Next Article in Journal
Structure and Bonding in Planar Hypercoordinate Carbon Compounds
Next Article in Special Issue
Synthesis and Characterization of Abasic β-Diol-C-nucleosides
Previous Article in Journal
Imidazolium-Modified Silica Gel for Highly Selective Preconcentration of Ag(I) from the Nitric Acid Medium
Previous Article in Special Issue
Carbohydrate-Based Azacrown Ethers in Asymmetric Syntheses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Greener Technique for Microwave-Assisted O-Silylation and Silyl Ether Deprotection of Uridine and Other Substrates

Department of Biotechnologies, Chemistry and Pharmacy, University of Siena, Via Aldo Moro 2, I-53100 Siena, Italy
*
Author to whom correspondence should be addressed.
Chemistry 2022, 4(4), 1714-1722; https://doi.org/10.3390/chemistry4040112
Submission received: 23 November 2022 / Revised: 9 December 2022 / Accepted: 11 December 2022 / Published: 14 December 2022
(This article belongs to the Special Issue Organic Chemistry Research in Italy)

Abstract

:
A single clean, good-yielding, environment-friendly microwave-assisted procedure for O-silylation of uridine with tert-butyldimethylsilyl chloride (TBDMSCl), 1,8-Diazabicyclo(5.4.0)undec-7-ene (DBU) and potassium nitrate as catalyst under solvent-free conditions is reported. Subsequent silyl ether deprotection is accomplished with a reusable acidic resin via microwave irradiation. Both the silylation and desilylation protocols have been applied to a panel of alcohols of pharmaceutical interest.

1. Introduction

The hydroxyl functional group can be found almost ubiquitously in compounds of pharmaceutical interest, such as nucleosides, carbohydrates, terpenes and steroids [1].
Silyl ethers are among the most frequently used protective groups for alcohols thanks to the variability of both their steric properties and cleavage conditions [2,3,4]. Important applications are described in the field of antiviral research, for the synthesis of nucleoside analogues [5,6]. Among them, molnupiravir has recently attracted interest due to its broad-spectrum antiviral activity, in particular against the SARS-CoV-2 virus. Corey-Venkateswarlu [7] described the use of a silyl chloride with imidazole in N,N’-dimethylformamide (DMF) to promote O-silylation of primary and secondary alcohols, and since then, many other silylation procedures with different silylating agents, bases and other additives have been developed [2,3,4,8,9]. However, despite being trustworthy and useful synthetic allies, the use of protecting groups often struggles to meet with green and clean chemistry principles [10]. In fact, among the bases and catalysts adopted, efficient results have been reported with pyridine or pyridine derivatives such as pyridine N-oxide, and 4-dimethylaminopyridine, iodine and lithium sulfide, known for their relevant toxicity for operators and the environment [3,11]. Thus, the development of safer and environmentally friendly procedures for both the introduction and cleavage of protecting groups is of considerable interest [12,13,14]. Many efforts have been recently made to limit the use of toxic and pollutant systems, including their replacement with safer chemicals or the use of less-impacting technologies. Solvent-free and microwave-assisted reactions are receiving increasing attention from the viewpoint of green and clean chemistry [15]. Recently, during our studies on the development of a cleaner uridine O-silylation protocol, we found that microwave irradiation made the reaction considerably fast and good-yielding. We report herein a new solvent-free microwave-assisted O-silylation procedure for uridine and other alcohols and a microwave-assisted cleavage of the obtained silyl ethers with a reusable acidic resin.

2. Materials and Methods

2.1. Materials and Instruments

All commercially available reagents and solvents were used as purchased without further purification (Sigma-Aldrich, St Louis, MO, USA), (Alfa-Aesar, Kandel, Germany) (Carlo Erba, Rodano, Milano, Italy). Dry DMF was used as purchased and dichloromethane (DCM) was distilled from CaH2 directly before use.
Microwave-assisted reactions were performed in a single mode using a CEM Discover Synthesis Unit (CEM Corp., Matthews, NC, USA)TM microwave. A continuous microwave power ranging from 0 to 300 W has been applied. A calibrated infrared temperature control was used to measure the temperature of the vessel. The method used for all experiments involved stirring the reaction mixture with a Teflon-coated magnetic stir bar inside the vessel.
Progress of the reaction was monitored by thin-layer chromatography (TLC) using Merck TLC silica gel 60 F254 using the appropriate eluent as described in the Supplementary Materials. Column chromatography was performed on Merck 60 Å silica gel, 230–400 mesh, for flash technique.
1H and 13C spectra were recorded at 400 MHz on a Bruker UltraShieldTM 400 instrument (400 MHz for 1H and 101 MHz for 13C NMR spectroscopy), or a 600 MHz on a Bruker UltraShieldTM 600 instrument (600 MHz for 1H and 242 MHz for 13C NMR spectroscopy). Chemical shifts are reported in parts per million (δ scale), relative to TMS peak.
Mass spectra (LC-MS) were acquired using an Agilent 1100 LC-MSD VL system (G1946C) with an electrospray ionization source (ESI). Data were obtained by direct injection with a 0.4 mL/min flow rate using a binary solvent system of 95/5 (v/v) methanol (MeOH)/H2O. Mass spectra were acquired in positive- or negative-mode scanning over the mass range 105–1500 m/z, using a variable fragmentor voltage of 10–70 mV (0–70).
The purity of final products was assessed by HPLC-MS and was above 90%.

2.2. General Procedure for the Synthesis of Silyl Ethers

The appropriate alcohol (1.0 mmol), KNO3 (0.3 mmol) and 1,8-Diazabicyclo(5.4.0)undec-7-ene (DBU) (2.7 mmol) were added to a microwave tube equipped with a magnetic stirrer. Then, tert-butyldimethylsilyl chloride (TBDMS-Cl) (1.5 mmol) was added to the mixture, the tube capped and the reaction was stirred in the microwave at 75 °C, 300 W, 100 PSI, on high-stirring mode and with the cooling option on. For entry 10, due to the volatility of both the starting alcohol and silyl ether obtained, the temperature was lowered to 60 °C. The completion of the reaction was monitored via TLC. The reaction mixture was washed with water and extracted with ethyl acetate (AcOEt) (3 × 25 mL). In the case of 1, the reaction mixture was treated with triethylamine phosphate buffer 0.5 M (pH = 7) and extracted with AcOEt (3 × 25 mL). The combined organic layers were washed with brine (saturated solution), dried on anhydrous Na2SO4 and evaporated under reduced pressure. The products were purified by silica gel flash column chromatography (petroleum ether (PE)/AcOEt 9:1 for entries 1 and 8; PE/AcOEt 95:5 for entries 2, 4, 5, 6, 7; PE/AcOEt 99:1 for entry 3, PE/AcOEt 98:2 for entry 9, PE/Et2O 98:2 for entry 10). The purified compounds were characterized by 1H and 13C NMR analysis and HPLC-MS (for details, see Supplementary Materials). Note: When performing scaling-up of the reaction conditions up to 1 g of starting uridine, the addition of the reagents was performed at 0 °C (exothermic reaction) and the microwave cooling function was turned on to prevent overheating.

2.3. General Procedure for the Synthesis of Alcohols via Deprotection of Silyl Ethers

The appropriate silyl ether (1.0 mmol), Dowex 50WX4-200 (2:1 w/w) and MeOH (0.5 mL) were added to a microwave tube without a magnetic stirrer. The mixture was allowed to incubate at rt in a vortex stirrer for 30–60 min at 160–200 rpm. Then, the reaction mixture was stirred in the microwave at 50–60 °C, 300 W and 100 PSI. The completion of the reaction was monitored via TLC. The reaction mixture was filtered with a vacuum-connected gooch funnel, and the Dowex beads were washed several times with MeOH. Then, the solvent was evaporated under reduced pressure. For derivative D4, additional work-up was performed, by treating with water and extracting with AcOEt (3 × 25 mL). The combined organic layers were washed with brine (saturated solution), dried on anhydrous Na2SO4 and evaporated under reduced pressure. The products were purified by silica gel flash column chromatography (DCM/MeOH 95:5→ 9:1 for entry D1, PE/AcOEt 99:1→ 95:5 for entry D2; PE/AcOEt 95:5 for entry D4). Entry D3 was characterized without further purification. The desired product in entry D5 has not been obtained. The purified compounds were characterized by 1H and 13C NMR analysis, and HPLC-MS (for details, see Supplementary Materials).

3. Results and Discussion

3.1. O-Silylation of Uridine

Preliminary experiments on O-silylation of uridine with TBDMSCl with different bases and additives (Scheme 1) were carried out in order to increase yields, reduce reaction times and find more sustainable reaction conditions. The adopted conditions and results are shown in Table 1 (entries 1aq).
The initial reaction system was composed by uridine, with TBDMSCl as the silylating agent, imidazole as the base and DCM or DMF as the solvent. Entries 1a and 1b were performed according to literature procedures [16,17], but very low yields, long reaction times and a mixture of disilylated and trisilylated compounds were observed.
The use of catalysts such as iodine or AgNO3 in silylation reactions has already been described in literature [2,11,18]. According to these studies, the addition of AgNO3 to the reagent system (entry 1c) considerably enhanced the yield and reduced the reaction time. In order to evaluate the role of the nitrate ion and its counterion on the reaction outcome, the experiment was conducted with KNO3 with excellent yield (97%, entry 1d). Due to concerns related to the use of silver [19,20], and the lower cost, KNO3 was chosen in place of AgNO3 for the rest of the experiments as a cheap and safer alternative.
Temperature control proved to be necessary for the silylation to proceed; lower temperatures are detrimental to the yield, reducing the yield from 97 to 47% (entry 1d and 1e, respectively).
Microwave irradiation gave excellent results (entry 1f) in terms of yield, reaction times and selectivity, yielding only the trisilylated uridine derivative, and was thus chosen as the heating method for the rest of the study.
After optimizing the heating method and catalyst, we replaced DMF with the cleaner 2-methyl tetrahydrofuran [21,22] (entries 1i1p), a non-toxic and sustainable solvent obtained from various lignocellulosic feedstocks. Then, we focused our attention on greener alternatives to imidazole. The best results were obtained with 1,8-diazabicyclo[5.4.0]undec-7-ene [23,24,25] (DBU), already included in the GSK-suggested environmentally friendly list of bases (1n1o) [26]. On the contrary, inorganic bases and other green organic alternatives such as 2,6 lutidine gave poor results (1g1h, 1p) [2]. To improve the atom economy of the reaction and avoid unnecessary stochiometric excess of reagents used, various attempts were made, reducing the quantity of potassium nitrate, silylating agent and base (1m1o, 1q). As shown in entry 1n1o, KNO3 can be considered catalytic for the reaction.
Finally, the reaction was performed in solvent-free conditions [27] (entry 1q), reducing TBDMSCl and DBU to 1.5 eq and 2.7 eq, respectively. Excellent yields were obtained, optimizing the protocol in terms of yield, reaction time and greenness. Differently from the procedures already described [16], we obtained the target triprotected uridine 1 without using toxic reagents, harmful solvents [2] and with significantly reduced reaction time (10 min vs. 18 h). Most importantly, the setup is extremely simple, and the reaction was performed in the presence of air and without the need of anhydrous solvents, as reported in classic procedures.
The reaction was satisfactorily applied on a medium scale (200 mg of starting uridine), obtaining a comparable yield and reaction time.

3.2. Scope Variation

The new solvent-free microwave-assisted procedure for uridine (Scheme 2) was then applied to a panel of representative aliphatic and aromatic substrates (Table 2), including alcohols of pharmaceutical interest such as eugenol (phenolic compounds), cholesterol (steroids), L-menthol and trans-sobrerol (terpenes), 4-pentyn-1-ol (alkyne).
It is noteworthy that the reaction conditions allowed the silylation of sterically hindered L-menthol with a very good yield (entry 4), which failed when performed in solution with the same reagent system and 2-methyltetrahydrofuran as solvent.
Trans-sobrerol was chosen to study the selectivity of the reaction when two differently hindered hydroxyl groups are simultaneously present. 1H NMR experiments confirmed the formation of silyl ether 5 and no evidence of disilylated compound was observed, due to the steric hindrance of the tertiary hydroxyl group. Application on benzylic substrates was successful (entries 6 and 8, 84% and 81% yield, respectively). The presence of an amino group in entry 8 was well tolerated, while the insertion of a 3-hydroxyl group in entry 7 gave a ratio of monosilylated and disilylated product of 3:1, indicating that both the phenolic group and the benzylic alcohol group compete for the silylation. Reaction of secondary alcohol in entry 9 required a longer time to react (20 min) and gave a moderate yield (43%), probably due to steric factors. The reaction conditions proved to be applicable also for the volatile alkyne 10 (62%). As for uridine protection, our sustainable solvent-free protocol proved to be effective for the panel of alcohols analyzed, eliminating hazardous reactants and solvents [8,27] and without the need of anhydrous procedures.

3.3. Microwave-Assisted Resin-Based Deprotection of a Panel of Silyl Ethers

As is well documented in literature, silyl ethers are typically cleaved with an excess of fluoride sources such as TBAF or KF, or under acidic conditions [3,28]. However, fluoride-mediated desilylation suffers from poor atom economy, due to the generation of waste fluorosilanes by-products [29]. In addition, the traditional THF/water/acid procedures usually suffer from lengthy reaction times, especially for bulkier silyl ethers such as TBS, TBDPS, or TIPS [30]. Some more environmentally friendly methods for silyl ether cleavage include examples with catalytic copper (II) salts [31,32], catalytic I2 [14,33], SO3H silica gel [34], alumina [35,36] kaolinite [37] and montmorillonite K-10 clay catalysis [38].
With the aim of developing also a green and mild silyl ether cleavage procedure which could benefit from either the use of reusable material or catalytic systems, various experiments with clay and resins were made on silyl ether 1. Montmorillonite K-10 gave a 56% yield of deprotected uridine, and similar results were obtained when montmorillonite K-10 was used in combination with KF [39] (53% yield). Switching to resin-based cleavage, the same trend was observed with Amberlyst 15 (50% yield) while Dowex 50WX4-200 resin gave a good yield of 83% and was thus chosen as the standard deprotection method (Scheme 3).
Cleavage of a representative panel of silyl ethers (compounds 15) was thus accomplished with Dowex 50WX4-200 acidic resin in methanol under microwave irradiation [30] at 60 °C, as shown in Table 3, with fair to good yields. Only a minimum amount of methanol was required to allow proper swelling of the acidic resin. The products have been isolated by simple filtration, avoiding work-ups and in some cases purifications. Most importantly, the resin could be recycled and reused for a further cleavage step, reducing the associated costs.
Deprotection of persilylated uridine D1 proceeded smoothly (83%), but a longer reaction time was required for a complete deprotection (100 min), while phenolic silyl ether D2 and secondary cholesteryl silyl ether D3 required shorter reaction times (20 min) for their cleavage, with fair to good yields (56% and 64%, respectively). As can be seen in entries D4 and D5, deprotection of hindered or sensitive-to-acidic conditions silyl ethers was troublesome. In particular, in entry D5, the desired alcohol 5 was not isolated, due to the formation of a complex mixture of by-products in acidic conditions.

4. Conclusions

In conclusion, we optimized the classical O-silylation protocols for uridine, developing a novel, high-yielding, clean and simple solvent-free microwave-assisted procedure. The protocol proved to be scalable and applicable to a panel of alcohols with good-to-excellent yields. In contrast to the methods previously described [16], we were able to produce the target alcohols with shorter reaction times and without the use of hazardous solvents or toxic reagents. Most importantly, the setup is very simple, the reaction was carried out in the presence of air, and anhydrous conditions were not required differently from traditional protocols.
Subsequent silyl ether deprotection was accomplished via microwave-assisted resin-based cleavage, with fair-to-good yields, making the entire sequence of protection and deprotection steps facile and sustainable. Despite the fact that the deprotection step is applicable mainly to acidic stable compounds, the solvent-free silylation and resin-based deprotection are cleaner and simpler procedures for the introduction and removal of silyl ether protecting groups, useful for phenyl, benzyl and hindered alcohols.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry4040112/s1, synthetic procedures and NMR.

Author Contributions

Conceptualization, C.P. and A.B.; methodology, C.P. and A.B.; software, C.P. and A.B.; investigation, C.P., F.P., C.V. and A.B.; writing—original draft preparation C.P.; writing—review and editing, A.B.; visualization, C.P. and A.B.; supervision, A.B. and E.D.; funding acquisition, E.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Tuscany Region “Bando Regionale Ricerca Salute 2018—TUSCAVIR.NET”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Synthetic procedures and NMR are reported in Supplementary Materials.

Acknowledgments

We gratefully acknowledge the Tuscany Region “Bando Regionale Ricerca Salute 2018—TUSCAVIR.NET” for the economic support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cramer, J.; Sager, C.P.; Ernst, B. Hydroxyl Groups in Synthetic and Natural Product Derived Therapeutics: A Perspective on a Common Functional Group. J. Med. Chem. 2019, 62, 8915–8930. [Google Scholar] [CrossRef] [PubMed]
  2. Bartoszewicz, A.; Kalek, M.; Stawinski, J. Iodine-Promoted Silylation of Alcohols with Silyl Chlorides. Synthetic and Mechanistic Studies. Tetrahedron 2008, 64, 8843–8850. [Google Scholar] [CrossRef]
  3. Wuts, P.G.M. Greene’s Protective Groups in Organic; John Wiley and Sons: Hoboken, NJ, USA, 2014; pp. 1–1360. [Google Scholar]
  4. Patschinski, P.; Zhang, C.; Zipse, H. The Lewis Base-Catalyzed Silylation of Alcohols-a Mechanistic Analysis. J. Org. Chem. 2014, 79, 8348–8357. [Google Scholar] [CrossRef]
  5. Seley-Radtke, K.L.; Yates, M.K. The Evolution of Nucleoside Analogue Antivirals: A Review for Chemists and Non-Chemists. Part 1: Early Structural Modifications to the Nucleoside Scaffold. Antiviral Res. 2018, 154, 66–86. [Google Scholar] [CrossRef] [PubMed]
  6. Somoza Àlvaro, S. Protecting Groups for RNA Synthesis: An Increasing Need for Selective Preparative Methods. Chem. Soc. Rev. 2008, 37, 2668–2675. [Google Scholar] [CrossRef] [PubMed]
  7. Corey, E.J.; Venkateswarlu, A. Protection of Hydroxyl Groups as Tert-Butyldimethylsilyl Derivatives. J. Am. Chem. Soc. 1972, 94, 6190–6191. [Google Scholar] [CrossRef]
  8. Jereb, M. Highly Atom Economical Uncatalysed and Iodine-Catalysed Silylation of Phenols, Alcohols and Carbohydrates, Using HMDS under Solvent-Free Reaction Conditions (SFRC). Tetrahedron 2012, 68, 3861–3867. [Google Scholar] [CrossRef]
  9. Ashraf, M.A.; Liu, Z.; Li, C.; Zhang, D. Recent Advances in Catalytic Silylation of Hydroxyl-Bearing Compounds: A Green Technique for Protection of Alcohols Using Si–O Bond Formations. Appl. Organomet. Chem. 2021, 35, 1–24. [Google Scholar] [CrossRef]
  10. Anastas, P.; Eghbali, N. Green Chemistry: Principles and Practice. Chem. Soc. Rev. 2010, 39, 301–312. [Google Scholar] [CrossRef]
  11. Hakimelahi, G.H.; Proba, Z.A.; Ogilvie, K.K. Nitrate Ion as Catalyst for Selective Silylations of Nucleosides. Tetrahedron Lett. 1981, 22, 4775–4778. [Google Scholar] [CrossRef]
  12. Giri, N.; Bowen, C.; Vyle, J.S.; James, S.L. Fast, Quantitative Nucleoside Protection under Solvent-Free Conditions. Green Chem. 2008, 10, 627–662. [Google Scholar] [CrossRef]
  13. Karimi, B.; Golshani, B. Mild and Highly Efficient Method for the Silylation of Alcohols Using Hexamethyldisilazane Catalyzed by Iodine under Nearly Neutral Reaction Conditions. J. Org. Chem. 2000, 65, 7228–7230. [Google Scholar] [CrossRef] [PubMed]
  14. Saxena, I.; Deka, N.; Sarma, J.C.; Tsuboi, S. A Convenient Method for Protection and Deprotection of Alcohols and Phenols as Alkylsilyl Ethers Catalyzed by Iodine under Microwave Irradiation. Synth. Commun. 2003, 33, 4185–4191. [Google Scholar] [CrossRef]
  15. Loupy, A. Solvent-Free Reactions. In Modern Solvents in Organic Synthesis; Springer: Berlin/Heidelberg, Germany, 1999. [Google Scholar]
  16. Painter, G.R.; Guthrie, D.B.; Bluemling, G.R.; Natchus, M.G. N4-Hydroxycytidine and Derivatives and Anti-Viral Uses Related Thereto. WO2016106050A1, 30 June 2016. [Google Scholar]
  17. Gadakh, B.; Vondenhoff, G.; Lescrinier, E.; Rozenski, J.; Froeyen, M.; Van Aerschot, A. Base Substituted 5′-O-(N-Isoleucyl)Sulfamoyl Nucleoside Analogues as Potential Antibacterial Agents. Bioorganic Med. Chem. 2014, 22, 2875–2886. [Google Scholar] [CrossRef] [PubMed]
  18. Ogilvie, K.K.; McGee, D.P.C.; Boisvert, S.M.; Hakimelahi, G.H.; Proba, Z.A. The Preparation of Protected Arabinonucleosides. Can. J. Chem. 1983, 61, 1204–1212. [Google Scholar] [CrossRef]
  19. Hadrup, N.; Sharma, A.K.; Loeschner, K. Toxicity of Silver Ions, Metallic Silver, and Silver Nanoparticle Materials after in Vivo Dermal and Mucosal Surface Exposure: A Review. Regul. Toxicol. Pharmacol. 2018, 98, 257–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Teran, C.G.; Sura, S.; Cabandugama, P.; Berson, C. Silver Nitrate Ingestion: Report of a Case with an Uneventful Course and Review of the Literature. Clin. Pract. 2011, 1, e43. [Google Scholar] [CrossRef]
  21. Pace, V.; Hoyos, P.; Castoldi, L.; Domínguez De María, P.; Alcántara, A.R. 2-Methyltetrahydrofuran (2-MeTHF): A Biomass-Derived Solvent with Broad Application in Organic Chemistry. Chem. Sus. Chem. 2012, 5, 1369–1379. [Google Scholar] [CrossRef]
  22. Cai, C.M.; Zhang, T.; Kumar, R.; Wyman, C.E. Integrated Furfural Production as a Renewable Fuel and Chemical Platform from Lignocellulosic Biomass. J. Chem. Technol. Biotechnol. 2014, 89, 2–10. [Google Scholar] [CrossRef]
  23. Shieh, W.C.; Dell, S.; Repič, O. 1,8-Diazabicyclo[5.4.0]Undec-7-Ene (DBU) and Microwave-Accelerated Green Chemistry in Methylation of Phenols, Indoles, and Benzimidazoles with Dimethyl Carbonate. Org. Lett. 2001, 3, 4279–4281. [Google Scholar] [CrossRef]
  24. Hermecz, I. Chemistry of Diazabicycloundecene (DBU) and Other Pyrimidoazepines. Adv. Heterocycl. Chem. 1987, 42, 83–202. [Google Scholar] [CrossRef]
  25. Nand, B.; Khanna, G.; Chaudhary, A.; Lumb, A.; Khurana, J.M. 1,8-Diazabicyclo[5.4.0]Undec-7-Ene (DBU): A Versatile Reagent in Organic Synthesis. Curr. Org. Chem. 2015, 19, 790–812. [Google Scholar] [CrossRef]
  26. Henderson, R.K.; Hill, A.P.; Redman, A.M.; Sneddon, H.F. Development of GSK’s Acid and Base Selection Guides. Green Chem. 2015, 17, 945–949. [Google Scholar] [CrossRef]
  27. Hatano, B.; Toyota, S.; Toda, F. Efficient Solvent-Free O-Silylation of Alcohols with R3SiCl. Green Chem. 2001, 3, 140–142. [Google Scholar] [CrossRef]
  28. Crouch, R.D. Selective Deprotection of Silyl Ethers. Tetrahedron 2013, 69, 2383–2417. [Google Scholar] [CrossRef]
  29. Wang, B.; Sun, H.X.; Chen, B.; Sun, Z.H. Practical, Environment-Benign and Atom Economic KOAc-Catalysed Deprotection of Aryl TIPS Ethers under Mild Fluoride-Free Conditions. Green Chem. 2009, 11, 1112–1114. [Google Scholar] [CrossRef]
  30. Crouch, R.; Williams, A. Rapid, Acid-Mediated Deprotection of Silyl Ethers Using Microwave Heating. Synth. Commun. 2006, 36, 959–964. [Google Scholar] [CrossRef]
  31. Tan, Z.P.; Wang, L.; Wang, J.B. Deprotection of T-Butyldimethylsiloxy (TBDMS) Protecting Group with Catalytic Copper (II) Chloride Dihydrate. Chin. Chem. Lett. 2000, 11, 753–756. [Google Scholar]
  32. González-Calderón, D.; Benítez-Puebla, L.J.; González-González, C.A.; Assad-Hernández, S.; Fuentes-Benítez, A.; Cuevas-Yáñez, E.; Corona-Becerril, D.; González-Romero, C. Selective Deprotection of TBDMS Alkyl Ethers in the Presence of TIPS or TBDPS Phenyl Ethers by Catalytic CuSO4·5H2O in Methanol. Tetrahedron Lett. 2013, 54, 5130–5132. [Google Scholar] [CrossRef]
  33. Vaino, A.R.; Szarek, W.A. A Mild and Efficient Method for the Deprotection of Tert-Butyldimethylsilyl Ethers Using Iodine in Methanol. Chem. Commun. 1996, 2351–2352. [Google Scholar] [CrossRef]
  34. Fujii, H.; Shimada, N.; Ohtawa, M.; Karaki, F.; Koshizuka, M.; Hayashida, K.; Kamimura, M.; Makino, K.; Nagamitsu, T.; Nagase, H. Deprotection of Silyl Ethers by Using SO3H Silica Gel: Application to Sugar, Nucleoside, and Alkaloid Derivatives. Tetrahedron 2017, 73, 5425–5429. [Google Scholar] [CrossRef]
  35. Varma, R.S.; Lamture, J.B.; Varma, M. Alumina-Mediated Cleavage of t-Butyldimethylsilyl Ethers. Tetrahedron Lett. 1993, 34, 3029–3032. [Google Scholar] [CrossRef]
  36. Feixas, J.; Capdevila, A.; Guerrero, A. Utilization of Neutral Alumina as a Mild Reagent for the Selective Cleavage of Primary and Secondary Silyl Ethers. Tetrahedron 1994, 50, 8539–8550. [Google Scholar] [CrossRef]
  37. Upadhya, T.T.; Daniel, T.; Sudalai, A.; Ravindranathan, T.; Sabu, K.R. Natural Kaolinitic Clay: A Mild and Efficient Catalyst for the Tetrahydropyranylation and Trimethylsilylation of Alcohols. Synth. Commun. 1996, 26, 4539–4544. [Google Scholar] [CrossRef]
  38. Zhang, Z.H.; Li, T.S.; Yang, F.; Fu, C.G. Montmorillonite Clay Catalysis XI1: Protection and Deprotection of Hydroxyl Group by Formation and Cleavage of Trimethylsilyl Ethers Catalysed by Catalysed by Montmorillonite K-10. Synth. Commun. 1998, 28, 3105–3114. [Google Scholar] [CrossRef]
  39. Ghanei, M.; Khalilzadeh, M.A.; Hashemi, M.M. Potassium Fluoride/K10-Montmorillonite Nanostructure as a Green and Reusable Catalyst under Mild Reaction Conditions for the Knovenagel Condensation. Orient. J. Chem. 2016, 32, 665–669. [Google Scholar] [CrossRef]
Scheme 1. O-silylation of uridine.
Scheme 1. O-silylation of uridine.
Chemistry 04 00112 sch001
Scheme 2. General procedure for solvent-free microwave-assisted O-silylation of uridine.
Scheme 2. General procedure for solvent-free microwave-assisted O-silylation of uridine.
Chemistry 04 00112 sch002
Scheme 3. General scheme for microwave-assisted resin-based silyl ether cleavage.
Scheme 3. General scheme for microwave-assisted resin-based silyl ether cleavage.
Chemistry 04 00112 sch003
Table 1. O-silylation of uridine with TBDMSCl using different catalysts, solvent bases and reaction setup.
Table 1. O-silylation of uridine with TBDMSCl using different catalysts, solvent bases and reaction setup.
EntryTBDMSCl
(mmol)
Catalyst
(mmol)
Base/Nucleophile
(mmol)
SolventHeating
Method
TimeYield
1a4.0DMAP
(0.1)
Imidazole
(4.0)
DCMNone3 d11%
1b6.0/Imidazole
(10.0)
DMFConventional48 h27%
1c12.0AgNO3
(11.0)
Imidazole
(26.0)
DMFConventional3 h86%
1d12.0KNO3
(11.0)
Imidazole
(26.0)
DMFConventional3 h97%
1e12.0KNO3
(11.0)
Imidazole
(26.0)
DMFNone3 h47%
1f12.0KNO3
(11.0)
Imidazole
(26.0)
DMFMW5 min99%
1g12.0KNO3
(11.0)
CsCO3
(26.0)
DMFMW5 min/
1h12.0KNO3
(11.0)
Na2CO3
(26.0)
DMFMW5 min/
1i12.0KNO3
(11.0)
Imidazole
(26.0)
2-Me-THFMW5 min99%
1j6.0KNO3
(6.0)
Imidazole
(26.0)
2-Me-THFMW10 min19%
1k9.0KNO3
(8.2)
Imidazole
(19.5)
2-Me-THFMW50 min88%
1l9.0KNO3
(8.2)
DBU
(19.5)
2-Me-THFMW5 min90%
1m9.0KNO3
(3.0)
DBU
(19.5)
2-Me-THFMW10 min89%
1n9.0KNO3
(1.0)
DBU
(19.5)
2-Me-THFMW10 min99%
1o6.0KNO3
(1.0)
DBU
(13.0)
2-Me-THFMW10 min90%
1p6.0KNO3
(1.0)
2,6-lutidine
(13.0)
2-Me-THFMW10 min21%
1q4.5KNO3
(1.0)
DBU
(8.0)
Solvent-freeMW10 min98%
Reaction conditions: All reactions were carried out at 75 °C unless otherwise stated. 2-Me-THF = 2-methyltetrahydrofuran; MW = Microwave.
Table 2. Silylation of alcohols using the TBDMSCl/DBU/KNO3 reagent system.
Table 2. Silylation of alcohols using the TBDMSCl/DBU/KNO3 reagent system.
EntryStarting AlcoholProductt (min)Yield a
1Chemistry 04 00112 i001Chemistry 04 00112 i0021098%
2Chemistry 04 00112 i003Chemistry 04 00112 i004582%
3Chemistry 04 00112 i005Chemistry 04 00112 i0064088%
4Chemistry 04 00112 i007Chemistry 04 00112 i0083086%
5Chemistry 04 00112 i009Chemistry 04 00112 i0102093%
6Chemistry 04 00112 i011Chemistry 04 00112 i012584%
7Chemistry 04 00112 i013Chemistry 04 00112 i014550%
8Chemistry 04 00112 i015Chemistry 04 00112 i0161081%
9Chemistry 04 00112 i017Chemistry 04 00112 i0182043%
10Chemistry 04 00112 i019Chemistry 04 00112 i020562% b
Reaction conditions: Starting alcohol (0.3 mmol, 1.0 eq), TBDMSCl (0.4 mmol, 1.5 eq), KNO3 (0.1 mmol, 0.3 eq), DBU (0.8 mmol, 2.7 eq), 75 °C, MW. a Isolated yield. b The reaction temperature was lowered to 60 °C due to the volatility of the starting material and the resulting silyl ether. Experimental details can be found in Supplementary Materials.
Table 3. Deprotection of the previously obtained silyl ethers 15 (Table 2) using the DowexWX4-200/MeOH reagent system.
Table 3. Deprotection of the previously obtained silyl ethers 15 (Table 2) using the DowexWX4-200/MeOH reagent system.
EntryStarting AlcoholProductt (min) Yield a
D1Chemistry 04 00112 i021
1
Chemistry 04 00112 i02210083%
D2Chemistry 04 00112 i023
2
Chemistry 04 00112 i0242056%
D3Chemistry 04 00112 i0253Chemistry 04 00112 i0262064%
D4Chemistry 04 00112 i027
4
Chemistry 04 00112 i0286040%
D5Chemistry 04 00112 i029
5
Chemistry 04 00112 i03040/ a
Reaction conditions: silyl ether (1.0 mmol, 1.0 eq); Dowex 50WX4-200 (2:1 w/w), MeOH (0.5 mL), MW, 60 °C. a Silyl ether 5 reacts slower, leading to a complex mixture of by-products. The desired alcohol D5 was not isolated. The reaction was carried out also at r.t. overnight to evaluate the influence of mild heating on the reaction outcome, but no changes were observed.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pasqualini, C.; Poggialini, F.; Vagaggini, C.; Brai, A.; Dreassi, E. A Greener Technique for Microwave-Assisted O-Silylation and Silyl Ether Deprotection of Uridine and Other Substrates. Chemistry 2022, 4, 1714-1722. https://doi.org/10.3390/chemistry4040112

AMA Style

Pasqualini C, Poggialini F, Vagaggini C, Brai A, Dreassi E. A Greener Technique for Microwave-Assisted O-Silylation and Silyl Ether Deprotection of Uridine and Other Substrates. Chemistry. 2022; 4(4):1714-1722. https://doi.org/10.3390/chemistry4040112

Chicago/Turabian Style

Pasqualini, Claudia, Federica Poggialini, Chiara Vagaggini, Annalaura Brai, and Elena Dreassi. 2022. "A Greener Technique for Microwave-Assisted O-Silylation and Silyl Ether Deprotection of Uridine and Other Substrates" Chemistry 4, no. 4: 1714-1722. https://doi.org/10.3390/chemistry4040112

Article Metrics

Back to TopTop