Translate this page into:
Computational design of quinazolin-4(3H)-one derivatives as multitarget inhibitors against angiogenesis and metabolic reprogramming to help overcome cancer resistance
* Corresponding authors: E-mail addresses: achraf.elallali@um6p.ma (A. El Allali), rachid.daoud@um6p.ma (R. Daoud)
-
Received: ,
Accepted: ,
Abstract
Dysfunction of receptor tyrosine kinases (RTKs) in angiogenesis significantly contributes to cancer progression. However, inhibiting primary angiogenesis pathways often leads to acquired resistance through various mechanisms, including alternative pathways. Targeting both angiogenesis and tumor energy metabolism, particularly via RTKs and phosphofructokinase-2/fructose-2,6-bisphosphatase 3 (PFKFB3), presents a promising strategy to slow tumor growth and propagation. The present study explored Quinazolin-4(3H)-one derivatives as kinase inhibitors using 3D-QSAR modeling to design more effective compounds. The developed CoMSIA/SHA model demonstrated high reliability, indicated by a determination coefficient (R2 = 0.995) and Leave-One-Out cross-validation coefficient (Q2 = 0.717). The model’s robustness and stability were further demonstrated by external validation (R2Pred = 0.832). Fifty-nine newly designed compounds with potent inhibitory activity were generated, with absorption, distribution, metabolism, excretion, and toxicity (ADMET) screening validating their favorable profiles. Compared to the most active molecule (molecule 39) in the dataset and the reference drug for each protein target, molecular docking revealed that four newly designed compounds exhibited better binding affinity (−8.5 to −11.5 kcal/mol) toward RTKs and PFKFB3. Molecular dynamics simulations confirmed the stability of these compounds in the binding pockets for 100 ns. In addition, the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) band gap energies and the molecular electrostatic potential (MEP) surface indicated that the four proposed compounds possessed favorable electronic properties, positioning them as good candidates for drug design.
Keywords
3D-QSAR
Angiogenesis
Molecular docking
Molecular dynamics
Multitargets
Quinazolin-4(3H)-one derivatives
PFKFB3
RTKs

1. Introduction
Angiogenesis is essential for normal physiological processes, including embryogenesis, tissue repair, wound healing, and fetal development [1,2]. However, it also contributes significantly to the advancement of various diseases, including cancer [3,4]. In the early 1970s, Folkman proposed that tumors depend on new blood vessel formation for growth and survival, suggesting that inhibiting angiogenesis could induce tumor regression by disrupting its nutrient supply [4]. This concept has driven the development of antiangiogenic therapies as an effective treatment for cancer and other diseases, including ocular, joint, and skin diseases [5,6]
A critical aspect of the angiogenic process is the activity of several receptor tyrosine kinases (RTKs), which promote cell growth and development [7]. As such, they have been identified as suitable targets for developing anti-angiogenesis drugs [8,9]. Among the various RTKs targeted in anti-angiogenesis strategies, the VEGFR family stands out due to its pivotal role in tumor angiogenesis, with VEGFR-1 and VEGFR-2 being particularly crucial in this process [9]. Similarly, angiogenic growth factors (Ang) and their tyrosine kinase with Ig and epidermal growth factor (EGF) homology domain-2 (TIE-2) have been linked to several processes such as blood vessel development [10]. Evidence from current studies has shown that the erythropoietin-producing hepatocyte receptor B4 (EPHB4) and its transmembrane-type ligand (ephrin B2) play a vital role in angiogenesis, vascular development, and pericyte recruitment [11]. Currently, clinical evaluation of angiogenesis inhibition as a strategy to limit tumor growth is underway [12]. Several molecules targeting angiogenic RTKs, including VEGFR-1, VEGFR-2, FGFRs, PDGFR, IGFR, RET, TIE-2, MET, KIT, EGFR, HER2, HER3, HER4, and EPHB4, have been identified. However, this approach faces challenges such as acquired resistance, limited efficacy, and toxicity [13]. These limitations highlight the urgent need for novel anti-angiogenic therapies that target different pathways, such as disrupting endothelial cell (EC) metabolism in the process of angiogenesis.
The metabolic activity of ECs, which line the vascular system, plays a crucial role in angiogenesis [14]. ECs predominantly derive energy from glycolysis, regulated by the PFKFB gene family, which encodes the enzyme PFKFB [15-17]. Among its isoforms, PFKFB3 exhibits the highest kinase/phosphatase activity and catalyzes the conversion of fructose-6-phosphate to fructose-2,6-bisphosphate, an activator of phosphofructokinase-1 (PFK-1), a key glycolytic enzyme [18,19]. Inhibiting PFKFB3 temporarily reduces glycolysis, effectively preventing pathological angiogenesis without adverse effects [19].
Targeting the energy metabolism and angiogenesis of tumors is a promising strategy for inhibiting tumor growth and metastasis. Currently, no drug targets both angiogenesis-related RTKs and PFKFB3, highlighting the potential of this dual approach to overcome cancer resistance and improve clinical outcomes.
To achieve lower costs and effective drug development, it is crucial to take advantage of computational methods such as 3D quantitative structure-activity relationship (QSAR) through the creation of comparative molecular field analysis (CoMFA) and Comparative Molecular Similarity Index Analysis (CoMSIA). These techniques examine pertinent descriptors to enhance the effectiveness of treatments and find appropriate candidates for drug development [20]. Furthermore, absorption, distribution, metabolism, excretion, and toxicity (ADMET) properties are essential in drug discovery and assessing the pharmacokinetics of new drug candidates [21]. Molecular docking and molecular dynamics simulations provide an initial guess for the binding modes of ligands to their target proteins and help understand the conformational flexibility of biomolecules, which is crucial for their function.
In recent years, quinazolin-4(3H)-ones have emerged as promising anti-angiogenic agents, primarily due to their potent VEGFR-2 inhibition [22]. This scaffold has been recognized as a versatile hinge-binding moiety for developing multitarget inhibitors targeting key angiogenic pathways, including VEGFR-2, Tie-2, and EphB4 [23-26]. Building on this foundation, the primary objective of this research is to investigate the potential of quinazolin-4(3H)-one derivatives as inhibitors of RTKs and PFKFB3, aiming to identify promising multitarget candidates that address these pathways simultaneously, offering a potential strategy to overcome acquired resistance in cancer treatment. To achieve this, 3D-QSAR and molecular docking were used to determine critical molecular descriptors influencing kinase activity and the interaction types with VEGFR1, VEGFR-2, FGFRs, PDGFR, IGFR, RET, TIE-2, MET, KIT, EGFR, HER2, HER3, HER4, EPHB4, and PFKFB3. Additionally, molecular dynamics simulations and density functional theory (DFT) analyses were performed to assess the stability of selected compounds and accurately predict and analyze the electronic structure and properties of drug molecules and their interactions with biological targets.
2. Materials and Methods
2.1. Data set
To build 3D-QSAR models for a set of 40 quinazolin-4(3H)-one derivative [27-32] (as indicated in Supplementary File 1 Table S1). The dataset was randomly divided into two subsets: one for training the model and the other for testing its performance.
2.2. Minimization and molecular alignment
The alignment technique utilized in comparative molecular field analysis (CoMFA) and comparative molecular similarity indices analysis (CoMSIA) analysis, as emphasized by Xiao et al [32], poses a substantial difficulty. The SYBYL-X 2.0 molecular modeling program was used to model all the substances being considered. The module for the designed benzothiazole derivatives was optimized using the Tripos force field. The SYBYL program framework was employed to assign Gasteiger-Huckel charges, with convergence set at 0.01 kcal/mol. Figure 1 illustrates the alignment of the 40 benzothiazole derivative compounds to a common core, with molecule 39 (the most potent in the dataset) acting as the reference compound.

- Alignment of 40 Quinazolin-4(3H)-one derivatives using molecule 39 as a template.
2.3. Methodology for CoMFA and CoMSIA descriptor generation
The various 3D-QSAR models were constructed using the Comparative Molecular Field Analysis (CoMFA) and Comparative Molecular Similarity Indices Analysis (CoMSIA) methods [33,34]. These methods were used to predict the anticancer activity (pIC50 Pred) of the studied quinazolin-4(3H)-one derivatives across various fields, including steric, electrostatic, hydrophobic, hydrogen bond donor, and acceptor interactions.
In the CoMFA method, all molecules were arranged on a grid lattice with a 2.0 Å spacing. An sp3 hybridized carbon atom with a charge +1 is used to model the electrostatic and steric field energies. The interactions’ cutoff energies have values of 30 kcal/mol, and the column filtering is set to a value of 2.0 kcal/mol [34].
The CoMSIA method computes different descriptors on a 2.0 Å grid. The cutoff energy for interactions is set around 30 kcal/mol, with an attenuation factor of 0.3. Column filtering is performed by evaluating molecular field energies, ensuring that the variation is less than 2.0 kcal/mol [35].
2.4. Partial least squares analysis
The 3D-QSAR models in this study are developed using the partial least squares (PLS) regression method, which establishes the correlation between two sets of variables [36]. In this investigation, the two variables are the CoMFA/CoMSIA descriptors and the pIC50 anticancer activity values. The calculation of the optimal number of components (N) and the determination of the cross-validation correlation coefficient (Q2) were performed via leave-one-out cross-validation [37]. Subsequently, non-cross-validation techniques were conducted to determine the determination coefficient (R2), the F-test value (F), and the standard error of estimation (SEE) [38].
Additionally, to assess the robustness of the developed models, they were externally validated using a set of four molecules [39]. The coefficient of determination (Rext2) for the test set was calculated using the following Eq. (1):
For the models to be considered acceptable, they must simultaneously meet the following criteria: Q2 > 0.5, R2 > 0.6, and Rext2 > 0.6 [40]
2.5. In silico pharmacokinetics absorption, distribution, metabolism, excretion, and toxicity (ADMET)
ADMET studies are pivotal for evaluating potential drug candidates in drug research and development. ADMET refers to a set of pharmacokinetic parameters that can be evaluated through in silico approaches. These parameters include the absorption of a compound in the human intestine, its ability to cross the blood-brain barrier (BBB) and affect the central nervous system (CNS), the body’s biotransformation of the compound, as well as the drug’s overall clearance and potential toxicity levels. These factors are crucial in assessing the drug’s suitability for therapeutic use [21].
2.6. Molecular docking
The 3D structures of all 18 protein targets, as outlined in Supplementary File 1 Table S2, were retrieved from the Protein Data Bank (RCSB) via https://www.rcsb.org/(accessed in May 2024). These structures were visualized using UCSF Chimera [41] and prepared with MGLtools (version 1.5.6, The Scripps Research Institute, La Jolla, CA, USA) [41]. Preprocessing involved removing water molecules, heteroatoms (hetatm), and co-crystallized ligands. Subsequently, polar hydrogen atoms were added, Gasteiger charges were assigned, and the structures were converted to pdbqt format for further analysis [42]. The grid box spacing was set to 0.375 Å, centered on the regions where co-crystallized ligands interact with active site residues. Docking simulations were conducted for all 18 protein targets, generating nine poses per protein-ligand complex based on docking affinity. The docking outcomes were visualized and analyzed using Discovery Studio Viewer to identify critical interactions between ligands and protein binding sites. For each protein target, the conformation with the lowest binding affinity (as indicated by docking scores) and the highest number of bonds was chosen as the initial binding mode for subsequent molecular dynamics simulations.
Next, we conducted redocking, a widely accepted technique for validating the molecular docking process. This method entails removing the crystallized ligand from the protein and conducting a docking analysis within the same binding site [42]. By comparing the alignment of the docked ligand with the original crystallized ligand, we can accurately assess the docking accuracy through the analysis of the root mean square deviation (RMSD) parameter using PyMOL [43].
2.7. Molecular dynamics simulations (DFT)
Molecular dynamics simulations (MDs) were carried out to examine the conformations of each complex with the lowest binding affinities, utilizing GROMACS version 2019.3 [44] on a high-performance computing cluster. Protein topologies were generated using the pdb2gmx module in GROMACS with the CHARMM27 force field [45], while ligand topologies were created using SwissParam to derive the necessary force field parameters [46]. Each complex was positioned within a dodecahedral box (1.0 nm) filled with TIP3P water molecules and neutralized with counter ions. Energy minimization was performed using the steepest descent method, ensuring that the maximum force (Fmax) remained below 1000 kJ/mol/nm [47]. The system was equilibrated at 300 Kelvin and 1 bar pressure through two consecutive 100 ps simulations employing the canonical NVT and isobaric NPT ensembles. Temperature and pressure were controlled using the Berendsen thermostat and the Parrinello-Rahman barostat. The MD simulations were conducted over 100 ns, totaling 50,000,000 steps, with coordinates recorded every 2 fs. The output trajectories were generated, and the corresponding data files were analyzed to gain a deeper understanding of the protein’s behavior.
2.8. Density functional theory
All calculations were performed within the DFT framework using the Gaussian software [48]. The electronic and structural characteristics of the selected compounds, including molecule 39, were examined at the B3LYP/6-311++G(d,p) level of theory. Various molecular parameters were computed, including the highest occupied molecular orbital (HOMO), the lowest unoccupied molecular orbital (LUMO), their respective energies (EHOMO and ELUMO), the energy gap (ΔEgap), ionization potential (I), electron affinity (A), electronegativity (χ), chemical hardness (η), chemical potential (μ), chemical softness (S), electrophilicity index (ω), dipole moment (D), and molecular electrostatic potential (MEP) maps.
3. Results and Discussion
3.1. Alignment of quinazolin-4(3H)-one derivatives and 3D-QSAR models generation
Molecule 39, which demonstrates the highest inhibitor activity was selected to construct a robust 3D-QSAR model, serve as a template for data alignment (Table S1), and facilitate the visualization of the contour maps in the CoMFA and CoMSIA analyses (Figure 1).
The primary goal of this step is to develop robust CoMFA and CoMSIA models based on the observed and predicted pIC50 values for the training and test sets. For the CoMFA model, steric and electrostatic fields were combined (Supplementary File 2 Table S1). In contrast, a total of 31 combinations of steric, electrostatic, hydrophobic, H-bond donor, and H-bond acceptor were employed in developing CoMSIA (Supplementary File 2 Tables S2-S3). The optimal model was determined by selecting the one with the highest coefficient of determination values for both non-cross-validation (R2) and cross-validation (Q2), along with the lowest standard error of estimate (SEE), the fewest principal components (N), and the most significant F-test value [49]. Table 1 lists the best CoMFA and CoMSIA models.
| Generated model | Fields description | Q2 | N | SEE | R2 | F |
|---|---|---|---|---|---|---|
| CoMFA/SE | Steric (S)- Electrostatic (E) | 0.650 | 5 | 0.078 | 0.997 | 1600.005 |
| CoMSIA/SEHDA | Steric (S)-Electrostatic (E)-Hydrophobic (H)- Donor (D) and Acceptor(A) of hydrogen bond | 0.580 | 2 | 0.378 | 0.923 | 156.755 |
| CoMSIA/SEHA | Steric (S)-Electrostatic (E)-Hydrophobic (H)-Acceptor(A) of hydrogen bond | 0.624 | 5 | 0.091 | 0.996 | 1176.535 |
| CoMSIA/SHDA | Steric (S)-Hydrophobic (H)- Donor (D) and Acceptor(A) of hydrogen bond | 0.630 | 10 | 0.059 | 0.999 | 1410.476 |
| CoMSIA/SEH | Steric (S)-Electrostatic (E)-Hydrophobic (H) | 0.615 | 4 | 0.133 | 0.991 | 677.977 |
| CoMSIA/SEA | Steric (S)-Electrostatic (E)-Acceptor(A) of hydrogen bond | 0.598 | 2 | 0.448 | 0.893 | 108.139 |
| CoMSIA/HDA | Hydrophobic (H)- Donor (D) and Acceptor(A) of hydrogen bond | 0.594 | 5 | 0.099 | 0.955 | 977.913 |
| CoMSIA/SHA | Steric (S)-Hydrophobic (H)- Acceptor(A) of hydrogen bond | 0.717 | 5 | 0.101 | 0.995 | 946.847 |
| CoMSIA/SA | Steric (S)-Acceptor(A) of hydrogen bond | 0.715 | 6 | 0.102 | 0.995 | 771.070 |
| CoMSIA/HA | Hydrophobic (H)- Acceptor(A) of hydrogen bond | 0.663 | 3 | 0.273 | 0.962 | 208.917 |
Q2 is the cross-validation R-squared, N is the sample size, SEE is the Standard error of estimate, R2 is the Coefficient of determination, F is the F-score.
Principle component analysis of the selected models indicated that the one with the highest Q2 (0.717) and R2 (0.995), the lowest SEE (0.101), and the most significant F-value (946.847) was the most robust and reliable. This model employed steric, hydrophobic, and hydrogen bond acceptor (SHA) fields and demonstrated the highest predictive power (R2Pred = 0.832). Therefore, the CoMSIA/SHA model exhibited good stability and predictability, as indicated in Table 2. The model’s predictive ability was validated according to Golbraikh and Tropsha’s statistical criteria [50-52], as shown in Table 3. The proposed CoMSIA/ steric, hydrophobic, acceptor (SHA) model was further validated based on these criteria.
| Compounds | pIC50 | pIC50pred | Steric (S) | Hydrophobic (H) | Acceptor (A) of hydrogen bond |
|---|---|---|---|---|---|
| 1 | 6.208 | 6.175 | 6.7286 | 5.7512 | 2.733 |
| 2 | 6.469 | 6.441 | 6.819 | 5.1842 | 3.0321 |
| 3 | 6.305 | 6.485 | 6.8839 | 5.7413 | 3.2838 |
| 4 | 6.142 | 5.947 | 6.4124 | 5.1059 | 3.0679 |
| 5 | 6.173 | 6.199 | 6.4267 | 5.5753 | 3.0626 |
| 6 | 6.147 | 6.151 | 6.5706 | 5.2854 | 3.0563 |
| 7 | 6.351 | 6.332 | 6.9012 | 5.0762 | 3.5752 |
| 9 | 5.509 | 5.62 | 6.3124 | 4.1196 | 3.3946 |
| 10 | 5.469 | 5.399 | 6.3238 | 4.419 | 3.6685 |
| 11 | 4.863 | 4.948 | 5.414 | 3.9463 | 3.0422 |
| 12 | 4.947 | 4.864 | 5.4212 | 4.5141 | 3.0412 |
| 14 | 4.893 | 5.025 | 5.7246 | 4.6434 | 3.0402 |
| 15 | 4.955 | 5.017 | 5.7356 | 5.1157 | 3.0388 |
| 16 | 4.979 | 5.059 | 5.9502 | 4.4366 | 3.0391 |
| 17 | 5.119 | 5.101 | 5.2691 | 4.6253 | 2.4432 |
| 19 | 5.032 | 5.099 | 5.5032 | 4.4255 | 2.443 |
| 20 | 5.268 | 5.197 | 5.5907 | 5.2306 | 2.4416 |
| 23 | 5.18 | 5.024 | 5.4181 | 3.9548 | 3.045 |
| 24 | 5.071 | 5.169 | 5.7321 | 4.6407 | 3.0436 |
| 25 | 5.032 | 4.997 | 5.6461 | 4.4409 | 3.5573 |
| 27 | 5.18 | 5.258 | 5.8253 | 5.5645 | 3.0601 |
| 29 | 5.31 | 5.108 | 5.9507 | 5.028 | 3.558 |
| 30 | 8.402 | 8.443 | 6.428 | 4.8528 | 3.9933 |
| 31 | 8.379 | 8.386 | 6.7686 | 5.4639 | 4.2416 |
| 32 | 8.476 | 8.42 | 6.4252 | 5.1326 | 3.7411 |
| 33 | 6.238 | 6.254 | 5.8835 | 4.7073 | 3.0543 |
| 36 | 8.48 | 8.516 | 6.7421 | 5.1592 | 3.2643 |
| 38 | 8.504 | 8.439 | 6.866 | 4.7454 | 3.665 |
| 39 | 8.509 | 8.517 | 6.4821 | 4.6171 | 3.6557 |
| 8* | 5.432 | 6.848 | 6.4254 | 4.9344 | 3.6677 |
| 13* | 4.971 | 4.98 | 5.6457 | 3.6986 | 3.0417 |
| 18* | 4.91 | 5.032 | 5.2785 | 5.1111 | 2.4442 |
| 21* | 5.071 | 5.105 | 5.5957 | 5.6696 | 2.4412 |
| 22* | 5.032 | 5.238 | 5.8241 | 5.0476 | 2.4419 |
| 26* | 5.125 | 5.149 | 5.5034 | 5.0494 | 3.063 |
| 28* | 5.337 | 5.109 | 5.649 | 4.4479 | 3.5632 |
| 34* | 6.124 | 6.595 | 6.6622 | 5.3319 | 3.0627 |
| 35* | 6.229 | 6.595 | 6.6695 | 5.3218 | 3.6918 |
| 37* | 8.443 | 7.794 | 6.9317 | 4.6499 | 3.6539 |
| 40* | 8.5 | 7.985 | 7.0483 | 5.1657 | 3.6604 |
| Statistical parameter | Score | Threshold | Comment |
|---|---|---|---|
| R2pred | 0.832 | More than 0.600 | Passed |
| R02: Determination coefficient for the plot of predicted against observed at zero intercept | 0.790 | More than 0.600 | Passed |
| R0’2: Determination coefficient of the plot of observed versus predicted at zero intercept. | 0.820 | More than 0.600 | Passed |
| 0.030 | Less than 0.300 | Passed | |
| 0.050 | Less than 0.100 | Passed | |
| 0.014 | Less than 0.100 | Passed | |
| K: Zero intercept slope of predicted against observed activity for the test set | 0.917 | Passed | |
| K’: Zero intercept slope of observed against predicted activity for the test set | 0.942 | Passed | |
| 0.661 | More than 0.600 | Passed | |
| 0.739 | More than 0.600 | Passed |
3.2. Graphical interpretation of the best COMSIA/SHA model
The validated 3D-QSAR model was visualized using CoMSIA/SHA contour maps, using molecule 39 as a template. The contour maps aid in an understanding of the structural characteristics of the compounds and provide insights into how molecular features contribute to their biological activities. Figure 2 illustrates the critical fields: steric (a), hydrophobic (b), and hydrogen bond acceptor (c). In the steric field (a), green contours represent regions where larger substituents are favorable, indicating that bulkier groups in these areas enhance activity, while yellow contours highlight regions where minimal steric bulk is preferred, suggesting that excessively bulky groups could negatively impact activity. The hydrophobic field (b) is represented by blue and white contours, with blue areas favoring positive hydrophobicity, indicating that more hydrophobic substituents in these regions could boost activity, whereas white contours signify regions where hydrophilic properties are more favorable, suggesting that hydrophobic groups should be minimized. The hydrogen bond acceptor field (c) shows magenta contours in regions that favor hydrogen bond acceptor groups, highlighting the importance of such functionalities for improving biological interactions. Conversely, red contours mark areas where hydrogen bond acceptor properties are disfavored, suggesting that such features could reduce activity if placed there. In summary, the structure–activity relationship (SAR) of potent compounds highlights the importance of steric and hydrophobic contributions and the hydrogen bond acceptor field. These contour maps collectively offer a strategic framework for designing more potent VEGFR-2 inhibitors.

- Contour maps (CoMSIA/SHA) of molecule 39 with 2 Å grid spacing; (a) Steric fields, (b) Hydrophobic fields, and (c) H-bond Acceptor fields.
The significant values of the CoMSIA field fractions outlined in Table 4 suggest that enhancing hydrogen bonding, steric, and hydrophobic interactions can significantly improve the efficacy of these compounds against tumor angiogenesis.
| COMSI/SHA Fields | Fraction |
|---|---|
| Steric (S) | 0.244 |
| Hydrophobic (H) | 0.277 |
| Acceptor (A) | 0.478 |
3.3. Design new inhibitor candidates against VEGFR-2 kinase based on the selected molecule 39
To design new VEGFR-2 kinase inhibitors by investigating the different structural features extracted from the selected CoMSIA/SHA models (Figure 3). We substituted the sulfonic amide group (–H2NO2S) with a 1-hydrosulfonyl-1H-1,2,3-triazole (–C2H2N3O2S) group, hydrosulfonylbenzene (–C6H5O2S), or 2-cyclopentylpropan-2-yl hydrosulfonylmethanoate (–C9H15O4S) at the R1 position. We added the proper substituents in the R2, R3, R4, R’4, R5, R’5, R6, R7, R8, and R9 positions according to the contour maps (Figure 3 and Supplementary File 1 Table S3). As a result, fifty-nine new Kinase inhibitors were proposed, and their VEGFR-2 kinase activity was computed using the proposed CoMFA and CoMSIA/SHA models (Table 5).

- CoMSIA/SHA SAR of molecule 39.
| Compound | Predicted pIC50 | |
|---|---|---|
| CoMFA | CoMSIA/SHA | |
| Pred1 | 8.698 | 8.892 |
| Pred2 | 8.667 | 9.097 |
| Pred3 | 8.068 | 8.891 |
| Pred4 | 7.354 | 8.563 |
| Pred5 | 8.244 | 8.502 |
| Pred6 | 7.259 | 8.506 |
| Pred7 | 8.064 | 9.481 |
| Pred8 | 7.815 | 8.777 |
| Pred9 | 8.001 | 8.671 |
| Pred10 | 7.64 | 9.702 |
| Pred11 | 7.821 | 8.534 |
| Pred12 | 8.009 | 9.387 |
| Pred13 | 7.204 | 8.966 |
| Pred14 | 7.792 | 8.895 |
| Pred15 | 8.073 | 8.546 |
| Pred16 | 7.42 | 9.112 |
| Pred17 | 6.943 | 8.626 |
| Pred18 | 7.126 | 8.832 |
| Pred19 | 8.103 | 8.652 |
| Pred20 | 8.029 | 8.577 |
| Pred21 | 7.94 | 10.611 |
| Pred22 | 7.441 | 8.794 |
| Pred23 | 7.902 | 9.712 |
| Pred24 | 7.436 | 8.845 |
| Pred25 | 7.839 | 9.006 |
| Pred26 | 8.12 | 9.072 |
| Pred27 | 7.2 | 8.948 |
| Pred28 | 7.79 | 10.285 |
| Pred29 | 7.769 | 9.644 |
| Pred30 | 7.732 | 8.749 |
| Pred31 | 7.694 | 9.081 |
| Pred32 | 7.615 | 8.841 |
| Pred33 | 7.329 | 9.607 |
| Pred34 | 7.402 | 8.566 |
| Pred35 | 7.963 | 8.59 |
| Pred36 | 7.788 | 8.807 |
| Pred37 | 7.057 | 8.96 |
| Pred38 | 7.839 | 8.633 |
| Pred39 | 7.785 | 8.513 |
| Pred40 | 7.884 | 8.682 |
| Pred41 | 7.77 | 8.712 |
| Pred42 | 7.412 | 8.878 |
| Pred43 | 7.478 | 8.707 |
| Pred44 | 7.429 | 8.864 |
| Pred45 | 7.393 | 8.756 |
| Pred46 | 7.602 | 9.086 |
| Pred47 | 7.092 | 8.552 |
| Pred48 | 7.814 | 8.664 |
| Pred49 | 7.769 | 8.613 |
| Pred50 | 7.799 | 9.672 |
| Pred51 | 7.52 | 8.881 |
| Pred52 | 7.859 | 8.91 |
| Pred53 | 6.322 | 9.262 |
| Pred54 | 7.38 | 8.904 |
| Pred55 | 7.332 | 8.974 |
| Pred56 | 7.326 | 8.777 |
| Pred57 | 7.082 | 8.694 |
| Pred58 | 7.352 | 8.537 |
| Pred59 | 7.48 | 8.546 |
3.4. Drug-likeness assessment and ADMET predictions
Several developed drugs fail in the clinical development stage because of problems like inadequate blood–brain barrier (BBB) permeability, toxicity, or insufficient efficacy. Consequently, predicting and optimizing the ADMET properties of new chemical compounds is crucial to avoid complications in clinical trials [53,54]. This study evaluated the ADMET pharmacokinetic parameters of the 59 newly designed compounds using pkCSM [55]. The ADMET properties and safety endpoints show the compounds’ high intestinal absorption rates, ranging from 38.533% to 100%, with compound Pred49 showing the highest value (Supplementary File 1 Table S4).
A Volume of Distribution at Steady State (VDss) value is classified as low when it falls below 0.71 L/kg, corresponding to a logarithmic value (log VDss) of less than -0.15. Conversely, a VDss value is considered high when it exceeds 2.81 L/kg, equating to a logarithmic value (log VDss) greater than 0.45 [36-58]. Based on these criteria, all proposed compounds, molecule 39, respect the defined barrier for low VDss values, suggesting limited tissue distribution for these molecules.
For BBB and the CNS permeability, compounds exhibiting a LogBB value below -1 indicate poor diffusion into the brain. In contrast, those with a LogBB above 0.3 are deemed capable of breaching the BBB. Furthermore, compounds with a LogPS value over -2 are considered effective in penetrating the CNS; however, those with a LogPS less than -3 are viewed as having difficulties in CNS barrier penetration [59-61]. Our analysis of the permeability results shows that the 59 newly designed compounds are predicted to fail to traverse these barriers.
Among 17 cytochrome P450 (CYP) families, 57 CYP genes were identified in humans, suggesting that CYP enzymes are the main ones responsible for the biotransformation of most foreign substances, including 70-80% of all drugs in clinical use [62-64]. Thus, most of the administered drugs undergo the initial phase of metabolism with these enzyme families. CYP1A2, 2C19, 2D6, 3A4, and 2C9 enzymes are used for our analysis. Therefore, all proposed ligands were identified as either inhibitors or substrates for CYP2D6 (Supplementary File 1 Table S4). Additionally, none of the newly designed compounds exhibited issues with clearance, which measures the rate of drug elimination from the body [65].
In the early stages of drug development, assessing the toxicity of proposed compounds is essential to ensure their safety and efficacy. Therefore, each newly designed compound examined in this study is subjected to a toxicity assessment using the AMES test [66]. This test is widely used due to its simplicity, cost-effectiveness, and ability to provide rapid results regarding a compound’s genotoxicity. As shown in Supplementary File 1 Table S4, the designed ligands were non-toxic. However, molecule 39, Pred11, Pred30, and Pred51 displayed some toxicity. These findings suggest that most of the compounds are promising for further development. Nevertheless, certain compounds need alterations to improve their non-toxicity. The high absorption, distribution, and permeability properties observed make them strong candidates for clinical evaluation.
3.5. Molecular docking (MD) study
Molecular docking and dynamics are central to designing and screening new bioactive molecules [67]. Molecular docking of the 18 protein targets with all newly designed molecules was performed to identify the optimal binding modes, allowing inhibition of the target proteins. This was identified by virtual screening to select the molecules with the highest binding affinity scores. The molecular docking procedures were initially validated as outlined in the methodology section. The Root Mean Square Deviation (RMSD) scores obtained during the validation of the chosen targets ranged from 0.2 to 1.7 Å. An RMSD value below 2.0 Å indicates a reliable predictive protocol for assessing protein-ligand interactions [68]. This indicates that the protocols used were appropriate for docking the proposed compound.
Additionally, parameters such as binding affinity and specific amino acid residues, along with the dimensions and coordinates of the grid box, were observed during validation. Lower Binding affinity values indicate better interaction between the ligand and the target [69]. Supplementary File 2 Table S4 presents the binding affinity values for the most favorable interaction poses of all 59 designed molecules with the selected receptor tyrosine kinase (RTK) targets and PFKFB3. Molecule 39 and the FDA-approved drug for each target were used as a control for comparison. This analysis shows that Pred3, Pred4, Pred34, and Pred49, exhibit better binding affinities than the reference compounds and molecule 39 (Table 6).
| Selected targets | Corresponding reference drug | RD binding affinity (Kcal/mol) | Binding affinity of ligands (Kcal/mol) | ||||
|---|---|---|---|---|---|---|---|
| Pred3 | Pred4 | Pred34 | Pred49 | Molecule 39 | |||
| FGFR1 | lenvatinib | -8.1 | -10 | -10 | -10.2 | -10 | -10 |
| FGFR2 | lenvatinib | -9 | -10.6 | -10.5 | -10.5 | -10.8 | -10.3 |
| FGFR3 | lenvatinib | -8.2 | -10.1 | -10.5 | -9.7 | -10 | -9.5 |
| FGFR4 | lenvatinib | -7.3 | -9.4 | -9.6 | -9.4 | -9.4 | -9 |
| PDGFR_alpha | Regorafenib | -9.8 | -10.4 | -10.1 | -10 | -9.7 | -9.5 |
| TIE2 | Rebastinib | -9.4 | -10 | -10 | -10.5 | -10 | -8.8 |
| KIT | Lenvatinib | -8.7 | -10.4 | -9.6 | -10 | -9.5 | -8.9 |
| MET | Quinoxaline | -9.3 | -10.4 | -10.1 | -10.5 | -10.2 | -9.2 |
| PFKFB3 | N aryl aminoquinoxaline | -7.2 | -10.4 | -10.9 | -10.4 | -10.9 | -9.5 |
| VEGFR1 | Sunitinib | -7.8 | -8.5 | -10.4 | -9 | -8.9 | -8.1 |
| EGFR | Erlotinib | -7.3 | -10.1 | -9.4 | -10.3 | -9.8 | -9.1 |
| VEGFR2 | Axitinib | -8.4 | -9.6 | -10 | -10 | -9.6 | -9.2 |
| EPHB4 | Dasatinib | -9.1 | -10.3 | -10 | -10.5 | -10.1 | -9.7 |
| HER2 | Afatinib | -9 | -11.5 | -10.8 | -11 | -11 | -10.5 |
| HER4 | Afatinib | -9 | -11 | -10.6 | -10.6 | -9.5 | -10 |
| HER3 | Sapitinib | -9 | -9.8 | -9.6 | -10 | -10 | -9 |
| IGFR | Ceritinib | -8.1 | -9 | -9.4 | -8.7 | -9 | -8.6 |
| RET | Sorafenib | -10 | -10.4 | -10.7 | -10.5 | -10.2 | -10.3 |
Furthermore, the 2D binding interaction of the compounds (Pred3, Pred4, Pred34, and Pred49), molecule 39, and the reference drug for each protein target showed a similar interaction in the binding pocket of all the targets. This is due to several amino acids participating in the same interactions compared to the food and drug administration (FDA) drugs and molecule 39 (Supplementary File 1 Table S5). Additionally, the proposed compounds demonstrate various interactions with different residues, especially in the regions where fragments have been attached to the core structure across all targeted proteins. These results emphasize the strong potential of the proposed compounds as multi-target drugs. Research indicates that multi-target compounds are particularly effective in treating complex diseases [70].
3.6. Molecular dynamics (MD)
Molecular dynamics simulations were conducted to assess the impact of the selected lead compounds (Pred3, Pred4, Pred34, and Pred49), molecule 39, and the reference drug for each target on their respective structures and stability within the binding pockets of each protein target. The optimal pose, which exhibited the most significant binding affinity for each target, was employed as the initial structure for the 100 ns molecular dynamics simulation [71]. The resultant trajectory data files were analyzed for RMSD, RMSF, Rg, and hydrogen bond.
During the simulation, the RMSD was calculated to determine the overall stability of the selected complexes [72] and taken as the primary indicator of the system’s convergence. A lower RMSD value signifies a higher level of system stability and vice versa [72]. Upon examining the RMSD graph shown in Figure 4, it is evident that all the protein-ligand complexes formed by the proposed compounds exhibited reduced fluctuation in their spectrum compared to the complexes formed by the reference drug and molecule 39, which indicates a minimal disruption in their conformational dynamics during the whole simulation. Pred49 shows a deviation between 65 and 80 ns for FGFR3, reaching approximately 0.37 nm before stabilizing. Similarly, for FGFR4, Pred49 displays a slight increase after 80 ns but stabilizes at around 0.35 nm until the end of the simulation. Meanwhile, in the case of KIT proto-oncogene receptor tyrosine kinase, a slight deviation was observed within the first 50 ns, where the RMSD briefly increased to 0.38 nm before stabilizing and maintaining a consistent value near 0.25 nm until 100 ns. These observations suggest that while minor fluctuations occurred, they were brief and quickly resolved, demonstrating that the proposed compounds exhibit greater stability than molecule 39 and reference drugs. Interestingly, this finding is consistent with the docking results, where the proposed compounds exhibited better binding affinities (Table 6), showing their strong potential to serve as promising candidates for inhibiting the selected RTKs and PFKFB3.

- RMSD profiles of the designed compounds, molecule 39, and the reference drug for the selected RTKs and PFKFB3.
RMSF plots during 100 ns were generated to investigate the binding affinity effects of the selected compounds on the dynamic structure of proteins and the behavior of their essential amino acids [73]. An increase in RMSF values indicates that the protein structure is more likely to have greater flexibility and lower residual stability [74].
RMSF data analysis of the selected complexes generally showed minimal fluctuations in the residues (<4 nm) during the simulations (Figure 5). Compared with the reference drugs and molecule 39, Pred49 exhibited significant fluctuations in residues 550-570 (FGFR3) and 860-870 (HER3). Additionally, Pred4 displayed higher fluctuations in residues 560-580 (FGFR2). Similarly, Pred3, Pred4, and Pred34 are associated with increased fluctuations of residues 760-780 within EPHB4. The fluctuations in these proposed compounds during the simulations could suggest that they assume different conformations to reach the most energetically stable state within the binding cavity of the targets [69]. Notably, compounds Pred3, Pred4, Pred34, and Pred49 displayed the lowest average RMSF values, suggesting they could be more effective inhibitors than the reference drugs.

- Comparison of RMSF profiles between the designed compounds, molecule 39, and the reference drug for each RTK target and PFKFB3.
The plots of radius of gyration (Rg) are calculated to assess how compact the protein changes upon binding to various compounds throughout the entire simulation [75]. A stable Rg value usually indicates a well-folded structure, while variations in the Rg value during the simulation suggest that the protein structure may be unfolding [76]. The plot shows that, except for Pred4, which has a larger EGFR Rg value, all the remaining proposed compound complexes have a lower Rg value, making them relatively stable compared to the reference drugs and molecule 39 complexes (Figure 6). Notably, the four proposed compounds display lower Rg values than reference compounds, suggesting they are more compact.

- Comparison of Radius of gyration profiles between the designed compounds, molecule 39, and the reference drug for each RTK target and PFKFB3.
Finally, hydrogen bonds play a crucial role in determining binding affinity and significantly contribute to the interaction between ligands and proteins. The quality and positioning of hydrogen bonds are equally important, as they directly influence the strength and stability of the ligand-protein complex. Furthermore, a greater number of hydrogen bonds generally enhances the overall stability and binding affinity of the complex [77]. To confirm the stability of the docked complexes, we assessed the hydrogen bonds between each protein target and selected compounds, including the reference drug and molecule 39, in a solvent environment during MD simulations using the Gromacs gmx H-bond module (Supplementary File 3). The comparison of Pred3, Pred4, Pred34, and Pred49 with reference drugs and molecule 39 across 18 protein targets demonstrates consistent and improved interaction profiles, as evidenced by H-bond analysis. Across all target proteins, these compounds exhibit a comparable or higher number of H-bonds than the reference drugs, indicating enhanced binding affinity and stability within the binding pockets throughout the simulation. Pred3 and Pred4 consistently displayed the highest H-bond numbers among the designed compounds, particularly for key targets such as FGFR1, FGFR3, KIT, PFKFB3, and VEGFR2, which are critical for their inhibitory activity. Pred34 and Pred49 also demonstrated strong H-bond interactions, with numbers comparable to Pred3 and Pred4 for several targets, including FGFR2, EGFR, and IGFR. This consistent performance across multiple targets highlights the versatility and effectiveness of the designed compounds in forming stable ligand-protein complexes, reinforcing their potential as selective modulators of protein activity.
3.7. DFT analysis
DFT studies were conducted on the best-selected compounds, including molecule 39, with optimized ground-state geometries calculated using the DFT-B3LYP/6-31G++(d, p) basis set [78]. Molecular orbital energies explain a molecule’s electrical and optical properties. Frontier molecular orbitals (FMOs), particularly the HOMO and the LUMO, are essential in demonstrating the active sites of compounds during molecular interactions [79]. The HOMO value indicates the region with the greatest energy and highly reactive electrons, while LUMO identifies where the atoms are most likely to accept electrons [80].
The HOMO/LUMO diagrams for the selected compounds have been shown in Figure 7; the red and green regions correspond to the positive and negative phases of the molecular orbitals, respectively. Instead of being located over the same atomic sites, the FMOs lie in different sites of each compound, indicating their polar nature and the feasibility of intramolecular interaction with multiple sites in the targeted binding pocket. The Pred3, Pred4, Pred34, and Pred49 compounds have energy gaps of 3.80, 3.53, 3.80, and 4.80 eV, respectively. The energy gap between HOMO and LUMO reflects the compound’s stability and reactivity. As a result, less energy is needed to move electrons from the ground state HOMO to the excited state LUMO (Figure 7). Notably, Pred3 and Pred34, with an energy gap of 3.80, are comparable to molecule 39, suggesting similar reactivity and stability (Figure 7). However, Pred4, with a lower energy gap of 3.53, is expected to be more reactive, while Pred49, with a higher energy gap of 4.80, indicates lower reactivity and higher stability. Numerous studies have demonstrated that a smaller HOMO–LUMO gap usually indicates increased reactivity and a greater potential for chemical interactions with biological molecular targets [81].

- FMOs (a) HOMO, LUMO, and HOMO-LUMO gap with corresponding energies, and density of states (DOS) spectra of (b) molecule 39, (c) Pred3, (d) Pred4, (e) Pred34, and (f) Pred49, respectively.
In addition to FMO, the global descriptive parameters of the compounds were calculated, including global softness, hardness, ionization energy, and electron affinity (Supplementary File 2, Table S4). These values were obtained from HOMO and LUMO energy. Global softness and hardness are indicators of stability that increase with hardness and decrease with softness. The reactivity of the ligands decreases with hardness and increases with softness under study [82]. All the studied compounds were found to be stable within permissible limits. Among them, Pred3 and Pred34 exhibit a comparable value to that of molecule 39. Pred49 is the most stable but least reactive, and Pred4 is the most reactive but least stable. The same description applies to global hardness, the opposite of global softness. The ligands’ ionization potential (IP) and electron affinity (EA) were identified. IP refers to the amount of energy needed to remove an electron from an atom of a compound in its ground state. Chemically, a compound with a lower IP is generally more stable but less reactive, while a higher IP indicates less stability but greater reactivity [83]. Furthermore, EA, which indicates how likely an electron is to gain a negative charge when attached to a neutral atom, was also evaluated, according to stable, FDA-approved, or commercialized drugs, which typically exhibit IP values between 4 and 15 eV and EA values between -3 and 7 eV [84]. The proposed compounds fall within these ranges, indicating that they possess acceptable electronic properties, making them promising candidates for drug design.
3.8. The molecular electrostatic potential
Molecular electrostatic potential (MEP) analysis of compounds Pred3, Pred4, Pred34, Pred49, and molecule 39 offers important information about their chemical reactivity and potential for molecular interactions (Figure 8). The analyses show that the proposed compounds exhibit notable electrostatic characteristics, primarily influenced by the positive charge of nitrogen and the negative charge of the oxygen atoms, which play critical roles in their interaction potential.

- MEP analyses of (a) Pred3, (b) Pred4, (c) Pred34, (d) Pred49, and (e) molecule 39.
In the designed compounds Pred3, Pred4, Pred34, and Pred49, the nitrogen atoms consistently contribute to regions of negative electrostatic potential, indicating their nucleophilic properties and allowing enhanced interactions with protein targets. These nucleophilic regions facilitate strong binding by forming electrostatic interactions and hydrogen bonds with the protein’s positively charged or electrophilic regions, such as active site residues of the targets, such as FGFRs and VEGFR2 (Figures 8a-e). The oxygen atoms of the four compounds exhibit regions of negative electrostatic potential, emphasizing their role as hydrogen bond acceptors and contributing to their interaction potential. Notably, the MEP profiles of Pred3, Pred4, Pred34, and Pred49 are identical, indicating a common electrostatic potential distribution that enhances their binding affinity and supports their potential as selective modulators of protein activity. Molecule 39 shows a similar MEP profile with comparable negative potentials near the nitrogen atoms and positive potentials near the oxygen atoms, supporting its nucleophilic activity and ability to accept hydrogen bonds.
Finally, the results of the DFT and MEP analyses offer a comprehensive view of the electronic, energetic, and electrostatic properties of the designed compounds Pred3, Pred4, Pred34, Pred49, and molecule 39. The DFT analysis provides insights into total and binding energies, HOMO-LUMO gaps, and other key parameters. In addition, the MEP analysis identifies regions prone to nucleophilic and electrophilic interactions. Together, these findings thoroughly understand the compounds’ chemical behavior. The distinct electronic configurations and reactivity profiles highlight the interest of the designed compounds Pred3, Pred4, Pred34, and Pred49 in cancer treatment.
4. Conclusions
In this study, a 3D-QSAR study was conducted using the CoMFA and CoMSIA methods to create a QSAR model that relates the biological activity of a series of Quinazolin-4(3H)-one derivatives with their VEGFR2 kinase inhibitory activity. Model stability was evaluated through external validation, revealing that the optimized CoMSIA/SHA model has good reliability and acceptable predictive accuracy (Q2 = 0.717, R2 = 0.995, R2pred = 0.832). Consequently, 59 new and improved anticancer inhibitors were designed based on the key structure shown on the contour maps of the best-selected model (CoMSIA/SHA). Further, the docking results revealed that four newly designed compounds showed better binding affinity ranging from −8.5 to −11.5 kcal/mol to the selected RTKs and PFKFB3, which were implicated in cancer progression through angiogenesis and energy metabolism. Based on findings from the molecular docking study, all the best-docked complexes for each protein target underwent a 100 ns simulation, confirming their stability in the binding pocket of the selected targets. In addition, DFT calculation by the HOMO and LUMO energy gap and the MEP surface indicate that all four proposed compounds have acceptable electronic properties, making them good candidates for drug design.
This study presents a comprehensive computational workflow for designing and evaluating novel multitarget inhibitors targeting cancer by addressing RTK-driven angiogenesis and PFKFB3-mediated metabolic reprogramming. However, to ensure the clinical relevance of these compounds, additional in vitro and in vivo experiments are crucial to validate the therapeutic efficacy of these compounds in cancer treatment. These investigations will offer valuable understanding of their molecular mechanisms of inhibition and increase their possible use as novel anticancer drug candidates.
Acknowledgment
The authors express their gratitude to the College of Computing at Mohamed VI Polytechnic University for providing access to the supercomputing resources (cc.um6p.ma/toubkal-super-computer) used in conducting the research presented in this paper.
CRediT authorship contribution statement:
S. Baammi: Conceptualization, Data curation, Formal analysis, Methodology, Visualization, Writing – original draft, Writing – review & editing. A. El Allali: Conceptualization, Methodology, Supervision, Writing – original draft, Writing – review & editing. R. Daoud: Conceptualization, Methodology, Supervision, Writing – original draft, Writing – review & 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.
Data availability
The article and the supplementary file contain all available data. The corresponding author can be contacted for further inquiries.
Declaration of generative AI and AI-assisted technologies in the writing process
The authors confirm that there was no use of AI-assisted technology for assisting in the writing 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_159_2024.
References
- Vascular endothelial growth factor and angiogenesis in the regulation of cutaneous wound repair. Advances in Wound Care. 2014;3:647-661. https://doi.org/10.1089/wound.2013.0517
- [Google Scholar]
- Angiogenesis: Basic pathophysiology and implications for disease. European Heart Journal. 2003;24:586-603. https://doi.org/10.1016/s0195-668x(02)00635-8
- [Google Scholar]
- Angiogenesis in disease. International Journal of Molecular Sciences. 2022;23:10962. https://doi.org/10.3390/ijms231810962
- [Google Scholar]
- Folkman, J., Kalluri, R., 2003. Beginning of angiogenesis research, https://www.ncbi.nlm.nih.gov/books/NBK13877/ (Last accessed October 21, 2024).
- Anti-angiogenic drugs: Involvement in cutaneous side effects and wound-healing complication. Advances in Wound Care. 2014;3:635-646. https://doi.org/10.1089/wound.2013.0496
- [Google Scholar]
- Antiangiogenic cancer treatment: The great discovery and greater complexity (Review) International Journal of Oncology. 2016;49:1773-1784. https://doi.org/10.3892/ijo.2016.3709
- [Google Scholar]
- Role of receptor tyrosine kinases mediated signal transduction pathways in tumor growth and angiogenesis-new insight and futuristic vision. International Journal of Biological Macromolecules. 2021;180:739-752. https://doi.org/10.1016/j.ijbiomac.2021.03.075
- [Google Scholar]
- Therapeutic advances of targeting receptor tyrosine kinases in cancer. Signal Transduction and Targeted Therapy. 2024;9:201. https://doi.org/10.1038/s41392-024-01899-w
- [Google Scholar]
- Vascular endothelial growth factor (VEGF) and its receptor (VEGFR) signaling in angiogenesis: A crucial target for anti- and pro-angiogenic therapies. Genes & Cancer. 2011;2:1097-1105. https://doi.org/10.1177/1947601911423031
- [Google Scholar]
- Orally active 4-amino-5-diarylurea-furo[2,3-d]pyrimidine derivatives as anti-angiogenic agent inhibiting VEGFR2 and Tie-2. Bioorganic & Medicinal Chemistry Letters. 2007;17:1773-1778. https://doi.org/10.1016/j.bmcl.2006.12.077
- [Google Scholar]
- Vascular endothelial growth factor-A inhibits EphB4 and stimulates delta-like ligand 4 expression in adult endothelial cells. The Journal of Surgical Research. 2013;183:478-486. https://doi.org/10.1016/j.jss.2013.01.009
- [Google Scholar]
- Computational insights into VEGFR1 inhibitors: Redefining cancer treatment through dual-targeted therapy. Biointerface Research in Applied Chemistry. 2024;14:149. https://doi.org/10.33263/BRIAC146.149.
- [Google Scholar]
- Mechanisms of resistance to anti-angiogenesis therapies. Biochimie. 2013;95:1110-1119. https://doi.org/10.1016/j.biochi.2013.03.002
- [Google Scholar]
- Endothelial cell metabolism. Physiological Reviews. 2018;98:3-58. https://doi.org/10.1152/physrev.00001.2017
- [Google Scholar]
- Metabolic reprogramming and interventions in angiogenesis. Journal of Advanced Research. 2025;70:323-338. https://doi.org/10.1016/j.jare.2024.05.001
- [Google Scholar]
- Hallmarks of endothelial cell metabolism in health and disease. Cell Metabolism. 2019;30:414-433. https://doi.org/10.1016/j.cmet.2019.08.011
- [Google Scholar]
- Hypoxic regulation of the 6‐phosphofructo‐2‐kinase/fructose‐2,6‐bisphosphatase gene family (PFKFB‐1–4) expression in vivo. FEBS Letters. 2003;554:264-270. https://doi.org/10.1016/s0014-5793(03)01179-7
- [Google Scholar]
- The molecular basis of targeting PFKFB3 as a therapeutic strategy against cancer. Oncotarget. 2017;8:62793-62802. https://doi.org/10.18632/oncotarget.19513
- [Google Scholar]
- Partial inhibition of the 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase-3 (PFKFB3) enzyme in myeloid cells does not affect atherosclerosis. Frontiers in Cell and Developmental Biology. 2021;9:695684. https://doi.org/10.3389/fcell.2021.695684
- [Google Scholar]
- 3D-QSAR in drug design--a review. Current Topics in Medicinal Chemistry. 2010;10:95-115. https://doi.org/10.2174/156802610790232260
- [Google Scholar]
- ADMET modeling approaches in drug discovery. Drug Discovery Today. 2019;24:1157-1165. https://doi.org/10.1016/j.drudis.2019.03.015
- [Google Scholar]
- Design, synthesis, molecular modeling, in vivo studies and anticancer evaluation of quinazolin-4(3H)-one derivatives as potential VEGFR-2 inhibitors and apoptosis inducers. Bioorganic Chemistry. 2020;94:103422. https://doi.org/10.1016/j.bioorg.2019.103422
- [Google Scholar]
- Discovery of new quinazolin-4(3H)-ones as VEGFR-2 inhibitors: Design, synthesis, and anti-proliferative evaluation. Bioorganic Chemistry. 2020;105:104380. https://doi.org/10.1016/j.bioorg.2020.104380
- [Google Scholar]
- A series of quinazolin-4(3H)-one-morpholine hybrids as anti-lung-cancer agents: Synthesis, molecular docking, molecular dynamics, ADME prediction and biological activity studies. Chemical Biology & Drug Design. 2024;104:e14599. https://doi.org/10.1111/cbdd.14599
- [Google Scholar]
- Novel quinazoline–chromene hybrids as anticancer agents: Synthesis, biological activity, molecular docking, dynamics and ADME studies. Chemistry in Life Sciences. 2023;356 https://doi.org/10.1002/ardp.202300423
- [Google Scholar]
- Synthesis, characterization and molecular docking studies of highly selective new hydrazone derivatives of anthranilic acid and their ring closure analogue quinazolin-4(3H)-ones against lung cancer cells A549. Journal of Molecular Structure. 2023;1282:135176. https://doi.org/10.1016/j.molstruc.2023.135176
- [Google Scholar]
- Design, synthesis and molecular modeling of new quinazolin-4(3H)-one based VEGFR-2 kinase inhibitors for potential anticancer evaluation. Bioorganic Chemistry. 2021;109:104695. https://doi.org/10.1016/j.bioorg.2021.104695
- [Google Scholar]
- Design, molecular docking, in vitro, and in vivo studies of new quinazolin-4(3H)-ones as VEGFR-2 inhibitors with potential activity against hepatocellular carcinoma. Bioorganic Chemistry. 2021;107:104532. https://doi.org/10.1016/j.bioorg.2020.104532
- [Google Scholar]
- Antiproliferative, antiangiogenic and apoptotic effect of new hybrids of quinazoline-4(3H)-ones and sulfachloropyridazine. European Journal of Medicinal Chemistry. 2023;245:114912. https://doi.org/10.1016/j.ejmech.2022.114912
- [Google Scholar]
- Exploration of N-alkyl-2-[(4-oxo-3-(4-sulfamoylphenyl)-3,4-dihydroquinazolin-2-yl)thio]acetamide derivatives as anticancer and radiosensitizing agents. Bioorganic Chemistry. 2019;88:102956. https://doi.org/10.1016/j.bioorg.2019.102956
- [Google Scholar]
- Design, synthesis, and anti-proliferative evaluation of new quinazolin-4(3H)-ones as potential VEGFR-2 inhibitors. Bioorganic & Medicinal Chemistry. 2021;29:115872. https://doi.org/10.1016/j.bmc.2020.115872
- [Google Scholar]
- Antitumor agents. 213. Modeling of epipodophyllotoxin derivatives using variable selection k nearest neighbor QSAR method. Journal of Medicinal Chemistry. 2002;45:2294-2309. https://doi.org/10.1021/jm0105427
- [Google Scholar]
- 3D-QSAR, molecular docking and ADMET studies of thioquinazolinone derivatives against breast cancer. Journal of the Indian Chemical Society. 2022;99:100675. https://doi.org/10.1016/j.jics.2022.100675
- [Google Scholar]
- 3D-QSAR, docking and ADMET properties of aurone analogues as antimalarial agents. Heliyon. 2020;6:e03580. https://doi.org/10.1016/j.heliyon.2020.e03580
- [Google Scholar]
- In silico design of novel Pyrazole derivatives containing thiourea skeleton as anti-cancer agents using: 3D QSAR, Drug-Likeness studies, ADMET prediction and molecular docking. Materials Today: Proceedings. 2021;45:7661-7674. https://doi.org/10.1016/j.matpr.2021.03.152
- [Google Scholar]
- Rational design of novel potential EGFR inhibitors by 3D-QSAR, molecular docking, molecular dynamics simulation, and pharmacokinetics studies. Chemical Data Collections. 2022;39:100851. https://doi.org/10.1016/j.cdc.2022.100851
- [Google Scholar]
- Studies on the antibacterial activities and molecular mechanism of GyrB inhibitors by 3D-QSAR, molecular docking and molecular dynamics simulation. Arabian Journal of Chemistry. 2022;15:103872. https://doi.org/10.1016/j.arabjc.2022.103872
- [Google Scholar]
- Molecular docking, dynamics simulations and 3D-QSAR modeling of arylpiperazine derivatives of 3,5-dioxo-(2H,4H)-1,2,4-triazine as 5-HT1AR agonists. Computational Biology and Chemistry. 2019;78:108-115. https://doi.org/10.1016/j.compbiolchem.2018.11.015
- [Google Scholar]
- Development, validation, and use of quantitative structure-activity relationship models of 5-hydroxytryptamine (2B) receptor ligands to identify novel receptor binders and putative valvulopathic compounds among common drugs. Journal of Medicinal Chemistry. 2010;53:7573-7586. https://doi.org/10.1021/jm100600y
- [Google Scholar]
- The importance of being earnest: Validation is the absolute essential for successful application and interpretation of QSPR models. QSAR & Combinatorial Science. 2003;22:69-77. https://doi.org/10.1002/qsar.200390007
- [Google Scholar]
- UCSF Chimera--a visualization system for exploratory research and analysis. Journal of Computational Chemistry. 2004;25:1605-1612. https://doi.org/10.1002/jcc.20084
- [Google Scholar]
- Potent VEGFR-2 inhibitors for resistant breast cancer: A comprehensive 3D-QSAR, ADMET, molecular docking and MMPBSA calculation on triazolopyrazine derivatives. Frontiers in Molecular Biosciences. 2023;10:1288652. https://doi.org/10.3389/fmolb.2023.1288652
- [Google Scholar]
- PyMOL | pymol.org, (n.d.). https://www.pymol.org/ (last accessed October 21, 2024).
- Study on the solubilization effect of 7-ethyl-10-hydroxycamptothecin based on molecular docking and molecular dynamics simulation. Journal of Molecular Modeling. 2023;29:58. https://doi.org/10.1007/s00894-023-05455-1
- [Google Scholar]
- Implementation of the CHARMM force field in GROMACS: Analysis of protein stability effects from correction maps, virtual interaction sites, and water models. Journal of Chemical Theory and Computation. 2010;6:459-466. https://doi.org/10.1021/ct900549r
- [Google Scholar]
- SwissParam: A fast force field generation tool for small organic molecules. Journal of Computational Chemistry. 2011;32:2359-2368. https://doi.org/10.1002/jcc.21816
- [Google Scholar]
- Molecular dynamics simulation, free energy landscape and binding free energy computations in exploration the anti-invasive activity of amygdalin against metastasis. Computer Methods and Programs in Biomedicine. 2020;195:105660. https://doi.org/10.1016/j.cmpb.2020.105660
- [Google Scholar]
- Molecular docking, ADME-Tox, DFT and molecular dynamics simulation of butyroyl glucopyranoside derivatives against DNA gyrase inhibitors as antimicrobial agents. Journal of Molecular Structure. 2024;1307:137930. https://doi.org/10.1016/j.molstruc.2024.137930
- [Google Scholar]
- Computational evaluation of 1,2,3-triazole-based VEGFR-2 inhibitors: Anti-angiogenesis potential and pharmacokinetic assessment. Journal of Biomolecular Structure & Dynamics. 2025;43:2549-2559. https://doi.org/10.1080/07391102.2023.2301686
- [Google Scholar]
- Towards designing of a potential new HIV-1 protease inhibitor using QSAR study in combination with molecular docking and molecular dynamics simulations. PloS One. 2023;18:e0284539. https://doi.org/10.1371/journal.pone.0284539
- [Google Scholar]
- In silico design of novel CDK2 inhibitors through QSAR, ADMET, molecular docking and molecular dynamics simulation studies. Journal of Biomolecular Structure & Dynamics. 2023;41:13646-13662. https://doi.org/10.1080/07391102.2023.2212304
- [Google Scholar]
- Multi-combined QSAR, molecular docking, molecular dynamics simulation, and ADMET of flavonoid derivatives as potent cholinesterase inhibitors. Journal of Biomolecular Structure & Dynamics. 2024;42:6027-6041. https://doi.org/10.1080/07391102.2023.2238314
- [Google Scholar]
- 3D-QSAR, molecular docking, simulation dynamic and ADMET studies on new quinolines derivatives against colorectal carcinoma activity. Journal of Biomolecular Structure & Dynamics. 2024;42:3682-3699. https://doi.org/10.1080/07391102.2023.2214233
- [Google Scholar]
- Antibacterial study of 3-(2-amino-6-phenylpyrimidin-4-yl)-N-cyclopropyl-1-methyl-1H-indole-2-carboxamide derivatives: CoMFA, CoMSIA analyses, molecular docking and ADMET properties prediction. Journal of Molecular Structure. 2019;1177:275-285. https://doi.org/10.1016/j.molstruc.2018.09.073
- [Google Scholar]
- pkCSM: Predicting small-molecule pharmacokinetic and toxicity properties using graph-based signatures. Journal of Medicinal Chemistry. 2015;58:4066-4072. https://doi.org/10.1021/acs.jmedchem.5b00104
- [Google Scholar]
- Prediction of human intestinal absorption by GA feature selection and support vector machine regression. International Journal of Molecular Sciences. 2008;9:1961-1976. https://doi.org/10.3390/ijms9101961
- [Google Scholar]
- Metabolism and pharmacokinetics characterization of metarrestin in multiple species. Cancer Chemotherapy and Pharmacology. 2020;85:805-816. https://doi.org/10.1007/s00280-020-04042-y
- [Google Scholar]
- Pharmacokinetic and molecular docking studies to pyrimidine drug using Mn3O4 nanoparticles to explore potential anti-Alzheimer activity. Scientific Reports. 2024;14:15436. https://doi.org/10.1038/s41598-024-65166-2
- [Google Scholar]
- Role of the blood-brain barrier in central nervous system insulin resistance. Frontiers in Neuroscience. 2019;13:521. https://doi.org/10.3389/fnins.2019.00521
- [Google Scholar]
- A blood–brain barrier overview on structure, function, impairment, and biomarkers of integrity. Fluids Barriers CNS. 2020;17:69. https://doi.org/10.1186/s12987-020-00230-3
- [Google Scholar]
- CuO nanoparticles for green synthesis of significant anti-helicobacter pylori compounds with in silico studies. Scientific Reports. 2024;14:1608. https://doi.org/10.1038/s41598-024-51708-1
- [Google Scholar]
- Stanley, L. A. Chapter 27 - Drug Metabolism; Badal, S., Rupika Delgoda, Eds.; Academic Press, 2017; pp. 527–545. https://doi.org/10.1016/B978-0-12-802104-0.00027-5
- Pharmacodynamics. In: Levine B.S., Kerrigan S., eds. Principles of Forensic Toxicology. Cham: Springer; 2020. https://doi.org/10.1007/978-3-030-42917-1_8
- [Google Scholar]
- Comparative analysis of substrate and inhibitor interactions with CYP3A4 and CYP3A5. Xenobiotica; the Fate of Foreign Compounds in Biological Systems. 2006;36:287-299. https://doi.org/10.1080/00498250500446208
- [Google Scholar]
- Mechanisms and consequences of drug-drug interactions. Clinical Pharmacology in Drug Development. 2017;6:118-124. https://doi.org/10.1002/cpdd.339
- [Google Scholar]
- The test that changed the world: The AMES test and the regulation of chemicals. Mutation research. Genetic toxicology and Environmental Mutagenesis. 2019;841:43-48. https://doi.org/10.1016/j.mrgentox.2019.05.007
- [Google Scholar]
- The potential role of in silico approaches to identify novel bioactive molecules from natural resources. Future medicinal Chemistry. 2017;9:1665-1686. https://doi.org/10.4155/fmc-2017-0124
- [Google Scholar]
- Is it reliable to take the molecular docking top scoring position as the best solution without considering available structural data? Molecules (Basel, Switzerland). 2018;23:1038. https://doi.org/10.3390/molecules23051038
- [Google Scholar]
- Structure-based molecular networking, molecular docking, dynamics simulation and pharmacokinetic studies of olax subscorpioidea for identification of potential inhibitors against selected cancer targets. Journal of Biomolecular Structure & Dynamics. 2024;42:1110-1125. https://doi.org/10.1080/07391102.2023.2198032
- [Google Scholar]
- Multi-target pharmacology: Possibilities and limitations of the “skeleton key approach” from a medicinal chemist perspective. Frontiers in Pharmacology. 2015;6 https://doi.org/10.3389/fphar.2015.00205
- [Google Scholar]
- Molecular Dynamics Simulation to Study Protein Conformation and Ligand Interaction. In: Saudagar P., Tripathi T., eds. Protein Folding Dynamics and Stability. Singapore: Springer; 2023. https://doi.org/10.1007/978-981-99-2079-2_6
- [Google Scholar]
- QSAR, ADMET, molecular docking, and dynamics studies of 1,2,4-triazine-3(2H)-one derivatives as tubulin inhibitors for breast cancer therapy. Scientific Reports. 2024;14:16418. https://doi.org/10.1038/s41598-024-66877-2
- [Google Scholar]
- Structure-based computational screening of 470 natural quercetin derivatives for identification of SARS-CoV-2 Mpro inhibitor. PeerJ. 2023;11:e14915. https://doi.org/10.7717/peerj.14915
- [Google Scholar]
- Phytochemical analysis and evaluation of antioxidant and antimicrobial properties of essential oils and seed extracts of Anethum graveolens from southern morocco: in vitro and in silico approach for a natural alternative to synthetic preservatives. Pharmaceuticals (Basel, Switzerland). 2024;17:862. https://doi.org/10.3390/ph17070862
- [Google Scholar]
- Unleashing Nature’s potential: A computational approach to discovering novel VEGFR-2 inhibitors from African natural compound using virtual screening, ADMET analysis, molecular dynamics, and MMPBSA calculations. Frontiers in Molecular Biosciences. 2023;10:1227643. https://doi.org/10.3389/fmolb.2023.1227643
- [Google Scholar]
- Structural and free energy landscape of novel mutations in ribosomal protein S1 (rpsA) associated with pyrazinamide resistance. Scientific Reports. 2019;9:7482. https://doi.org/10.1038/s41598-019-44013-9
- [Google Scholar]
- High-throughput virtual screening approach of natural compounds as target inhibitors of plasmepsin-II. Journal of Biomolecular Structure & Dynamics. 2023;41:10070-10080. https://doi.org/10.1080/07391102.2022.2152871
- [Google Scholar]
- Synthesis, characterization, DFT and TD-DFT study of novel bis(5,6-diphenyl-1,2,4-triazines) Journal of Molecular Structure. 2021;1226:129345. https://doi.org/10.1016/j.molstruc.2020.129345
- [Google Scholar]
- Evaluating frontier orbital energy and HOMO/LUMO gap with descriptors from density functional reactivity theory. Journal of molecular modeling. 2017;23:3. https://doi.org/10.1007/s00894-016-3175-x
- [Google Scholar]
- Predicting the degradation potential of Acid blue 113 by different oxidants using quantum chemical analysis. Heliyon. 2019;5:e02396. https://doi.org/10.1016/j.heliyon.2019.e02396
- [Google Scholar]
- A drug design strategy based on molecular docking and molecular dynamics simulations applied to development of inhibitor against triple-negative breast cancer by scutellarein derivatives. PloS One. 2023;18:e0283271. https://doi.org/10.1371/journal.pone.0283271
- [Google Scholar]
- DFT, molecular docking and molecular dynamics simulation studies on some newly introduced natural products for their potential use against SARS-CoV-2. Journal of Molecular Structure. 2021;1242:130733. https://doi.org/10.1016/j.molstruc.2021.130733
- [Google Scholar]
- Synthesis, crystal structure, DFT, Hirshfeld surface analysis, energy frameworks and in-Silico drug-targeting PFKFB3 kinase of novel triazolequinoxalin derivative (TZQ) as a therapeutic Strategy against cancer. Heliyon. 2023;9:e21312. https://doi.org/10.1016/j.heliyon.2023.e21312
- [Google Scholar]
- Computer-aided discovery of bis-indole derivatives as multi-target drugs against cancer and bacterial infections: DFT, docking, virtual screening, and molecular dynamics studies. Journal of Molecular Liquids. 2020;320:114375. https://doi.org/10.1016/j.molliq.2020.114375
- [Google Scholar]
