Next Article in Journal
Enzyme Molecules in Solitary Confinement
Next Article in Special Issue
Unusual Structure-Energy Correlations in Intramolecular Diels–Alder Reaction Transition States
Previous Article in Journal
Identification, Characterization, and Immobilization of an Organic Solvent-Stable Alkaline Hydrolase (PA27) from Pseudomonas aeruginosa MH38
Previous Article in Special Issue
Synthesis with Perfect Atom Economy: Generation of Furan Derivatives by 1,3-Dipolar Cycloaddition of Acetylenedicarboxylates at Cyclooctynes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enantiomerically Pure Phosphonated Carbocyclic 2'-Oxa-3'-Azanucleosides: Synthesis and Biological Evaluation

by
Roberto Romeo
1,*,
Caterina Carnovale
1,*,
Salvatore V. Giofrè
1,†,
Giulia Monciino
2,†,
Maria A. Chiacchio
2,†,
Claudia Sanfilippo
3,† and
Beatrice Macchi
4,†
1
Dipartimento Scienze del Farmaco e Prodotti per la Salute, Università di Messina, Via S.S. Annunziata, 98168 Messina, Italy
2
Dipartimento Scienze del Farmaco, Università di Catania, Viale A. Doria 6, 95125 Catania, Italy
3
Istituto di Chimica Biomolecolare del CNR, Via P. Gaifami 18, 95126 Catania, Italy
4
Dipartimento di Medicina dei Sistemi, Università di Roma "Tor Vergata", 00133 Roma, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2014, 19(9), 14406-14416; https://doi.org/10.3390/molecules190914406
Submission received: 30 July 2014 / Revised: 26 August 2014 / Accepted: 29 August 2014 / Published: 12 September 2014
(This article belongs to the Special Issue Cycloaddition Chemistry)

Abstract

:
Starting from enantiomeric pure 1-[(3S,5R)- and 1-[(3R,5S)-3-(hydroxymethyl)-2-methylisoxazolidin-5-yl]-5-methylpyrimidine-2,4(1H,3H)-diones (−)7a and (+)7b, obtained by lipase-catalyzed resolution, pure diethyl{[(3S,5R)-2-methyl-5-(5-methyl-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)isoxazolidin-3-yl]methyl}phosphonate (−)12a and diethyl{[(3R,5S)-2-methyl-5-(5-methyl-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)isoxazolidin-3-yl]methyl}phosphonate (+)12b have been synthesized. The obtained compounds showed no cytotoxic activity versus the U937 cell line in comparison with AZT, and were poorly able to inhibit HIV infection in vitro.

1. Introduction

Continuous efforts in the development of new antiviral agents are a consequence of the urgent demand for new therapeutic agents in which improved biological activity against viruses is matched with low toxicity towards host cells. In this context, particular interest has been focused on the synthesis and the biological activity of nucleoside analogues, in which structural modifications of the heterocyclic bases and/or the sugar moiety of natural nucleosides have been performed [1,2,3,4,5,6,7,8,9,10,11,12,13]. A remarkable impulse for research on nucleoside analogues arose from the urgent need to find a therapeutic approach to combat human immunodeficiency virus (HIV) infections. In the eighties, research on antiretrovirals was very fruitful and in 1986 the in vitro anti-HIV activity of the prototype antiretroviral drug, 3'-azido-3'-deoxythymidine (AZT, zidovudine) was demonstrated by Mitsuya and Broder [14,15].
Since the discovery of AZT, a number of nucleoside analogues have been designed, sharing structural similarities with each other and mimicking endogenous nucleosides. In this context, modified nucleosides, in which alternative carbon or heterocyclic systems replaces the furanose ring, have attracted special interest by virtue of their biological action as antiviral or anticancer agents [16,17,18,19,20,21,22,23,24,25]. Among these, N,O-nucleosides 17, characterized by the presence of an isoxazolidine moiety (carbocyclic-2'-oxo-3'-azanucleosides), have emerged as an interesting class of dideoxynucleoside analogues endowed with potential pharmacological activity (Figure 1) [26,27,28,29,30,31,32]. These compounds, which mimic natural nucleosides, exert their antiviral activity by competitive reversible inhibition of reverse transcriptase, acting as viral DNA chain terminators, or they behave as antimetabolites, competing with physiological nucleosides and consequently interacting with a large number of intracellular targets to induce cytotoxicity.
Figure 1. N,O-modified nucleosides.
Figure 1. N,O-modified nucleosides.
Molecules 19 14406 g001
The biological activity of nucleoside analogues (NA) showing antiviral properties is strictly linked to their conversion, through cellular enzymes, to the corresponding biologically active triphosphate form, which interacts with viral RT or interferes with cell growth, slowing the cell cycle progression. One of the metabolic drawbacks of nucleoside analogues (NA) is the retention of their stability following the triphosphorylation inside the host cell. To overcome the instability of triphosphate nucleic acid (NA), several strategies have been proposed to increase their resistance toward phosphohydrolases or to ensure a more efficient phosphorylation within the target cells [33,34,35,36,37,38,39,40,41,42,43,44,45,46].
Instability of the phosphate forms of nucleoside analogues (NA) has been, at least partially, overcome by the introduction in their molecular structures of phosphonate groups. With the aim of bypassing the first limiting step of phosphorylation [47,48,49], we have synthesized a series of racemic phosphonated N,O-nucleosides (PCOANs) 4, as mimics of monophosphate nucleosides [29,30,31], by exploitation of the 1,3-dipolar cycloaddition methodology [50,51], starting from nitrones containing a phosphonic group.
PCOANs 4 show low levels of cytotoxicity and exert a specific inhibitor activity on two different RT: these compounds have been proposed to ensure long lasting control of HTLV-1, an oncogen retrovirus associated with adult leukemia/limphoma (ATLL) and with myelopathy, tropical spastic paraparesis. By considering the epidemiological and pathogenetic relevance of HIV infection and that HTLV-1 and HIV are both lymphotropic retroviruses, we decided to investigate the effect of PCOANs on an in vitro model of HIV transmission.
Both enantiomeric purity and absolute configuration could be key factors in determining the physiological activity of a drug [52]; thus, in this paper, we have investigated the development of an efficient approach to obtain PCOANs in enantiomerically pure form, in order to test the biological activity of pure enantiomers.

2. Results and Discussion

Our synthetic approach to the pure enantiomers diethyl{[(3S,5R)-2-methyl-5-(5-methyl-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)isoxazolidin-3-yl]methyl}phosphonate (12a) and diethyl-{[(3R,5S)-2-methyl-5-(5-methyl-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)isoxazolidin-3-yl]methyl}phosphonate (12b) is described in Scheme 1 and Scheme 2.
Scheme 1. Synthesis of racemic N,O-thymine nucleoside and its double enzymatic resolution.
Scheme 1. Synthesis of racemic N,O-thymine nucleoside and its double enzymatic resolution.
Molecules 19 14406 g003
Reagents and conditions: (a) Lipozyme, 1,4-dioxane, vinyl benzoate, 45 °C; (b) NH4OH/MeOH, 50 °C, 8 h, lipozyme, vinyl propionate, 45 °C; (c) NH4OH/MeOH, 50 °C, 8 h.
Thus, the racemic N,O-nucleoside isoxazolidine 4, obtained by 1,3-dipolar cycloaddition reaction of C-[(tert-butyldiphenylsilyl)oxy]-N-methylnitrone with vinyl acetate, followed by Hilbert–Jones nucleosidation with silylated thymine and TBAF treatment, was converted into the enantiomeric pure 1-[(3S,5R)-3-(hydroxymethyl)-2-methylisoxazolidin-5-yl]-5-methylpyrimidine-2,4(1H,3H)-dione (−)4a and 1-[(3R,5S)-3-(hydroxymethyl)-2-methylisoxazolidin-5-yl]-5-methylpyrimidine-2,4(1H,3H)-dione (+)4b by lipase-catalyzed resolution in 1,4-dioxane [53].
Scheme 2. Synthesis of enantiopure phosphonated N,O-thymine nucleoside.
Scheme 2. Synthesis of enantiopure phosphonated N,O-thymine nucleoside.
Molecules 19 14406 g004
Reagents and conditions: (a) Tosyl chloride, TEA, CH2Cl2, rt, 24 h; (b) NaI, acetone, reflux, 72 h; (c) P(OEt)3, 180 °C, 16 h.
The subsequent reaction of (−)4a with tosyl chloride affords the corresponding tosylate (−)10a, which was transformed into the 2-methyl-5-(5-methyl-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)isoxazolidin-3-yl]methyl}phosphonate (−)12a by reaction with NaI in acetone, followed by triethylphosphite treatment (Scheme 2).
Analogously, compound (+)4b was converted in (+)12b by similar reactions. HPLC chromatography showed the absence of any diasteromeric product, which could be originated from the sequence of the reaction. NOE measurements confirm that the configurations of the reagents is maintained in the obtained products.
The absolute configuration of (3S,5R) was assigned to compound (−)12a according to its origin from (−)4a in a reaction pathway which does not involve any change to the stereogenic centers. CD spectra, reported in Figure 2, confirm this assumption: the CD spectrum obtained for (−)12a was found to be a good approximation matching that registered for (3S,5R) (−)4a, thus allowing the assignment of the same absolute configuration to the stereogenic centers of both compounds. Analogous data confirm as (3R,5S) the absolute configuration of (+)12b.
The biological activity of enantiomers (−)12a and (−)12b was tested by assessing their cytotoxic effect towards a monocytoid human cell line, U937, and their specific activity on HIV infection in vitro. 3'-Azido-2',3'-dideoxythymidine (AZT) was used as positive control, since it is a prototype of nucleoside analogs (NRTI) acting as chain terminator and endowed with potent anti-HIV activity in vitro. The obtained results indicated that both enantiomers 12 showed no cytotoxicity versus the U937 cell line in comparison with AZT, which exhibited a IC50 of 800 µM. On the other hand, the compounds were poorly able to inhibit HIV infection in vitro, showing a IC50 > of 1000 µM with respect to AZT. Although the tested molecules showed poor specific anti-HIV activity, their low toxicity encourages us to continue search for cytostatic compounds in the phosphonated N,O-nucleoside class with special attention to different heterocyclic systems as mimics of natural nucleobases.
Figure 2. CD Spectra recorded in methanol of (−)4a, (−)12a and (+)12b.
Figure 2. CD Spectra recorded in methanol of (−)4a, (−)12a and (+)12b.
Molecules 19 14406 g002

3. Experimental Section

3.1. General Information

Solvents and reagents were used as received from commercial sources. Melting points were determined with a Kofler apparatus. Thin layer chromatographic separations were performed on Merck silica gel 60-F254 precoated aluminum plates (Merk, Darmstadt, Germany). Flash chromatography was accomplished on Merck silica gel (200–400 mesh). Preparative separations were carried out by a Buchi C-601MPLC (BUCHI Italia S.r.l., Milano, Italy), using Merck silica gel 0.040–0.063 mm and the eluting solvents were delivered by a pump at the flow rate of 3.5–7.0 mL/min. HRMS were determined with a TSQ Quantum XLS Triple Quadrupole GC-MS/MS (Thermo Scientific, Waltham, MA, USA). 1H-NMR (500 MHz) and 13C-NMR (125 MHz) spectra were recorded in CDCl3, on a Varian 500 instrument (Agilent Technologies, Palo Alto, CA, USA). Chemical shifts (δ) are reported in ppm relative to TMS and coupling constants (J) in Hz. CD spectra 3 were registered at 20 °C in methanol (0.1 cm cell length) on a JASCO J-810 spectropolarimeter (JASCO, Europe S.r.l., Lecco, Italy).

3.2. Synthesis of ((3S,5R)-5-(3,4-Dihydro-5-methyl-2,4-dioxopyrimidin-1(2H)-yl)-2-methylisoxa-zolidin- 3-yl)methyl 4-methylbenzenesulphonate (−);)10a

To a solution of compound (−)4a (0.400 g, 1.6 mmol) in CH2Cl2 (20 mL), Et3N (0.25 mL, 1.8 mmol) was added. TsCl (0.349 g, 1.8 mmol) was then added slowly at 0 °C and the reaction mixture was stirred at room temperature for 24 h. The solvent was removed under vacuum and the residue was purified by MPLC, using CH2Cl2/MeOH 99:1 as eluent, to give (−)10a in 70% yield (0.44 g). Compound (−)10a was recrystallized from hexane. White solid, m.p. 134–136 °C, Molecules 19 14406 i001 = −84.9 (c 0.25, MeOH). 1H-NMR (CDCl3): δ = 1.77 (s, 3H); 1.93–2.04 (m, 1H), 2.32 (s, 3H); 2.63 (s, 3H); 2.77–2.91 (m, 2H); 3.89–4.07 (m, 2H), 5.94–6.02 (m, 1H), 7.23 (d, J = 8.2 Hz, 2H), 7.46 (s, 1H); 7.63 (d, J = 8.2 Hz, 2H), 9.37 (bs, 1H). 13C-NMR (CDCl3): δ = 8.54, 17.48, 36.53, 40.26, 61.81, 63.64, 89.98, 106.58, 123.67, 125.90, 131.73, 138.66, 144.09, 151.95, 160.09. HRMS: calcd for C17H21N3O6SNa+ 418.1043, found 418.1050.

3.3. Synthesis of (3R,5S)-1-(3-(Iodomethyl)-2-methylisoxazolidin-5-yl)-5-methylpyrimidin-2,4-(1H,3H) dione (−)11a

Compound (−)10a (0.32 g, 0.81 mmol) was added to a solution of NaI (0.600g, 4 mmol) in acetone (20 mL) and the reaction mixture was refluxed for 24 h. After cooling, acetone was removed in vacuo and chloroform was added to the residue. The resulting organic phase was filtered and evaporated. The residue was purified by MPLC, using CH2Cl2: MeOH 99:1 as eluent, to give the compound (−)11a in 93% yield (0.260 g). Compound (−)11a was recrystallized from hexane. Pale yellow solid; m.p. 157–159 °C. Molecules 19 14406 i001 = −88.50 (c 0.5, MeOH); 1H-NMR (CDCl3): δ = 1.96 (s, 3H), 2.20 (ddd, J = 13.7, 9.2, 4.1 Hz, 1H), 2.51–2.64 (m, 1H), 2.70 (s, 3H), 3.06 (m, 1H), 3.11 (dd, J = 10.7, 4.1Hz, 1H), 3.25 (dd, J = 10.7, 3.2 Hz, 1H), 6.16 (dd, J = 7.8, 4.1 Hz, 1H), 7.76 (s, 1H), 8.54 (bs, 1H). 13C-NMR (CDCl3): δ = 8.44; 25.45, 38.99, 41.18, 63.28, 77.44, 110.20, 132.30, 150.20, 163.4. HRMS: calcd for C10H14N3O3INa+ 373.9972, found 373.9981

3.4. Synthesis of Diethyl{(1'R,4'S)-1'-[[(5-methyl-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl]-3'-methyl-2'-oxa-3'-aza-cyclopent-4'-yl]}methylphosphonate (−)12a

Compound (−)12a was recrystallized from hexane. White solid, m.p. 134–136 °C. Molecules 19 14406 i001 = −83.4 (c 0.50, MeOH); 1H-NMR (CDCl3): δ= 1.30 (dt, 6H, J = 3.6 and 7.1 Hz), 1.89 (ddd, 1H, J = 10.2, 15.0 and 18.3 Hz, H5'a), 1.93 (d, 3H, J = 1.3Hz), 2.08 (ddd, 1H, J = 3.2, 15.0 and 20.6 Hz, H5'b), 2.23 (ddd, 1H,J = 4.6, 10.1 and 13.8 Hz, H6'a), 2.75 (s, 3H, N-CH3), 2.98 (dddd, 1H, J = 3.2, 7.0, 10.1 and 10.2 Hz, H4'), 3.18 (ddd, 1H, J = 7.0, 7.9 and 13.8 Hz, , H6'b), 4.10–4.17 (m, 4H), 6.20 (dd, 1H, J = 4.6 and 7.9 Hz, H1'), 7.66 (q, 1H, J = 1.3 Hz, H6), 9.56 (bs, 1H, NH); 13C-NMR (CDCl3); δ: 12.57, 16.33, 16.38, 27.88 (d, J = 143.2 Hz), 42.68, 44.35, 61.93, 61.98, 63.37, 81.94, 110.66, 135.96, 150.56, 164.11 CD: λext 200 (Δε +1.49), 206 (Δε +3.82), 219 (Δε −0.11), 233(Δε +0.84), 249 (Δε +0.38), 275 (Δε +2.82), 290 (Δε +0.67), 307 (Δε −0.42), 360 (Δε −0.08).
An analogous reaction pathway, starting from (+)4b, afforded enantiomerically pure (+)10b, (+)11b and (+)12b. CD spectra of (+)12b: λext 200 (Δε −1.33), 205 (Δε −4.47), 219 (Δε −0.14), 233(Δε −1.24), 249 (Δε −0.70), 275 (Δε −3.19), 290 (Δε −0.96), 305 (Δε +0.16), 360 (Δε +0.20).

3.5. Biological Assay

HIV infection was carried on by using a stable T cell line (CEM) containing a plasmid encoding a green fluorescence protein (GFP) driven by the HIV-1 long terminal repeat [54]. Infection was carried on as previously b shown with some modification [55]. Briefly, 5 × 105 CEM-GFP were infected with a volume of supernatant from HIV chronically infected H9 cells equivalent to 20 ng/mL of HIV p24, for 2 h in 100 µL in DMSO in presence of 1000, 100, 10 and 1 µM concentration of compounds. Then medium was added and the cultures were incubated for 72 h. The inhibition was assessed on the basis of GFP expression in the different culture conditions. Cytotoxicity assays were performed by MTS assay kit (Promega Corporation, Madison, WI, USA). Briefly Inhibition of cell metabolic activity revealed by reduction of the oxidative burst was detected through formazan product formation, using a commercial colorimetric kit (MTS [3,4-(5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium salt]) (Cell Titer 96 Aqueous One Solution; Promega). The assay was performed by seeding 1 × 104 U937 cells in the presence or absence of the different compounds at concentrations ranging from 1000 to 1 µM.

4. Conclusions

In this work, an efficient synthetic pathway towards diethyl{[(3S,5R)-2-methyl-5-(5-methyl-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)isoxazolidin-3-yl]methyl}-phosphonate (12a) and diethyl{[(3R,5S)2-methyl-5-(5-methyl-2,4-dioxo-3,4-dihydro-pyrimidin-1(2H)-yl)isoxazolidin-3-yl]methyl}phosphonate (12b) has been developed, starting from enantiomerically pure 1-[(3S,5R)-3-(hydroxymethyl)-2-methylisoxazolidin-5-yl]-5-methylpyrimidine-2,4(1H,3H)-dione ((−)4a) and 1-[(3R,5S)-3-(hydroxymethyl)-2-methylisoxazolidin-5-yl]-5-methylpyrimidine-2,4(1H,3H)-dione ((+)4b) by tosylation, iodination and Arbuzov reaction. The obtained compounds show no cytotoxicity versus the U937 cell line in comparison with AZT, but they were found to be poorly able to inhibit HIV infection in vitro.

Acknowledgments

We gratefully acknowledge the Italian Ministry of Education, Universities, and research (MIUR), the Universities of Messina and Catania, and the Interuniversity Consortium for Innovative Methodologies and Processes for Synthesis (CINMPIS) for partial financial support.

Author Contributions

RR, CC and SVG designed research; BM performed biological data; RR, CC, SVG, GM and MAC performed research and analyzed the data; CS performed CD spectra of compounds; RR, CC and SVG wrote the paper. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, X.; Zhan, P.; de Clercq, E.; Liu, X. The HIV-1 non-nucleoside reverse transcriptase inhibitors (Part 5): Capravirine and its analogues. Curr. Med. Chem. 2012, 19, 6138–6149. [Google Scholar] [CrossRef]
  2. Saag, M.S. New and investigational antiretroviral drugs for HIV infection: Mechanisms of action and early research findings. Top. Antivir. Med. 2012, 20, 162–167. [Google Scholar]
  3. De Clercq, E. Highlights in the discovery of antiviral drugs: A personal retrospective. J. Med. Chem. 2010, 53, 1438–1450. [Google Scholar] [CrossRef]
  4. Mehellou, Y.; de Clercq, E. Twenty-six years of anti-HIV drug discovery: Where do we stand and where do we go. J. Med. Chem. 2010, 53, 521–538. [Google Scholar] [CrossRef]
  5. Stambasky, J.; Hocek, M.; Kocovsky, P. C-nucleosides: Synthetic strategies and biological applications. Chem. Rev. 2009, 109, 6729–6764. [Google Scholar] [CrossRef]
  6. Galmarini, C.M.; Popowycz, F.; Joseph, B. Cytotoxic nucleoside analogues: Different strategies to improve their clinical efficacy. Curr. Med. Chem. 2008, 15, 1072–1082. [Google Scholar] [CrossRef]
  7. Kool, E.T. Modified DNA bases: Probing base-pair recognition by polymerases. Modified nucleosides. In Biochemistry,Biotechnology and Medicine; Herdewijn, P., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2008. [Google Scholar]
  8. Balestrieri, E.; Matteucci, C.; Ascolani, A.; Piperno, A.; Romeo, R.; Romeo, G.; Chiacchio, U.; Mastino, A.; Macchi, B. Effect of phosphonated carbocyclic 2'-oxa-3'-aza-nucleoside on human T-cell leukemia virus type 1 infection in vitro. Antimicrob. Agents Chemother. 2008, 52, 54–64. [Google Scholar] [CrossRef] [Green Version]
  9. De Clercq, E. New approaches toward anti-HIV chemotherapy. J. Med. Chem. 2005, 48, 1297–1313. [Google Scholar] [CrossRef]
  10. Sari, O.; Roy, V.; Agrofoglio, L.A. Nucleosides modified at the base moiety. In Chemical Synthesis of Nucleoside Analogues; 1st ed.; Merino, P., Ed.; Wiley-VCH: Hoboken, NJ, USA, 2013; p. 49. [Google Scholar]
  11. Romeo, R.; Carnovale, C.; Rescifina, A.; Chiacchio, M.A. Phosphonated nucleoside analogues. In Chemical Synthesis of Nucleoside Analogues; 1st ed.; Merino, P., Ed.; Wiley-VCH: Hoboken, NJ, USA, 2013; p. 163. [Google Scholar]
  12. Chiacchio, U.; Corsaro, A.; Giofrè, S.V.; Romeo, G. Isoxazolidinyl nucleosides. In Chemical Synthesis of Nucleoside Analogues; 1st ed.; Merino, P., Ed.; Wiley-VCH: Hoboken, NJ, USA, 2013; p. 781. [Google Scholar]
  13. Peyrottes, S.; Perigaud, C. Mononucleotide prodrug synthetic strategies. In Chemical Synthesis of Nucleoside Analogues; 1st ed.; Merino, P., Ed.; Wiley-VCH: Hoboken, NJ, USA, 2013; p. 229. [Google Scholar]
  14. Mitsuya, H.; Weinhold, K.J.; Furman, P.A.; St. Clair, M.H.; Lehrman, S.N.; Gallo, R.C.; Bolognesi, D.; Barry, D.W.; Broder, S. 3'-Azido-3'-deoxythymidine (BW A509U): An antiviral agent that inhibits the infectivity and cytopathic effect of human T-lymphotropic virus type III/lymphadenopathy-associated virus in vitro. Proc. Natl. Acad. Sci. USA 1985, 82, 7096–7100. [Google Scholar] [CrossRef]
  15. Mitsuya, H.; Broder, S. Inhibition of the in vitro infectivity and cytopathic effect of human T-lymphotrophic virus type III/lymphadenopathy-associated virus (HTLV-III/LAV) by 2',3'-dideoxynucleosides. Proc. Natl. Acad. Sci. USA 1986, 83, 1911–1915. [Google Scholar] [CrossRef]
  16. Romeo, G.; Chiacchio, U.; Corsaro, A.; Merino, P. Chemical synthesis of heterocyclic-sugar nucleoside analogues. Chem. Rev. 2010, 110, 3337–3370. [Google Scholar] [CrossRef]
  17. Chiacchio, U.; Padwa, A.; Romeo, G. Cycloaddition methodology: A useful entry towards biologically active heterocycle. Curr. Org. Chem. 2009, 13, 422–443. [Google Scholar] [CrossRef]
  18. Piperno, A.; Chiacchio, M.A.; Iannazzo, D.; Romeo, R. Synthesis and biological activity of phosphonated nucleosides: Part 1. Furanose, carbocyclic and heterocyclic analogues. Curr. Med. Chem. 2006, 13, 3675–3695. [Google Scholar] [CrossRef]
  19. Merino, P.; Tejero, T.; Unzurrunzaga, F.J.; Franco, S.; Chiacchio, U.; Saita, M.G.; Iannazzo, D.; Piperno, A.; Romeo, G. An efficient approach to enantiomeric isoxazolidinyl analogues of tiazofurin based on nitrone cycloadditions. Tetrahedron Asymmetry 2005, 16, 3865–3876. [Google Scholar] [CrossRef]
  20. Chiacchio, U.; Genovese, F.; Iannazzo, D.; Librando, V.; Merino, P.; Rescifina, A.; Romeo, R.; Procopio, A.; Romeo, G. Diastereoselective synthesis of homo-N,O-nucleosides. Tetrahedron 2004, 60, 441–448. [Google Scholar] [CrossRef]
  21. Chiacchio, U.; Corsaro, A.; Iannazzo, D.; Piperno, A.; Pistarà, V.; Rescifina, A.; Romeo, R.; Sindona, G.; Romeo, G. Diastereo- and enantioselective synthesis of N,O-nucleosides. Tetrahedron Asymmetry 2003, 14, 2717–2723. [Google Scholar] [CrossRef]
  22. Chiacchio, U.; Corsaro, A.; Mates, J.; Merino, P.; Piperno, A.; Rescifina, A.; Romeo, G.; Tejero, T. Isoxazolidine analoguese of pseudouridine: A new class of modified nucleosides. Tetrahedron 2003, 59, 4733–4738. [Google Scholar] [CrossRef]
  23. Chiacchio, U.; Corsaro, A.; Pistarà, V.; Rescifina, A.; Iannazzo, D.; Piperno, A.; Romeo, G.; Romeo, R.; Grassi, G. Diastereoselective synthesis of N,O-psiconucleosides, a new class of modified nucleosides. Eur. J. Org. 2002, 7, 1206–1212. [Google Scholar]
  24. Chiacchio, U.; Corsaro, A.; Iannazzo, D.; Piperno, A.; Procopio, A.; Rescifina, A.; Romeo, G.; Romeo, R. A stereoselective approach to isoxazolidinyl nucleosides. Eur. J. Org. Chem. 2001, 10, 1893–1898. [Google Scholar]
  25. Chiacchio, U.; Gumina, G.; Rescifina, A.; Romeo, R.; Uccella, N.; Casuscelli, F.; Piperno, A.; Romeo, G. Modified dideoxynucleosides: Synthesis of 2'-N-alkyl-3'-hydroxyalkyl-1',2'-isoxazolidinyl thymidine and 5-fluorouridine derivatives. Tetrahedron 1996, 52, 8889–8898. [Google Scholar] [CrossRef]
  26. Romeo, R.; Giofrè, S.V.; Macchi, B.; Balestrieri, E.; Mastino, A.; Merino, P.; Carnovale, C.; Romeo, G.; Chiacchio, U. Truncated reverse isoxazolidinyl nucleosides: A new class of allosteric HIV-1 reverse transcriptase inhibitors. ChemMedChem 2012, 7, 565–569. [Google Scholar] [CrossRef]
  27. Romeo, R.; Carnovale, C.; Giofrè, S.V.; Romeo, G.; Macchi, B.; Frezza, C.; Marino-Merlo, F.; Pistarà, V.; Chiacchio, U. Truncated phosphonated C-1'-branched N,O-nucleosides: A new class of antiviral agents. Bioorg. Med. Chem. 2012, 20, 3652–3657. [Google Scholar] [CrossRef] [Green Version]
  28. Piperno, A.; Giofrè, S.V.; Iannazzo, D.; Romeo, R.; Romeo, G.; Chiacchio, U.; Rescifina, A.; Piotrowska, D.G. Synthesis of C-4'-truncated phosphonated carbocyclic 2'-oxa-3'-azanucleosides as antiviral agents. J. Org. Chem. 2010, 75, 2798–2805. [Google Scholar] [CrossRef]
  29. Chiacchio, U.; Rescifina, A.; Iannazzo, D.; Piperno, A.; Romeo, R.; Borrello, L.; Sciortino, M.T.; Balestrieri, E.; Macchi, B.; Mastino, A.; et al. Phosphonated carbocyclic 2'-Oxa-3'-azanucleosides as new antiretroviral agents. J. Med. Chem. 2007, 50, 3747–3750. [Google Scholar] [CrossRef] [Green Version]
  30. Chiacchio, U.; Iannazzo, D.; Piperno, A.; Romeo, R.; Romeo, G.; Rescifina, A.; Saglimbeni, M. Synthesis and biological evaluation of phosphonated carbocyclic 2'-oxa-3'-aza-nucleosides. Bioorg. Med. Chem. 2006, 14, 955–959. [Google Scholar] [CrossRef]
  31. Chiacchio, U.; Balestrieri, E.; Macchi, B.; Iannazzo, D.; Piperno, A.; Rescifina, A.; Romeo, R.; Saglimbeni, M.; Sciortino, M.T.; Valveri, V.; et al. Synthesis of phosphonated carbocyclic 2'-Oxa-3'-aza-nucleosides: Novel inhibitors of reverse transcriptase. J. Med. Chem. 2005, 48, 1389–1394. [Google Scholar] [CrossRef] [Green Version]
  32. Chiacchio, U.; Corsaro, A.; Iannazzo, D.; Piperno, A.; Pistarà, V.; Rescifina, A.; Romeo, R.; Valveri, V.; Mastino, A.; Romeo, G. Enantioselective syntheses and cytotoxicity of N,O-nucleosides. J. Med. Chem. 2003, 46, 3696–3702. [Google Scholar] [CrossRef]
  33. Hecker, S.J.; Erion, M.D. Prodrugs of phosphates and phosphonates. J. Med.Chem. 2008, 51, 2328–2345. [Google Scholar] [CrossRef]
  34. McGuigan, C.; Derudes, M.; Bugert, J.J.; Andrei, G.; Snoecke, R.; Balzarini, J. Successful kinase bypass with new acyclovir phosphoramidate prodrugs. Bioorg. Med. Chem. Lett. 2008, 18, 4364–4367. [Google Scholar] [CrossRef]
  35. He, G.-X.; Krise, J.P.; Oliyai, R. Prodrugs: Challenges and Rewards; Springer: New York, NY, USA, 2007; pp. 223–264. [Google Scholar]
  36. Perrone, P.; Luoni, G.M.; Kelleher, M.R.; Daverio, F.; Angell, A.; Mulready, S.; Congiatu, C.; Rajyageru, S.; Martin, J.A.; Leveque, V.; et al. Application of the phosphoramidate protide approach to 4'-azidouridine confers sub-micromolar potency versus hepatitis C virus on an inactive nucleoside. J. Med. Chem. 2007, 50, 1840–1849. [Google Scholar] [CrossRef]
  37. Ariza, M.E. Current prodrug strategies for the delivery of nucleotides into cells. Drug Del. Rev. 2005, 2, 273–387. [Google Scholar]
  38. Congiatu, C.; McGuigan, C.; Jiang, W.G.; Davies, G.; Mason, M.D. Naphthyl phosphoramidate derivatives of BVdU as potential anticancer agents: Design, synthesis and biological evaluation. Nucleosides Nucleotides Nucleic Acids 2005, 24, 485–489. [Google Scholar] [CrossRef]
  39. McGuigan, C.; Harris, S.A.; Daluge, S.M.; Gudmundsson, K.S.; McLean, E.W.; Burnette, T.C.; Marr, H.; Hazen, R.; Condreay, L.D.; Johnson, L.; et al. Application of phosphoramidate pronucleotide technology to abacavir leads to a significant enhancement of antiviral potency. J. Med. Chem. 2005, 48, 3504–3515. [Google Scholar] [CrossRef]
  40. Mackman, R.L.; Cihlar, T. Prodrug strategies in the design of nucleoside and nucleotide antiviral therapeutics. Annu. Rep. Med. Chem. 2004, 39, 305–321. [Google Scholar] [CrossRef]
  41. Schultz, C. Prodrugs of biologically active phosphate esters. Bioorg. Med. Chem. 2003, 11, 885–898. [Google Scholar] [CrossRef]
  42. De Clercq, E. The clinical potential of the acyclic (and cyclic) nucleoside phosphonates.The magic of the phosphonate bond. Biochem. Pharmacol. 2011, 82, 99–109. [Google Scholar] [CrossRef]
  43. De Clercq, E.; Holy, A. Acyclic nucleoside phosphonates: A key class of antiviral drugs. Nat. Res. Drug Discov. 2005, 4, 928–940. [Google Scholar] [CrossRef]
  44. Deville-Bonne, D.; El Amri, C.; Meyer, P.; Chen, Y.X.; Agrofoglio, L.A. Human and viral nucleoside/nucleotide kinases involved in antiviral drug activation: Structural and catalytic properties. Antivir. Res. 2010, 86, 101–120. [Google Scholar] [CrossRef]
  45. De Clercq, E. In search of a selective therapy of viral infections. Antivir. Res. 2010, 85, 19–24. [Google Scholar] [CrossRef]
  46. De Clercq, E. Antiviral drug discovery: Ten more compounds, and ten more stories (part B). Med. Res. Rev. 2009, 29, 571–610. [Google Scholar] [CrossRef]
  47. Gallier, F.; Alexandre, J.A.C.; El Amri, C.; Deville-Bonne, D.; Peyrotts, S.; Perigaud, C. 5',6'-Nucleoside phosphonate analogues architecture: Synthesis and comparative evaluation towards metabolic enzymes. ChemMedChem. 2011, 6, 1094–1106. [Google Scholar] [CrossRef]
  48. Meurillon, M.; Gallier, F.; Peyrottes, S.; Périgaud, C. Developing an efficient route to the synthesis of nucleoside1-alkynylphosphonates. Tetrahedron 2009, 65, 6039–6046. [Google Scholar] [CrossRef]
  49. Gallier, F.; Peyrottes, S.; Perigaud, C. Ex-chiral-pool synthesis of β-hydroxyphosphonate nucleoside analogues. Eur. J. Org. Chem. 2007, 925–933. [Google Scholar] [CrossRef]
  50. Chiacchio, U.; Corsaro, A.; Iannazzo, D.; Piperno, A.; Procopio, A.; Rescifina, A.; Romeo, G.; Romeo, R. Intramolecular cycloadditions of α-allyloxycarbonylnitrones: Stereoselective synthesis of 3-amino-2(5H)furanones. J. Org. Chem. 2002, 67, 4380–4383. [Google Scholar] [CrossRef]
  51. Chiacchio, U.; Corsaro, A.; Pistarà, V.; Rescifina, A.; Romeo, G.; Romeo, R. An asymmetric approach to pyrrolidinone and pyrrolizidinone systems by intramolecular oxime-olefin cycloaddition. Tetrahedron 1996, 52, 7875–7884. [Google Scholar] [CrossRef]
  52. Hutt, A.J. Drug chirality and its pharmacological consequences. In Introduction to the Principles of Drug Design and Action; Smith, H.J., Williams, H., Eds.; CRC Press: Boca Raton, FL, USA, 2006; pp. 117–183. [Google Scholar]
  53. Carnovale, C.; Iannazzo, D.; Nicolosi, G.; Piperno, A.; Sanfilippo, C. Preparation of isoxazolidinyl nucleoside enantiomers by lipase-catalysed kinetic resolution. Tetrahedron Asymmetry 2009, 20, 425–430. [Google Scholar] [CrossRef]
  54. Gervaix, A.; West, D.; Leoni, L.M.; Richman, D.D.; Wong-Staal, F.; Corbeil, J. A new reporter cell line to monitor HIV infection and drug susceptibility in vitro. Proc. Natl. Acad. Sci. USA 1997, 94, 4653–4658. [Google Scholar]
  55. Balestrieri, E.; Pizzimenti, F.; Ferlazzo, A.; Giofrè, S.V.; Iannazzo, D.; Piperno, A.; Romeo, R.; Chiacchio, M.; Macchi, B. Antiviral activity of seed extract from Citrus bergamia towards human retroviruses. Bioorg. Med. Chem. 2011, 19, 2084–2089. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds are available from the authors.

Share and Cite

MDPI and ACS Style

Romeo, R.; Carnovale, C.; Giofrè, S.V.; Monciino, G.; Chiacchio, M.A.; Sanfilippo, C.; Macchi, B. Enantiomerically Pure Phosphonated Carbocyclic 2'-Oxa-3'-Azanucleosides: Synthesis and Biological Evaluation. Molecules 2014, 19, 14406-14416. https://doi.org/10.3390/molecules190914406

AMA Style

Romeo R, Carnovale C, Giofrè SV, Monciino G, Chiacchio MA, Sanfilippo C, Macchi B. Enantiomerically Pure Phosphonated Carbocyclic 2'-Oxa-3'-Azanucleosides: Synthesis and Biological Evaluation. Molecules. 2014; 19(9):14406-14416. https://doi.org/10.3390/molecules190914406

Chicago/Turabian Style

Romeo, Roberto, Caterina Carnovale, Salvatore V. Giofrè, Giulia Monciino, Maria A. Chiacchio, Claudia Sanfilippo, and Beatrice Macchi. 2014. "Enantiomerically Pure Phosphonated Carbocyclic 2'-Oxa-3'-Azanucleosides: Synthesis and Biological Evaluation" Molecules 19, no. 9: 14406-14416. https://doi.org/10.3390/molecules190914406

Article Metrics

Back to TopTop