diastereoisomer).
2-(4-Chlorophenyl)-4-phenyltetrahydrofuran (3ae) [11]: yellow oil; purification by flash chromatography (hexane/EtOAc), 36% yield; (cis/trans) = 60:40; 1H NMR (400 MHz, CDCl3): cis isomer: δH = 7.40–7.30 (m, 9H), 5.07 (dd, J = 10.1, 5.8 Hz, 1H), 4.38 (t, J = 8.2 Hz, 1H), 4.03 (t, J = 8.5 Hz, 1H), 3.78–3.61 (m, 1H), 2.86–2.72 (m, 1H), 1.98 (dd, J = 12.4, 10.4 Hz, 1H) ppm; further signals for the trans isomer: δH = 5.22 (dd, J = 7.7, 5.9 Hz, 1H), 4.47 (dd, J = 8.4, 7.5 Hz, 1H), 3.97 (t, J = 8.2 Hz, 1H), 3.54 (t, J = 7.6 Hz, 1H), 2.50 (dt, J = 12.6, 7.7 Hz, 1H), 2.35–2.24 (m, 1H) ppm; 13C NMR (101 MHz, CDCl3): cis isomer: δC = 141.5, 141.2, 133.0, 129.9, 128.8, 128.6, 128.3, 128.2, 127.3, 127.1, 126.7, 81.1, 75.1, 45.9, 43.8 ppm; further signals for the trans isomer: δC = 142.1, 142.7, 132.8, 129.8, 128.7, 128.5, 128.3, 128.1, 127.2, 126.9, 126.7, 126.4, 79.9, 72.7, 44.3, 42.7 ppm; MS (EI): m/z 258 (M+, 18%), 228 (25), 193 (100), 180 (15), 178 (22), 167 (22), 154 (16), 139 (27), 117 (64), 115 (54), 104 (17), 91 (27), 77 (13).
3-Phenyl-1-oxaspiro[4.5]decane (3ag): yellow solid, purification by flash chromatography (hexane/EtOAc), 45% yield; 1H NMR (400 MHz, CDCl3): δH = 7.36–7.31 (m, 3H), 7.27–7.23 (m, 2H), 4.23 (t, J = 8.0 Hz, 1H), 3.80 (t, J = 17.6 Hz, 1H), 3.51 (tt, J = 17.6, 8.8 Hz, 1H), 2.30 (dd, J = 12.4, 8.2 Hz, 1H), 1.78 (dd, J = 12.4, 10.5 Hz, 1H), 1.74–1.48 (m, 10H) ppm; 13C NMR (126 MHz, CDCl3): δC = 141.9, 128.6, 128.5, 128.1, 127.3, 126.5, 83.3, 72.9, 45.0, 38.3, 37.3, 25.61, 23.8, 23.8 ppm; MS (EI): m/z 216 (M+, 55%), 174 (25), 173 (100), 160 (40), 118 (18), 117 (28), 104 (41), 91 (26), 55 (73); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C15H20O 216.1514, found 216.1514.
3,7a-Diphenyloctahydrobenzofuran (3ai): orange oil; purification by flash chromatography (hexane/EtOAc); 34% yield; diastereomeric ratio = 55:45; 1H NMR (300 MHz, CDCl3): major isomer: δH = 7.58 (dd, J = 8.5, 1.1 Hz, 2H), 7.54–7.48 (m, 2H), 7.44–7.35 (m, 6H), 7.34–7.28 (m, 4H), 7.23 (m, 6H), 7.16–7.08 (m, 2H), 4.40 (m, 2H), 4.00–3.91 (m, 1H), 3.59 (m, 1H), 2.67 (dd, J = 12.0, 5.5 Hz, 1H), 2.13–1.85 (m, 4H), 1.83–1.52 (m, 12H) ppm; furher signals for the minor isomer: δH = 4.27 (t, J = 8.6 Hz, 1H), 3.42 (td, J = 9.6, 5.5 Hz, 1H), 2.57 (dt, J = 11.4, 5.7 Hz, 1H) ppm; 13C NMR (75 MHz, CDCl3): major isomer: δC = 148.6, 141.2, 128.6, 128.2, 128.1, 126.9, 126.4, 125.9, 84.8, 72.6, 51.2, 46.6, 35.8, 24.2, 21.9, 20.0 ppm; further signals for minor isomer: δC = 146.5, 138.2, 128.3, 128.1, 127.9, 126.7, 126.2, 124.5, 86.8, 67.3, 49.4, 46.4, 38.3, 24.6, 22.1, 20.0 ppm; MS (EI): m/z 278 (M+, 85%), 236 (19), 235 (100), 221 (18), 115 (14), 105 (67), 91 (25), 77 (18); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C20H22O 278.1671, found 278.1667.
3-Phenyl-3, 3a, 4, 8b-tetrahydro-2H-indeno [1,2-b]furan (3aj) [11]: yellow oil; purification by flash chromatography (hexane/EtOAc), 40% yield; (cis/trans) = 60:40; 1H NMR (300 MHz, CDCl3): cis isomer: δH = 7.53–7.47 (m, 2H), 7.39–7.29 (m, 8H), 5.73 (d, J = 7.2 Hz, 1H), 4.13 (dd, J = 8.6, 6.7 Hz, 1H), 3.92 (dd, J = 8.6, 7.7 Hz, 1H), 3.26–3.16 (m, 2H), 3.12–3.03 (m, 1H), 3.01–2.91 (m, 1H) ppm; further signals for the trans isomer: δH = 7.51–7.46 (m, 1H), 7.36–7.24 (m, 6H), 7.14–7.09 (m, 2H), 5.69 (d, J = 6.8 Hz, 1H), 4.27–4.21 (m, 1H), 3.82–3.73 (m, 2H), 3.59–3.47 (m, 1H), 2.83 (dd, J = 17.4, 9.3 Hz, 1H), 2.60 (dd, J = 17.4, 4.8 Hz, 1H) ppm; 13C NMR (75 MHz, CDCl3): mixture of isomers, δC = 142.5, 141.9, 141.7, 141.6, 128.8, 128.7, 128.4, 128.4, 127.5, 127.2, 127.0, 126.7, 126.5, 125.6, 125.2, 124.4, 87.9, 87.8, 74.6, 68.2, 53.3, 50.1, 48.5, 45.5, 38.7, 36.6 ppm; MS (EI): m/z 236 (M+, 34%), 207 (17), 206 (100), 205 (27), 128 (20), 115 (27), 91 (86).
3-Phenyl-2,3,3a,8b-tetrahydrofuro[3,2-b]benzofuran (3al): yellow oil; purification by flash chromatography (hexane/EtOAc), 40% estimated yield, mixture of isomers (not purely isolated); 1H NMR (400 MHz, CDCl3): δH = 7.50–7.46 (m, 1H), 7.40–7.38 (m, 3H), 7.35–7.30 (m, 3H), 6.99 (td, J = 7.4, 0.9 Hz, 1H), 6.88 (d, J = 8.3 Hz, 1H), 5.81 (d, J = 6.0 Hz, 1H), 5.19 (dd, J = 6.0, 1.2 Hz, 1H), 4.08 (dd, J = 9.1, 2.3 Hz, 1H), 3.96 (dd, J = 9.1, 5.5 Hz, 1H), 3.67 (d, J = 5.5 Hz, 1H) ppm; MS (EI): m/z 238 (M+, 34%), 220 (66), 219 (43), 208 (45), 207 (100), 191 (15), 189 (17), 178 (19), 165 (13), 131 (24), 117 (12); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C16H14O 238.0994, found 238.0990.
2,4–Dimethyl–2,4–diphenyltetrahydrofuran (3ba) [29]: yellow solid; purification by flash chromatography (hexane/EtOAc), 62% yield; (cis/trans) = 50:50; the cis isomer is highlighted in bold; 1H NMR (300 MHz, CDCl3): δH = 7.54–7.49 (m, 2H), 7.45–7.39 (m, 2H), 7.39–7.30 (m, 8H), 7.26–7.21 (m, 4H), 7.20–7.10 (m, 4H), 4.30 (d, J = 8.6 Hz, 1H), 4.10 (dd, J = 8.4, 1.0 Hz, 1H), 4.06 (d, J = 8.5 Hz, 1H), 3.96 (dd, J = 8.4, 0.6 Hz, 1H), 2.72 (d, J = 12.6 Hz, 1H), 2.58 (s, 2H), 2.47 (dd, J = 12.6, 1.1 Hz, 1H), 1.70 (s, 3H), 1.59 (s, 6H), 1.50 (s, 3H) ppm; MS (EI): m/z 252(M+, 0.08%), 237 (100), 207 (12), 129 (13), 117 (29), 105 (97), 91 (14), 77 (15).
2–(4–Methoxyphenyl)–2,4–dimethyl-4-phenyl tetrahydrofuran (3bb): white solid; purification by flash chromatography (hexane/EtOAc), 58% yield; (cis/trans) = 45:55; the cis isomer is highlighted in bold; 1H NMR (300 MHz, CDCl3): δH = 7.36–7.30 (m, 3H), 7.27–7.22 (m, 2H), 7.16–7.09 (m, 2H), 6.87–6.82 (m, 2H), 4.28 (d, J = 8.6 Hz, 1H), 4.07 (d, J = 9.1 Hz, 1H), 4.03 (s, 1H), 3.95 (d, J = 8.4 Hz, 1H), 3.84 (s, 3H), 3.80 (s, 3H), 2.70 (d, J = 12.6 Hz, 1H), 2.55 (d, J = 2.2 Hz, 1H), 2.42 (d, J = 12.5 Hz, 1H), 1.67 (s, 3H), 1.58 (s, 3H), 1.48 (s, 3H), 1.23 (s, 3H) ppm; Only major isomer is given, 13C NMR (101 MHz, CDCl3): δH = 157.9, 147.7, 141.6, 128.3, 126.0, 125.9, 125.6, 113.5, 84.6, 77.8, 55.3, 53.9, 48.9, 32.8, 29.9 ppm; MS (EI): m/z 282(M+, 9%), 268 (16), 267 (83), 135 (100), 117 (13); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C19H22O2 282.1620, found 282.1620.
3-Methyl-3-phenyl-1-oxaspiro[4,5]decane (3bg): colourless oil; purification by flash chromatography (hexane/EtOAc), 30% estimated yield (not purely isolated); 1H NMR (300 MHz, CDCl3): δH = 7.36–7.30 (m, 4H), 7.26–7.23 (m, 1H), 4.05 (d, J = 8.7 Hz, 1H), 3.94 (d, J = 8.5 Hz, 1H), 2.12 (d, J = 12.6 Hz, 1H), 1.98 (dd, J = 12.6, 0.7 Hz, 1H), 1.79–1.7 (m, 5H), 1.61 (d, J = 2.7 Hz, 2H) 1.51 (d, J = 6.0 Hz, 3H), 1.46 (s, 3H) ppm; MS (EI): m/z 230(M+, 43%), 216 (16), 215 (100), 187 (59), 118 (19), 117 (25), 91 (14), 55 (30); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C16H22O 230.1671, found 230.1672.
3-Methyl-3-phenyl-2,3,3a,8b-tetrahydrofuro [3,2-b] benzofuran (3bl): Inseparable mixture of regioisomers. The major isomer data is highlighted in bold. Yellow solid; purification by flash chromatography (hexane/EtOAc), 35% estimated yield (not purely isolated); 1H NMR (400 MHz, CDCl3): δH = 7.60–7.49 (m, 3H), 7.48–7.38 (m, 3H), 7.38–7.30 (m, 8H), 7.27–7.21 (m, 2H), 7.00–6.88 (m,2H), 5.60 (d, J = 6.0 Hz, 1H), 5.15 (d, J = 6.0 Hz, 1H), 4.31 (d, J = 9.3 Hz, 1H), 4.19 (d, J = 12.4 Hz, 1H), 3.91 (d, J = 5.7 Hz, 1H), 3.82 (d, J = 12.2 Hz, 1H), 3.64 (d, J = 9.3 Hz, 1H), 3.60 (d, J = 12.4 Hz, 1H), 1.61 (s, 3H), 1.52 (s, 3H), ppm; MS (EI): m/z 252 (M+, 71%), 237 (14), 221 (15), 207 (100), 194 (20), 178 (15), 145 (23), 131 (41), 129 (11), 118 (36), 115 (16), 105 (14), 91 (20), 89(15), 77 (14); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C17H16O2 252.1150, found 252.1149.
4a-Phenyl-1,2,3,4,4a,4b,9b,10a-octahydrobenzofuro[3,2-b]benzofuran (3cl): white solid; purification by flash chromatography (hexane/EtOAc), 38% yield; diastereomeric ratio = 55:45; 1H NMR (400 MHz, CDCl3): δH = 7.53–7.47 (m, 3H), 7.45–7.34 (m, 3H), 7.34–7.30 (m, 1H), 7.02–6.90 (m, 2H), 5.54 (d, J = 7.8 Hz, 1H), 5.15 (d, J = 7.9 Hz, 1H), 4.43 (t, J = 2.5 Hz, 1H), 2.17–2.03 (m, 1H), 2.00–1.85 (m, 1H), 1.51–1.39 (m, 3H), 1.05–0.82(m, 3H) ppm; 13C NMR (400 MHz, CDCl3): δC =130.7, 130.4, 128.5, 128.2, 128.0, 127.0, 126.4, 126.3, 126.2, 126.0, 120.9, 120.7, 110.0, 109.7, 96.8, 93.9, 80.7, 80.5, 78.9, 77.2, 75.3, 32.2, 29.7, 28.7, 26.5, 25.1, 21.1, 20.4, 20.2 ppm; MS (EI): m/z 292 (M+, 17%), 274 (20), 208 (17), 207 (100), 194 (25), 91 (11); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C20H20O2 292.1463, found 292.1460.
Ethyl 3,5-dimethyl-3,5-diphenyltetrahydrofuran-2-carboxylate (3da): yellow solid; purification by flash chromatography (hexane/EtOAc), 43% yield; diastereomeric ratio = 65:35; 1H NMR (300 MHz, CDCl3): δH = 7.68–7.62 (m, 2H), 7.54–7.49 (m, 2H), 7.42–7.34 (m, 6H), 7.26 (d, J = 2.0 Hz, 1H), 5.08 (s, 1H), 4.21 (qd, J = 7.1, 1.5 Hz, 2H), 2.75 (d, J = 12.9 Hz, 1H), 2.65 (d, J = 12.8 Hz, 1H), 1.61 (s, 3H), 1.23 (t, J = 7.1 Hz, 3H) 1.22 (t, J = 7.1 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): only major isomer is given, δC = 170.6, 148.7, 145.4, 128.3, 128.1, 126.4, 126.3, 126.3, 124.6, 84.7, 84.3, 60.7, 56.6, 31.8, 29.7, 23.5, 14.2 ppm; MS (EI): m/z 324 (M+, 0.13%), 309 (86), 251 (48), 233 (18), 207 (27), 173 (14), 133 (17), 129 (16), 105 (100), 91 (15), 77 (15); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C21H24O3 324.1725, found 324.1715.
Ethyl 3-methyl-3,5,5-triphenyltetrahydrofuran-2-carboxylate (3dc): yellow oil; purification by flash chromatography (hexane/EtOAc), 60% yield; diastereomeric ratio = 90:10; 1H NMR (300 MHz, CDCl3): δH = 7.70–7.64 (m, 3H), 7.52–7.48 (m, 3H), 7.39–7.30 (m, 6H), 7.24–7.18 (m, 3H), 4.96 (s, 1H), 4.14 (q, J = 7.1 Hz, 2H), 3.33 (d, J = 13.1 Hz, 1H), 3.01 (d, J = 13.0 Hz, 1H), 1.34 (s, 3H), 1.20 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 170.3, 147.2, 147.0, 145.1, 128.5, 128.3, 128.2, 128.1, 127.8, 126.8, 126.6, 126.5, 126.4, 126.3, 125.7, 125.5, 125.1, 87.6, 85.1, 60.7, 56.0, 51.2, 23.9, 14.2 ppm; MS (EI): m/z 386 (M+, 0.75%), 314 (15), 313 (59), 309 (45), 295 (24), 269 (26), 206 (12), 196 (75), 191 (30), 181 (12), 178 (16), 167 (86), 165 (30), 133 (24), 105 (100), 91 (1), 77 (19); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C26H26O3 386.1882, found 386.1884.
Ethyl-3-methyl-3,5-diphenyltetrahydrofuran-2-carboxylate (3dd): orange solid; purification by flash chromatograhpy (hexane/EtOAc), 47% yield; diastereomeric ratio = 90:10; signals for the major isomer: 1H NMR (400 MHz, CDCl3): δH = 7.66–7.60 (m, 1H), 7.55 (dd, J = 8.4, 1.1 Hz, 1H), 7.48–7.29 (m, 8H), 5.05 (d, J = 5.6 Hz, 1H), 5.01 (s, 1H), 4.36–4.25 (m, 2H), 2.71 (dd, J = 12.8, 5.6 Hz, 1H), 2.28 (dd, J = 12.8, 10.5 Hz, 1H), 1.46 (s, 3H), 1.35 (td, J = 7.1, 4.4 Hz, 3H) ppm; signals for the minor isomer: 1H NMR (300 MHz, CDCl3): δH = 7.54–7.48 (m, 2H), 7.44–7.32 (m, 8H), 5.48 (dd, J =10.4, 5.5 Hz, 1H), 5.06 (s, 1H), 4.28 (q, J = 7.2 Hz, 2H), 2.57 (dd, J = 12.4, 5.6 Hz, 1H), 2.49–2.39 (m, 1H), 1.77 (s, 3H), 1.44 (s, 3H) ppm; 13C NMR (75 MHz, CDCl3): δC = 172.2, 145.9, 141.5, 128.7, 128.4, 128.3, 128.0, 127.6, 126.7, 126.7, 126.5, 126.1, 125.8, 86.0, 81.3, 60.8, 51.2, 48.3, 24.7, 14.3 ppm; MS (EI): m/z 310 (M+, 0.59%), 237 (14), 191 (100), 147 (12), 145 (26), 120 (18), 115 (23), 105 (45), 91 (27), 77 (12); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C20H22O3 310.1569, found 310.1565.
Ethyl 3-methyl-3-phenyl-2,3,3a,8b-tetrahydrofuro[3,2-b]benzofuran-2-carboxylate (3dl): yellow oil; purification by flash chromatography (hexane/EtOAc), 44% yield; diastereomeric ratio = 70:30; 1H NMR (300 MHz, CDCl3): δH = 7.62–7.29 (m, 7H), 7.06–6.80 (m, 2H), 5.70 (d, J = 6.2 Hz, 1H), 5.17 (d, J = 6.2 Hz, 1H), 4.94 (s, 1H), 4.05 (qd, J = 7.1, 1.4 Hz, 2H), 1.56 (s, 3H), 1.07 (t, J = 7.1 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 169.2, 160.3, 139.0, 131.0, 127.9, 127.9, 126.9, 126.5, 125.0, 121.4, 109.9, 93.5, 81.8, 79.9, 61.1, 54.5, 20.8, 14.0 ppm; MS (EI): m/z 234 (M+, 40%), 251 (30), 221 (23), 208 (16), 207 (100), 178 (13), 145 (16), 133 (43), 118 (23), 105 (92); HRMS (GC/MS-EI/Q-TOF): m/z calcd. for C20H20O4 324.1362, found 324.1361.

4. Conclusions

In conclusion, we have described a new methodology for the straightforward synthesis of substituted tetrahydrofurans based on the reaction of electron-rich alkenes with epoxides mediated by fluorinated alcohol 1,1,1,3,3,3-hexafluoroisopropanol (HFIP). Although the yields achieved are moderate in most of the cases, the procedure can be envisioned as environmentally benign as it has a perfect atom economy and the reactants are readily available from raw materials such as alkenes with minimum manipulation. Using this methodology, not only densely substituted furans, but also spiro- and polycyclic compounds containing furan moiety were obtained. Additionally, preliminary mechanistic studies point towards a purely ionic pathway (SN1-type) or SN2-like mechanism depending on the nucleophilicity of the alkene employed.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/25/15/3464/s1, Experimental procedures, mechanism elucidation tests, compound characterization data, and copies of NMR spectra for all new compounds.

Author Contributions

N.L. and A.B. conceived of and designed the experiments. N.L. performed the experiments. A.B. supervised the experiments. N.L. and A.B. wrote the paper. Both authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministerio de Economía, Industria y Competitividad (CTQ2015-66624-P and CTQ2017-88171-P) and Spanish Ministerio de Ciencia, Innovación y Universidades (PGC2018-096616-B-I00) and the University of Alicante (UADIF17-27, VIGROB-285, VIGROB-316/19, UADIF19-106) and we would like to thank them for the financial support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fang, X.; Hu, X. Advances in the synthesis of lignan natural products. Molecules 2018, 23, 3385–3407. [Google Scholar]
  2. Lorente, A.; Lamariano-Merketegi, J.; Albericio, F.; Álvarez, M. Tetrahydrofuran-containing macrolides: A fascinating gift from the deep sea. Chem. Rev. 2013, 113, 4567–4610. [Google Scholar]
  3. Wolfe, J.P.; Hay, M.B. Recent advances in the stereoselective synthesis of tetrahydrofurans. Tetrahedron 2007, 63, 261–290. [Google Scholar]
  4. Gharpure, S.J.; Vishwakarma, D.S.; Nanda, S.K. Lewis acid mediated “endo-dig” hydroalkoxylation–reduction on internal alkynols for the stereoselective synthesis of cyclic ethers and 1,4-Oxazepanes. Org. Lett. 2017, 19, 6534–6537. [Google Scholar]
  5. Rodríguez, A.D.; Cóbar, O.M.; Padilla, O.L. The calyxolanes:  New 1,3-diphenylbutanoid metabolites isolated from the Caribbean marine sponge. Calyx Podatypa. J. Nat. Prod. 1997, 60, 915–917. [Google Scholar]
  6. Pauli, L.; Tannert, R.; Scheil, R.; Pfaltz, A. Asymmetric hydrogenation of furans and benzofurans with Iridium–Pyridine–Phosphinite catalysts. Chem. Eur. J. 2015, 21, 1482–1487. [Google Scholar]
  7. Chaimanee, S.; Pohmakotr, M.; Kuhakarn, C.; Reutrakul, V.; Soorukram, D. Asymmetric synthesis of ent-fragransin C1. Org. Biomol. Chem. 2017, 15, 3985–3994. [Google Scholar]
  8. Clive, D.L.J.; Stoffman, E.J.L. Total synthesis of (–)-conocarpan and assignment of the absolute configuration by chemical methods. Chem. Commun. 2007, 2151–2153. [Google Scholar]
  9. Wang, L.; Li, Z.; Lu, L.; Zhangk, W. Synthesis of spiro[furan-3,3′-indolin]-2′-ones by PET-catalyzed [3+2] reactions of spiro[indoline-3,2′-oxiran]-2-ones with electron-rich olefins. Tetrahedron 2012, 68, 1483–1491. [Google Scholar]
  10. Shuler, W.G.; Combee, L.A.; Falk, I.D.; Hilinski, M.K. Intermolecular electrophilic addition of epoxides to alkenes: [3+2] cycloadditions catalyzed by lewis acids. Eur. J. Org. Chem. 2016, 3335–3346. [Google Scholar]
  11. Hilt, G.; Bolze, P.; Harms, K. An improved catalyst system for the iron--catalyzed intermolecular ring--expansion reactions of epoxides. Chem. Eur. J. 2007, 13, 4312–4325. [Google Scholar]
  12. Hilt, G.; Walter, C.; Bolze, P. Iron--salen complexes as efficient catalysts in ring expansion reactions of epoxyalkenes. Adv. Synth. Catal. 2006, 348, 1241–1247. [Google Scholar]
  13. Hilt, G.; Bolze, P.; Kieltsch, I. An iron-catalysed chemo- and regioselective tetrahydrofuran synthesis. Chem. Commun. 2005, 15, 1996–1998. [Google Scholar]
  14. Pérez, J.M.; Maquilón, C.; Ramón, D.J.; Baeza, A. Hexafluoroisopropanol-promoted metal-free allylation of silyl enol ethers with allylic alcohols. Asian J. Org. Chem. 2017, 6, 1440–1444. [Google Scholar]
  15. Trillo, P.; Baeza, A.; Nájera, C. Fluorinated alcohols as promoters for the metal-free direct substitution reaction of allylic alcohols with nitrogenated, silylated, and carbon nucleophiles. J. Org. Chem. 2012, 77, 7344–7354. [Google Scholar] [PubMed]
  16. An, X.-D.; Xiao, J. Fluorinated alcohols: Magic reaction medium and promoters for organic synthesis. Chem. Rec. 2020, 20, 142–161. [Google Scholar] [CrossRef] [PubMed]
  17. Colomer, I.; Chamberlain, A.E.R.; Haughey, M.B.; Donohoe, T.J. Hexafluoroisopropanol as a highly versatile solvent. Nat. Rev. Chem. 2017, 1, 88. [Google Scholar]
  18. Bégué, J.-P.; Bonnet-Delpon, D.; Crousse, B. Fluorinated alcohols: A new medium for selective and clean reaction. Synlett 2004, 2004, 18–29. [Google Scholar]
  19. Wencel-Delord, J.; Colobert, F. A remarkable solvent effect of fluorinated alcohols on transition metal catalysed C–H functionalizations. Org. Chem. Front. 2016, 3, 394–400. [Google Scholar]
  20. Sugiishi, T.; Matsugi, M.; Hamamoto, H.; Amii, H. Enhancement of stereoselectivities in asymmetric synthesis using fluorinated solvents, auxiliaries, and catalysts. RSC Adv. 2015, 5, 17269–17282. [Google Scholar]
  21. Khaksar, S. Fluorinated alcohols: A magic medium for the synthesis of heterocyclic compounds. J. Fluor. Chem. 2015, 172, 51–61. [Google Scholar]
  22. Shuklov, I.A.; Dubrovina, N.V.; Börner, A. Fluorinated alcohols as solvents, cosolvents and additives in homogeneous catalysis. Synthesis 2007, 19, 2925–2943. [Google Scholar]
  23. Tian, Y.; Xu, X.; Zhang, L.; Qu, J. Tetraphenylphosphonium Tetrafluoroborate/1,1,1,3,3,3-Hexafluoroisopropanol (Ph4PBF4/HFIP) effecting epoxide-initiated cation–olefin polycyclizations. Org. Lett. 2016, 18, 268–271. [Google Scholar] [PubMed]
  24. Li, G.-X.; Qu, J. Friedel–crafts alkylation of arenes with epoxides promoted by fluorinated alcohols or water. Chem. Commun. 2010, 46, 2653–2655. [Google Scholar]
  25. Westermaier, M.; Mayr, H. Regio- and stereoselective ring--opening reactions of epoxides with indoles and pyrroles in 2,2,2--Trifluoroethanol. Chem. Eur. J. 2008, 14, 1638–1647. [Google Scholar]
  26. Das, U.; Crousse, B.; Kesavan, V.; Bonnet-Delpon, V.D.; Bégué, J.-P. Facile ring opening of oxiranes with aromatic amines in fluoro alcohols. J. Org. Chem. 2000, 65, 6749–6751. [Google Scholar]
  27. Wang, Z. Comprehensive Organic Name Reactions and Reagents Vol. 2; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2010; p. 892. [Google Scholar]
  28. Neimann, K.; Neumann, R. Electrophilic activation of hydrogen peroxide:  Selective oxidation reactions in perfluorinated alcohol solvents. Org. Lett. 2000, 2, 2861–2863. [Google Scholar]
  29. Sultana, S.; Devi, N.R.; Saikia, A.K. Synthesis of substituted Tetrahydropyrans and Tetrahydrofurans via intramolecular hydroalkoxylation of alkenols. Asian J. Org. Chem. 2015, 4, 1281–1288. [Google Scholar]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. Natural products and bioactive molecules containing tetrahydrofuran moiety.
Figure 1. Natural products and bioactive molecules containing tetrahydrofuran moiety.
Molecules 25 03464 g001
Scheme 1. Synthesis of substituted tetrahydrofurans by reaction between alkenes and epoxides.
Scheme 1. Synthesis of substituted tetrahydrofurans by reaction between alkenes and epoxides.
Molecules 25 03464 sch001
Scheme 2. Reaction between styrene oxide and electron-rich alkenes a.
Scheme 2. Reaction between styrene oxide and electron-rich alkenes a.
Molecules 25 03464 sch002
Scheme 3. One-pot oxidation/ring opening reaction.
Scheme 3. One-pot oxidation/ring opening reaction.
Molecules 25 03464 sch003
Scheme 4. Possible operating reaction mechanisms.
Scheme 4. Possible operating reaction mechanisms.
Molecules 25 03464 sch004
Scheme 5. Mechanism elucidation tests.
Scheme 5. Mechanism elucidation tests.
Molecules 25 03464 sch005
Table 1. Optimization of the reaction parameters a.
Table 1. Optimization of the reaction parameters a.
Molecules 25 03464 i001
EntrySolventTemp (°C)Conv. (%) b
1H2O4513
2iPrOH450
3HFIP4572
4TFE453
5None452
6HFIP/CH2Cl2 (9/1)4550
7HFIP/CH2Cl2 (1/1)4515
8HFIP/CH2Cl2 (1/9)4523
9HFIP2530
Molecules 25 03464 i002
a All reactions were carried out using 0.15 mmol of 1a and 0.25 mmol of 2a in 150 µL of the solvent at 45 °C for 20 h. b Conversion towards formation of 3aa, determined by GC-MS.
Table 2. Reaction between styrene oxide and electron-rich alkenes a.
Table 2. Reaction between styrene oxide and electron-rich alkenes a.
Molecules 25 03464 i003
Molecules 25 03464 i004
a All reactions were carried out using 0.15 mmol of 1a and 0.25 mmol of 2a in 150 µL of the solvent. b Estimation of the yield from 1H NMR data. c Conversion towards formation of tetrahydrofuran, determined by 1H NMR and/or GC-MS. d Diastereomeric ratio determined by 1H NMR and/or GC-MS of the crude compounds. e Product decomposition after purification was observed.

Share and Cite

MDPI and ACS Style

Llopis, N.; Baeza, A. HFIP-Promoted Synthesis of Substituted Tetrahydrofurans by Reaction of Epoxides with Electron-Rich Alkenes. Molecules 2020, 25, 3464. https://doi.org/10.3390/molecules25153464

AMA Style

Llopis N, Baeza A. HFIP-Promoted Synthesis of Substituted Tetrahydrofurans by Reaction of Epoxides with Electron-Rich Alkenes. Molecules. 2020; 25(15):3464. https://doi.org/10.3390/molecules25153464

Chicago/Turabian Style

Llopis, Natalia, and Alejandro Baeza. 2020. "HFIP-Promoted Synthesis of Substituted Tetrahydrofurans by Reaction of Epoxides with Electron-Rich Alkenes" Molecules 25, no. 15: 3464. https://doi.org/10.3390/molecules25153464

Article Metrics

Back to TopTop