Next Article in Journal
Synthesis, Spectroscopy and Electrochemistry in Relation to DFT Computed Energies of Ferrocene- and Ruthenocene-Containing β-Diketonato Iridium(III) Heteroleptic Complexes. Structure of [(2-Pyridylphenyl)2Ir(RcCOCHCOCH3]
Previous Article in Journal
Screening of the Hepatotoxic Components in Fructus Gardeniae and Their Effects on Rat Liver BRL-3A Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthetic, Structural, and Anticancer Activity Evaluation Studies on Novel Pyrazolylnucleosides

1
Bioorganic Laboratory, Department of Chemistry, University of Delhi, Delhi 110 007, India
2
Medicinal Chemistry Laboratory, Department of Chemistry, Acadia University, Wolfville, NS B4P 2R6, Canada
3
SUN Pharmaceuticals R&D, Gurgaon, Sarhaul, Sector-18, Haryana-122 015, India
4
Sri Venkateswara College, Benito Juarez Road, Dhaula Kuan, University of Delhi, Delhi 110 021, India
5
Department of Chemistry and Environmental Science, Medgar Evers College, The City University of New York, 1638 Bedford Avenue, Brooklyn, NY 11225, USA
6
Laboratory of Synthetic Chemistry, Leidos Biomedical Research Inc., Frederick National Laboratory for Cancer Research, Frederick, MD 2170, USA
7
Department of Radiation Oncology, Stanford University, 1050A Arastradero Road, A252, Palo Alto, CA 94304, USA
8
Chimie ParisTech, PSL University, CNRS Institute of Chemistry for Life and Health Sciences—i-CLeHS, 11 rue Pierre et Marie Curie, F-75005 Paris, France
9
Nucleic Acid Center, Department of Physics, Chemistry and Pharmacy, University of Southern Denmark, Campusvej 55, DK-5230 Odense, Denmark
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(21), 3922; https://doi.org/10.3390/molecules24213922
Submission received: 13 September 2019 / Revised: 22 October 2019 / Accepted: 28 October 2019 / Published: 30 October 2019

Abstract

:
The synthesis of novel pyrazolylnucleosides 3ae, 4ae, 5ae, and 6ae are described. The structures of the regioisomers were elucidated by using extensive NMR studies. The pyrazolylnucleosides 5ae and 6ae were screened for anticancer activities on sixty human tumor cell lines. The compound 6e showed good activity against 39 cancer cell lines. In particular, it showed significant inhibition against the lung cancer cell line Hop-92 (GI50 9.3 µM) and breast cancer cell line HS 578T (GI50 3.0 µM).

Graphical Abstract

1. Introduction

The chemistry of nucleosides has been extensively studied and several analogs have been found to exhibit potential as fungicidal, antitumor, and antiviral agents [1,2,3,4,5,6,7]. Modifications in both the heterocyclic bases and the sugar moieties have led to active and safer nucleoside analogues that have found applications as agents effective against human immunodeficiency virus (HIV), the causative agent of acquired immune deficiency syndrome (AIDS), and also against viral infections caused by the herpes simplex virus (HSV types 1 and 2), varicella zoster virus (VZV), hepatitis C virus (HCV), human cytomegalovirus (HCMV), and Epstein-Barr virus (EBV) [8,9]. Nucleoside and nucleotide modifications resulted in an increased interest in the regio- and stereoselective synthesis of nucleosides [10,11]. Moreover, modified nucleosides and nucleotides with a restricted conformation have been used to reach a particular conformation of a rotamer to study the affinity of a biomacromolecule for its natural ligand as well as the molecular recognition in an oligonucleotide chain (RNA/DNA) [12,13,14,15]. Similar studies of anti-sense and anti-gene oligonucleotides (ONs) as potential and selective inhibitors of gene expression [16,17,18,19] and their use as anti-tumor or anti-viral agents [20,21,22,23] have also influenced the developments in the field of nucleic acid-based drugs. Among the nucleoside analogues with significant biological activities, dideoxynucleoside-based compounds such as 2′,3-dideoxycytidine (ddC) [24], 2,3-dideoxyinosine (ddI) [24], and 3-azidothymidine (AZT) [25] are effective therapeutic agents for the treatment of AIDS, while ribavirin (virazole) [26,27] is an antiviral drug (Figure 1). Similarly, other dideoxynucleosides such as d4T (2,3-didehydro-3-deoxythymidine, stavudine) [28,29] and AZddU (3-azido-2,3-dideoxyuridine) [30] have gone through clinical studies. Different nucleosides isolated from nature such as oxazinomycin [31], pyrazofurin [32,33], showdomycin [34,35], formycin A, and formycin B [36,37] have shown antibiotic properties and have also been found to exhibit anticancer and/or antiviral activities (Figure 1). These examples and new developments in the chemistry and biology of these compounds and their analogs [38,39,40,41,42] have motivated us to work in this area and have led us to investigate the interesting chemical and biological properties of novel nucleoside-based compounds.
Despite the developments in nucleoside chemistry, the clinical use of nucleosides has some drawbacks due to their side-effects and primary or acquired drug resistance [43]. Therefore, the search for the design of new and effective nucleoside-based analogues continues to motivate studies in this field. Our efforts toward this goal have led to the synthesis of twenty novel pyrazolylnucleosides with new structures and potent antiviral and antitumoral behaviors. In this regard, two regioisomers for each pyrazole derivative have been produced and characterized using spectroscopic techniques such as 1H NMR, 13C NMR, NOESY, HMBC, IR, and mass spectroscopy. All compounds were evaluated against the National Cancer Institute (NCI)’s panel of 60 human tumor cell lines for their anticancer activities, and for their antiviral activities against representative viruses.

2. Results and Discussion

2.1. Chemical Synthesis of the Nucleoside Analogues 5 and 6

In our approach, the synthesis of the target compounds 5ae and 6ae was achieved in two steps starting from the already reported 3-cyanomethyl-5-aryl-1H-pyrazoles 1ae [44] and the well-known 2-deoxy-3,5-di-O-p-toluoyl-α-d-ribofuranosyl chloride (2) [45] (Scheme 1). Treatment of pyrazoles 1ae with sodium hydride in acetonitrile and the subsequent addition of chlorosugar 2 gave two regioisomers of the modified pyrazolyl nucleosides 3ae (by coupling 2 with the N-1 nitrogen of the pyrazole derivatives) and 4ae (by coupling chlorosugar 2 with the N-2 nitrogen of the pyrazole derivatives) in 58–65% yields. The products 3ae and 4ae were de-toluoylated by using sodium methoxide in methanol, resulting in the formation of the corresponding modified nucleosides 5ae and 6ae in 70–80% yields, respectively.
Pyrazoles with no substitution on either of the two N ring atoms can be alkylated to produce two regioisomers under strongly basic conditions. The anion generated produces the two resonance forms that react with chlorosugar 2 in the glycosidation step, leading to mixtures of two regioisomeric pyrazolyl nucleosides (i.e., 3ae and 4ae) (Scheme 2).

2.2. Structural Identification of the Isomeric Pyrazolyl Nucleoside Analogues 36

The structural identification of the isomeric disubstituted pyrazolyl nucleosides, based on their 1H NMR spectral data (See Supplementary Materials), has been reported in the literature [46,47,48]. It has been observed that the anomeric protons in the isomeric pyrazolyl nucleosides exhibit different proton chemical shifts. The anomeric proton adjacent to N-1 of the pyrazole compounds 3ae and 5ae would appear downfield when compared to the anomeric proton adjacent to N-2 of the pyrazole compounds 4ae and 6ae [46,47,48]. In addition to using 1H NMR chemical shifts, extensive NOE and 2D NMR experiments have been employed to confirm the structures of the isomeric pyrazolyl nucleosides. Using 1H and 13C NMR studies, we confirmed the positional assignments of the isomeric pyrazolyl nucleosides synthesized in the present work (Table 1, Table 2 and Table 3). The anomeric proton of the analogues 5ae generally appeared at 0.14–0.19 ppm upfield when compared to the corresponding proton in the 1H NMR spectra of its corresponding 6 series isomers. This is in agreement with observations for other 1,5- and 1,3-disubstituted pyrazole nucleosides [46,47,48]. We found that the effect was less pronounced in the 3 series of nucleosides as compared to the 4 series nucleosides where this difference was in the range of 0.04–0.14 ppm. The 13C chemical shift of the –CH2CN bearing pyrazole carbon atom can help distinguish between positional isomers. In the 3 and 5 series of nucleosides, this carbon signal was 4–11 ppm downfield relative to the corresponding signal in the 4 and 6 series nucleosides. In most cases, the aryl bearing pyrazole carbon in the 3 and 5 series was 4–7 ppm upfield relative to the corresponding carbon in the 4 and 6 series. However, in the case of 3d versus 4d, the difference was small and reversed, making the aryl bearing pyrazole carbon less reliable for use in positional assignments.
In order to confirm the above positional assignments, we performed 2D NMR experiments (NOESY and HMBC) on the two pairs of compounds (i.e., one pair consisting of the ditoluoyl protected nucleosides 3d and 4d and another pair consisting of the deprotected nucleosides 5d and 6d) (Table 4). The results of the HMBC experiments rely on the fact that the anomeric proton may see the aryl bearing pyrazole carbon in 3d and 5d, but not in 4d and 6d, while the NOESY results would explain the spatial proximity of the anomeric proton to either the protons in the –CH2CN group or to the ortho protons of the aryl group. Our results verified the positional properties at the pyrazole ring in the four compounds and are summarized in Table 4.
The isomers of the 3 series nucleosides generally had higher Rf values on the TLC than the corresponding isomer of the 4 series of nucleosides. The same was observed for the isomers of the 5 series relative to those in the 6 series of nucleosides.

2.3. Anticancer Activity of the Isomeric Pyrazolyl Nucleoside Analogues 5ae and 6ae

The anticancer and toxicity of compounds 5ae and 6ae were evaluated by using the National Cancer Institute′s 60 human cancer cell lines. Compounds 5ae, where the sugar is attached to N-1 of the pyrazole ring, were found to be inactive against all the cell lines irrespective of the substituent on the aromatic ring. In the other series, compounds 6a, 6b, and 6c with 4-methyl, 4-methoxy, and 4-fluoro substituents, respectively, at the aromatic ring were also inactive. However, compounds 6d and 6e with 4-chloro and 4-bromo substituents at the aromatic ring, respectively, showed inhibition against multiple anticancer cell lines. The GI50 (cytostatic parameter) and LC50 (toxicity parameter) values of 6d and 6e for the selected cell lines are given in Table 5. Compound 6d showed inhibition in 19 cell lines, and was most active against the renal cancer cell line UO-31 and breast cancer cell line HS 578T with GI50 < 20 µM in both cases. The most active compound was 6e, which showed moderate inhibition in 39 cell lines. It showed significant inhibition against lung cancer cell line Hop-92 with a GI50 of 9.3 µM and breast cancer cell line HS 578T with a GI50 of 3.0 µM. Most importantly, both compounds (6d and 6e) did not show any toxicity, even at the highest concentration tested, as indicated by their high LC50 values.
The pyrazolyl nucleosides 5ae and 6ae were also examined for their antiviral activities against a number of viruses such as Simplex virus type 1 (HSV-1) and type 2 (HSV-2), thymidine kinase-deficient (TK-) strains of HSV-1, Vaccinia virus, para-influenza-3 virus, Sindbis virus, Coxsackie virus, Punta toro virus, vesicular stomatitis virus (VSV), Coxsackie virus B4 (CV-B4), respiratory syncytial virus (RSV), feline corona virus, and feline herpes virus. However, none of these compounds showed any significant activity against any of these viruses.

3. Experimental

3.1. General Information

Reactions were conducted under an atmosphere of nitrogen in anhydrous solvents. Column chromatography was carried out by using silica gel (100–200 mesh). Melting points were determined in a concentrated H2SO4 acid bath and were uncorrected. Analytical TLCs were performed on pre-coated Merck silica gel 60F254 plates; the spots were detected either by using UV light or by charring with 4% alcoholic sulfuric acid. The IR spectra were recorded on a Perkin-Elmer 2000 FT-IR spectrophotometer (Waltham, MA, USA). The optical rotations were measured with a Bellingham-Stanley AD 220 polarimeter and the concentrations expressed as g/mL. The 1H-NMR and 13C-NMR spectra were recorded on a Bruker Avance 300 spectrometer (Billerica, MA, USA) at 300 and 75.5 MHz, respectively in CDCl3, DMSO-d6, or CD3CN. All 2D measurements were performed in acetone-d6 on a Bruker Avance 400 spectrometer (Billerica, MA, USA). Chemical shifts are relative to internal TMS. Assignments were based on COSY, NOESY, HMBC (by using Bruker’s microprogram inv4gslplrnd, which includes the low-pass J-filter to suppress one-bond correlations), HSQC, and JRES spectra. The chemical shift values were reported as δ ppm relative to TMS used as the internal standard and the coupling constants (J) were measured in Hz. The ESI-HRMS spectra of all compounds were recorded on a JEOL JMS-AX505W high-resolution mass spectrometer (Tokyo, Japan) in positive ion mode by using the matrix HEDS (bis-hydroxyethylsulfide) doped with sodium acetate. Acetonitrile was used after distillation over freshly ignited potassium carbonate.

3.2. Synthesis and Characterization

3.2.1. Synthesis of 2′-Deoxy-1′-(3-cyanomethyl-5-aryl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (3ae) and 2′-deoxy-1′-(3-aryl-5-cyanomethyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (4ae)

A solution of 3-cyanomethyl-5-arylpyrazole (1ae, 15 mmol) in acetonitrile (190 mL) was added into a stirred mixture of sodium hydride (22.5 mmol) in acetonitrile (30 mL) under a nitrogen atmosphere at 30–35 °C, and continuously stirred for 30 min. The reaction mixture was cooled to 0 °C and 1-α-chloro-3,5-di-O-toluoyl-1,2-dideoxyribose (2, 15 mmol) was added and the contents stirred at 0 °C for 2.5 to 3 h. The progress of the reaction was monitored by using silica gel TLC. On completion, the reaction mixture was poured over ice-cold water and extracted with ethyl acetate (3 × 50 mL). The combined organic layer was washed with water (2 × 50 mL), dried over anhydrous sodium sulfate, the solvent was removed under reduced pressure, and the residue thus obtained was column chromatographed over silica gel with ethyl acetate (8–10%) in petroleum ether as the eluent to afford the pyrazolyl nucleosides 3ae and 4ae in 31 to 35% and 24 to 29% yields, respectively.
2′-Deoxy-1′-(3-cyanomethyl-5-p-methylphenyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (3a): Obtained as a white semisolid in a 34% yield. [ α ] D 18 = −115.46° (c 0.25, CHCl3); Rf = 0.53 (20% ethyl acetate in petroleum ether); IR (Nujol): 2923, 2256 (CN), 1721 (COO), 1611, 1452, 1269, 1177 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.41 (9H, s, 3 × Ar-CH3), 2.45–2.51(1H, m, C-2Hα), 3.45–3.52 (1H, m, C-2Hβ), 3.65 (2H, s, CH2CN), 4.55–4.67 (3H, m, C-4H and C-5 Hα+β), 5.86 (1H, brs, C-3H), 6.13 (1H, t, J = 6.1 Hz, C-1H), 6.33 (1H, s, C-4H), 7.21–7.24 (4H, brs, 4 × Ar-H), 7.28 (2H, d, J = 7.8 Hz, Ar-H), 7.42 (2H, d, J = 7.8 Hz, Ar-H), 7.90 (2H, d, J = 8.0 Hz, Ar-H), 8.00 (2H, d, J = 8.0, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 15.6, 19.3, 19.7, 19.7, 34.6, 62.4, 73.9, 80.6, 84.5, 103.8, 115.2, 124.4, 124.7, 125.2, 127.0, 127.1, 127.1, 127.5, 127.7, 127.9, 137.2, 139.9, 142.1, 144.6, 141.7, 164.0, 164.3; HRMS (ESI): Calculated for C33H31N3O5 [M + Na]+ 572.2156, found [M + Na]+ 572.2163.
2′-Deoxy-1′-(3-cyanomethyl-5-p-methoxyphenyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (3b): Obtained as a semisolid in a 31% yield. [ α ] D 18 = −119.46° (c 0.25, CHCl3); Rf = 0.31 (20% ethyl acetate in petroleum ether); IR (Nujol): 2926, 2256 (CN), 1724 (CO), 1610, 1505, 1254, 1178 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.40 (6H, s, 2 × Ar-CH3), 2.46–2.50 (1H, m, C-2Hα), 3.48–3.53 (1H, m, C-2Hβ), 3.64 (2H, s, CH2CN), 3.84 (3H, s, OCH3), 4.54–4.62 (3H, m, C- C-4H & 5Hα+β), 5.87 (1H, brs, C-3H), 6.10 (1H, brs, C-1H), 6.30 (1H, s, C-4H), 6.99 (2H, d, J = 8.1, Ar-H), 7.21–7.23 (4H, m, Ar-H), 7.46 (2H, d, J = 8.0 Hz, Ar-H), 7.89 (2H, d, J = 7.6 Hz, Ar-H), 7.99 (2H, d, J = 7.6, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 19.7, 23.7, 23.8, 38.7, 57.5, 66.5, 78.0, 84.7, 88.6, 107.8, 116.4, 119.3, 123.8, 128.9, 129.4, 131.1, 131.2, 131.8, 132.0, 132.6, 144.0, 145.8, 146.2, 148.5, 162.4, 168.1, 168.4; HRMS (ESI): Calculated for C33H31N3O6 [M + Na]+ 588.2105, found [M + Na]+ 588.2127.
2′-Deoxy-1′-(3-cyanomethyl-5-p-flurophenyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (3c): Obtained as a white semisolid in a 35% yield. [ α ] D 18 = −63.64° (c 0.25, CHCl3); Rf = 0.54 (20% ethyl acetate in petroleum ether); IR (Nujol): 2925, 2256 (CN), 1723 (COO), 1610, 1506, 1453, 1177, 1104 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.40 (6H, s, 2 × Ar-CH3), 2.48–2.52 (1H, m, C-2Hα), 3.50-3.54 (1H, m, C-2Hβ), 3.64 (2H, s, CH2CN), 4.56–4.66 (3H, m, C-4H & C-5H), 5.87 (1H, brs, C-3H), 6.05 (1H, t, J = 5.8 Hz, C-1H), 6.33 (1H, s, C-4H), 7.13–7.24 (6H, m, Ar-H), 7.50–7.55 (2H, m, Ar-H), 7.90 (2H, d, J = 7.9 Hz, Ar-H), 7.98 (2H, d, J = 7.9 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 17.5, 21.5, 21.6, 36.4, 64.2, 75.7, 82.6, 86.3, 106.1, 115.9 (d, J = 22 Hz), 116.9, 125.4, 126.6, 127.1, 128.9, 129.0, 129.6, 129.8, 131.0 (d, J = 8 Hz), 141.9, 143.7, 144.1, 145.4, 163.1 (d, J = 247 Hz), 165.9, 166.1; HRMS (ESI): Calculated for C32H28N3O5F [M + Na]+ 576.1905, found [M + Na]+ 576.1916.
2′-Deoxy-1′-(3-cyanomethyl-5-p-chlorophenyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (3d): Obtained as a yellow solid in a 32% yield. [ α ] D 18 = −109.79° (c 0.25, CHCl3); Mp = 114–115 °C; Rf= 0.52 (20% ethyl acetate in petroleum ether); IR (Nujol): 2924, 2257 (CN), 1728 (COO), 1610, 1494, 1177, 1108 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.40 (6H, s, 2 × Ar-CH3), 2.46–2.52 (1H, m, C-2Hα), 3.50–3.57 (1H, m, C-2Hβ), 3.63 (2H, s, CH2CN), 4.51–4.66 (3H, m, C-4H & C-5H), 5.86–5.88 (1H, m, C-3H), 6.05 (1H, t, J = 6.1 Hz, C-1H), 6.35 (1H, s, C-4H), 7.21–7.24 (4H, m, Ar-H), 7.43–7.50 (4H, m, Ar-H), 7.90 (2H, d, J = 7.9 Hz, Ar-H), 7.98 (2H, d, J = 7.9 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 17.4, 21.5, 21.6, 36.4, 64.1, 75.6, 82.7, 86.4, 106.1, 116.9, 126.4, 126.6, 127.6, 128.9, 129.0, 129.1, 129.6, 129.7, 130.3, 135.3, 141.9, 143.7, 144.1, 145.2, 165.9, 166.1; HRMS (ESI): Calculated for C32H28N3O5Cl [M + Na]+ 592.1610, found [M + Na]+ 592.1625.
2′-Deoxy-1′-(3-cyanomethyl-5-p-bromophenyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (3e): Obtained as a white solid in a 35% yield. [ α ] D 18 = −96.35° (c 0.25, CHCl3); Mp = 118–110 °C; Rf = 0.47 (20% ethyl acetate in petroleum ether); IR (Nujol): 2921, 2255 (CN), 1714 (COO), 1610, 1463, 1267, 1176, 1101 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.41 (6H, s, 2 × Ar-CH3), 2.46–2.54 (1H, m, C-2Hα), 3.48–3.57 (1H, m, C-2Hβ), 3.64 (2H, s, CH2CN), 4.51–4.66 (3H, m, C-4H & C-5H), 5.85–5.87 (1H, m, C-3H), 6.05 (1H, t, J = 6.1 Hz, C-1H), 6.35 (1H, s, C-4H), 7.22 (4H, d, J = 7.9 Hz, Ar-H), 7.42 (2H, d, J = 8.3 Hz, Ar-H), 7.60 (2H, d, J = 8.3 Hz, Ar-H), 7.90 (2H, d, J = 8.0 Hz, Ar-H), 7.98 (2H, d, J = 8.0 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 17.8, 21.9, 21.9, 36.8, 64.5, 76.0, 83.1, 86.8, 106.5, 117.3, 123.9, 127.0, 127.4, 128.5, 129.3, 129.4, 130.0, 130.1, 130.9, 132.4, 142.3, 144.0, 144.4, 145.6, 166.2, 166.5; HRMS (ESI): Calculated for C32H28N3O5Br [M + Na]+ 636.1105, found [M + Na]+ 636.1113.
2′-Deoxy-1′-(3-p-methylphenyl-5-cyanomethyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (4a): Obtained as a white semisolid in a 26% yield. [ α ] D 18 = +25.10° (c 0.25, CHCl3); Rf = 0.36 (20% ethyl acetate in petroleum ether); IR (Nujol): 2923, 2256 (CN), 1721 (COO), 1611, 1452, 1269, 1177 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.31, 2.38 and 2.42 (9H, 3s, 3H each, 3 × Ar-CH3), 2.65-2.70 (1H, m, C-2Hα), 3.74–3.78 (1H, m, C-2Hβ), 3.91 (2H, s, CH2CN), 4.36 (1H, q, J = 6.3 Hz, C-4H), 4.54–4.60 (2H, m, C-5H), 5.85-5.87 (1H, m, C-3H), 6.17 (1H, t, J = 5.7 Hz, C-1H), 6.59 (1H, s, C-4H), 7.03 (2H, d, J = 8.0 Hz, Ar-H), 7.18 (2H, d, J = 7.9 Hz, Ar-H), 7.26 (2H, d, J = 8.1 Hz, Ar-H), 7.63 (2H, d, J = 8.0 Hz, Ar-H), 7.83 (2H, d, J = 8.1 Hz, Ar-H), 7.95 (2H, d, J = 8.0 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 15.28, 21.2, 21.5, 21.6, 36.0, 63.6, 75.0, 83.2, 87.0, 105.3, 115.3, 125.6, 126.6, 126.7, 128.9, 129.0, 129.1, 129.2, 129.6, 129.7, 133.0, 137.9, 143.5, 144.2, 151.0, 165.9, 166.0; HRMS (ESI): Calculated for C33H31N3O5 [M + Na]+ 572.2156, found [M + Na]+ 572.2155.
2′-Deoxy-1′-(3-p-methoxyphenyl-5-cyanomethyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (4b): Obtained as a white solid in a 24% yield. [ α ] D 18 = +42.05° (c 0.25, CHCl3); Mp = 88–90 °C; Rf= 0.24 (20% ethyl acetate in petroleum ether); IR (Nujol): 2924, 2256 (CN), 1724 (COO), 1610, 1456, 1270, 1178 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.32 and 2.42 (6H, 2s, 3H each, 2 × Ar-CH3), 2.65–2.69 (1H, m, C-2Hα), 3.73–3.77 (1H, m, C-2Hβ), 3.83 (3H, s, Ar-OCH3), 3.90 (2H, s, CH2CN), 4.34–4.37 (1H, m, C-4H), 4.56–4.59 (2H, m, C-5H), 5.87 (1H, brs, C-3H), 6.16 (1H, brs, C-1H), 6.55 (1H, s, C-4H), 6.91 (2H, d, J = 7.6 Hz, Ar-H), 7.05 (2H, d, J = 7.3 Hz, Ar-H), 7.26 (2H, d, J = 6.9 Hz, Ar-H), 7.67 (2H, d, J = 7.7 Hz, Ar-H), 7.84 (2H, d, J = 7.3 Hz, Ar-H), 7.96 (2H, d, J = 7.4 Hz, Ar-H); 13C (75.5 MHz, CDCl3): δ 15.2, 21.5, 21.6, 36.0, 55.2, 63.6, 75.0, 83.1, 86.9, 105.0, 113.9, 115.4, 125.2, 126.6, 126.9, 128.9, 129.1, 129.5, 129.6, 129.7, 133.0, 143.5, 144.1, 150.8, 159.6, 165.8, 165.9; HRMS (ESI): Calculated for C33H31N3O6 [M + Na]+ 588.2105, found [M + Na]+ 588.2134.
2′-Deoxy-1′-(3-p-flurophenyl-5-cyanomethyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (4c): Obtained as a white semisolid in a 25% yield. [ α ] D 18 = +14.07° (c 0.25, CHCl3); Rf = 0.47 (20% ethyl acetate in petroleum ether); IR (Nujol): 2925, 2257 (CN), 1725 (COO), 1610,1522,1443,1272, 1178, 1102 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.31, and 2.42 (6H, 2s, 3H each, 2 × Ar-CH3), 2.66–2.70 (1H, m, C-2Hα), 3.73–3.77 (1H, m, C-2Hβ), 3.92 (2H, s, CH2CN), 4.33–4.35 (1H, m, C-4H), 4.58–4.63 (2H, m, C-5H), 5.87 (1H, brs, C-3H), 6.17 (1H, t, J = 5.1 Hz, C-1H), 6.56 (1H, s, C-4H), 7.02–7.08 (4H, m, Ar-H), 7.26 (2H, d, J = 7.9 Hz, Ar), 7.67–7.71 (2H, m, Ar-H), 7.83 (2H, d, J = 8.0 Hz, Ar-H), 7.96 (2H, d, J = 8.0 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 17.7, 23.9, 24.0, 38.5, 66.0, 77.3, 85.7, 89.5, 107.7, 117.8 (d, J = 22 Hz), 128.0, 129.0 (d, J = 7.5 Hz), 129.7, 129.8, 131.0, 131.4, 131.6, 132.0, 132.1, 135.8, 146.0, 146.7, 152.5, 165.4 (d, J = 250 Hz), 168.2, 168.4; HRMS (ESI): Calculated for C32H28N3O5F [M + Na]+ 576.1905, found [M + Na]+ 576.1915.
2′-Deoxy-1′-(3-p-chlorophenyl-5-cyanomethyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (4d): Obtained as a yellow solid in a 24% yield. [ α ] D 18 = +34.06° (c 0.25, CHCl3); Mp = 109–110 °C; Rf = 0.40 (20% ethyl acetate in petroleum ether); IR (nujol): 2924, 2257 (CN), 1716 (COO), 1459, 1377, 1176, cm−1; 1H NMR (300 MHz, CDCl3): δ 2.33 and 2.45 (6H, 2s, 3H each, 2 × Ar-CH3), 2.65–2.74 (1H, m, C-2Hα), 3.74–3.85 (1H, m, C-2Hβ), 3.94 (2H, s, CH2CN), 4.33–4.38 (1H, m, C-4H), 4.57–4.65 (2H, m, C-5H), 5.88–5.90 (1H, m, C-3H), 6.19 (1H, t, J = 5.8 Hz, C-1H), 6.59 (1H, s, C-4H), 7.04 (2H, d, J = 7.9 Hz, Ar-H), 7.29 (2H, d, J = 8.0 Hz, Ar-H), 7.35 (2H, d, J = 8.4 Hz, Ar-H), 7.66 (2H, d, J = 8.4 Hz, Ar-H), 7.82 (2H, d, J = 8.1 Hz, Ar-H), 7.98 (2H, d, J = 8.0 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 15.7, 21.9, 22.0, 36.5, 63.9, 75.3, 83.8, 87.6, 105.8, 116.5, 127.0, 127.3, 129.1, 129.4, 129.6, 130.0, 130.1, 130.2, 131.3, 133.9, 134.3, 144.1, 144.7, 150.3, 166.2, 166.4; HRMS (ESI): Calculated for C32H28N3O5Cl [M + Na]+ 592.1610, found [M + Na]+ 592.1621.
2′-Deoxy-1′-(3-p-bromophenyl-5-cyanomethyl-)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (4e): Obtained as a white solid in a 29% yield. [ α ] D 18 = +39.82° (c 0.25, CHCl3); Mp = 118–120 °C; Rf = 0.25 (20% ethyl acetate in petroleum ether); IR (Nujol): 2923, 2255 (CN), 1716 (COO), 1611, 1464, 1270, 1177 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.31 and 2.42 (6H, 2s, 3H each, 2 × Ar-CH3), 2.62–2.66 (1H, m, C-2Hα), 3.79–3.82 (1H, m, C-2Hβ), 3.92 (2H, s, CH2CN), 4.32–4.36 (1H, m, C-4H), 4.56–4.62 (2H, m, C-5H), 5.86–5.88 (1H, m, C-3H), 6.17 (1H, t, J = 5.7 Hz, C-1H), 6.57 (1H, s, C-4H), 7.03 (2H, d, J = 7.9 Hz, Ar-H), 7.26 (2H, d, J = 7.9 Hz, Ar-H), 7.48 (2H, d, J = 8.3 Hz, Ar-H), 7.57 (2H, d, J = 8.4 Hz, Ar-H), 7.80 (2H, d, J = 8.0 Hz, Ar-H), 7.96 (2H, d, J = 8.0 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 15.2, 21.5, 21.6, 36.0, 63.4, 74.8, 83.3, 87.1, 105.3, 115.3, 122.0, 126.5, 126.6, 127.2, 128.9, 129.1, 129.5, 129.7, 131.3, 131.6, 133.5, 143.6, 144.2, 149.8, 166.0, 166.1; HRMS (ESI): Calculated for C32H28N3O5Br [M + Na]+ 636.1105, found [M + Na]+ 636.1116.

3.2.2. Synthesis of 2′-Deoxy-1′-(3-cyanomethyl-5-aryl)pyrazolyl-β-D-ribofuranose (5ae) and 2′-deoxy-1′-(3-aryl-5-cyanomethyl)pyrazolyl-β-d-ribofuranose (6ae)

2′-Deoxy-1′-(3-cyanomethyl-5-aryl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (3ae) or 2′-deoxy-1′-(3-aryl-5-cyanomethyl)pyrazolyl-3′,5′-di-O-toluoyl-β-d-ribofuranose (4ae, 1.3 mmol) was suspended in dry methanol (18 mL), then a mixture of NaOMe and MeOH (10 mL, 1:3, v/v) was added to the resulting solution, and the contents stirred at room temperature for 3–4 h. The progress of the reaction was monitored by using silica gel TLC. On completion, NH4Cl was added to the reaction mixture to adjust the pH to 8. One-third of the solvent was removed under reduced pressure and the reaction mixture was poured in water and extracted with ethyl acetate (3 × 30 mL), the organic layer was dried over anhydrous sodium sulfate, the solvent was removed under reduced pressure, and the residue thus obtained was subjected to column chromatography over silica gel with methanol (2–2.5%) in chloroform as the eluent to afford 2′-deoxy-1′-(3-cyanomethyl-5-aryl)pyrazolyl-β-d-ribofuranose (5ae) and 2′-deoxy-1′-(3-aryl-5-cyanomethyl)pyrazolyl-β-d-ribofuranose (6ae).
2′-Deoxy-1′-(3-cyanomethyl-5-p-methylphenyl)pyrazolyl-β-d-ribofuranose (5a): Obtained as a white semisolid in a 63% yield. [ α ] D 30 = −87.56° (c 0.25, MeOH); Rf = 0.3 (7% methanol in chloroform); IR (Nujol): 3391 (OH), 2936, 2258 (CN), 1613, 1506, 1456, 1253, 1180, 1055 cm−1; 1H NMR (300 MHz, CDCl3): δ 2.10–2.18 (1H, m, C-2Hα), 2.39 (3H, s, Ar-CH3), 2.87–2.91 (1H, m, C-2Hβ), 3.53 (1H, brs, C-5Hα), 3.61 (1H, brs, C-5Hβ), 3.84 (1H, brs, C-4H), 3.90 (2H, s, CH2CN), 4.44 (1H, brs, C-3H), 4.49 (1H, brs, C-3OH), 5.13 (1H, brs, C-5OH), 5.99 (1H, brs, C-1H), 6.30 (1H, s, C-4H), 7.29 (2H, d, J = 7.3, Ar-H), 7.40 (2H, d, J = 7.7 Hz, Ar-H); 13C NMR (75.5 MHz, CDCl3): δ 17.3, 21.1, 39.9, 63.0, 71.6, 86.0, 88.3, 105.3, 117.7, 126.5, 128.9, 129.5, 138.8, 142.0, 145.9; HRMS (ESI): Calculated for C17H19N3O3 [M + Na]+ 336.1319, found [M + Na]+ 336.1288.
2′-Deoxy-1′-(3-cyanomethyl-5-p-methoxyphenyl)pyrazolyl-β-d-ribofuranose (5b): Obtained as a white semisolid in a 66% yield. [ α ] D 30 = −94.35° (c 0.25, MeOH); Rf = 0.48 (7% methanol in chloroform); IR (Nujol): 3391 (OH), 2932, 2254 (CN), 1612, 1510, 1457, 1253, 1183, 1056 cm1; 1H NMR (300 MHz, DMSO-d6): δ 2.17–2.21 (1H, m, C-2′Hα), 2.89–2.93 (1H, m, C-2′Hβ), 3.57–3.61 (1H, m, C-5′Hα), 3.65–3.70 (1H, m, C-5′Hβ), 3.83-3.85 (5H, m, Ar-OCH3 & CH2CN), 3.92-3.94 (1H, m, C-4′H), 4.53 (1H, brs, C-3′H), 4.64 (1H, t, J = 6.6, C-5′OH), 5.09 (1H, d, J = 4.1 Hz, C-3′OH), 6.02 (1H, t, J = 6.3 Hz, C-1′H), 6.26 (1H, s, C-4H), 7.00 (2H, d, J = 8.6 Hz, Ar-H), 7.42 (2H, d, J = 8.6 Hz, Ar-H); 13C NMR (75.5 MHz, DMSO-d6): δ 17.3, 39.9, 55.3, 63.1, 71.8, 86.0, 88.4, 105.1, 114.3, 117.4, 121.5, 130.3, 141.8, 145.8, 160.1; HRMS (ESI): Calculated C17H19N3O4 [M + Na]+ 352.1268, found [M + Na]+ 352.1248.
2′-Deoxy-1′-(3-cyanomethyl-5-p-fluorophenyl)pyrazolyl-β-d-ribofuranose (5c): Obtained as a semisolid in a 63% yield. [ α ] D 30 = −153.52° (c 0.25, MeOH); Rf = 0.41 (7% methanol in chloroform); IR (Nujol): 3393 (OH), 2932, 2256 (CN), 1611, 1508, 1451, 1255, 1182, 1051 cm−1; 1H NMR (300 MHz, DMSO-d6): δ 2.18–2.22 (1H, m, C-2Hα), 2.91–2.95 (1H, m, C-2Hβ), 3.61 (1H, brs, C-5Hα), 3.65 (1H, brs, C-5Hβ), 3.87 (2H, s, CH2CN), 3.93 (1H, brs, C-4H), 4.53 (2H, brs, C-3H and OH), 5.08 (1H, brs, C-5OH), 5.97 (1H, brs, C-1H), 6.32 (1H, s, C-4H), 7.22 (2H, brs, Ar-H), 7.53 (2H, brs, Ar-H); 13C NMR (75.5 MHz, DMSO-d6): δ 17.3, 39.9, 63.0, 71.7, 86.1, 88.4, 105.7, 115.9 (d, J = 22 Hz), 117.3, 125.6, 131.1 (d, J = 8 Hz), 141.9, 144.9, 162.8 (d, J = 246 Hz); HRMS (ESI): Calculated for C16H16N3O3F [M + Na]+ 340.1068, found [M + Na]+ 340.1058.
2′-Deoxy-1′-(3-cyanomethyl-5-p-chlorophenyl)pyrazolyl-β-d-ribofuranose (5d): Obtained as a semisolid in a 63% yield. [ α ] D 30 = −124.74° (c 0.25, MeOH); Rf = 0.46 (7% methanol in chloroform); IR (Nujol): 3393 (OH), 2932, 2256 (CN), 1611, 1498, 1451, 1254, 1181, 1048 cm−1; 1H NMR (300 MHz, DMSO-d6): δ 2.22–2.24 (1H, m, C-2Hα), 2.92–2.96 (1H, m, C-2Hβ), 3.59–3.61 (1H, m, C-5Hα), 3.68–3.72 (1H, m, C-5Hβ), 3.85 (2H, brs, CH2CN), 3.97 (1H, brs, C-4H), 4.57 (2H, brs, C-3H and OH), 5.09 (1H, brs, C-5OH), 6.00 (1H, brs, C-1H), 6.34 (1H, s, C-4H), 7.24 (2H, d, J = 7.5 Hz, Ar-H), 7.87 (2H, d, J = 8.5 Hz, Ar-H); 13C (75.5 MHz, DMSO-d6): δ 17.3, 39.9, 63.1, 71.8, 86.2, 88.5, 105.7, 117.8, 127.8, 129.3, 130.4, 134.7, 141.9, 144.7; HRMS (ESI): Calculated for C16H16N3O3Cl [M + Na]+ 356.0772, found [M + Na]+ 356.0770.
2′-Deoxy-1′-(3-cyanomethyl-5-p-bromophenyl)pyrazolyl-β-d-ribofuranose (5e): Obtained as a semisolid in a 64% yield. [ α ] D 30 = −105.15° (c 0.25, MeOH); Rf = 0.53 (7% methanol in chloroform); IR (Nujol): 3387 (OH), 2935, 2249 (CN), 1612, 1504, 1454, 1248, 1179, 1048 cm−1; 1H NMR (300 MHz, DMSO-d6): δ 2.20–2.24 (1H, m, C-2Hα), 2.90–2.96 (1H, m, C-2Hβ), 3.59–3.61 (1H, m, C-5Hα), 3.68–3.72 (1H, m, C-5Hβ), 3.85 (2H, brs, CH2CN), 3.97 (1H, brs, C-4H), 4.56 (2H, brs, C-3H & OH ), 5.07 (1H, brs, C-5OH), 6.00 (1H, t, J = 6.06 Hz, C-1H), 6.34 (1H, s, C-4H), 7.42 (2H, d, J = 8.1 Hz, Ar-H), 7.61 (2H, d, J = 8.0 Hz, Ar-H); 13C (75.5 MHz, DMSO-d6): δ 16.5, 39.0, 62.2, 71.0, 85.3, 87.6, 104.8, 116.3, 122.2, 128.2, 129.8, 131.0, 141.1, 143.9; HRMS (ESI): Calculated for C16H16N3O3Br [M + Na]+ 400.0267, found [M + Na]+ 400.0232.
2′-Deoxy-1′-(5-cyanomethyl-3-p-methylphenyl)pyrazolyl-β-d-ribofuranose (6a): Obtained as a white semisolid in a 61% yield. [ α ] D 30 = +34.78° (c 0.25, MeOH); Rf = 0.3 (7% methanol in chloroform); IR (Nujol): 3392 (OH), 2934, 2258 (CN), 1611, 1504, 1452, 1254, 1179, 1051 cm1; 1H NMR (300 MHz, CDCl3): δ 2.23–2.27 (1H, m, C-2′Hα), 2.31 (3H, s, Ph-CH3), 2.77–2.83 (1H, m, C-2′Hβ), 3.36–3.40 (1H, m, C-5′Hα), 3.49–3.51 (1H, m, C5′Hβ), 3.82–3.86 (1H, m, C-4′H), 4.32 (2H, s, CH2CN), 4.46 (1H, brs, C-3′H), 4.86 (1H, t, J = 5.4 Hz, C-5′OH), 5.29 (1H, d, J = 4.0 Hz, C-3′OH), 6.13 (1H, t, J = 5.9 Hz, C-1′H), 6.72 (1H, s, C-4H), 7.22 and 7.69 (4H, 2d, 2H each, J = 7.8 Hz, & 7.9 Hz, ArH); 13C NMR (75.5 MHz, DMSO-d6): δ 14.6, 21.1, 39.8, 62.6, 71.4, 86.5, 88.4, 104.0, 117.5, 125.5, 129.6, 130.1, 135.0, 137.7, 150.2; HRMS (ESI): Calculated for C17H19N3O3 [M + Na]+ 336.1319, found [M + Na]+ 336.1290.
2′-Deoxy-1′-(5-cyanomethyl-3-p-methoxyphenyl)pyrazolyl-β-d-ribofuranose (6b): Obtained as a semisolid in a 62% yield. [ α ] D 30 = +56.37o (c 0.25, MeOH); Rf = 0.39 (7% methanol in chloroform); IR (Nujol): 3385 (OH), 2937, 2258 (CN), 1610, 1501, 1448, 1251, 1183, 1053 cm1; 1H NMR (300 MHz, DMSO-d6): δ 2.40–2.42 (1H, m, C-2′Hα), 2.94–2.98 (1H, m, C-2′Hβ), 3.60–3.64 (1H, m, C-5′Hα), 3.72–3.76 (1H, m, C-5′Hβ), 3.81 (3H, s, Ph-OCH3), 3.87 (1H, brs, C-4′H), 4.08 (2H, s, CH2CN), 4.61 (1H, brs, C-3′H), 5.01–5.05 (2H, m, C-5′OH & C-3′OH), 6.19 (1H, t, J = 6.0 Hz, C-1′H), 6.56 (1H, s, C-4H), 6.91 and 7.66 (4H, 2d, 2H each, J = 8.4 Hz and 8.4 Hz, Ar-H); 13C NMR (75.5 MHz, DMSO-d6): δ 14.9, 39.9, 55.1, 63.2, 71.9, 86.9, 88.9, 103.6, 114.0, 116.1, 126.7, 129.3, 133.7, 150.7, 159.6; HRMS (ESI): Calculated for C17H19N3O4 [M + Na]+ 352.1268, found [M + Na]+ 352.1237.
2′-Deoxy-1′-(5-cyanomethyl-3-p-fluorophenyl)pyrazolyl-β-d-ribofuranose (6c): Obtained as a semisolid in a 59% yield. [ α ] D 30 = +47.97° (c 0.25, MeOH); Rf = 0.36 (7% methanol in chloroform); IR (Nujol): 3391 (OH), 2928, 2249 (CN), 1610, 1503, 1448, 1252, 1179, 1055 cm1; 1H NMR (300 MHz, CDCl3): δ 2.24–2.28 (1H, m, C-2′Hα), 2.85–2.89 (1H, m, C-2′Hβ), 3.39–3.41 (1H, m, C-5′Hα), 3.48–3.54 (1H, m, C-5′ Hβ), 3.82–3.87 (1H, m, C-4′H), 4.33 (2H, s, CH2CN), 4.46 (1H, brs, C-3′H), 4.81 (1H, t, J = 5.4 Hz, C-5′OH), 5.27 (1H, d, J = 4.0 Hz, C-3′OH), 6.16 (1H, t, J = 6.0 Hz, C-1′H), 6.76 (1H, s, C-4H), 7.24 (2H, t, J = 7.6 Hz), 7.83 (2H, q, J = 5.3 Hz, Ar-H); 13C NMR (75.5 MHz, DMSO-d6): δ 15.2, 40.3, 63.1, 71.9, 87.0, 89.0, 104.7, 116.5 (d, J = 22 Hz), 117.9, 128.1 (d, J = 8 Hz), 130.0, 135.8, 149.8, 162.9 (d, J = 243 Hz); HRMS (ESI): Calculated for C16H16N3O3F [M + Na]+ 340.1068, found [M + Na]+ 340.1038.
2′-Deoxy-1′-(3-p-chlorophenyl-5-cyanomethyl-)pyrazolyl-β-d-ribofuranose (6d): Obtained as a semisolid in a 61% yield. [ α ] D 30 = +37.58° (c 0.25, MeOH); Rf = 0.36 (7% methanol in chloroform); IR (Nujol): 3387 (OH), 2935, 2249 (CN), 1612, 1504, 1454, 1248, 1179, 1048 cm1; 1H NMR (300 MHz, CD3CN): δ 2.36–2.43 (1H, m, C-2′Hα), 2.88–2.96 (1H, m, C-2′Hβ), 3.43 (1H, brs, C-4′H), 3.56–3.63 (1H, m, C-5′Hα), 3.69–3.73 (1H, m, C-5′ Hβ), 4.05–4.11 (4H, m, CH2CN, C-3′H & C-5′OH), 4.62 (1H, brs, C-3′OH), 6.17 (1H, t, J = 6.1 Hz, C-1′H), 6.72 (1H, s, C-4H), 7.44 (2H, d, J = 8.5 Hz, Ar-H), 7.77 (2H, d, J = 8.5 Hz, Ar-H); 13C NMR (75.5 MHz, CD3CN): δ 14.4, 39.9, 62.6, 71.7, 86.6, 88.6, 103.7, 116.9, 126.6, 128.5, 130.9, 133.1, 134.6, 149.6; HRMS (ESI): Calculated for C16H16N3O3Cl [M + Na]+ 356.0772, found [M + Na]+ 356.0757.
2′-Deoxy-1′-(3-p-bromophenyl-5-cyanomethyl)pyrazolyl-β-d-ribofuranose (6e): Obtained as a white solid in a 63% yield. [ α ] D 30 = +50.37° (c 0.25, MeOH); Mp = 74–75 °C; Rf = 0.39 (7% methanol in chloroform); IR (Nujol): 3385 (OH), 2937, 2258 (CN), 1610, 1501, 1448, 1251, 1183, 1053 cm1; 1H NMR (300 MHz, DMSO-d6): δ 2.34–2.39 (1H, m, C-2′Hα), 2.92–2.98 (1H, m, C-2′Hβ), 3.47 (1H, brs, C-4′H), 3.58 (1H, brs, C-5′Hα), 3.95 (1H, brs, C-5′ Hβ), 4.24 (2H, s, CH2CN), 4.51 (1H, brs, 3′H), 4.77 (1H, brs, C-3′OH), 5.20 (1H, brs, C-5′OH), 6.18 (1H, s, C-1′H), 6.71 (1H, s, C-4H), 7.55 (2H, brs, Ar-H), 7.69 (2H, brs, Ar-H); 13C NMR (75.5 MHz, DMSO-d6): δ 14.9, 39.7, 62.7, 71.4, 86.9, 88.6, 104.3, 116.8, 121.5, 127.3, 129.2, 131.7, 134.9, 149.2; HRMS (ESI): Calculated for C16H16N3O3Br [M + Na]+ 400.0267, found [M + Na]+ 400.0234.

3.3. NCI-60 Human Tumor Cell Line Screen

Details of the methodology are described at http://dtp.nci.nih.gov/branches/btb/ivclsp.html. Briefly, the panel was organized into nine subpanels representing a diverse histology: leukemia, melanoma, and cancers of the lung, colon, kidney, ovary, breast, prostate, and central nervous system. The cells were grown in supplemented RPM1 1640 medium for 24 h. For the five dose study, the test compounds were dissolved in DMSO and incubated with cells at five concentrations with 10-fold dilutions, the highest being 10−4 M and the others being 10−5, 10−6, 10−7, and 10−8 M. The assay was terminated by the addition of cold trichloroacetic acid, and the cells were fixed and stained with sulforhodamine B. The bound stain was solubilized, and the absorbance was read on an automated plate reader. The cytostatic parameter, which determines the 50% growth inhibition (GI50) of the tumor cells, was calculated from time zero, control growth, and the absorbance at the five concentration levels. The cytotoxic parameter, lethal concentration (LC50 is the concentration of a drug resulting in a 50% reduction in the measured protein at the end of the drug treatment when compared to that at the beginning), indicating a net loss of cells following the treatment, was calculated as the average of two independent experiments.

4. Conclusions

We successfully achieved the synthesis of twenty novel pyrazolyl nucleosides 3ae, 4ae, 5ae, and 6ae, which have been characterized in detail by using various NMR spectroscopic techniques such as 1H NMR, 13C NMR, NOESY, HMBC, etc. The pyrazolyl nucleosides 5ae and 6ae were screened for anticancer activities on 60 human tumor cell lines, and we identified one compound 6e, which showed good activity against 39 cancer cell lines and showed significant inhibition against lung cancer cell line Hop-92 (GI50 9.3 µM) and breast cancer cell line HS 578T (GI50 3.0 µM). Our studies have demonstrated the potential of newly synthesized pyrazolyl-nucleosides that will be pursued further.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/24/21/3922/s1, Scanned copies of the NMR and IR spectra of selected compounds, as appropriate are given in the “Supplementary Information”.

Author Contributions

Y.Y., K.K. and D.S. synthesized the compounds; V.K. and S.V.M. compiled the NCI-60 activity; A.J. interpreted the NMR data and drafted the manuscript; J.W., C.L., A.K.P. and V.S.P. planned and designed the whole study and finalized the manuscript.

Funding

Please add: This research received no external funding; See the “Acknowledgements Section” for details.

Acknowledgments

We would like to thank the Department of Biotechnology (DBT, Govt. of India, New Delhi), the University of Delhi (DU, India), and the Nucleic Acid Center (NAC, University of Southern Denmark, Odense, Denmark) for their financial assistance. The authors are grateful to Carl-Erik Olsen, University of Copenhagen, Denmark for carrying out extensive NMR spectral studies and to Eric De Clercq, Rega Institute for Medical Research, K.U. Leuven, Belgium for carrying out the antiviral evaluation studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yates, M.K.; Seley-Radtke, K.L. The evolution of antiviral nucleoside analogues: A review for chemists and non-chemists. Part II: Complex modifications to the nucleoside scaffold. Antivir. Res. 2019, 162, 5–21. [Google Scholar] [CrossRef] [PubMed]
  2. Singh, S.; Bhattarai, D.; Veeraswamy, G.; Choi, Y.; Lee, K. Nucleosides with modified sugar ring: Synthesis and biological activities. Curr. Org. Chem. 2016, 20, 856–897. [Google Scholar] [CrossRef]
  3. Zhou, L.; Zhang, H.; Tao, S.; Ehteshami, M.; Cho, J.H.; McBrayer, T.R.; Tharnish, P.; Whitaker, T.; Amblard, F.; Coats, S.J.; et al. Synthesis and evaluation of 2,6-modified purine 2′-C-methyl ribonucleosides as inhibitors of HCV replication. ACS Med. Chem. Lett. 2016, 7, 17–22. [Google Scholar] [CrossRef] [PubMed]
  4. De Clercq, E. Milestones in the discovery of antiviral agents: Nucleosides and nucleotides. Acta Pharm. Sin. B. 2012, 2, 535–548. [Google Scholar] [CrossRef]
  5. Zhou, C.; Chattopadhyaya, J. The synthesis of therapeutic locked nucleos(t)ides. Curr. Opin. Drug. Discov. Devel. 2009, 12, 876–898. [Google Scholar] [CrossRef]
  6. Wan, J.; Xia, Y.; Liu, Y.; Wang, M.; Rocchi, P.; Yao, J.; Qu, F.; Neyts, J.; Iovanna, J.L.; Peng, L. Discovery of novel arylethynyltriazole ribonucleosides with selective and effective antiviral and antiproloferative activity. J. Med. Chem. 2009, 52, 1144–1155. [Google Scholar] [CrossRef]
  7. Chu, C.K.; Baker, D.C. (Eds.) Nucleosides and Nucleotides as Antitumor and Antiviral Agents; Plenum Press: New York, NY, USA, 1993; pp. 245–264. [Google Scholar]
  8. Prakash, T.P. An overview of sugar-modified oligonucleosides for antisense therapeutics. Chem. Biodivers. 2011, 8, 1616–1641. [Google Scholar] [CrossRef]
  9. Yu, R.Z.; Grundy, J.S.; Geary, R.S. Clinical pharmacokinetics of second generation antisense oligonucleotides. Expert Opin. Drug Metab. Toxicol. 2013, 9, 169–182. [Google Scholar] [CrossRef]
  10. Watts, J.K. Locked nucleic acid: Tighter is different. Chem. Commun. 2013, 49, 5618–5620. [Google Scholar] [CrossRef]
  11. 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]
  12. Len, C.; Postel, D. Synthesis of 2’,3’-didehydro-2’,3’-dideoxynucleosides via nucleoside route. Curr. Org. Synth. 2006, 3, 261–281. [Google Scholar] [CrossRef]
  13. Len, C.; Mackenzie, G. Synthesis of 2’,3’-didehydro-2’,3’-dideoxynucleosides having variations at either or both of the 2’- and 3’-positions. Tetrahedron 2006, 62, 9085–9107. [Google Scholar] [CrossRef]
  14. Len, C.; Mondon, M.; Lebreton, J. Synthesis of cyclonucleosides having a C-C bridge. Tetrahedron 2008, 64, 7453–7475. [Google Scholar] [CrossRef]
  15. Wengel, J. Synthesis of 3’-C- and 4’-C-branched oligodeoxynucleotides and the development of locked nucleic acid (LNA). Acc. Chem. Res. 1999, 32, 301–310. [Google Scholar] [CrossRef]
  16. Petersen, M.; Wengel, J. LNA: A versatile tool for therapeutics and genomics. Trends Biotechnol. 2003, 21, 74–81. [Google Scholar] [CrossRef]
  17. Kalra, N.; Babu, B.R.; Parmar, V.S.; Wengel, J. Conformationally controlled high-affinity targeting of RNA or DNA by novel 2’-amino-DNA/LNA mixmers and pyrenyl-functionalized 2’-amino-DNA. Org. Biomol. Chem. 2004, 2, 2885–2887. [Google Scholar] [CrossRef]
  18. Babu, B.R.; Hrdlicka, P.J.; McKenzie, C.J.; Wengel, J. Optimizing DNA targeting using N,N-bis(2-pyridylmethyl)-|?-alanyl 2’-amino-LNA. Chem. Commun. 2005, 1705–1707. [Google Scholar] [CrossRef]
  19. Hrdlicka, P.J.; Jepsen, J.S.; Nielsen, C.; Wengel, J. Synthesis and biological evaluation of nucleobase-modified analogs of the anticancer compounds 3’-C-ethynyluridine (EUrd) and 3’-C-ethynylcytidine (ECyd). Bioorg. Med. Chem. 2005, 13, 1249–1260. [Google Scholar] [CrossRef]
  20. Prakash, T.P.; Prhavc, M.; Eldrup, A.B.; Cook, P.D.; Carroll, S.S.; Olsen, D.B.; Stahlhut, M.W.; Tomassini, J.E.; MacCoss, M.; Galloway, S.M.; et al. Synthesis and evaluation of S-acyl-2-thioethyl esters of modified nucleoside 5’-monophosphates as inhibitors of hepatitis C virus RNA replication. J. Med. Chem. 2005, 48, 1199–1210. [Google Scholar] [CrossRef]
  21. Sharma, V.K.; Watts, J.K. Oligonucleotide therapeutics: Chemistry, delivery and clinical progress. Future Med. Chem. 2015, 7, 2221–2242. [Google Scholar] [CrossRef]
  22. Reid, G.; Wielinga, P.; Zelcer, N.; de Haas, M.; van Deemter, L.; Wijnholds, J.; Balzarini, J.; Borst, P. Characterization of the transport of nucleoside analog drugs by the human multidrug resistance proteins MRP4 and MRP5. Mol. Pharmacol. 2003, 63, 1094–1103. [Google Scholar] [CrossRef] [PubMed]
  23. Parmar, V.S.; Jain, S.C.; Jha, A.; Kumar, N.; Kumar, A.; Vats, A.; Jha, H.N.; Mukherjee, S.; Singh, S.K.; Jennings, K.R.; et al. Synthetic and mass spectral fragmentation studies on trisubstituted 2H-pyran-2-ones and comparative EIMS behavior of biologically active 3,5-disubstituted pyrazoles and isoxazoles. Indian J. Chem. 1997, 36B, 872–879. [Google Scholar] [CrossRef]
  24. Sharma, V.K.; Sharma, R.K.; Singh, P.K.; Singh, S.K. An engrossing history of azidothymidine. Immun. Endoc. Metab. Agents Med. Chem. 2015, 15, 168–175. [Google Scholar] [CrossRef]
  25. Sidwell, R.W.; Huffman, J.H.; Khare, G.P.; Allen, L.B.; Witkowski, J.T.; Robins, R.K. Broad-spectrum antiviral activity of virazole: 1-β-D-ribofuranosyl-1,2,4-triazole-3-carboxamide. Science 1972, 177, 705–706. [Google Scholar] [CrossRef] [PubMed]
  26. Storer, R.; Ashton, C.J.; Baxter, A.D.; Hann, M.M.; Marr, C.L.P.; Mason, A.M.; Mo, C.L.; Myers, P.L.; Noble, S.A.; Penn, C.R.; et al. The synthesis and antiviral activity of 4-fluoro-1-β-D-ribofuranosyl-1H-pyrazole-3-carboxamide. Nucleosides Nucleotides 1999, 18, 203–216. [Google Scholar] [CrossRef] [PubMed]
  27. Balzarini, J.; Kang, G.J.; Dalal, M.; Herdewijn, P.; De Clercq, E.; Broder, S.; Johns, D.G. The anti-HTLV-III (anti-HIV) and cytotoxic activity of 2’,3’-didehydro-2’,3’-dideoxynucleosides: A comparison with their parental 2’,3’-dideoxyribonucleosides. Mol. Pharmacol. 1987, 32, 162–167. [Google Scholar]
  28. Hamamoto, Y.; Nakashima, H.; Matsui, T.; Mateuda, A.; Ueda, T.; Yamamoto, N. Inhibitory effect of 2’,3’-didehydro-2’,3’-dideoxynucleosides on infectivity, cytopathic effects, and replication of human immunodeficiency vírus. Antimicrob. Agents Chemother. 1987, 31, 907–910. [Google Scholar] [CrossRef]
  29. Lin, T.S.; Guo, J.Y.; Schinazi, R.F.; Chu, C.K.; Xiang, J.N.; Prusoff, W.H. Synthesis and antiviral activity of various 3’azido analogues of pyrimidine deoxyribonucleosides against Human Immunodeficiency Virus (HIV-1, HTLV-III/LAV). J. Med. Chem. 1988, 31, 336–340. [Google Scholar] [CrossRef]
  30. Haneishi, T.; Okazaki, T.; Hata, T.; Tamura, C.; Nomura, M.; Naito, A.; Seki, I.; Arai, M. Oxazinomycin, a new carbon-linked nucleoside antibiotic. J. Antibiot. 1971, 24, 797–799. [Google Scholar] [CrossRef]
  31. Sweeney, M.J.; Davis, F.A.; Gutowski, G.E.; Hamill, R.L.; Hoffman, D.H.; Poore, G.A. Experimental antitumor activity of pyrazomycin. Cancer Res. 1973, 33, 2619–2623. [Google Scholar]
  32. Harusawa, S.; Matsuda, C.; Araki, L.; Kurihara, T. Efficient and β-steroselective synthesis of pyrazole C-nucleosides. Synthesis 2006, 793–798. [Google Scholar] [CrossRef]
  33. Nishimura, H.; Mayama, M.; Komatsu, Y.; Kato, H.; Shimaoka, N.; Tanaka, Y. Showdomycin, a new antibiotic from a Streptomyces Sp. J. Antibiot. Ser. A. 1964, 17, 148–155. [Google Scholar]
  34. Michalik, D.; Peseke, K. Syntheses of pyrazole iso-C-nucleosides. J. Carbohydr. Chem. 2000, 19, 1049–1057. [Google Scholar] [CrossRef]
  35. Hori, M.; Ito, E.; Takita, T.; Koyama, G.; Takeuchi, T.; Umezawa, H. A new antibiotics, Formycin. J. Antibiot. Ser. A. 1964, 17, 96–99. [Google Scholar]
  36. Zhou, J.; Yang, M.; Akdag, A.; Wang, H.; Schneller, S.W. Carbocyclic 4’-epi-formycin. Tetrahedron 2008, 64, 433–438. [Google Scholar] [CrossRef]
  37. Moriyama, K.; Suzuki, T.; Negishi, K.; Graci, J.D.; Thompson, C.N.; Cameron, C.E.; Watanabe, M. Effects of introduction of hydrophobic group on ribavirin base on mutation induction and anti-RNA viral activity. J. Med. Chem. 2008, 51, 159–166. [Google Scholar] [CrossRef]
  38. Manfredini, S.; Baraldi, P.G.; Bazzanini, R.; Durini, E.; Vertuani, S.; Pani, A.; Marceddu, T.; Demontis, F.; Vargiu, L.; La Calla, P. Pyrazole related nucleosides 5. Synthesis and biological activity of 2’-deoxy-2’,3’-dideoxy- and acyclo-analogues of 4-iodo-1-β-D-ribofuranosyl-3-carboxymethyl pyrazole (IPCAR). Nucleosides Nucleotides Nucleic Acids 2000, 19, 705–722. [Google Scholar] [CrossRef]
  39. Rauter, A.P.; Figueiredo, J.A.; Ismael, M.I.; Justino, J. Synthesis of new pseudo-C-nucleosides containing pyrazole rings in their structure. J. Carbohydr. Chem. 2004, 23, 513–528. [Google Scholar] [CrossRef]
  40. Zhang, J.; Visser, F.; King, K.M.; Baldwin, S.A.; Young, J.D.; Cass, C.E. The role of nucleoside transporters in cancer chemotherapy with nucleoside drugs. Cancer Metastasis Rev. 2007, 26, 85–110. [Google Scholar] [CrossRef]
  41. Podgorska, M.; Kocbuch, K.; Pawelczyk, T. Recent advances in studies on biochemical and structural properties of equilibrative and concentrative nucleoside transporter. Acta Biochim. Pol. 2005, 52, 749–758. [Google Scholar]
  42. Uhlmann, E.; Peyman, A. Antisense oligonucleotides: A new therapeutic principle. Chem. Rev. 1990, 90, 544–584. [Google Scholar] [CrossRef]
  43. Dhimitruka, I.; Santa Lucia, J., Jr. Efficient preparation of 2-deoxy-3,5-di-O-p-toluoyl-α-D-ribofuranosyl chloride. Synlett 2004, 335–337. [Google Scholar] [CrossRef]
  44. Chenon, M.T.; Coupry, C.; Grant, D.M.; Pugmire, R.J. Carbon-13 magnetic resonance study of solvent stabilized tautomerism in pyrazoles. J. Org. Chem. 1977, 42, 659–661. [Google Scholar] [CrossRef]
  45. Bergstrom, D.E.; Zhang, P.; Johnson, W.T. Design and synthesis of heterocyclic carboxamides as natural nucleic acid base mimics. Nucleosides Nucleotides 1996, 15, 59–68. [Google Scholar] [CrossRef]
  46. Schneller, S.W. Carbocyclic nucleosides (Carbanucleosides) as new therapeutic leads. Curr. Top. Med. Chem. 2002, 2, 1087–1092. [Google Scholar] [CrossRef]
  47. Larsen, J.S.; Zahran, M.A.; Pedersen, E.B.; Nielsen, C. Synthesis of triazenopyrazole derivatives as potential inhibitors of HIV-1. Monatsh. Chem. 1999, 130, 1167–1173. [Google Scholar] [CrossRef]
  48. Tominaga, Y.; Ushirogochi, A.; Matsuda, Y.; Koboyashi, G. Synthesis and reactions of 6-aryl- and 6-styryl-3-cyano-4-methylthio-2H-pyran-2-ones. Chem. Pharm. Bull. 1984, 32, 3384–3395. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from the corresponding author, V.S.P.
Figure 1. Selected nucleoside analogues with antiviral and/or antitumoral activities: 3’-azidothymidine (AZT or ZDV), showdomycin, ribavirin, pyrazofurin, formycin A and formycin B.
Figure 1. Selected nucleoside analogues with antiviral and/or antitumoral activities: 3’-azidothymidine (AZT or ZDV), showdomycin, ribavirin, pyrazofurin, formycin A and formycin B.
Molecules 24 03922 g001
Scheme 1. Synthesis of pyrazolylnucleoside analogues 5ae and 6ae.
Scheme 1. Synthesis of pyrazolylnucleoside analogues 5ae and 6ae.
Molecules 24 03922 sch001
Scheme 2. Mechanistic explanation for the regioisomeric alkylation of pyrazole derivatives 1ae. The IUPAC numbering scheme of 3ae is retained in 4ae (and in 5ae and 6ae) for easy comparison.
Scheme 2. Mechanistic explanation for the regioisomeric alkylation of pyrazole derivatives 1ae. The IUPAC numbering scheme of 3ae is retained in 4ae (and in 5ae and 6ae) for easy comparison.
Molecules 24 03922 sch002
Table 1. Chemical shift values of the anomeric protons in the 1H NMR spectra and chemical shift values of the Ar- and CH2CN bearing carbons of the pyrazole ring in the 13C NMR spectra of the isomeric pyrazolyl nucleosides 3ae, 4ae, 5ae, and 6ae in CDCl3 on a Bruker Avance 300 spectrometer.
Table 1. Chemical shift values of the anomeric protons in the 1H NMR spectra and chemical shift values of the Ar- and CH2CN bearing carbons of the pyrazole ring in the 13C NMR spectra of the isomeric pyrazolyl nucleosides 3ae, 4ae, 5ae, and 6ae in CDCl3 on a Bruker Avance 300 spectrometer.
CompoundC-1′ H Shift in the 1H NMR Spectrum (ppm)Ar- Bearing C Shifts in the 13C NMR Spectrum (ppm)–CH2CN Bearing C Shifts in the 13C NMR Spectrum (ppm)
Series 3Series 5Series 3Series 5Series 3Series 5
a6.135.99144.6145.9141.7142.0
b6.106.02146.2145.8144.0141.8
c6.055.97145.4144.9141.9141.9
d6.056.00145.2144.7143.7141.9
e6.056.00145.6143.9142.3141.1
Series 4Series 6Series 4Series 6Series 4Series 6
a6.176.13151.0150.2137.9137.7
b6.166.19150.8150.7133.0133.7
c6.176.16152.5149.8135.8135.8
d6.196.17144.1149.6134.3134.6
e6.176.18149.8149.2133.5134.9
Table 2. Chemical shift and coupling constants from the 1H-NMR of compounds 3d, 4d, 5d, and 6d in acetone-d6.a Please refer to structures in Table 4 for skeleton numbering.
Table 2. Chemical shift and coupling constants from the 1H-NMR of compounds 3d, 4d, 5d, and 6d in acetone-d6.a Please refer to structures in Table 4 for skeleton numbering.
ProtonChemical Shifts (δ)Coupling Constants (Hz)
3d4d5d6d3d4d5d6d
H1′6.296.486.096.29dd 5.7, 6.5dd 6.5, 5.2t* 6.4t* 6.2
H2′α2.722.792.312.42ddd
14.0, 6.7, 4.1
ddd
14.0, 4.5, 6.6
ddd
13.5, 3.7, 6.8
ddd
13.5, 3.7, 6.7
H2′β3.543.633.013.00ddd
14.0, 6.8, 5.6
ddd
14.0, 5.2, 6.7
dt*
13.5, 6.0
dt*
13.5, 5.9
H3′5.935.974.664.70mmmm
H4′4.624.654.024.07mmmm
H5′α4.544.453.613.59mdd 13.2, 6.4ddd 11.8, 4.8, 7.7ddd 11.9, 4.1, 8.1
H5′β4.634.633.703.71mmdt* 11.9, 4.7dt* 11.9,4.1
H46.506.826.466.83brs, 1Ht* 0.8t* 0.5t 0.8
H63.904.293.964.31brs, 2H19.0, 0.8d 0.518.9, 0.8
H2″7.677.877.607.87
H3″ 7.577.46
a The chemical shift values presented in Table 2 were all measured in acetone-d6 (on an Bruker Avance 400 instrument) and may therefore differ from those in Table 1, which were measured in CDCl3 (on Bruker Avance 300 instrument). * Due to the spatial interactions between different protons as explicitly shown in Table 4, these protons exhibited multiplicities as shown here when the spectra were recorded in acetone-d6 on a Bruker Avance 400 instrument as against those recorded in CDCl3, CD3CN or DMSO-d6 as given in the Experimental section for the corresponding protons where they appeared as broad singlets or ill resolved multiplets when their spectra were recorded on a Bruker Avance 300 instrument.
Table 3. Chemical shift values from the 13C-NMR of compounds 3d, 4d, 5d, and 6d in acetone-d6.a Please refer to the structures in Table 4 for skeleton numbering.
Table 3. Chemical shift values from the 13C-NMR of compounds 3d, 4d, 5d, and 6d in acetone-d6.a Please refer to the structures in Table 4 for skeleton numbering.
Chemical Shifts (δ) (Carbon Position)
3d5d4d6d
87.6 (C1′)87.4 (C1′)87.6 (C1′)88.0 (C1′)
37.1 (C2′)40.5 (C2′)37.0 (C2′)41.1 (C2′)
76.6 (C3′)73.2 (C3′)76.1 (C3′)73.0 (C3′)
83.5 (C4′)89.8 (C4′)83.7 (C4′)90.0 (C4′)
65.1 (C5′)64.2 (C5′)64.7 (C5′)63.9 (C5′)
143.8 (C3)143.6 (C3)136.3 (C3)136.1 (C3)
107.2 (C4)106.7 (C4)105.4 (C4)104.9 (C4)
146.0 (C5)145.8 (C5)150.6 (C5)150.5 (C5)
17.6 (C6)17.6 (C6)15.3 (C6)15.2 (C6)
118.1 (CN)118.0 (CN)117.0 (CN)117.1 (CN)
166.4 (C6′′′) 166.5 (C6′′′)
166.5 (C6′′′′) 166.5 (C6′′′′)
*129.3 (C1′′)#134.2 (C1′′)
*131.6 (C2′′)#127.9 (C2′′)
*129.9 (C3′′)#129.7 (C3′′)
*135.6 (C4′′)#132.6 (C4′′)
a The chemical shift values presented in Table 3 were all measured in acetone-d6 (on an Bruker Avance 400 instrument) and may therefore differ from those in Table 1, which were measured in CDCl3 (on Bruker Avance 300 instrument). * Benzene ring Carbons: Quaternary C: 145.0, 144.7, 135.6, 129.3, 128.3, 128.1; CHs: 131.6, 130.5, 130.5, 130.0, 130.0, 129.9. # Benzene ring Carbons: Quaternary C: 145.0, 144.5, 134.1, 132.7, 128.1; CHs: 130.5, 130.4, 130.1, 129.9, 129.5.
Table 4. Results from the NOESY and HMBC spectra of the pairs 3d4d and 5d6d *.
Table 4. Results from the NOESY and HMBC spectra of the pairs 3d4d and 5d6d *.
Molecules 24 03922 i001
Blue and red arrows indicate the NOESY and HMBC cross peaks, respectively.
3d4d5d6d
NOESY H1′-H6oxox
NOESY H1′-H2′′xoxo
NOESY H4-H2′′xxxx
HMBC H1′-C5xoxo
HMBC H1′-C3oxox
* The presence of a cross peak is indicated by (x). When it contributes to the positional verification, it is underlined (x). See Table 1, Table 2 and Table 3 for positions of the relevant protons and carbons in the NMR spectra of the compounds 3d, 4d, 5d and 6d.
Table 5. Antitumor activity (GI50/µM) a and toxicity (LC50/µM) b data of 6d and 6e with the anticancer drug dasatinib as a positive control.
Table 5. Antitumor activity (GI50/µM) a and toxicity (LC50/µM) b data of 6d and 6e with the anticancer drug dasatinib as a positive control.
Panels/Cell Lines6d6eDasatinib
GI50LC50GI50LC50GI50LC50
Leukemia
CCRF-CEM>67.5>67.546.3>57.55.3>100
K-562>67.5>67.540.0>57.50.01>100
MOLT-4>67.5>67.530.9>57.54.1>100
RPMI-8226>67.5>67.534.0>57.54.899.5
SR>67.5>67.525.5>57.53.189.3
Non-Small Cell Lung Cancer
A549/ATCC44.8>67.529.3>57.50.0575.5
HOP-92>67.5>67.59.3>57.50.01>100
NCI-H322M29.8>67.530.2>57.50.0427.0
NCI-H522>67.5>67.525.7>57.50.0655.1
Colon Cancer
HCT-116>67.5>67.547.7>57.511.869.8
HCT-15>67.5>67.545.1>57.50.671.1
CNS Cancer
SF-26833.5>67.520.5>57.50.0775.5
SF-295>67.5>67.524.8>57.51.146.8
SNB-1924.6>67.518.5>57.511.976.4
SNB-7553.5>67.510.5>57.50.0146.3
U251>67.5>67.536.4>57.53.252.4
Melanoma
LOX IMVI>67.5>67.549.7>57.50.012.8
M14>67.5>67.557.1>57.53.151.5
MDA-MB-435>67.5>67.553.0>57.54.165.5
SK-MEL-2>67.5>67.539.6>57.51.171.8
SK-MEL-560.9>67.511.8>57.54.544.3
UACC-6237.8>67.510.5>57.52.847.0
Ovarian Cancer
IGROV132.6>67.521.5>57.50.0180.0
OVCAR-3>67.5>67.539.1>57.50.274.6
OVCAR-5>67.5>67.544.4>57.50.0286.3
Renal Cancer
A49822.6>67.518.8>57.50.0316.0
ACHN>67.5>67.545.1>57.50.01520
CAK-122.9>67.518.8>57.50.015.1
RXF 39338.0>67.534.6>57.50.0310.4
SN12C48.4>67.524.3>57.50.0544.4
UO-3119.4>67.516.6>57.50.0182.0
Prostate Cancer
PC-337.4>67.520.0>57.50.292.5
DU-145>67.5>67.545.1>57.50.14.9
Breast Cancer
MCF748.0>67.532.8>57.512.671.4
MDA-MB-231ATCC>67.5>67.555.8>57.50.0138.0
HS 578T16.6>67.53.0>57.50.01>100
BT-54949.9>67.530.0>57.55.949.8
T-47D26.6>67.519.4>57.50.291.4
MDA-MB-46831.6>67.513.8>57.50.098.4
a GI50: 50% Growth inhibition, concentration of drug resulting in a 50% reduction in net protein increase when compared with the control cells. b LC50: Lethal concentration, concentration of drug lethal to 50% of cells.

Share and Cite

MDPI and ACS Style

Yadav, Y.; Sharma, D.; Kaushik, K.; Kumar, V.; Jha, A.; Prasad, A.K.; Len, C.; Malhotra, S.V.; Wengel, J.; Parmar, V.S. Synthetic, Structural, and Anticancer Activity Evaluation Studies on Novel Pyrazolylnucleosides. Molecules 2019, 24, 3922. https://doi.org/10.3390/molecules24213922

AMA Style

Yadav Y, Sharma D, Kaushik K, Kumar V, Jha A, Prasad AK, Len C, Malhotra SV, Wengel J, Parmar VS. Synthetic, Structural, and Anticancer Activity Evaluation Studies on Novel Pyrazolylnucleosides. Molecules. 2019; 24(21):3922. https://doi.org/10.3390/molecules24213922

Chicago/Turabian Style

Yadav, Yogesh, Deepti Sharma, Kumar Kaushik, Vineet Kumar, Amitabh Jha, Ashok K. Prasad, Christophe Len, Sanjay V. Malhotra, Jesper Wengel, and Virinder S. Parmar. 2019. "Synthetic, Structural, and Anticancer Activity Evaluation Studies on Novel Pyrazolylnucleosides" Molecules 24, no. 21: 3922. https://doi.org/10.3390/molecules24213922

Article Metrics

Back to TopTop