Abstract
A novel approach for the formation of anomeric carbon-functionalized furanoside systems was accomplished through the employment of an oxo-rhenium catalyst. The transformation boasts a broad range of nucleophiles including allylsilanes, enol ethers, and aromatics in addition to sulfur, nitrogen, and hydride donors, able to react with an oxocarbenium ion intermediate derived from furanosidic structures. The excellent stereoselectivities observed followed the Woerpel model, ultimately providing 1,3-cis-1,4-trans systems. In the case of electron-rich aromatic nucleophiles, an equilibration occurs at the anomeric center with the selective formation of 1,3-trans-1,4-cis systems. This anomalous result was rationalized through density functional theory calculations. Different oxocarbenium ions such as those derived from dihydroisobenzofuran, pyrrolidine, and oxazolidine heterocycles can also be used as a substrate for the oxo-Re-mediated allylation reaction.
Introduction
Decorated tetrahydrofurans (THFs) are prevalent motifs in the structures of a myriad of natural products. In particular, the trans-1,4-dialkyl-substituted THF ring is a common functionality in biologically active compounds, polyether antibiotics, acetogenins, C-glycosides, and amphidinolides.1Figure 1 illustrates a few examples of natural products that contain this THF ring systems, such as haterumalides NA 1, amphidinolide F 2, (−)-obtusallene III 3, and oxylipid 4.
Figure 1.
Representative members of the trans-2,5-dialkyl-substituted THF family.
Among the different synthetic strategies to assembly-substituted THFs,1 one of the classical approaches is the venerable Sakurai–Hosomi reaction (Scheme 1), which operates through the exposure of a suitable THF acetal, such as 6, to a strong Lewis acid (e.g., TMSOTf, SnCl4, BF3·Et2O, or TiCl4) and soft, nucleophilic allylsilanes to afford the corresponding allylated THF product 7.2
Scheme 1. General Approach to the Trisubstituted THF Core via the Sakurai–Hosomi Reaction.
However, this reaction is limited by the highly corrosive and moisture-sensitive Lewis acids that are required in stoichiometric amounts to provide the desired products. Therefore, an affordable and reliable catalytic system that could promote an efficient and stereoselective reaction would be advantageous considering both the commercial availability and the low synthetic manipulation of THF acetals. To the best of our knowledge, only two catalytic approaches have been developed for the allylation of THF acetals. In 2009, Yadav reported a protocol using molecular iodine in CH2Cl2 at cryogenic temperatures.3 In the same year, Fiestad delivered an alternative approach to solve the innate stereochemical problems of cyclic, five-membered oxocarbenium ions, developing an allylation reaction on a 2,7-dioxabicyclo[2.2.1]heptane ring system which uses a catalytic amount of TiCl4 as a Lewis acid.4 As a part of our program aimed to explore the chemistry of rhenium (V),5 herein, we report a new talent of oxo–rhenium complex 8, namely, its ability to promote the formation of an oxocarbenium species in a catalytic fashion (Scheme 2).
Scheme 2. Catalytic Approaches to the Allylation Reaction via a Five-Membered Ring-Oxocarbenium Intermediate.
DCE = dichloroethane and DCM = dichloromethane.
Results and Discussion
In the course of our studies on rhenium (V) chemistry, we discovered that Lewis acidic oxo-rhenium complex 8 was catalytically competent in the promotion of Meyer Schuster rearrangements and one-pot Meyer Schuster/Diels–Alder reaction sequences. Indeed, the 14e– complex 8 exhibited an efficiency even higher than that of hallmark Lewis acids (e.g. Cu(OTf)2, BF3·Et2O) usually employed in the Meyer Schuster rearrangement.5a−5c
Therefore, we reasoned that substoichiometric metal-oxo complex 8 could promote the Sakurai–Hosomi reaction between suitable THF acetals and allyltrimethylsilane.5d
In the event, the reaction of simple ethoxyl-acetal 9 (Table 1)6 with allyltrimethylsilane (250 mol %) in the presence of 15 mol % catalyst 8 in CH2Cl2 was initially carried out in a range of temperatures from −50 to −30 °C, but no significant trace of product was observed. Only after 30 min at room temperature, the expected products 10 (Table 1) were obtained, albeit in moderate 45% isolated yields and as a 55:45 trans–cis epimeric mixture at C-1 (Table 1, entry 1).7
Table 1. Optimization of Reaction Conditions: General Reaction Conditions: 9 (1 mmol), Allyltrimethylsilane (2.5 mmol), Oxo-Re Catalyst 8 (15 mol %), and Cocatalyst (5 mol %) Were Reacted in CH2Cl2 (10 mL) at −50 °Ca.
| entry | oxo-rhenium catalyst | cocatalyst | time (h) | yield (%) |
|---|---|---|---|---|
| 1b | Re(O)Cl3(OPPh3)(Me2S) 8 | none | 0.5 | 45 |
| 2 | Re(O)Cl3(OPPh3)(Me2S) 8 | Cu(OTf)2 | 12 | 92 |
| 3c | Re(O)Cl3(OPPh3)(Me2S) 8 | Cu(OTf)2 | 16 | n.d. |
| 4d | Re(O)Cl3(OPPh3)(Me2S) 8 | CuI | 26 | 80 |
| 5 | Re(O)Cl3(OPPh3)(Me2S) 8 | CoCl2 | 26 | 85 |
| 6e | none | CoCl2–Me2S | 12 | n.d. |
| 7e | none | Co(OTf)2–Me2S | 12 | n.d. |
| 8 | (dppe)ReOCl3 | none | 12 | n.d. |
| 9 | (dppp)ReOCl3 | none | 12 | n.d. |
| 10 | (dppm)ReOCl3 | none | 12 | n.d. |
Temperature was then increased to −30 °C for a given time indicated in the table.
Without the cocatalyst and the reaction mixture was warmed to rt for 0.5 h.
15 mol % Cu cocatalyst.
30 mol % of cocatalyst.
5 mol % of Me2S, without catalyst 8, has been used.
Substrate 9 has been used as a 1:1 mixture at C1 by virtue of the control experiment and the literature precedent that revealed that the stereoselectivity was not influenced by the anomeric composition of the starting glycoside.8
We speculate that this low reactivity could be attributed to the inability of acetal 9 to displace the dimethylsulfide ligand from the complex 8, a pre-equilibrium necessary to initiate the catalytic cycle (vide infra). We hypothesized that a thiophilic cocatalyst would facilitate the displacement of the sulfide ligand to form the active, oxophilic 12e– Re complex and by consequence improve the reaction kinetics. A brief screening of copper salts was performed, as they maintain a high affinity for sulfur.9 We were delighted to find that catalyst 8 (15 mol %) in the presence of Cu(OTf)2 (5 mol %) afforded the desired product 10 even at low temperature (−50 to −30 °C) with 92% isolated yield (Table 1, entry 2). Increasing the catalyst loading of Cu(OTf)2 to 15 mol %, no reaction was observed (Table 1, entry 3). Switching to a softer copper (I) salt, such as CuI,10 the reaction proceeded promptly but only when the catalyst loading was increased to 35 mol % of both Re catalyst 8 and CuI cocatalyst (Table 1, entry 4). This behavior could be ascribed to the stronger electrophilicity of Cu(II) compared to that of Cu(I), which causes Cu(II) to bind Me2S molecules more strongly. The essential role of the Me2S scavenging in the catalytic cycle was further demonstrated by employing CoCl2 instead of Cu(OTf)2.11
Indeed, allylated products 10 have been obtained in a slightly lower isolated yield (85%) using CoCl2 in the presence of 8 (15 mol %) (Table 1, entry 5). A background reaction promoted by the Lewis acidity of the cocatalyst was ruled out by carrying out the allylation reaction with the sole Cu(OTf)2 and CoCl2 in the presence of an equimolecular amount, with respect to the cocatalyst, of Me2S (Table 1, entries 6 and 7). CoCl2 was inactive to promote the allylation reaction, while Cu(OTf)2 induced an extensive decomposition of acetal 9.
Not surprisingly, when alternative 14e– Re-complexes, such as (dppm)ReOCl3, (dppe)ReOCl3, and (dppp)ReOCl3, were employed, no reactions were observed, even upon warming the reaction mixture at room temperature (Table 1, entries 8–10).
With an allylation protocol in hands, we tested the versatility of our synthetic method by reacting oxo-Re catalyst 8 with different nucleophiles using (3S,4R)-bis-TBDPS-protected methoxy-acetal 11 (Scheme 3).
Scheme 3. Oxo-Rhenium-Mediated Allylation of O-Methyl bis-TBDSO-Protected Furanoside Acetals with Various Nucleophiles.

Conditions: 1.0 equiv of furanoside 11, 2.5 equiv of nucleophile, 15 mol % oxo-Re complex 8, 5 mol % Cu(OTf)2, CH2Cl2 (0.05 M), from −50 °C to −30 °C; (a) reaction temperature was increased to 0 °C; (b) 30 mol % oxo-Re complex 8 and 15 mol % Cu(OTf)2 were used; and (c) 1.5 equiv of furan nucleophile was used. Unless otherwise noted, alkylated furanosides displayed a 1,3-cis - 1,4-trans stereochemistry. Diastereoselectivities, except for compound 16, were determined from the 1H NMR spectrum.
Allyltrimethylsilane was added to an oxocarbenium ion intermediate with an exquisite 99:1, 1,3-syn (1,4-anti) stereoselectivity with a 93% isolated yield. Relative stereochemistry of 12 has been established by chemical correlation with the known Borhan allylated diol, previously reported as diol 4.12 This complete selectivity is remarkable in light of the fact that other catalytic approaches3 afford the corresponding TBS-protected 12 in 90:10 selectivity. Indeed, in order to achieve the same level of stereoselectivity, the reaction requires an equimolar amount of Lewis acids such as BF3 etherate13 or SnBr4.14 We could anticipate that the stereochemical outcome, for products from 13 to 24, is in good agreement with the inside attack model proposed by Woerpel and co-workers for ribose-derived analogues.8a,15
The same levels of both stereoselectivities and yields have been observed using substituted allylsilanes such as 2-bromoallyltrimethylsilane, ethyl 3-(trimethylsilyl)-4-pentenoate, and methallyltrimethylsilane, delivering alkylated products 13, 14, and 15, respectively. The latter was obtained with a 1,3-syn stereoselectivity of 95:5, a remarkable result in light of the poor selectivity obtained by Yadav for an analogous compound.3 The use of 2-(trimethylsilyl)-2,3-pentadiene afforded product 16, with a pendant alkyne, in good yield and high 1,3-syn stereoselectivity and 7:3 (UPLC analysis) at a C-4′ (S)-stereocenter.16 Noncarbon-based nucleophiles such as nitrogen (Me3SiN3), sulfur (Me3SiSPh), and hydride (Et3SiH)17 afforded the corresponding azido- (17) and thio- (19) glycosides and 1,4-anhydro-2-deoxy-glycitol (18) in high yield and good selectivity. Other TMS-enol ether nucleophiles proved to be excellent partners in the oxocarbenium coupling; glycosylation with acetal 11 proceeded smoothly under the previous C-allylation conditions. All the reactions provided the desired C-glycosides (20–24) in high yields and stereoselectivity. An NOE enhancement observed by NMR analysis of 23 (see the Supporting Information for details) supported its 1,3-cis, 1,4-trans relative configurations, thus establishing the expected stereochemical Woerpel outcome. Of note, 1,3-syn stereoselectivity has been imposed without aid of either a directing group18 or by ortho-alkynylbenzoate glycosyl derivatives.19 This showcases the high atom economy of our glycosidation protocol. When vinylogous heterocyclic TMS-ether 2-(trimethylsilyloxy)furan was used,20 an interesting stereochemical inversion was observed. Indeed, glycoside 25 was obtained in 57% isolated yield higher than a 99:1 mixture of diastereoisomers, in which the 1,3-trans stereoisomer was formed rather than the 1,3-cis stereoisomer.21 The synthesis of aryl-C-glycosides via Lewis acid catalysis has been reported in the literature to afford primarily the β-stereoisomer (1,4-cis in our scenario) over the α-stereoisomer (1,4-trans).22 Albeit the origin of observed stereoselectivity has been attributed to a Lewis/Brønsted acid thermodynamic isomerization, a clear mechanistic picture is not described in the literature.22a−22c,22e,23 The formation of 25 therefore could be explained by an initial Friedel–Crafts reaction on the electron-rich 2-(trimethylsilyloxy)furan and then epimerization followed by furan ring dearomatization. Not surprisingly, a deviation from the Woerpel stereochemical model was observed when electron-rich, aromatic C-nucleophiles were subjected to our glycosidation protocol. In the event, anisole (para regioisomer only), 1,3-benzodioxole, and furan afforded the corresponding aryl-C-glycosides 26, 27, and 28, respectively, in high yields and stereoselectivity. Vorbüggen stereoselectivity, that is, 1,3-anti and 1,4-syn, was inferred by NOE experiments (see the Supporting Information) and tentatively explained using quantum chemical calculations (vide infra).
Next, the rhenium oxo-complex-mediated allylation procedure was next extended to more decorated and synthetically useful THF acetal substrates (Scheme 4). Acetates and benzyl ethers on carbons C-4 and C-5 and tosylate on C-4 (ribose numbering system)1 are well-tolerated, affording the corresponding allylated diacetate 29, bis-benzyl ethers 30, and tosylate 31, in high yields and selectivities. The potential of this catalytic Re-mediated process compares favorably with those reported in the literature that typically require a stoichiometric amount of the Lewis acid trimethylsilyltrifluoromethanesulfonate (TMSOTf).24 The stereochemical behavior predicted by Woerpel was demonstrated by reacting the corresponding C-3 epimer of 11 (not shown) under our Re-mediated allylation reaction. Glycoside 32 was thus obtained in 86% isolated yield as a 73:27 1,3-cis/1,3-trans mixture of C-1 isomers, respectively. The stereochemical inversion could be ascribed to the pseudo-equatorial orientation—instead of the preferred pseudo-axial one—of the C-3 alkoxy group in the five-membered-ring oxocarbenium systems, as observed by Woerpel.8a,15b
Scheme 4. Oxo-Rhenium-Mediated Allylation Reaction on Decorated O-Alkyl Furanoside Acetals.

Conditions: 1 equiv of furanoside substrate, 2.5 equiv of allyltrimethyl silane, 15 mol % oxo–Re complex 8, 5 mol % Cu(OTf)2, CH2Cl2 (0.05 M), from −50 to −30 °C. Unless otherwise noted, reactions were carried out on the corresponding O-methyl furanoside. (a) Reaction was carried out on the corresponding O-allyl acetal; (b) 30 mol % oxo-Re complex 8 and 15 mol % Cu(OTf)2 were used. Unless otherwise noted, alkylated furanosides displayed a 1,3-cis/1,4-trans stereochemistry. Diastereoselectivities were determined from the 1H NMR spectrum.
Fully decorated THF afforded the corresponding allylated products 33 and 34 in 93% and 77% isolated yields, respectively. The 91:9 stereoselectivity in favor of the 1,3-cis stereoisomer of 33, confirmed by two-dimensional NOESY experiments, is of particular importance in light of the conflicting results found in the literature and which prompted Vogel to publish an in-depth examination of the synthesis of 33 using alternative approaches.25 Our robust and reliable allylation protocol could therefore be used to prepare a stereochemically defined allyl-C-glycoside such as 33 without any doubt about the C-1 allyl stereochemistry. Finally, aryl-C-glycoside could be prepared on a decorated THF ring as demonstrated by the high yield and stereoselective preparation of compound 35. This complete stereoselectivity is due to an anchimeric assistance of the acetate group at the C-2 position of the ribose, which shields the α-face of the oxocarbenium intermediate.26a The regioselectivity of the Friedel–Craft reaction is in line with the data in the literature, as only the ortho product has been obtained. This is intriguing, in light of the fact that the stereoselectivity in our case is complete.26b
Another key feature of 31 and 32 formation lies in the nature of substituent at C-1, namely, the allyloxy group. This group could be easily ionized under our Re catalysis, to form the corresponding oxocarbenium ion. Of note, the direct use of allyloxy-protected glycosides in glycosidation reactions, for example, allows us to save the deprotection-activation sequence procedure which usually involves the use of Pd, Ir, or even Hg salts.27
Dihydroisobenzofuran, pyrrolidine, and oxazolidine heterocycles could also be employed as substrates for this oxo-Re-mediated allylation (Scheme 5). In the event, 1-methoxy-1,3-dihydroisobenzofuran 36 and 2-methoxy-1-methoxycarbonylpyrrolidine 38(28) afforded the corresponding allylated product 37 and 39 in 86 and 90% isolated yields, respectively. Interestingly, when chiral29a (4S)-4-benzyl-2-methoxy-3-tosyl-1,3-oxazolidine 40 was subjected to the allylation reaction, the corresponding 2,4-cis diastereoisomer 41 was obtained in more than a 99:1 ratio in 63% isolated yield. According to the data available in the literature, stereoisomer 41 corresponds to the thermodynamic product.29b
Scheme 5. Oxo-Rhenium-Mediated Allylation Reaction on Different Oxocarbenium Ions.
Conditions: 1 equiv of furanoside substrate, 2.5 equiv of allyltrimethyl silane, 15 mol % oxoRe complex 8, 5 mol % Cu(OTf)2, CH2Cl2 (0.05 M), from −50 to −30 °C. Diastereoselectivities were determined from the 1H NMR spectrum.
Computational Studies
To shed some light on the peculiar stereochemical behavior of aryl-C-glycoside formation, some computational density functional theory (DFT) calculations have been carried out.
The reaction starts with the initial approach of the anisole reactant toward the α- or β-face of the oxocarbenium species 5, by forming two quasi-isoenergetic adducts, adduct-α and adduct-β, respectively (Scheme 6a). The system undergoes the Friedel–Crafts reaction passing through the attachment to the α side, a largely favored (ΔΔE = 3.22 kcal/mol) approach over the β one, as indicated by the energy of the TS1-α and TS1-β transition states (TSs), −1.32 and 1.90 kcal/mol, respectively.
Scheme 6. (a) Proposed Mechanism for the Aryl-C-glycoside Formation Based on DFT Calculations (ΔE in kcal/mol); (b) Relative Energy between Products 26-TMS-β and 26-TMS-α.
[B3LYP/6-31G(d,p)&6-31+G(d,p)/CH2Cl2(PCM)//B3LYP/6-31G(d,p)&6-31+G(d,p)]
After the first TS, the two next Wheland intermediates are reported as INT1-α and INT1-β, which maintain the same distribution in selectivity as in the previous TSs.
The results obtained so far, seemed to be in disagreement with the experimental data: literature methods for C-1 epimerization of the furanoside derivative require harsh conditions in order to be accomplished. Strong acids, such as benzenesulfonic acid or TFA, in refluxing toluene22g or in over stoichiometric concentrations are usually required.23b Under these conditions, the formation of a carbocation intermediate has been demonstrated by NMR studies.23b However, in our case, the equilibration occurred at −30 °C in the presence of a catalytic amount of a Lewis acid; therefore, a different scenario should be operative.
Specifically, INT1 species could undergo an intramolecular hydrogen transfer of the two corresponding Wheland cationic intermediates to the furan oxygen atom and the concomitant opening of the THF ring. The TSs obtained for this transformation appear to be the rate-determining step in both the cases, with an activation energy of 25.00 kcal/mol to overcome TS2-α and 25.30 kcal/mol for TS2-β. The resulting intermediates INT2 correspond to the opened-ring cationic structures, where the alcohol obtained from the protonation of the furane oxygen is opposed to the trigonal planar cationic counterpart (see Scheme 6a). Also, in this step, the α-approach is favored over the β-one. These two transition states are fundamental to explain the reactivity of this reaction: indeed, the population of INT2-α as a main intermediate occurs under a kinetic-dominated scenario. Furthermore, INT2-α can be easily interconverted in the more stable INT2-β through TS3-α-β with a very low energy barrier (2.71 kcal/mol) under thermodynamic equilibration. This negligible C–C bond energy rotation is the key step for the equilibration between the two C-1 epimers, thus resulting in the formation of the β-arylated product as a main isomer. By applying a Boltzmann population analysis of the direct (α to β, ΔE = 2.71 kcal/mol) and reverse (β to α, ΔE = 3.40 kcal/mol) reactions at the equilibrium, we found that theoretical values of 76:24 (α/β) are fully consistent with the selectivity trend observed in our experiments, 78:22 (α/β) (Scheme 6a).
In Scheme 6b, the structures of the final deprotonated products 26-TMS-α and 26-TMS-β are reported together with the relative energy, showing the higher stability of the β product over the α one (ΔΔE = 1.54 kcal/mol).
Conclusions
In summary, a new methodology for the formation of oxocarbenium ions and subsequent addition of a wide range of nucleophiles was developed, exploiting the advantage of using commercially available THF O-methoxy-acetals as starting materials in combination with a Re (V) complex catalyst. The peculiarity of this method is the use of catalytic amounts of a stable and easy to handle Re (V) catalyst as a Lewis acid, together with the Cu(OTf)2 cocatalyst, achieving high yields and stereoselectivities. The wide scope for both the nucleophile and the oxocarbenium species paves the way for further investigation in this research field, also opening to a fast and reliable approach for other heterocyclic substrates. Moreover, the computational explanation of the diastereoselectivity observed in the case of electron-rich aromatic nucleophiles uncovered the presence of an equilibration step, which is ultimately responsible for the regulation of the observed anti-Woerpel selectivity.
Experimental Section
General Methods
All the reactions were performed in glassware which has been dried in an oven at 140 °C for at least 3 h prior to use and allowed to cool in a desiccator over self-indicating silica gel pellets. Anhydrous solvents were distilled from appropriate drying agents. Allyltrimethylsilane, anisole, methyl trimethylsilyl dimethylketene acetal, γ-decalactone, 1-O-methyl-2-deoxy-d-ribose, 2-deoxy-d-ribose, methyl d-ribofuranoside, methyl d-arabinofuranoside, methyl 3,5-di-O-acetyl-2-deoxy-d-ribofuranoside S29, 1-methoxy-1,3-dihydro-2-benzofuran 36, 2-methoxy-pyrrolidine-1-carboxylic acid methylester 38, and complex ReOCl3(OPPh3)(SMe2) 8 were purchased and used as received. The reactions were carried out under a slightly positive static pressure of argon. Routine monitoring of reactions was performed using GF- 254 Merck (0.25 mm) aluminum-supported TLC plates. Compounds were visualized by UV irradiation at a wavelength of 254 nm or stained by exposure to a 0.5% solution of vanillin in H2SO4–EtOH followed by charring. Flash column chromatography was performed using Kieselgel 60 Merck (40–63 μm). Yields are reported for chromatographically and spectroscopically pure isolated compounds. NMR spectra were recorded on 200, 300, and 400 MHz spectrometers. Chemical shifts are reported in δ units relative to the employed solvent; the main abbreviations used have the following meaning: s = singlet, d = doublet, t = triplet, q = quartet, qu = quintuplet, m = multiplet, and br = broad. Coupling constants (J) are given in Hz. The multiplicity (in parentheses) of each carbon atom was determined by DEPT experiments, while * indicates the presence of anomers. A relative configuration of the anomeric center of compounds Ba-trans and Bb-cis was determined with NOE experiments.
Synthesis of Acetals
(5R)-2-Ethoxy-5-hexyltetrahydrofuran (9)
DIBAL-H (2.50 mL, 1 M solution in hexane, 2.50 mmol, 1.2 equiv) was added dropwise to a solution of γ-decalactone (0.35 g, 2.10 mmol, 1.0 equiv) in dry CH2Cl2 (13 mL) at −78 °C under an argon atmosphere. After been stirred for 30 min, the mixture was quenched by adding a saturated solution of NH4Cl (50 mL). CH2Cl2 (80 mL) was then added, and the resulting mixture was warmed to rt before the addition of saturated solution of Rochelle’s salt. The layers were separated, and the aqueous phase was extracted with CH2Cl2 (4 × 10 mL). The combined organic phases were washed with brine, dried with MgSO4, filtered, and concentrated under reduced pressure. The resulting residue was dissolved in CH2Cl2/EtOH (2:1, 24 mL) and cooled to 4 °C, and then, PTSA (cat.) was added. After 12 h, excess solid NaHCO3 was added, and the resulting mixture was stirred for 15′, filtered, and concentrated under reduced pressure. The residue was dissolved in CH2Cl2 (50 mL), and a saturated solution of NaHCO3 (50 mL) was added. The layers were separated, and the aqueous phase was extracted with CH2Cl2 (3 × 20 mL). The combined organic phases were dried with Na2SO4, filtered, and concentrated under reduced pressure. The resulting residue was purified by flash chromatography on silica gel. Elution with hexane/EtOAc (98:2) gave compound 9 (386 mg, 92%), colorless oil as a 1:1 mixture of unseparable anomers; 1H NMR (300 MHz, CD2Cl2): δ 5.10 (dd, J = 4.9, 2.0 Hz, 1H), 5.03 (dd, J = 3.7, 1.5 Hz, 1 H), 4.01 (m, 1H), 3.71 (m, 1H), 3.43 (dq, J = 16.1, 8.6 Hz, 1H), 2.04–1.33 (m, 14H), 1.12 (t, J = 7.1 Hz, 3H), 0.92 (t, J = 6.9 Hz, 3H) ppm; 13C{1H} NMR (75 MHz, CD2Cl2): δ 104.4 (CH), 104.1* (CH), 81.2 (CH), 78.6 (CH), 63.2 (CH2), 62.7* (CH2), 38.5 (CH2), 36.4* (CH2), 34.0 (CH2), 33.1 (CH2), 32.7* (CH2), 30.3 (CH2), 30.2* (CH2), 27.1 (CH2), 26.9 (CH2), 23.4 (CH2), 15.9 (CH3), 14.7 (CH3) ppm; IR (film, cm–1): 2980, 1464, 1369, 1107, 739; HRMS (ESI) m/z: [M + H]+ calcd. for C12H25O2, 201.1849; found, 201.1852.
tert-Butyl(((2R,3S)-3-((tert-butyldiphenylsilyl)oxy)-5-methoxytetrahydrofuran-2-yl)methoxy)diphenylsilane (11)
The commercially available 1-O-methyl-2-deoxy-d-ribose (1000 mg, 6.75 mmol) was dissolved
in dry CH2Cl2 (67.5 mL) at room temperature
under an argon atmosphere;
then, imidazole (1379 mg, 20.25 mmol, 3 equiv) and TBDPSCl (3.80 mL,
14.85 mmol, 2.2 equiv) were added under magnetic stirring. The resulting
solution has been stirred for 10 h and then was quenched by adding
a saturated solution of NaHCO3 (50 mL). The layers were
separated, and the aqueous phase was extracted with CH2Cl2 (3 × 50 mL). The combined organic phases were
dried with Na2SO4, filtered, and concentrated
under reduced pressure. The resulting residue was purified by flash
chromatography on silica gel. Elution with hexane/EtOAc (99:1) gave
compound 11 (3922 mg, 95%), colorless oil as a 1:1 mixture
of unseparable anomers. The spectroscopic data correspond with those
reported in the literature (See the Supporting Information for 1H NMR spectra).30
(2R,3S,5S)-5-(Allyloxy)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-ol (Ba) and (2R,3S,5R)-5-(allyloxy)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-ol (Bb)
2-Deoxy-d-ribose (134.1 mg, 1.0
mmol) was dissolved
in CH2Cl2/2-propen-1-ol (15:1, 16 mL) and cooled
to 4 °C, and then, PTSA (cat.) was added. After 14 h, excess
solid NaHCO3 was added, and the resulting mixture was stirred
for 15′ and concentrated under reduced pressure. The residue
was dissolved in CH2Cl2 (50 mL), and a saturated
solution of NaHCO3 (50 mL) was added. The layers were separated,
and the aqueous phase was extracted with CH2Cl2 (3 × 20 mL). The combined organic phases were dried with Na2SO4, filtered, and concentrated under reduced pressure.
The resulting residue was purified by flash chromatography on silica
gel. Elution with hexane–EtOAc (2:8) gave a desired allyl acetal A as a 1:1 mixture of unseparable anomers. Allyl acetal A (120.1 mg, 0.7 mmol) was dissolved in dry CH2Cl2 (6 mL) at room temperature under an argon atmosphere;
then, imidazole (120.1 mg, 1.7 mmol, 3.0 equiv) and TBDPSCl (0.16
mL, 0.61 mmol, 1.1 equiv) were added under magnetic stirring. The
resulting solution has been stirred for 16 h and then was quenched
by adding a saturated solution of NaHCO3 (4 mL). The layers
were separated, and the aqueous phase was extracted with CH2Cl2 (3 × 8 mL). The combined organic phases were
dried with Na2SO4, filtered, and concentrated
under reduced pressure. The resulting residue was purified by flash
chromatography on silica gel. Elution with a gradient from hexane/MTBE
(99:1) to hexane/MTBE (9:1) gave compound Ba-trans (139 mg) and compound Bb-cis (88.1 mg), colorless oil, overall yield = 96%, trans/cis ratio = 1.58/1;
(Ba)-trans
1H NMR (300 MHz, CD2Cl2): δ 7.70 (m, 4H), 7.40 (m, 6H), 5.85 (m, 1H), 5.35–5.15 (m, 3H), 4.28 (dd, J = 13.0, 5.1 Hz, 2H), 4.18 (t, J = 4.9 Hz, 1H), 4.02 (dd, J = 13.0, 5.9 Hz, 1H), 3.80 (dd, J = 10.9, 3.8 Hz, 1H), 3.68 (dd, J = 10.9, 4.8 Hz, 1H), 2.80 (dd, J = 10.6, 1.9 Hz, 1H), 2.25 (ddd, J = 13.6, 6.1, 4.7 Hz, 1H), 1.05 (s, 9H); 13C{1H} NMR (75 MHz, CD2Cl2): δ 136.5 (CH), 135.5 (CH), 134.2 (C), 130.6 (CH), 134.6 (CH), 128.6 (CH), 117.3 (), 104.8 (CH), 88.8 (CH), 73.9 (CH), 68.9 (CH2), 65.4 (CH2), 42.1 (CH2), 27.5 (C), 20.0 (C); IR (film, cm–1): 2933, 1564, 1472, 1432, 1115; HRMS (ESI) m/z: [M + H]+ calcd. for C24H33O4Si, 413.2143; found, 413.2139; [α]D20 + 58.0, c = 0.96 in CH2Cl2.
(Bb)-cis
1H NMR (300 MHz, CD2Cl2): δ 7.75 (m, 4H), 7.40 (m, 6H), 5.85 (m, 1H), 5.25–5.05 (m, 3H), 4.50 (dt, J = 6.3, 4.2 Hz, 1H), 4.15 (dd, J = 13.0, 5.0 Hz, 1H), 3.90 (m,1H), 3.80 (dd, J = 10.3, 5.4 Hz, 1H), 3.70 (dd, J = 10.3, 7.3 Hz, 1H), 2.24 (ddd, J = 13.4, 6.7, 2.3 Hz, 1H), 2.08 (m, 1H), 1.05 (s, 9H); 13C{1H} NMR (75 MHz, CD2Cl2): δ 136.3 (CH), 135.5 (CH), 134.2 (CH), 130.5 (C), 130.4 (C), 128.5 (CH), 128.4 (CH), 116.9 (CH2), 104.1 (CH), 86.9 (CH), 73.9 (CH), 69.0 (CH2), 66.3 (CH2), 42.1 (CH2), 27.4 (CH3), 19.9 (C); IR (film, cm–1): 2931, 1568, 1475, 1431, 1118; HRMS (ESI) m/z: [M + H]+ calcd. for C24H33O4Si, 413.2143; found, 413.2139; [α]D20 = −41.7, c = 0.94 CH2Cl2.
(2R,3S)-3-(benzyloxy)-2-((benzyloxy)methyl)-5-methoxytetrahydrofuran (S30)
The commercially available 1-O-methyl-2-deoxy-d-ribose (200 mg, 1.35 mmol) was dissolved
in dry DMF (6.75
mL) at room temperature under an argon atmosphere. Then, the solution
was cooled down to 0 °C, and NaH (71 mg, 2.97 mmol, 2.2 equiv)
was added under magnetic stirring. The resulting solution was stirred
for 30 min, and then, benzyl-bromide (481 μL, 4.05 mmol, 3 equiv)
was added dropwise. The so-obtained mixture was then warmed to room
temperature and left under stirring overnight. The quench was operated
by the addition of a saturated solution of NH4Cl (10 mL).
The layers were separated, and the aqueous phase was extracted with
Et2O (3 × 10 mL). The combined organic phases were
dried with Na2SO4, filtered, and concentrated
under reduced pressure. The resulting residue was purified by flash
chromatography on silica gel. Elution with hexane/EtOAc (9:1) gave
compound S30 (353 mg, 80%), colorless oil as a 1:1 mixture
of unseparable anomers. The spectroscopic data correspond with those
reported in the literature (see the Supporting Information for 1H NMR spectra).31
(2R,3S,5S)-5-(Allyloxy)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl 4-methylbenzenesulfonate (S31)
Free
alcohol Ba (100.1 mg, 0.24 mmol) was dissolved
in dry pyridine (4.0 mL) at room temperature under an argon atmosphere;
then, tosyl chloride (185.0 mg, 0.96 mmol, 4.0 equiv) and 4-dimethylaminopyridine
(2.9 mg, 0.02 mmol, 0.1 equiv) were added under magnetic stirring.
The resulting solution has been stirred for 15 h and then was quenched
by adding a saturated solution of NaHCO3 (5 mL). The layers
were separated, and the aqueous phase was extracted with Et2O (3 × 5 mL). The combined organic phases were dried with Na2SO4, filtered, and concentrated under reduced pressure.
The resulting residue was purified by flash chromatography on silica
gel. Elution with hexane/AcOEt (9:1) gave compound S31 (122.3 mg, 90%) as colorless oil; 1H NMR (300 MHz, CD3CN): δ 7.75 (d, J = 8.3 Hz, 2H), 7.61
(m, 4H), 7.50–7.40 (m, 8H), 6.05–5.85 (m, 1H), 5.32–5.12
(m, 3H), 5.02 (dt, J = 7.6, 1.5 Hz, 1H), 4.25–4.15
(m, 2H), 4.05–3.91 (m, 1H), 3.60–3.48 (dd, J = 11.3, 3.0 Hz, 2H), 2.45 (s, 3H), 2.25-2.37-1.90 (m, 2H), 1.08
(s, 9H) ppm; 13C{1H} NMR (75 MHz, CD3CN): δ 145.5 (C), 135.4 (CH), 135.3 (CH), 134.8 (CH), 133.0
(C), 132.9 (C), 130.1 (CH), 129.9 (CH), 129.9 (CH), 127.8 (CH), 127.7
(CH), 115.9 (CH2), 103.2 (CH), 83.5 (CH), 80.6 (CH), 67.9
(CH2), 63.2 (CH2), 39.2 (CH2), 26.1
(CH3), 20.7 (CH), 18.7 (C) ppm; IR (film, cm–1): 3068, 2855, 1471, 1424, 1261, 1102, 1004, 940, 713; HRMS (ESI) m/z: [M + H]+ calcd. for C31H39O6SSi, 567.2231; found, 567.2233;
[α]D20 = +65.3, c = 2.72 CH2Cl2.
Synthetical Sequence for the Preparation of Inverted Acetal S32

(2R,3R)-5-(Allyloxy)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl 4-nitrobenzoate (C)
To
a magnetically stirred solution of anomeric mixture of Ba and Bb (80.9 mg, 0.196 mmol) in dry toluene
(2 mL) under an Ar atmosphere at −78 °C were added PPh3 (66.8 mg, 0.255 mmol, 1.3 equiv), 4-NO2-C6H4-CO2H (36.1 mg, 0.216 mmol, 1.1 equiv),
and DEAD (41 mL, 0.262 mmol). The reaction mixture was allowed to
warm to room temperature, and the reaction was completed after 16
h. The reaction mixture was thus concentrated under vacuum. The residue
was purified by column chromatography on silica gel using hexane/AcOEt
92:8 as an eluent to afford the ester C as white foam
(97.9 mg, yield = 89%); 1H NMR (300 MHz, CDCl3): δ 8.40–8.09 (m, 4H), 7.75–7.66 (m, 3H), 7.65–7.57
(m, 2H), 7.51–7.25 (m, 8H), 5.96 (dddd, J =
17.3, 10.6, 5.8, 5.0 Hz, 1H), 5.82 (dt, J = 5.3,
3.7 Hz, 1H), 5.41 (dd, J = 5.0, 3.7 Hz, 1H), 5.29
(dq, J = 17.3, 1.8 Hz, 1H), 5.15 (dq, J = 10.5, 1.6 Hz, 1H), 4.46 (td, J = 6.2, 4.0 Hz,
1H), 4.23 (ddt, J = 13.2, 5.0, 1.6 Hz, 1H), 4.10–3.91
(m, 4H), 2.50–2.40 (m, 2H), 1.00 (s, 9H); 13C{1H} NMR (75 MHz, CDCl3): δ 164.9 (C), 151.9
(C), 136.7 (C), 136.6 (C), 136.3 (C), 134.4 (C), 131.9 (CH), 131.1
(CH), 131.0 (CH), 129.0 (CH), 128.9 (CH), 124.8 (C), 116.9 (CH2), 103.7 (CH), 80.6 (CH), 76.3 (CH), 69.6 (CH2),
63.0 (CH2), 41.4 (CH2), 27.41 (CH3), 20.01 (C); IR (film, cm–1): 3071, 2932, 2857,
1719, 1529, 1479, 1429, 1281, 1103, 1003, 935; HRMS (ESI) m/z: [M + H]+ calcd. for C31H36NO7Si2, 562.2256; found,
562.2253.
(2R,3R)-5-(Allyloxy)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-ol (D)
To a magnetically stirred solution of C (236
mg, 0.42 mmol) in dry MeOH (8 mL) were added few crystals of K2CO3. The reaction mixture was stirred at room temperature,
and the reaction was completed after 16 h. Methanol was removed in
vacuo, and the residue was purified on silica gel using hexane/AcOEt
80:20 as an eluent to afford the desired alcohol D as
colorless oil (164.6 mg, yield = 95%); 1H NMR (300 MHz,
(CD3)2CO): δ 7.86–7.67 (m, 4H),
7.56–7.36 (m, 6H), 5.95 (ddt, J = 17.1, 10.6,
5.4 Hz, 1H), 5.39–5.02 (m, 3H), 4.48 (p, J = 4.4 Hz, 1H), 4.28–3.78 (m, 5H), 2.14 (t, J = 4.4 Hz, 2H), 1.07 (s, 9H); 13C{1H} NMR (75
MHz, (CD3)2CO): δ 136.9 (CH), 136.8 (CH),
136.7 (CH), 136.7 (CH), 134.9 (C), 134.8 (C), 130.9 (CH), 129.0 (CH),
128.9 (CH), 116.5 (CH2), 103.9 (CH), 82.80 (CH), 72.21
(CH), 69.3 (CH2), 64.3 (CH2), 44.1 (CH2), 27.7 (CH3), 27.5 (CH3), 20.1 (C); IR (film,
cm–1): 2934, 1566, 1475, 1431, 1117; HRMS (ESI) m/z: [M + H]+ calcd. for C24H33O4Si, 413.2143; found, 413.2145.
(((2R,3R)-5-(Allyloxy)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl)oxy) (tert-butyl)diphenylsilane (S32)
The epimeric alcohol D (42 mg, 0.1 mmol) was
dissolved in dry CH2Cl2 (2 mL) under Ar, and
at room temperature, imidazole (28 mg, 0.41 mmol, 4.1 equiv) and TBDPSCl
(32 mL, 0.11 mmol, 1.1 equiv) were added. The reaction mixture was
stirred at room temperature for 16 h and was quenched with H2O (2 mL). The layers were separated, and the aqueous layer was extracted
with CH2Cl2 (3 × 5 mL). The combined organic
layers were dried over Na2SO4, filtered, and
concentrated under vacuum. The residue was purified by column chromatography
on silica gel using hexane/MTBE 95:5 as an eluent to afford the product S32 as colorless oil (59.9 mg, yield = 92%); 1H
NMR (300 MHz, (CD3)2CO): δ 7.77 (ddt, J = 12.8, 6.3, 1.8 Hz, 4H), 7.64 (ddt, J = 6.6, 3.7, 1.9 Hz, 4H), 7.52–7.32 (m, 12H), 6.00–5.76
(m, 1H), 5.28–5.02 (m, 3H), 4.65 (dt, J =
6.0, 3.9 Hz, 1H), 4.20–3.88 (m, 5H), 2.15–2.01 (m, H,
overlapped with solvent), 1.85 (ddd, J = 13.8, 6.2,
2.8 Hz, 1H), 1.10 (s, 9H), 1.00 (s, 9H) ppm; 13C{1H} NMR (75 MHz, (CD3)2CO): δ 136.9 (CH),
136.8 (CH), 136.8 (CH), 136.7 (CH), 136.7 (CH), 136.5 (CH), 135.1
(C), 135.0 (C), 134.9 (C), 134.4 (C), 131.1 (CH), 131.1 (CH), 130.9
(CH), 129.1 (CH), 128.9 (CH), 128.9 (CH), 116.6 (CH2),
103.5 (CH), 82.9 (CH), 74.6 (CH), 69.2 (CH2), 65.0 (CH2), 43.7 (CH2), 27.7 (CH3), 27.6 (CH3), 20.2 (C), 20.1 (C) ppm; IR (film, cm–1): 3065, 2858, 1470, 1421, 1257, 1106, 1002, 941, 751, 713, 601;
HRMS (ESI) m/z: [M + H]+ calcd. for C40H51O4Si2, 651.3320; found, 651.3322.
(2R,3R,4R)-2-(Acetoxymethyl)-5-methoxytetrahydrofuran-3,4-diyl diacetate (S33)
The commercially available methyl d-ribofuranoside
(90.7 mg, 0.55 mmol) was dissolved in dry CH2Cl2 (8 mL) at room temperature under an argon atmosphere; then, pyridine
(0.35 μL, 3.92 mmol, 7.1 equiv) and Ac2O (0.20 μL,
1.93 mmol, 3.5 equiv) were added under magnetic stirring. The resulting
solution has been stirred for 15 h and then was quenched by adding
a saturated solution of NaHCO3 (5 mL). The layers were
separated, and the aqueous phase was extracted with CH2Cl2 (3 × 5 mL). The combined organic phases were
dried with Na2SO4, filtered, and concentrated
under reduced pressure. The resulting residue was purified by flash
chromatography on silica gel. Elution with hexane/MTBE (9:1) gave
compound S33 (145.3 mg, 91%) as colorless oil as a mixture
of unseparable anomers. The spectroscopic data correspond with those
reported in the literature (See the Supporting Information for 1H NMR spectra).32
(2R,3R,4S)-3,4-bis(benzyloxy)-2-((benzyloxy)methyl)-5-methoxytetrahydrofuran (S34)
The commercially available
methyl d-arabinofuranoside
was involved in the same procedure abovereported for the synthesis
of compound S30, by giving compound S34 in
85% of isolated yield as colorless oil as a mixture of unseparable
anomers. The spectroscopic data correspond with those reported in
the literature (see the Supporting Information for 1H NMR spectra).33
General Procedure for Allylation with Allyl-TMS
A suitable acetal (0.08 mmol) was dissolved in dry CH2Cl2 (30 mM) at room temperature under an argon atmosphere; then, allyltrimethylsilane (0.40 mmol, 5.0 equiv or less, see manuscript for details) was added under magnetic stirring. The resulting solution was cooled at −50 °C, and complex 8 (15 mol %) followed by Cu(OTf)2 (5 mol %) were added in one portion. The reaction mixture was allowed to warm at −30 °C and stirred at this temperature until the starting material was completely consumed. The reaction was then quenched with a saturated solution of NaHCO3 (2 mL). The layers were separated, and the aqueous phase was extracted with CH2Cl2 (3 × 4 mL). The combined organic phases were dried with Na2SO4, filtered, and concentrated under reduced pressure. The resulting residue was purified by flash chromatography on silica gel. Elution with a suitable solvent mixture (see conditions for each compound) gave the final product (see yield for each compound).
(2S,5R)-2-Allyl-5-hexyltetrahydrofuran (10a) and (2R,5R)-2-allyl-5-hexyltetrahydrofuran (10b)
See Table 1 for yields associated to each reaction conditions; for entry 2, purification with hexane/MTBE (99:1) gave 186.2 mg, 95%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 5.82 (ddt, J = 10.1, 2.7, 0.9 Hz, 1H), 5.12 (dd, J = 2.1, 0.9 Hz, 1H), 5.04 (dd, J = 2.7, 0.9 Hz, 1H), 3.95 (dt, J = 32.0, 7.4 Hz, 1H), 3.89 (dt, J = 43.3, 6.2 Hz, 1H), 2.37 (m, 1H), 2.24 (m, 1H), 1.99 (m, 2H), 1.59–1.29 (m, 12H), 0.89 (t, J = 6.9 Hz, 3H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 134.6 (CH), 116.1 (CH2), 79.1 (CH), 78.5 (CH), 77.9 (CH), 77.3 (CH), 34.0 (CH2), 39.9 (CH2), 35.6 (CH2), 35.5 (CH2), 31.4 (CH2), 31.3 (CH2), 30.9 (CH2), 30.4 (CH2), 29.9 (CH2), 28.9 (CH2), 25.6 (CH2), 22.1 (CH2), 13.6 (C) ppm; IR (film, cm–1): 3070, 2974, 2869, 1641; HRMS (ESI) m/z: [M + H]+ calcd. for C13H25O, 197.1900; found, 197.1903.
(((2R,3S,5R)-5-Allyl-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl)oxy)(tert-butyl)diphenylsilane (12)
Purification with hexane/MTBE (98:2) gave 59.1 mg, yield 93%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.75 (m, 8H), 7.45 (m, 12H), 6.00 (dddd, J = 17.2, 10.2, 6.9, 6.9 Hz, 1H), 5.27 (dd, J = 16.5, 0.8 Hz, 1H), 5.20 (dd, J = 9.6, 0.8 Hz, 1H), 4.62 (ddd, J = 6.8, 3.5, 3.5 Hz, 1H), 4.20 (m, 2H), 3.62 (dd, J = 10.9, 3.5 Hz, 1H), 3.40 (dd, J = 10.9, 3.8 Hz, 1H), 2.65 (ddd, J = 13.9, 6.9, 6.9 Hz, 1H), 2.52 (ddd, J = 13.9, 6.9, 6.9 Hz, 1H), 2.19 (ddd, J = 8.4, 6.0, 1.8 Hz, 1H), 1.89 (ddd, J = 12.5, 5.6, 4.9 Hz, 1H), 1.20 (s, 9H), 1.05 (s, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 135.7 (CH), 135.5 (CH), 135.4 (CH), 135.3 (CH), 133.7 (C), 133.6 (C), 133.3 (C), 129.6 (CH), 129.4 (CH), 129.4 (CH), 127.6 (CH), 127.5 (CH), 116.5 (CH2), 86.6 (CH), 78.6 (CH), 74.8 (CH), 64.4 (CH2), 40.8 (CH2), 40.3 (CH2), 26.8 (CH3), 26.6 (CH3), 19.0 (C) ppm; IR (film, cm–1): 3069, 2933, 2855, 1475, 1431, 1261, 1113, 1002, 935, 777, 741, 702, 614; HRMS (ESI) m/z: [M + H]+ calcd. for C40H51O3Si2, 635.3371; found, 635.3370; [α]D20 + 38.0, c = 1.9, CH2Cl2.
((2R,3S,5R)-3-Acetoxy-5-allyltetrahydrofuran-2-yl)methyl acetate (29)
Purification with hexane/AcOEt (9:1) gave 22.3 mg, yield 92%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 5.82 (m, 1H), 5.07 (m, 3H), 4.18 (m, 4H), 2.25–2.45 (m, 3H), 2.08 (s, 3H), 2.06 (s, 3H), 1.75 (dt, J = 13.1, 5.3 Hz, 1H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 170.6 (C), 170.5 (C), 134.0 (CH), 117.4 (CH2), 80.8 (CH), 78.1 (CH), 75.6 (CH), 63.9 (CH2), 39.9 (CH2), 36.7 (CH2), 20.9 (CH3), 20.7 (CH3) ppm; IR (film, cm–1): 2953, 1749, 1494, 1451, 1367, 1243, 1171, 1125, 1085, 1026, 992, 896, 768, 721; HRMS (ESI) m/z: [M + H]+ calcd. for C12H19O5, 243.1227; found, 243.1230; [α]D20 + 34.7, c = 1.16, CH2Cl2.
(2R,3S,5R)-5-Allyl-3-(benzyloxy)-2-((benzyloxy)methyl)tetrahydrofuran (30)
Purification with hexane/MTBE (97:3) gave 30.1 mg, yield 89%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.57–7.10 (m, 10H), 5.86 (ddt, J = 17.2, 10.2, 7.0 Hz, 1H), 5.23–5.02 (m, 2H), 4.67–4.46 (m, 4H), 4.29–4.07 (m, 3H), 3.65–3.47 (m, 2H), 2.54 (dtt, J = 13.2, 6.5, 1.4 Hz, 1H), 2.45–2.23 (m, 2H), 1.82 (ddd, J = 12.5, 7.0, 5.2 Hz, 1H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 138.2 (C), 138.1 (C), 134.7 (CH), 128.3 (CH), 128.2 (CH), 127.5 (CH), 127.5 (CH), 127.5 (CH), 116.9 (CH2), 82.1 (CH), 80.7 (CH), 78.1 (CH), 73.3 (CH2), 71.5 (CH2), 70.7 (CH2), 40.3 (CH2), 37.2 (CH2) ppm; IR (film, cm–1): 3061, 2927, 1445, 1367, 1219, 1081, 1017, 957, 737, 703; HRMS (ESI) m/z: [M + H]+ calcd. for C22H27O3, 339.1955; found, 339.1953; [α]D20 + 39.5, c = 1.5, CH2Cl2.
(2R,3S,5R)-5-Allyl-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl 4-methylbenzenesulfonate (31)
Purification with hexane/MTBE (92:8) gave 51.1 mg, yield 93%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.75 (d, J = 8.3 Hz, 2H), 7.62 (m, 4H), 7.40 (m, 6H), 7.25 (d, J = 8.3 Hz, 2H), 5.85 (m, 1H), 5.25 (dt, J = 3.9, 0.8 Hz, 1H), 5.16–5.07 (m, 2H), 4.24 (quintet, J = 6.7 Hz, 1H), 4.16 (m, 1H), 3.60 (dd, J = 11.0, 4.0 Hz, 2H), 2.50 (s, 3H), 2.25–2.37 (m, 3H), 1.90 (m, 1H), 1.08 (s, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 144.7 (C), 135.4 (CH), 134.2 (CH), 133.6 (C), 132.9 (C), 132.8 (C), 129.8 (CH), 129.7 (CH), 129.6 (CH), 127.7 (CH), 127.6 (CH), 127.6 (CH), 117.2 (CH2), 83.35 (CH), 82.3 (CH), 78.6 (CH), 63.9 (CH2), 40.1 (CH2), 37.6 (CH2), 26.6 (CH3), 21.5 (CH3), 19.0 (C) ppm; IR (film, cm–1): 2856, 1473, 1429, 1268, 1098, 1009, 951; HRMS (ESI) m/z: [M + H]+ calcd. for C31H39O5SSi, 551.2282; found, 551.2285; [α]D20 + 34.7, c = 0.6 CH2Cl2.
(((2R,3R,5S)-5-Allyl-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl)oxy)(tert-butyl)diphenylsilane (32)
Purification with hexane/MTBE (98:2) gave 54.2 mg, yield 86%, as colorless oil; 1H NMR (300 MHz, CD2Cl2): δ 7.81–7.27 (m, 20H), 6.01–5.74 (m, 1H), 5.29–4.97 (m, 2H), 4.64–4.39 (m, 1H), 4.25–3.75 (m, 4H), 2.51 (dtt, J = 14.5, 6.5, 1.4 Hz, 1H), 2.41–2.25 (m, 1H), 2.14–1.59 (m, 2H), 1.11 (s, 9H), 1.01 (s, 9H) ppm; 13C{1H} NMR (75 MHz, CD2Cl2): δ 136.7 (CH), 136.6 (CH), 136.5 (CH), 136.5 (CH), 136.4 (CH), 136.1 (CH), 135.8 (CH), 134.8 (CH), 134.7 (CH), 134.6 (C), 134.1 (C), 134.0 (C), 130.5 (CH), 130.5 (CH), 130.4 (CH), 130.3 (CH), 130.3 (CH), 128.5 (CH), 128.4 (CH), 128.4 (CH), 128.3 (CH), 117.1 (CH2), 116.9 (CH2), 83.9 (CH), 82.5 (CH), 77.7 (CH), 74.8 (CH), 64.6 (CH2), 43.3 (CH2), 41.4 (CH2), 41.15 (CH2), 27.5 (CH3), 27.5 (CH3), 27.5 (CH3), 27.4 (CH3), 19.9 (C), 19.8 (C) ppm; IR (film, cm–1): 3068, 2934, 2853, 1473, 1435, 1262, 1112, 1000, 936, 774, 742, 701, 618; HRMS (ESI) m/z: [M + H]+ calcd. for C40H51O3Si2, 635.3371; found, 635.3375.
(2R,3R,4S,5R)-2-(Acetoxymethyl)-5-Allyltetrahydrofuran-3,4-diyl diacetate (33)
Purification with hexane/AcOEt (85:15) gave 27.9 mg, yield 93%, as pale-yellow oil; 1H NMR (300 MHz, CDCl3): δ 5.74 (ddt, J = 17.1, 10.2, 7.0 Hz, 1H), 5.46 (dd, J = 4.6, 3.4 Hz, 1H), 5.34–5.25 (m, 1H), 5.18–5.04 (m, 2H), 4.39–4.02 (m, 4H), 2.56–2.29 (m, 2H), 2.20–2.01 (m, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 170.6 (C), 169.7 (C), 169.6 (C), 132.9 (CH), 117.7 (CH2), 78.8 (CH), 76.5 (CH), 72.2 (CH), 72.0 (CH), 63.6 (CH2), 33.8 (CH2), 20.7 (CH3), 20.4 (CH3), 20.4 (CH3) ppm; IR (film, cm–1): 2957, 1745, 1494, 1447, 1367, 1243, 1101, 1083, 1047, 891, 771, 723; HRMS (ESI) m/z: [M + H]+ calcd. for C14H21O7, 301.1282; found, 301.1283; NOESY correlations (see the Supporting Information).
(2R,3R,4R,5R)-2-Allyl-3,4-bis(benzyloxy)-5-((benzyloxy)methyl)tetrahydrofuran (34)
Purification with hexane/MTBE (9:1) gave 34.2 mg, yield 77%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.46–7.19 (m, 15H), 5.94–5.72 (m, 1H), 5.24–4.97 (m, 2H), 4.69–4.32 (m, 4H), 4.17–3.99 (m, 2H), 3.99–3.79 (m, 2H), 3.73–3.46 (m, 2H), 2.52 (tq, J = 7.0, 1.4 Hz, 2H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 138.2 (C), 137.8 (C), 137.7 (C), 134.8 (CH), 134.1 (C), 128.3 (CH), 128.3 (CH), 128.3 (CH), 128.2 (CH), 128.2 (CH), 127.7 (CH), 127.6 (CH), 127.6 (CH), 127.5 (CH), 127.5 (CH), 127.4 (CH), 116.8 (CH2), 83.7 (CH), 82.7 (CH), 82.5 (CH), 80.9 (CH), 73.2 (CH2), 71.3 (CH2), 71.2 (CH2), 70.5 (CH2), 33.07 (CH2) ppm; IR (film, cm–1): 2957, 1445, 1367, 1209, 1081, 1017, 737, 703; HRMS (ESI) m/z: [M + H]+ calcd. for C29H33O4, 445.2373; found, 445.2378; NOESY correlations (see the Supporting Information).
General Procedures for the Glycosylation Scope
Procedure A
The acetal for the specific reaction (0.2 mmol) was dissolved in dry CH2Cl2 (4 mL) at room temperature under an argon atmosphere; then, the trimethylsilyl-substituted alkyl-based reagent (0.5 mmol, 2.5 equiv) was added under magnetic stirring. The resulting solution has been cooled at −50 °C, and complex 5 (15 mol %) followed by Cu(OTf)2 (5 mol %) were added in one portion. The reaction mixture was allowed to warm at −30 °C and stirred at this temperature for 12 h (except for compound 16 which was warmed up to 0 °C). The reaction was then quenched with a saturated solution of NaHCO3 (5 mL). The layers were separated, and the aqueous phase was extracted with CH2Cl2 (3 × 7 mL). The combined organic phases were dried with Na2SO4, filtered, and concentrated under reduced pressure. The resulting residue was purified by flash chromatography on silica gel.
Procedure B
The acetal for the specific reaction (0.2 mmol) was dissolved in dry CH2Cl2 (4 mL) at room temperature under an argon atmosphere; then, the trimethylsilyl-substituted aromatic reagent (0.5 mmol, 2.5 equiv) was added under magnetic stirring (except for compound 28 where furan was used in 1.5 equiv). The resulting solution has been cooled at −50 °C, and complex 5 (30 mol %) followed by Cu(OTf)2 (15 mol %) were added in one portion. The reaction mixture was allowed to warm at −30 °C and stirred at this temperature for 12 h. The reaction was then quenched with a saturated solution of NaHCO3 (5 mL). The layers were separated, and the aqueous phase was extracted with CH2Cl2 (3 × 7 mL). The combined organic phases were dried with Na2SO4, filtered, and concentrated under reduced pressure. The resulting residue was purified by flash chromatography on silica gel.
Methyl 3-(((2R,4S,5R)-4-((tert-butyldiphenylsilyl)oxy)-5-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-2-yl)methyl)but-3-enoate (13-Procedure A)
Purification with hexane/MTBE (85:15) gave 104.5 mg, yield 74%, as colorless oil; 1H NMR (400 MHz, (CD3)2CO): δ 7.67–7.47 (m, 8H), 7.43–7.25 (m, 12H), 5.62–5.46 (m, 2H), 4.49 (ddd, J = 6.6, 4.6, 3.4 Hz, 1H), 4.07–3.89 (m, 2H), 3.54 (s, 3H), 3.44 (dd, J = 11.0, 3.4 Hz, 1H), 3.32 (dd, J = 11.0, 4.0 Hz, 1H), 2.97 (dt, J = 4.2, 1.1 Hz, 2H), 2.45–2.22 (m, 2H), 2.04 (dt, J = 13.4, 6.7 Hz, 1H), 1.72 (ddd, J = 12.6, 6.4, 4.6 Hz, 1H), 1.00 (s, 9H), 0.85 (s, 9H) ppm; 13C{1H} NMR (101 MHz, (CD3)2CO: δ 171.5 (C), 135.7 (CH), 135.6 (CH), 135.5 (CH), 135.4 (CH), 133.6 (C), 133.6 (C), 133.3 (C), 130.6 (CH), 129.9 (CH), 129.9 (CH), 129.7 (CH), 129.6 (CH), 127.8 (CH), 127.7 (CH), 127.6 (CH), 124.3 (CH), 86.4 (CH), 78.5 (CH), 74.9 (CH), 64.4 (CH2), 50.9 (CH3), 40.3 (CH2), 39.3 (CH2), 37.4 (CH2), 26.5 (CH3), 26.2 (CH3), 18.8 (C), 18.7 (C) ppm; IR (film, cm–1): 2937, 1746, 1651, 1562, 1475, 1429, 1113, 1001, 826, 729; HRMS (ESI) m/z: [M + H]+ calcd. for C43H55O5Si2, 707.3583; found, 707.3581; [α]D20 + 33.6, c = 0.72, CH2Cl2.
(((2R,3S,5S)-5-(2-bromoallyl)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl)oxy)(tert-butyl)diphenylsilane (14-Procedure A)
Purification with hexane/MTBE (98:2) gave 140.9 mg, yield 99%, as colorless oil; 1H NMR (400 MHz, CDCl3): δ 7.60–7.39 (m, 8H), 7.38–7.16 (m, 12H), 5.61 (q, J = 1.2 Hz, 1H), 5.42 (d, J = 1.6 Hz, 1H), 4.44 (dt, J = 6.0, 2.9 Hz, 1H), 4.36 (tdd, J = 7.4, 5.9, 4.7 Hz, 1H), 3.97 (td, J = 3.7, 2.4 Hz, 1H), 3.38 (dd, J = 11.0, 3.8 Hz, 1H), 3.19 (dd, J = 11.0, 3.6 Hz, 1H), 2.88 (ddd, J = 14.5, 7.3, 1.1 Hz, 1H), 2.62 (ddd, J = 14.5, 5.9, 1.1 Hz, 1H), 2.08 (ddd, J = 13.4, 7.5, 6.2 Hz, 1H), 1.69 (ddd, J = 13.0, 4.8, 3.2 Hz, 1H), 0.99 (s, 9H), 0.84 (s, 9H) ppm; 13C{1H} NMR (101 MHz, CDCl3): δ 135.8 (CH), 135.8 (CH), 135.6 (CH), 135.6 (CH), 133.7 (C), 133.6 (C), 133.3 (C), 133.2 (C), 131.1 (C), 129.8 (CH), 129.6 (CH), 129.6 (CH), 127.7 (CH), 127.6 (CH), 118.5 (CH2), 87.2 (CH), 77.4 (CH), 77.1 (CH), 77.0 (CH), 76.7 (CH), 75.0 (CH), 64.5 (CH2), 48.3 (CH2), 39.9 (CH2), 27.0 (CH3), 26.8 (CH3), 19.1 (C), 19.1 (C) ppm; IR (film, cm–1): 3067, 2931, 2861, 1459, 1257, 1119, 1008, 931; HRMS (ESI) m/z: [M + H]+ calcd. for C40H50O3Si2Br, 713.2476; found, 713.2478; [α]D20 + 32.2, c = 0.68, CH2Cl2.
tert-butyl(((2R,3S,5R)-3-((tert-butyldiphenylsilyl)oxy)-5-(2-methylallyl)tetrahydrofuran-2-yl)methoxy)diphenylsilane (15-Procedure A)
Purification with hexane/MTBE (9:1) gave 112.7 mg, yield 87%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.78–7.23 (m, 20H), 4.90–4.73 (m, 2H), 4.53 (ddd, J = 6.8, 4.2, 3.0 Hz, 1H), 4.27 (p, J = 6.7 Hz, 1H), 4.07 (q, J = 3.5 Hz, 1H), 3.52 (dd, J = 11.0, 3.6 Hz, 1H), 3.43–3.23 (m, 1H), 2.54 (dd, J = 14.0, 7.2 Hz, 1H), 2.34 (dd, J = 14.1, 6.4 Hz, 1H), 2.11 (dt, J = 13.2, 6.8 Hz, 1H), 1.84–1.71 (m, 4H), 1.17–1.02 (s, 9H), 0.94 (s, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 143.1 (C), 135.7 (CH), 135.6 (CH), 135.5 (CH), 135.4 (CH), 133.8 (C), 133.7 (C), 133.3 (C), 133.2 (C), 129.6 (CH), 129.4 (CH), 129.4 (CH), 127.6 (CH), 127.5 (CH), 127.4 (CH), 111.9 (CH2), 86.5 (CH), 77.5 (CH), 74.8 (CH), 64.4 (CH2), 44.7 (CH2), 40.7 (CH2), 26.8 (CH3), 26.6 (CH3), 22.7 (C), 19.0 (C) ppm; IR (film, cm–1): 3070, 2933, 2857, 1479, 1429, 1261, 1115, 1002, 935, 702, 614; HRMS (ESI) m/z: [M + H]+ calcd. for C41H53O3Si2, 649.3528; found, 649.3529; [α]D20 + 37.4, c = 0.5, CH2Cl2; NOESY correlations (see the Supporting Information).
tert-butyl(((2R,3S,5S)-3-((tert-butyldiphenylsilyl)oxy)-5-((S)-pent-3-yn-2-yl)tetrahydrofuran-2-yl)methoxy)diphenylsilane (16-Procedure A)
Purification with hexane/MTBE (99:1) gave 108.2 mg, yield 82%, as colorless oil; 1H NMR (400 MHz, CDCl3): δ 7.55–7.30 (m, 8H), 7.29–7.05 (m, 12H), 4.29 (ddd, J = 6.1, 4.7, 3.3 Hz, 1H), 3.85 (q, J = 3.6 Hz, 1H), 3.64 (dt, J = 8.5, 6.3 Hz, 1H), 3.32 (dd, J = 11.0, 3.6 Hz, 1H), 3.22–3.14 (m, 1H), 2.59 (ddq, J = 9.0, 6.8, 2.3 Hz, 1H), 1.97–1.87 (m, 2H), 1.62 (d, J = 2.4 Hz, 3H), 1.03 (s, 9H), 0.89 (s, 9H) ppm; 13C{1H} NMR (101 MHz, CDCl3): δ 135.9 (CH), 135.8 (CH), 135.8 (CH), 135.7 (CH), 135.6 (CH), 135.6 (CH), 135.5 (CH), 133.9 (C), 133.8 (C), 133.5 (C), 133.4 (C), 129.8 (CH), 129.7 (CH), 129.5 (CH), 129.5 (CH), 129.4 (CH), 127.7 (CH), 127.6 (CH), 127.6 (CH), 127.6 (CH), 86.9 (CH), 82.9 (C), 81.3 (C), 74.5 (CH), 64.4 (CH2), 39.4 (CH), 32.3 (CH3), 26.8 (CH3), 26.6 (CH3), 19.2 (C), 19.1 (C), 3.7 (CH3) ppm; IR (film, cm–1): 3071, 2923, 2857, 1447, 1477, 1425, 1263, 1112, 1003, 937, 706; HRMS (ESI) m/z: [M + H]+ calcd. for C42H53O3Si2, 661.3528; found, 661.3526.
(((2R,3S,5S)-5-Azido-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl)oxy)(tert-butyl)diphenylsilane (17-Procedure A)
Purification with hexane/MTBE (7:3) gave 114.3 mg, yield 90%, as pale-yellow oil; 1H NMR (300 MHz, CDCl3): δ 7.82–7.22 (m, 20H), 5.49 (dd, J = 6.4, 1.7 Hz, 1H), 4.47 (ddt, J = 6.9, 4.7, 2.8 Hz, 1H), 4.29 (q, J = 3.0 Hz, 1H), 3.63–3.41 (m, 1H), 3.26 (dd, J = 11.3, 3.3 Hz, 1H), 2.32–1.94 (m, 2H), 1.10 (s, 9H), 0.98 (s, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 135.7 (CH), 135.7 (CH), 135.6 (CH), 135.4 (CH), 135.4 (CH), 133.42 (C), 133.4 (C), 133.0 (C), 132.9 (C), 129.8 (CH), 129.7 (CH), 129.6 (CH), 127.7 (CH), 127.6 (CH), 91.9 (CH), 89.0 (CH), 73.4 (CH), 63.7 (CH2), 26.8 (CH3), 26.6 (CH3), 19.0 (C), 18.9 (C) ppm; IR (film, cm–1): 3071, 2937, 2859, 2137, 1487, 1427, 1261, 1111, 1009, 931, 703; HRMS (ESI) m/z: [M + H]+ calcd. for C37H46O3N3Si2, 636.3072; found, 636.3075.
tert-butyl(((2R,3S)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl)oxy)diphenylsilane (18-Procedure A)
Purification with hexane/MTBE (99:1) gave 114.1 mg, yield 96%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.89–7.53 (m, 8H), 7.49–7.22 (m, 12H), 4.47 (td, J = 3.9, 2.0 Hz, 1H), 4.14–3.92 (m, 3H), 3.42 (qd, J = 10.9, 4.3 Hz, 2H), 1.87 (dtd, J = 8.7, 4.6, 1.4 Hz, 2H), 1.11 (s, 9H), 0.99 (s, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 135.7 (CH), 135.5 (CH), 135.4 (CH), 133.9 (C), 133.7 (C), 133.3 (C), 133.3 (C), 129.6 (CH), 129.5 (CH), 129.4 (CH), 127.6 (CH), 127.6 (CH), 127.5 (CH), 127.5 (CH), 87.1 (CH), 75 (CH), 67.5 (CH2), 64.5 (CH2), 35.6 (CH2), 26.9 (CH3), 26.7 (CH3), 19.1 (C), 19.0 (C) ppm; IR (film, cm–1): 3067, 2939, 2861, 1485, 1421, 1264, 1112, 1011, 927, 705; HRMS (ESI) m/z: [M + H]+ calcd. for C37H47O3Si2, 595.3058; found, 595.3059; [α]D20 + 23.7, c = 1.7, CH2Cl2.
tert-butyl(((2R,3S,5R)-3-((tert-butyldiphenylsilyl)oxy)-5-(phenylthio)tetrahydrofuran-2-yl)methoxy)diphenylsilane (19-Procedure A)
Purification with hexane/AcOEt (99:1) gave 134.8 mg, yield 96%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.80–7.15 (m, 25H), 5.68 (dd, J = 7.5, 4.1 Hz, 1H), 4.47 (dt, J = 8.0, 4.0 Hz, 1H), 4.40–4.27 (m, 1H), 3.70 (dd, J = 11.3, 2.7 Hz, 1H), 3.54–3.41 (m, 1H), 2.49 (dt, J = 14.1, 7.2 Hz, 1H), 2.12 (dt, J = 13.6, 4.0 Hz, 1H), 1.15 (s, 9H), 0.98 (s, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 137.0 (C), 135.8 (CH), 135.7 (CH), 135.6 (CH), 135.6 (CH), 135.5 (CH), 135.5 (CH), 135.5 (CH), 135.1 (C), 133.6* (C), 133.6* (C), 133.5* (C), 133.4* (C), 133.3* (C), 133.2* (C), 133.1* (C), 133.1* (C), 131.2* (CH), 130.4 (CH), 129.7 (CH), 129.7* (CH), 129.5* (C), 129.5 (CH), 128.7* (CH), 128.7 (CH), 127.7 (CH), 127.6 (CH), 127.5* (CH), 127.5* (CH), 127.5* (CH), 126.8* (CH), 126.4 (CH), 88.8* (CH), 87.2 (CH), 86.2 (CH), 85.8* (CH), 74.3* (CH), 72.8 (CH), 64.1* (CH2), 63.4 (CH2), 42.5 (CH2), 41.3* (CH2), 29.6 (CH2), 26.8 (CH3), 26.7* (CH3), 26.6 (CH3), 19.1 (C), 19.1 (C) ppm; IR (film, cm–1): 3134, 3071, 3050, 3015, 2957, 2931, 2893, 2857, 2773, 2739, 2710, 2316, 1960, 1891, 1825, 1775, 1731, 1660, 1587, 1472, 1427, 1390, 1256, 1112, 702, 612, 505, 436, 423, 413; HRMS (ESI) m/z: [M + H]+ calcd. for C43H51O3SSi2, 703.3092; found, 703.3094.
Methyl 2-((2R,4S,5R)-4-((tert-butyldiphenylsilyl)oxy)-5-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-2-yl)-2-methylpropanoate (20-Procedure A)
Purification with hexane/MTBE (85:15) gave 102.7 mg, yield 74%, as colorless oil; 1H NMR (400 MHz, CDCl3): δ 7.60–7.42 (m, 8H), 7.38–7.16 (m, 12H), 4.44 (ddd, J = 7.1, 6.2, 4.6 Hz, 1H), 4.16 (dd, J = 9.3, 6.8 Hz, 1H), 3.85–3.78 (m, 1H), 3.56 (s, 3H), 3.54–3.48 (m, 1H), 3.31–3.22 (m, 1H), 1.92 (dt, J = 12.6, 7.0 Hz, 1H), 1.72 (ddd, J = 12.6, 9.3, 6.2 Hz, 1H), 1.19 (s, 3H), 1.08 (s, 3H), 0.98 (d, J = 9.4 Hz, 9H), 0.84 (d, J = 5.1 Hz, 9H) ppm; 13C{1H} NMR (101 MHz, CDCl3): δ 177.0 (C), 135.8 (CH), 135.8 (CH), 135.8 (CH), 135.7 (CH), 135.6 (CH), 135.6 (CH), 133.8 (C), 133.7 (C), 133.6 (C), 133.4 (C), 129.7 (CH), 129.5 (CH), 129.5 (CH), 127.7 (CH), 127.7 (CH), 127.7 (CH), 127.6 (CH), 127.59, 127.6 (CH), 86.4 (CH), 83.0 (CH), 73.8 (CH), 64.2 (CH2), 51.8 (CH3), 46.0 (C), 36.9 (CH2), 26.9 (CH3), 26.7 (CH3), 21.1 (CH3), 20.5 (CH3), 19.1 (C), 19.0 (C) ppm; IR (film, cm–1): 2935, 1748, 1653, 1561, 1475, 1429, 1113, 824, 738; HRMS (ESI) m/z: [M + H]+ calcd. for C42H55O5Si2, 695.3583; found, 695.3585.
2-((2S,4S,5R)-4-((tert-butyldiphenylsilyl)oxy)-5-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-2-yl)-1-phenylethan-1-one (21-Procedure A)
Purification with hexane/AcOEt (9:1) gave 95.4 mg, yield 67%, as colorless oil; 1H NMR (400 MHz, CDCl3): δ 7.96–7.87 (m, 2H), 7.61–7.11 (m, 23H), 4.70–4.59 (m, 1H), 4.48–4.41 (m, 1H), 4.04 (td, J = 4.0, 2.2 Hz, 1H), 3.54 (dd, J = 16.4, 6.2 Hz, 1H), 3.33 (dd, J = 11.0, 4.2 Hz, 1H), 3.27 (d, J = 7.1 Hz, 1H), 3.25–3.18 (m, 1H), 2.21 (ddd, J = 13.4, 7.6, 6.1 Hz, 1H), 1.75 (ddd, J = 13.1, 4.4, 2.9 Hz, 1H), 0.98 (s, 9H), 0.83 (s, 9H) ppm; 13C{1H} NMR (101 MHz, CDCl3): δ 198.9 (C), 198.4* (C), 137.3 (C), 137.2* (C), 135.8* (CH), 135.8 (CH), 135.6 (CH), 135.6 (CH), 135.6 (CH), 133.9* (C), 133.9* (C), 133.60 (C), 133.4 (C), 133.3 (C), 133.1 (CH), 129.8 (CH), 129.7 (CH), 129.6* (CH), 129.6 (CH), 128.6 (CH), 128.3 (CH), 128.3 (CH), 127.8 (CH), 127.8 (CH), 127.7 (CH), 127.6 (CH), 87.9* (CH), 87.2 (CH), 75.9 (CH), 75.6* (CH), 75.4 (CH), 75.2* (CH), 64.5 (CH2), 64.3* (CH2), 45.7 (CH2), 44.5* (CH2), 42.0* (CH2), 40.8 (CH2), 27.0 (CH3), 26.8 (CH3), 26.8 (CH3), 19.2* (C), 19.1 (C), 19.1 (C) ppm; IR (film, cm–1): 3070, 3050, 2998, 2957, 2931, 2893, 2857, 1961, 1897, 1823, 1684, 1597, 1472, 1448, 1428, 1389, 1361, 1306, 1263, 1208, 1112, 1039, 999, 938, 822, 741, 702, 612, 506, 446, 413; HRMS (ESI) m/z: [M + H]+ calcd. for C45H53O4Si2, 713.3477; found, 713.3474.
1-((2S,4S,5R)-4-((tert-butyldiphenylsilyl)oxy)-5-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-2-yl)propan-2-one (22-Procedure A)
Purification with hexane/AcOEt (9:1) gave 101.5 mg, yield 78%, as colorless oil; 1H NMR (400 MHz, CDCl3): δ 7.61–7.14 (m, 20H), 4.47–4.33 (m, 2H), 3.97 (q, J = 3.7 Hz, 1H), 3.35 (dd, J = 11.0, 3.9 Hz, 1H), 3.20 (dd, J = 11.0, 3.9 Hz, 1H), 2.93 (dd, J = 16.0, 7.3 Hz, 1H), 2.66 (dd, J = 16.0, 6.0 Hz, 1H), 2.17–2.04 (m, 4H), 1.63 (ddd, J = 13.0, 5.0, 3.5 Hz, 1H), 0.99 (s, 9H), 0.84 (s, 9H) ppm; 13C{1H} NMR (101 MHz, CDCl3): δ 207.8 (C), 135.8 (CH), 135.8 (CH), 135.6 (CH), 135.6 (CH), 135.5 (CH), 133.6 (C), 133.5 (C), 133.3 (C), 133.2 (C), 129.8 (CH), 129.8 (CH), 129.6 (CH), 127.8 (CH), 127.7 (CH), 127.6 (CH), 87.0 (CH), 75.3 (CH), 75.0 (CH), 64.4 (CH2), 50.6 (CH2), 40.8 (CH2), 30.8 (CH3), 27.0 (CH3), 26.7 (CH3), 19.1 (C), 19.1 (C) ppm; IR (film, cm–1): 3134, 3071, 3049, 2998, 2958, 2931, 2893, 2858, 2709, 1961, 1892, 1825, 1715, 1589, 1472, 1428, 1360, 1242, 1189, 1111, 1007, 936, 823, 741, 703, 612, 506, 443; HRMS (ESI) m/z: [M + H]+ calcd. for C40H51O4Si2, 651.3320; found, 651.3324.
(S)-2-((2R,4S,5R)-4-((tert-butyldiphenylsilyl)oxy)-5-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-2-yl)-2-((trimethylsilyl)oxy)cyclobutan-1-one (23-Procedure A)
Purification with hexane/MTBE (94:6) gave 132.0 mg, yield 88%, as colorless oil; 1H NMR (400 MHz, CD2Cl2): δ 7.56–7.13 (m, 20H), 5.21–5.14 (m, 1H), 4.36 (ddd, J = 7.1, 5.7, 4.2 Hz, 1H), 3.92–3.84 (m, 1H), 3.40 (dd, J = 11.2, 2.8 Hz, 1H), 3.19 (dd, J = 11.2, 3.9 Hz, 1H), 2.78–2.45 (m, 3H), 2.03–1.91 (m, 1H), 1.90–1.79 (m, 2H), 0.94 (s, 9H), 0.79 (s, 9H), 0.00 (s, 9H) ppm; 13C{1H} NMR (101 MHz, CD2Cl2): δ 210.8 (C), 209.8* (C), 135.5* (CH), 135.5 (CH), 135.4 (CH), 135.4* (CH), 135.3* (CH), 135.2 (CH), 135.2 (CH), 135.2* (CH), 135.1* (CH), 133.4* (C), 133.3 (C), 133.3 (C), 133.2* (C), 133.0 (C), 132.9 (C), 129.5 (CH), 129.4* (CH), 129.3 (CH), 129.2 (CH), 129.2 (CH), 129.2 (CH), 127.4 (CH), 127.4 (CH), 127.3* (CH), 127.3* (CH), 127.3 (CH), 127.3 (CH), 127.2* (CH), 125.1* (CH), 93.8 (C), 93.2* (C), 87.8* (CH), 87.7* (CH), 87.0 (CH), 85.6* (CH), 80.9* (CH), 80.3* (CH), 79.6 (CH), 79.0* (CH), 74.4* (CH), 73.3 (CH), 72.6* (CH), 64.1 (CH2), 63.6* (CH2), 40.2 (CH2), 40.1* (CH2), 36.6 (CH2), 36.1* (CH2), 33.8* (CH), 29.8 (CH), 29.4* (CH), 26.4 (CH3), 26.4* (CH3), 26.2* (CH3), 26.2 (CH3), 23.8 (CH2), 23.3* (CH2), 20.5* (CH3), 18.7 (C), 18.6 (C), 0.9 (CH3) ppm; IR (film, cm–1): 3640, 3135, 3071, 3050, 2999, 2957, 2931, 2894, 2858, 2710, 1960, 1891, 1789, 1661, 1589, 1472, 1428, 1389, 1362, 1308, 1252, 1180, 1111, 1008, 980, 937, 844, 740, 702, 612, 504, 421; HRMS (ESI) m/z: [M + H]+ calcd. for C44H59O5Si3, 751.3665; found, 751.3668; NOESY correlations (see the Supporting Information).
(R)-2-((2R,4S,5R)-4-((tert-butyldiphenylsilyl)oxy)-5-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-2-yl)cyclohexan-1-one (24-Procedure A)
Purification with hexane/AcOEt (9:1) gave 114.5 mg, yield 83%, as colorless oil; 1H NMR (400 MHz, CDCl3): δ 7.56–7.40 (m, 8H), 7.35–7.18 (m, 12H), 4.37 (dt, J = 6.4, 3.3 Hz, 1H), 4.10 (ddd, J = 9.1, 7.2, 5.2 Hz, 1H), 3.96–3.83 (m, 1H), 3.33 (dd, J = 10.9, 4.0 Hz, 1H), 3.21 (dd, J = 11.0, 4.2 Hz, 1H), 2.74 (ddd, J = 11.8, 9.2, 5.4 Hz, 1H), 2.43 (dt, J = 13.4, 3.9 Hz, 1H), 2.31–2.20 (m, 3H), 2.00 (d, J = 9.2 Hz, 1H), 1.81 (dp, J = 18.9, 6.4, 5.2 Hz, 2H), 1.73–1.53 (m, 2H), 1.41 (dtd, J = 22.6, 11.2, 10.2, 4.5 Hz, 1H), 0.97 (s, 9H), 0.84 (s, 9H) ppm; 13C{1H} NMR (101 MHz, CDCl3): δ 212.8 (C), 135.9* (CH), 135.8 (CH), 135.8 (CH), 135.8* (CH), 135.7* (CH), 135.6 (CH), 133.7 (C), 133.6 (C), 133.4 (C), 133.4 (C), 129.8 (CH), 129.8 (CH), 129.5 (CH), 129.5 (CH), 127.7 (CH), 127.7* (CH), 127.6* (CH), 127.6 (CH), 127.6 (CH), 86.4 (CH), 78.0 (CH), 74.9 (CH), 64.4 (CH2), 56.8 (CH), 42.8 (CH2), 42.0 (CH2), 40.5 (CH2), 31.4 (CH2), 28.4 (CH2), 27.0 (CH2), 27.0 (CH3), 26.7 (CH3), 25.0* (CH2), 24.9 (CH2), 19.1 (C), 19.1 (C) ppm; IR (film, cm–1): 3071, 3049, 2956, 2931, 2893, 2858, 2739, 2710, 1960, 1891, 1825, 1709, 1589, 1472, 1428, 1390, 1264, 1112, 999, 970, 823, 740, 703, 612, 506; HRMS (ESI) m/z: [M + H]+ calcd. for C43H55O4Si2, 691.3633; found, 691.3630.
(R)-5-((2R,4S,5R)-4-((tert-butyldiphenylsilyl)oxy)-5-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-2-yl)furan-2(5H)-one (25-Procedure B)
Purification with hexane/AcOEt (9:1) gave 74.3 mg, yield 55%, as colorless oil; 1H NMR (400 MHz, CDCl3): δ 7.68 (dd, J = 5.7, 1.5 Hz, 1H), 7.60–7.17 (m, 20H), 6.10 (dd, J = 5.7, 1.9 Hz, 1H), 5.17 (dt, J = 8.5, 1.7 Hz, 1H), 4.39 (td, J = 4.1, 1.8 Hz, 1H), 4.03 (tt, J = 3.8, 2.1 Hz, 1H), 3.78 (dt, J = 8.6, 5.6 Hz, 1H), 3.31 (dd, J = 11.1, 4.0 Hz, 1H), 3.13 (dd, J = 11.1, 3.7 Hz, 1H), 2.20–2.10 (m, 2H), 0.99 (s, 9H), 0.83 (s, 9H) ppm; 13C{1H} NMR (101 MHz, CDCl3): δ 173.2 (C), 156.2 (CH), 135.8 (CH), 135.7 (CH), 135.5 (CH), 135.5 (CH), 135.5 (CH), 133.4 (C), 133.1 (C), 133.1 (C), 133.0 (CH), 130.0 (CH), 129.9 (CH), 129.7 (CH), 127.9 (CH), 127.8 (CH), 127.7 (CH), 127.6 (CH), 121.7 (CH), 88.4 (CH), 85.0 (CH), 80.6 (CH), 75.0 (CH), 64.4 (CH2), 38.4 (CH2), 27.0 (CH3), 26.7 (CH3), 19.1 (C), 19.0 (C) ppm; IR (film, cm–1): 3071, 3049, 2956, 2931, 2894, 2858, 1787, 1760, 1589, 1472, 1428, 1390, 1362, 1155, 1113, 1046, 1007, 950, 886, 822, 703, 613, 506, 442, 410; HRMS (ESI) m/z: [M + H]+ calcd. for C41H49O5Si2, 677.3113; found 677.3117; NOESY correlations (see the Supporting Information).
tert-butyl(((2R,3S,5R)-3-((tert-butyldiphenylsilyl)oxy)-5-(4-methoxyphenyl)tetrahydrofuran-2-yl)methoxy)diphenylsilane (26-Procedure B)
Purification with hexane/MTBE (93:7) gave 99.4 mg, yield 71%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.79–7.20 (m, 22H), 6.88–6.79 (m, 2H), 5.25 (dd, J = 11.0, 4.7 Hz, 1H), 4.61 (d, J = 5.2 Hz, 1H), 4.16 (td, J = 3.9, 1.5 Hz, 1H), 3.81 (s, 3H), 3.60 (dd, J = 11.0, 4.0 Hz, 1H), 3.51–3.33 (m, 1H), 2.22–2.10 (m, 1H), 1.85 (ddd, J = 12.7, 11.0, 5.2 Hz, 1H), 1.14 (s, 9H), 0.99 (d, J = 5.1 Hz, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 158.9 (C), 135.7 (CH), 135.5 (CH), 135.5 (CH), 133.9 (C), 133.8 (C), 133.7 (C), 133.2 (C), 133.1 (C), 129.6 (CH), 129.5 (CH), 129.4 (CH), 127.6 (CH), 127.6 (CH), 127.5 (CH), 127.5 (CH), 127.5 (CH), 113.5 (CH), 88.0 (CH), 80.0 (CH), 75.72 (CH), 64.4 (CH2), 55.2 (CH3), 44.6 (CH2), 26.9 (CH3), 26.67 (CH3), 19.1 (C) ppm; IR (film, cm–1): 3067, 2929, 2861, 1491, 1247, 1117, 1019, 931, 827, 734; HRMS (ESI) m/z: [M + H]+ calcd. for C44H53O4Si2, 701.3477; found, 701.3479; NOESY correlations (see the Supporting Information).
(((2R,3S,5R)-5-(benzo[d][1,3]dioxol-5-yl)-2-(((tert-butyldiphenylsilyl)oxy)methyl)tetrahydrofuran-3-yl)oxy)(tert-butyl)diphenylsilane (27-Procedure B)
Purification with hexane/MTBE (92:8) gave 100.0 mg, yield 70%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.76–7.53 (m, 8H), 7.52–7.26 (m, 12H), 7.08–6.71 (m, 3H), 5.96 (dd, J = 10.3, 1.3 Hz, 2H), 5.28–4.97 (m, 1H), 4.74–4.49 (m, 1H), 4.31–4.11 (m, 1H), 3.72–3.54 (m, 1H), 3.52–3.38 (m, 1H), 2.51–1.72 (m, 2H), 1.10 (dd, J = 23.2, 1.3 Hz, 8H), 0.99 (dd, J = 4.4, 1.3 Hz, 9H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 147.5 (C), 146.7 (C), 135.8 (CH), 135.7 (CH), 135.6 (CH), 135.5 (CH), 135.4 (CH), 133.8 (C), 133.7 (C), 133.1 (CH), 133.0 (CH), 129.6 (CH), 129.5 (CH), 129.5 (CH), 129.4 (CH), 127.6 (CH), 127.6 (CH), 127.5 (CH), 119.5 (CH), 107.8 (CH), 106.7 (CH), 100.8 (CH2), 88.1 (CH), 80.2 (CH), 75.7 (CH), 64.4 (CH2), 44.7 (CH2), 26.9 (CH3), 26.7 (CH3), 19.0 (C) ppm; IR (film, cm–1): 3071, 2921, 2857, 1502, 1481, 1463, 1233, 1115, 1092, 1041, 939, 739; HRMS (ESI) m/z: [M + H]+ calcd. for C44H51O5Si2, 715.3270; found, 715.3269; NOESY correlations (see the Supporting Information).
tert-Butyl(((2R,3S,5R)-3-((tert-butyldiphenylsilyl)oxy)-5-(furan-2-yl)tetrahydrofuran-2-yl)methoxy)diphenylsilane (28-Procedure B)
Purification with hexane/MTBE (9:1) gave 75.4 mg, yield 57%, as colorless oil; 1H NMR (400 MHz, (CD3)2CO): δ 7.78–7.29 (m, 21H), 6.39–6.30 (m, 2H), 5.25 (dd, J = 10.4, 5.4 Hz, 1H), 4.74–4.63 (m, 1H), 4.12 (td, J = 4.3, 1.7 Hz, 1H), 3.54–3.46 (m, 1H), 3.41 (dd, J = 10.9, 4.1 Hz, 1H), 2.25–2.09 (m, 2H), 1.13 (s, 9H), 0.95 (s, 9H) ppm; 13C{1H} NMR (101 MHz, (CD3)2CO): δ 154.1 (C), 142.4 (CH), 135.7 (CH), 135.7 (CH), 135.7 (CH), 135.5 (CH), 135.5 (CH), 135.4 (CH), 133.6 (C), 133.5 (C), 133.1 (C), 133.1 (C), 130.0 (CH), 129.9 (CH), 129.7 (CH), 129.6 (CH), 127.9 (CH), 127.7 (CH), 127.6 (CH), 110.1 (CH), 107.3 (CH), 87.8 (CH), 75.3 (CH), 73.2 (CH), 64.3 (CH2), 40.0 (CH2), 26.5 (CH3), 26.4 (CH3), 18.76 (C) ppm; IR (film, cm–1): 3067, 2939, 2861, 1693, 1483, 1421, 1382, 1167, 1061, 992, 745; HRMS (ESI) m/z: [M + H]+ calcd. for C41H49O4Si2, 661.3164; found, 661.3166.
(2R,3R,4S,5S)-2-(Acetoxymethyl)-5-(2-methoxyphenyl)tetrahydrofuran-3,4-diyl diacetate (35-Procedure B)
Purification with hexane/AcOEt (8:2) gave 57.8 mg, yield 79%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.52 (ddd, J = 7.6, 1.8, 0.8 Hz, 1H), 7.32–7.20 (m, 1H), 6.97 (td, J = 7.5, 1.1 Hz, 1H), 6.81 (dd, J = 8.2, 1.0 Hz, 1H), 5.83 (dd, J = 4.4, 3.1 Hz, 1H), 5.56 (d, J = 3.1 Hz, 1H), 5.45 (dd, J = 8.1, 4.5 Hz, 1H), 4.51–4.37 (m, 2H), 4.29–4.17 (m, 1H), 3.80 (s, 3H), 2.23–1.99 (m, 6H), 1.77 (s, 3H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 170.7 (C), 169.7 (C), 169.3 (C), 155.7 (C), 128.6 (CH), 127.2 (CH), 124.1 (C), 119.9 (CH), 109.5 (CH), 76.8 (CH), 76.5 (CH), 72.6 (CH), 72.0 (CH), 63.6 (CH2), 55.1 (CH3), 20.8 (CH3), 20.4 (CH3), 20.2 (CH3) ppm; IR (film, cm–1): 2961, 1743, 1497, 1446, 1367, 1245, 1043, 1047, 891, 701; HRMS (ESI) m/z: [M + H]+ calcd. for C18H23O8, 367.1387; found, 367.1384; NOESY correlations (see the Supporting Information). The spectroscopic data are in agreement with those reported in the literature26b but refer to a single anomeric compound.
Allylation Scope on Different Heterocyclic Cores
The procedure herein adopted is the same as described for the previous allylation reactions.
1-Allyl-1,3-dihydroisobenzofuran (37)
Purification with hexane/MTBE (99:1) gave 27.6 mg, yield 86%, as colorless oil; the spectroscopic data correspond with those reported in the literature (See the Supporting Information for 1H NMR spectra).34
Methyl 2-allylpyrrolidine-1-carboxylate (39)
Purification with hexane/MTBE (78:22) gave 30.4 mg, yield 90%, as colorless oil; the spectroscopic data correspond with those reported in the literature (See the Supporting Information for 1H NMR spectra).35
(2S,4S)-2-Allyl-4-benzyl-3-tosyloxazolidine (41)36
Purification with hexane/MTBE (7:3) gave 45.0 mg, yield 63%, as colorless oil; 1H NMR (300 MHz, CDCl3): δ 7.83–7.70 (m, 2H), 7.42–7.14 (m, 7H), 5.89 (ddt, J = 17.2, 10.1, 7.0 Hz, 1H), 5.30–5.13 (m, 2H), 5.00 (dd, J = 7.5, 3.2 Hz, 1H), 3.84 (ddq, J = 10.1, 7.1, 3.6 Hz, 1H), 3.74 (dd, J = 9.1, 3.1 Hz, 1H), 3.28–3.09 (m, 2H), 2.91–2.65 (m, 2H), 2.57–2.36 (m, 4H) ppm; 13C{1H} NMR (75 MHz, CDCl3): δ 144.1 (C), 137.3 (C), 134.1 (C), 132.3 (CH), 129.8 (CH), 129.7 (CH), 129.4 (CH), 129.2 (CH), 128.6 (CH), 128.6 (CH), 128.5 (CH), 127.7 (CH), 126.7 (CH), 118.4 (CH2), 91.81, 68.8 (CH2), 60.50, 40.9 (CH2), 40.46 (CH2), 21.43 (CH3) ppm; IR (film, cm–1): 2985, 2923, 2885, 2254, 1598, 1456, 1337, 1232, 1157, 1093, 904, 725; HRMS (ESI) m/z: [M + H]+ calcd. for C20H24NO3S, 358.1471; found, 358.1475; [α]D20 – 65.0, c = 0.55 in CH2Cl2; NOESY correlations (see the Supporting Information).
Large-Scale Synthesis
The allylation reaction of 11 to give 12 was tested on a larger scale to prove the scale up of the synthesis. Specifically, an identical procedure for the synthesis of 12 was followed, using 750 mg of the starting acetal 11 (1.2 mmol) and the corresponding percentages of Re catalyst 8 (15 mol %, 0.18 mmol, 117 mg) and Cu(OTf)2 (5 mol %, 0.06 mmol, 22 mg). Purification with hexane/MTBE (98:2) gave 700 mg of product 12 (1.1 mmol) with an isolated yield 92%, as colorless oil. NMR spectra correspond with those obtained for the sub-millimolar scale.
Computational Methods
All structures studied during the DFT investigation of the anisole reaction mechanism were optimized with the Gaussian 09 program package.37 Previously reported studies on the allylation reaction of five-membered ring oxocarbenium ions, by Woerpel and co-workers, showed that the use of B3LYP could results in difficulties in locating the transitions states when allyltrimethylsilane was involved in the reaction, thus suggesting the use of M06-2X.38 Considering the different types of reactants involved, we decided to use the more parametrized B3LYP functional anyway, by specifying a differentiated basis set.39 This resulted in the absence of any problem in locating the corresponding TSs. Specifically, C and H atoms were described using the 6-31G(d,p) basis set, whereas O and Si atoms were described with the 6-31+G(d,p) basis set.40 Vibrational frequencies were computed at the same level of theory to verify that the optimized structures were minima or TSs. For TS, IRC calculations were computed at the same level of theory in order to confirm the structures at the ends. Solvent effects were then evaluated by running single-point calculations at the same level of theory, the polarizable continuum solvent model (PCM) was used for CH2Cl2. NPA charges on C1 (furanoside numbering system) are calculated in CH2Cl2 at the same level of theory. TMS was used to model the presence of a TBDPS-protecting group.41
Acknowledgments
We thank Regione Lombardia - project VIPCAT (Value Added Innovative Protocols for Catalytic Transformations - CUP: E46D17000110009) and the Italian MIUR (funds PRIN 2017) for financial support, Professor Mariella Mella for the NMR spectra measurement, and Professors Giovanni Vidari and Lucio Toma for helpful discussions.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.1c00706.
1H, 13C{1H} NMR spectra for all new compounds, NOESY correlations and spectra, computational details, summary tables of geometrical features, Cartesian coordinates of all computed structures (PDF).
The authors declare no competing financial interest.
This paper was published ASAP on May 25, 2021 with an an incorrect surname for author Sirwan T. Othman. The corrected version was reposted on June 1, 2021.
Supplementary Material
References
- a Furanoside numbering system;; b 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. 10.1021/cr3004778. [DOI] [PubMed] [Google Scholar]; c Fernandes R. A.; Gorve D. A.; Pathare R. S. Emergence of 2,3,5-trisubstituted tetrahydrofuran natural products and their synthesis. Org. Biomol. Chem. 2020, 18, 7002–7025. 10.1039/d0ob01542c. [DOI] [PubMed] [Google Scholar]; d de la Torre A.; Cuyamendous C.; Bultel-Poncé V.; Durand T.; Galano J.-M.; Oger C. Recent advances in the synthesis of tetrahydrofurans and applications in total synthesis. Tetrahedron 2016, 72, 5003–5025. 10.1016/j.tet.2016.06.076. [DOI] [Google Scholar]; e Fernandes R. A.; Pathare R. S.; Gorve D. A. Advances in Total Synthesis of Some 2,3,5-Trisubstituted Tetrahydrofuran Natural Products. Chem.—Asian J. 2020, 15, 2815–2837. 10.1002/asia.202000753. [DOI] [PubMed] [Google Scholar]; f Lu Q.; Harmalkar D. S.; Choi Y.; Lee K. An Overview of Saturated Cyclic Ethers: Biological Profiles and Synthetic Strategies. Molecules 2019, 24, 3778. 10.3390/molecules24203778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Yamamoto Y.; Asao N. Selective reactions using allylic metals. Chem. Rev. 1993, 93, 2207–2293. 10.1021/cr00022a010. [DOI] [Google Scholar]; b Colvin E. W.Allylsilanes. Silicon in Organic Synthesis; London: Butterworth, 1981, p 97. [Google Scholar]; c Hosomi A. Characteristics in the reactions of allylsilanes and their applications to versatile synthetic equivalents. Acc. Chem. Res. 1988, 21, 200–206. 10.1021/ar00149a004. [DOI] [Google Scholar]; d Langkopf E.; Schinzer D. Uses of Silicon-Containing Compounds in the Synthesis of Natural Products. Rev 1995, 95, 1375–1408. 10.1021/cr00037a011. [DOI] [Google Scholar]; e Harmange J.-C.; Figadère B. Synthetic routes to 2,5-disubstituted tetrahydrofurans. Tetrahedron: Asymmetry 1993, 4, 1711–1754. 10.1016/s0957-4166(00)80408-5. [DOI] [Google Scholar]; f Lade J. J.; Pardeshi S. D.; Vadagaonkar K. S.; Murugan K.; Chaskar A. C. The remarkable journey of catalysts from stoichiometric to catalytic quantity for allyltrimethylsilane inspired allylation of acetals, ketals, aldehydes and ketones. RSC Adv. 2017, 7, 8011–8033. 10.1039/c6ra27813b. [DOI] [Google Scholar]
- Yadav J. S.; Reddy B. V. S.; Reddy A. S.; Reddy C. S.; Raju S. S. Highly diastereoselective allylation of lactols and their ethers using molecular iodine. Tetrahedron Lett. 2009, 50, 6631–6634. 10.1016/j.tetlet.2009.09.025. [DOI] [Google Scholar]
- Friestad G. K.; Lee H. J. Trans-2,5-Disubstituted Tetrahydrofurans via Addition of Carbon Nucleophiles to the Strained Bicyclic Acetal 2,7-Dioxabicyclo[2.2.1]heptane. Org. Lett. 2009, 11, 3958–3961. 10.1021/ol901613k. [DOI] [PubMed] [Google Scholar]
- a Bugoni S.; Porta A.; Valiullina Z.; Zanoni G.; Vidari G. Dual ReVCatalysis in One-Pot Consecutive Meyer-Schuster and Diels-Alder Reactions. Eur. J. Org. Chem. 2016, 2016, 4900–4906. 10.1002/ejoc.201600926. [DOI] [Google Scholar]; b Mattia E.; Porta A.; Merlini V.; Zanoni G.; Vidari G. One-Pot Consecutive Reactions Based on the Synthesis of Conjugated Enones by the Re-Catalysed Meyer-Schuster Rearrangement. Chem.—Eur. J. 2012, 18, 11894–11898. 10.1002/chem.201201639. [DOI] [PubMed] [Google Scholar]; c Stefanoni M.; Luparia M.; Porta A.; Zanoni G.; Vidari G. A Simple and Versatile Re-Catalyzed Meyer-Schuster Rearrangement of Propargylic Alcohols to α,β-Unsaturated Carbonyl Compounds. Chem.—Eur. J. 2009, 15, 3940–3944. 10.1002/chem.200802622. [DOI] [PubMed] [Google Scholar]; d Complex 8 can be purchased directly from the vendors or easily synthesized following the procedure ofAbu-Omar M. M.; Khan S. I. Molecular Rhenium(V) Oxotransferases: Oxidation of Thiols to Disulfides with Sulfoxides. The Case of Substrate-Inhibited Catalysis. Inorg. Chem. 1998, 37, 4979–4985. 10.1021/ic980348j. [DOI] [PubMed] [Google Scholar]
- Palm U.; Mosandl A.; Bensch W. Stereoisomeric flavour compounds XLV: Structure analysis of 2-alkoxy-5-alkyl-tetrahydrofurans. Chirality 1991, 3, 76–83. 10.1002/chir.530030113. [DOI] [Google Scholar]
- Rhenium catalyst 8 was used in the synthesis of glycosides by Toste and subsequently by Zhu. However, the peculiar mechanism of Nu-H activation via a protonated Re-oxo species cannot be operative in our scenario. For more details, see:; a Sherry B. D.; Loy R. N.; Toste F. D. Rhenium(V)-Catalyzed Synthesis of 2-Deoxy-α-glycosides. J. Am. Chem. Soc. 2004, 126, 4510–4511. 10.1021/ja031895t. [DOI] [PubMed] [Google Scholar]; b Baryal K. N.; Adhikari S.; Zhu J. Catalytic Stereoselective Synthesis of β-Digitoxosides: Direct Synthesis of Digitoxin and C1′-epi-Digitoxin. J. Org. Chem. 2013, 78, 12469–12476. 10.1021/jo4021419. [DOI] [PubMed] [Google Scholar]
- Data not reported. For literature example, see:; a Larsen C. H.; Ridgway B. H.; Shaw J. T.; Smith D. M.; Woerpel K. A. StereoselectiveC-Glycosylation Reactions of Ribose Derivatives: Electronic Effects of Five-Membered Ring Oxocarbenium Ions. J. Am. Chem. Soc. 2005, 127, 10879–10884. 10.1021/ja0524043. [DOI] [PubMed] [Google Scholar]; b Shenoy S. R.; Woerpel K. A. Investigations into the Role of Ion Pairing in Reactions of Heteroatom-Substituted Cyclic Oxocarbenium Ions. Org. Lett. 2005, 7, 1157–1160. 10.1021/ol0500620. [DOI] [PubMed] [Google Scholar]
- a Pan F.; Shi Z.-J. Recent Advances in Transition-Metal-Catalyzed C-S Activation: From Thioester to (Hetero)aryl Thioether. ACS Catal. 2014, 4, 280–288. 10.1021/cs400985m. [DOI] [Google Scholar]; b for Cu(II) DMS complexes, see:; c Sigel H.; Scheller K. H.; Rheinberger V. M.; Fischer B. E. Comparison of the ligating properties of disulphides and thioethers: dimethyl disulphide, dimethyl sulphide, and related ligands. J. Chem. Soc., Dalton Trans. 1980, 1022–1028. 10.1039/dt9800001022. [DOI] [Google Scholar]; d Black J. R.; Levason W. Synthesis and solution multinuclear magnetic resonance studies of homoleptic copper(I) complexes of sulfur, selenium and tellurium donor ligands. J. Chem. Soc., Dalton Trans. 1994, 3225–3230. 10.1039/dt9940003225. [DOI] [Google Scholar]; e Knorr M.; Bonnot A.; Lapprand A.; Khatyr A.; Strohmann C.; Kubicki M. M.; Rousselin Y.; Harvey P. D. Reactivity of CuI and CuBr toward Dialkyl Sulfides RSR: From Discrete Molecular Cu4I4S4and Cu8I8S6Clusters to Luminescent Copper(I) Coordination Polymers. Inorg. Chem. 2015, 54, 4076–4093. 10.1021/acs.inorgchem.5b00327. [DOI] [PubMed] [Google Scholar]
- For Cu(I) DMS complexes, see:Liebeskind L. S.; Srogl J. Thiol Ester–Boronic Acid Coupling. A Mechanistically Unprecedented and General Ketone Synthesis. J. Am. Chem. Soc. 2000, 122, 11260–11261. 10.1021/ja005613q. [DOI] [Google Scholar]
- For a quantitative scale of oxophilicity and thiophilicity, see:; a Kepp K. P. A Quantitative Scale of Oxophilicity and Thiophilicity. Inorg. Chem. 2016, 55, 9461–9470. for a comparison between Cu(II) and Co(II) see: 10.1021/acs.inorgchem.6b01702. [DOI] [PubMed] [Google Scholar]; b Dickerson T. J.; Reed N. N.; LaClair J. J.; Janda K. D. A Precipitator for the Detection of Thiophilic Metals in Aqua. J. Am. Chem. Soc. 2004, 126, 16582–16586. 10.1021/ja045798r. [DOI] [PubMed] [Google Scholar]
- Shomaker J. M.; Borhan B. Total Synthesis of Haterumalides NA and NC via a Chromium-Mediated Macrocyclization. J. Am. Chem. Soc. 2008, 130, 12228–12229. 10.1021/ja8043695. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Tran V. T.; Woerpel K. A. Nucleophilic Addition to Silyl-Protected Five-Membered Ring Oxocarbenium Ions Governed by Stereoelectronic Effects. J. Org. Chem. 2013, 78, 6609–6621. 10.1021/jo400945j. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Gil A.; Lorente A.; Albericio F.; Álvarez M. Stereoselective Allylstannane Addition for a Convergent Synthesis of a Complex Molecule. Lett 2015, 17, 6246–6249. 10.1021/acs.orglett.5b03252. [DOI] [PubMed] [Google Scholar]
- a Nyberg K.; Servin R. Large-scale Laboratory Electrolysis in Organic Systems. III. The Synthesis of alpha-Methoxyalkylamides. Cyclic Acylimmonium Precursors. Acta Chem. Scand., Ser. B 1976, 30, 640–648. 10.3891/acta.chem.scand.30b-0640. [DOI] [Google Scholar]; b Shono T. Electroorganic chemistry in organic synthesis. Tetrahedron 1984, 40, 811–850. 10.1016/s0040-4020(01)91472-3. [DOI] [Google Scholar]
- a Beaver G. M.; Buscagan T. M.; Lavinda O.; Woerpel K. A. Stereoelectronic Model to Explain Highly Stereoselective Reactions of Seven-membered Ring Oxocarbenium Ion Intermediates. Angew. Chem., Int. Ed. 2016, 55, 1816–1819. 10.1002/anie.201507806. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Larsen C. H.; Ridgway B. H.; Shaw J. T.; Woerpel K. A. A Stereoelectronic Model To Explain the Highly Stereoselective Reactions of Nucleophiles with Five-Membered-Ring Oxocarbenium Ions. J. Am. Chem. Soc. 1999, 121, 12208–12209. 10.1021/ja993349z. [DOI] [Google Scholar]
- a Buckle M. J. C.; Fleming I.; Gil S.; Pang K. L. C. Accurate determinations of the extent to which the SE2′ reactions of allyl-, allenyl- and propargylsilanes are stereospecifically. Org. Biomol. Chem. 2004, 2, 749–769. 10.1039/b313446f. [DOI] [PubMed] [Google Scholar]; b Marshall J. A.; Maxson K. Stereoselective Synthesis of Stereotriad Subunits of Polyketides through Additions of Nonracemic Allenylsilanes to (R)- and (S)-2-Methyl-3-oxygenated Propanals. J. Org. Chem. 2000, 65, 630–633. 10.1021/jo991543y. [DOI] [PubMed] [Google Scholar]; c Brawn R. A.; Panek J. S. Stereoselective C-Glycosidations with Achiral and Enantioenriched Allenylsilanes. Org. Lett. 2010, 12, 4624–4627. 10.1021/ol1019629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Bennek J. A.; Gray G. R. An efficient synthesis of anhydroalditols and allylic-glycosides. J. Org. Chem. 1987, 52, 892–897. 10.1021/jo00381a030. [DOI] [Google Scholar]; b Sassaman B. M.; Surya Prakash G. K.; Olah G. A.; Donald P.; Loker K. B. Ionic Hydrogenation with Organosilanes under Acid-Free Conditions. Synthesis of Ethers, Alkoxysilanes, Thioethers, and Cyclic Ethers via rganosilyl Iodide and Triflate Catalyzed Reductions of Carbonyl Compounds and Their Derivatives. Tetrahedron 1988, 44, 3771–3780. 10.1016/s0040-4020(01)86635-7. [DOI] [Google Scholar]
- Yuh-ichiro I.; Hideki K.; Ken’ichi F.; Tatsuo O.; Koichi N. Stereoselective β-C- and β-S-Glycosylation of 2-Deoxyribofuranose Derivatives Controlled by the 3-Hydroxy Protective Group. Bull. Chem. Soc. Jpn. 1989, 62, 845–852. 10.1246/bcsj.62.845. [DOI] [Google Scholar]
- Chen X.; Wang Q.; Yu B. Gold(I)-catalyzed C-glycosylation of glycosyl ortho-alkynylbenzoates: the role of the moisture sequestered by molecular sieves. Chem. Commun. 2016, 52, 12183–12186. 10.1039/c6cc07218f. [DOI] [PubMed] [Google Scholar]
- Figadère B.; Peyrat J.-F.; Cavé A. Replicative Chirons: Stereoselective Synthesis of Oligo-Tetrahydrofuranic Lactones via C-Glycosylation with [(Trimethylsilyl)oxy]furan. J. Org. Chem. 1997, 62, 3428–3429. 10.1021/jo970212n. [DOI] [Google Scholar]
- Structure of 25 has been established by comparison of the 1H-chemical shifts of literature related compounds and NOE experiments:Zanardi F.; Battistini L.; Rassu G.; Auzzas L.; Pinna L.; Marzocchi L.; Acquotti D.; Casiraghi G. Modular Approach toward the Construction of the Core Motifs of Annonaceous Acetogenins and Variants Thereof. J. Org. Chem. 2000, 65, 2048–2064. 10.1021/jo991568x. [DOI] [PubMed] [Google Scholar]
- For a review on aryl-C-glycoside, see:; a Kitamura K.; Ando Y.; Matsumoto T.; Suzuki K. Total Synthesis of Aryl C-Glycoside Natural Products: Strategies and Tactics. Chem. Rev 2018, 118, 1495–1598. 10.1021/acs.chemrev.7b00380. [DOI] [PubMed] [Google Scholar]; b Štambasky J.; Hocek M.; Kočovsky P. C-Nucleosides: Synthetic Strategies and Biological Applications. Chem. Rev. 2009, 109, 6729–6764. 10.1021/cr9002165. [DOI] [PubMed] [Google Scholar]; c Chaudhuri N. C.; Ren R. X.-F.; Kool E. T. C-Nucleosides Derived from Simple Aromatic HydrocarbonsSelected examples of aryl-C-glycoside preparation via Friedel-Crafts reaction can be found here. Synlett 1997, 4, 341–347. 10.1055/s-1997-788. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Hainke S.; Arndt S.; Seitz O. Concise synthesis of aryl-C-nucleosides by Friedel–Crafts alkylation. Org. Biomol. Chem. 2005, 3, 4233–4238. 10.1039/b509846g. [DOI] [PubMed] [Google Scholar]; e Spadafora M.; Mehiri M.; Burger A.; Benhida R. Friedel–Crafts and modified Vorbrüggen ribosylation. A short synthesis of aryl and heteroaryl-C-nucleosides. Tetrahedron Lett. 2008, 49, 3967–3971. 10.1016/j.tetlet.2008.04.105. [DOI] [Google Scholar]; f Ohrui H.; Kuzuhara H.; Emoto S. Facile Syntheses of C-Glycosides of Aromatic Compounds. Agric. Biol. Chem. 1972, 36, 1651–1653. 10.1080/00021369.1972.10860455. [DOI] [Google Scholar]; g Ren R. X.-F.; Chaudhuri N. C.; Paris P. L.; Rumney S.; Kool E. T. Naphthalene, Phenanthrene, and Pyrene as DNA Base Analogues: Synthesis, Structure, and Fluorescence in DNA. J. Am. Chem. Soc. 1996, 118, 7671–7678. 10.1021/ja9612763. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Jiang Y. L.; Stivers J. T. Efficient epimerization of pyrene and other aromatic C-nucleosides with trifluoroacetic acid in dichloromethane. Tetrahedron Lett. 2003, 44, 85–88. 10.1016/s0040-4039(02)02481-4. [DOI] [Google Scholar]; b Jiang Y. L.; Stivers J. T. Novel epimerization of aromatic C-nucleosides with electron-withdrawing substituents with trifluoroacetic acid–benzenesulfonic acid using mild conditions. Tetrahedron Lett. 2003, 44, 4051–4055. 10.1016/s0040-4039(03)00848-7. [DOI] [Google Scholar]
- Diederichsen U.; Biro C. M. Phe and Asn side chains in DNA double strands. Bioorg. Med. Chem. Lett. 2000, 10, 1417–1420. 10.1016/s0960-894x(00)00231-6. [DOI] [PubMed] [Google Scholar]
- Martinez H. O.; Reinke H.; Michalik D.; Vogel C. Peracetylated β-Allyl C-Glycosides of D-Ribofuranose and 2-Deoxy-D-ribofuranose in the Chemical Literature: Until Now, Mirages in the Literature. Synthesis 2009, 11, 1834–1840. 10.1055/s-0029-1216794. [DOI] [Google Scholar]
- a Marzag H.; Alaoui S.; Amdouni H.; Martin A. R.; Bougrin K.; Benhida R. Efficient and selective azidation of per-O-acetylated sugars using ultrasound activation: application to the one-pot synthesis of 1,2,3-triazole glycosides. New J. Chem. 2015, 39, 5437–5444. 10.1039/c5nj00624d. [DOI] [Google Scholar]; b Tachallait H.; Filho M. S.; Marzag H.; Bougrin K.; Demange L.; Martin A. R.; Benhida R. A straightforward and versatile FeCl3 catalyzed Friedel–Crafts C-glycosylation process. Application to the synthesis of new functionalized C-nucleosides. New J. Chem. 2019, 43, 5551–5558. 10.1039/c8nj06300a. [DOI] [Google Scholar]
- a Chen L.; Zhu Y.; Kong F. Synthesis of α-Manp-(1→2)-α-Manp-(1→3)-α-Manp-(1→3)-Manp, the tetrasaccharide repeating unit of Escherichia coli O9a, and α-Manp-(1→2)-α-Manp-(1→2)-α-Manp-(1→3)-α-Manp-(1→3)-Manp, the pentasaccharide repeating unit of E. coli O9 and Klebsiella O3. Carbohydr. Res. 2002, 337, 383–390. 10.1016/s0008-6215(02)00011-3. [DOI] [PubMed] [Google Scholar]; b Bélot F.; Wright K.; Costachel C.; Phalipon A.; Mulard L. A. Blockwise Approach to Fragments of the O-Specific Polysaccharide of Shigella flexneri Serotype 2a: Convergent Synthesis of a Decasaccharide Representative of a Dimer of the Branched Repeating Unit. J. Org. Chem. 2004, 69, 1060–1074. 10.1021/jo035125b. [DOI] [PubMed] [Google Scholar]
- Boto A.; Hernández R.; Suárez E. Tandem Radical Decarboxylation–Oxidation of Amino Acids: A Mild and Efficient Method for the Generation of N-Acyliminium Ions and Their Nucleophilic Trapping. J. Org. Chem. 2000, 65, 4930–4937. 10.1021/jo000356t. [DOI] [PubMed] [Google Scholar]
- a Moulins J. R.; Hughes J. A.; Doyle L. E.; Cameron T. S.; Burnell D. J. Two Rearrangement Pathways in the Geminal Acylation of 2-Methoxyoxazolidines Leading to Substituted 1,4-Oxazines. Eur. J. Org. Chem. 2015, 6, 1325–1332. 10.1002/ejoc.201403371. [DOI] [Google Scholar]; b Conde-Friebose K.; Hoppe D. Synthesis of Enantiomerically Pure 2,4-trans-Substituted 3-Arenesulfonyl-1,3-oxazolidines by Lewis Acid Mediated Substitution of the 2-Methoxy Derivatives and their Epimerization. Synlett 1990, 2, 99–102. 10.1055/s-1990-21000. [DOI] [Google Scholar]
- Kim J.; Gil J. M.; Greenberg M. M. Synthesis and Characterization of Oligonucleotides Containing the C4′-Oxidized Abasic Site Produced by Bleomycin and Other DNA. Damaging Agents. Angew. Chem., Int. Ed. 2003, 42, 5882–5885. 10.1002/anie.200352102. [DOI] [PubMed] [Google Scholar]
- Ludek O. R.; Marquez V. E. A greener enantioselective synthesis of the antiviral agent North-methanocarbathymidine (N-MCT) from 2-deoxy-d-ribose. Tetrahedron 2009, 65, 8461–8467. 10.1016/j.tet.2009.08.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taverna-Porro M.; Bouvier L. A.; Pereira C. A.; Montserrat J. M.; Iribarren A. M. Chemoenzymatic preparation of nucleosides from furanoses. Tetrahedron Lett. 2008, 49, 2642–2645. 10.1016/j.tetlet.2008.02.087. [DOI] [Google Scholar]
- Zeng J.; Vedachalam S.; Xiang S.; Liu X.-W. Direct C-Glycosylation of Organotrifluoroborates with Glycosyl Fluorides and Its Application to the Total Synthesis of (+)-Varitriol. Org. Lett 2011, 13, 42–45. 10.1021/ol102473k. [DOI] [PubMed] [Google Scholar]
- Asai S.; Yabe Y.; Goto R.; Nagata S.; Monguchi Y.; Kita Y.; Sajiki H.; Sawama Y. Gold-Catalyzed Benzylic Azidation of Phthalans and Isochromans and Subsequent FeCl3-Catalyzed Nucleophilic Substitutions. Chem. Pharm. Bull. 2015, 63, 757–761. 10.1248/cpb.c15-00347. [DOI] [PubMed] [Google Scholar]
- Yoshida J.-I.; Suga S.; Suzuki S.; Kinomura N.; Yamamoto A.; Fujiwara K. Direct Oxidative Carbon–Carbon Bond Formation Using the “Cation Pool” Method. 1. Generation of Iminium Cation Pools and Their Reaction with Carbon Nucleophiles. J. Am. Chem. Soc. 1999, 121, 9546–9549. 10.1021/ja9920112. [DOI] [Google Scholar]
- Synthesis of precursor 40 was performed in accordance with the procedure described in ref (29a).
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Keith T.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam J. M.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas O.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian 09, Revision B.01; Gaussian, Inc.: Wallingford, CT, 2010.
- Lavinda O.; Tran V. T.; Woerpel K. A. Effect of Conformational Rigidity on the Stereoselectivity of Nucleophilic Additions to Five-membered Ring Bicyclic Oxocarbenium Ion Intermediates. Org. Biomol. Chem. 2014, 12, 7083–7091. 10.1039/c4ob01251h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Becke A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. 10.1063/1.464913. [DOI] [Google Scholar]; b Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. 10.1103/physrevb.37.785. [DOI] [PubMed] [Google Scholar]
- a Hehre W. J.; Ditchfield R.; Pople J. A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar]; b Hariharan P. C.; Pople J. A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. 10.1007/bf00533485. [DOI] [Google Scholar]; c Hehre W. J.; Ditchfield R.; Pople J. A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257–2261. 10.1063/1.1677527. [DOI] [Google Scholar]; d Francl M. M.; Pietro W. J.; Hehre W. J.; Binkley J. S.; Gordon M. S.; DeFrees D. J.; Pople J. A. Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements. J. Chem. Phys. 1982, 77, 3654–3665. 10.1063/1.444267. [DOI] [Google Scholar]; e Gordon M. S.; Binkley J. S.; Pople J. A.; Pietro W. J.; Hehre W. J. Self-consistent molecular-orbital methods. 22. Small split-valence basis sets for second-row elements. J. Am. Chem. Soc. 1982, 104, 2797–2803. 10.1021/ja00374a017. [DOI] [Google Scholar]; f Spitznagel G. W.; Clark T.; Schleyer P.; Hehre W. J. An evaluation of the performance of diffuse function-augmented basis sets for second row elements, Na-Cl. J. Comput. Chem. 1987, 8, 1109–1116. 10.1002/jcc.540080807. [DOI] [Google Scholar]
- All the information related to the geometrical features of the species involved in the reaction can be found in the Supporting Information.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







