5.2
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original Article
2025
:18;
2412025
doi:
10.25259/AJC_241_2025

Iron (III) triflate catalyzed, one-pot three-component synthesis of novel α-aminophosphonate of 5-hydrazineyl-2-(phenylsulfonyl)pyridine as potent antiproliferative agents against MCF-7 and HCT-116: In vitro and in silico investigations on molecular docking and dynamics simulation

Department of Chemistry, College of Science, Taibah University, 30002 Al-Madinah Al-Munawarah, Saudi Arabia.

* Corresponding author: E-mail address: gjohani@taibahu.edu.sa (G. Aljohani)

Licence
This is an open-access article distributed under the terms of the Creative Commons Attribution-Non Commercial-Share Alike 4.0 License, which allows others to remix, transform, and build upon the work non-commercially, as long as the author is credited and the new creations are licensed under the identical terms.

Abstract

Cancer statistics over the past decade reveal a significant global burden, with increasing incidence and mortality rates. Among the most commonly diagnosed cancers are colorectal and breast cancer. The present work aims to design new drug-like compounds as potential agents against colorectal and breast cancer, using the Kabachnik–Fields reaction as a synthetic approach. α-Aminophosphonate phenylsulfonyl derivatives (4a–f) were synthesized via a one-pot, three-component reaction involving 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1), benzaldehyde derivatives (2a–f), and diethyl phosphite (3). Reaction conditions were optimized using various Lewis acids, with iron(III) trifluoromethanesulfonate identified as the most effective catalyst, affording excellent yields in a short time. The novel derivatives were characterized using FT-IR, 1H NMR, 13C NMR, 31P NMR, and mass spectrometry. The in vitro antiproliferative activity of the α-aminophosphonates was evaluated against two human cancer cell lines: breast (MCF-7) and colon (HCT-116), using Doxorubicin (DM2) as a standard and the MTT assay for comparison. Compounds 4b and 4e emerged as selectively potent antiproliferative candidates against the HCT-116 cell line, while all compounds showed significantly lower cytotoxic activity against the MCF-7 cell line. In silico molecular docking simulations of compounds 4b, 4e and DM2 with CDK2 (PDB: 8FP0) were performed using GNINA software. The results revealed that both compounds occupy the same region of DM2 and interact with the ATP-binding site, suggesting a potential to affect CDK2 function. Notably, compound 4e exhibited more extensive hydrophobic interactions, indicating a stronger binding affinity compared to 4b. Molecular dynamics simulations supported these findings, showing that compound 4b did not remain stably bound to CDK2 throughout the simulation period.

Keywords

α-aminophosphonate
Antiproliferative
Iron triflate
Kabachnik-fields reaction
Molecular docking
Molecular dynamics

1. Introduction

Cancer is indeed one of the most serious diseases worldwide, representing a significant global burden with increasing incidence and mortality rates. In 2022, nearly 20 million new cancer cases were reported, alongside approximately 9.7 million deaths, indicating a persistent rise in cancer prevalence [1]. By 2050, researchers estimate that there may be 35 million new cancer cases globally. The most commonly diagnosed cancers with the highest mortality include lung, breast, colorectal and lung cancer [1,2]. Consequently, the development of effective cancer treatments is a research priority, aimed at designing drugs with high efficacy and minimal drawbacks [3,4]. Most commercial anticancer drugs are associated with side effects and drug resistance—key challenges that can potentially be addressed using multi-target drugs (MTDs) [5,6]. This strategy involves incorporating two or more distinct pharmacophores from different medications into a single hybrid molecule, offering a promising approach for the treatment of complex diseases [7,8]. MTDs often demonstrate improved efficacy, enhanced safety profiles, and more convenient administration [9,10].

One class of compounds of interest is α-aminophosphonate derivatives, which are organic molecules featuring both an amino group (–NH₂), imparting basic and nucleophilic properties, and a phosphonate group (–P(O)(OR)₂) attached to the α-carbon. These compounds can be synthesized through various methods, including the Kabachnik–Fields reaction and the direct phosphorylation of amino acids [11-13]. Due to their structural features, α-aminophosphonates have shown potential in medicinal chemistry, with applications in antimicrobial [13,14], antiviral [15], anti-COVID-19 [16] and antitumor [17-19]. Their ability to interact with phosphate-dependent biological pathways, especially in cellular signaling, also suggests potential in neurological applications [20-22].

Similarly, phenylsulfonyl pyridine derivatives have demonstrated anticancer activity against a variety of cancers, including breast cancer, hepatocellular carcinoma, pancreatic cancer, myeloid leukemia, pulmonary carcinoma, and brain cancer [23-26]. The phenylsulfone moiety also exhibits a wide range of therapeutic properties, including anti-HBV, antimicrobial, and antiviral activities, [27,28].

Substructures are commonly found in commercial drugs, reflecting their versatility in medicinal chemistry. Pyridine derivatives are widely utilized for their diverse biological activities [29-31], while phosphonates contribute to therapeutic efficacy in specific applications [13]. Representative examples of drugs containing these substructures are illustrated in Figure 1.

Examples of drugs bearing pyridine, phosphonate substructure.
Figure 1.
Examples of drugs bearing pyridine, phosphonate substructure.

Iron is one of the most abundant, cost-effective, and low-toxicity metals in the Earth’s crust. It also offers unique chemical versatility, functioning as both a Lewis acid and a transition metal catalyst [32,33]. In recent years, there has been remarkable growth in the number of synthetic transformations catalyzed by iron complexes [34,35]. Lewis acids, such as metal triflates, exhibit tolerance to water—an advantage in reactions where water is generated as a byproduct and could otherwise deactivate or degrade the catalyst [36,37].

As highlighted in the literature, combining hydrazineyl, phenylsulfonyl, and α-aminophosphonate functionalities within a single molecular framework offers broad biological potential. These hybrid compounds represent promising candidates for therapeutic development across multiple fields, including antimicrobial therapy, cancer treatment, neuroprotection, and anti-inflammatory applications, as illustrated in Figure 2.

Structure of the target phenylsulfonyl-α-aminophosphonate.
Figure 2.
Structure of the target phenylsulfonyl-α-aminophosphonate.

Despite significant advances in cancer treatment, the development of effective therapies remains hindered by drug resistance, toxicity, and limited efficacy of single-target drugs. There is an urgent need for novel multi-target compounds that can overcome these limitations. In this study, we aim to design and synthesize new hybrid molecules combining hydrazineyl, phenylsulfonyl, and α-aminophosphonate pharmacophores, integrating pyridine and phosphonate moieties known for their broad biological activities. Utilizing environmentally friendly Lewis acid catalysis, we propose an innovative synthetic route starting from 5-chloro-2-(phenylsulfonyl)pyridine. Furthermore, we will evaluate their antiproliferative effects against cancer cell lines and employ molecular docking and dynamic simulations to elucidate their mechanism of action. This approach is expected to contribute new candidates for anticancer drug development with improved efficacy and safety profiles. The hybrid molecules were specifically developed based on a dual approach: (1) structural similarity to clinically approved anticancer drugs known to inhibit kinases, and (2) a targeted mode of action involving the inhibition of CDK2, a key regulator of cell cycle progression. This mechanism was supported by molecular docking and dynamics simulations, which confirmed favorable interactions of the designed compounds within CDK2’s active site.

2. Material and Methods

2.1. Materials

The melting point of new compounds was conducted on an Electrothermal melting point apparatus using an open capillary tube. A Shimadzu FT-IR 8400S spectrophotometer is used to determine IR peaks. Thin layer chromatography (TLC) is utilized to follow the reactions using 0.2 mm silica gel F254 plates (Merck) with ethyl acetate: petroleum ether elution. Bruker NMR spectrometer (300, 75, and 121.4 MHz for 1H, 13C, and 31P NMR, respectively) is used to characterize the chemical structures of the synthesized compounds by 1H, 13C, and 31P NMR. Chemical shifts are expressed in parts per million (ppm) using the deuterated DMSO peak as an internal reference. Mass spectra (EI-MS) were obtained with ISQ (Single Quadrupole MS, Thermo Scientific). The purity of all newly synthesized samples was achieved by column chromatography using Silica gel and ethyl acetate: petroleum ether as eluent. 5-Chloro-2-(phenylsulfonyl)pyridine, hydrazine 99%, aldehydes, diethylphopshite Iron(III) trifluoromethanesulfonate 90%, and other solvents were purchased from Sigma Aldrich. The chemical names given for the prepared compounds are according to the IUPAC system. The reported yields are based on pure materials isolated by column chromatography.

2.2. Synthesis methods

2.2.1. Synthesis of 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1)

5-Chloro-2-(phenylsulfonyl)pyridine (9.09 mmol) reacted with hydrazine 99% (36.4 mmol) in ethanol (40 mL) at reflux temperature for 5h; the reaction completion was monitored using TLC (Scheme 1). The reaction was cooled, evaporated and triturated with diethyl ether and the crude was purified using column chromatography with ethyl acetate/petroleum ether (30:70) to give compound 1 in 53% yield (1.2g), which is used in the next reaction directly.

Synthesis of 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1).
Scheme 1.
Synthesis of 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1).

2.2.2. Synthesis of α-aminophosphonates (4a-f)

A solution of compound 1 and diethyl phosphite (3) in 1,2-Dichloroethane (DCE) was added to the corresponding benzaldehyde compounds (2a-f). The reaction mixture was refluxed using Iron triflate Fe(OTf)3 for 2 h to produce α-aminophosphonates (4a-f).

2.2.2.1. Diethyl (phenyl(2-(6-(phenylsulfonyl)pyridin-3-yl)hydrazineyl)methyl)phosphonate (4a)

Compound 4a was obtained as pale yellow crystals, elution: EtOAc: petroleum ether (60-80°C)/(70/30); mp 250-252°C and yield 88%. 1H NMR (300 MHz, DMSO) δ 11.07 (br s, 1H, NH), 8.04 (s, 1H, CHpyridine.), 7.62 (d, J = 8.3 Hz, 1H, CHpyridine), 7.53-7.34 (m, 5H, CHarom.), 7.27-7.03 (m, 4H, CHarom.), 6.87 (d, J = 8.6 Hz, 1H, CHpyridine), 6.84 (d, J = 8.6 Hz, 1H, CHarom), 5.18-5.07 (dd, J = 20.1, 13.6 Hz, 1H, -CH-P=O), 4.06 – 3.81 (m, 4H, 2 OCH2), 1.11 (t, J = 6 Hz, 3H, CH3), 1.02 (t, J = 7.5 Hz, 3H, CH3).13C NMR (75 MHz, DMSO) δ 152.7, 151.2, 151.1, 146.1, 139.3, 136.5, 128.7, 128.6, 128.5, 127.9, 117.2, 112.8, 112.6, 62.9 (2×OCH2), 54.7 and 52.7 (CH-P), 16.7 (CH3). 31P NMR (121 MHz, DMSO) δ 20.7. MS (m/z): M+475.59 (10%).

2.2.2.2. Diethyl ((2-(6-(phenylsulfonyl)pyridin-3-yl)hydrazineyl)(p-tolyl)methyl)phosphonate (4b)

Compound 4b was obtained as pale yellow crystals, elution: EtOAc: petroleum ether (60-80°C)/(75/25); mp 160-162°C and yield 90%. IR (cm-1, ʋ): 3271 (NH), 1596, 1523 (C=N), 1228 (P=O), 1151 (SO2). 1H NMR (300 MHz, DMSO) δ 11.11 (br s, 1H, NH), 8.04 (s, 1H, CHpyridine), 7.62 (d, J = 8.3 Hz, 1H, CHpyridine), 7.55-7.34 (m, 4H, CHarom.), 7.22-7.05 (m, 4H, CHarom.), 6.87 (d, J = 8.6 Hz, 1H, CHpyridine), 6.83 (d, J = 8.6 Hz, 1H, CHarom.), 5.09 (dd, J = 20.1, 12.0 Hz, 1H, -CH-P=O), 4.03 – 3.43 (m, 4H, 2 OCH2), 2.72 (s, 1H, NH), 2.24 (s, 3H, PhCH3), 1.12 (t, J = 6 Hz, 3H, CH3), 1.03 (t, J = 7.5Hz, 3H, CH3).13C NMR (75 MHz, DMSO) δ 152.6, 151.1, 146.1, 139.2, 137.0, 133.2, 128.9, 128.5, 128.4 128.3, 127.7 117.1, 112.6, 62.6 (2×OCH2), 54.2 and 52.3 (-CH-P), 20.9 (PhCH3), 16.7 (2CH2CH3). 31P NMR (121 MHz, DMSO) δ 20.6. MS (m/z): M+ 488.07 (<5%).

2.2.2.3. Diethyl ((4-hydroxyphenyl)(2-(6-(phenylsulfonyl)pyridin-3-yl)hydrazineyl)methyl) phosphonate (4c)

Compound 4c was obtained as yellow crystals, elution: EtOAc: petroleum ether (60-80°C)/(65/35); mp 202-204°C and yield 79%. IR (cm-1, ʋ): 3390 (OH), 3271 (NH), 1626, 1589 (C=N), 1224 (P=O), 1128 (SO2). 1H NMR (300 MHz, DMSO) δ 11.07 (s, 1H, NH), 9.41 ((s, 1H, OH), 8.05 (s, 1H, CHpyridine.), 7.95 (d, J = 8.3 Hz, 1H, CHpyridine), 7.54-7.34 (m, 4H, CHarom.), 7.22-7.05 (m, 4H, CHarom.), 6.88 (d, J = 8.6 Hz, 1H, CHpyridine), 6.86 (d, J = 8.6 Hz, 1H, CHarom), 5.05-4.91 (dd, J = 24.0, 12.0 Hz, 1H, -CH-P=O), 4.00-3.67 (m, 4H, 2 CH2), 2.87 (s, 1H, NH), 1.13-1.00 (dt, J = 28.8, 15.0 Hz, 6H, 2 CH3).13C NMR (75 MHz, DMSO) δ 162.5 (s), 157.1 (s), 152.5 (s), 151.1 (d, J = 12.75 Hz), 146.1 (s), 139.1 (s), 129.6 (s), 129.5 – 128.5 (m), 127.5 (s), 117.1 (s), 112.6 (d, J = 5.25 Hz), 62.5 – 62.4 (CH2), 53.9, 51.9 (d, C-P, JC-P = 151.5 Hz ), 16.4 (CH3). 31P NMR (121 MHz, DMSO) δ 20.6.

2.2.2.4. Diethyl ((4-nitrophenyl)(2-(6-(phenylsulfonyl)pyridin-3-yl)hydrazineyl)methyl)phosphonate (4d)

Compound 4d was obtained as yellow crystals, elution: EtOAc: petroleum ether (60-80°C)/(50/50); mp 235-237°C and yield 75%. IR (cm-1, ʋ): 3269 (NH), 1636, 1595 (C=N), 1235 (P=O), 1150 (SO2). 1H NMR (300 MHz, DMSO) δ 11.04 (s, 1H, NH), 8.22 (s, 1H, CHpyridine.), 8.20 (d, J = 8.3 Hz, 1H, CHpyridine), 7.80-7.36 (m, 4H, CHarom.), 7.22-7.07 (m, 4H, CHarom.), 6.89 (d, J = 8.6 Hz, 1H, CHpyridine), 6.85 (d, J = 8.6 Hz, 1H, CHarom), 5.50-5.42 (dd, J = 24.0, 12.0 Hz, 1H, -CH-P=O), 4.04-3.81 (m, 4H, 2 CH2), 2.87 (s, 1H, NH), 1.15-1.04 (dt, J = 28.8, 15.0 Hz, 6H, 2 CH3).13C NMR (75 MHz, DMSO) δ 162.5 (s), 152.6 (s), 150.7 (d, J = 12.75 Hz), 147.2 (s), 144.6 (s), 139.2 (s), 129.6 (s), 128.5 – 123.4 (m), 116.9 (s), 112.7 (d, J = 5.25 Hz), 63.2 – 62.8 (CH2), 54.3, 52.3 (d, C-P, JC-P = 151.5 Hz ), 16.4 (CH3). 31P NMR (121 MHz, DMSO) δ 20.5 (s, P=O group).

2.2.2.5. Diethyl ((4-bromophenyl)(2-(6-(phenylsulfonyl)pyridin-3-yl)hydrazineyl)methyl)phosphonate (4e)

Compound 4e was obtained as yellow crystals, elution: EtOAc: petroleum ether (60-80°C)/(65/35); mp 215-217°C and yield 55%. IR (cm-1, ʋ): 3279 (NH), 1634, 1594 (C=N), 1231 (P=O), 1152 (SO2). 1H NMR (300 MHz, DMSO) δ 11.08 (s, 1H, NH), 8.00 (d, J = 8.3 Hz, 1H, CHpyridine), 7.57 (s, 1H, CHpyridine.), 7.54-7.41 (m, 4H, CHarom.), 7.03-6.82 (m, 4H, CHarom.), 6.81 (d, J = 8.6 Hz, 1H, CHpyridine), 6.78 (d, J = 8.6 Hz, 1H, CHarom), 5.21-5.09 (dd, J = 24.0, 12.0 Hz, 1H, -CH-P=O), 3.99-3.72 (m, 4H, 2 CH2), 2.81 (s, 1H, NH), 1.09-0.98 (dt, J = 28.8, 15.0 Hz, 6H, 2 CH3).13C NMR (75 MHz, DMSO) δ 152.7 (s), 150.1 (d, J = 12.75 Hz), 141.1 (s), 139.2 (s), 136.7 (s), 131.3 (s), 130.7 – 121.1 (m), 117.3 (s), 112.7 (d, J = 5.25 Hz), 63.0 – 62.7 (CH2), 54.0, 52.0 (d, C-P, JC-P = 151.5 Hz ), 16.5 (CH3). 31P NMR (121 MHz, DMSO) δ 20.96 (s, P=O group). MS (m/z): M+, M+1 554.78, 553.00 (25%).

2.2.2.6. Diethyl ((4-fluorophenyl)(2-(6-(phenylsulfonyl)pyridin-3-yl)hydrazineyl)methyl)phosphonate (4f)

Compound 4f was obtained as yellow crystals, elution: EtOAc: petroleum ether (60-80°C)/(77/23); mp 198-200°C and yield 87%. IR (cm-1, ʋ): 3283 (NH), 1594 (C=N), 1229 (P=O), 1153 (SO2). 1H NMR (300 MHz, DMSO) δ 11.14 (s, 1H, NH), 8.05 (d, J = 8.3 Hz, 1H, CHpyridine), 7.95 (s, 1H, CHpyridine.), 7.64-7.53 (m, 4H, CHarom.), 7.29-6.89 (m, 4H, CHarom.), 6.86 (d, J = 8.6 Hz, 1H, CHpyridine), 6.83 (d, J = 8.6 Hz, 1H, CHarom), 5.26-5.15 (dd, J = 24.0, 12.0 Hz, 1H, -CH-P=O), 4.03-3.72 (m, 4H, 2 CH2), 2.87 (s, 1H, NH), 1.13-1.00 (dt, J = 28.8, 15.0 Hz, 6H, 2 CH3).13C NMR (75 MHz, DMSO) δ 163.4 (s), 160.1 (s), 152.6 (s), 150.0 (d, J = 12.75 Hz), 146.6 (s), 139.1 (s), 132.6 (s), 130.5 (s), 130.4 – 127.9 (m), 117.0 (s), 112.6 (d, J = 5.25 Hz), 62.8 – 62.6 (CH2), 53.7-51.7 (d, C-P, JC-P = 151.5 Hz ), 16.4 (CH3). 31P NMR (121 MHz, DMSO) δ 21.92 (s, P=O group). MS (m/z): M+, 493.98 (20%).

2.3. In silico methods

2.3.1. Ligands and protein preparation

The molecular structure of compounds 4b/4e was accurately sketched using ChemDraw Ultra software. The initial structural representation, provided in Simplified Molecular Input Line Entry System (SMILES) notation, was subsequently converted to the Protein Data Bank (PDB) format. This conversion was facilitated by a custom Python script employing the RDKit cheminformatics toolkit, an open-source library for molecular manipulation and analysis (https://www.rdkit.org/).

The structures of MD2 and CDK2 was acquired from the RCSB, a curated archive of experimentally determined three-dimensional structures of biological macromolecules (https://www.rcsb.org/ligand/DM2/https://www.rcsb.org/structure/8FP0). The retrieved CDK2 structure (PDB ID: 8FP0) was refined using UCSF Chimera, a molecular visualization and analysis system. This refinement process involved the removal of extraneous polypeptide chains and bound ligands to isolate the target protein. Subsequently, any unresolved regions within the CDK2 structure were modeled using the loop modeling functionality within MODELLER, a comparative protein structure modeling program [38]. Finally, to optimize the refined CDK2 structure, energy minimization was performed using the CHARMM force field, implemented within the OpenMM molecular dynamics simulation toolkit. This minimization protocol consisted of 100 steps.

2.3.2. Molecular docking

The interaction between 4b/4e and DM2 chemical structures and the CDK2 protein was gauged using a computational docking approach. Molecular docking was performed with GNINA, a docking program that utilizes a convolutional neural network (CNN) ensemble to improve prediction accuracy [39]. A blind docking protocol was employed, where the search space, defined by a grid box, encompassed the entire CDK2 protein structure. Ten ligand poses in PDBQT format were generated, and the pose exhibiting the highest binding affinity was selected for further analysis.

To characterize the CDK2-ligand interactions, PyMOL was used to generate complex files representing the top-ranked pose. The Protein-Ligand Interaction Profiler (PLIP) web server (https://plip-tool.biotec.tu-dresden.de) was then used to analyze these complex files and identify key binding residues within the protein-ligand complexes. Complementary 2D interaction diagrams were generated using PoseView (https://proteins.plus/) to visualize the specific interactions. Furthermore, functional domains of CDK2 were identified using the InterPro database (https://www.ebi.ac.uk/interpro/), providing context for the observed interactions based on known protein family and domain information.

2.3.3. Molecular dynamic simulations

Molecular dynamics (MD) simulations were conducted using Maestro version 12.0.012 (Schrödinger, LLC, New York, NY). The CDK2-ligand complex was prepared using the Protein Preparation Wizard, which included preprocessing steps, energy minimization, and removal of crystallographic water molecules. The system was solvated using the simple point charge (SPC) water model, and counterions were added to neutralize the charge. The MD simulations were performed for 100 ns under constant temperature conditions (300 K).

Post-simulation trajectory analysis was performed to evaluate the stability and dynamics of the complex. Root mean square deviation (RMSD) and root mean square fluctuation (RMSF) were calculated using the interaction diagram module. Additionally, the radius of gyration (rGyr), polar surface area (PSA), and solvent-accessible surface area (SASA) were determined. A trajectory-based 3D structure analysis was conducted over the entire simulation period using the trj2mae.py utility script within Maestro.

Binding free energies were calculated using the Prime module. The calculations were performed on every 100th trajectory frame, providing a decomposition of the binding free energy into van der Waals, electrostatic, polar solvation, non-polar solvation, and entropic contributions. This decomposition allowed for a detailed understanding of the energetic factors driving ligand binding.

Principal Component Analysis (PCA) was carried out using the bio3d package in R to analyze the dominant modes of motion within the simulated complexes. The trajectory files were first converted from .xtc to .dcd format. The analysis focused on the C-alpha atoms, and their coordinates were extracted from the trajectory. The pca.xyz function, employing the singular value decomposition (SVD) method, was applied to these coordinates. The first two principal components (PCs), capturing the largest variance in the data, were used to represent the essential dynamics and conformational changes, providing insights into the complexes’ stability and flexibility.

Dynamic Cross-Correlation Matrix (DCCM) analysis was employed to investigate the correlated motions of Cα atoms within the protein-ligand complex using the bio3d package in R. The analysis, performed on the 0-100 ns trajectory, calculated the cross-correlation coefficient (Mxy) between the displacement vectors of all Cα atom pairs. The resulting cross-correlation values, ranging from -1 (perfect anti-correlation) to +1 (perfect correlation), with 0 indicating no correlation, were visualized as a heatmap. This allowed for identifying regions exhibiting correlated or anti-correlated motions, providing insights into the internal dynamics of the complex.

3. Results and Discussion

3.1. Chemistry

5-Chloro-2-(phenylsulfonyl)pyridine reacted with hydrazine (99%) in ethanol at reflux temperature to afford 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1), which is used directly in the synthesis of α-aminophosphonate derivatives.

The one-pot three-component Kabachnik-Fields reaction comprising 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1), benzaldehyde (2a), and diethyl phosphite (3) was run as a model reaction for optimization. The model reaction was studied in the presence of available Lewis acids as catalysts, solvents and temperature, as presented in Table 1, Scheme 2.

Optimization condition of α-aminophosphonate derivatives 4a.
Scheme 2.
Optimization condition of α-aminophosphonate derivatives 4a.
Table 1. Synthesis of diethyl (phenyl(2-(6-(phenylsulfonyl)pyridin-3-yl)hydrazineyl)methyl)phosphonate (4a) under optimization conditions.
Entry Catalyst mol % Solvent Reaction conditions Time Isolated yield
1 - - DCM r.t. overnight 5%
2 - - DCM Reflux 10h 15%
3 FeCl3 5 DCM Reflux 10h 10%
4 FeCl3.6H2O 5 DCM Reflux 15h 10%
5 Fe(OTf)2 5 DCM Reflux 3h 35%
6 Fe(OTf)3 5 DCM Reflux 1h 60%
7 Fe(OTf)3 5 DCE Reflux 2h 90%
8 Bi(OTf)3 5 DCM Reflux 5h 55%
9 Bi(OTf)3 5 DCE Reflux 1h 65%

First, 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1), benzaldehyde (2a), and diethyl phosphite (3) were reacted without a catalyst in dichloromethane (DCM) at room temperature overnight (entry 1). The resulting compound 4a was obtained with a low yield (5%), while its yield increased three times using the reflux condition (entry 2).

On the other hand, compound 4a is afforded in 10% yield using a catalytic amount of iron chloride in DCM at reflux temperature for 10h (entry 3). Notably, using FeCl3.6H2O as a catalyst does not alter the product yield (entry 4). The yield percentage increased to 35% when Fe(OTf)2 was used as a catalyst in DCM for three hours (entry 5).

Iron(III) trifluoromethanesulfonate is used in DCM at reflux for one hour to obtain the target compound 4a in 60% yield (entry 6). Interestingly, the most abundant yield of 4a (90%) is afforded using Fe(OTf)3 in dichloroethane (DCE) under reflux conditions (entry 7). Replacement of Fe(OTf)3 with Bi(OTf)3 affords the desired product in 55% yield in DCM (entry 8) and 65% yield (entry 9) at reflux temperature for five hours and one hour, respectively.

These findings suggest that using Iron triflate Fe(OTf)3 in DCE under reflux conditions was the most effective condition to be adopted to achieve α-aminophosphonate derivatives (entry 7). Hence, this optimum condition was applied to one-pot three-component Kabachnik-Fields reaction of 5-hydrazineyl-2-(phenylsulfonyl)pyridine (1), diethyl phosphite (3) and p-substituted benzaldehyde 2b-f to get 4b-f (Scheme 3)

Synthesis of the target α-aminophosphonate derivatives 4b-f.
Scheme 3.
Synthesis of the target α-aminophosphonate derivatives 4b-f.

Spectroscopic analyses of 1H, 13C, 31P NMR, and mass spectroscopy confirmed the structures of the resulting compounds (see the supplementary file). Considering compound 4b as an example, it was obtained as a pale-yellow crystal with a yield of 90%. The 1H NMR spectrum of 4b revealed the presence of a broad singlet signal at δH 11.11 ppm related to the NH proton. The aromatic phenyl and pyridine ring protons were observed on their expected chemical shift between δH 8.05–6.80. Interestingly, the spectrum showed an ABX spin system at δH5.09 (dd, J = 20.1, 12.0 Hz, 1H) attributed to the methine proton adjacent to the phosphonate (-P(O)(OEt)2) group. This diagnostic signal is due to the coupling of the methine proton with the P atom (2JH-P = 20.1 Hz) and the adjacent NH (3JH-H = 12.0 Hz). Consequently, the Kabachnik-Fields Reaction is confirmed as this pattern is consistent with numerous α-phosphonohydrazine derivatives described in the literature [40]. Moreover, the presence of ethyl groups of phosphate ester was confirmed due to the presence of a multiplet signal at the upfield between δH 4.03 and 3.43 (correlated to two methylene protons) and two triplets signals (associated with two methyl protons) at δH1.12 and 1.03, respectively. A singlet signal integrating for three protons at δH 2.24 confirms the presence of methyl group on the phenyl ring.

supplementary file

The 13C NMR spectrum showed 13 signals at the upfield between ẟC 152.6 and 112.6 of carbon signals corresponding to aromatic and pyridine carbon. This is complemented with the 4b structure due to the symmetrical feature in the two phenyl rings. Moreover, the disappearance of carbonyl carbon of benzaldehyde confirms the condensation reaction. Besides, the spectrum disclosed the presence of four signals resonated between ẟC 62.8 and 62.5 related to (2 OCH2) methylene carbons due to the coupling of methylene carbons with the P atom. Besides, the methine carbon (-CH-P) was observed as two signals at ẟC 54.2 and 52.3 (d, C-P, JC-P = 151.5 Hz). Two signals of methyl carbons were detected upfield at ẟC 20.9 and 16.7, which were assigned to (PhCH3) and (2CH2CH3), respectively. The elucidation of 4b was confirmed by its 31P NMR spectrum, which exhibited a singlet signal at 20.62 ppm. Moreover, the mass spectral data agree with the molecular formula of 4b.

3.2. Biological evaluation

3.2.1. In-vitro cytotoxic activity

The resulted α-aminophosphonate phenylsulfonyl derivatives 4a-f were evaluated as antiproliferative agents to study their effects on tumor cells using the 3-[4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) assay against two types of human cancer cell lines breast (MCF-7) and colon (HCT-116), which is based on the cleavage of the tetrazolium salt by mitochondrial dehydrogenases in viable cells [18,41]. The percentages of intact cells were calculated and compared to those of the control activities of these compounds against the two carcinoma cell lines in comparison with DM2 as a standard drug. While DM2 primarily exerts its anticancer effects through topoisomerase inhibition, it was selected as a reference due to its well-established efficacy in comparison to the newly designed hybrid agents. Although the chemical structure of DM2 differs from our hybrid compounds, its broad-spectrum anticancer activity serves as a useful benchmark for evaluating the antiproliferative potential of our compounds [42,43].

It is apparent from Figure 3 and Table 2 that all the investigated compounds showed no significant cytotoxic activity against MCF-7 human breast cancer cells when compared to DM2.

Dose-dependent cytotoxic activities of compounds (4a-4f) against MCF-7 cancer cells according to the MTT assay.
Figure 3.
Dose-dependent cytotoxic activities of compounds (4a-4f) against MCF-7 cancer cells according to the MTT assay.
Table 2. IC50 of the synthesized compounds against MCF-7 and HCT-116 cancer cell lines according to the MTT assay.
Compound code IC50 (µM) ± SD
MCF-7 HCT-116
4a 63.4 ± 4.5 21.1 ± 3.6
4b 57.3 ± 3.9 9.5 ± 1.5
4c 58.8 ± 5.5 15.0 ± 2.6
4d 70.6 ± 5.8 31.3 ± 4.1
4e 62.9 ± 5.6 9.2 ± 1.4
4f 67.9 ± 5.1 28.3 ± 4.1
DM2 26.9 ± 3.9 9.4 ± 3.9

It is interesting to note that the compounds (4e) and (4b) were the most potent cytotoxic compounds against HCT-116 cell lines comparable to DM2 with IC50 9.2 and 9.5 µg/ml, respectively, as illustrated in Figure 4 and Table 2.

Dose-dependent cytotoxic activities of compounds (4a-4f) against HCT-116 cancer cells according to the MTT assay.
Figure 4.
Dose-dependent cytotoxic activities of compounds (4a-4f) against HCT-116 cancer cells according to the MTT assay.

Structure-Activity Relationship (SAR) for In-Vitro Cytotoxic Activity:

The SAR analysis of the α-aminophosphonate phenylsulfonyl derivatives (4a-f) provides insights into the structural features responsible for the observed cytotoxic activity against human cancer cell lines, particularly the breast cancer (MCF-7) and colon cancer (HCT-116) cell lines.

The compounds 4a-f displayed varying cytotoxicity, but none were as effective against MCF-7 breast cancer cells when compared to DM2. This suggests that the presence of para substituted phenyl group in the hybrid compounds may not significantly influence activity against this particular cell line.

Compounds 4e and 4b emerged as the most potent cytotoxic agents against the HCT-116 colon cancer cell line, showing IC50 values of 9.2 µg/ml and 9.5 µg/ml, respectively. These values were comparable to the cytotoxicity of DM2.

The differences in cytotoxicity may be contributed to the presence of one of the following:

The presence of a phenylsulfonyl group in the structure of the compounds could enhance solubility and facilitate interactions with the cell membrane, improving cellular uptake [44,45].

The differences in cytotoxicity between compounds could be attributed to the substitutions on the aromatic rings. For example, Br in compound 4e or methyl in 4b could influence interactions with cellular targets or modify the compound’s stability.

The ability of these compounds to engage in hydrophobic interactions within the cancer cell environment may also be critical for their efficacy, particularly in the context of the breast cancer cells, where membrane penetration and cellular interactions are important for the drug’s effectiveness [46].

3.2.2. In-Silico investigations

3.2.2.1. Molecular docking

The prospective candidates 4e, 4b and the reference drug DM2 were evaluated through in silico docking studies to gain further insight into the binding patterns of the target ligands with cyclin-dependent kinase 2 (CDK2), whose dysregulation is linked to various cancers, making it a significant target for therapeutic interventions [47,48]. CDK2 regulates the cell cycle, particularly the transition from the G1 to the S phase [49]. CDK2’s involvement in cancer is multifaceted, influencing cell proliferation, drug resistance, and the development of targeted therapies [50,51]. Nemours research reported that alterations in CDK2 signaling pathways are associated with cancers such as colon, breast, ovarian, and prostate cancer [52-54].

The experimentally obtained CDK2 structure was docked against 4b /4e and the reference drug DM2 using the GNINA, yielding binding affinity scores of -8.6 kcal/mol, -8.4 kcal/mol and -9.3 kcal/mol. Based on the molecular docking of DM2 with CDK2, we observed that the chosen drug binds to the same region as 4e and 4b within minimal differences in binding residues as depicted in Figures 5 and 6.

The shared binding pocket of the protein CDK2 (PDB ID: 8FP0) is visualized with the docked poses of the three ligands: 4e (yellow), 4b (green), and DM2 (pink).
Figure 5.
The shared binding pocket of the protein CDK2 (PDB ID: 8FP0) is visualized with the docked poses of the three ligands: 4e (yellow), 4b (green), and DM2 (pink).
Interaction analysis of 4b/4e and DM2 with CDK2. (a) Visualization of CDK2 domains and structure. The 2D interaction diagrams of (b) 4e, (c) 4b and (d) DM2 with CDK2 residues highlights hydrophobic interactions (green) and hydrogen bonds (purple arrows).
Figure 6.
Interaction analysis of 4b/4e and DM2 with CDK2. (a) Visualization of CDK2 domains and structure. The 2D interaction diagrams of (b) 4e, (c) 4b and (d) DM2 with CDK2 residues highlights hydrophobic interactions (green) and hydrogen bonds (purple arrows).

Upon docking the two compounds, 4b and 4e, it was revealed that both compounds engage in hydrophobic interactions and hydrogen bonding within the protein structure. 4e interacts hydrophobically with ILE (10), VAL (18), ALA (31), PHE (82), LYS (88), LYS (89), GLN (131), LEU (134), and GLU (162). Additionally, it forms hydrogen bonds with GLU (12), GLN (131), and ASN (132). In contrast, 4b interacts hydrophobically with ILE (10), VAL (18), ALA (31), GLN (131), and LEU (134), thereby showing less hydrophobic interactions, while forming hydrogen bonds with GLU (12), GLN (131), and ASN (132), similar to 4e.

Critically, both compounds interact with GLN (131), a residue within the active site (residues 123-135). Additionally, ILE (10), GLU (12) and VAL (18), located in the ATP-binding site, participate in hydrophobic interactions with both compounds. However, 4e establishes additional hydrophobic contacts with PHE (82), LYS (88), LYS (89), and GLU (162), interactions absent in 4b binding. Furthermore, both compounds form hydrogen bonds with GLU (12) in the ATP-binding region and ASN (132) near the C-terminal of the active site.

Both 4b and 4e interact with CDK2’s ATP-binding and active sites, implying they could affect CDK2 function. However, 4e’s more extensive hydrophobic interactions suggest a more substantial binding efficacy than 4b. The 2D interactions between CDK2- 4e, CDK2- 4b and CDK2- DM2 are shown in Figures 6(b-d).

3.2.3. Molecular dynamics simulation

The conformational stability of the CDK2- 4e complex was assessed via molecular dynamics simulations. The complex’s root-mean-square deviation (RMSD) remained within a narrow range of 2.0 Å to 3.5 Å throughout the simulation period, indicating a generally stable conformation. No abrupt conformational changes were observed. Following an initial equilibration period of approximately 4 ns, during which the RMSD stabilized around 2.5 Å, the complex maintained high stability, exhibiting only minor fluctuations. A gradual decrease in RMSD from 3.5 Å to 3.1 Å was noted between 60 ns and the end of the simulation. In comparison, the CDK2- 4b complex exhibited similar conformational RMSD values to CDK2- 4e until approximately 40 ns. However, after this point, the RMSD of CDK2- 4b gradually increased, diverging significantly from that of CDK2- 4e. By the end of the simulation, the RMSD of CDK2- 4b was approximately 1 Å higher than that of CDK2- 4e. This difference suggests that, while both complexes show initial stability, CDK2- 4b exhibits signs of minor instability and deviates from its initial conformation over approximately half of the simulation period. RMSD visualizations are shown in Figure 7(a).

Conformational and interaction analysis of CDK2 bound to 4b and 4e. (a) RMSD trajectories over time for CDK2-4b and CDK2-4e complexes. (b) RMSF analysis showing fluctuations at key residues. (c) SASA of the ligands throughout the simulation. (d) Ligand RMSD (e-f) Ligand stability as observed from the Rg and PSA plots. (g) Time-series analysis of protein-ligand contact frequencies with key interacting residues identified. (h) Residual interaction plots of 4b and 4e with CDK2.
Figure 7.
Conformational and interaction analysis of CDK2 bound to 4b and 4e. (a) RMSD trajectories over time for CDK2-4b and CDK2-4e complexes. (b) RMSF analysis showing fluctuations at key residues. (c) SASA of the ligands throughout the simulation. (d) Ligand RMSD (e-f) Ligand stability as observed from the Rg and PSA plots. (g) Time-series analysis of protein-ligand contact frequencies with key interacting residues identified. (h) Residual interaction plots of 4b and 4e with CDK2.

The RMSF analysis revealed that both complexes have similar fluctuations at key residue regions of the protein, such as around the 45th amino acid and the C-terminal. However, CDK2- 4b was observed to have higher fluctuations on the N-terminal when compared to CDK2- 4e. Similarly, several minor differential fluctuations were observed between 150-250 amino acid regions, Figure 7(b).

In a contrasting observation, it was revealed that both ligands show a different accessibility to the solvent through SASA analysis, especially after the 25th ns CDK2- 4b, whereby gradually increased from 350 Å to ∼800.0 Å whereas within the same timeframe, CDK2- 4e gradually decreased till 80th ns to 250 Å before going to around 400 Å at the end of the simulation, as shown in Figure 7(c,d). Interestingly, both ligands showed a similar radius of gyration (∼5.0 Å) and polar surface area (∼145-155.0 Å) were nearly identical, Figure 7(e,f).

Interestingly, the ligand 4e demonstrated highly significant stability while showing a sudden peak at the end of the simulation period, around 90-100 ns. The observed high energy elevation between 90-100 ns appears to stem from a sudden alteration in the protein residues interacting with the ligand. This involves a transition to using VAL-18, HIS-84, and LEU-83 more prominently, while interactions with GLY-11 and LYS-9, consistent throughout the simulation, become less significant. Interestingly, 3D trajectory analysis indicates the ligand remains securely in its binding pocket despite this energetic fluctuation. Despite this peak, the ligand remained relatively stable overall. In contrast, 4b exhibited marked instability, reaching peaks of 600.0 Å before settling back to ∼35 Å by the end of the simulation. Even then, the ligand continued to fluctuate substantially throughout the simulation, suggesting a high instability of the ligand when bound to CDK2, indicating a potential imbalance in interactions, as illustrated in Figure 7(d).

Protein-ligand contact time-series analysis revealed consistent interactions between the ligands and several key residues. The most prominent interacting residues for 4b throughout the simulation included 15-TYR, 33-LYS, 132-ASN, 131-GLN, 165-THR, 14-THR, 6-LYS, 9-LYS, 10-ILE and 16-GLY, which are comparatively more when compared to 4e which included 12-GLU, 20-LYS, 10-ILE, 131-GLN and 33-LYS. Interestingly, 12-GLU showed the highest interaction factor of 56%, followed by 33-LYS, both of which are in the protein’s active site. In CDK2- 4b, residue 15-TYR had the most prominent interaction at 27.83%, followed by 33-LYS (16.87%), 132-ASN(13.55%), 131-GLN (9.98%), 16-GLY (9.85%), 10-ILE (8.99%), 9-LYS(4.68%), 6-LYS(3.57%) 14-THR (2.96%), and 165-THR (1.72%). In CDK2- 4e, other notable interactions included 131-GLN (11.12%), 10-ILE (5.61%), and 20-LYS (2.13%), shown in Figure 7(g-h).

Additional DCCM analysis revealed the correlated motions of residues in CDK2 when bound to 4b and 4e. The overall similarity in the correlation patterns for both ligands suggests that the global dynamics of CDK2 remain relatively consistent regardless of the ligand bound. However, subtle differences in the strength and extent of these correlations may indicate ligand-specific effects on the local flexibility of certain protein regions, potentially influencing CDK2’s functional state, Figure 8(a) and (b).

Global dynamics and binding analysis of CDK2 with 4b and 4e. (a) DCCM analysis shows correlated motions of residues for both complexes. (b) PCA scatter plots illustrate the conformational changes and flexibility of CDK2 upon binding each ligand. (c-e) MMGBSA analysis comparing the binding free energy profiles of the two complexes, highlighting stability shifts and binding affinities over time, indicating that 4b does not stay bound to the CDK2 protein over the simulation period.
Figure 8.
Global dynamics and binding analysis of CDK2 with 4b and 4e. (a) DCCM analysis shows correlated motions of residues for both complexes. (b) PCA scatter plots illustrate the conformational changes and flexibility of CDK2 upon binding each ligand. (c-e) MMGBSA analysis comparing the binding free energy profiles of the two complexes, highlighting stability shifts and binding affinities over time, indicating that 4b does not stay bound to the CDK2 protein over the simulation period.

PCA provides insight into the major conformational changes experienced by CDK2 in the presence of each ligand. The scatter plots show the distribution of simulation snapshots along the first two principal components (PC1 and PC2), representing the system’s dominant motions. The clustering and spread of these points indicate how the conformational landscape of CDK2 is modulated by each ligand. For CDK2 bound to 4b, the data points show a more defined clustering, suggesting that the protein explores a narrower range of conformations, implying a relatively stable binding mode. In contrast, for 4e, a wider spread of points is observed, which suggests greater conformational flexibility and possibly less stable binding. Greater conformational diversity may indicate that 4e allows CDK2 to sample a broader range of states, potentially influencing its enzymatic activity. If excessive flexibility is induced, it could lead to decreased binding affinity or altered functional dynamics of CDK2, Figure 8(c) and (d).

The MMGBSA analysis estimates the binding free energy for each ligand throughout the simulation. Lower (more negative) binding free energy values correspond to more potent and more favorable interactions between the ligand and CDK2. The CDK2- 4b complex exhibits a relatively stable binding energy profile, fluctuating around a more favorable range (∼ -35 to -25 kcal/mol); however, after the 70th ns, the ligand completely detaches from the protein, indicating a completely unstable binding and potentially meaningless use case of this ligand as an inhibitor for CDK2. In contrast, CDK2- 4e shows larger fluctuations and a less favorable average binding energy, as it starts at around -57 kcal/mol and then dips to -25 kcal/mol. It then gradually increases for the remainder of the simulation and reaches back to -45 kcal/mol. A differential shift around the same timeframe is observed where 4b loses all its interactions while 4e gains more free energy, indicating a stronger bound complex, Figure 8(e). Based on these observations, 4e serves as a potential inhibitor against CDK2 due to strong binding affinity and bound conformation at the protein’s active site.

4. Conclusions

In this study, a novel series of α-aminophosphonate derivatives incorporating hydrazineyl and phenylsulfonyl functionalities were successfully synthesized via a one-pot Kabachnik–Fields reaction catalyzed by iron triflate under environmentally friendly conditions. Structural elucidation was confirmed through comprehensive spectroscopic analyses. The antiproliferative activity of the synthesized compounds was evaluated against MCF-7 and HCT-116 cancer cell lines, revealing that compounds 4b and 4e demonstrated significant cytotoxicity toward colon cancer cells, with IC₅₀ values comparable to the standard drug DM2.

In silico docking studies indicated that both compounds interact with critical active ATP-binding sites of CDK2, a key regulator of the cell cycle and occupy the same region of DM2. Notably, compound 4e formed more extensive hydrophobic and hydrogen-bonding interactions than 4b, suggesting stronger binding affinity. Molecular dynamics simulations further supported these findings, with the CDK2–4e complex exhibiting higher structural stability and lower conformational fluctuation over time compared to the CDK2–4b complex.

Together, the experimental and computational results underscore the potential of these newly synthesized hybrid molecules, particularly compound 4e, as promising candidates for further development as anticancer agents targeting CDK2. Future work will focus on in vivo studies and optimization to enhance their therapeutic potential and selectivity.

Acknowledgment

The author extended her appreciation to Taibah University, represented by the Deanship of Scientific Research, for funding this project NO. (Rc-442/28).

CRediT authorship contribution statement

Ghadah Aljohani: Conceptualizing, Methodology, Data curation, Investigations, Software, Validation, Writing- Original draft, Reviewing and Editing.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Declaration of generative AI and AI-assisted technologies in the writing process

The authors confirm that there was no use of artificial intelligence (AI)-assisted technology for assisting in the writing or editing of the manuscript and no images were manipulated using AI.

Supplementary data

Supplementary material to this article can be found online at https://dx.doi.org/10.25259/AJC_241_2025.

References

  1. , , , , , , . Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA: A Cancer Journal for Clinicians. 2024;74:229-263. https://doi.org/10.3322/caac.21834
    [Google Scholar]
  2. , , , , , , , . Cancer statistics: The United States vs. worldwide. Journal of Clinical Oncology. 2024;42:e23276. https://doi.org/10.1200/jco.2024.42.16_suppl.e23276
    [Google Scholar]
  3. , , , . Comparison of anticancer drug toxicities: Paradigm shift in adverse effect profile. Life (Basel, Switzerland). 2021;12:48. https://doi.org/10.3390/life12010048
    [Google Scholar]
  4. , , , , . Potentials and future perspectives of multi-target drugs in cancer treatment: The next generation anti-cancer agents. Cell Communication and Signaling : CCS. 2024;22:228. https://doi.org/10.1186/s12964-024-01607-9
    [Google Scholar]
  5. , , , . Exploring the next generation of antibody-drug conjugates. Nature Reviews. Clinical Oncology. 2024;21:203-223. https://doi.org/10.1038/s41571-023-00850-2
    [Google Scholar]
  6. , , , . Drug repurposing and nanotechnology for topical skin cancer treatment: Redirecting toward targeted and synergistic antitumor effects. ACS Pharmacology & Translational Science. 2025;8:308-338. https://doi.org/10.1021/acsptsci.4c00679
    [Google Scholar]
  7. , , , . Designing multi-targeted agents: An emerging anticancer drug discovery paradigm. European Journal of Medicinal Chemistry. 2017;136:195-211. https://doi.org/10.1016/j.ejmech.2017.05.016
    [Google Scholar]
  8. , , , . What are the challenges with multi-targeted drug design for complex diseases? Expert Opinion on Drug Discovery. 2022;17:673-683. https://doi.org/10.1080/17460441.2022.2072827
    [Google Scholar]
  9. , ed. Polypharmacology: Strategies for multi-target drug Discovery. John Wiley & Sons; . ISBN: 978-1-394-18286-2
  10. , , . Computational multitarget drug design. Journal of Chemical Information and Modeling. 2017;57:403-412. https://doi.org/10.1021/acs.jcim.6b00491
    [Google Scholar]
  11. , , , . Recent developments in functional polymers via the kabachnik–fields reaction: The state of the art. Molecules. 2024;29:727. https://doi.org/10.3390/molecules29030727
    [Google Scholar]
  12. , , , , , . Asymmetric synthesis of tetrasubstituted α-aminophosphonic acid derivatives. Molecules. 2021;26:3202. https://doi.org/10.3390/molecules26113202
    [Google Scholar]
  13. , , . Remarkable potential of the α-aminophosphonate/phosphinate structural motif in medicinal chemistry. Journal of Medicinal Chemistry. 2011;54:5955-5980. https://doi.org/10.1021/jm200587f
    [Google Scholar]
  14. , , , . Synthesis of novel benzo[d] imidazole bearing α-aminophosphonate and their antimicrobial evaluation. Results in Chemistry. 2024;8:101586. https://doi.org/10.1016/j.rechem.2024.101586
    [Google Scholar]
  15. , , , , . Design, synthesis, docking, and antiviral evaluation of some novel pyrimidinone-based α-aminophosphonates as potent H1N1 and HCoV-229E inhibitors. Bioorganic Chemistry. 2024;147:107353. https://doi.org/10.1016/j.bioorg.2024.107353
    [Google Scholar]
  16. , , , , , , . Synthesis, DFT investigation, ADME-T properties, molecular docking and molecular dynamics simulation of new α-aminophosphonate inhibitor targeting Mpro and RdRp enzymes in SARS-CoV-2. Journal of Molecular Structure. 2024;1315:138842. https://doi.org/10.1016/j.molstruc.2024.138842
    [Google Scholar]
  17. , , , , , , , , , , , , . Design and synthesis of novel 20(S)-α-aminophosphonate derivatives of camptothecin as potent antitumor agents. Bioorganic Chemistry. 2021;114:105065. https://doi.org/10.1016/j.bioorg.2021.105065
    [Google Scholar]
  18. , , , , . Cytotoxic activity of α-aminophosphonic derivatives coming from the tandem kabachnik–fields reaction and acylation. Pharmaceuticals. 2023;16:506. https://doi.org/10.3390/PH16040506.
    [Google Scholar]
  19. . Design, synthesis, cytotoxicity and docking evaluation of novel α‐aminophosphonates quinoline thiohydrazineyl derivatives. ChemistrySelect. 2024;9:e202401180. https://doi.org/10.1002/slct.202401180
    [Google Scholar]
  20. , , , , , , , . Green approach to the synthesis of α-aminophosphonate-tetrahydroisoquinoline hybrids and their anti-cholinesterase activity. Bioorganic Chemistry. 2024;143:107008. https://doi.org/10.1016/j.bioorg.2023.107008
    [Google Scholar]
  21. , , , , , , , , , . Synthesis and anti-alzheimer potential of novel α-amino phosphonate derivatives and probing their molecular interaction mechanism with acetylcholinesterase. European Journal of Medicinal Chemistry. 2023;253:115288. https://doi.org/10.1016/j.ejmech.2023.115288
    [Google Scholar]
  22. , , , , , , , , . Hydroxyphenylamine phosphonate derivatives: Synthesis, X-ray crystallographic analysis, and evaluation of theirs anti-alzheimer effects and antioxidant activities. Journal of Molecular Structure. 2021;1225:129121. https://doi.org/10.1016/j.molstruc.2020.129121
    [Google Scholar]
  23. , , , . Microwaves assisted synthesis of antitumor agents of novel azoles, azines, and azoloazines pendant to phenyl sulfone moiety and molecular docking for VEGFR-2 kinase. Journal of Molecular Structure. 2022;1249:131657. https://doi.org/10.1016/j.molstruc.2021.131657
    [Google Scholar]
  24. , , , , , , . Discovery of 3,5-dimethyl-4-sulfonyl-1 h-pyrrole-based myeloid cell leukemia 1 inhibitors with high affinity, selectivity, and oral bioavailability. Journal of Medicinal Chemistry. 2021;64:11330-11353. https://doi.org/10.1021/acs.jmedchem.1c00682
    [Google Scholar]
  25. , , , , , , , . Synthesis and anticancer effects evaluation of 1-alkyl-3-(6-(2-methoxy-3-sulfonylaminopyridin-5-yl)benzo[d]thiazol-2-yl)urea as anticancer agents with low toxicity. Bioorganic & Medicinal Chemistry. 2015;23:6477-6485. https://doi.org/10.1016/j.bmc.2015.08.013
    [Google Scholar]
  26. , , , , , . N-phenylsulfonyl-3,5-bis(arylidene)-4-piperidone derivatives as activation NF-κB inhibitors in hepatic carcinoma cell lines. European Journal of Medicinal Chemistry. 2018;155:531-544. https://doi.org/10.1016/j.ejmech.2018.06.027
    [Google Scholar]
  27. , , , , , , , , , . Design and synthesis of pyridine-pyrazole-sulfonate derivatives as potential anti-HBV agents. MedChemComm. 2016;7:832-836. https://doi.org/10.1039/c6md00008h
    [Google Scholar]
  28. , . Pyridine compounds with antimicrobial and antiviral activities. International Journal of Molecular Sciences. 2022;23:5659. https://doi.org/10.3390/ijms23105659
    [Google Scholar]
  29. , , , , , , , , , . A decade of pyridine-containing heterocycles in US FDA approved drugs: A medicinal chemistry-based analysis. RSC Medicinal Chemistry. 2024;16:12-36. https://doi.org/10.1039/d4md00632a
    [Google Scholar]
  30. , , , . Nano-synthesis, characterization, theoretical, biological and catalytic studies of Mn(II), Cu(II), Cd(II), and Th(IV) complexes based on antipyrine azo dye ligand. Journal of Molecular Structure. 2025;1321:140273. https://doi.org/10.1016/j.molstruc.2024.140273
    [Google Scholar]
  31. , , , , . Novel nanoparticle‐sized Cu(II), Ru(III), and Th(IV) chelates derived from quinaldine azo dye ligand: DFT studies, molecular docking, and applications in drug development and catalysis. Applied Organometallic Chemistry. 2025;39:e7849. https://doi.org/10.1002/aoc.7849
    [Google Scholar]
  32. , , . Metal triflates as catalysts in organic synthesis: determination of their lewis acidity by mass spectrometry. Chem Plus Chem. 2022;87:e202200037. https://doi.org/10.1002/cplu.202200037
    [Google Scholar]
  33. , , , . Iron-catalyzed miyaura borylation of aryl chlorides and triflates. Organic Letters. 2025;27:197-201. https://doi.org/10.1021/acs.orglett.4c04171
    [Google Scholar]
  34. , , . Iron(III) triflate as a green catalyst for janus polymerization to prepare block polyesters. Macromolecules. 2023;56:7389-7395. https://doi.org/10.1021/acs.macromol.3c01359
    [Google Scholar]
  35. , , , , , , , . Iron triflate salts as highly active catalysts for the solvent‐free oxidation of cyclohexane. European Journal of Organic Chemistry. 2020;2020:3552-3559. https://doi.org/10.1002/ejoc.202000263
    [Google Scholar]
  36. , , , . Design, synthesis and biological evaluation of novel α-aminophosphonate oxadiazoles via optimized iron triflate catalyzed reaction as apoptotic inducers. European Journal of Medicinal Chemistry. 2019;180:310-320. https://doi.org/10.1016/j.ejmech.2019.07.029
    [Google Scholar]
  37. , , , , , . Microwave-assisted synthesis of novel sulfonamide-based compounds bearing α-aminophosphonate and their antimicrobial properties. Journal of Molecular Structure. 2022;1266:133553. https://doi.org/10.1016/j.molstruc.2022.133553
    [Google Scholar]
  38. , , , , , , , , , , . Tools for comparative protein structure modeling and analysis. Nucleic Acids Research. 2003;31:3375-3380. https://doi.org/10.1093/nar/gkg543
    [Google Scholar]
  39. , , , , , , , . GNINA 1.0: Molecular docking with deep learning. Journal of Cheminformatics. 2021;13:43. https://doi.org/10.1186/s13321-021-00522-2
    [Google Scholar]
  40. , , , . Synthesis of functionalised phosphonates by three-component reaction between phosphites, dialkyl acetylenedicarboxylates and pyrrole, indole or benzotriazole. Journal of Chemical Research. 2007;2007:291-295. https://doi.org/10.3184/030823407x215618
    [Google Scholar]
  41. , , , , , . A new class of anticancer activity with computational studies for a novel bioactive aminophosphonates based on pyrazole moiety. Scientific Reports. 2023;13:14680. https://doi.org/10.1038/s41598-023-40265-8
    [Google Scholar]
  42. , , , , , , . Doxorubicin: The good, the bad and the ugly effect. Current Medicinal Chemistry. 2009;16:3267-3285. https://doi.org/10.2174/092986709788803312
    [Google Scholar]
  43. , , , , , . Doxorubicin—An agent with multiple mechanisms of anticancer activity. Cells. 2023;12:659. https://doi.org/10.3390/cells12040659
    [Google Scholar]
  44. , , , , , , . Design and synthesis of phenyl sulfide-based cationic amphiphiles as membrane-targeting antimicrobial agents against gram-positive pathogens. Journal of Medicinal Chemistry. 2022;65:14221-14236. https://doi.org/10.1021/acs.jmedchem.2c01437
    [Google Scholar]
  45. , , . Theoretical insights into the uptake of sulfonamides onto phospholipid bilayers: Mechanisms, interaction and toxicity evaluation. Journal of Hazardous Materials. 2022;435:129033. https://doi.org/10.1016/j.jhazmat.2022.129033
    [Google Scholar]
  46. , , , , , , , , . The anticancer efficacy of paclitaxel liposomes modified with low-toxicity hydrophobic cell-penetrating peptides in breast cancer: An in vitro and in vivo evaluation. RSC Advances. 2018;8:24084-24093. https://doi.org/10.1039/c8ra03607a
    [Google Scholar]
  47. , , , . Cyclin-dependent kinase-2 as a target for cancer therapy: Progress in the development of CDK2 inhibitors as anti-cancer agents. Current Medicinal Chemistry. 2015;22:237-263. https://doi.org/10.2174/0929867321666141106113633
    [Google Scholar]
  48. , , , , , , . Targeting CDK2 in cancer: Challenges and opportunities for therapy. Drug Discovery Today. 2020;25:406-413. https://doi.org/10.1016/j.drudis.2019.12.001
    [Google Scholar]
  49. , , , , , , , , . Cyclin-dependent protein kinases and cell cycle regulation in biology and disease. Signal Transduction and Targeted Therapy. 2025;10:11. https://doi.org/10.1038/s41392-024-02080-z
    [Google Scholar]
  50. , , . Targeting CDK2 to combat drug resistance in cancer therapy. Future Oncology (London, England). 2024;20:3325-3341. https://doi.org/10.1080/14796694.2024.2416382
    [Google Scholar]
  51. , . Cdks, cyclins and CKIs: Roles beyond cell cycle regulation. Development (Cambridge, England). 2013;140:3079-3093. https://doi.org/10.1242/dev.091744
    [Google Scholar]
  52. , , , . Cyclin dependent kinase 2 (CDK2) inhibitors in oncology clinical trials: A review. Journal of Immunotherapy and Precision Oncology. 2025;8:47-54. https://doi.org/10.36401/JIPO-24-22
    [Google Scholar]
  53. , , . Cyclin-dependent kinase inhibitors and their therapeutic potential in colorectal cancer treatment. Frontiers in Pharmacology. 2021;12:757120. https://doi.org/10.3389/fphar.2021.757120
    [Google Scholar]
  54. , , , , , , , , , , , , . Cyclin-dependent kinase 2 (CDK2) inhibitors and others novel CDK inhibitors (CDKi) in breast cancer: Clinical trials, current impact, and future directions. Critical Reviews in Oncology/Hematology. 2024;196:104324. https://doi.org/10.1016/j.critrevonc.2024.104324
    [Google Scholar]
Show Sections