Next Article in Journal
1-(2,5-Dimethoxy-4-nitrophenyl)piperidine
Previous Article in Journal
Synthesis of Norabietyl and Nordehydroabietyl Imidazolidine-2,4,5-Triones and Their Activity against Tyrosyl-DNA Phosphodiesterase 1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis and In Silico Drug-Likeness Modeling of 5-FU/ASA Hybrids

by
Wilson Castrillón-López
,
Andrés F. Yepes
and
Wilson Cardona-Galeano
*
Chemistry of Colombian Plants Group, Institute of Chemistry, Faculty of Exact and Natural Sciences, University of Antioquia, Calle 70 No. 52—21, A.A 1226, Medellín 050010, Colombia
*
Author to whom correspondence should be addressed.
Molbank 2023, 2023(4), M1745; https://doi.org/10.3390/M1745
Submission received: 27 September 2023 / Revised: 17 November 2023 / Accepted: 23 November 2023 / Published: 27 November 2023
(This article belongs to the Section Organic Synthesis)

Abstract

:
A series of 5-FU-ASA hybrids were synthesized with good yields using click chemistry as the key step. The structures of these compounds were elucidated by spectroscopic analysis. Finally, an optimal pharmacokinetic profile was also estimated for each synthetized hybrid. Taken together, hybrids 4ah could be used as starting points for further pharmacological studies concerning therapeutic cancer intervention.

Graphical Abstract

1. Introduction

Molecular hybridization is the combination of two or more pharmacophores in one molecule. This has emerged as a promising strategy in medicinal chemistry in the search of new therapeutic alternatives to treat colorectal cancer [1,2], which is the second deadliest and widely diagnosed cancer in the world, accounting for 10% of all cancers [3]. Treatment used clinically for colorectal cancer (CRC) is based on 5-fluorouracil (5-FU), an antimetabolite of the pyrimidine analogue type. In addition, acetylsalicylic acid (ASA) an analgesic and anti-inflammatory agent [4] has shown chemopreventive potential over colon cancer [5]. On the other hand, click chemistry refers to a group of reactions that are fast, simple to use, easy to purify, versatile, regiospecific, and provide high product yields. An example of these reactions is the CuI-catalyzed Huisgen 1,3-dipolar cycloaddition of azides and terminal alkynes to form 1,2,3-triazoles. This methodology has been applied to the synthesis of diverse pharmaceutical agents suitable for medicinal chemistry, including hybrids molecules, and drugs discovery [6,7,8,9,10].
Based on the anticancer potential of both 5-FU and ASA and the need for new thera-peutic alternatives for the treatment of CRC, we designed and synthesized several hydrolyzable conjugated hybrids 5-FU-ASA (Figure 1), which were obtained via click reaction between different ASA-alkylazides and propargyl-5-FU. Moreover, pharmacokinetic modelling studies were conducted aiming at investigating the potential of the synthetized hybrids as drug-like molecules.

2. Results and Discussion

2.1. Chemistry

The synthesis of the hybrids is shown in Scheme 1. It began with the esterification of ASA with 1,ω-dibromoalkanes (ω = 3, 4, 5, 8, 9, and 12) producing ASA-bromoalkyls 1ah with yields ranging between 32% and 57%. Compounds 1ah were treated with sodium azide leading to the formation of the ASA-alkylazides 2ah with 72–96% yields. These compounds have already been reported [11]. Reaction of 5-FU with propargyl bromide led to the 5-FU-alkyne (3) with a 70% yield [12]. Finally, ultrasound assisted click reactions between azides 2ah and alkyne 3 led to the formation of hybrids 4ah with 60–80% yields [10]. The use of sonication in the Huisgen reaction reduces the reaction time, which is, on average, 24 h [13,14], and in our case, 1 h.
The structures of all compounds have been established by a combined study of HRMS-ESI (m/z), 1H-NMR, and 13C-NMR spectra. HRMS-ESI (m/z) spectra showed characteristic [M + H]+ peaks corresponding to their molecular weights. The assignments of all signals to individual H or C-atoms were carried out based on typical δ-values and J-constants. The 1H-NMR spectra of the hybrids dissolved in DMSO-d6 showed signals belonging to 5-FU (CH–C–F) around 8.10 ppm as a doublet. 5-FU-CH2-triazolyl (4.87 ppm). The acetyl group of ASA was observed around 2.25 ppm. 13C-NMR spectra of the hybrids showed signals corresponding to the carbonyl groups of the ASA around 169 and 163 ppm. The signal corresponding to the C-F of 5-FU appeared as a doublet around 141 and 139 ppm. The triazolyl ring exhibited signals around 142.2 and 124 ppm.

2.2. In Silico Pharmacokinetic Analysis of Conjugates 4ah

In silico calculations currently represent one of the most feasible and fast methodologies for accessing the pharmacokinetic and physicochemical properties of novel drug potential candidates. Particularly, a rigorous analysis of this data can be used as an efficient filter in the development of new oncolytic candidates. In this line, early predictions of biopharmaceutical profiling could enhance the probability of success in drug discovery settings followed by a significant reduction of time and money, thereby promoting further preclinical and clinical experiments for a lead compound [15,16]. Based on its chemical structures, the SwissADME web tool was used to evaluate seven key biopharmaceutical parameters for conjugates 4ah and then compared against those of 95% of approved drugs (Table 1) in order to investigate if novel hybrids can be regarded as valid development starting points for cancer drug discovery. Notably, favorable pharmacokinetics indices were found for conjugates compared to 95% of FDA-approved drugs. According to Lipinski’s rule of five [16,17] (there should be no more than one violation), the synthetized hybrids could apparently be used as oral systemic drugs in humans. Thus, the degree of lipophilicity (calculated as logPo/w) was predicted to be in about 0.61–5.19, fitting well within the ideal range for lipid-based formulations [18] (−2.0 to 6.0).
In addition, we also calculated the PSA parameter, which together with the logPo/w value correlates passive molecular transport through membranes and drug-membrane interactions. [19] Based on our analysis, for all tested compounds, an estimated PSA value was found to be 138.19 Å2, which fits well within the recommended range for oral drugs candidates (7.0 to 200 Å2).
Finally, the pan-assay interference compounds (PAINS) filter which provides specific information on the potential toxicity of promiscuous compounds [20], showed that the novel hybrids could be used as safe starting prototypes for drug cancer development. In sum, our modelling showed that these novel molecules are projected to have the best pharmacokinetic qualities for further chemical biology projects.

3. Materials and Methods

3.1. Chemical Synthesis

5-FU (≥98.0%) and Aspirin (≥99.0%, Acetylsalicylic acid) were purchased from AK scientific and chemicals (Union City, CA, USA). Ultrasound equipment (BRANSON) was used to assist the reactions. NMR spectra were recorded on an AMX 300 instrument (Bruker, Billerica, MA, USA) operating at 300 MHz for 1H and 75 MHz for 13C. The signals of the deuterated solvents were used as references and the chemical shifts (δ) were displayed in ppm. TMS was used as an internal standard. Coupling constants (J) are given in Hertz (Hz). HRMS was obtained using a Bruker Impact II UHR-Q-TOF mass spectrometer (Bruker Daltonik GmbH, Bremen Germany) in positive mode. For column chromatography and thin layer chromatography (TLC), silica gel 60 (0.063–0.200 mesh, Merck, Whitehouse Station, NJ, USA) and precoated silica gel plates (Merck 60 F254 0.2 mm) were used.

General Procedure for the Synthesis of 5-FU-ASA Hybrids 4a4h

Hybrids 4ah were synthesized following a procedure described in the literature [11]. In a 10 mL flat-bottomed flask, ASA-alkylazides (2ah) (1 mmol), propargyl-5-FU (3) (1 mmol), and DMF (5 mL) were placed, then the mixture was sonicated for 5 min to 40 °C. After this time, a mixture of ascorbic acid (0.5 mmol), copper acetate (0.5 mmol), DMF (1 mL), and water (1 mL) was added, and the reaction mixture sonicated for 1 h to 40 °C. Then, 10% HCl was added and extracted with ethyl acetate. The organic phase was dried on anhydrous sodium sulfate, filtered, and concentrated under reduced pressure, and the residue was subjected to crystallization (MeOH:H2O, 1:1 ratio). Finally, the solid obtained was purified by preparative chromatography on silica gel to obtain compounds 4ah. The 1H, 13C NMR, and MS spectra of all hybrids can be found in the Supplementary Materials.
2-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)ethyl 2-acetoxybenzoate (4a): Yield 71%, solid yellow, m.p. 136–139 °C; 1H RMN (300 MHz, DMSO-d6): δ 8.19 (s, 1H; triazole-H), 8.14 (d, J = 6.7 Hz, 1H; 5-FU-H), 7.81 (dd, J = 7.8, 1.7 Hz, 1H), 7.68 (td, J = 7.8, 1.7 Hz, 1H), 7.40 (td, J = 7.6, 1.2 Hz, 1H), 7.23 (dd, J = 8.2, 1.1 Hz, 1H), 4.91 (s, 2H; triazole-CH2–), 4.75 (t, J = 7.0 Hz, 2H, –OCH2–), 4.62 (t, J = 6.2 Hz, 2H, –NCH2–), 2.20 (s, 3H). 13C RMN (75 MHz, DMSO-d6): δ 169.55 (C=O), 163.80 (C=O), 158.30 and 157.97 (F–C–C=O), 150.61, 150.01, 142.85 (triazolyl), 141.68 y 138.59 (F–C), 135.03, 131.62, 130.48 and 130.04 (CH–C–F), 126.76, 124.60 (triazolyl), 124.54, 122.86, 63.60, 48.98, 43.18, 21.08. HRMS-ESI (m/z): calcd for C18H17FN5O6 [M + H]+: 418.1118, found: 418.1157.
3-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)propyl 2-acetoxybenzoate (4b): Yield 80%, solid yellow, m.p. 131–134 °C; 1H RMN (300 MHz, DMSO-d6): δ 8.14 (s, 1H; triazole–H), 8.08 (d, J = 6.6 Hz, 1H; 5-FU-H), 7.94 (dt, J = 7.9, 1.2 Hz, 1H), 7.69 (td, J = 7.4, 1.8 Hz, 1H), 7.42 (td, J = 7.5, 1.0 Hz, 1H), 7.25 (dd, J = 8.1, 1.1 Hz, 1H), 4.88 (s, 2H; triazole-CH2–), 4.49 (t, J = 7.0 Hz, 2H, –OCH2–), 4.23 (t, J = 6.2 Hz, 2H, –NCH2–), 2.28 (s, 3H), 2.28–2.23 (m, 2H). 13C RMN (75 MHz, DMSO-d6): δ 169.65 (C=O), 164.28 (C=O), 150.52, 142.93 (triazolyl), 141.01 and 138.23 (F–C), 134.85, 131.79, 130.17 and 129.72 (CH–C–F), 126.74, 124.48 (triazolyl), 124.14, 123.31, 62.53, 47.01, 43.18, 29.26, 21.23. HRMS-ESI (m/z): calcd for C19H19FN5O6 [M + H]+: 432.1275, found: 432.1316.
4-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)butyl 2-acetoxybenzoate (4c): Yield 60%, solid yellow, m.p. 126–129 °C; 1H RMN (300 MHz, DMSO-d6): 8.09 (s, 1H; triazole-H), 8.02 (d, J = 6.5 Hz, 1H; 5-FU-H), 7.94 (dt, J = 7.8, 1.5 Hz, 1H), 7.68 (td, J = 7.3, 1.2 Hz, 1H), 7.42 (td, J = 7.5, 1.0 Hz, 1H), 7.24 (dd, J = 8.0, 1.4 Hz, 1H), 4.87 (s, 2H; triazole-CH2–), 4.40 (t, J = 7.0 Hz, 2H, –OCH2–), 4.24 (t, J = 6.5 Hz, 2H, –NCH2–), 2.25 (s, 3H), 1.94–1.89 (m, 2H), 1.67–1.63 (m, 2H). 13C RMN (75 MHz, DMSO-d6): δ 169.60 (C=O), 164.49 (C=O), 157.44 and 157.06 (F–C–C=O), 150.39, 143.04 (triazolyl), 141.17 and 139.04 (F–C), 134.74, 131.67, 129.77 and 129.32 (CH–C–F), 126.78, 124.49 (triazolyl), 123.96, 123.59, 49.40, 43.19, 26.78, 24.43, 21.18. HRMS-ESI (m/z): calcd for C20H21FN5O6 [M + H]+: 446.1431, found: 446.1470.
5-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)pentyl 2-acetoxybenzoate (4d): Yield 70%, solid yellow, m.p. 127–130 °C; 1H RMN (300 MHz, DMSO-d6): δ 8.09 (d, 2H; FU-H, triazole-H), 7.92 (dd, J = 7.7, 1.7 Hz, 1H), 7.68 (td, J = 7.3, 1.2 Hz, 1H), 7.42 (td, J = 7.6, 1.1 Hz, 1H), 7.24 (dd, J = 8.1, 1.1 Hz, 1H), 4.88 (s, 2H; triazole-CH2–), 4.35 (t, J = 7.1 Hz, 2H, –OCH2–), 4.21 (t, J = 6.5 Hz, 2H, –NCH2–), 2.27 (s, 3H), 1.89–1.84 (m, 2H), 1.72–1.68 (m, 2H), 1.39–1.31 (m, 2H). 13C RMN (75 MHz, DMSO-d6): δ 169.59 (C=O), 164.49 (C=O), 157.25 and 156.97 (F–C–C=O), 153.08, 152.20, 150.40, 142.78 (triazolyl), 141.09 and 138.04 (F–C), 134.71, 131.64, 130.20 and 130.04 (CH–C–F), 126.77, 124.49 (triazolyl), 123.89, 123.64, 65.06, 49.68, 43.20, 29.72, 27.91, 22.83, 21.22. HRMS-ESI (m/z): calcd for C21H23FN5O6 [M + H]+: 460.1588, found: 460.1621
6-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)hexyl 2-acetoxybenzoate (4e): Yield 75%, solid yellow, m.p. 109–111 °C; 1H RMN (300 MHz, DMSO-d6): δ 8.11 (d, J = 6.6 Hz, 1H; 5-FU-H), 8.09 (s, 1H; triazole-H), 7.93 (dd, J = 7.9, 1.7 Hz, 1H), 7.68 (td, J = 7.8, 1.7 Hz, 1H), 7.42 (td, J = 7.6, 1.2 Hz, 1H), 7.24 (dd, J = 8.1, 1.1 Hz, 1H), 4.88 (s, 2H; triazole-CH2–), 4.33 (t, J = 7.1 Hz, 2H, –OCH2–), 4.21 (t, J = 6.6 Hz, 2H, –NCH2–), 2.27 (s, 3H), 1.84–1.79 (m, 2H), 1.68–1.62 (m, 2H), 1.41–1.36 (m, 2H), 1.31–1.26 (m, 2H). 13C RMN (75 MHz, DMSO-d6): δ 169.59 (C=O), 164.54 (C=O), 158.53 and 158.38 (F–C–C=O), 150.39, 142.72 (triazolyl), 141.80 and 138.74 (F–C), 134.69, 131.65, 130.35 and 129.90 (CH–C–F), 126.78, 124.49 (triazolyl), 123.85, 123.68, 65.21, 49.76, 49.05, 43.23, 30.00, 28.36, 25.94, 25.22, 21.22. HRMS-ESI (m/z): calcd for C22H25FN5O6 [M + H]+: 474.1744, found: 474.1783
8-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)octyl 2-acetoxybenzoate (4f): Yield 70%, solid yellow, m.p. 90–94 °C; 1H RMN (300 MHz, DMSO-d6): δ 8.12 (d, J = 6.6 Hz, 1H; 5-FU-H), 8.08 (s, 1H; triazole-H), 7.93 (dd, J = 7.8, 1.7 Hz, 1H), 7.68 (td, J = 7.7, 1.6 Hz, 1H), 7.42 (td, J = 7.5, 1.1 Hz, 1H), 7.24 (dd, J = 8.1, 1.2 Hz, 1H), 4.88 (s, 2H; triazole-CH2–), 4.32 (t, J = 7.1 Hz, 2H, –OCH2–), 4.21 (t, J = 6.6 Hz, 2H, –NCH2–), 2.27 (s, 3H), 1.82–1.76 (m, 2H), 1.68–1.63 (m, 2H), 1.35–1.22 (m, 8H). 13C RMN (75 MHz, DMSO-d6): δ 169.58 (C=O), 164.55 (C=O), 158.71 and 157.62 (F–C–C=O), 150.39, 142.69 (triazolyl), 141.79 and 138.73 (F–C), 134.68, 131.62, 129.91 and 129.85 (CH–C–F), 126.78, 124.50 (triazolyl), 123.84, 123.71, 65.32, 49.81, 43.22, 30.07, 28.92, 28.71, 28.50, 26.22, 25.73, 21.23. HRMS-ESI (m/z): calcd for C24H29FN5O6 [M + H]+: 502.2057, found: 502.2092.
9-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)nonyl 2-acetoxybenzoate (4g): Yield 68%, solid yellow, m.p. 93–95 °C; 1H RMN (300 MHz, DMSO-d6): δ 8.15 (d, J = 6.7 Hz, 1H; 5-FU-H), 8.09 (s, 1H; triazole-H), 7.93 (dd, J = 7.8, 1.3 Hz, 1H), 7.68 (td, J = 7.7, 1.6 Hz, 1H), 7.41 (td, J = 7.6, 1.1 Hz, 1H), 7.24 (dd, J = 8.1, 1.2 Hz, 1H), 4.89 (s, 2H; triazole-CH2–), 4.31 (t, J = 7.2 Hz, 2H, –OCH2–), 4.22 (t, J = 6.6 Hz, 2H, –NCH2–), 2.27 (s, 3H), 1.81–1.76 (m, 2H), 1.68–1.63 (m, 2H), 1.37–1.20 (m, 10H). 13C RMN (75 MHz, DMSO-d6): δ 169.57 (C=O), 164.56 (C=O), 158.28 and 157.92 (F–C–C=O), 150.39, 149.97, 142.59 triazolyl), 141.68 and 138.64 (F–C), 134.68, 131.62, 130.57 and 130.12 (CH–C–F), 126.77, 124.50 (triazolyl), 123.85, 123.71, 65.34, 49.83, 43.22, 30.08, 29.18, 28.98, 28.72, 28.52, 26.24, 25.77, 21.22. HRMS-ESI (m/z): calcd for C25H31FN5O6 [M + H]+: 516.2214, found: 516.2242.
12-(4-((5-fluoro-2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)methyl)-1H-1,2,3-triazol-1-yl)dodecyl 2-acetoxybenzoate (4h): Yield 70%, solid yellow, m.p. 99–102 °C. 1H RMN (300 MHz, DMSO-d6): δ 8.13 (d, J = 6.6 Hz, 1H; 5-FU-H), 8.08 (s, 1H; triazole-H), 7.93 (dd, J = 7.8, 1.7 Hz, 1H), 7.68 (td, J = 7.8, 1.7 Hz, 1H), 7.41 (td, J = 7.6, 1.1 Hz, 1H), 7.24 (dd, J = 8.1, 1.2 Hz, 1H), 4.89 (s, 2H; triazole-CH2–), 4.31 (t, J = 7.1 Hz, 2H, –OCH2–), 4.22 (t, J = 6.6 Hz, 2H, –NCH2–), 2.27 (s, 3H), 1.80–1.75 (m, 2H), 1.68–1.64 (m, 2H), 1.36–1.21 (m, 16H). 13C RMN (75 MHz, DMSO-d6): δ 169.56 (C=O), 164.56 (C=O), 158.44 and 157.80 (F–C–C=O), 150.39, 142.62 (triazolyl), 141.72 and 138.06 (F–C), 134.68, 131.61, 130.03 and 129.84 (CH–C–F), 126.77, 124.51 (triazolyl), 123.86, 123.72, 65.35, 49.84, 43.21, 30.08, 29.35, 29.30, 29.08, 28.80, 28.53, 26.27, 25.80, 21.22. HRMS-ESI (m/z): calcd for C28H37FN5O6 [M + H]+: 558.2683, found: 558.2716.

3.2. Drug-Likeness Modelling

To investigate the drug-likeness characteristics, the SWISSADME online software was employed aiming at estimating seven crucial physicochemical and pharmacokinetic descriptors related to drug development and medicinal chemistry friendliness [21]. The properties analyzed include the number of rotatable bonds, donor–acceptor groups, topological polar surface area (TPSA), Lipinski’s rule of five, the lipophilicity index logPo/w (octanol/water participation coefficient) value, and the pan-assay interference compounds (PAINS) alerts.

4. Conclusions

In this work, the authors synthesized eight new 5-FU-ASA hybrids using, as a key step, Huisgen 1,3-dipolar cycloaddition, an example of click chemistry, with good yields. These compounds were characterized by NMR and high-resolution masses. Cytotoxicity studies of these compounds are underway. Concerning pharmacokinetic modelling, we have seen that the novel hybrids possess good drug-like attributes with similar values to those of the majority of marketed drugs, such as “a balanced” hydrophilic-lipophilic character, and do not inflict more than one violation of LRFV, making these compounds valid candidates in future anti-cancer drug-development projects.

Supplementary Materials

Supplementary data (1H, 13C NMR and MS spectra of compounds 4a4h) associated with this article can be found, in the online version. Figure S1a: 1H NMR of compound 4a, Figure S1b: 13C NMR of compound 4a and Figure S1c: MS spectra of compound 4a; Figure S2a: 1H NMR of compound 4b, Figure S2b: 13C NMR of compound 4b and Figure S2c: MS spectra of compound 4b; Figure S3a: 1H NMR of compound 4c, Figure S3b: 13C NMR of compound 4c and Figure S3c: MS spectra of compound 4c; Figure S4a: 1H NMR of compound 4d, Figure S4b: 13C NMR of compound 4d and Figure S4c: MS spectra of compound 4d; Figure S5a: 1H NMR of compound 4e, Figure S5b: 13C NMR of compound 4e and Figure S5c: MS spectra of compound 4e; Figure S6a: 1H NMR of compound 4f, Figure S6b: 13C NMR of compound 4f and Figure S6c: MS spectra of compound 4f; Figure S7a: 1H NMR of compound 4g, Figure S7b: 13C NMR of compound 4g and Figure S7c: MS spectra of compound 4g; Figure S8a: 1H NMR of compound 4h, Figure S8b: 13C NMR of compound 4h and Figure S8c: MS spectra of compound 4h.

Author Contributions

W.C.-L.: Synthesis and characterization of hybrid molecules. A.F.Y.: in silico studies, analysis, writing—original Draft. W.C.-G.: resources, supervision, project administration, funding acquisition, writing—original draft, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Antioquia and the Ministry of Science MINCIENCIAS, through the Program: NanoBioCáncer 2.0 GAT 2.0. Código: 121092092332, grant: 621-2022, project number 92355.

Data Availability Statement

The data presented in this study is available in this article and Supporting Information.

Acknowledgments

The authors thank University of Antioquia and MINCIENCIAS for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meunier, B. Hybrid molecules with a dual mode of action: Dream or reality? Acc. Chem. Res. 2008, 41, 69–77. [Google Scholar] [CrossRef] [PubMed]
  2. De Oliveira Pedrosa, M.; Duarte da Cruz, R.M.; de Oliveira Viana, J.; de Moura, R.O.; Ishiki, H.M.; Barbosa Filho, J.M.; Diniz, M.F.; Scotti, M.T.; Scotti, L.; Bezerra Mendonca, F. Hybrid Compounds as Direct Multitarget Ligands: A Review. Curr. Top. Med. Chem. 2017, 17, 1044–1079. [Google Scholar] [CrossRef] [PubMed]
  3. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. Available online: https://acsjournals.onlinelibrary.wiley.com/doi/10.3322/caac.21660 (accessed on 1 June 2021). [CrossRef]
  4. Alqahtani, Z.; Jamali, F. Clinical outcomes of aspirin interaction with other non-steroidal anti-inflammatory drugs: A systematic review. J. Pharm. Pharm. Sci. 2018, 21, 29854. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, W.; Yan, Y.; Chen, M.; Luo, G.; Hao, J.; Pan, J.; Hu, S.; Guo, P.; Li, W.; Wang, R.; et al. Aspirin enhances the sensitivity of colon cancer cells to cisplatin by abrogating the binding of NF-κB to the COX-2 promoter. Aging 2020, 12, 611–627. [Google Scholar] [CrossRef] [PubMed]
  6. Kaur, J.; Saxena, M.; Rishi, N. An Overview of Recent Advances in Biomedical Applica-tions of Click Chemistry. Bioconjug. Chem. 2021, 32, 1455–1471. [Google Scholar] [CrossRef] [PubMed]
  7. Li, X.; Xiong, Y. Application of “Click” Chemistry in Biomedical Hydrogels. ACS Omega 2022, 7, 36918–36928. [Google Scholar] [CrossRef]
  8. Kumar, A.; Kumar Yadav, A.; Mishra, V.; Kumar, D. Recent Advance-ments in Triazole-based Click Chemistry in Cancer Drug Discovery and Development. Syn. Open 2023, 7, 186–208. [Google Scholar]
  9. Moreno-Quintero, G.; Betancur-Zapata, E.; Herrera-Ramírez, A.; Cardona-Galeano, W. New Hybrid Scaffolds Based on 5-FU/Curcumin: Synthesis, Cytotoxic, Antiproliferative and Pro-Apoptotic Effect. Pharmaceutics 2023, 15, 1221. [Google Scholar] [CrossRef]
  10. Moreno-Quintero, G.; Castrillón-Lopez, W.; Herrera-Ramirez, A.; Yepes-Pérez, A.F.; Quintero-Saumeth, J.; Cardona-Galeano, W. Synthesis and Chemopreventive Potential of 5-FU/Genistein Hybrids on Colorectal Cancer Cells. Pharmaceuticals 2022, 15, 1299. [Google Scholar] [CrossRef]
  11. Gómez-R, L.; Moreno-Q, G.; Herrera-R, A.; Castrillón-L, W.; Yepes, A.F.; Cardona, G.W. New Hybrid Scaffolds Based on ASA/Genistein: Synthesis, Cytotoxic Effect, Molecular Docking, Drug-likeness and in silico ADME/tox Modeling. J. App. Pharm. Sci. 2022, 12, 15–30. [Google Scholar]
  12. Belkharchach, S.; Ighachane, H.; Lachgar, A.; Ait Ali, M.; Lazrek, H.B. Efficient and selective catalytic N-Alkylation of pyrimidine by ammonium Sulfate@Hydro-thermal carbone under eco-friendly conditions. J. Chem. Sci. 2020, 132, 138. [Google Scholar] [CrossRef]
  13. Kuan, H.; Xie, Y.; Guo, Y.; Gianoncelli, A.; Ribaudo, G.; Coghi, P. (2R, 4S, 5S) 1-(4-(4-(((7-Chloroquinolin-4-yl)amino)methyl)-1H-1,2,3-triazol-1-yl)-5-(hydroxymethyl)tetrahydrofuran-2-yl)-5-methylpyrimidine-2,4(1H,3H)-dione. Molbank 2023, 2023, M1681. [Google Scholar] [CrossRef]
  14. Karabulut, H.R.F.; Karatavuk, A.O.; Ozyildirim, H.; Doğanlar, O.; Doğanlar, Z.B. Synthesis of novel dimeric compounds containing triazole using click method and their selective antiproliferative and proapoptotic potential via mitochondrial apoptosis signaling. Med. Chem. Res. 2020, 29, 643–655. [Google Scholar] [CrossRef]
  15. Iwata, H. Application of in Silico Technologies for Drug Target Discovery and Pharmacokinetic Analysis. Chem. Pharm. Bull. 2023, 71, 398–405. [Google Scholar] [CrossRef]
  16. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef] [PubMed]
  17. Lipinski, C.A. Drug-like properties and the causes of poor solubility and poor permeability. J. Pharmacol. Toxicol. Methods 2000, 44, 235–249. [Google Scholar] [CrossRef]
  18. Ditzinger, F.; Price, D.J.; Ilie, A.R.; Köhl, N.J.; Jankovic, S.; Tsakiridou, G.; Aleandri, S.; Kalantzi, L.; Holm, R.; Nair, A.; et al. Lipophilicity and hydrophobicity considerations in bio-enabling oral formulations approaches—A PEARRL review. J. Pharm. Pharmacol. 2019, 71, 464–482. [Google Scholar] [CrossRef]
  19. Ertl, P.; Rohdem, B.; Selzer, P. Fast calculation of molecular polar surface area as a sum of fragment-based contributions and its application to the prediction of drug transport properties. J. Med. Chem. 2000, 43, 3714–3717. [Google Scholar] [CrossRef]
  20. Baell, J.B.; Nissink, J.W.M. Seven Year Itch: Pan-Assay Interference Compounds (PAINS) in 2017-Utility and Limitations. ACS Chem. Biol. 2018, 13, 36–44. [Google Scholar] [CrossRef]
  21. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A free web tool to evaluate pharmacokinetics, drug-likeness and medicinal chemistry friendliness of small molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Design of 5-FU-ASA hybrids.
Figure 1. Design of 5-FU-ASA hybrids.
Molbank 2023 m1745 g001
Scheme 1. Synthesis of 5-FU-ASA hybrids.
Scheme 1. Synthesis of 5-FU-ASA hybrids.
Molbank 2023 m1745 sch001
Table 1. Lipinski’s rule and pharmacokinetic score for the synthesized conjugates 4ah.
Table 1. Lipinski’s rule and pharmacokinetic score for the synthesized conjugates 4ah.
Comp. (n)MW aLogPo/w bR.B. cn-ON dn-OHNH eTPSA fLRFV gPAINS h
4a417.350.619101138.1900
4b431.380.8810101138.1900
4c445.411.1511101138.1900
4d459.431.6612101138.1900
4e473.462.1613101138.1900
4f501.513.1615101138.1910
4g515.543.6816101138.1910
4h557.625.1919101138.1910
a Molecular weight of the hybrid (150–500). b Parameter calculated as consensus log P for octanol/water (−2 to 6.5). c Number of rotatable bonds (0–10). d Estimated n-ON number of hydrogen bond acceptors <10. e n-OHNH number of hydrogen bond donors ≤5. f Topological polar surface area (7.0–200 Å2). g Lipinski’s rule of five violations. h Identification of potentially problematic fragments for pan-assay interference compounds (PAINS).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Castrillón-López, W.; Yepes, A.F.; Cardona-Galeano, W. Synthesis and In Silico Drug-Likeness Modeling of 5-FU/ASA Hybrids. Molbank 2023, 2023, M1745. https://doi.org/10.3390/M1745

AMA Style

Castrillón-López W, Yepes AF, Cardona-Galeano W. Synthesis and In Silico Drug-Likeness Modeling of 5-FU/ASA Hybrids. Molbank. 2023; 2023(4):M1745. https://doi.org/10.3390/M1745

Chicago/Turabian Style

Castrillón-López, Wilson, Andrés F. Yepes, and Wilson Cardona-Galeano. 2023. "Synthesis and In Silico Drug-Likeness Modeling of 5-FU/ASA Hybrids" Molbank 2023, no. 4: M1745. https://doi.org/10.3390/M1745

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