Next Article in Journal
4-[(3,5-Dimethyl-1H-pyrazol-1-yl)methyl]-4-methyl-2-phenyl-4,5-dihydrooxazole
Previous Article in Journal
1-(Imidazo[1,2-a]pyridin-1-ium-1-yl)-2,3,4-trioxocyclobutan-1-ide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

2,6-Bis[4-(4-butylphenyl)-1H-1,2,3-triazol-1-yl]-9-dodecyl-9H-purine

Faculty of Materials Science and Applied Chemistry, Riga Technical University, P. Valdena Str. 3, LV-1048 Riga, Latvia
*
Author to whom correspondence should be addressed.
Molbank 2019, 2019(3), M1073; https://doi.org/10.3390/M1073
Submission received: 21 June 2019 / Accepted: 12 July 2019 / Published: 14 July 2019
(This article belongs to the Section Organic Synthesis)

Abstract

:
Target 2,6-bis[4-(4-butylphenyl)-1H-1,2,3-triazol-1-yl]-9-dodecyl-9H-purine has been prepared via a Cu(I)-catalyzed azide-alkyne 1,3-dipolar cycloaddition reaction between 2,6-diazido-9-dodecyl-9H-purine and 4-n-butyl(phenylacetylene) in a 29% yield. The obtained compound was fully characterized by NMR, IR and HRMS.

Graphical Abstract

1. Introduction

Purine derivatives are widely studied due to their extensive applications in biology, biochemistry and medicine. Purine and purine nucleosides are applicable in anticancer therapy [1,2,3], as antivirals [4,5,6] and as agonists and antagonists of adenosine receptors [7,8]. As a result, the development of novel purine derivatives is very in demand.
In 2013, we developed a method for the synthesis of 2,6-bis(triazolyl)purine nucleosides, which contained ribose, arabinopyranose and arabinofuranose as sugar moieties [9]. We successfully showed that these 2,6-bis(triazolyl)purine nucleosides participate in aromatic nucleophilic reactions with different N- and S-nucleophiles and that the triazolyl ring in purine C6 positions acts as a good leaving group [9,10].
With this information in hand, we decided to adapt the discovered conditions to synthesize 2,6-bis(triazolyl)purine derivatives with different alkyl groups in the N9 position of purine. We replaced the monosaccharide moiety with the simple alkylgroups and avoided the unwanted deprotection of acetyl protecting groups from hydroxyl moieties during chemical transformations.

2. Results and Discussion

The target product was obtained in three steps from 2,6-dichloropurine (1) (Scheme 1). Firstly, the Mitsunobu reaction between 2,6-dichloropurine and dodecanol was used for the synthesis of 2,6-dichloro-9-dodecyl-9H-purine. The latter was obtained in 72 h with a 48% yield. Next, chloro substituents were exchanged with azides in a nucleophilic aromatic substitution reaction yielding compound 2 in a 32% yield over two steps [11]. Then, 2,6-diazido-9-dodecyl-9H-purine 2 was used in a CuAAC reaction with 4-n-butyl(phenylacetylene). CuSO4·5H2O and sodium ascorbate as a catalytic system was used in the presence of 10% AcOH aq. solution. The reaction was carried out in DMF at 60 °C for 11 h, in a flask covered from light. Purification of product 3 by silicagel column chromatography using gradient of 1%→10% of MeCN/CHCl3 mixture gave a light yellow solid in a 29% yield. Product 3 was fully characterized by IR and 1H- and 13C-NMR (Figure 1, Figure 2 and Figure 3).
The obtained compound 3 can be further exploited in aromatic nucleophilic substitution reactions with different nucleophiles such as N-, S- and O-nucleophiles.

3. Materials and Methods

3.1. General Information

Reactions and purity of the synthesized compounds were monitored by TLC using Silicagel 60 F254 aluminum plates (Merck, Darmstadt, Germany). Visualization was accomplished by UV light. Column chromatography was performed using Silicagel 60 (0.040–0.063 mm) (Merck). Yields of products refer to chromatographically and spectroscopically homogeneous materials. Anhydrous dimethylformamide was obtained by distillation over CaH2, tetrahydrofuran was obtained by distillation over sodium. Commercial reagents were used as received.
NMR spectra were recorded on Bruker Avance 300 (300 MHz for 1H and 75.5 MHz for 13C, respectively). The proton signal for residual non-deuterated solvent (δ 7.26 ppm for CDCl3) and carbon signal (δ 77.1 ppm for CDCl3) were used as an internal reference for 1H-NMR and 13C-NMR spectra, respectively. Coupling constants are reported in Hz. Chemical shifts of signals are given in ppm and multiplicity assigned as s: singlet; d: doublet; t: triplet; m: multiplet.
The infrared spectra were recorded on a Perkin Elmer Spectrum BX. Wavelengths are given in cm−1. High resolution mass spectrometry (HRMS) analyses were carried out on a Dual-ESI Q-TOF 6520 (Agilent Technologies, Foster City, CA, USA) mass spectrometer and Agilent 1290 Infinity series UPLC system equipped with column Extend C18 RRHD 2.1 × 50 mm, 1.8 µm connected to an Agilent 6230 TOF LC/MS mass spectrometer.
For HPLC analysis we used an Agilent Technologies 1200 Series chromatograph equipped with an Agilent XDB-C18 (4.6 × 50 mm, 1.8 µm) column and Phenomenex Gemini NX (4.6 × 100 mm, 3 µm) column. Eluent A: 0.01 M KH2PO4 solution with 6% v/v MeCN added; eluent B: 0.1% TFA solution with 5% v/v MeCN added; eluent C: MeCN.

3.2. Synthesis of 2,6-Bis[4-(4-butylphenyl)-1H-1,2,3-triazol-1-yl]-9-dodecyl-9H-purine (3)

In a flask protected from daylight, 10% AcOH aq. solution (1.0 mL), CuSO4∙5H2O (24 mg, 0.096 mmol, 8 mol %) and sodium ascorbate (37 mg, 0.19 mmol, 15 mol %) were added to a solution of 2,6-diazido-9-dodecyl-9H-purine 2 (471 mg, 1.27 mmol, 1.0 equiv.) and 4-n-butyl(phenylacetylene) (1.00 mL, 5.73 mmol, 4.5 equiv.) in DMF (10 mL) and stirred for 11 h at 60 °C. The reaction mixture was evaporated under the reduced pressure and the residue was dissolved in DCM (20 mL); the organic phase was washed with brine (5 × 10 mL) and subsequently dried over anh. Na2SO4 and evaporated. Silicagel column chromatography (CHCl3/MeCN, gradient 1%→10%) provided desired compound.
Yield 250 mg, 29%.
Light yellow solid, Rf = 0.18 (CHCl3/MeCN = 20/1).
mp = 205–208 °C (decomposes).
IR (KBr) ν (cm−1): 2924, 2853, 1607, 1590.
1H-NMR (300 MHz, CDCl3) δ (ppm): 9.24 (s, 1H, H-C(triazole)), 8.88 (s, 1H, H-C(triazole)), 8.28 (s, 1H, H-C(8)), 7.96–7.84 (m, 4H, Ar), 7.32–7.26 (m, 4H, Ar), 4.42 (t, 2H, 3J = 7.2 Hz, (-CH2-)), 2.66 (t, 4H, 3J = 7.6 Hz, 2 × (-CH2-)), 2.01 (quintet, 2H, 3J = 6.7 Hz, (-CH2-)), 1.64 (quintet, 4H, 3J = 7.8 Hz, 2 × (-CH2-)), 1.48–1.31 (m, 8H, 4 × (-CH2-)), 1.31–1.16 (m, 14H, 7 × (-CH2-)), 0.95 (t, 6H, 3J = 7.3 Hz, 2 × (-CH3)), 0.86 (t, 3H, 3J = 6.6 Hz, (-CH3)).
13C-NMR (75.5 MHz, CDCl3) δ (ppm): 155.9, 148.7, 148.5, 148.3, 147.2, 145.4, 144.1, 143.8, 129.1 (2C), 127.3, 127.0, 126.3, 126.1, 122.2, 119.2, 118.6, 44.9, 35.6 (2C)*, 33.7 (2C)*, 32.0, 29.9, 29.7 (2C)*, 29.6 (2C), 29.4, 29.2, 26.8, 22.8, 22.5 (2C)*, 14.2, 14.1 (2C). (* The signal was assigned from HSQC spectrum.)
HRMS (ESI): calcd for [C41H54N10 + Na+] 709.4431, found 709.4417.

Author Contributions

M.T. and Ē.B. designed the experiments; A.Š. performed the experiments; A.Š., I.N., Ē.B. and M.T. analyzed the IR, HRMS and NMR spectral data and wrote the manuscript. All authors read and approved the final manuscript.

Funding

This research received no external funding.

Acknowledgments

Authors thank Latvian-Ukrainian bilateral cooperation project No W3956 “From the theory of azidopurine-tetrazolopurine tautomerism to its applications in medicinal and materials chemistry” for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Robak, P.; Robak, T. Older and New Purine Nucleoside Analogs for Patients with Acute Leukemias. Cancer Treat. Rev. 2013, 39, 851–861. [Google Scholar] [CrossRef] [PubMed]
  2. Kowalska, A.; Latocha, M.; Pluta, K. Synthesis and Anticancer Activity of Thiosubstituted Purines. Med. Chem. Res. 2015, 24, 3107–3116. [Google Scholar] [CrossRef] [PubMed]
  3. 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]
  4. Andrei, G.; Carter, K.; Janeba, Z.; Sampath, A.; Schang, L.M.; Tarbet, E.B.; Vere Hodge, R.A.; Bray, M.; Esté, J.A. Highlights of the 30th International Conference on Antiviral Research. Antivir. Res. 2017, 145, 184–196. [Google Scholar] [CrossRef] [PubMed]
  5. Manvar, A.; Shah, A. Diversity Oriented Efficient Access of Trisubstituted Purines via Sequential Regioselective Mitsunobu Coupling and SNAr Based C6 Functionalizations. Tetrahedron 2013, 69, 680–691. [Google Scholar] [CrossRef]
  6. De Clercq, E. Recent Highlights in the Development of New Antiviral Drugs. Curr. Opin. Microbiol. 2005, 8, 552–560. [Google Scholar] [CrossRef] [PubMed]
  7. Fredholm, B.B.; Ijzerman, A.P.; Jacobson, K.A.; Linden, J.; Mu, C.E. International Union of Basic and Clinical Pharmacology. LXXXI. Nomenclature and Classification of Adenosine Receptors—An Update. Pharmocol. Rev. 2011, 63, 1–34. [Google Scholar] [CrossRef] [PubMed]
  8. Minetti, P.; Tinti, M.O.; Carminati, P.; Castorina, M.; Assunta, M.; Cesare, D.; Di Serio, S.; Gallo, G.; Ghirardi, O.; Giorgi, F.; et al. 2-n-Butyl-9-methyl-8-[1,2,3]triazol-2-yl-9H-purin-6-ylamine and Analogues as A2A Adenosine Receptor Antagonists. Design, Synthesis, and Pharmacological Characterization. J. Med. Chem. 2005, 48, 6887–6896. [Google Scholar] [CrossRef] [PubMed]
  9. Kovaļovs, A.; Novosjolova, I.; Bizdēna, Ē.; Bižāne, I.; Skardziute, L.; Kazlauskas, K.; Jursenas, S.; Turks, M. 1,2,3-Triazoles as Leaving Groups in Purine Chemistry: A Three-Step Synthesis of N6-Substituted-2-triazolyl-adenine Nucleosides and Photophysical Properties Thereof. Tetrahedron Lett. 2013, 54, 850–853. [Google Scholar] [CrossRef]
  10. Novosjolova, I.; Bizdēna, Ē.; Turks, M. Application of 2,6-Diazidopurine Derivatives in the Synthesis of Thiopurine Nucleosides. Tetrahedron Lett. 2013, 54, 6557–6561. [Google Scholar] [CrossRef]
  11. Šišuļins, A.; Bucevičius, J.; Tseng, Y.-T.; Novosjolova, I.; Traskovskis, K.; Bizdēna, Ē.; Chang, H.-T.; Tumkevičius, S.; Turks, M. Synthesis and Fluorescent Properties of N(9)-Alkylated 2-Amino-6-triazolylpurines and 7-Deazapurines. Beilstein J. Org. Chem. 2019, 15, 474–489. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis of 2,6-bis(triazolyl)purine derivative 3.
Scheme 1. Synthesis of 2,6-bis(triazolyl)purine derivative 3.
Molbank 2019 m1073 sch001
Figure 1. 1H-NMR (300 MHz, CDCl3) spectrum.
Figure 1. 1H-NMR (300 MHz, CDCl3) spectrum.
Molbank 2019 m1073 g001
Figure 2. 13C-NMR (75.5 MHz, CDCl3) spectrum.
Figure 2. 13C-NMR (75.5 MHz, CDCl3) spectrum.
Molbank 2019 m1073 g002
Figure 3. IR spectrum (KBr) ν (cm−1).
Figure 3. IR spectrum (KBr) ν (cm−1).
Molbank 2019 m1073 g003

Share and Cite

MDPI and ACS Style

Šišuļins, A.; Bizdēna, Ē.; Turks, M.; Novosjolova, I. 2,6-Bis[4-(4-butylphenyl)-1H-1,2,3-triazol-1-yl]-9-dodecyl-9H-purine. Molbank 2019, 2019, M1073. https://doi.org/10.3390/M1073

AMA Style

Šišuļins A, Bizdēna Ē, Turks M, Novosjolova I. 2,6-Bis[4-(4-butylphenyl)-1H-1,2,3-triazol-1-yl]-9-dodecyl-9H-purine. Molbank. 2019; 2019(3):M1073. https://doi.org/10.3390/M1073

Chicago/Turabian Style

Šišuļins, Andrejs, Ērika Bizdēna, Māris Turks, and Irina Novosjolova. 2019. "2,6-Bis[4-(4-butylphenyl)-1H-1,2,3-triazol-1-yl]-9-dodecyl-9H-purine" Molbank 2019, no. 3: M1073. https://doi.org/10.3390/M1073

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