Next Article in Journal
Ultrasound-Assisted Preparation of Maillard Reaction Products Derived from Hydrolyzed Soybean Meal with Meaty Flavor in an Oil-In-Water System
Next Article in Special Issue
Recent Advances on Natural Aryl-C-glycoside Scaffolds: Structure, Bioactivities, and Synthesis—A Comprehensive Review
Previous Article in Journal
Recent Research Progress on Natural Stilbenes in Dendrobium Species
Previous Article in Special Issue
Not Just Anticoagulation—New and Old Applications of Heparin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Acid Catalyzed Stereocontrolled Ferrier-Type Glycosylation Assisted by Perfluorinated Solvent

1
College of Chemistry and Chemical Engineering and Henan Key Laboratory of Function-Oriented Porous Materials, Luoyang Normal University, Luoyang 471934, China
2
Hubei Key Laboratory of Natural Products Research and Development, Key Laboratory of Functional Yeast (China National Light Industry), College of Biological and Pharmaceutical Sciences, China Three Gorges University, Yichang 443002, China
3
Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2022, 27(21), 7234; https://doi.org/10.3390/molecules27217234
Submission received: 24 September 2022 / Revised: 15 October 2022 / Accepted: 18 October 2022 / Published: 25 October 2022
(This article belongs to the Special Issue Carbohydrate-Based Drugs Discovery)

Abstract

:
Described herein is the first application of perfluorinated solvent in the stereoselective formation of O-/S-glycosidic linkages that occurs via a Ferrier rearrangement of acetylated glycals. In this system, the weak interactions between perfluoro-n-hexane and substrates could augment the reactivity and stereocontrol. The initiation of transformation requires only an extremely low loading of resin-H+ and the mild conditions enable the accommodation of a broad spectrum of glycal donors and acceptors. The ‘green’ feature of this chemistry is demonstrated by low toxicity and easy recovery of the medium, as well as operational simplicity in product isolation.

Graphical Abstract

1. Introduction

Facile and stereoselective construction of glycosidic linkages has always been one of the major focal points in the carbohydrate research community. Among these, the 2,3-unsaturated O-glycosides have attracted great attention because of their wide occurrence in bioactive molecules (Figure 1a) [1,2] and the potential for rapid functionalization [3,4]. Over the past several decades, various efficient methods have been established for forging such core scaffolds with a Ferrier rearrangement [5,6] that employs readily accessible glycals, and O-nucleophiles emerging as the most robust strategy [7,8,9,10]. Owing to the mild conditions and short reaction times, Lewis acids are the catalyst class of choice to promote this type of transformation [7,8,9,10], while Brønsted acids [7,8,9,10] and transition metal catalysts were also found to be effective [7,8,9,10,11,12]. Alternatively, a Ferrier-type O-glycosylation could be mediated by single-electron transfer reagents [13,14] via a radical pathway. These developments notwithstanding, a predominant α-selectivity in the formation of O-glycosidic linkages is normally dictated by multiple factors, including the conformation of glycal, anomeric effect, as well as the solvent effect in most cases [10,15,16] (Figure 1b).
In this context, palladium-catalyzed O-glycosylation with glycals as donors offers complementary and more programmable access by which excellent stereocontrol could be governed through the rational selection of the leaving group [17,18], ligand [19], or palladium source [20]. In this paradigm, tactics such as the addition of zinc reagent to render a softer acceptor [17,19], modification of glycal to activate the donor [21,22], or application of decarboxylative pathway to formally activate both reactants [23,24] are invoked to improve the performance of these reactions (Figure 1c). A review of these systems suggested that by incorporating a catalyst that could bring the donor and acceptor together through noncovalent interactions, the reaction might be catalytically mediated via a stereoselective manifold. Inspired by the recent advance in stereoselective O-glycosylation by means of bifunctional H-bond catalysis with O-acceptor [25], we envisioned devising a novel catalytic system to mimic this activation mode with other less-explored weak interactions.
Perfluorinated hydrocarbons displaying low chemical activity, low toxicity and low miscibility with common organic solvents have been recognized as a class of useful reaction mediums in various research fields [26,27], particularly in molecular-oxygen-involved aerobic oxidation reactions [28,29,30,31,32,33,34]. Wide application potential is also found in biphase catalysis by virtue of their unique physical properties [35,36]. Moreover, fluorous solvents could engage in diverse weak interactions such as π–πF, C–F···H hydrogen bond, C–F···C=O, and anion-πF, which play essential roles in the promotion of chemical transformations by enhancing reactivity and stereoselectivity as well as the design of functional materials [26,27,37,38,39,40,41]. In carbohydrate chemistry, it has been found that introducing a perfluorinated solvent could improve the reaction outcome [42,43,44,45]. These findings led us to postulate that the weak interactions stemming from perfluorinated solvent could be leveraged to improve the acid-catalyzed Ferrier-type glycosylation reaction (Figure 1d). On account of the weak acidic condition compared to traditional acid-catalyzed Ferrier rearrangement, the translation of this design into an effective process would further enable stereocontrol and broadens the substrate generality. Notably, the use of perfluorinated solvent could additionally ease the isolation of the glycoside products and the recovery of the reaction medium and assistor.

2. Results and Discussion

2.1. Optimization of Reaction Conditions

Based on these design criteria, the study on this stereocontrolled glycosylation commenced by employing tri-O-acetylated glucal 1a as the donor, while ethanol 2a serves as both the acceptor and solvent (Table 1). TFE (trifluoroethanol) was first attempted as the additive, which might promote glycosylation through acidic proton or/and other noncovalent weak interactions with 2a [46]. Encouragingly, the O-glycosidic product 3a was provided in 45% yield after 6 h at 100 °C (entry 1). The use of PFD (1H,1H,2H,2H-perfluoro-1-decanol) with a longer perfluorinated alkyl chain improved the yield to 55%, indicating the dominant role of the fluorine effect (entry 2). This speculation was further corroborated by the enhanced chemical yield when PFH (perfluoro-n-hexane) without an acidic proton was used as the catalyst (entry 3). Nonetheless, a significant decrease in conversion was observed when PFTEA (perfluoro-triethylamine) [47] was utilized, implying that the basic environment could retard the progress of this transformation (entry 4). It should be noted that high α-selectivity was detected for the generated O-glycosidic product for all these reactions (α:β > 20:1). Unsurprisingly, less than 10% yield and poor stereoselectivity (α:β = 5:1) was obtained in the absence of additive (entry 5). These results illustrated the positive effect of weak interactions on both efficiency and stereocontrol. As more complex glycosyl acceptors may not be accessed as easily and well-suited for use in solvent quantities, the reaction using stoichiometric glycosyl acceptors was evaluated in PFH due to the environmental friendliness and recyclability. However, under this set of conditions, only a trace amount of 3a was detected (entry 6). Exogenous proton was introduced, and notably, 0.6 wt% of H+ type sulfonic resin (resin-H+) was sufficient to deliver a quantitative amount of glycosylated α-3a (entry 7). Meanwhile, when CH2Cl2 was used as the solvent, low yield (16%) and poor stereoselectivity (α:β =1.5:1) were delivered (entry 8). Similarly, the stereoselectivity was decreased (α:β = 7:1) when PFH was substituted by ethanol (entry 9), and no 3a was obtained with less amount of PFH (10%) and n-hexane as a solvent, further affirming our hypothesis (entry 10). Other solvents were also screened, but no satisfactory results could be observed (entries 11–13). Lowering the temperature to 80 °C led to appreciable erosion of chemical yield (entry 14), whereas a prolonged reaction time of 14 h led again to a good yield (entry 15). A trace amount of 3a was detected when the temperature was further decreased to 60 °C (entry 16). The absolute configuration of 3a was determined by X-ray crystallographic analysis.

2.2. Substrate Scope

With the optimized conditions in hand, the substrate generality with respect to glycosyl acceptors was evaluated using glucal 1a as the standard donor. As depicted in Scheme 1a, various types of glycosyl acceptors, including alkyl, allyl, benzyl, and propargyl alcohols, could give the desired glycosidic products in excellent yield with high stereocontrol at the anomeric center (3b-3o, α:β > 20:1). It is noteworthy that sterically hindered (3f and 3j) and structurally rigid (3o) alcohols that are unreactive reactants for conventional Ferrier rearrangement approaches could convert efficiently to respective O-glycosylation products. Subsequently, phenols with different substituents and substitution patterns were examined, and the glycosidic 3p-3ac was synthesized smoothly (Scheme 1b). Compared to aliphatic alcohol acceptors, the yields and stereoselectivities deteriorated in most cases, probably due to the strong background reaction catalyzed by an acidic hydroxyl group of phenols. Apart from O-nucleophiles, S-nucleophiles were also applicable for this reaction (Scheme 1c). Although all the tested substrates reacted well with 1a to give compounds 4a-4e in good yields, the stereochemical outcome varied greatly. For instance, a 1:1 α:β mixture was detected for 4a (from n-butylthiol) while 4b (from t-butylthiol) was generated with an α:β ratio > 20:1. Likewise, thiophenol with electron-withdrawing group delivered S-glycosidic 4c in poor stereocontrol while 4d with an electron-donating group on thiophenol was obtained with α:β ratio of 10:1. When 2-methylbenzenethiol was utilized, the desired glycosylation product 4e was formed in 75% yield with 6:1 α:β selectivity. Additionally, C-3 substitution products 4c’ and 4e’ were isolated alongside 6% and 8% yields, respectively. The absolute configurations of 3aa, 3ab, 4e, and 4e’ were determined by X-ray crystallographic analysis, and those of other products in this scheme were assigned by analogy. Water also functioned well as an acceptor in the developed reaction, giving α-5 an 87% yield.
Subsequently, the generality of this glycosylation method was studied with other types of glycal donors (Scheme 2). Firstly, d-galactal 1b, C-4 epimer of 1a was employed, and the results were summarized in Scheme 2a. A series of alcohols were examined, and these reactions invariably gave only 6a-6e in excellent yields and α:β > 20:1. Phenols, thiols, and thiophenols were also applicable to afford 6f-6i in good yields and stereoselectivities. As a C-3 epimer of 1a, the combination of d-allal 1c with selected glycosyl acceptors forged the corresponding products in more than 80% yield (3a, 3aa, 4b, and 4c). Interestingly, remarkable α-selectivities were detected for all of these reactions, same with the case for glucal 1a (Scheme 2b). l-Rhamnal 1d was also verified to be a competent donor for this transformation, and 7a-7d was established with excellent outcomes (Scheme 2c). However, when the pentose substrates were employed in this procedure, such as D-xylal 1e or d-arabinal 1f (a pair of C-3 epimers) as glycosyl donors, poor α:β ratios were observed for these reactions (Scheme S1, 8a-8d), indicating the direct significance of C-5 substitution in stereoinduction.
To demonstrate the practicality of the developed glycosylation strategy, the reactions of 1a with an array of functional molecules as acceptors were investigated (Scheme 3a). First, glycosylated product 9a with a long alkyl chain was prepared in 90% chemical yield with α:β > 20:1, indicating the potential utility in lipidosome assembly. A fluorous tag containing long-chain linear perfluorocarbon was well tolerated to afford 9b with the same level of outcome. Glycosylation with sugar alcohol delivered disaccharide 9c in 80% yield with α:β selectivity of 12:1. When phenol derived from tetraphenylethylene with aggregation-induced emission attribute was reacted, 9d could be generated in moderate yield with α:β = 9:1. Furthermore, the reaction operated smoothly on bioactive diosgenin to generate the C-O bond formation product 9e with perfect stereochemical control.
A gram-scale reaction between 1a and 2a was also implemented under the standard conditions, in which the synthetic efficiency and stereocontrol observed for the small-scale reaction were perfectly preserved (Scheme S2). Additionally, given the ease of isolation and good recyclability of organofluorine solvent, the recycling experiments were conducted to reinforce the utility of this strategy. After the completion of each reaction, the target product was easily isolated by phase separation, and the recovered reaction system (bottom phase) was reused successively. As summarized in Scheme 3b, when ethanol 2a was used to react with donor 1a, the stereoselectivity (α:β > 20:1) was perfectly preserved, and the chemical yield was maintained at a good level (>70%) even after a repetition of this procedure for seven times. Similar results were obtained by using 3,4-dimethylphenol 2q as a glycosyl acceptor for the recycling experiment.

3. Materials and Methods

The detailed procedure of the synthesis and characterization of the products are given in Supplementary Materials.

4. Conclusions

In conclusion, an acid-catalyzed stereoselective Ferrier-type glycosylation assisted by perfluorinated solvent has been established. A wide range of glycal donors and glycosyl acceptors are well accommodated to provide structurally diverse O- and S-glycosylated linkages products in good efficiency for most cases. The utilization of perfluoro-n-hexane as the solvent improves the reaction conditions, increases the yield, and enhances the stereocontrol at the anomeric center. Notably, the turnover of this procedure is achieved with a minimal amount of resin-H+. Aside from experimental ease in isolating products, the use of low toxic and recyclable perfluorinated solvent highlights the environmental friendliness of the developed method.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27217234/s1, CCDC 2132603, 2160183, 2160185, 2160188, and 2161131 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif (accessed on 22 March 2022), by emailing da-ta_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033, synthesis and characterization of all compounds described in this paper. References [48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68] are cited in the supplementary materials.

Author Contributions

Conceptualization, Z.L. and S.X.; methodology, Z.L.; software, Y.L.; validation, Y.L. and M.Z.; formal analysis, Y.L., M.Z., Y.S., X.J., R.J. and Y.W.; investigation, Y.L., M.Z., Y.S., X.J., R.J. and Y.W.; resources, Y.L. and Y.S.; data curation, Y.L.; writing—original draft preparation, Z.L. and S.X.; writing—review and editing, Z.L., S.X. and Y.F.; visualization, Y.L.; supervision, Y.F.; project administration, Z.L.; funding acquisition, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

We thank the financial support from Shenzhen Special Funds (JCYJ20180305123508258), the National Natural Science Foundation of China (U1304206, 21801112), the Natural Science Foundation of Henan Province (212300410374), the Science and Technology Project of Henan Province (212102210549), and the Key Scientific Research Project of Higher Education of Henan Province (18A150012, 19A150003, 19A150004, 13A150799).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Polkowski, K.; Popiołkiewicz, J.; Krzeczyński, P.; Ramza, J.; Pucko, W.; Zegrocka-Stendel, O.; Boryski, J.; Skierski, J.S.; Ma-zurek, A.P.; Grynkiewicz, G. Cytostatic and Cytotoxic Activity of Synthetic Genistein Glycosides Against Human Cancer Cell Lines. Cancer Lett. 2004, 203, 59–69. [Google Scholar] [CrossRef] [PubMed]
  2. Popiołkiewicz, J.; Polkowski, K.; Skierski, J.S.; Mazurek, A.P. In Vitro Toxicity Evaluation in the Development of New Anticancer Drugs–Genistein Glycosides. Cancer Lett. 2005, 229, 67–75. [Google Scholar] [CrossRef] [PubMed]
  3. Babu, R.S.; Zhou, M.; O’Doherty, G.A. De Novo Synthesis of Oligosaccharides Using a Palladium-Catalyzed Glycosylation Reaction. J. Am. Chem. Soc. 2004, 126, 3428–3429. [Google Scholar] [CrossRef] [PubMed]
  4. Ghosh, A.K.; Veitschegger, A.M.; Nie, S.; Relitti, N.; MacRae, A.J.; Jurica, M.S. Enantioselective Synthesis of Thailanstatin A Methyl Ester and Evaluation of in Vitro Splicing Inhibition. J. Org. Chem. 2018, 83, 5187–5198. [Google Scholar] [CrossRef] [PubMed]
  5. Ferrier, R.J. Unsaturated Carbohydrates. Part II. Three Reactions Leading to Unsaturated Glycopyranosides. J. Chem. Soc. 1964, 5443–5449. [Google Scholar] [CrossRef]
  6. Ferrier, R.J.; Prasad, N. Unsaturated Carbohydrates. Part XI. Isomerisation and Dimerisation of Tri-O-acetyl-D-glucal. J. Chem. Soc. C 1969, 581–586. [Google Scholar] [CrossRef]
  7. Gómez, A.M.; Lobo, F.; Uriel, C.; López, J.C. Recent Developments in the Ferrier Rearrangement. Eur. J. Org. Chem. 2013, 32, 7221–7262. [Google Scholar] [CrossRef] [Green Version]
  8. Gómez, A.M.; Lobo, F.; Miranda, S.; López, J.C. A Survey of Recent Synthetic Applications of 2,3-Dideoxy-Hex-2-enopyranosides. Molecules 2015, 20, 8357–8394. [Google Scholar] [CrossRef] [Green Version]
  9. Jiang, N.; Wu, Z.; Dong, Y.; Xu, X.; Liu, X.; Zhang, J. Progress in the Synthesis of 2,3-Unsaturated Glycosides. Curr. Org. Chem. 2020, 24, 184–199. [Google Scholar] [CrossRef]
  10. Bennett, C.S.; Galan, M.C. Methods for 2-Deoxyglycoside Synthesis. Chem. Rev. 2018, 118, 7931–7985. [Google Scholar] [CrossRef]
  11. McKay, M.J.; Nguyen, H.M. Recent Advances in Transition Metal-Catalyzed Glycosylation. ACS Catal. 2012, 2, 1563–1595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bauer, E.B. Transition Metal Catalyzed Glycosylation Reactions—An Overview. Org. Biomol. Chem. 2020, 18, 9160–9180. [Google Scholar] [CrossRef] [PubMed]
  13. Gómez, A.M.; Valverde, S.; Fraser-Reid, B. A Route to Unsaturated Spiroketals from Phenylthio Hex-2-enopyranosides via Sequential alkylation, Allylic Rearrangement and Intramolecular Glycosidation. J. Chem. Soc. Chem. Commun. 1991, 1207–1208. [Google Scholar] [CrossRef]
  14. Rafiee, E.; Tangestaninejad, S.; Habibi, M.H.; Mirkhani, V. A Mild, Efficient and α-Selective Glycosidation by Using Potassium Dodecatungstocobaltate Trihydrate as Catalyst. Bioorg. Med. Chem. Lett. 2004, 14, 3611–3614. [Google Scholar] [CrossRef]
  15. Leng, W.-L.; Yao, H.; He, J.-X.; Liu, X.-W. Venturing beyond Donor-Controlled Glycosylation: New Perspectives toward Anomeric Selectivity. Acc. Chem. Res. 2018, 51, 628–639. [Google Scholar] [CrossRef]
  16. Alabugin, I.V.; Kuhn, L.; Medvedev, M.G.; Krivoshchapov, N.V.; Vil’, V.A.; Yaremenko, I.A.; Mehaffy, P.; Yarie, M.; Terent’ev, A.O.; Zolfigol, M.A. Stereoelectronic Power of Oxygen in Control of Chemical Reactivity: The Anomeric Effect is Not Alone. Chem. Soc. Rev. 2021, 50, 10253–10345. [Google Scholar] [CrossRef]
  17. Schuff, B.P.; Mercer, G.J.; Nguyen, H.M. Palladium-Catalyzed Stereoselective Formation of α-O-Glycosides. Org. Lett. 2007, 9, 3173–3176. [Google Scholar] [CrossRef] [PubMed]
  18. Xiang, S.-H.; Hoang, K.L.; He, M.J.; Tan, Y.-J.; Liu, X.-W. Reversing the Stereoselectivity of a Palladium-Catalyzed O-Glycosylation through an Inner-sphere or Outer-sphere Pathway. Angew. Chem. Int. Ed. 2015, 54, 604–607. [Google Scholar]
  19. Kim, H.; Men, H.; Lee, C. Stereoselective Palladium-Catalyzed O-Glycosylation Using Glycals. J. Am. Chem. Soc. 2004, 126, 1336–1337. [Google Scholar] [CrossRef] [PubMed]
  20. Yao, H.; Zhang, S.; Leng, W.-L.; Leow, M.-L.; Xiang, S.; He, J.; Liao, H.; Hoang, K.L.M.; Liu, X.-W. Catalyst-Controlled Stereoselective O-Glycosylation: Pd(0) vs Pd(II). ACS Catal. 2017, 7, 5456–5460. [Google Scholar] [CrossRef]
  21. Comely, A.C.; Eelkema, R.; Minnaard, A.J.; Feringa, B.L. De Novo Asymmetric Bio- and Chemocatalytic Synthesis of Saccharides–Stereoselective Formal O-Glycoside Bond Formation Using Palladium Catalysis. J. Am. Chem. Soc. 2003, 125, 8714–8715. [Google Scholar] [CrossRef]
  22. Babu, R.S.; O’Doherty, G.A. A Palladium-Catalyzed Glycosylation Reaction: The de Novo Synthesis of Natural and Unnatural Glycosides. J. Am. Chem. Soc. 2003, 125, 12406–12407. [Google Scholar] [CrossRef] [PubMed]
  23. Xiang, S.; Lu, Z.; He, J.; Hoang, K.L.M.; Zeng, J.; Liu, X.-W. β-Type Glycosidic Bond Formation by Palladium-Catalyzed Decarboxylative Allylation. Chem. Eur. J. 2013, 19, 14047–14051. [Google Scholar] [CrossRef]
  24. Wang, Q.; Lai, M.; Luo, H.; Ren, K.; Wang, J.; Huang, N.; Deng, Z.; Zou, K.; Yao, H. Stereoselective O-Glycosylation of Glycals with Arylboronic Acids Using Air as the Oxygen Source. Org. Lett. 2022, 24, 1587–1592. [Google Scholar] [CrossRef]
  25. Loh, C.C.J. Exploiting Non-Covalent Interactions in Selective Carbohydrate Synthesis. Nat. Rev. Chem. 2021, 5, 792–815. [Google Scholar] [CrossRef]
  26. Betzemeier, B.; Knochel, P. Modern Solvents in Organic Synthesis: Perfluorinated Solvents–A Novel Reaction Medium in Organic Chemistry. Top. Curr. Chem. 1999, 206, 60–78. [Google Scholar]
  27. Berger, R.; Resnati, G.; Metrangolo, P.; Weberd, E.; Hulliger, J. Organic Fluorine Compounds: A Great Opportunity for Enhanced Materials Properties. Chem. Soc. Rev. 2011, 40, 3496–3508. [Google Scholar] [CrossRef] [PubMed]
  28. Klement, I.; Knochel, P. Selective Oxidation of Zinc Organometallics to Hydroperoxides Using Oxygen in Perfluorohexanes. Synlett 1995, 27, 1113–1114. [Google Scholar] [CrossRef]
  29. Brown, H.C.; Negishi, E. Bisborolane. Highly Elusive Bisboracyclane. J. Am. Chem. Soc. 1971, 93, 6682–6683. [Google Scholar] [CrossRef]
  30. Brown, H.C.; Midland, M.M.; Kabalka, G.W. Stoichiometrically Controlled Reaction of Organoboranes with Oxygen Under very Mild Conditions to Achieve Essentially Quantitative Conversion into Alcohols. J. Am. Chem. Soc. 1971, 93, 1024–1025. [Google Scholar] [CrossRef]
  31. Barton, D.H.R.; Jang, D.O.; Jaszberenyi, J.C. An Improved Radical Chain Procedure for the Deoxygenation of Secondary and Primary Alcohols using Diphenylsilane as Hydrogen Atom Donor and Triethylborane-Air as Initiator. Tetrahedron Lett. 1990, 31, 4681–4684. [Google Scholar] [CrossRef]
  32. Klement, I.; Lütjens, H.; Knochel, P. Transition Metal Catalyzed Oxidations in Perfluorinated Solvents. Angew. Chem. Int. Ed. 1997, 36, 1454–1456. [Google Scholar] [CrossRef]
  33. Tada, N.; Cui, L.; Ishigami, T.; Ban, K.; Miura, T.; Uno, B.; Itoh, A. Facile Aerobic Photooxidative Oxylactonization of Oxocarboxylic Acids in Fluorous Solvents. Green Chem. 2012, 14, 3007–3009. [Google Scholar] [CrossRef]
  34. Karimi, M.; Sadeghi, S.; Mohebali, H.; Azarkhosh, Z.; Safarifard, V.; Mahjoub, A.; Heydari, A. Fluorinated Solvent-Assisted Photocatalytic Aerobic Oxidative Amidation of Alcohols via Visible-Light-Mediated HKUST-1/Cs-POMoW Catalysis. New J. Chem. 2021, 45, 14024–14035. [Google Scholar] [CrossRef]
  35. Horváth, I.T.; Rábi, J. Facile Catalyst Separation Without Water: Fluorous Biphases Hydroformylation of Olefins. Science 1994, 266, 72–75. [Google Scholar] [CrossRef]
  36. Maayan, G.; Fish, R.H.; Neumann, R. Polyfluorinated Quaternary Ammonium Salts of Polyoxometalate Anions: Fluorous Biphasic Oxidation Catalysis with and without Fluorous Solvents. Org. Lett. 2003, 5, 3547–3550. [Google Scholar] [CrossRef]
  37. O’Hagan, D. Understanding Organofluorine Chemistry. An Introduction to the C–F Bond. Chem. Soc. Rev. 2008, 37, 308–319. [Google Scholar] [CrossRef]
  38. Champagne, P.A.; Desroches, J.; Paquin, J.-F. Organic Fluorine as a Hydrogen-Bond Acceptor: Recent Examples and Applications. Synthesis 2015, 47, 306–322. [Google Scholar]
  39. Yu, J.-S.; Liu, Y.-L.; Tang, J.; Wang, X.; Zhou, J. Highly Efficient “On Water” Catalyst-Free Nucleophilic Addition Reactions Using Difluoroenoxysilanes: Dramatic Fluorine Effects. Angew. Chem. Int. Ed. 2014, 53, 9512–9516. [Google Scholar] [CrossRef]
  40. Cao, Z.; Wang, W.; Liao, K.; Wang, X.; Zhou, J.; Ma, J. Catalytic Enantioselective Synthesis of Cyclopropanes Featuring Vicinal All-Carbon Quaternary Stereocenters with a CH2F Group; Study of Influence of C-F···H-N Interactions on Reactivity. Org. Chem. Front. 2018, 5, 2960–2968. [Google Scholar] [CrossRef]
  41. Myers, K.E.; Kumar, K. Fluorophobic Acceleration of Diels-Alder Reactions. J. Am. Chem. Soc. 2000, 122, 12025–12026. [Google Scholar] [CrossRef]
  42. Piscelli, B.A.; Sanders, W.; Yu, C.; Maharik, N.A.; Lebl, T.; Cormanich, R.A.; O’Hagan, D. Fluorine-Induced Pseudo-Anomeric Effects in Methoxycyclohexanes through Electrostatic 1,3-Diaxial Interactions. Chem. Eur. J. 2020, 26, 11989–11994. [Google Scholar] [CrossRef] [PubMed]
  43. Misbahi, K.; Lardic, M.; Ferrières, V.; Noiret, N.; Kerbal, A.; Plusquellec, D. Unexpected fluorous solvent effect on oxidation of 1-thioglycosides. Tetrahedron Asymmetry 2001, 12, 2389–2393. [Google Scholar] [CrossRef]
  44. Oikawa, M.; Tanak, T.; Fukud, N.; Kusumoto, S. One-Pot Preparation and Activation of Glycosyl Trichloroacetimidates: Operationally Simple Glycosylation Induced by Combined Use of Solid-Supported, Reactivity-Opposing Reagents. Tetrahedron Lett. 2004, 45, 4039–4042. [Google Scholar] [CrossRef]
  45. Farrán, A.; Cai, C.; Sandoval, M.; Xu, Y.; Liu, J.; Hernáiz, M.J.; Linhardt, R.J. Green Solvents in Carbohydrate Chemistry: From Raw Materials to Fine Chemicals. Chem. Rev. 2015, 115, 6811–6853. [Google Scholar] [CrossRef]
  46. Di Salvo, A.; David, M.; Crousse, B.; Bonnet-Delpon, D. Self-Promoted Nucleophilic Addition of Hexafluoro-2-propanol to Vinyl Ethers. Adv. Synth. Catal. 2006, 348, 118–124. [Google Scholar] [CrossRef]
  47. Nakano, H.; Kitazume, T. Organic Reactions without an Organic Medium–Utilization of Perfluorotriethylamine as a Reaction Medium. Green Chem. 1999, 1, 21–22. [Google Scholar] [CrossRef]
  48. Gorityala, B.K.; Lorpitthaya, R.; Bai, Y.; Liu, X.-W. ZnCl2/Alumina Impregnation Catalyzed Ferrier Rearrangement: An Expedient Synthesis of Pseudoglycosides. Tetrahedron 2009, 65, 29–30. [Google Scholar] [CrossRef]
  49. Zhou, J.; Chen, H.; Shan, J.; Li, J.; Yang, G.; Chen, X.; Xin, K.; Zhang, J.; Tang, J. FeCl3·6H2O/C: An Efficient and Recyclable Catalyst for the Synthesis of 2,3-Unsaturated O- and S-Glycosides. J. Carbohydr. Chem. 2014, 33, 313–325. [Google Scholar] [CrossRef]
  50. Santra, A.; Guchhait, G.; Misra, A. Nitrosyl Tetrafluoroborate Catalyzed Preparation of 2,3-Unsaturated Glycosides and 2-Deoxyglycosides of Hindered Alcohols, Thiols, and Sulfonamides. Synlett 2013, 24, 581–586. [Google Scholar] [CrossRef]
  51. Ruan, Z.; Dabideen, D.; Blumenstein, M.; Mootoo, D.R.A. Modular Synthesis of the Bis-Tetrahydrofuran Core of Rolliniastatin from Pyranoside Precursors. Tetrahedron 2000, 56, 9203–9211. [Google Scholar] [CrossRef]
  52. Yadav, J.S.; Reddy, B.V.S.; Pandey, S.K. Ceric (IV) Ammonium Nitrate-Catalyzed Glycosidation of Glycals: A Facile Synthesis of 2,3-Unsaturated Glycosides. New J. Chem. 2001, 25, 538–540. [Google Scholar] [CrossRef]
  53. Srinivas, B.; Reddy, T.R.; Radha Krishna, P.; Kashyap, S. Copper (II) Triflate as a Mild and Efficient Catalyst for Ferrier Glycosylation: Synthesis of 2,3-Unsaturated O-Glycosides. Synlett 2014, 25, 1325–1330. [Google Scholar] [CrossRef]
  54. Saeeng, R.; Siripru, O.; Sirion, U. IBr-Catalyzed O-Glycosylation of D-Glucals: Facile Synthesis of 2,3-Unsaturated-O-Glycosides. Heterocycles 2015, 91, 849–861. [Google Scholar]
  55. Bound, D.J.; Bettadaiah, B.K.; Srinivas, P. ZnBr2-Catalyzed and Microwave-Assisted Synthesis of 2,3-Unsaturated Glucosides of Hindered Phenols and Alcohols. Synth. Commun. 2014, 44, 2565–2576. [Google Scholar] [CrossRef]
  56. Frappa, I.; Sinou, D. An Easy and Efficient Preparation of Aryl α-O-Δ2−Glycosides. Synth. Commun. 1995, 25, 2941–2951. [Google Scholar] [CrossRef]
  57. Sun, G.; Qiu, S.; Ding, Z.; Chen, H.; Zhou, J.; Wang, Z.; Zhang, J. Magnetic Core-Shell Fe3O4@C-SO3H as an Efficient and Renewable ‘Green Catalyst’ for the Synthesis of O-2,3-Unsaturated Glycopyranosides. Synlett 2017, 28, 347–352. [Google Scholar]
  58. Babu, B.S.; Balasubramanian, K.K. Indium Trichloride Catalyzed Glycosidation. An Expeditious Synthesis of 2,3-Unsaturated Glycopyranosides. Tetrahedron Lett. 2000, 41, 1271–1274. [Google Scholar] [CrossRef]
  59. Yadav, J.S.; Reddy, B.V.S.; Murthy, C.V.S.R.; Kumar, G.M. Scandium Triflate Catalyzed Ferrier Rearrangement: An Efficient Synthesis of 2,3-Unsaturated Glycopyranosides. Synlett 2000, S11210, 1450–1451. [Google Scholar] [CrossRef]
  60. Stevanović, D.; Pejović, A.; Damljanović, I.; Vukićević, R.D.; Vukić ević, M.; Bogdanovic, G.A. Anodic Generation of a Zirconium Catalyst for Ferrier Rearrangement and Hetero Michael Addition. Tetrahedron Lett. 2012, 53, 6257–6260. [Google Scholar] [CrossRef]
  61. Stevanović, D.; Pejović, A.; Damljanović, I.; Minić, A.; Bogdanović, G.A.; Vukićević, M.; Radulović, N.S.; Vukićević, R.D. Ferrier Rearrangement Promoted by an Electrochemically Generated Zirconium Catalyst. Carbohydr. Res. 2015, 407, 111–121. [Google Scholar] [CrossRef]
  62. Bhagavathy, S.; Ajay, K.B.; Kalpattu, K.B. Microwave-induced, Montmorillonite K10-Catalyzed Ferrier Rearrangement of Tri-O-Acetyl-D-Galactal: Mild, Eco-Friendly, Rapid Glycosidation with Allylic Rearrangement. Tetrahedron Lett. 2002, 43, 6795–6798. [Google Scholar]
  63. Grynkiewicz, G.; Priebe, W.; Zamojski, A. Synthesis of Alkyl 4, 6-di-O-Acetryl-2,3-Dideoxy-a-D-Three-hex-2-Enopyranosides from 3,4,6-tri-O-Acetyl-1,5-Anhydro-2-Droxy D-lyxo-hex-1-enitol(3,4,6-tri-O-acetyl-e-galactal). Carbohydr. Res. 1979, 68, 33–41. [Google Scholar] [CrossRef]
  64. Chen, P.; Zhang, D.D. Sm(OTf)3 as a Highly Effificient Catalyst for the Synthesis of 2,3-Unsaturated O- and S-Pyranosides from Glycals and the Temperature-Dependent Formation of 4-O-Acetyl-6-Deoxy-2,3-Unsaturated S-Pyranosides and 4-O-Acetyl-6-Deoxy-3-Alkylthio Glycals. Tetrahedron 2014, 70, 8505–8510. [Google Scholar] [CrossRef]
  65. Tatina, M.B.; Mengxin, X.; Peilin, R.; Judeh, Z.M.A. Robust Perfluorophenylboronic Acid-Catalyzed Stereoselective Synthesis of 2,3-Unsaturated O-, C-, N- and S-linked Glycosides. Beilstein J. Org. Chem. 2019, 15, 1275–1280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Kim, B.H.; Jacobs, P.B.; Elliott, R.L.; Curran, D.P. A Gentral- Synthetic Approach to S113 Optically Active Iridoid Aglpoaes. The Total Synthesis of Beta-Ethyl Descarbomethoxyverbenalol, Ethyl Catalpot, and (-)-Specionin. Tetrahedron 2014, 44, 3079–3092. [Google Scholar]
  67. Bartlett, M.J.; Peter, T.; Northcote, P.T.; Lein, M.; Harvey, J.E. 13C NMR Analysis of 3,6-Dihydro-2H-pyrans: Assignment of Remote Stereochemistry Using Axial Shielding Effects. J. Org. Chem. 2014, 79, 5521–5532. [Google Scholar] [CrossRef]
  68. Khan, A.T.; Sidick Basha, R.S.; Lal, M. Bromodimethylsulfonium Bromide (BDMS) Catalyzed Synthesis of 2,3-Unsaturated-O-Glycosides via Ferrier Rearrangement. Arkivoc 2012, 2013, 201–212. [Google Scholar] [CrossRef]
Figure 1. Motivation and reaction design for acid-catalyzed stereocontrolled Ferrier-type glycosylation assisted by perfluorinated solvent. (a) Representative bioactive molecules with 2,3-unsaturated O-glycoside scaffold; (b) Conventional approaches to access 2,3-unsaturated O-glycoside scaffold; (c) Strategies to activate donor or/and acceptor for Pd-catalyzed O-glycosylation; (d) This work: acid-catalyzed stereocontrolled O-glycosylation assisted by perfluorinated solvent.
Figure 1. Motivation and reaction design for acid-catalyzed stereocontrolled Ferrier-type glycosylation assisted by perfluorinated solvent. (a) Representative bioactive molecules with 2,3-unsaturated O-glycoside scaffold; (b) Conventional approaches to access 2,3-unsaturated O-glycoside scaffold; (c) Strategies to activate donor or/and acceptor for Pd-catalyzed O-glycosylation; (d) This work: acid-catalyzed stereocontrolled O-glycosylation assisted by perfluorinated solvent.
Molecules 27 07234 g001
Scheme 1. Substrate generality with respect to glycosyl acceptors. (a) Substrate scope with respect to alcohols; (b) Substrate scope with respect to phenols; (c) Substrate scope of S-glycosyl acceptors.
Scheme 1. Substrate generality with respect to glycosyl acceptors. (a) Substrate scope with respect to alcohols; (b) Substrate scope with respect to phenols; (c) Substrate scope of S-glycosyl acceptors.
Molecules 27 07234 sch001
Scheme 2. Substrate generality with respect to glycosyl donors. (a) Substrate scope with respect ot galactal; (b) Substrate scope with respect to allal; (c) Substrate scope with respect to rhamnal.
Scheme 2. Substrate generality with respect to glycosyl donors. (a) Substrate scope with respect ot galactal; (b) Substrate scope with respect to allal; (c) Substrate scope with respect to rhamnal.
Molecules 27 07234 sch002
Scheme 3. Reactions with representative functional molecules as glycosyl acceptors and recycling experiments. (a) Functional molecules as acceptors; (b) Recycling experiment investigation.
Scheme 3. Reactions with representative functional molecules as glycosyl acceptors and recycling experiments. (a) Functional molecules as acceptors; (b) Recycling experiment investigation.
Molecules 27 07234 sch003
Table 1. Optimization of the reaction conditions.
Table 1. Optimization of the reaction conditions.
Molecules 27 07234 i001
Entry aAdditiveCatalyst bSolventTemp. (°C)Yield (%) cStereoselectivity (α:β)
1TFE-Ethanol10045>20:1
2PFD-Ethanol10055>20:1
3PFH-Ethanol10060>20:1
4PFTEA-Ethanol10015>20:1
5--Ethanol100<105:1
6--PFH100trace-
7-resin-H+PFH10096>20:1
8-resin-H+CH2Cl2100161.5:1
9-resin-H+Ethanol100857:1
10PFHresin-H+Hexane100--
11PFHresin-H+Toluene100trace-
12PFHresin-H+DCE100trace-
13PFHresin-H+DMF100--
14-resin-H+PFH8055>20:1
15 d-resin-H+PFH8095>20:1
16-resin-H+PFH60trace-
a Unless otherwise specified, all reactions were performed with 1a (0.138 mmol, 1 equiv), 2a (1.2 equiv), additive (10 mol%), catalyst (0.2 mg, 0.6 wt%) for 6 h under N2 in 0.5 mL solvent. b Resin-H+: sulfonic polystyrene type resin. c Isolated yields. d 14 h. DCE: dichloride ethane, DMF: N, N-dimethylformamide.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, Z.; Li, Y.; Xiang, S.; Zuo, M.; Sun, Y.; Jiang, X.; Jiao, R.; Wang, Y.; Fu, Y. Acid Catalyzed Stereocontrolled Ferrier-Type Glycosylation Assisted by Perfluorinated Solvent. Molecules 2022, 27, 7234. https://doi.org/10.3390/molecules27217234

AMA Style

Lu Z, Li Y, Xiang S, Zuo M, Sun Y, Jiang X, Jiao R, Wang Y, Fu Y. Acid Catalyzed Stereocontrolled Ferrier-Type Glycosylation Assisted by Perfluorinated Solvent. Molecules. 2022; 27(21):7234. https://doi.org/10.3390/molecules27217234

Chicago/Turabian Style

Lu, Zhiqiang, Yanzhi Li, Shaohua Xiang, Mengke Zuo, Yangxing Sun, Xingxing Jiang, Rongkai Jiao, Yinghong Wang, and Yuqin Fu. 2022. "Acid Catalyzed Stereocontrolled Ferrier-Type Glycosylation Assisted by Perfluorinated Solvent" Molecules 27, no. 21: 7234. https://doi.org/10.3390/molecules27217234

Article Metrics

Back to TopTop