Introduction
Rheumatoid arthritis (RA) is a complex autoimmune condition involving multiple genetic polymorphisms and is associated with a chronic inflammatory reaction in the joints and probable association with destructive bone erosion [1]. The prevalence of this condition is estimated to impact between 0.1% to 2.0% of the global population [2]. RA’s aetiology is mostly distinguished by synovial proliferation and inflammation, auto-antibody production, cartilage, and bone degradation. Tumor Necrosis Factor-alpha (TNF- α), Interleukin-1 (IL-1), and IL-6 are three key pro-inflammatory cytokines that are responsible for RA, among which TNF-α has been recognized as a crucial modulator of inflammatory reactions [3]. Secreted by Th1 cells and macrophages, TNF-α was initially thought to play a synergistic role in the development of RA [4] However, follow-up studies indicated that excessive activation of TNF-α signaling led to arthritis even in the absence of functional T and B cells [5]. Later studies found that even the membrane-bound form of TNF-α (mTNF-α) can lead to the full expression of arthritis [6, 7]. Additionally, synovial fibroblasts, activated by TNF-α, release pro-inflammatory cytokines such as IL-6, IL-1β which further accelerates bone erosion [7]. The multi-directional involvement of TNF- α in RA pathogenesis has made it a prime target for managing the disease’s systemic effects as there is no recognized curative treatment currently. In recent years, a multifaceted treatment regimen has successfully decreased disease activity and alleviated the onset of systemic effects. Disease-modifying anti-rheumatic drugs (DMARDs), encompassing conventional synthetic, biologic, and targeted synthetic medications, demonstrate promising results but are associated with adverse effects. Biological DMARDs are expensive and impose a significant financial burden on the patients [8]. Additionally, conventional synthetic DMARDs, for instance, methotrexate, are associated with significant side effects, including infections, malignancies, hepatotoxicity, and teratogenesis in pregnancy [9]. Conventional non-steroidal anti-inflammatory drugs (NSAIDs) like naproxen and celecoxib can provide temporary pain relief but fall short in slowing disease progression [10].
The continuous progress in medical research has resulted in the acquisition of several bioactive molecules derived from medicinal plants that exhibit enhanced effectiveness and safety characteristics [11] against TNF-α. Triptolide, a diterpenoid epoxide produced by the plant Tripterygium wilfordii Hook F, demonstrated significant pharmacological effects particularly immunosuppressive impacts in the treatment of RA [12]. Another study focused on Paeonia lactiflora Pallas for its anti-inflammatory, palliative, and immune system-enhancing properties, utilizing the isolated total glycoside of paeony (TGP) from its roots [13]. The authors observed a therapeutic response in 71.7 percent of TGP-treated patients. Various other molecules such as Aloe barbadensis, Alstonia scholarisa, Berberis lycium Royle, Cinnamomum verum, Citrus limon, Coriander sativum, Curcuma longathe, etc. have been used as herbal interventions for the management of RA [14]. Natural substances also have drawbacks. Triptolide modulates immune cell activity and cytokine expression, but it has been associated with hepatotoxicity, nephrotoxicity, reproductive system toxicity, cardiotoxicity, splenic toxicity, pulmonary toxicity, and gastrointestinal tract toxicity [15]. Additionally, the quest for natural compounds with enhanced efficacy and assured safety has predominantly relied on manual methods, particularly literature reviews. This approach has overlooked many natural drug candidates that can potentially treat RA having minimal toxicity. Unfortunately, these candidates are not studied in a structured manner, limiting their potential application in clinical cases. In summary, manual screening limits the discovery of entire sets of natural ligands or inhibitors against specific proteins or receptors.
Virtual screening techniques, computational assessment of drug molecules against a particular target based on structural similarity, are currently used in the industry for rapid drug discovery [16]. More recently, a ligand-derived active pharmacophore model against TNF-α was generated to screen an in-house database consisting of 10,000 synthetic and natural compounds [17]. The authors identified five compounds based on computational screening and narrowed them down to three potential inhibitors after in-vitro validation and cytotoxicity assays. However, the authors determined the pharmacophore structure based on only 26 known ligands against TNF-α. This small number may preclude important structural features from the model and prevent an exhaustive search for all possible inhibitors. Additionally, their work was not tailored exclusively to plant-derived ligands as their database contained a mixture of synthetic and natural compounds. Deep learning (DL) based virtual screening models can handle large volumes of structural data and can provide more accurate predictions than the conventional pharmacophore technique. A virtual screening approach to precisely find natural therapeutic agents against the SARS-CoV-2 main protease (Mpro) protein through Selleckchem was established using a DL model [18]. The study identified two compounds, Palmatine and Sauchinone, as promising inhibitors of Mpro protein. Such methodologies can be adapted to find a comprehensive list of plant-derived bioactive molecules targeting TNF-α. This paper presents a DL-based novel computational screening methodology to assess the potential of natural ligands in inhibiting the TNF-α protein. We deployed our model on the Selleckchem database which is a curated natural product library. The best performing TNF-α inhibitors are validated through binding affinity analysis and Molecular Dynamics (MD) simulation studies. A complete workflow is described in Fig 1. Moreover, the DL-based virtual screening approach can be integrated with in-vivo and in-vitro assays to create a complete drug discovery pipeline for any protein/target of interest.
[Figure omitted. See PDF.]
The model was developed using the ChEMBL dataset, with 2AZ5 as the target protein. After validation, the model was used to screen the Selleckchem Natural Compounds database. Compounds were prioritized based on their predicted pIC50 values, and after additional filtering, four compounds were chosen for MD simulation studies.
Materials and methods
Refining target protein—TNF-α (2AZ5)
TNF-α’s three-dimensional conformation was obtained from the Protein Data Bank [19]. The PDB ID: 2AZ5 structure of the TNF-alpha protein has been extensively studied in literature for different virtual screening studies [20]. The structure is complexed with a small molecule inhibitor and has become a benchmark for drug discovery studies aiming to inhibit the inflammatory activity of the TNF-alpha protein. The structure [21] was found to be incomplete due to the absence of protein residues which is summarized in S1 Table and visually depicted in S1 Fig. PDBsum [22] and the FirstGlance in the Jmol web server [23] aided in visualizing the absent regions. The SWISS-MODEL server [24] was utilized for homology modeling. This method utilizes the structural similarity of a closely related protein to predict the 3D structure of an unspecified protein. The FASTA sequence of 2AZ5 was uploaded to the server and the template search resulted in the identification of 2E7A [25], which displayed the highest similarity to the structure of 2AZ5. Furthermore, it outperformed other protein models in terms of evaluation metrics provided by SWISS-MODEL. The resulting model had GMQE score of 0.87, QMEANDisCO (Global) score of 0.82±0.05, and Ramachandran outlier of 0.46%, confirming the validity of its structure.
Deep learning: Advanced predictive modeling
In our predictive DL model, ChEMBL ID-1825 [26] which corresponded to compounds against our target protein (2AZ5), served as a training dataset and contained 953 compounds in total. The dataset underwent filtration based on protein compatibility and reported IC50 values, resulting in the retention of compounds having a set of unique canonical SMILES [27]. A negative logarithmic conversion was used to create a more uniformly distributed collection of IC50 data, resulting in pIC50. The PaDEL software [28] was employed to generate PubChem fingerprints [29] from the canonical SMILES. PubChem fingerprints contained 881 binary descriptors which indicate the existence of specific chemical compound groups [30]. To enhance the training dataset, we implemented variance thresholding to retain essential descriptors and reduced the amount from 881 to 235 in the training dataset. The same pre-processing techniques were applied during testing on the Selleckchem library [31]. The variance thresholding on the test dataset resulted in 251 essential descriptors. To ensure data uniformity, the descriptors from the train and test datasets were merged, and compared, and the repeating ones were eliminated to preserve 342 binary descriptors. The data pre-processing pipeline in both the train and test cases is highlighted in Fig 2.
[Figure omitted. See PDF.]
Panel A represents the characterization of compounds using PubChem descriptors (881) of both training (ChEMBL) and testing (Selleckchem) datasets and Panel B depicts variance thresholding to retain essential features.
We used RandomizedSearchCV for hyperparameter tuning [32] to find the best combination of parameters that optimize the given performance metrics. Finally, the best model incorporating the following parameters mentioned in Table 1 was later deployed on the test dataset consisting of 2563 compounds and the same 342 binary descriptors.
[Figure omitted. See PDF.]
We modified the model’s structure by modifying hyperparameters including the number of neurons, activation functions, and initializers for each hidden layer to match the data’s individual properties. In the hidden layer, certain activation functions are adept at managing non-linearities, whereas other different activation functions are more effective at handling negative values, while the output layer uses a linear function. Choosing a weight initializer and a learning rate of 0.0001 helped our model converge and train faster without sacrificing stability. The "Adam" optimizer effectively adjusted the weights of the neural network while training.
Optimizing models: Virtual screening assessment
The predictive model was assessed and reconstructed employing several statistical metrics. To evaluate the model’s performance, we employed Mean squared error (MSE), Mean absolute percentage error (MAPE), and Mean absolute error (MAE) [33], which are employed to quantify the level of error present in statistical models. The regression model was utilized to analyze the Selleckchem natural product library, a subset of the Selleck database consisting of 2563 (July 2023) natural compounds, for virtual screening.(1)(2)(3)k is the number of observations, yi represents the actual observed values, represents the estimated values.
Active site and binding site confirmation
The protein structure was analyzed using the CASTp [34] and DeepSite [35] web servers to determine its topological properties and predict the binding sites. This allowed us to identify and describe surface pockets and clefts, which are crucial for understanding the functional areas of the protein. These tools not only annotated the predicted binding pockets but also provided information regarding the involved interacting amino acid residues. The residues are represented in S2 and S3 Tables and S2 Fig.
Drug likeness analysis
Following the virtual screening, the retrieved compounds underwent additional refinement by assessing their drug-likeness utilizing the SwissADME web server [36], which offers Lipinski filter, Veber filter, Ghose filter, Egan filter, and so on. These filters helped ensure that the selected compounds met specific criteria associated with drug-likeness [37].
Molecular docking of screened compounds
Autodock Vina was utilized to conduct molecular docking between 68 screened compounds and the target protein [38]. Kollman charges and Polar hydrogen atoms were introduced into the 3D structure of the receptor during preprocessing. While merging non-polar hydrogen atoms in the ligand structures, the ligands underwent Gasteiger partial charges. To facilitate docking, a three-dimensional grid box was positioned at X = -4.173, Y = -6.686, and Z = 2.774 grid points. The grid spacing for the X, Y, and Z coordinates was set at 22 x 22 x 22 Å, respectively. For predicting the accurate result, the number of exhaustiveness was set to eight.
ADMET analysis
The ADMET attributes, encompassing absorption, distribution, metabolism, excretion, and toxicity, were performed using the ADMETlab 2.0 [39] designed for systematic study of compounds’ ADMET properties. It utilizes a vast ADMET database with over 280,000 entries. This platform provides support for the calculation of 88 ADMET-related endpoints across seven distinct categories.
Retrospective validation of virtual screening protocol
Retrospective validation based on ligand shapes, fingerprints, and binding affinities have been reported in prior literature to distinguish between active ligands and decoy molecules [40, 41]. Previous reports have shown that validation based on 2D fingerprints similarity is fast and have shown similar or superior performance compared to the 3D shape based methods [42]. Therefore, we assessed our virtual screening using molecular fingerprints and utilized a subset of the DUD-E [41, 43] (Directory of Useful Decoys, Enhanced) database, which is widely recognized as a benchmark for evaluating virtual screening methodologies. To mitigate bias between active compounds and decoys, the decoys were selected based on similar physical properties but distinctive topological features. DUD-E is specifically designed to evaluate both structure-based and 3D ligand-based virtual screening techniques, with decoys deliberately chosen to have low 2D similarity to the active compounds [43]. We selected the top 20 active compounds for validation and generated 25 decoys for each of the selected active compounds.
The validation process employed two types of molecular fingerprint descriptors, Morgan and Layered fingerprints [44, 45], both calculated using the RDKit Python package. Morgan fingerprints were generated with a radius of 2 and a fingerprint size of 2048 bits, while Layered fingerprints were computed with a minimum path length of 1, a maximum path length of 9, and a fingerprint size of 2048 bits. For benchmarking purposes, we used 500 decoy molecules from the DUD-E dataset and included Schisantherin A [46], a natural TNF-alpha inhibitor, as a reference control and for internal thresholding. The anti-inflammatory activity of Schisantherin A against TNF-alpha is well-documented in both computational and in-vitro studies [47]. ROC plot showing the relationship between the True Positive Rate (TP) and False Positive Rate (FPR) was constructed for both the descriptors. Additionally, we calculated Enrichment Factor at top 1% based on fingerprint descriptors. All the active and decoy molecules also underwent molecular docking by the method outlined previously. The molecules were then ranked based on their binding affinity scores and the Enrichment Factor at top 1% was calculated to assess model performance in distinguishing between active and decoy compounds.
MD simulation and post MD simulation
The GROMACS 23.1 [48, 49] package was utilized to perform molecular dynamics simulations, where CHARMM 36 force field [50] was applied to generate topology files utilizing the CGenFF server and pdb2gmx for both the ligand and protein. A dodecahedral system with periodic boundary conditions was constructed using box vectors of length (4.517, 6.309, 5.533 nm). The TIP3P water model was used to solvate the system and neutralize it by adding 4 positive sodium (Na+) ions. Each solute within the box was positioned at the center, ensuring a minimum distance of 1 nm from the edge of the box. A collective total of 50,000 steps were taken by all systems during energy minimization and was conducted at 10 KJ/mol using the steepest descent algorithm with a Verlet cut-off scheme [51] and Particle Mesh Ewald (PME) [52] for Coulombic interactions. Subsequently, position restraints were applied during the equilibration step. The system underwent NVT equilibration and NPT equilibration at 300 K for 200 picoseconds (ps) [53]. Both systems maintained a constant number of particles (N) and temperature (T). The only distinction is that the volume of the NVT system remains constant, while the pressure of the NPT system remains constant. Finally, molecular dynamics simulation for apo-protein and protein-ligand complexes was executed for 200 nanoseconds (ns) with a time interval of 2 femtoseconds (fs). The generated trajectories were analyzed for root mean square deviation (RMSD), root mean square fluctuation (RMSF), radius of gyration (Rg), hydrogen bonding (H-Bonds), and solvent-accessible surface area (SASA), protein-ligand complexes using GROMACS applications.
Principal component analysis (PCA) and dynamic cross-correlation matrix (DCCM)
The protein’s functional activity was analyzed by studying its dynamic motions using Principal Component Analysis (PCA) [54] on the last 50 ns of MD simulation trajectories, using the Bio3d [55] package. PCA reduces the complexity of protein motion data by computing eigenvectors and eigenvalues. These are then used to identify and select different modes of motion, facilitating a more comprehensive understanding of the underlying dynamics [56]. Dynamic cross-correlation matrix (DCCM) [57] represents movements within a protein involving the coordinated shifts of α-carbon backbone atoms, influencing the transition of protein conformations between different functional states. Both PCA and DCCM analyses were executed on aligned frames of C-α atoms from the respective MD trajectories, employing the fit.xyz function within the Bio3d package.
Binding free energy calculation using gmx_MMGBSA
The computational method known as Molecular Mechanics Generalized Born Surface Area (MM/GBSA) [58, 59] is extensively employed in the estimation of the binding free energy of protein-ligand complexes. In this study, the gmx_MMGBSA [59] package was employed for conducting free energy calculations, utilizing a single trajectory generated by GROMACS with the CHARMM-36 force field. The evaluation was carried out using the following equation:(4)(5)(6)(7)(8)
In Eq (4), GComplex represents the total binding energy of the complex, while Greceptor and Gligand denote the energies associated with the receptor and ligand, respectively. Further elucidation of the changes in conformational entropy (–TΔS) and enthalpy (ΔH) upon ligand binding is provided by additional Eqs (5–8). Here, ΔEMM, TΔS, and ΔGsolv, represent alterations in gas-phase molecular mechanics energy, conformational entropy, and solvation free energy, respectively. Additionally, ΔEbond stands for the energy of bonded interactions, assumed to be zero in dynamic simulations, while ΔEvdW, and ΔEelec signify van der Waals and electrostatic interaction energies, respectively. Determination of the solvation free energy (ΔGsolv) involves summing polar ΔGGB (electrostatic solvation energy) and non-polar ΔGSA between the solute and the continuum solvent. The non-polar component is assumed to be proportional to the molecule’s solvent-accessible surface area (SASA). The impact of entropy (ΔS) on the overall energy is negligible when comparing the binding states of ligands with the protein, hence it may be disregarded. Parameters utilized in this study include igb = 8, saltcon = 0.150, and intdiel = 8.
Results
Deep learning: Advancing predictive modeling
In the pursuit of constructing a robust regression model, we undertook an extensive approach utilizing PubChem fingerprint assays and implementing various preprocessing steps to establish standard values. To accurately assess DL models’ performance and maintain their unbiased attitude toward the data, multiple evaluation metrics were implemented and yielded good performance with MSE (0.6), MAPE (10%), MAE (0.5), and the total time utilized for this model was 82.61 seconds. From Fig 3, it can be observed that there is little to less overfitting occurring between the training and testing set. The ChEMBL dataset was split into training and testing sets in an 80:20 ratio. Implementing dropout rates for each layer and hyper-parameterizing regularization terms: L1 (0.00113) and L2 (0.00015), proved effective in counteracting overfitting. The strategic application of the early stopping callback method, terminating training after 120 epochs, also helped in preventing overfitting by discerning the optimal point of model proficiency on the testing set. After this, the DL model was employed to screen a subset of the Selleckchem database, which encompasses 2563 natural compounds.
[Figure omitted. See PDF.]
(a) MSE, (b) MAPE, (c) MAE, and (d) Loss.
The predicted model yielded a mean predicted pIC50 value of 5.276, with a range between 4.64 and 7.67. Compounds falling into the upper fourth quartile based on predicted values were exclusively selected. This criterion led to the inclusion of 186 compounds meeting the specified requirements. Despite their elevated PIC50 values, the majority of these compounds were excluded due to their inadequacy in Drug-likeness analysis. As a result, only 68 compounds were chosen for subsequent and more thorough investigation. The structures of 12 of the predicted compounds are depicted in Fig 4 along with their pIC50 values.
[Figure omitted. See PDF.]
The structures and the pIC50 values of the eligible candidates are included (CID 442977: Imperialine, CID 6070: Veratramine, CID 10098: Jervine, and CID 5390854: Gelsemine).
Molecular docking
In the course of the investigation, molecular docking studies were conducted on 68 compounds, against our target protein resulting in binding energies between -10.4 and -5.1 kcal mol-1. Compounds chosen were docked into the active site of 2AZ5 utilizing AutoDock Vina, which predicts the optimal binding mode and reduced binding energy. Docking scores and pIC50 values of 68 compounds are summarized in Fig 5.
[Figure omitted. See PDF.]
ADMET analysis and drug-likeness analysis of final 4 compounds
ADME and toxicology properties.
High docking scores suggest a strong binding affinity between proteins and ligands. In total 23 Compounds were selected within the docking score range of -10.4 to -8.7 and subjected to ADMET analysis. Despite their high docking scores, a substantial proportion of the compounds exhibited harmful effects according to the ADMET evaluation. The identified cohort included only four compounds—Imperialine, Veratramine, Jervine, and Gelsemine—demonstrating commendable properties presented in Table 2.
[Figure omitted. See PDF.]
In terms of absorption, all compounds were found to have better membrane permeability (Caco-2) around or greater than -5.15 log cm/s. Imperialine and Gelsemine act as P-gp substrates rather than P-gp inhibitors and contribute to the drug elimination process facilitated by enzymes. On the other hand, Veratramine and Jervine act as both P-gp inhibitors and substrates. The dual role of being both a substrate and an inhibitor may result in complex interactions with other drugs, affecting their bioavailability and therapeutic effects. In terms of Human Intestinal Absorption (HIA), molecules with HIA >30% were categorized as HIA- (Category 0) and those with HIA < 30% as HIA+ (Category 1). All our compounds belong to Category 0, signifying elevated HIA with values approaching zero. Drugs exerting their effects within the central nervous system (CNS) necessitate traversal of the blood–brain barrier to access their molecular targets. Apart from Gelsemine, the remaining three compounds exhibited satisfactory BBB values within the range of 0 to 0.7. Imperialine displayed an exceptional clearance rate exceeding 15 ml/min/kg, while veratramine, Jervine, and Gelsemine exhibited moderate clearance. Despite differences in clearance rates, all compounds showed a moderate half-life, reflecting a balanced pharmacokinetic profile with neither excessively rapid nor prolonged elimination.
Toxicology.
From a quantitative perspective, four characteristics were predicted: hepatotoxicity (H-HT), carcinogenicity, AMES toxicity, and HERG blockers of acute toxicity. The hERG gene, responsible for encoding a potassium ion channel, is more strongly linked to severe cardiotoxicity. The AMES test is utilized to determine the mutagenic characteristics of a chemical, specifically its capacity to cause genetic harm and mutations. Upon comprehensive toxicity analysis, it was found that only the above four compounds exhibited favorable results across the majority of the evaluated properties, indicating a promising profile for further development. Conversely, those compounds that did not meet the established safety thresholds displayed multiple adverse toxicity indicators, such as Sanguinarine showcasing -10.1 kcal/mol binding affinity exhibited AMES toxicity, carcinogenicity along with respiratory toxicity, and many more, necessitating its exclusion from consideration. This rigorous selection process was crucial in ensuring the identification of potential candidates with the highest likelihood of safety and efficacy for further investigation and development as natural products.
Drug likeness analysis.
The compounds that successfully passed ADMET analysis also showed promising drug-likeness results. Among the four compounds, Imperialine was reported to have more than 70 atoms, not adhering to the Ghose filter, but complying with the other four filters. However, all of our test compounds followed the Lipinski rule of five, ensuring proper absorption or permeation and thus qualifying as potential drug candidates against our target protein. The summary of Drug Likeness results is given in Table 3. The docked poses of these 4 compounds are represented in Fig 6.
[Figure omitted. See PDF.]
(a) Imperialine, (b) Veratramine, (c) Jervine, and (d) Gelsemine.
[Figure omitted. See PDF.]
In the context of protein-ligand interactions, it is apparent that Imperialine, Veratramine, Jervine, and Gelsemine exhibited binding affinity comprising the active site residues of 2AZ5. To gain a better understanding of the binding mechanisms of ADMET-approved compounds within the protein’s active sites, we conducted a 2D interaction analysis for the docked complexes as depicted in Fig 7.
[Figure omitted. See PDF.]
(a) Imperialine, (b) Veratramine, (c) Jervine, and (d) Gelsemine.
Pertaining to the protein-ligand complex, the equilibrium among various bond types, such as carbon-hydrogen bonds, alkyl bonds, and conventional hydrogen bonds, plays a pivotal role in influencing conformational dynamics, affinity, and biological function. Given the inherently weaker nature of carbon-hydrogen bonds in comparison to conventional hydrogen bonds, the stability of a ligand is anticipated to increase with a greater prevalence of conventional hydrogen bonds in the complex. Among the four selected drug candidates, Imperialine exhibited the formation of two conventional hydrogen bonds with ARG94 and LYS103, in conjunction with a carbon-hydrogen bond with CYS92. Additionally, it established four alkyl hydrophobic interactions with CYS92, PRO97, and ARG94, and thus contributed to enhanced structural flexibility. The 2AZ5-Veratramine complex displayed a diverse array of bond types, manifesting attractive charge interactions with GLU107 and amide-pi stacked bonds with PRO91. The pi-stacked interactions with the aromatic amino acid residues in the protein influenced the specificity of the binding. This complex further exhibited five alkyl bonds with CYS92, CYS60, and LYS103 which further stabilized the complex. Notably, two conventional hydrogen bonds with LYS89 and GLU101, along with a carbon-hydrogen bond with SER90, fortified the interaction profile. In the case of the 2AZ5-Jervine complex, alkyl-pi interactions, and a non-covalent bond mediated by van der Waals forces, contributed to heightened durability and prevented ligand dissociation. This complex was stabilized by one conventional hydrogen bond with GLU101, along with six alkyl and pi-alkyl hydrophobic interactions involving TRP105, CYS92, LYS103, and CYS60. Additionally, a carbon-hydrogen bond with CYS92 further enriched the interaction landscape. Conversely, the 2AZ5-Gelsemine complex diverged from the aforementioned patterns, featuring two carbon-hydrogen bonds with ARG94 and CYS92, without engaging in conventional hydrogen bonding. Alkyl hydrophobic interactions with CYS92 and ARG94, coupled with van der Waals interactions with TRP105, influenced protein conformation changes and overall dynamics of the complex.
Retrospective validation
ROC curve analysis revealed a robust predictive performance for Morgan fingerprints, achieving an area under the curve (AUC) of 0.90, while Layered fingerprints demonstrated reasonable accuracy with an AUC of 0.80 as can be observed from Fig 8.
[Figure omitted. See PDF.]
a) ROC plot based on Tanimoto scores for MF and Layered fingerprint descriptors b) Heat map showing Tanimoto similarity between control (index 1), actives (index 2–21) and decoys (22–50) and c) Box plots with distribution of Tanimoto scores between active-control and decoy-control for both the fingerprint descriptors.
The Enrichment Factor at 1% was 24.81 for Morgan and 12.40 for Layered, highlighting the model’s ability to prioritize active compounds over decoys. The calculated Enrichment Factor (EF) at 1% for binding affinity was 18.18. Considering our limitations in using paid software and that the ROC-AUC metric may not always be optimal for evaluating binding affinity or docking scores in virtual screening, we used EF at top 1% to assess the performance of our model in prioritizating top-ranking compounds [60]. A similarity analysis using Tanimoto coefficients compared active compounds and decoys with Schisantherin A. For clarity, we visualized the first 50 compounds, starting with Schisantherin A as the control, followed by 20 active compounds and 29 decoys. The heatmap reveals a clear distinction between the control, active, and decoy compounds. The control compound (index 1) shows notable similarity with many of the active compounds (indices 2–21), while the active compounds themselves form clusters of moderate-to-high Tanimoto similarity, indicating shared structural features. In contrast, the decoys (indices 22–50) generally display lower similarity to both the control and active compounds, with only a few exceptions. Overall, the demarcations in the heat map indicate that active compounds are more similar to each other and the control, while the decoys remain largely dissimilar, validating the effectiveness of the virtual screening process. The boxplots demonstrate statistically significant differences in similarity values between active and decoy compounds for both Morgan (p <10−5) and Layered fingerprint (p <10−6) techniques. These results highlight that active compounds consistently show greater similarity than decoys, validating the effectiveness of the virtual screening process.
MD simulation
An exhaustive analysis of the MD trajectories of 2AZ5 was carried out, encompassing both its apoprotein state as well as its complex formation with four specific ligands over a simulation period of 200 ns. This facilitated the assessment of protein fluctuation and the influence of ligand binding on protein residues in these two states. It also aided in the identification of stable conformational sub-states and binding modes, as well as the detection of instabilities under dynamic circumstances to evaluate the potential of a natural ligand as a therapeutic candidate.
Root mean square deviation (RMSD).
The RMSD values were computed for the complexes and 2AZ5 to evaluate the stability of each structure during the MD simulation. Overall, the four complexes exhibited low RMSD values corresponding to that of the apoprotein (Apoprotein: 0.254 ± 0.03 nm) demonstrated in Table 4. The time plot of the RMSD values for the protein backbone, shown in Fig 9, indicates that all four complexes exhibited fluctuations around 0.2 nm. While the RMSD values varied during the first 150 ns, they subsequently stabilized, indicating consistent behavior across the complexes. Notably, the ligand complexes—2AZ5-Imperialine, 2AZ5-Veratramine, 2AZ5-Jervine, and 2AZ5-Gelsemine—demonstrated diminished fluctuations compared to the native protein, which is indicative of their potential inhibitory capabilities.
[Figure omitted. See PDF.]
(a) RMSD, (b) RMSF, (c) Rg values plotted for the native apo-protein -black and apo-protein docked with selected bioactive molecules, (1) Imperialine-red, (2) Veratramine-green, (3) Jervine-yellow, and (4) Gelsemine- purple.
[Figure omitted. See PDF.]
Root mean square fluctuation (RMSF).
The complexes exhibited heightened stability relative to the unbound state. By observing the mobility of key residues of 2AZ5, it was visible that the critical residues at the binding site such as PRO91, CYS92, and LYS103 of chain A and SER90, PRO91, CYS92, GLN93 of chain B, had stable fluctuations with RMSF < 0.2 nm (Table 5). For the whole 200 ns simulation, the complexes exhibited lower fluctuations than the apoprotein in general. This was in full agreement with the RMSD values found earlier. Therefore, the results strongly suggested that protein-ligand complexes were stable at their binding site. Some atoms in these four complexes along with the apoprotein had high RMSF values but mostly exhibited values centered around 2 nm as observed in Fig 9(b). Certain active site residues with elevated RMSF may be naturally flexible or have undergone structural modifications, giving clues about conformational alterations linked to enzymatic activity.
[Figure omitted. See PDF.]
Radius of gyration (Rg).
The assessment of the compactness for all four complexes over the entire simulation duration was discerned by examining the Radius of Gyration (Rg) data, as illustrated in Fig 9(c). The Rg values exhibited a consistent trend throughout the MD simulation, indicating a persistently folded structure. As outlined in Table 4, it is evident that all the systems displayed comparable and uniform values of Rg suggesting compactness and stability throughout the entire simulation period.
Hydrogen bonds (H-bonds).
Hydrogen bonds played a key role in maintaining the stability of the complex, with higher numbers of these bonds correlating with increased protein stability. Fig 10(a) illustrates the evolution of hydrogen bond formations over time. Throughout the simulation, the number of hydrogen bonds fluctuated between 2 and 9, reflecting changes in molecular interactions. The moderate level of hydrogen bonds observed in the complexes of Imperialine, Veratramine, and Gelsemine with 2AZ5 suggested that their hydrogen bonding networks remained stable during the simulation. In contrast, Jervine showed lower overall efficiency, consistently forming only two hydrogen bonds throughout most of the simulation, although there were brief moments where it formed up to six hydrogen bonds. A detailed examination of the intermolecular hydrogen bonds between the protein and ligand indicates sustained engagement throughout the entire 200 ns simulation, confirming the stability of these complexes.
[Figure omitted. See PDF.]
(a) H-bonds, and (b) SASA values plotted for the native apo-protein (TNF-α in the human body)-black and apo-protein docked with selected bioactive molecules, (1) Imperialine-red, (2) Veratramine-green, (3) Jervine-yellow, and (4) Gelsemine–purple.
Solvent accessible surface area (SASA).
Lastly, SASA analysis in Fig 10(b) demonstrated that all ligand complexes maintained relatively stable surface areas, reinforcing their enhanced stability. The binding of the ligand did not introduce any instability in the 3D structure as demonstrated by fluctuations around a stable value. The surface area of the complex can impact its stability by modulating the creation of water bridges (intermolecular hydrogen bonds) between the protein and the ligand. The results of the SASA are in good agreement with the binding energy calculations as Jervine demonstrated the maximum SASA also showed the least favorable binding energy values (Table 6). The opposite held for Veratramine which exhibited the lowest SASA and binding energy values.
[Figure omitted. See PDF.]
Essential dynamics and dynamic cross-correlation matrix analysis
The RMSD of the protein backbone stabilized at approximately 0.2 nm, indicating that the structure achieved relative stability after 150 ns during a 200 ns molecular simulation conducted at 300 K. Given these RMSD results, the final 50 ns of the simulation trajectory were utilized to calculate DCCM and to perform PCA.
Principal component analysis
In 200 ns simulation trajectories, the top 20 PCs of the apo-system, 2AZ5-Imperialine, 2AZ5-Veratramine, 2AZ5-Jervine, and 2AZ5-Gelsemine system accounted for 84%, 74.3%, 77.3%, 71.6%, and 74.8% of the total variation, respectively. It was observed that the first few eigenvectors contributed mostly to elucidate the overall dynamics of the protein. The generated 2D plots indicated two distinct conformational states, denoted by blue (unstable) and red (stable) dots, with white dots representing intermediate states [61]. Fig 11 reveals that the percentage of variation in the 2AZ5 structure exhibited a significant decline corresponding to the first four eigenvectors, indicating the conformational changes caused by the docked ligand. Moreover, the eigenvalues for both docked complexes remained relatively constant from 5 to 20 eigenvectors, pointing to a lack of major changes within this range. These observations suggest that the 2AZ5 structure, when bound to the docked compounds, initially showed considerable flexibility, which then decreased over time. Notably, all systems demonstrated compact and correlated motions for the 2AZ5 structure, indicating a rigid and stable state for the docked complex.
[Figure omitted. See PDF.]
(a) 2AZ5, docked with (b) Imperialine, (c)Veratramine, (d) Jervine, and (e) Gelsemine.
Dynamic cross-correlation matrix
DCCM captured the synchronized displacements of Cα backbone atoms that impact the transition of protein conformations across different functional states. It exhibited the collective movement of Cα atoms and, by analyzing their relative motions, revealed the connections between residues and domains. Positive values within the matrix signified a favorable correlation, denoting the concerted movement of residues in a parallel direction. Conversely, negative values indicated anti-correlations, signaling the anti-parallel displacement of residues within complexes. In the accompanying figures (Fig 12), the intensity of the color reflected the strength of the correlation, with red representing a positive correlation, white denoting no correlation, and blue indicating a negative correlation.
[Figure omitted. See PDF.]
2AZ5 complexed with (a) 2AZ5 apo form, (b) Imperialine, (c) Veratramine, (d) Jervine, and (e) Gelsemine on α-carbon atoms.
Upon examining the DCCM of the apo-protein in Fig 12(a), a mostly negative correlation in the overall conformation was evident. This had been diminished notably in the docked complexes The shift in motion toward a slightly positive correlated motion or a lack of negative correlation implied a decrease in overall residue movement. This pointed to improved stability thereby leading to reduced conformational flexibility.
MM/GBSA
The MM/GBSA method was used to estimate the total binding free energy (ΔGbind) for the complexes based on the last 10 ns of the simulation trajectory. The complexes containing Veratramine, Gelsemine, and Imperialine exhibited ΔGbind values of –54.91 kcal/mol, –32.57 kcal/mol, and -24.23 kcal/mol, respectively. The observed low total binding free energy was attributed to the significant contributions from both ΔGgas and ΔEvdW. Notably, the polar solvation energy was unable to make a positive contribution to the overall binding in any of the investigated complexes. Both electrostatic energy and non-polar solvation energy contributed similarly to the binding free energy. Conversely, Jervine displayed higher ΔGbind values of—8.35 kcal/mol, indicating a weaker interaction with the 2AZ5 active site for these compounds. Through comprehensive analyses encompassing complex and ligand stability, as well as MM/GBSA binding free energies, Imperialine, Veratramine, and Gelsemine were identified as the most promising compounds.
Discussion
We scrutinized multiple possible inhibitors of TNF-α from a wide range of natural substances of the Selleckchem database through molecular docking, drug-likeness, and ADMET analysis process. Consequently, we have narrowed down our selection to four probable compounds: Imperialine, Veratramine, Jervine, and Gelsemine. Imperialine, from Bulbs of Fritillaria cirrhosa, is a steroidal alkaloid that has been scientifically proven to suppress inflammatory responses and mitigate pulmonary functional and structural impairment [62]. Additionally, Veratramine and Jervine, also steroidal alkaloids, have been recognized for their pharmacological activities. Among these, Veratramine has been reported to lower blood pressure and act as an anti-thrombotic agent [63]. Jervine, a constituent of Veratrum album, has demonstrated significant anti-inflammatory effects, ranging from 50.4% to 73.5%, against inflammation biomarkers such as TNF-α, IL-1β, and MPO [64]. Gelsemine, a natural alkaloid found in the Gelsemium genus of plants within the Loganiaceae family, has also been shown to possess anxiolytic effects and reduce the levels of pro-inflammatory cytokines [65]. Prior research has documented the use of natural alkaloids to yield promising outcomes in the management of inflammatory conditions such as RA. Therefore, the compounds identified through our screening process are well-validated and could emerge as promising therapeutic agents for treating RA and reducing its systemic symptoms.
Our investigation commenced with the selection of Protein ID 2AZ5 based on validation scores and recent literature references. Addressing missing residues, we aligned it with the template structure 2E7A to ensure accessibility. This technique can be extended to investigate an unknown protein of interest by aligning its structure with relatively matched ones from relevant databases. After the identification of structural homologs, the structural hits can be used to virtually screen databases containing potential inhibitors by integrating DL techniques. In a broader context, DL models enable comprehensive analyses across diverse databases of interest. Leveraging a curated dataset from the ChEMBL repository, we meticulously trained our model to discern intricate molecular features indicative of potential binding affinities by filtering the dataset based on their pIC50 values. Our preprocessing steps ensured data standardization, reducing redundancy which was reflected in the excellent performance of our selected compounds as evidenced by our MD simulation, Essential Dynamics, and MM-GBSA analysis. Overall, our DL model demonstrated good performance and had MAPE value of 10%. This outcome can be further improved by using a larger training dataset or incorporating data from multiple different datasets. To ensure that the top hits were not accidentally eliminated, we selected the compounds falling in the upper fourth quartile after deploying our model on the Selleckchem database. This confirmed that the top hits were all investigated for detailed molecular docking studies, drug-likeness, and ADMET analysis. Additionally, when investigating any known or unknown protein rather than TNF-α, these stringent screening criteria may be altered, such as changing the variance threshold, the desired range of predicted values, Drug-likeness properties, cut-off values of binding affinity, etc. to get the highest number of desired targets. Most importantly, our proposed virtual screening model is well validated as it was capable of successfully distinguishing between active ligands and decoy molecules based on two popular fingerprint descriptors.
Notably, our exploration extended beyond conventional docking analyses, encompassing Molecular Dynamics simulations to assess the dynamic behavior and stability of all four protein-ligand complexes over extended timeframes. These simulations provided valuable insights into the conformational dynamics of TNF-α and its interactions with potential inhibitors, shedding light on the mechanisms underlying inhibition. Furthermore, PCA & DCCM analysis offered a holistic view of the collective conformational changes induced by the selected compounds, illuminating their potential impact on TNF-α activity. In addition, complexes containing Imperialine, Veratramine, and Gelsemine exhibited remarkable results in binding energy analysis. Conversely, Jervine displayed weaker interactions with the protein active site, as indicated by their higher ΔGbind values. Through comprehensive analyses encompassing complex and ligand stability, as well as MM/GBSA binding free energies, Imperialine, Veratramine, and Gelsemine emerged as the most promising compounds for further investigation. Their exceptional binding affinities and favorable interaction profiles make them attractive candidates for therapeutic intervention in inflammatory diseases such as RA. As with any computational studies, our study aimed to optimize between computing power and cost. This optimization meant that we were only able to utilize a fraction of the 200 ns simulation for PCA, DCCM, and MM/GBSA studies. Overall, our study underscores the power of computational methodologies in accelerating drug discovery efforts, offering a systematic approach to identify promising candidates for therapeutic intervention, and contributing to the ongoing quest for effective treatments. Moving forward, our research sets the stage for further investigations, including experimental validation of identified compounds and exploration of their efficacy in preclinical models.
Conclusion
In this study, we present a comprehensive and validated virtual screening pipeline that leverages deep learning techniques to rapidly search databases for natural compounds overcoming the limitation of manual screening and facilitating the exploration of potential drug candidates for RA. Our focus was on identifying natural compounds capable of inhibiting TNF-α, as potential therapeutic agents exhibiting superior safety and efficacy compared to existing drugs. From the approximately 2563 substances within the Selleckchem natural product library, our study employed a DL-based model for screening, followed by assessment through ADMET prediction and docking to initially identify four potential candidates: Imperialine, Veratramine, Jervine, and Gelsemine. Subsequent application of MD simulation and essential dynamics was conducted for these selected compounds. The MM/GBSA approach was then employed to identify top-hit compounds by calculating the binding capacity of each compound to the target, ultimately leading to the identification of Imperialine, Veratramine, and Gelsemine as the most promising TNF-α inhibitors.
Supporting information
S1 Fig. Missing amino acid representation among 4 chains of native 2AZ5 structure.
https://doi.org/10.1371/journal.pone.0303954.s001
(TIFF)
S2 Fig. Active site prediction using, (A) CASTp and (B) DeepSite.
https://doi.org/10.1371/journal.pone.0303954.s002
(TIFF)
S1 Table. Missing amino acid residues among 4 chains of native protein (TNF-α) and their position range.
https://doi.org/10.1371/journal.pone.0303954.s003
(DOCX)
S2 Table. Characterization of protein active sites using CASTp server.
https://doi.org/10.1371/journal.pone.0303954.s004
(DOCX)
S3 Table. Characterization of protein active sites using Deep Site server.
https://doi.org/10.1371/journal.pone.0303954.s005
(DOCX)
S1 Dataset.
https://doi.org/10.1371/journal.pone.0303954.s006
(XLSX)
S1 File. Processed protein FASTA sequence.
https://doi.org/10.1371/journal.pone.0303954.s007
(DOCX)
Acknowledgments
The authors are grateful to the Department of Biomedical Engineering, MIST for providing support to successfully conduct this research.
References
1. 1. Weyand CM, Klimiuk PA, Goronzy JJ. Heterogeneity of rheumatoid arthritis: from phenotypes to genotypes. Springer Semin Immunopathol. 1998;20(1–2):5–22. pmid:9836366
* View Article
* PubMed/NCBI
* Google Scholar
2. 2. Almutairi K, Nossent J, Preen D, Keen H, Inderjeeth C. The global prevalence of rheumatoid arthritis: a meta-analysis based on a systematic review. Rheumatol Int. 2021 May 11;41(5):863–77. pmid:33175207
* View Article
* PubMed/NCBI
* Google Scholar
3. 3. Nerurkar L, Siebert S, McInnes IB, Cavanagh J. Rheumatoid arthritis and depression: an inflammatory perspective. Lancet Psychiatry. 2019 Feb;6(2):164–73. pmid:30366684
* View Article
* PubMed/NCBI
* Google Scholar
4. 4. van de Loo AA, van den Berg WB. Effects of murine recombinant interleukin 1 on synovial joints in mice: measurement of patellar cartilage metabolism and joint inflammation. Ann Rheum Dis. 1990 Apr 1;49(4):238–45. pmid:2339905
* View Article
* PubMed/NCBI
* Google Scholar
5. 5. Keffer J, Probert L, Cazlaris H, Georgopoulos S, Kaslaris E, Kioussis D, et al. Transgenic mice expressing human tumour necrosis factor: a predictive genetic model of arthritis. EMBO J. 1991 Dec;10(13):4025–31. pmid:1721867
* View Article
* PubMed/NCBI
* Google Scholar
6. 6. Georgopoulos S, Plows D, Kollias G. Transmembrane TNF is sufficient to induce localized tissue toxicity and chronic inflammatory arthritis in transgenic mice. J Inflamm. 1996;46(2):86–97. pmid:8734789
* View Article
* PubMed/NCBI
* Google Scholar
7. 7. Kondo N, Kuroda T, Kobayashi D. Cytokine Networks in the Pathogenesis of Rheumatoid Arthritis. Int J Mol Sci. 2021 Oct 10;22(20):10922. pmid:34681582
* View Article
* PubMed/NCBI
* Google Scholar
8. 8. Drosos AA, Pelechas E, Kaltsonoudis E, Voulgari P V. Therapeutic Options and Cost-Effectiveness for Rheumatoid Arthritis Treatment. Curr Rheumatol Rep. 2020 Aug 26;22(8):44. pmid:32591916
* View Article
* PubMed/NCBI
* Google Scholar
9. 9. Baghdadi LR. Effect of methotrexate use on the development of type 2 diabetes in rheumatoid arthritis patients: A systematic review and meta-analysis. PLoS One. 2020 Jul 6;15(7):e0235637. pmid:32628710
* View Article
* PubMed/NCBI
* Google Scholar
10. 10. Jahid M, Khan KU, Rehan-Ul-Haq , Ahmed RS. Overview of Rheumatoid Arthritis and Scientific Understanding of the Disease. Mediterr J Rheumatol. 2023 Sep;34(3):284–91. pmid:37941854
* View Article
* PubMed/NCBI
* Google Scholar
11. 11. Radu AF, Bungau SG. Management of Rheumatoid Arthritis: An Overview. Cells. 2021 Oct 23;10(11):2857. pmid:34831081
* View Article
* PubMed/NCBI
* Google Scholar
12. 12. Fan D, Guo Q, Shen J, Zheng K, Lu C, Zhang G, et al. The Effect of Triptolide in Rheumatoid Arthritis: From Basic Research towards Clinical Translation. Int J Mol Sci. 2018 Jan 26;19(2):376. pmid:29373547
* View Article
* PubMed/NCBI
* Google Scholar
13. 13. Zhang L, Wei W. Anti-inflammatory and immunoregulatory effects of paeoniflorin and total glucosides of paeony. Pharmacol Ther. 2020 Mar;207:107452. pmid:31836457
* View Article
* PubMed/NCBI
* Google Scholar
14. 14. Singh S, Singh TG, Mahajan K, Dhiman S. Medicinal plants used against various inflammatory biomarkers for the management of rheumatoid arthritis. Journal of Pharmacy and Pharmacology. 2020 Sep 10;72(10):1306–27. pmid:32812250
* View Article
* PubMed/NCBI
* Google Scholar
15. 15. Cheng Y, Zhao Y, Zheng Y. Therapeutic potential of triptolide in autoimmune diseases and strategies to reduce its toxicity. Chin Med. 2021 Nov 7;16(1):114. pmid:34743749
* View Article
* PubMed/NCBI
* Google Scholar
16. 16. Rabaan AA, Halwani MA, Garout M, Alotaibi J, AlShehail BM, Alotaibi N, et al. Exploration of phytochemical compounds against Marburg virus using QSAR, molecular dynamics, and free energy landscape. Mol Divers. 2023 Nov 5; pmid:37925643
* View Article
* PubMed/NCBI
* Google Scholar
17. 17. Zia K, Ashraf S, Jabeen A, Saeed M, Nur-e-Alam M, Ahmed S, et al. Identification of potential TNF-α inhibitors: from in silico to in vitro studies. Sci Rep. 2020 Dec 1;10(1):20974.
* View Article
* Google Scholar
18. 18. Joshi T, Joshi T, Pundir H, Sharma P, Mathpal S, Chandra S. Predictive modeling by deep learning, virtual screening and molecular dynamics study of natural compounds against SARS-CoV-2 main protease. J Biomol Struct Dyn. 2021 Nov 22;39(17):6728–46. pmid:32752947
* View Article
* PubMed/NCBI
* Google Scholar
19. 19. Burley SK, Berman HM, Bhikadiya C, Bi C, Chen L, Costanzo L Di, et al. Protein Data Bank: the single global archive for 3D macromolecular structure data. Nucleic Acids Res. 2019 Jan 8;47(D1):D520–8. pmid:30357364
* View Article
* PubMed/NCBI
* Google Scholar
20. 20. Parves MdR, Mahmud S, Riza YM, Sujon KM, Uddin MAR, Chowdhury MdIA, et al. Inhibition of TNF-Alpha Using Plant-Derived Small Molecules for Treatment of Inflammation-Mediated Diseases. In: The 1st International Electronic Conference on Biomolecules: Natural and Bio-Inspired Therapeutics for Human Diseases. Basel Switzerland: MDPI; 2020. p. 13.
21. 21. Xu S, Peng H, Wang N, Zhao M. Inhibition of TNF-α and IL-1 by compounds from selected plants for rheumatoid arthritis therapy: <i>In vivo</i> and <i>in silico</i> studies. Tropical Journal of Pharmaceutical Research. 2018 Mar 7;17(2):277.
* View Article
* Google Scholar
22. 22. Shivaleela B, Srushti SC, Shreedevi SJ, Babu RL. Thalidomide-based inhibitor for TNF-α: designing and Insilico evaluation. Futur J Pharm Sci. 2022 Dec 7;8(1):5.
* View Article
* Google Scholar
23. 23. Brunocilla C, Console L, Rovella F, Indiveri C. Insights into the Transport Cycle of LAT1 and Interaction with the Inhibitor JPH203. Int J Mol Sci. 2023 Feb 17;24(4):4042. pmid:36835453
* View Article
* PubMed/NCBI
* Google Scholar
24. 24. Mahuri M, Paul M, Sahu SK, Thatoi H. In silico homology modeling, docking and sequence analysis of some bacterial laccases to unravel enzymatic specificity towards lignin biodegradation. J Biomol Struct Dyn. 2023 Aug 13;41(12):5757–75.
* View Article
* Google Scholar
25. 25. Zaka M, Abbasi BH, Durdagi S. Novel tumor necrosis factor-α (TNF-α) inhibitors from small molecule library screening for their therapeutic activity profiles against rheumatoid arthritis using target-driven approaches and binary QSAR models. J Biomol Struct Dyn. 2019 Jun 13;37(9):2464–76.
* View Article
* Google Scholar
26. 26. Padalino G, Coghlan A, Pagliuca G, Forde-Thomas JE, Berriman M, Hoffmann KF. Using ChEMBL to Complement Schistosome Drug Discovery. Pharmaceutics. 2023 Apr 28;15(5):1359. pmid:37242601
* View Article
* PubMed/NCBI
* Google Scholar
27. 27. Weininger D, Weininger A, Weininger JL. SMILES. 2. Algorithm for generation of unique SMILES notation. J Chem Inf Comput Sci. 1989 May 1;29(2):97–101.
* View Article
* Google Scholar
28. 28. Warszycki D, Struski Ł, Śmieja M, Kafel R, Kurczab R. Pharmacoprint: A Combination of a Pharmacophore Fingerprint and Artificial Intelligence as a Tool for Computer-Aided Drug Design. J Chem Inf Model. 2021 Oct 25;61(10):5054–65. pmid:34547888
* View Article
* PubMed/NCBI
* Google Scholar
29. 29. Liu Y, Bi M, Zhang X, Zhang N, Sun G, Zhou Y, et al. Machine Learning Models for the Classification of CK2 Natural Products Inhibitors with Molecular Fingerprint Descriptors. Processes. 2021 Nov 19;9(11):2074.
* View Article
* Google Scholar
30. 30. Prabha S, Sasikumar S, Surendra S, Chennakeshava P, Reddy YSM. Drug Discovery Analysis Using Machine Learning Bioinformatics. In 2023. p. 419–31.
31. 31. Hight SK, Clark TN, Kurita KL, McMillan EA, Bray W, Shaikh AF, et al. High-throughput functional annotation of natural products by integrated activity profiling. Proceedings of the National Academy of Sciences. 2022 Dec 6;119(49). pmid:36449542
* View Article
* PubMed/NCBI
* Google Scholar
32. 32. Kurnaz TF, Erden C, Kökçam AH, Dağdeviren U, Demir AS. A hyper parameterized artificial neural network approach for prediction of the factor of safety against liquefaction. Eng Geol. 2023 Jun;319:107109.
* View Article
* Google Scholar
33. 33. Chicco D, Warrens MJ, Jurman G. The coefficient of determination R-squared is more informative than SMAPE, MAE, MAPE, MSE and RMSE in regression analysis evaluation. PeerJ Comput Sci. 2021 Jul 5;7:e623. pmid:34307865
* View Article
* PubMed/NCBI
* Google Scholar
34. 34. Miah MM, Tabassum N, Afroj Zinnia M, Islam ABMMdK. Drug and Anti-Viral Peptide Design to Inhibit the Monkeypox Virus by Restricting A36R Protein. Bioinform Biol Insights. 2022 Jan 19;16:117793222211411. pmid:36570327
* View Article
* PubMed/NCBI
* Google Scholar
35. 35. Jiménez J, Doerr S, Martínez-Rosell G, Rose AS, De Fabritiis G. DeepSite: protein-binding site predictor using 3D-convolutional neural networks. Bioinformatics. 2017 Oct 1;33(19):3036–42. pmid:28575181
* View Article
* PubMed/NCBI
* Google Scholar
36. 36. Riyadi PH, Romadhon , Sari ID, Kurniasih RA, Agustini TW, Swastawati F, et al. SwissADME predictions of pharmacokinetics and drug-likeness properties of small molecules present in Spirulina platensis. IOP Conf Ser Earth Environ Sci. 2021 Oct 1;890(1):012021.
* View Article
* Google Scholar
37. 37. Yalcin S. Molecular Docking, Drug Likeness, and ADMET Analyses of Passiflora Compounds as P-Glycoprotein (P-gp) Inhibitor for the Treatment of Cancer. Curr Pharmacol Rep. 2020 Dec 18;6(6):429–40.
* View Article
* Google Scholar
38. 38. Butt SS, Badshah Y, Shabbir M, Rafiq M. Molecular Docking Using Chimera and Autodock Vina Software for Nonbioinformaticians. JMIR Bioinform Biotech. 2020 Jun 19;1(1):e14232. pmid:38943236
* View Article
* PubMed/NCBI
* Google Scholar
39. 39. Xiong G, Wu Z, Yi J, Fu L, Yang Z, Hsieh C, et al. ADMETlab 2.0: an integrated online platform for accurate and comprehensive predictions of ADMET properties. Nucleic Acids Res. 2021 Jul 2;49(W1):W5–14. pmid:33893803
* View Article
* PubMed/NCBI
* Google Scholar
40. 40. Mysinger MM, Carchia M, Irwin John J, Shoichet BK. Directory of Useful Decoys, Enhanced (DUD-E): Better Ligands and Decoys for Better Benchmarking. J Med Chem. 2012 Jul 26;55(14):6582–94. pmid:22716043
* View Article
* PubMed/NCBI
* Google Scholar
41. 41. Khan SU, Ahemad N, Chuah LH, Naidu R, Htar TT. Sequential ligand- and structure-based virtual screening approach for the identification of potential G protein-coupled estrogen receptor-1 (GPER-1) modulators. RSC Adv. 2019;9(5):2525–38. pmid:35520492
* View Article
* PubMed/NCBI
* Google Scholar
42. 42. Cereto-Massagué A, Ojeda MJ, Valls C, Mulero M, Garcia-Vallvé S, Pujadas G. Molecular fingerprint similarity search in virtual screening. Methods. 2015 Jan;71:58–63. pmid:25132639
* View Article
* PubMed/NCBI
* Google Scholar
43. 43. Ni B, Wang H, Khalaf HKS, Blay V, Houston DR. AutoDock-SS: AutoDock for Multiconformational Ligand-Based Virtual Screening. J Chem Inf Model. 2024 May 13;64(9):3779–89. pmid:38624083
* View Article
* PubMed/NCBI
* Google Scholar
44. 44. Zhou H, Skolnick J. Utility of the Morgan Fingerprint in Structure-Based Virtual Ligand Screening. J Phys Chem B. 2024 Jun 6;128(22):5363–70. pmid:38783525
* View Article
* PubMed/NCBI
* Google Scholar
45. 45. Pattanaik L, Coley CW. Molecular Representation: Going Long on Fingerprints. Chem. 2020 Jun;6(6):1204–7.
* View Article
* Google Scholar
46. 46. Boyenle ID, Adelusi TI, Ogunlana AT, Oluwabusola RA, Ibrahim NO, Tolulope A, et al. Consensus scoring-based virtual screening and molecular dynamics simulation of some TNF-alpha inhibitors. Inform Med Unlocked. 2022;28:100833.
* View Article
* Google Scholar
47. 47. Ci X, Ren R, Xu K, Li H, Yu Q, Song Y, et al. Schisantherin A Exhibits Anti-inflammatory Properties by Down-Regulating NF-κB and MAPK Signaling Pathways in Lipopolysaccharide-Treated RAW 264.7 Cells. Inflammation. 2010 Apr 14;33(2):126–36.
* View Article
* Google Scholar
48. 48. Abraham MJ, Murtola T, Schulz R, Páll S, Smith JC, Hess B, et al. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX. 2015 Sep;1–2:19–25.
* View Article
* Google Scholar
49. 49. Loschwitz J, Jäckering A, Keutmann M, Olagunju M, Olubiyi OO, Strodel B. Dataset of AMBER force field parameters of drugs, natural products and steroids for simulations using GROMACS. Data Brief. 2021 Apr;35:106948. pmid:33855133
* View Article
* PubMed/NCBI
* Google Scholar
50. 50. da Silva TU, Pougy K de C, Albuquerque MG, da Silva Lima CH, Machado S de P. Development of parameters compatible with the CHARMM36 force field for [Fe 4 S 4] 2+ clusters and molecular dynamics simulations of adenosine-5’-phosphosulfate reductase in GROMACS 2019. J Biomol Struct Dyn. 2022 May 24;40(8):3481–91.
* View Article
* Google Scholar
51. 51. Lui LH, van der Walle CF, Brocchini S, Velayudhan A. Discovering Novel Small Molecule Compound for Prevention of Monoclonal Antibody Self-Association. Antibodies. 2022 Jun 8;11(2):40. pmid:35735359
* View Article
* PubMed/NCBI
* Google Scholar
52. 52. Darden T, York D, Pedersen L. Particle mesh Ewald: An N ⋅log (N) method for Ewald sums in large systems. J Chem Phys. 1993 Jun 15;98(12):10089–92.
* View Article
* Google Scholar
53. 53. Rout AK, Acharya V, Maharana D, Dehury B, Udgata SR, Jena R, et al. Insights into structure and dynamics of extracellular domain of Toll-like receptor 5 in Cirrhinus mrigala (mrigala): A molecular dynamics simulation approach. PLoS One. 2021 Jan 14;16(1):e0245358. pmid:33444377
* View Article
* PubMed/NCBI
* Google Scholar
54. 54. Gewers FL, Ferreira GR, De Arruda HF, Silva FN, Comin CH, Amancio DR, et al. Principal Component Analysis. ACM Comput Surv. 2022 May 31;54(4):1–34.
* View Article
* Google Scholar
55. 55. Grant BJ, Skjærven L, Yao X. The Bio3D packages for structural bioinformatics. Protein Science. 2021 Jan 17;30(1):20–30. pmid:32734663
* View Article
* PubMed/NCBI
* Google Scholar
56. 56. Chanana S, Thomas CS, Zhang F, Rajski SR, Bugni TS. hcapca: Automated Hierarchical Clustering and Principal Component Analysis of Large Metabolomic Datasets in R. Metabolites. 2020 Jul 21;10(7):297. pmid:32708222
* View Article
* PubMed/NCBI
* Google Scholar
57. 57. dos Santos Nascimento IJ, De Souza M, Medeiros DC, de Moura RO. Dynamic Cross-Correlation Matrix (DCCM) Reveals New Insights to Discover New NLRP3 Inhibitors Useful as Anti-Inflammatory Drugs. In: ECMC 2022. Basel Switzerland: MDPI; 2022. p. 84.
58. 58. Tuccinardi T. What is the current value of MM/PBSA and MM/GBSA methods in drug discovery? Expert Opin Drug Discov. 2021 Nov 2;16(11):1233–7. pmid:34165011
* View Article
* PubMed/NCBI
* Google Scholar
59. 59. Tran QH, Nguyen QT, Vo NQH, Mai TT, Tran TTN, Tran TD, et al. Structure-based 3D-Pharmacophore modeling to discover novel interleukin 6 inhibitors: An in silico screening, molecular dynamics simulations and binding free energy calculations. PLoS One. 2022 Apr 6;17(4):e0266632. pmid:35385549
* View Article
* PubMed/NCBI
* Google Scholar
60. 60. Wójcikowski M, Ballester PJ, Siedlecki P. Performance of machine-learning scoring functions in structure-based virtual screening. Sci Rep. 2017 Apr 25;7(1):46710. pmid:28440302
* View Article
* PubMed/NCBI
* Google Scholar
61. 61. Bahuguna A, Bharadwaj S, Bajpai VK, Shukla S, Won DW, Park I, et al. Insights into cyclooxygenase-2 inhibition by isolated bioactive compounds 3-caffeoyl-4-dihydrocaffeoyl quinic acid and isorhamnetin 3-O-β-D-glucopyranoside from Salicornia herbacea. Phytomedicine. 2021 Sep;90:153638.
* View Article
* Google Scholar
62. 62. Wang D, Du Q, Li H, Wang S. The Isosteroid Alkaloid Imperialine from Bulbs of Fritillaria cirrhosa Mitigates Pulmonary Functional and Structural Impairment and Suppresses Inflammatory Response in a COPD-Like Rat Model. Mediators Inflamm. 2016;2016:1–17.
* View Article
* Google Scholar
63. 63. Yin L, Xia Y, Xu P, Zheng W, Gao Y, Xie F, et al. Veratramine suppresses human HepG2 liver cancer cell growth in vitro and in vivo by inducing autophagic cell death. Oncol Rep. 2020 May 25;44(2):477–86.
* View Article
* Google Scholar
64. 64. Dumlu FA, Aydin T, Odabasoglu F, Berktas OA, Kutlu Z, Erol HS, et al. Anti-inflammatory and antioxidant properties of jervine, a sterodial alkaloid from rhizomes of Veratrum album. Phytomedicine. 2019 Mar;55:191–9. pmid:30668429
* View Article
* PubMed/NCBI
* Google Scholar
65. 65. Yu H, Tang MH, Zeng ZY, Huang SJ, Zheng XF, Liu ZY. Suppressive Effects of Gelsemine on Anxiety-like Behaviors Induced by Chronic Unpredictable Mild Stress in Mice. Brain Sci. 2022 Jan 30;12(2):191. pmid:35203954
* View Article
* PubMed/NCBI
* Google Scholar
Citation: Nabi T, Riyed TH, Ornob A (2024) Deep learning based predictive modeling to screen natural compounds against TNF-alpha for the potential management of rheumatoid arthritis: Virtual screening to comprehensive in silico investigation. PLoS ONE 19(12): e0303954. https://doi.org/10.1371/journal.pone.0303954
About the Authors:
Tasnia Nabi
Contributed equally to this work with: Tasnia Nabi, Tanver Hasan Riyed
Roles: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Resources, Validation, Visualization, Writing – original draft, Writing – review & editing
Affiliation: Department of Biomedical Engineering, Military Institute of Science and Technology (MIST), Dhaka, Bangladesh
Tanver Hasan Riyed
Contributed equally to this work with: Tasnia Nabi, Tanver Hasan Riyed
Roles: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Software, Validation, Visualization, Writing – original draft, Writing – review & editing
Affiliation: Department of Biomedical Engineering, Military Institute of Science and Technology (MIST), Dhaka, Bangladesh
Akid Ornob
Roles: Conceptualization, Formal analysis, Investigation, Project administration, Resources, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing
E-mail: [email protected]
Affiliation: Department of Biomedical Engineering, Military Institute of Science and Technology (MIST), Dhaka, Bangladesh
ORICD: https://orcid.org/0000-0002-4590-5255
[/RAW_REF_TEXT]
[/RAW_REF_TEXT]
[/RAW_REF_TEXT]
1. Weyand CM, Klimiuk PA, Goronzy JJ. Heterogeneity of rheumatoid arthritis: from phenotypes to genotypes. Springer Semin Immunopathol. 1998;20(1–2):5–22. pmid:9836366
2. Almutairi K, Nossent J, Preen D, Keen H, Inderjeeth C. The global prevalence of rheumatoid arthritis: a meta-analysis based on a systematic review. Rheumatol Int. 2021 May 11;41(5):863–77. pmid:33175207
3. Nerurkar L, Siebert S, McInnes IB, Cavanagh J. Rheumatoid arthritis and depression: an inflammatory perspective. Lancet Psychiatry. 2019 Feb;6(2):164–73. pmid:30366684
4. van de Loo AA, van den Berg WB. Effects of murine recombinant interleukin 1 on synovial joints in mice: measurement of patellar cartilage metabolism and joint inflammation. Ann Rheum Dis. 1990 Apr 1;49(4):238–45. pmid:2339905
5. Keffer J, Probert L, Cazlaris H, Georgopoulos S, Kaslaris E, Kioussis D, et al. Transgenic mice expressing human tumour necrosis factor: a predictive genetic model of arthritis. EMBO J. 1991 Dec;10(13):4025–31. pmid:1721867
6. Georgopoulos S, Plows D, Kollias G. Transmembrane TNF is sufficient to induce localized tissue toxicity and chronic inflammatory arthritis in transgenic mice. J Inflamm. 1996;46(2):86–97. pmid:8734789
7. Kondo N, Kuroda T, Kobayashi D. Cytokine Networks in the Pathogenesis of Rheumatoid Arthritis. Int J Mol Sci. 2021 Oct 10;22(20):10922. pmid:34681582
8. Drosos AA, Pelechas E, Kaltsonoudis E, Voulgari P V. Therapeutic Options and Cost-Effectiveness for Rheumatoid Arthritis Treatment. Curr Rheumatol Rep. 2020 Aug 26;22(8):44. pmid:32591916
9. Baghdadi LR. Effect of methotrexate use on the development of type 2 diabetes in rheumatoid arthritis patients: A systematic review and meta-analysis. PLoS One. 2020 Jul 6;15(7):e0235637. pmid:32628710
10. Jahid M, Khan KU, Rehan-Ul-Haq , Ahmed RS. Overview of Rheumatoid Arthritis and Scientific Understanding of the Disease. Mediterr J Rheumatol. 2023 Sep;34(3):284–91. pmid:37941854
11. Radu AF, Bungau SG. Management of Rheumatoid Arthritis: An Overview. Cells. 2021 Oct 23;10(11):2857. pmid:34831081
12. Fan D, Guo Q, Shen J, Zheng K, Lu C, Zhang G, et al. The Effect of Triptolide in Rheumatoid Arthritis: From Basic Research towards Clinical Translation. Int J Mol Sci. 2018 Jan 26;19(2):376. pmid:29373547
13. Zhang L, Wei W. Anti-inflammatory and immunoregulatory effects of paeoniflorin and total glucosides of paeony. Pharmacol Ther. 2020 Mar;207:107452. pmid:31836457
14. Singh S, Singh TG, Mahajan K, Dhiman S. Medicinal plants used against various inflammatory biomarkers for the management of rheumatoid arthritis. Journal of Pharmacy and Pharmacology. 2020 Sep 10;72(10):1306–27. pmid:32812250
15. Cheng Y, Zhao Y, Zheng Y. Therapeutic potential of triptolide in autoimmune diseases and strategies to reduce its toxicity. Chin Med. 2021 Nov 7;16(1):114. pmid:34743749
16. Rabaan AA, Halwani MA, Garout M, Alotaibi J, AlShehail BM, Alotaibi N, et al. Exploration of phytochemical compounds against Marburg virus using QSAR, molecular dynamics, and free energy landscape. Mol Divers. 2023 Nov 5; pmid:37925643
17. Zia K, Ashraf S, Jabeen A, Saeed M, Nur-e-Alam M, Ahmed S, et al. Identification of potential TNF-α inhibitors: from in silico to in vitro studies. Sci Rep. 2020 Dec 1;10(1):20974.
18. Joshi T, Joshi T, Pundir H, Sharma P, Mathpal S, Chandra S. Predictive modeling by deep learning, virtual screening and molecular dynamics study of natural compounds against SARS-CoV-2 main protease. J Biomol Struct Dyn. 2021 Nov 22;39(17):6728–46. pmid:32752947
19. Burley SK, Berman HM, Bhikadiya C, Bi C, Chen L, Costanzo L Di, et al. Protein Data Bank: the single global archive for 3D macromolecular structure data. Nucleic Acids Res. 2019 Jan 8;47(D1):D520–8. pmid:30357364
20. Parves MdR, Mahmud S, Riza YM, Sujon KM, Uddin MAR, Chowdhury MdIA, et al. Inhibition of TNF-Alpha Using Plant-Derived Small Molecules for Treatment of Inflammation-Mediated Diseases. In: The 1st International Electronic Conference on Biomolecules: Natural and Bio-Inspired Therapeutics for Human Diseases. Basel Switzerland: MDPI; 2020. p. 13.
21. Xu S, Peng H, Wang N, Zhao M. Inhibition of TNF-α and IL-1 by compounds from selected plants for rheumatoid arthritis therapy: <i>In vivo</i> and <i>in silico</i> studies. Tropical Journal of Pharmaceutical Research. 2018 Mar 7;17(2):277.
22. Shivaleela B, Srushti SC, Shreedevi SJ, Babu RL. Thalidomide-based inhibitor for TNF-α: designing and Insilico evaluation. Futur J Pharm Sci. 2022 Dec 7;8(1):5.
23. Brunocilla C, Console L, Rovella F, Indiveri C. Insights into the Transport Cycle of LAT1 and Interaction with the Inhibitor JPH203. Int J Mol Sci. 2023 Feb 17;24(4):4042. pmid:36835453
24. Mahuri M, Paul M, Sahu SK, Thatoi H. In silico homology modeling, docking and sequence analysis of some bacterial laccases to unravel enzymatic specificity towards lignin biodegradation. J Biomol Struct Dyn. 2023 Aug 13;41(12):5757–75.
25. Zaka M, Abbasi BH, Durdagi S. Novel tumor necrosis factor-α (TNF-α) inhibitors from small molecule library screening for their therapeutic activity profiles against rheumatoid arthritis using target-driven approaches and binary QSAR models. J Biomol Struct Dyn. 2019 Jun 13;37(9):2464–76.
26. Padalino G, Coghlan A, Pagliuca G, Forde-Thomas JE, Berriman M, Hoffmann KF. Using ChEMBL to Complement Schistosome Drug Discovery. Pharmaceutics. 2023 Apr 28;15(5):1359. pmid:37242601
27. Weininger D, Weininger A, Weininger JL. SMILES. 2. Algorithm for generation of unique SMILES notation. J Chem Inf Comput Sci. 1989 May 1;29(2):97–101.
28. Warszycki D, Struski Ł, Śmieja M, Kafel R, Kurczab R. Pharmacoprint: A Combination of a Pharmacophore Fingerprint and Artificial Intelligence as a Tool for Computer-Aided Drug Design. J Chem Inf Model. 2021 Oct 25;61(10):5054–65. pmid:34547888
29. Liu Y, Bi M, Zhang X, Zhang N, Sun G, Zhou Y, et al. Machine Learning Models for the Classification of CK2 Natural Products Inhibitors with Molecular Fingerprint Descriptors. Processes. 2021 Nov 19;9(11):2074.
30. Prabha S, Sasikumar S, Surendra S, Chennakeshava P, Reddy YSM. Drug Discovery Analysis Using Machine Learning Bioinformatics. In 2023. p. 419–31.
31. Hight SK, Clark TN, Kurita KL, McMillan EA, Bray W, Shaikh AF, et al. High-throughput functional annotation of natural products by integrated activity profiling. Proceedings of the National Academy of Sciences. 2022 Dec 6;119(49). pmid:36449542
32. Kurnaz TF, Erden C, Kökçam AH, Dağdeviren U, Demir AS. A hyper parameterized artificial neural network approach for prediction of the factor of safety against liquefaction. Eng Geol. 2023 Jun;319:107109.
33. Chicco D, Warrens MJ, Jurman G. The coefficient of determination R-squared is more informative than SMAPE, MAE, MAPE, MSE and RMSE in regression analysis evaluation. PeerJ Comput Sci. 2021 Jul 5;7:e623. pmid:34307865
34. Miah MM, Tabassum N, Afroj Zinnia M, Islam ABMMdK. Drug and Anti-Viral Peptide Design to Inhibit the Monkeypox Virus by Restricting A36R Protein. Bioinform Biol Insights. 2022 Jan 19;16:117793222211411. pmid:36570327
35. Jiménez J, Doerr S, Martínez-Rosell G, Rose AS, De Fabritiis G. DeepSite: protein-binding site predictor using 3D-convolutional neural networks. Bioinformatics. 2017 Oct 1;33(19):3036–42. pmid:28575181
36. Riyadi PH, Romadhon , Sari ID, Kurniasih RA, Agustini TW, Swastawati F, et al. SwissADME predictions of pharmacokinetics and drug-likeness properties of small molecules present in Spirulina platensis. IOP Conf Ser Earth Environ Sci. 2021 Oct 1;890(1):012021.
37. Yalcin S. Molecular Docking, Drug Likeness, and ADMET Analyses of Passiflora Compounds as P-Glycoprotein (P-gp) Inhibitor for the Treatment of Cancer. Curr Pharmacol Rep. 2020 Dec 18;6(6):429–40.
38. Butt SS, Badshah Y, Shabbir M, Rafiq M. Molecular Docking Using Chimera and Autodock Vina Software for Nonbioinformaticians. JMIR Bioinform Biotech. 2020 Jun 19;1(1):e14232. pmid:38943236
39. Xiong G, Wu Z, Yi J, Fu L, Yang Z, Hsieh C, et al. ADMETlab 2.0: an integrated online platform for accurate and comprehensive predictions of ADMET properties. Nucleic Acids Res. 2021 Jul 2;49(W1):W5–14. pmid:33893803
40. Mysinger MM, Carchia M, Irwin John J, Shoichet BK. Directory of Useful Decoys, Enhanced (DUD-E): Better Ligands and Decoys for Better Benchmarking. J Med Chem. 2012 Jul 26;55(14):6582–94. pmid:22716043
41. Khan SU, Ahemad N, Chuah LH, Naidu R, Htar TT. Sequential ligand- and structure-based virtual screening approach for the identification of potential G protein-coupled estrogen receptor-1 (GPER-1) modulators. RSC Adv. 2019;9(5):2525–38. pmid:35520492
42. Cereto-Massagué A, Ojeda MJ, Valls C, Mulero M, Garcia-Vallvé S, Pujadas G. Molecular fingerprint similarity search in virtual screening. Methods. 2015 Jan;71:58–63. pmid:25132639
43. Ni B, Wang H, Khalaf HKS, Blay V, Houston DR. AutoDock-SS: AutoDock for Multiconformational Ligand-Based Virtual Screening. J Chem Inf Model. 2024 May 13;64(9):3779–89. pmid:38624083
44. Zhou H, Skolnick J. Utility of the Morgan Fingerprint in Structure-Based Virtual Ligand Screening. J Phys Chem B. 2024 Jun 6;128(22):5363–70. pmid:38783525
45. Pattanaik L, Coley CW. Molecular Representation: Going Long on Fingerprints. Chem. 2020 Jun;6(6):1204–7.
46. Boyenle ID, Adelusi TI, Ogunlana AT, Oluwabusola RA, Ibrahim NO, Tolulope A, et al. Consensus scoring-based virtual screening and molecular dynamics simulation of some TNF-alpha inhibitors. Inform Med Unlocked. 2022;28:100833.
47. Ci X, Ren R, Xu K, Li H, Yu Q, Song Y, et al. Schisantherin A Exhibits Anti-inflammatory Properties by Down-Regulating NF-κB and MAPK Signaling Pathways in Lipopolysaccharide-Treated RAW 264.7 Cells. Inflammation. 2010 Apr 14;33(2):126–36.
48. Abraham MJ, Murtola T, Schulz R, Páll S, Smith JC, Hess B, et al. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX. 2015 Sep;1–2:19–25.
49. Loschwitz J, Jäckering A, Keutmann M, Olagunju M, Olubiyi OO, Strodel B. Dataset of AMBER force field parameters of drugs, natural products and steroids for simulations using GROMACS. Data Brief. 2021 Apr;35:106948. pmid:33855133
50. da Silva TU, Pougy K de C, Albuquerque MG, da Silva Lima CH, Machado S de P. Development of parameters compatible with the CHARMM36 force field for [Fe 4 S 4] 2+ clusters and molecular dynamics simulations of adenosine-5’-phosphosulfate reductase in GROMACS 2019. J Biomol Struct Dyn. 2022 May 24;40(8):3481–91.
51. Lui LH, van der Walle CF, Brocchini S, Velayudhan A. Discovering Novel Small Molecule Compound for Prevention of Monoclonal Antibody Self-Association. Antibodies. 2022 Jun 8;11(2):40. pmid:35735359
52. Darden T, York D, Pedersen L. Particle mesh Ewald: An N ⋅log (N) method for Ewald sums in large systems. J Chem Phys. 1993 Jun 15;98(12):10089–92.
53. Rout AK, Acharya V, Maharana D, Dehury B, Udgata SR, Jena R, et al. Insights into structure and dynamics of extracellular domain of Toll-like receptor 5 in Cirrhinus mrigala (mrigala): A molecular dynamics simulation approach. PLoS One. 2021 Jan 14;16(1):e0245358. pmid:33444377
54. Gewers FL, Ferreira GR, De Arruda HF, Silva FN, Comin CH, Amancio DR, et al. Principal Component Analysis. ACM Comput Surv. 2022 May 31;54(4):1–34.
55. Grant BJ, Skjærven L, Yao X. The Bio3D packages for structural bioinformatics. Protein Science. 2021 Jan 17;30(1):20–30. pmid:32734663
56. Chanana S, Thomas CS, Zhang F, Rajski SR, Bugni TS. hcapca: Automated Hierarchical Clustering and Principal Component Analysis of Large Metabolomic Datasets in R. Metabolites. 2020 Jul 21;10(7):297. pmid:32708222
57. dos Santos Nascimento IJ, De Souza M, Medeiros DC, de Moura RO. Dynamic Cross-Correlation Matrix (DCCM) Reveals New Insights to Discover New NLRP3 Inhibitors Useful as Anti-Inflammatory Drugs. In: ECMC 2022. Basel Switzerland: MDPI; 2022. p. 84.
58. Tuccinardi T. What is the current value of MM/PBSA and MM/GBSA methods in drug discovery? Expert Opin Drug Discov. 2021 Nov 2;16(11):1233–7. pmid:34165011
59. Tran QH, Nguyen QT, Vo NQH, Mai TT, Tran TTN, Tran TD, et al. Structure-based 3D-Pharmacophore modeling to discover novel interleukin 6 inhibitors: An in silico screening, molecular dynamics simulations and binding free energy calculations. PLoS One. 2022 Apr 6;17(4):e0266632. pmid:35385549
60. Wójcikowski M, Ballester PJ, Siedlecki P. Performance of machine-learning scoring functions in structure-based virtual screening. Sci Rep. 2017 Apr 25;7(1):46710. pmid:28440302
61. Bahuguna A, Bharadwaj S, Bajpai VK, Shukla S, Won DW, Park I, et al. Insights into cyclooxygenase-2 inhibition by isolated bioactive compounds 3-caffeoyl-4-dihydrocaffeoyl quinic acid and isorhamnetin 3-O-β-D-glucopyranoside from Salicornia herbacea. Phytomedicine. 2021 Sep;90:153638.
62. Wang D, Du Q, Li H, Wang S. The Isosteroid Alkaloid Imperialine from Bulbs of Fritillaria cirrhosa Mitigates Pulmonary Functional and Structural Impairment and Suppresses Inflammatory Response in a COPD-Like Rat Model. Mediators Inflamm. 2016;2016:1–17.
63. Yin L, Xia Y, Xu P, Zheng W, Gao Y, Xie F, et al. Veratramine suppresses human HepG2 liver cancer cell growth in vitro and in vivo by inducing autophagic cell death. Oncol Rep. 2020 May 25;44(2):477–86.
64. Dumlu FA, Aydin T, Odabasoglu F, Berktas OA, Kutlu Z, Erol HS, et al. Anti-inflammatory and antioxidant properties of jervine, a sterodial alkaloid from rhizomes of Veratrum album. Phytomedicine. 2019 Mar;55:191–9. pmid:30668429
65. Yu H, Tang MH, Zeng ZY, Huang SJ, Zheng XF, Liu ZY. Suppressive Effects of Gelsemine on Anxiety-like Behaviors Induced by Chronic Unpredictable Mild Stress in Mice. Brain Sci. 2022 Jan 30;12(2):191. pmid:35203954
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2024 Nabi et al. This is an open access article distributed under the terms of the Creative Commons Attribution License: http://creativecommons.org/licenses/by/4.0/ (the “License”), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Rheumatoid arthritis (RA) affects an estimated 0.1% to 2.0% of the world’s population, leading to a substantial impact on global health. The adverse effects and toxicity associated with conventional RA treatment pathways underscore the critical need to seek potential new therapeutic candidates, particularly those of natural sources that can treat the condition with minimal side effects. To address this challenge, this study employed a deep-learning (DL) based approach to conduct a virtual assessment of natural compounds against the Tumor Necrosis Factor-alpha (TNF-α) protein. TNF-α stands out as the primary pro-inflammatory cytokine, crucial in the development of RA. Our predictive model demonstrated appreciable performance, achieving MSE of 0.6, MAPE of 10%, and MAE of 0.5. The model was then deployed to screen a comprehensive set of 2563 natural compounds obtained from the Selleckchem database. Utilizing their predicted bioactivity (pIC50), the top 128 compounds were identified. Among them, 68 compounds were taken for further analysis based on drug-likeness analysis. Subsequently, selected compounds underwent additional evaluation using molecular docking (< − 8.7 kcal/mol) and ADMET resulting in four compounds posing nominal toxicity, which were finally subjected to MD simulation for 200 ns. Later on, the stability of complexes was assessed via analysis encompassing RMSD, RMSF, Rg, H-Bonds, SASA, and Essential Dynamics. Ultimately, based on the total binding free energy estimated using the MM/GBSA method, Imperialine, Veratramine, and Gelsemine are proven to be potential natural inhibitors of TNF-α.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer