Next Article in Journal
Effects of Sequential Induction Combining Thermal Treatment with Ultrasound or High Hydrostatic Pressure on the Physicochemical and Mechanical Properties of Pea Protein–Psyllium Hydrogels as Elderberry Extract Carriers
Previous Article in Journal
Disturbance of Immune Microenvironment in Androgenetic Alopecia through Spatial Transcriptomics
Previous Article in Special Issue
Stereoselective Synthesis and Catalytical Application of Perillaldehyde-Based 3-Amino-1,2-diol Regioisomers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Natural and Sugar-Modified Nucleosides Using the Iodine/Triethylsilane System as N-Glycosidation Promoter

by
Martina Cimafonte
1,†,
Anna Esposito
2,†,
Maria De Fenza
1,
Francesco Zaccaria
1,
Daniele D’Alonzo
1,* and
Annalisa Guaragna
1
1
Department of Chemical Sciences, University of Naples Federico II, I-80126 Naples, Italy
2
Department of Chemical, Materials and Production Engineering, University of Naples Federico II, I-80125 Naples, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(16), 9030; https://doi.org/10.3390/ijms25169030
Submission received: 17 July 2024 / Revised: 11 August 2024 / Accepted: 12 August 2024 / Published: 20 August 2024
(This article belongs to the Special Issue Recent Trends in Stereoselective Synthesis and Chiral Catalysis)

Abstract

:
The reagent system based on the combined use of Et3SiH/I2 acts as an efficient N-glycosidation promoter for the synthesis of natural and sugar-modified nucleosides. An analysis of reaction stereoselectivity in the absence of C2-positioned stereodirecting groups revealed high selectivity with six-membered substrates, depending on the nucleophilic character of the nucleobase or based on anomerization reactions. The synthetic utility of the Et3SiH/I2-mediated N-glycosidation reaction was highlighted by its use in the synthesis of the investigational drug apricitabine.

1. Introduction

Whether in their monomeric or oligomeric form, nucleosides represent key biomolecules involved in a wide variety of research areas, spanning from chemistry to biology and from biotechnology to medicine [1,2,3,4,5]. Nucleosides have been the cornerstone of antiviral and anticancer therapy over the past five decades [6,7]. From early and well-known examples (zidovudine, lamivudine, acyclovir) up to the latest drugs (sofosbuvir, remdesivir, molnupiravir), nucleosides have become the standards of care for the treatment of many long-standing deadly diseases [8], while acting as first-line therapies for emerging infections of concern [9,10]. On the other hand, oligonucleotides are currently recognized as being among the most promising and versatile tools in modern drug discovery and diagnostics, aimed at the treatment, prevention and detection of various pathological conditions [11,12]. As illustrative examples in the field, mRNA vaccines have emerged as highly effective alternatives to conventional vaccines [13], boosted by the exceptional performance of Comirnaty (Pfizer-BioNTech) and Spikevax (Moderna) in counteracting the spread of COVID-19 [14,15]. Many of these notable successes have been achieved through the strategic replacement of natural entities with biomimetic analogues, built to resemble the structure of their physiological counterparts while altering (silencing or amplifying) their functions [16,17]. The demand for diverse nucleoside structures with unique chemical properties has driven intense efforts aimed at the development of efficient synthetic procedures, enabling their preparation in large amounts with high purity levels [18,19,20,21,22,23,24]. A crucial aspect of nucleoside synthesis focuses on the development of new protocols for the N-glycosylation reaction (Scheme 1). The latter is based on the formation of an N-glycosidic bond, achieved by a coupling reaction between a sugar and a nucleobase, both suitably activated [25].
Considering that bioactive nucleosides are typically those with a β-configuration at their anomeric centre (i.e., the nucleobase and the hydroxymethyl group are involved in a cis stereochemical relationship), most N-glycosylation methods focus on stereoselective reactions toward β-nucleosides, regardless of their belonging to the d- or l-series. However, while synthetic strategies, from the earliest Vorbrüggen reaction [25] up to the recent Au-catalyzed [26] or epoxide-based methods [27], typically achieve β-selectivity by exploiting neighbouring group participation at the C2 position [28], the search for stereoselective approaches is much more challenging in the case of the synthesis of 2′-deoxynucleosides and their analogues. Only a few methods able to fulfil this purpose have been devised so far, based on remote stereocontrol [29], in situ complexation [30,31], SN2 reactions [32] and anomerization processes [33,34]. In this field, we have provided an example of a remotely controlled N-glycosidation reaction [35] using the well-known Et3SiH/I2 and PHMS/I2 systems [36,37], as exemplified by the coupling reaction of oxathiolane 1 with (5-fluoro)-cytosine en route to the synthesis of the antiretroviral drugs lamivudine (3TC) and emtricitabine (FTC) (Scheme 2a). The N-glycosidation reaction provided excellent results in terms of chemical efficiency and β-l-stereoselectivity (up to >99% de) as a result of the remote assistance of the menthyl ester group [38] (Scheme 2a). In addition, the nature of the protective group at the N4 position of cytosine or 5-fluorocytosine was demonstrated to have a deep influence on the stereoselective outcome of the reaction, as the electron-withdrawing group dramatically improved β- l-selectivity.
Based on these preliminary data, our ongoing efforts aimed to achieve selective transformations from halogen-containing substrates [39,40,41,42,43,44] led us to explore the stereoselective potential of the Et3SiH/I2 system in nucleoside synthesis beyond participating effects of the substrate. Particularly, the Et3SiH/I2-mediated N-glycosidation reaction has been studied using a variety of sugars and sugar-like substrates, lacking hydroxyl groups at the C2 and/or C3 positions (Scheme 2b). The last ones have been prepared by exploiting known synthetic procedures or slightly modified to improve their reaction yields (see Supplementary Materials for synthetic details). These compounds represent useful substrates for preparing modified nucleosides of biological/pharmacological interest [45], either to be employed as such or within more complex oligonucleotide structures [33,46]. Various pyrimidine nucleobases have been introduced to probe their influence on the stereoselective outcome of the reaction [35].

2. Results and Discussion

The general N-glycosidation procedure involved, as step 1 (Scheme 2b), the reaction of the deoxysugar substrate (an anomeric acetate) with the reactive species from the Et3SiH/I2 system (HI and Et3SiI), which was in turn obtained by premixing the two reagents and waiting for the resulting solution to discolour [47]. Attention was first focused on this activation step, to study its efficiency and stereoselectivity and to identify the actual iodinating agent. The treatment of five- and six-membered (di)deoxyglycosides 3–6 (Scheme 3) with the premixed Et3SiH/I2 rapidly (<1 h) led, as early as by −78 °C, to the disappearance of the starting sugars.
However, NMR analysis [48] of the crude mixture from furanosides 3 and 4 indicated the formation of a variety of species, and none of them were characterized by the presence of the typical downshielded anomeric protons (δ ~7 ppm) of glycosyl iodides [49]. On the other hand, the reaction of hexopyranosides 5 and 6 led to the exclusive formation of α-glycosyl iodides 7 and 8, as demonstrated by their small J1–2 1H NMR values (e.g., for compound 7, J1,2ax = 2.6 Hz, δ = 7.18 ppm; Scheme 3) [34]. To determine whether both in situ-generated HI and Et3SiI could work as activating agents, 5 was also treated with a substoichiometric amount (0.6 eq) of a 1:1 mixture of Et3SiH/I2. Under these conditions, at −78 °C, we only observed a ~50% conversion of 5 into the corresponding iodide 7, while at 0 °C a much faster and complete conversion was found. This evidence demonstrated that both reagents enable glycoside activation, although Et3SiI, which is supposed to be the less reactive species [50], could be effectively exploited only at relatively high temperatures.
N-glycosidation reactions using thymine and (N4-benzoyl)cytosine as model nucleobases were then performed (Scheme 4 and Scheme 5; see Supplementary Materials for chemical structures and the synthesis of sugars and sugar-like substrates, Schemes S1–S5). After the iodination step was achieved (1.2 eq Et3SiH, 1.2 eq I2 at −78 °C) and TLC showed the disappearance of the starting glycoside acetate [51] (30 min), a solution of the nucleobase, previously silylated with bis-(trimethylsilyl)-acetamide (BSA) (DCM, 40 °C, 1–6 h), was added dropwise via a cannula under an argon flow. Glycosyl iodide consumption in favour of the corresponding nucleoside was commonly observed after 1–2 h. High-yielding conversions were observed in all cases (Scheme 4).
As an example, the reaction of dideoxyhexopyranoside 5 with thymine provided the corresponding nucleoside 10a with a 91% yield. The same substrate, when reacted with persilylated N4-benzoylcytosine, provided deoxycytidine analogue 10b with an 80% yield. The reaction of peracetylated 2-deoxyglucoside 6 with thymine required relatively higher temperatures (−20 °C), providing the corresponding nucleoside 11a with a 79% yield. Thymidine 12a and the nucleosides with a dideoxyribose (13a and 13b) and an oxathiolane (14a) moiety were similarly obtained under the same conditions (70–84%). Conversely, oxathiolane nucleoside 15a was obtained from the corresponding acetyl derivative only after keeping the reaction at 0 °C, and much more sluggishly [52] (24 h; Scheme 4). Looking at reaction stereoselectivity, only a very weak β-selectivity was observed under these conditions. As an example, thymine dideoxynucleoside 10a was attained from the corresponding sugar substrate 5 with an α:β ratio of 1:1. The reaction of the same substrate with N4-benzoylcytosine furnished nucleoside 10b with an α:β ratio of 1:2. Similar selectivities were observed for nucleosides bearing deoxyglucoside (12a), (di)deoxyribofuranoside (13a and 13b) and oxathiolane moieties (14a and 15a).
Looking for an improvement of the reaction stereoselectivity, N-glycosidation studies were performed by modifying parameters such as temperature, reaction time and the amount of activating reagents and/or nucleobases used (Scheme 5a). In most cases, hexopyranoside 5 was used as a model substrate. As an example, the treatment of 5 with Et3SiH/I2 was conducted, changing the step 1 temperature from −78 °C to −20 °C for 1 h or 16 h (Scheme 5a); afterwards, the reaction mixture was cooled to −78 °C and the addition of the persilylated nucleobase was performed at the same temperature. In this case, only a slight improvement in the α:β ratio (from 1:1 to 1:3) was observed in the reaction with thymine (compare with Scheme 4). A further increase in temperature of the iodination step up to rt did not further enhance the overall selectivity. Next, a substoichiometric amount (0.5 eq) of a 1:1 mixture of Et3SiH and I2 at room temperature was used. The α:β ratio still did not increase beyond 1:3. Reactions were eventually performed by changing the nucleobase (and keeping the temperature of step 1 at −20 °C). When uracil was used in place of thymine, the corresponding dideoxyuridine 10c was obtained from 5 with an even lower selectivity (α:β = 1:2) than thymine nucleoside 10a. The use of 5-fluorocytosine provided nucleoside 10d with an increase in its α:β ratio to 1:4. Conversely, the same reaction carried out with cytosine provided the corresponding deoxycytidine analogue 10e with an α:β ratio of 1:11. A further change in N-glycosidation conditions was eventually performed. As soon as the conversion of the starting acetates into the corresponding nucleosides was detected by TLC, we found that if the reaction mixtures were warmed to higher temperatures and left under stirring for longer times, the anomerization of α/β-nucleosides took place (Scheme 5b). As an example, the reaction of 5 with thymine, left at −78 °C for 1 h and then allowed to gradually reach room temperature for an additional 16 h, provided nucleoside 10a with an α:β ratio of 1:25 (>95% yield). The α:β ratio obtained from the reaction of 5 with N4-benzoylcytosine, under the same conditions, was even higher (1:50, 67% after chromatographic purification). In the reaction of 2-deoxyglucose 6 with thymine, the anomeric α:β ratio increased from 1:1 to 1:7 in the formation of 11a, but only after the reaction temperature was raised to 60 °C for 16 h. The same temperature did not significantly increase the α:β ratio in the reaction of 5 with cytosine (from 1:11 to 1:12; compare with Scheme 4).
The data reported in Scheme 5 deserve some comments. Starting from the observation that reactions with glycosyl iodides typically proceed under SN2 conditions [49], the presence of α/β anomeric mixtures of nucleosides demonstrate that the α-oriented glycosyl iodide is not the only reactive species in the solution. Indeed, even though an NMR analysis at room temperature indicated the presence of as the only anomeric glycosyl iodide, the occurrence at lower temperatures of the more reactive [53] β-glycosyl iodide , obtained under SN2 conditions from glycoside acetate , is expected (Scheme 6a).
This hypothesis could explain why, in the first trials, a complete lack of stereoselectivity was observed (Scheme 4). At higher temperatures, the formation of a certain amount of iodide by the anomerization of can take place as long as the reaction proceeds, considering that the concentration of iodide ion released from the substrate progressively increases (Scheme 6b).
While the overall stereoselectivity of the reaction was primarily influenced by the presence of , a strong nucleophile effect further affecting the α:β ratio was also observed. It was closely dependent on the electron density of the nucleobase (Scheme 6c). The highest β-selectivity was achieved with cytosine, which has the highest electron density at its N1 position. On the other hand, a linear decrease in selectivity was found as the electron density at N1 decreased, from 5-fluorocytosine to thymine and uracil [54] (Scheme 5 and Scheme 6).
Fruitful for a mechanistic analysis, we also observed that, when not strictly anhydrous conditions were used during the iodination step, the reaction smoothly proceeded to the undesired α,α-disaccharide 16 (Scheme 5a), which was detected as the only stereoisomer by NMR analysis (see Supplementary Materials). Accordingly, a reasonable hypothesis for the observed stereoselectivity assumes that weak nucleophiles (e.g., H2O) are able to discriminate between the more reactive α-directing β-glycosyl iodide and the more stable β-directing α-glycosyl iodide. With moderate nucleophiles (e.g., uracil, thymine and fluorocytosine), a low discrimination capacity is expected between the two glycosyl iodides, leading to low anomeric selectivity toward the β-isomers. With the stronger nucleophile cytosine, faster reactions and no discrimination at all between the two electrophilic species are conceivable. The attack will be therefore directed toward the most abundant species, i.e., α-glycosyl iodide , thus providing a high β-selectivity (Scheme 6c).
The results reported in Scheme 5 also demonstrate that hexopyranosides 5 and 6 are substrates for anomerization reactions. The increase in the β:α ratio demonstrates that no remote assistance of the C6-acetyl group occurs with these substrates. It is worth recalling that these substrates were already found to be amenable to anomerization reactions when using TfOTMS as an N-glycosidation promoter [34,55]. However, our conditions represent largely preferable alternatives to the above approach, as they provide higher β-selectivities while requiring lower reaction temperatures, thereby minimizing the formation of the side products encountered when using the previous method [34]. Compared to the TfOTMS-based reaction, the milder conditions required by the Et3SiH/I2 method clearly suggest an involvement of the iodide ion in promoting the anomerization rection, as depicted in Scheme 6d.
The Et3SiH/I2-mediated N-glycosidation reaction was also tested on five-membered sugar-like substrates, modifying the reaction temperature, time, the amount of reagents used and the nucleobase (Scheme 7). The reaction again demonstrated high efficiency with all substrates (reaction yields: from 70 to >95%). Regarding reaction stereoselectivity, no relevant β-selectivity could be achieved with these substrates (α:β from 1:1 to 1:2). An enhancement of the β:α ratios by anomerization in the reactions of five-membered sugar substrates was observed, although it was much lower. As an example, after oxathiolanyl thymidine 15a was obtained (0 °C, 24 h, α:β = 1:2, Scheme 4), the reaction mixture was left at room temperature for an additional 24 h, with an increase in the α:β ratio only up to 1:3 observed (Scheme 7). Under similar conditions, after dideoxy-thymidine or cytidine were formed (13a and 13b) (−78 °C, 2 h), they were kept at rt for 24 h. Unexpectedly, an inversion in the α:β ratio (from 1:2 to 2:1) [56] was found in this case.
To highlight the usefulness of the Et3SiH/I2 reagent system in the synthesis of bioactive compounds, the method was eventually applied to improve the synthesis of the antiretroviral nucleoside apricitabine [57,58] (Scheme 8). We first studied if the enantiomerically enriched [59] sulfoxide 17 could serve as a substrate for Et3SiH/I2 N-glycosidation. However, after 1 h at −78 °C, the treatment of 17 with Et3SiH/I2 only provided the corresponding sulfide. As witnessed by the color change of the reaction mixture (from colorless to purple), the reaction most probably proceeds via formation of molecular iodine (see Supplementary Materials for mechanistic details, Scheme S6). On the other hand, the addition of an organic base (TEA) drove the reaction to the formation of the desired nucleoside 15b, although in limited amounts (21%, 43% based on the recovery of the starting material) and with low β-selectivity. Conversely, the reaction of acetate 18 with cytosine ran much more efficiently (67%), although still with low selectivity (α:β = 1:2). When N4-acetylcytosine was employed, a higher stereoselectivity was observed, although it was still not excellent (α:β = 1:4). After the removal of the protective groups (MeONa), nucleoside 15c could be converted into the desired apricitabine [58]. It is worth underlining that the above approach led to a β-selectivity higher than that reached by the synthetic method used for the industrial production of apricitabine (α:β = 1:2.9) [58].

3. Materials and Methods

3.1. General Methods

All moisture-sensitive reactions were performed under an argon atmosphere using oven-dried glassware. TLC (precoated silica gel plate F254, Merk Life Science S.r.l., Milan, Italy) was used to monitor reactions and compounds were detected by exposure to ultraviolet radiation and iodine vapour and by spraying a 5% ethanolic solution of sulfuric acid. Purifications of the compounds were performed by column chromatography (Merck Kieselgel 60, 70–230 mesh, Merk Life Science S.r.l., Milan, Italy). Combustion analyses were performed using a Flash Smart V elemental analyzer (Thermo Scientific Inc., Waltham, MA, USA). NMR spectrometers operating at 400 MHz (Bruker DRX, Bruker AVANCE, Bruker Corp., Billerica, MA, USA) or 500 MHz (Varian Inova equipped with a VnmrJ 4.0 software, Agilent Technologies, Santa Clara, CA, USA) were used to record NMR spectra. CDCl3 solutions were employed unless otherwise specified. Coupling constant values (J) were reported in Hz. ESI-MS spectra were recorded on a Shimadzu LCMS-8040 system with ESI interface, triple-quadrupole mass analyzer (Shimadzu Corporation, Kyoto, Japan) and Shimadzu LC-MS solution Workstation (version 5.97) software for data processing. Sugars and sugar-like substrates were prepared by exploiting known synthetic procedures [60,61,62,63,64,65,66,67,68,69,70,71,72,73,74] or slightly modified to improve their reaction yields (see Supplementary Materials for details; Schemes S1–S5). Copies of 1H and 13C NMR spectra for new compounds are provided in the Supplementary Material (Figures S1–S8).

3.2. N-Glycosidation Reactions

General procedure. Step 1: To a solution of I2 (0.12 mmol) in anhydrous DCM (0.5 mL), kept under an argon atmosphere, Et3SiH (0.12 mmol) was added at rt. The solution was stirred at the same temperature for 30 min. Then, the mixture was transferred via cannula into a solution of anomeric sugar acetate (0.10 mmol) in anhydrous DCM (0.5 mL) at the given temperature and stirred at the same temperature for 30 min. Step 2: N,O-bis(trimethylsilyl)acetamide (BSA) (0.32 mmol) was added to a suspension of the nucleobase (0.13 mmol) in anhydrous DCM (0.5 mL), kept under an argon atmosphere. The resulting suspension was warmed to 40 °C and stirred at the same temperature until a clear solution was formed (1–6 h). The mixture obtained in step 2 was then transferred via cannula into the solution obtained in step 1 at the given temperature and stirred therein for an appropriate amount of time under an argon atmosphere. Afterwards, the resulting solution was extracted with DCM and washed with brine and a few drops of an aqueous Na2S2O3 solution. The collected organic layers were dried (Na2SO4) and the solvent evaporated under reduced pressure. α:β ratios were assigned by a comparative 1H NMR analysis of the crude reaction mixtures with the literature data [75,76,77,78,79,80,81] (see Supplementary Materials for details).

4. Conclusions

The search for stereoselective methods capable of installing the N–glycosidic bond of natural and synthetic nucleosides without the involvement of stereodirecting groups still represents a challenging topic. Herein, a substantial extension of the N-glycosidation strategy relying on the combined use of Et3SiH/I2 has been performed. Various five- and six-membered deoxy and dideoxysugars and sugar-like substrates have been synthesized and tested in N-glycosidation reactions, with the aim of exploring the stereoselectivity of this reaction beyond the stereodirecting effects caused by neighbouring group participation. Our results indicate that Et3SiH/I2 is an excellent N-glycosidation reagent for the synthesis of natural and synthetic nucleosides, as it promotes, in most cases, stereoselective coupling reactions, especially in the case of six-membered sugar substrates. In this context, a “nucleophile effect” affecting the stereoselective outcome of the reactions has been found, and a mechanistic hypothesis has been formulated accordingly. The reaction has also been conveniently used for the synthesis of nucleoside structures with consolidated biological activity, such as apricitabine, with the finding that its β-selectivity was higher than that reported in the corresponding preparation processes used on an industrial scale. This result, combined with the low cost and high stability of the reagents, as well as with the high chemical efficiency of the reactions, contributes to making this methodology an attractive alternative to the existing N-glycosidation methods.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25169030/s1.

Author Contributions

D.D.: conceptualization, investigation, writing—original draft preparation; M.C., M.D.F. and F.Z.: investigation; A.E.: investigation, writing—original draft preparation, writing—review and editing; A.G.: supervision, funding acquisition, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article and in the Supplementary Materials section.

Acknowledgments

D.D. is deeply thankful to Alfonso Iadonisi for his invaluable assistance in understanding the reactivity of glycosyl iodides throughout the N-glycosidation reactions. M.D.F. is grateful to the Italian MUR for being granted a research associate position (PONR&I 2014–2020, CUP: E65F21003010003). A.E acknowledges the European Cystic Fibrosis Society (ECFS) and Cystic Fibrosis (CF) Europe for their funding (ECFS/CF Europe Post-Doctoral Research Fellowship).

Conflicts of Interest

The authors declare no conflicts of interest.

References and Notes

  1. Seley-Radtke, K.L.; Thames, J.E.; Waters, C.D. Broad Spectrum Antiviral Nucleosides—Our Best Hope for the Future, 1st ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2021; Volume 57, ISBN 9780323915113. [Google Scholar]
  2. De Clercq, E. Selected Milestones in Antiviral Drug Development. Viruses 2024, 16, 169. [Google Scholar] [CrossRef] [PubMed]
  3. Egli, M.; Manoharan, M. Chemistry, Structure and Function of Approved Oligonucleotide Therapeutics. Nucleic Acids Res. 2023, 51, 2529–2573. [Google Scholar] [CrossRef] [PubMed]
  4. Kulkarni, J.A.; Witzigmann, D.; Thomson, S.B.; Chen, S.; Leavitt, B.R.; Cullis, P.R.; van der Meel, R. The Current Landscape of Nucleic Acid Therapeutics. Nat. Nanotechnol. 2021, 16, 630–643. [Google Scholar] [CrossRef] [PubMed]
  5. Shelton, J.; Lu, X.; Hollenbaugh, J.A.; Cho, J.H.; Amblard, F.; Schinazi, R.F. Metabolism, Biochemical Actions, and Chemical Synthesis of Anticancer Nucleosides, Nucleotides, and Base Analogs. Chem. Rev. 2016, 116, 14379–14455. [Google Scholar] [CrossRef] [PubMed]
  6. Jordheim, L.P.; Durantel, D.; Zoulim, F.; Dumontet, C. Advances in the Development of Nucleoside and Nucleotide Analogues for Cancer and Viral Diseases. Nat. Rev. Drug Discov. 2013, 12, 447–464. [Google Scholar] [CrossRef] [PubMed]
  7. De Clercq, E.; Li, G. Approved Antiviral Drugs over the Past 50 Years. Clin. Microbiol. Rev. 2016, 29, 695–747. [Google Scholar] [CrossRef] [PubMed]
  8. Ami, E.; Ohrui, H. Intriguing Antiviral Modified Nucleosides: A Retrospective View into the Future Treatment of COVID-19. ACS Med. Chem. Lett. 2021, 12, 510–517. [Google Scholar] [CrossRef] [PubMed]
  9. Li, G.; De Clercq, E. Therapeutic Options for the 2019 Novel Coronavirus (2019-NCoV). Nat. Rev. Drug Discov. 2020, 19, 149–150. [Google Scholar] [CrossRef] [PubMed]
  10. Roy, V.; Agrofoglio, L.A. Nucleosides and Emerging Viruses: A New Story. Drug Discov. Today 2022, 27, 1945–1953. [Google Scholar] [CrossRef] [PubMed]
  11. Roberts, T.C.; Langer, R.; Wood, M.J.A. Advances in Oligonucleotide Drug Delivery. Nat. Rev. Drug Discov. 2020, 19, 673–694. [Google Scholar] [CrossRef] [PubMed]
  12. Doudna, J.A. The Promise and Challenge of Therapeutic Genome Editing. Nature 2020, 578, 229–236. [Google Scholar] [CrossRef]
  13. Gu, Y.; Duan, J.; Yang, N.; Yang, Y.; Zhao, X. mRNA Vaccines in the Prevention and Treatment of Diseases. MedComm 2022, 3, e167. [Google Scholar] [CrossRef]
  14. Fang, E.; Liu, X.; Li, M.; Zhang, Z.; Song, L.; Zhu, B.; Wu, X.; Liu, J.; Zhao, D.; Li, Y. Advances in COVID-19 MRNA Vaccine Development. Signal Transduct. Target. Ther. 2022, 7, 94. [Google Scholar] [CrossRef] [PubMed]
  15. The importance of chemical modifications in mRNA therapy has recently received the deserved worldwide recognition, culminating in the assignment of 2023 Nobel Prize in Physiology or Medicine to K. Karikó and D. Weissman.
  16. Jadhav, V.; Vaishnaw, A.; Fitzgerald, K.; Maier, M.A. RNA Interference in the Era of Nucleic Acid Therapeutics. Nat. Biotechnol. 2024, 42, 394–405. [Google Scholar] [CrossRef]
  17. Wan, W.B.; Seth, P.P. The Medicinal Chemistry of Therapeutic Oligonucleotides. J. Med. Chem. 2016, 59, 9645–9667. [Google Scholar] [CrossRef]
  18. D’Alonzo, D.; Guaragna, A. Stereoselective Methods in the Synthesis of Bioactive Oxathiolane and Dioxolane Nucleosides. In Chemical Synthesis of Nucleoside Analogues; Merino, P., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2013; pp. 727–780. [Google Scholar]
  19. Esposito, A.; Giovanni, C.; De Fenza, M.; Talarico, G.; Chino, M.; Palumbo, G.; Guaragna, A.; D’Alonzo, D. A Stereoconvergent Tsuji–Trost Reaction in the Synthesis of Cyclohexenyl Nucleosides. Chem. Eur. J. 2020, 26, 2597–2601. [Google Scholar] [CrossRef] [PubMed]
  20. Esposito, A.; Talarico, G.; De Fenza, M.; D’Alonzo, D.; Guaragna, A. Stereoconvergent Synthesis of Cyclopentenyl Nucleosides by Palladium-Assisted Allylic Reaction. Eur. J. Org. Chem. 2022, 2022, e202200708. [Google Scholar] [CrossRef]
  21. De Fenza, M.; Esposito, A.; D’Alonzo, D.; Guaragna, A. Synthesis of Piperidine Nucleosides as Conformationally Restricted Immucillin Mimics. Molecules 2021, 26, 1652. [Google Scholar] [CrossRef]
  22. Wang, H.-J.; Zhong, Y.-Y.; Xiao, Y.-C.; Chen, F.-E. Chemical and Chemoenzymatic Stereoselective Synthesis of β-Nucleosides and Their Analogues. Org. Chem. Front. 2022, 9, 1719–1741. [Google Scholar] [CrossRef]
  23. Ojeda-Porras, A.C.; Roy, V.; Agrofoglio, L.A. Chemical Approaches to Carbocyclic Nucleosides. Chem. Rec. 2022, 22, e202100307. [Google Scholar] [CrossRef]
  24. Meanwell, M.; Silverman, S.M.; Lehmann, J.; Adluri, B.; Wang, Y.; Cohen, R.; Campeau, L.-C.; Britton, R. A Short de Novo Synthesis of Nucleoside Analogs. Science 2020, 369, 725–730. [Google Scholar] [CrossRef] [PubMed]
  25. Vorbruggen, H.; Ruh-Polenz, C. Handbook of Nucleoside Synthesis; John Wiley & Sons: Hoboken, NJ, USA, 2001. [Google Scholar]
  26. Liu, R.; Chen, Y.; Zheng, J.; Zhang, L.; Xu, T.; Xu, P.; Yang, Y. Synthesis of Nucleosides and Deoxynucleosides via Gold(I)-Catalyzed N -Glycosylation of Glycosyl (Z)-Ynenoates. Org. Lett. 2022, 24, 9479–9484. [Google Scholar] [CrossRef]
  27. Downey, A.M.; Pohl, R.; Roithová, J.; Hocek, M. Synthesis of Nucleosides through Direct Glycosylation of Nucleobases with 5- O-Monoprotected or 5-Modified Ribose: Improved Protocol, Scope, and Mechanism. Chem. Eur. J. 2017, 23, 3910–3917. [Google Scholar] [CrossRef]
  28. Kaspar, F.; Stone, M.R.L.; Neubauer, P.; Kurreck, A. Route Efficiency Assessment and Review of the Synthesis of β-Nucleosides via N-Glycosylation of Nucleobases. Green Chem. 2021, 23, 37–50. [Google Scholar] [CrossRef]
  29. Mear, S.J.; Nguyen, L.V.; Rochford, A.J.; Jamison, T.F. Synthesis of (±)-Emtricitabine and (±)-Lamivudine by Chlorotrimethylsilane–Sodium Iodide-Promoted Vorbrüggen Glycosylation. J. Org. Chem. 2022, 87, 2887–2897. [Google Scholar] [CrossRef]
  30. Choi, W.B.; Wilson, L.J.; Yeola, S.; Liotta, D.C.; Schinazi, R.F. In Situ Complexation Directs the Stereochemistry of N-Glycosylation in the Synthesis of Thialanyl and Dioxolanyl Nucleoside Analogs. J. Am. Chem. Soc. 1991, 113, 9377–9379. [Google Scholar] [CrossRef]
  31. Choi, W.-B.; Yeola, S.; Liotta, D.C.; Schinazi, R.F.; Painter, G.R.; Davis, M.; St. Clair, M.; Furman, P.A. Synthesis, Anti-Human Immunodeficiency Virus, and Anti-Hepatitis B Virus Activity of Pyrimidine Oxathiolane Nucleosides. Bioorg. Med. Chem. Lett. 1993, 3, 693–696. [Google Scholar] [CrossRef]
  32. Zhong, M.; Nowak, I.; Robins, M.J. Regiospecific and Highly Stereoselective Coupling of 6-(Substituted-Imidazol-1-yl)Purines with 2-Deoxy-3,5-Di-O-(p-Toluoyl)-α-d-Erythro-Pentofuranosyl Chloride. Sodium-Salt Glycosylation in Binary Solvent Mixtures:  Improved Synthesis of Cladribine. J. Org. Chem. 2006, 71, 7773–7779. [Google Scholar] [CrossRef] [PubMed]
  33. D’Alonzo, D.; Amato, J.; Schepers, G.; Froeyen, M.; Van Aerschot, A.; Herdewijn, P.; Guaragna, A. Enantiomeric Selection Properties of β-HomoDNA: Enhanced Pairing for Heterochiral Complexes. Angew. Chem. Int. Ed. 2013, 52, 6662–6665. [Google Scholar] [CrossRef]
  34. D’Alonzo, D.; Guaragna, A.; Van Aerschot, A.; Herdewijn, P.; Palumbo, G. Toward β-Homo-DNA: Stereoselective de Novo Synthesis of β-l- Erythro -Hexopyranosyl Nucleosides. J. Org. Chem. 2010, 75, 6402–6410. [Google Scholar] [CrossRef]
  35. Caso, M.F.; D’Alonzo, D.; D’Errico, S.; Palumbo, G.; Guaragna, A. Highly Stereoselective Synthesis of Lamivudine (3TC) and Emtricitabine (FTC) by a Novel N -Glycosidation Procedure. Org. Lett. 2015, 17, 2626–2629. [Google Scholar] [CrossRef] [PubMed]
  36. Adinolfi, M.; Iadonisi, A.; Ravidà, A.; Schiattarella, M. Efficient and Direct Synthesis of Saccharidic 1,2-Ethylidenes, Orthoesters, and Glycals from Peracetylated Sugars via the In Situ Generation of Glycosyl Iodides with I2/Et3SiH. Tetrahedron Lett. 2003, 44, 7863–7866. [Google Scholar] [CrossRef]
  37. Adinolfi, M.; Iadonisi, A.; Pastore, A.; Valerio, S. The I2/Et3SiH System: A Versatile Combination with Multiple Applications in Carbohydrate Chemistry. Pure Appl. Chem. 2011, 84, 1–10. [Google Scholar] [CrossRef]
  38. Goodyear, M.D.; Hill, M.L.; West, J.P.; Whitehead, A.J. Practical Enantioselective Synthesis of Lamivudine (3TCTM) via a Dynamic Kinetic Resolution. Tetrahedron Lett. 2005, 46, 8535–8538. [Google Scholar] [CrossRef]
  39. La Gatta, S.; Leone, L.; Maglio, O.; De Fenza, M.; Nastri, F.; Pavone, V.; Chino, M.; Lombardi, A. Unravelling the Structure of the Tetrahedral Metal-Binding Site in METP3 through an Experimental and Computational Approach. Molecules 2021, 26, 5221. [Google Scholar] [CrossRef]
  40. D’Alonzo, D.; De Fenza, M.; Pavone, V.; Lombardi, A.; Nastri, F. Selective Oxidation of Halophenols Catalyzed by an Artificial Miniaturized Peroxidase. Int. J. Mol. Sci. 2023, 24, 8058. [Google Scholar] [CrossRef]
  41. Esposito, A.; D’Alonzo, D.; D’Errico, S.; De Gregorio, E.; Guaragna, A. Toward the Identification of Novel Antimicrobial Agents: One-Pot Synthesis of Lipophilic Conjugates of N-Alkyl d-and l-Iminosugars. Mar. Drugs 2020, 18, 572. [Google Scholar] [CrossRef]
  42. Loreto, D.; Esposito, A.; Demitri, N.; Guaragna, A.; Merlino, A. Reactivity of a Fluorine-Containing Dirhodium Tetracarboxylate Compound with Proteins. Dalt. Trans. 2022, 51, 3695–3705. [Google Scholar] [CrossRef]
  43. Loreto, D.; Esposito, A.; Demitri, N.; Guaragna, A.; Merlino, A. Digging into Protein Metalation Differences Triggered by Fluorine Containing-Dirhodium Tetracarboxylate Analogues. Dalt. Trans. 2022, 51, 7294–7304. [Google Scholar] [CrossRef]
  44. Chino, M.; La Gatta, S.; Leone, L.; De Fenza, M.; Lombardi, A.; Pavone, V.; Maglio, O. Dye Decolorization by a Miniaturized Peroxidase Fe-MimochromeVI*a. Int. J. Mol. Sci. 2023, 24, 11070. [Google Scholar] [CrossRef]
  45. D’Alonzo, D.; De Fenza, M.; Palumbo, G.; Romanucci, V.; Zarrelli, A.; Di Fabio, G.; Guaragna, A. Synthesis of β- L -2′-Fluoro-3′-Thiacytidine (F-3TC) Stereoisomers: Toward a New Class of Oxathiolanyl Nucleosides? Synthesis 2017, 49, 998–1008. [Google Scholar] [CrossRef]
  46. D’Alonzo, D.; Froeyen, M.; Schepers, G.; Di Fabio, G.; Van Aerschot, A.; Herdewijn, P.; Palumbo, G.; Guaragna, A. 1′,5′-Anhydro-l-Ribo-Hexitol Adenine Nucleic Acids (α-l-HNA-A): Synthesis and Chiral Selection Properties in the Mirror Image World. J. Org. Chem. 2015, 80, 5014–5022. [Google Scholar] [CrossRef] [PubMed]
  47. Reaction efficiency was indeed affected by the addition order of the reagents. if I2 was added soon after Et3SiH during glycosyl iodide formation, a considerable amount (50%) of pyran 9 was detected after subsequent nucleobase addition.
  48. NMR spectra of all glycosyl iodides were acquired at room temperature.
  49. Gervay-Hague, J. Taming the Reactivity of Glycosyl Iodides to Achieve Stereoselective Glycosidation. Acc. Chem. Res. 2016, 49, 35–47. [Google Scholar] [CrossRef] [PubMed]
  50. Compare the reactivity of Me3SiI (TMSI) in glycosyl iodide formation; Lam, S.N.; Gervay-Hague, J. Efficient Route to 2-Deoxy β-O-Aryl-d-Glycosides via Direct Displacement of Glycosyl Iodides. Org. Lett. 2003, 5, 4219–4222. [Google Scholar] [CrossRef] [PubMed]
  51. Due to the inherent instability of the deoxyglycosyl iodides, TLC analysis always displayed formation of lower spots than starting sugar acetates, presumably due to formation of hydrolysis products (hemiacetal sugars) of the corresponding glycosyl iodides.
  52. This is most likely a consequence of the stability of the thionium ion intermediate, conceivably formed during activation of the starting acetate.
  53. The formation of β-glycosyl iodide intermediates has been postulated to be formed at −5/0 °C or at lower temperatures, see: Davis, R.A.; Lin, C.-H.; Gervay-Hague, J. Chemoenzymatic Synthesis of Cholesteryl-6-O-Tetradecanoyl-α-d-Glucopyranoside: A Product of Host Cholesterol Efflux Promoted by Helicobacter Pylori. Chem. Commun. 2012, 48, 9083–9085. [Google Scholar] [CrossRef] [PubMed]
  54. Kamel, M.; Mohammadifard, K. Thermodynamic and Reactivity Descriptors Studies on the Interaction of Flutamide Anticancer Drug with Nucleobases: A Computational View. Chem. Rev. Lett. 2021, 4, 54–65. [Google Scholar] [CrossRef]
  55. Augustyns, K.; Rozenski, J.; Van Aerschot, A.; Janssen, G.; Herdewijn, P. Synthesis of 2,4-Dideoxy-β-d-Erythro-Hexopyranosyl Nucleosides. J. Org. Chem. 1993, 58, 2977–2982. [Google Scholar] [CrossRef]
  56. The different anomerization capacity of five- and six-membered sugar and sugar-like substrates is clearly related to the energetic differences between β- and α-anomers. In the case of five-membered nucleosides, the small stability differences between quasi-equatorial/quasi-axial positions must be invoked for the minor selectivity.
  57. Cahn, P.; Altclas, J.; Martins, M.; Losso, M.; Cassetti, I.; Cooper, D.; Cox, S. Antiviral Activity of Apricitabine in Treatment-Experienced HIV-1-Infected Patients with M184V Who Are Failing Combination Therapy. HIV Med. 2011, 12, 334–342. [Google Scholar] [CrossRef] [PubMed]
  58. Marcuccio, S.M.; Epa, R.; White, J.M.; Deadman, J.J. A New Process for Synthesis of Apricitabine, 2-(R)-Hydroxymethyl-4-(R)-(Cytosin-1′-yl)-1,3-Oxathiolane, an Anti-HIV NRTI. Org. Process Res. Dev. 2011, 15, 763–773. [Google Scholar] [CrossRef]
  59. Caputo, R.; Guaragna, A.; Palumbo, G.; Pedatella, S. A New Strategy for the Asymmetric Synthesis of 1,3-Oxathiolane-Based Nucleoside Analogues. Eur. J. Org. Chem. 1999, 1999, 1455–1458. [Google Scholar] [CrossRef]
  60. Belleau, B.; Nguyen-ba, N.; Kong, L.C.C.; Mansour, T.; Jin, H.; Brasili, L. Substituted 1,3-Oxathiolanes and Substituted 1,3-Dithiolanes with Antiviral Properties. DE Patent DE69122264D1, 24 October 1996. [Google Scholar]
  61. Mansour, T.; Haulon, J. Substituted 1,3-Oxathiolanes with Antiviral Properties. EP Patent EP1473294A2, 3 November 2004. [Google Scholar]
  62. Mansour, T.S.; Jin, H.; Wang, W.; Hooker, E.U.; Ashman, C.; Cammack, N.; Salomon, H.; Belmonte, A.R.; Wainberg, M.A. Anti-Human Immunodeficiency Virus and Anti-Hepatitis-B Virus Activities and Toxicities of the Enantiomers of 2′-Deoxy-3′-Oxa-4′-Thiocytidine and Their 5-Fluoro Analogs in Vitro. J. Med. Chem. 1995, 38, 1–4. [Google Scholar] [CrossRef]
  63. 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. [Google Scholar] [CrossRef]
  64. Jun, S.J.; Moon, M.S.; Lee, S.H.; Cheong, C.S.; Kim, K.S. Selective Monodeacetylation of Methyl 2,3,5-Tri-O-Acetyl-d-Arabinofuranoside Using Biocatalyst. Tetrahedron Lett. 2005, 46, 5063–5065. [Google Scholar] [CrossRef]
  65. 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. [Google Scholar] [CrossRef]
  66. Koreeda, M.; Houston, T.A.; Shull, B.K.; Klemke, E.; Tuinman, R.J. Iodine-Catalyzed Ferrier Reaction 1. A Mild and Highly Versatile Glycosylation of Hydroxyl and Phenolic Groups. Synlett 1995, 1995, 90–92. [Google Scholar] [CrossRef]
  67. Sasaki, K.; Wakamatsu, T.; Matsumura, S.; Toshima, K. Synthesis of Hexopyranosyl Acetates and 2,3-Disubstituted Tetrahydropyrans via Chemoselective Hydrogenation of Hex-2-Enopyranosyl Acetates. Tetrahedron Lett. 2006, 47, 8271–8274. [Google Scholar] [CrossRef]
  68. Regueira, J.; Dantas, C.; de Freitas, J.; da Silva, A.; Freitas Filho, J.; Menezes, P.; Freitas, J. Stereoselective Synthesis of 2,3-Unsaturated Pseudoglycosides Promoted by Ultrasound. Synthesis 2016, 48, 1069–1078. [Google Scholar] [CrossRef]
  69. Banaag, A.R.; Tius, M.A. Traceless Chiral Auxiliaries for the Allene Ether Nazarov Cyclization. J. Org. Chem. 2008, 73, 8133–8141. [Google Scholar] [CrossRef]
  70. Jung, S.; Inoue, A.; Nakamura, S.; Kishi, T.; Uwamizu, A.; Sayama, M.; Ikubo, M.; Otani, Y.; Kano, K.; Makide, K.; et al. Conformational Constraint of the Glycerol Moiety of Lysophosphatidylserine Affords Compounds with Receptor Subtype Selectivity. J. Med. Chem. 2016, 59, 3750–3776. [Google Scholar] [CrossRef]
  71. Sarkar, B.; Pramanik, T.; Jayaraman, N. Cyclic Disaccharide Formation Enforced by a Ring Contraction: 2,3-Dideoxy Pyranoside Glycoside Donor to a Furanoside Macrocycle. J. Org. Chem. 2023, 88, 670–674. [Google Scholar] [CrossRef]
  72. Lu, Y.-S.; Li, Q.; Zhang, L.-H.; Ye, X.-S. Highly Direct α-Selective Glycosylations of 3,4-O-Carbonate-Protected 2-Deoxy- and 2,6-Dideoxythioglycosides by Preactivation Protocol. Org. Lett. 2008, 10, 3445–3448. [Google Scholar] [CrossRef]
  73. Yakura, T.; Horiuchi, Y.; Nishimura, Y.; Yamada, A.; Nambu, H.; Fujiwara, T. Efficient Oxidative Cleavage of Tetrahydrofuran-2-methanols to γ-Lactones by a 2-Iodobenzamide Catalyst in Combination with Oxone®. Adv. Synth. Catal. 2016, 358, 869–873. [Google Scholar] [CrossRef]
  74. Hu, L.; Schaufelberger, F.; Zhang, Y.; Ramström, O. Efficient Asymmetric Synthesis of Lamivudine via Enzymatic Dynamic Kinetic Resolution. Chem. Commun. 2013, 49, 10376–10378. [Google Scholar] [CrossRef]
  75. Böhringer, M.; Roth, H.; Hunziker, J.; Göbel, M.; Krishnan, R.; Giger, A.; Schweizer, B.; Schreiber, J.; Leumann, C.; Eschenmoser, A. Oligonucleotide Aus 2′,3′-Dideoxy-β-d-glucopyranosyl-Bausteinen (‘Homo-DNS’): Herstellung. Helv. Chim. Acta 1992, 75, 1416–1477. [Google Scholar] [CrossRef]
  76. Khan, A.R.; Mulligan, K.X.; Redda, K.K.; Ollapally, A.P. Synthesis of 3′-azido-2′,3′-dideoxy-4′-ketohexopyranoid analogues as possible antiviral nucleosides. Synth. Commun. 2002, 32, 1023–1030. [Google Scholar] [CrossRef]
  77. Chandra, T.; Broderick, W.E.; Broderick, J.B. An Efficient Deprotection of N -Trimethylsilylethoxymethyl (SEM) Groups from Dinucleosides and Dinucleotides. Nucleos. Nucleot. Nucl. Acids 2010, 29, 132–143. [Google Scholar] [CrossRef]
  78. Palomino, E.; Meltsner, B.R.; Kessel, D.; Horwitz, J.P. Synthesis and in Vitro Evaluation of Some Modified 4-Thiopyrimidine Nucleosides for Prevention or Reversal of AIDS—Associated Neurological Disorders. J. Med. Chem. 1990, 33, 258–263. [Google Scholar] [CrossRef] [PubMed]
  79. Jo, K.; Frode, R.; Kjell, U. 2′,3′ Dideoxyribofuranoxide Derivatives. WO Patent WO1988007532A1, 6 October 1988. [Google Scholar]
  80. Kaulinya, L.T.; Liepin’sh; Lidak, M.Y.; Zhuk, R.A. Analogs of Pyrimidine Nucleosides. Racemic 2,3-Dideoxynucleosides and Their Derivatives. Chem. Heterocycl. Compd. 1982, 18, 85–93. [Google Scholar] [CrossRef]
  81. Coates, J.A.V.; Mutton, I.M.; Penn, C.R.; Storer, R.; Williamson, C. 1,3-Oxathiolane Nucleoside Analogues. WO Patent WO1991017159A1, 14 November 1991. [Google Scholar]
Scheme 1. General scheme of the N-glycosidation reaction.
Scheme 1. General scheme of the N-glycosidation reaction.
Ijms 25 09030 sch001
Scheme 2. Silane/I2-mediated N-glycosidation reactions.
Scheme 2. Silane/I2-mediated N-glycosidation reactions.
Ijms 25 09030 sch002
Scheme 3. Et3SiI and HI as activating agents of glycoside acetates.
Scheme 3. Et3SiI and HI as activating agents of glycoside acetates.
Ijms 25 09030 sch003
Scheme 4. Et3SiH/I2-mediated N-glycosidation: preliminary screening. α:β ratios were assigned by comparative 1H NMR analysis with literature data (see Supplementary Materials for details).
Scheme 4. Et3SiH/I2-mediated N-glycosidation: preliminary screening. α:β ratios were assigned by comparative 1H NMR analysis with literature data (see Supplementary Materials for details).
Ijms 25 09030 sch004
Scheme 5. (a) Effect of temperature, time, nucleobase and eq. of activating reagents on the stereoselectivity of the N-glycosidation of hexopyranosides 5 and 6. (b) Substrate-dependent effect of the temperature of step 2 on the stereoselective outcome of the N-glycosidations due to anomerization reactions. α:β ratios were assigned by comparative 1H NMR analyses with the literature data (see Supplementary Materials for details). a If a non-anhydrous reaction solvent was used, up to 90% of disaccharide 16 was recovered (see Supplementary Materials).
Scheme 5. (a) Effect of temperature, time, nucleobase and eq. of activating reagents on the stereoselectivity of the N-glycosidation of hexopyranosides 5 and 6. (b) Substrate-dependent effect of the temperature of step 2 on the stereoselective outcome of the N-glycosidations due to anomerization reactions. α:β ratios were assigned by comparative 1H NMR analyses with the literature data (see Supplementary Materials for details). a If a non-anhydrous reaction solvent was used, up to 90% of disaccharide 16 was recovered (see Supplementary Materials).
Ijms 25 09030 sch005
Scheme 6. Stereoselectivity of the N-glycosidation reaction: mechanistic hypotheses.
Scheme 6. Stereoselectivity of the N-glycosidation reaction: mechanistic hypotheses.
Ijms 25 09030 sch006
Scheme 7. N-Glycosidation reactions of five-membered substrates.
Scheme 7. N-Glycosidation reactions of five-membered substrates.
Ijms 25 09030 sch007
Scheme 8. The use of the Et3SiH/I2-mediated N-glycosidation reaction in the synthesis of antiretroviral nucleoside apricitabine.
Scheme 8. The use of the Et3SiH/I2-mediated N-glycosidation reaction in the synthesis of antiretroviral nucleoside apricitabine.
Ijms 25 09030 sch008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cimafonte, M.; Esposito, A.; De Fenza, M.; Zaccaria, F.; D’Alonzo, D.; Guaragna, A. Synthesis of Natural and Sugar-Modified Nucleosides Using the Iodine/Triethylsilane System as N-Glycosidation Promoter. Int. J. Mol. Sci. 2024, 25, 9030. https://doi.org/10.3390/ijms25169030

AMA Style

Cimafonte M, Esposito A, De Fenza M, Zaccaria F, D’Alonzo D, Guaragna A. Synthesis of Natural and Sugar-Modified Nucleosides Using the Iodine/Triethylsilane System as N-Glycosidation Promoter. International Journal of Molecular Sciences. 2024; 25(16):9030. https://doi.org/10.3390/ijms25169030

Chicago/Turabian Style

Cimafonte, Martina, Anna Esposito, Maria De Fenza, Francesco Zaccaria, Daniele D’Alonzo, and Annalisa Guaragna. 2024. "Synthesis of Natural and Sugar-Modified Nucleosides Using the Iodine/Triethylsilane System as N-Glycosidation Promoter" International Journal of Molecular Sciences 25, no. 16: 9030. https://doi.org/10.3390/ijms25169030

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop