Next Article in Journal
Bioactive Phytochemicals from Mulberry: Potential Anti-Inflammatory Effects in Lipopolysaccharide-Stimulated RAW 264.7 Macrophages
Next Article in Special Issue
In Silico Screening of Novel α1-GABAA Receptor PAMs towards Schizophrenia Based on Combined Modeling Studies of Imidazo [1,2-a]-Pyridines
Previous Article in Journal
Core Histones Are Constituents of the Perinuclear Theca of Murid Spermatozoa: An Assessment of Their Synthesis and Assembly during Spermiogenesis and Function after Gametic Fusion
Previous Article in Special Issue
Docking-Based 3D-QSAR Studies for 1,3,4-oxadiazol-2-one Derivatives as FAAH Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploration of Novel Xanthine Oxidase Inhibitors Based on 1,6-Dihydropyrimidine-5-Carboxylic Acids by an Integrated in Silico Study

1
Hubei Key Laboratory of Novel Reactor and Green Chemical Technology, School of Chemical Engineering and Pharmacy, Wuhan Institute of Technology, Wuhan 430205, China
2
Hubei Key Laboratory of Plasma Chemistry and Advanced Materials, Wuhan Institute of Technology, Wuhan 430205, China
3
School of Materials Science and Engineering, Zhengzhou University, No. 100 Science Avenue, Zhengzhou 450001, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(15), 8122; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22158122
Submission received: 29 May 2021 / Revised: 19 July 2021 / Accepted: 22 July 2021 / Published: 29 July 2021
(This article belongs to the Special Issue QSAR and Chemoinformatics in Molecular Modeling and Drug Design 2.0)

Abstract

:
Xanthine oxidase (XO) is an important target for the effective treatment of hyperuricemia-associated diseases. A series of novel 2-substituted 6-oxo-1,6-dihydropyrimidine-5-carboxylic acids (ODCs) as XO inhibitors (XOIs) with remarkable activities have been reported recently. To better understand the key pharmacological characteristics of these XOIs and explore more hit compounds, in the present study, the three-dimensional quantitative structure–activity relationship (3D-QSAR), molecular docking, pharmacophore modeling, and molecular dynamics (MD) studies were performed on 46 ODCs. The constructed 3D-QSAR models exhibited reliable predictability with satisfactory validation parameters, including q2 = 0.897, R2 = 0.983, rpred2 = 0.948 in a CoMFA model, and q2 = 0.922, R2 = 0.990, rpred2 = 0.840 in a CoMSIA model. Docking and MD simulations further gave insights into the binding modes of these ODCs with the XO protein. The results indicated that key residues Glu802, Arg880, Asn768, Thr1010, Phe914, and Phe1009 could interact with ODCs by hydrogen bonds, π-π stackings, or hydrophobic interactions, which might be significant for the activity of these XOIs. Four potential hits were virtually screened out using the constructed pharmacophore model in combination with molecular dockings and ADME predictions. The four hits were also found to be relatively stable in the binding pocket by MD simulations. The results in this study might provide effective information for the design and development of novel XOIs.

1. Introduction

Gout is a clinical syndrome accompanied by some chronic and recurrent symptoms, such as pain, inflammation, and swelling [1,2,3,4]. Such an unhealthy condition originates from the superabundant presence of serum uric acid (SUA), which has been defined as hyperuricemia (HUA) and regarded as the primary cause of gout [5,6,7]. In the last decade, the dramatic increase of HUA patients has driven HUA to be the second metabolic disease following by the type II diabetes [8]. Uric acid (UA) and reactive oxygen species (ROS), as the oxidative end-products of xanthine and hypoxanthine, are generated in the purine scavenging pathway under the catalysis of xanthine oxidase (XO) [9,10]. The excess UA and ROS usually associate with many pathogeneses, such as inflammation, atherosclerosis, and carcinogenesis [11,12,13]. Clinical evidence has indicated that XO was a promising target for effectively treating with several diseases, especially with the HUA-associated conditions [8,9,10,11,12,13,14].
In the past few decades, XO inhibitors (XOIs) have been widely used in the clinical treatment of HUA. Allopurinol, as a XOI pioneer, was discovered for HUA treatment in the 1960s [15]. However, allopurinol and its analogues with the semblable purine backbone have been restrained to use in some cases, owing to severe and life-threatening side effects like fever, rashes, kidney toxicity, and Stevens–Johnsons syndrome [16,17,18]. Febuxostat and topiroxostat, as novel and effective non-purine XOIs, were approved for marketing in 2009 and 2013, respectively [19,20]. However, these drugs are not regarded as a precise and safe therapy because of their untoward effects on patients, which has prompted researchers to explore and develop novel XOIs with alternate scaffolds and fewer adverse reactions [21,22,23,24]. Various chemotypes of non-purine XOIs have been reported in recent years, such as benzoflavone derivatives [20], 2-aryl/heteroaryl-4-quinolones [11], and 1-hydroxy-2-phenyl-4-pyridyl-1H-imidazoles [18]. In addition, the replacement of the thiazole ring of febuxostat with other heterocyclic rings (magenta, Figure 1), such as pyrazole (e.g., Y-700) [7], imidazole [18], triazole [9], isoxazole [12], thiazole [19], and selenazole [5], could produce more alternatives for novel XOIs.
As shown in Figure 1, febuxostat and its analogues interact with XO through two key pharmacophores, including an aromatic moiety (blue) and a nitrogen-containing heterocyclic ring (red) [6,14]. The linker of two pharmacophores could be modified by the extension of carbon chain or the introduction of heteroatoms like nitrogen. The carboxylate moiety of the nitrogen-containing heterocyclic ring (red) in febuxostat and its analogues act as the most tightly binding part of XOIs with the protein [15,23]. It has been confirmed that the presence of an electron-withdrawing substituent at the 3′ position of the febuxostat, such as a cyano or a nitro, resulted in an increase of inhibition efficiency [5,7]. The tetrazole ring could serve as a hydrogen-bond acceptor (HBA) to participate in forming privileged interactions with some residues in the XO binding pocket [13]. Recently, a novel series of febuxostat analogues, 2-substituted 6-oxo-1,6-dihydropyrimidine-5-carboxylic acids (ODCs), have been designed and synthesized based on bioisosteric replacement and ring enlargement strategies [3,4]. Some of these analogues exhibited better inhibitory activities than febuxostat against the XO protein in vitro.
To explore the structure–activity relationships (SARs) of these novel ODC XOIs and find more ideal candidates, in this study, 46 ODCs were selected for the integrated modeling studies. Three-dimensional quantitative SAR (3D-QSAR) models, including comparative molecular field analysis (CoMFA) and comparative molecular similarity indices analysis (CoMSIA), and pharmacophore models, were constructed using these ODCs. Meanwhile, molecular dockings were served to obtain the possible binding conformations of these XOIs and to explore their action mechanism. To explore novel XOI scaffolds, virtual screening was then carried out by pharmacophore model; molecular dockings; and predictions of absorption, distribution, metabolism, and excretion (ADME). Molecular dynamics (MD) simulations were subsequently performed to verify the rationality of the docking method, and to further analyze the effects and stability of hit compounds in the XO binding pocket. This study might provide important information and more alternatives for the design and development of novel XOIs.

2. Results and Discussion

2.1. CoMFA and CoMSIA Statistical Results

A dataset of 46 ODCs was selected for 3D-QSAR modelling, and their structures and biological activities represented by half-maximal inhibitory concentrations (IC50s) were given in Table 1. The internal and external validation parameters of the CoMFA and CoMSIA models were summarized in Table 2 and Table 3, respectively, using the same training (35 molecules) and test (11 molecules) sets. The CoMFA model gave rational parameters with a cross-validation correlation coefficient (q2) of 0.897, an optimum number of components (ONC) of 7, a standard error of estimate (SEE) of 0.050, a non-cross-validated correlation coefficient (R2) of 0.983, an F-statistic values (F) of 229.50, and a predictive correlation coefficient (rpred2) of 0.948 in the partial least squares (PLS) analysis. The external validation parameters of the CoMFA model, including a root-mean-squared error (RMSE) of 0.074, a Δrm2 of 0.052, and a r m 2 ¯ of 0.864, were also considered to meet the requirements. The contributions of steric and electrostatic fields were 77.3% and 22.7%, respectively.
Different combination modes of steric, electrostatic, hydrophobic, hydrogen-bond donor (HBD), and HBA fields were used to construct different CoMSIA models, and the q2 values of all possible combinations were summarized in Figure S1. The statistical parameters of the six best CoMSIA models were shown in Table 2. Eventually, the CoMSIA-SEHDA model was chosen as the optimal predictive CoMSIA model for the further analyses. As for the CoMSIA-SEHDA model, the q2, SEE, R2, F, rpred2, RMSE, Δrm2, and r m 2 ¯ were 0.922, 0.041, 0.990, 212.26, 0.840, 0.130, 0.118, and 0.717, respectively. The contributions of steric, electrostatic, hydrophobic, HBD, and HBA fields were 10.5%, 24.8%, 37.2%, 19.3%, and 8.2%, respectively. All above statistical parameters indicated that the constructed CoMFA and CoMSIA models could be used for the following study, and the electrostatic, hydrophobic, and HBD fields might be significant for the improvement of ODCs activity.
The obtained CoMFA and CoMSIA models were then applied to predict the bioactivities of the training and test compounds. The actual pIC50s (−logIC50), predicted pIC50s, and their residuals were listed in Table 1. All the residuals were smaller than 0.4, suggesting that the CoMFA and CoMSIA models exhibited good predictivity. To further exhibit the relationships between the actual and predicted activities of all compounds, the scatter plots were depicted in Figure 2. As shown in Figure 2, the two outlier points were related to compounds 41 and 42, whose predicted activities based on the CoMSIA model were slightly lower than their actual activity. All residual values (41: 0.2199; 42: 0.3296) were in the reasonable range. The statistic points of other compounds exhibited great linear correlation, indicating that the 3D-QSAR models possessed high quality for the activity prediction of ODCs.

2.2. Contour Maps of the CoMFA and CoMSIA Models

The CoMFA and CoMSIA contour maps with the most potent compound 44 as a reference molecule were shown in Figure 3 and Figure 4, respectively. As shown in Figure 3, the sterically advantageous and disadvantageous contours were colored in green and yellow, respectively. A medium green contour surrounding the R1 position of compound 44 in both CoMFA and CoMSIA models indicated that bulky substituents at this position might be beneficial to the activity. This was supported by the activity orders as follows: 29 (iso-pentyl) > 28 (iso-butyl) > 27 (iso-propyl) and 4 (iso-pentyl) > 3 (iso-butyl) > 2 (iso-propyl). Another small green contour neared the meta-position of the phenyl ring (R1) in the CoMFA and CoMSIA models, demonstrating that big substituents for structural modification at this place might be favorable for the increment of activity. In the steric contours of the CoMFA model, a large green contour covering the R3 position of the pyrimidine ring highlighted the importance of large groups in this region. The fact that an imino substituent at the R3 position was better for the activity than an oxygen atom could explain this result, as illustrated in the activity orders: 43 (R3 = NH) > 32 (R3 = O) and 44 (R3 = NH) > 31 (R3 = O).
The electrostatic maps of the CoMFA and CoMSIA models were shown in Figure 3; the blue contours denoted that electropositive groups were favor in these regions, whereas the red contours were the opposite. It could be observed that a medium blue contour in the CoMFA model and a large blue contour in the CoMSIA model lied in the R1 position, which were congruent with the following activity order: 5 (allyl) > 8 (propinyl) and 4 (iso-pentyl) > 7 (iso-pentenyl), attributed to the existence of electropositive substituents. A medium red contour in the CoMFA model and a small red contour in the CoMSIA model appeared over the pata- or meta-position of the phenyl ring (R1), indicating that the occupation of electronegative groups at this region was advantageous for increasing the inhibitory activity. This result was in accordance with the experimental data: 20 (m-fluorobenzyl) > 12 (benzyl) and 36 (p-bromobenzyl) > 37 (p-tert-butylbenzyl). Two small blue contours in the CoMFA model and one small blue contour in the CoMSIA model were located near the pyrimidine ring, which were in accordance with the actual activity orders: 45 (R3 = NH) > 36 (R3 = O) and 46 (R3 = NH) > 38 (R3 = O), as the electropositivity of the oxygen is weaker than that of the nitrogen.
Regarding the hydrophobic contours of the CoMSIA model (Figure 4a), the yellow and white contours represented favor and unfavored regions for the hydrophobic groups, respectively. The existence of a big yellow contour at the R1 position revealed that the substituents of hydrophobic groups in this position tended to increase the activity, such as the following activity orders: 13 (p-methylbenzyl) > 12 (benzyl), 38 (p-methylbenzyl) > 33 (benzyl), and 1 (methyl) > 26 (H). There was a white contour covering the R2 position of the pyrimidine ring, suggesting that the hydrophobic groups in this position might not be beneficial for the activity improvement. For instance, compounds 29, 31, 36, and 38 without substituent at the R2 position exhibited higher activity than the corresponding compounds 39, 40, 41, and 42 with a methyl group, respectively.
The HBD and HBA contour maps of the CoMSIA model are shown in Figure 4. For the HBD contours, the cyan contours suggested HBD groups were advantageous for the activity at the corresponding regions. For the HBA contours, the magenta regions represented that the occupation of HBA substituents surrounding these regions might facilitate the bioactivity improvement, whereas the red contours mean the contrary effect. Two small cyan contours and a medium-sized magenta contour were observed over the carboxyl group of the pyrimidine ring in the HBD and HBA contours, respectively. These observations were consistent with a previous study that the carboxylate group of febuxostat analogues was essential for the inhibitory activity since it could serve as HBDs or HBAs to form critical hydrogen bonds with the XO protein [22]. Besides, there was another medium cyan contour near the pyrimidine ring in the HBD contours, suggesting that the nitrogen atoms of the pyrimidine ring might serve as HBDs to interact with the protein. One medium magenta was observed near the benzene ring of the common skeleton in the HBA contour map. This manifested that the HBA existence in the common scaffold, such as a tetrazole ring or a cyano group, might be important for the inhibitory activity.
According to the analyses of the 3D-QSAR models, the appropriate substituents for improving the inhibitory activity of these ODC XOIs at specific positions might be concluded as follows: (1) bulky, positive charged, and/or hydrophobic groups at the R1 position; (2) negative charged groups at the para- or meta-position of the phenyl ring at the R1 position; (3) hydrophilic groups at the R2 position; (4) bulky and/or positive charged groups at the R3 position; (5) a tetrazole ring or a cyano group at the phenyl ring as HBAs; (6) a carboxyl group of the pyrimidine ring as HBDs or HBAs; and (7) the nitrogen atom of the pyrimidine ring as HBDs.

2.3. Molecular Docking

To validate the reliability of the docking method and explore the key interactions, the co-crystallized febuxostat was extracted and then redocked into the XO protein (PDB ID: 1N5X). As shown in Figure 5a, the XO protein was a homodimer, in which one of subunit (blue cartoon) was selected to generate binding pocket (highlighted as gray surface) by the surflex-docking method. Febuxostat and 46 ODCs were successively docked, and their docking scores were summarized in Table S1. The docking superimposition of compounds febuxostat, 44, and 26 was highlighted in Figure 5b to compare their differences in the binding positions. As shown in Figure 6a, the conformation of the redocked febuxostat almost completely overlapped with that of the crystal ligand, with a root-mean-square deviation (RMSD) value of 0.867 Å (<2 Å) and a docking score of 8.22. This indicated that the used docking method and the related parameters were reasonable. The carboxylate group of febuxostat formed hydrogen bonds with Arg880 (Arg880-N-H…O=C, 2.1 Å) and Thr1010 (Thr1010-N-H…O=C, 2.0 Å), and the cyano group interacted with Asn768 by a hydrogen bond (Asn768-N-H…N≡C, 1.9 Å). These hydrogen bonds might be significant for maintaining the binding conformation of febuxostat. The phenyl ring embedded into a hydrophobic pocket that was generated by the surrounding amino acid residues Leu648, Phe649, Leu873, Val1011, Phe1013, Leu1014, Phe914, and Phe1009. Moreover, the π-π stacking was also found between the thiazole ring of febuxostat and the phenyl groups of the aromatic residues Phe914 (face-to-face) and Phe1009 (face-to-edge). These interactions were also important for the binding of febuxostat to the XO protein, as reported in a previous study [12,23]. The above-mentioned results indicated that the docking method was reliable and could be used for the following experiments.
Forty-six ODC XOIs (Table 1) were then docked into the binding site using the same pattern. The docking modes of the most active compound 44 and the least active compound 26 are shown in Figure 6. The docking score of compound 44 (10.69) was much higher than that of compound 26 (7.85), which was in agreement with their experimental activities. In general, we found that the docking score order of these XOIs was accordant with their inhibition potencies. Moreover, it was noted that compounds febuxostat, 44, and 26 have similar docking orientations and binding interactions in the pocket. Similar to febuxostat, compound 44 (Figure 6b) formed hydrogen bonds with Arg880 (Arg880-N-H…O=C, 2.4 Å; Arg880-N-H…O=C, 2.2 Å) by the carboxylate group. The additional hydrogen bonds were formed between Glu802 and the nitrogen atom of the pyrimidine ring (Glu802-C=O…H-N-C, 1.9 Å; Glu802-C=O…H-N-C, 2.6 Å; Glu802-C=O…H-N=C, 1.9 Å).
The docking result of compound 26 (Figure 6c) showed that the carboxylate group, the nitrogen atom of the pyrimidine ring, and the tetrazole group were generated the semblable hydrogen bonds with Arg880 (Arg880-N-H…O=C, 2.3 Å; Arg880-N-H…O=C, 1.8 Å), Thr1010 (Thr1010-C-O-H…O=C, 2.5 Å), Glu802 (Glu802-C=O…H-N-C, 2.1 Å), and Asn768 (Asn768-N-H…N=C, 1.9 Å), respectively. Furthermore, it could be observed that some hydrogen bonds were formed between the nitrogen atom of the pyrimidine ring and Ser876 (Ser876-O-H…N=C, 2.6 Å), and the nitrogen atoms of the tetrazole group and Lys771 (Lys771-N-H…N=C, 2.4 Å; Lys771-N-H…N=C, 2.0 Å).
Compared to the conformation of febuxostat, the changes in the docking results of compound 44 might be caused by the deeper binding position and the rotation of the benzene ring. These variations of compound 44 might have caused the purine ring to form a hydrogen bond with Glu802. The previous study indicated that Glu802 as a key residue could form at least one hydrogen bond with the amino group or the nitrogen-containing heterocyclic ring of febuxostat analogues [8,15]. In comparison with compound 44, compound 26 was found to loss partial interactions with Glu802, which might be considered as the primary reason caused its decreased potency.
An online web service Protein Contacts Atlas (http://www.mrc-lmb.cam.ac.uk/pca/ (accessed on September 2020)) was also used for validating key residues interacted with the ligand febuxostat in XO protein [25]. It could be seen from Figure S2 that there were an inner circle and an outer circle, representing directly and indirectly interacted residues with febuxostat (central blue note), respectively [26]. The circle size of each residue represented the number of atomic contacts. It is worth noting that Phe914, Glu802, Phe1009, Arg880, Leu648, and Thr1010 might play important roles for febuxostat binding, consistent with the docking results.

2.4. Pharmacophore Model

The pharmacophore models were built using different selections of 12 compounds, and the partial modeling results were displayed in Table S2. In order to make the pharmacophores more reasonable, febuxostat, which has structural similarity with ODCs, was given preference to construct models. The selection of remaining 11 molecules was followed the principles of relatively higher structural diversity and better activity. Twenty pharmacophore models were established by aligning and comparing the common features from a set of 12 active compounds (Table 1), and their statistical results were listed in Table 4. By comparing potential models based on different selections, the FEATS values were almost unaffected. If the modeling molecules possessed too high structural difference, it might cause poor N_HITS, low SPECIFICITY, and high ENERGY. However, too low structural difference might cause high SPECIFICITY, enrichment factor (EF), and Güner-Henry score (GH), which would be likely detrimental for further screening. The model_6 was considered as the optimal pharmacophore model as it gave relative fitting parameters, including SPECIFICITY = 5.709 (>5), N_HITS ≈ 12, FEATS = 8, PARETO = 0, ENERGY = 9.41, STERICS = 1864.20, HBOND = 486.00, and MOL_QRY = 104.79 [27]. Additionally, the calculated parameters of a decoy set method for the model_6 could be concluded as follows: GH = 0.75 (0.6 < GH < 0.8) and EF = 120.56 (> 1). These also indicated that the model_6 has powerful ability to differentiate benign from inert compounds [28]. Therefore, the model_6 was chosen to analyze the key pharmacological features and was applied for the following virtual screening.
The pharmacophore features were shown in Figure 7 with the alignment of 12 XOIs. There were 3 green, 2 magenta, 2 cyan, and 1 blue spheres, which represented 3 hydrogen-bond acceptor atoms (HAs), 2 hydrogen-bond donor atoms (HDs), 2 hydrophobic centers (HYs), and 1 negative center (NC), respectively. It could be observed that the HDs covered the nitrogen atoms of the pyridine ring, suggesting that the nitrogen-containing heterocyclic structure might be indispensable for the activity. There were 2 HAs at the oxygen atoms of -OR1 and the carboxyl group, respectively, and another HA at the tetrazole or the cyano group, indicating that HA groups might be essential for activity in this position. The common scaffolds of febuxostat analogues were reported to be necessary to maintain the activity, which was consistent with the pharmacophore results that 2 HYs were located at the center of the 2 aromatic rings. These pharmacophore characteristics were generally in agreement with the 3D-QSAR and docking results. A graphical SAR summary of these ODC XOIs based on the results of the 3D-QSAR models, the molecular dockings, and the optimal pharmacophore model was shown in Figure 8.

2.5. Virtual Screening and Docking Analysis

The best pharmacophore model was converted into a UNITY query for the virtual screening from the ZINC purchasable database. The screened compounds (QFIT > 35) from the pharmacophore were then docked into the XO protein, and 37 compounds (docking score > 10) were obtained. We found that several screened compounds contained a similar indole ring that corresponded to the benzene ring of febuxostat analogues, which might provide new ideas for designing novel XOIs. The ADME properties of screened hits, compound 44, and febuxostat were predicted by SwissADME (http://www.swissadme.ch (accessed on December 2019)), and a portion of data was summarized in Table 5 and Table S3. Finally, the screened hits VS12 (ZINC71763654), VS16 (ZINC09234838), VS19 (ZINC59369278), and VS26 (ZINC16688904) were obtained through the following criteria: 150 < molecular weight (MW) < 500 g/mol, 20 < topological polar surface area (TPSA) < 130 Å2, high gastrointestinal (GI) absorption, inexistent blood–brain barrier (BBB) permeability, synthetic accessibility (SA) score < 4.5, Lipinski violations = 0, and lower inhibitory characteristics with cytochrome P450 (CYP450) [21,29,30,31].
The chemical structures and docking scores of compounds VS12, VS16, VS19, and VS26 were summarized in Table 6, and their docking conformations were shown in Figure 9. Compounds VS12, VS16, and VS26 could form similar hydrogen bonds with residues Arg880, Thr1010, and Glu802, which were similar with the docking results of febuxostat and compound 44. In the previous docking results, residues Arg880, Thr1010, and Glu802 have also been proved to be essential for XOI binding. As for compound VS19, it could be observed to bind stably by hydrogen bonds with Arg880, Thr1010, Asn768, and Leu648. Compared with febuxostat and compound 44, the nitrogen-containing heterocyclic rings of compounds VS12, VS16, VS19, and VS26 were also observed to form π-π stacking interactions with Phe914 and Phe1009. These docking results demonstrated that the 4 screened compounds might have the potential to become novel XOIs, which might provide new ideas for designing novel XOI chemotypes.

2.6. Molecular Dynamics Simulations

The best docking conformations of compounds febuxostat, 44, VS12, VS16, VS19, and VS26 were subjected to MD simulations based on the repaired XO protein. To further analyze the stability of these protein–ligand complexes, the time-dependent behavior of complexes XO-VS12, XO-VS16, XO-VS19, and XO-VS26 in the 50 ns simulation trajectories were analyzed using XO-febuxostat and XO-44 as the comparisons, and their results were shown in Figure 10 and Figure 11, Figures S3–S5.
The RMSD values that could reflect the differences between the initial and current conformations were suited to evaluate the convergence and stability of the systems [32]. The comparisons between the initial and final conformations of 6 complexes were shown in Figure S3. According to Figure 10a, the RMSD values of backbone atoms in the 6 complexes could become stable at approximately 25 ns and varied around 0.2 nm. In complex XO-febuxostat, the RMSD values of backbone atoms were higher than those of the other complexes. As shown in Figure 10b, the ligands (febuxostat, 44, VS12, VS16, VS19, and VS26) with low RMSD values could become stable during MD simulations. Compared to febuxostat, the RMSD results manifested that compounds VS12, VS16 and VS26 might be more favorable for binding to XO.
The root-mean-square fluctuation (RMSF) values (Figure 10c) were used to evaluate the flexibility of the protein side chains, and the RMSF values of >0.35 nm were considered as high flexibility [33]. The RMSF values of the most residues in the 6 systems showed similar fluctuation, and the critical residues in all complexes exhibited low flexibility, illustrating that the key interactions of all compounds in the binding pocket might similar. The gyration radius (Rg) values could estimate the protein compactness, and the numerical results of each system over time were summarized in Figure 10d. All complexes showed slight fluctuations, and the Rg values finally stabilized between 3.17 and 3.21 nm. The above evidence suggested that the protein conformations of all complexes were basically stable, which was consistent with the RMSD results [28].
The hydrogen-bond number were carried out to assess the stability of the complexes from another perspective [34]. Figure S4 showed the hydrogen-bond numbers between the ligand and protein in complexes XO-febuxostat, XO-44, XO-VS12, XO-VS16, XO-VS19, and XO-VS26 fluctuated within 0–4, 0–6, 0–7, 0–4, 0–5, and 0–4, respectively, during the MD simulations. Furthermore, the hydrogen-bond numbers in complexes XO-febuxostat, XO-44, XO-VS12, XO-VS16, XO-VS19, and XO-VS26 maintained in 2, 3, 3, 2, 1, and 1, respectively. The above results suggested that compound VS12 might bind tightly with XO by more hydrogen bonds than the other virtual-screened compounds.
The binding free energy was calculated by the Molecular Mechanics-Poisson Bolzmann Surface Area (MM-PBSA) method, and the corresponding values are listed in Table 7. The binding free energies of compounds febuxostat, 44, VS12, VS16, VS19, and VS26 in the XO protein were −97.04, −95.30, −88.31, −159.71, −95.91, and −150.84 kJ/mol, respectively. These results indicated that the binding strengths of compounds VS16 and VS26 might be stronger than those of compounds febuxostat and 44 [35]. It was observed that binding energies of VS16 and VS26 with XO were more favorable due to their higher van der Waals interactions (ΔEvdw: VS16 = −220.50 kJ/mol; VS26 = −216.89 kJ/mol). In addition, VS16 exhibited relatively low polar solvation energy (ΔGPB: 109.67 kJ/mol), and VS26 showed relatively good electrostatic interaction (ΔEele: −40.34 kJ/mol), respectively. As for all 6 systems, the ΔEvdw played an important role in the ΔGbinding, the contribution of ΔEele was nearly counteracted by the ΔGPB, and the values of SASA energy (ΔGSA) in each complex were semblable. These results indicated that the nonpolar energy might be the main driving force for the binding of these XOIs. To understand the stability of complexes, the binding energies obtained during the MD simulations versus time were shown in Figure S5. It was observed that the binding energies of all complexes reached stability after 45 ns. Essentially, the stability of compounds VS12, VS16, VS19, and VS26 in the XO binding pocket were verified. The above analyses proved that the screened compounds might have powerful potential to be XOI hits.

3. Materials and Methods

3.1. Molecules Set and Optimization

All calculations and simulations were performed using SYBYL-X 2.1 software (Tripos Inc., St. Louis, MO, USA) running on Windows 7 workstations, unless otherwise specified. A dataset of 46 ODCs (Table 1) was selected from the published literature [3,4] for 3D-QSAR modelling. The energy and geometry minimizations of all ODC compounds were applied in the Tripos force field and the Gasteiger–Hückel charges under the energy gradient convergence criterion of 0.005 kcal/(mol·Å) and the maximum iteration coefficient of 10,000 [36].

3.2. Molecular Modeling and Molecular Alignment

The CoMFA and CoMSIA models were used to help us to better understand the relationship between the characteristics of molecular structures and their activities. SYBYL-X 2.1 software was used for the CoMFA and CoMSIA models. Different physicochemical properties used for the CoMFA and CoMSIA models are calculated by the same lattice boxes with the same sp3 carbon probe (default probe atom) [28]. Thereinto, the CoMFA model incorporates two different descriptor fields: steric and electrostatic fields. The steric and electrostatic fields were calculated using Lennard-Jones potential and Coulombic potential, respectively [37]. In CoMSIA, five different similarity fields, including steric, electrostatic, hydrophobic, HBD, and HBA fields, were calculated using the Gaussian function [38]. Besides, the CoMFA and CoMSIA models are transformed to contour maps using the field type of “StDev*Coeff”, which are useful for analyzing the SARs [6]. Forty-six compounds used for the 3D-QSAR models were randomly divided into a training set of 35 molecules for the model generation and a test set of 11 molecules for the model validation. The molecular alignment quality has a great effect on the predictability and robustness of the models [39]. In this study, the compound 44 with the highest potency was selected as a template, and the other molecules were superimposed onto the common skeleton (colored as magenta) by automatic alignment and manual adjustment. The alignment results of all ODCs were shown in Figure 11.

3.3. Model Validation

PLS regression analysis method was performed for establishing the 3D-QSAR models. The first step detecting the reproducibility and robustness of the CoMFA and CoMSIA models was the internal validation. In the PLS analysis, a series of parameters, including q2, ONC, R2, F, and SEE, were calculated to assess the predictive ability of models [36]. To further judge the feasibility of the constructed models, external validation parameters, including r02, r02, k, k’, rm2, r’m2, rpred2, Δrm2, r m 2 ¯ , and RMSE, were further taken into consideration [39]. r02 (predicted vs. actual pIC50) and r02 (actual vs. predicted pIC50) are the correlation coefficients of regression lines with a zero intercept, and k (predicted vs. actual pIC50) and k’ (actual vs. predicted pIC50) are the slopes of regression lines, respectively. rm2, rpred2, and RMSE are calculated according to the following Equations (1)–(3), respectively.
r m 2 = r 2 × 1 r 2 r 0 2  
r p r e d 2 = 1 y p r e d ( t e s t ) y ( t e s t ) 2 y ( t e s t ) y ( t r a i n i n g ) ¯ 2
R M S E = 1 n y ( t e s t ) y p r e d ( t e s t ) 2
The y(test), ypred(test), and y ( training ) ¯ represent the actual pIC50 value of each test set compound, the predicted pIC50 value of each test set compound, and the mean pIC50 value of the training set compounds, respectively [40]. An appropriate model should satisfy the following conditions: q2 > 0.5, R2 > 0.8, r 2 r 0 2 or   r 0 2 r 2 < 0.1, 0.85 ≤ k (or k’) ≤ 1.15, Δrm2 < 0.2, r m 2 ¯ > 0.5, and rpred2 > 0.6 [41].

3.4. Molecular Docking

Molecular docking served as a helpful tool to obtain the reasonable binding conformations of bioactive molecules and to identify core residues in the active site of target protein. The crystal structure of bovine XO protein (PDB ID: 1N5X), a very close homologue of human XO enzyme, was used for molecular docking by the surflex-docking package of SYBYL-X 2.1 with default parameters [1]. The sequence alignment of bovine (Bos taurus) and human (Homo sapiens) XO with approximately 90% sequence identity was shown in Figure S6, and particularly in the febuxostat binding site, the key amino acids were the same, which was consistent with the reported literatures [1,20]. Before docking, an online web service (http://www.mrc-lmb.cam.ac.uk/pca/ (accessed on September 2020)) was used to explore the non-covalent contacts between ligand and protein [25]. After the pretreatment steps of the original protein, including hydrogenating, adding electric charges, extracting the crystallographic ligands, and removing water and other unnecessary atoms, the applicable docking pocket was generated with a threshold of 5 Å by a “ligand” mode. The crystallographic ligand was first redocked into the pocket to examine the dependability of the docking method. The conformation differences between the redocked and original ligands were evaluated by the RMSD values [17]. The RMSD < 2.0 Å is considered as a reference criterion, indicating that the docking method is reasonable. The same docking method was then applied on 46 ODC XOIs (Table 1), and the appropriate docking conformations with different docking scores were then obtained [42]. Finally, the conformations of febuxostat, the most active compound 44, and the least active compound 26 were used for further analyses. The docking visualization was completed by PyMOL software (DeLano Scientific LLC, San Carlos, California, USA).

3.5. Pharmacophore Model

Pharmacophore model could be used to extract the key chemical information of active compounds, and to automatically generate pharmacophore features, for instance, HAs, HDs, HYs, and NCs [23]. Twelve molecules (Table 1) containing febuxostat with relatively high activities and diverse structures were selected to construct pharmacophore models by the Genetic Algorithm with Linear Assignment of Hypermolecular Alignment of Datasets (GALAHAD) module of SYBYL-X 2.1. Twenty models with different parameters were generated, of which the models with optimal SPECIFICITY, N_HITS, ENERGY, STERICS, HBOND, and MOL_QRY values were chosen for further studies. In addition, a decoy set method was used to assess the searching active molecules capability of pharmacophore models. The potential pharmacophore models were performed to screen a decoy set database. The decoy set database was composed of 6234 inactive compounds (downloaded from http://dud.docking.org/r2/ (accessed on September 2019)) and 35 active ODCs (Table 1), except for the 11 ODCs and febuxostat that used for constructing this model [28]. The EF and GH values used for evaluating the reliability of the models were calculated as follows:
E F = H a / H t A / D
G H = H a ( 3 A + H t ) 4 H t A 1 H t H a D A
in which Ha, Ht, A, and D represent the number of true positive compounds in the hit list, the number of all compounds in the hit list, the number of true positive compounds in the database, and the number of all compounds in the database, respectively [43]. When EF > 1 and 0.6 < GH < 0.8, the model is eligible to be chosen for further analyses and virtual screening.

3.6. Virtual Screening

A multi-stage virtual screening was carried out against the purchasable ZINC15 database (http://zinc15.docking.org (accessed on September 2019)) [44] by the combination of the optimal pharmacophore model, molecular dockings, and ADME predictions. The pharmacophore features from the best pharmacophore model were extracted to use as a search query for the first-round screening. The QFIT parameters in the range of 0 to 100 were obtained to assess the matching degree of screened compounds with the pharmacophore features [33]. The compounds with a QFIT > 35 were then selected for the second-round screening of docking. The compounds with a docking score > 10 were selected for the third-round screening [13]. In the third-round screening, the ADME profiling was then applied to assess pharmacokinetic and pharmacodynamic properties of the second-round screened compounds by a web tool of SwissADME (http://www.swissadme.ch (accessed on December 2019)) [30]. Among the predicted ADME properties, following properties were considered preferentially to get the satisfied compounds, including MW, TPSA, GI absorption, BBB permeability, inhibitory ability assessment of CYP450, Lipinski violations, and SA score [10,29,30,31]. Finally, the hits compounds with desired pharmacophore compliance, preferable docking scores, and ideal ADME evaluation results were further investigated for their binding interactions and stability in the XO protein by MD simulations and post-analysis experiments.

3.7. Molecular Dynamics Simulations

To further verify the stability of the docked hits compounds in the XO protein, 50 ns MD simulations were performed on different complexes using GROMACS 2016.5 software (Uppsala University, Stockholm University, and the Royal Institute of Technology, Sweden). The overall XO protein was a homodimer, in which one of subunit was selected for simulations. Residue deficiencies in 166–191 and 532–536 (loop region) were repaired by the protein loop search module of SYBYL-X 2.1 software, and the residues 1326–1332 were reasonably removed to solve the problem of the residues 1317–1325 deficiencies in the subunit terminal. The repaired models with good fit and high homology were further evaluated using PROCHECK and ProSa servers. Compared with the evaluation results of the original and optimal repaired proteins (Figure S7), the residues in disallowed regions were unchanged, and the Z-score value showed little difference, indicating the repaired protein was eligible enough for simulations. The protein topology files were generated by the pdb2gmx program under the AMBER99SB force field [1]. The ligand topological files were obtained by the ACPYPE program [35]. The protein–ligand complexes were positioned into the center of a cubic box with a side length of 13.7 nm, and there was a buffering distance of approximately 12 Å between the protein periphery and the box edges. This box filled with water, and five additional chloride ions, were added into box to reach the neutralization. The steepest gradient descent method was then used to minimize the energy of each system within 1 ns (50,000 steps) with a convergence criterion of 10 kJ/mol [32]. The protein–ligand complex was subjected to 100 ps simulation to achieve the NVT and NPT equilibrium at 300 K and 1 atm, respectively [28]. Finally, 50 ns MD simulations were performed for all systems, and their trajectories were recorded at every 10 ps (5000 steps) for post-analysis [26]. Many post-analyses, such as RMSD, RMSF, Rg, and hydrogen-bond numbers, were conducted to investigate the stability or variation of all complexes during the dynamic environment. The equilibrium trajectory (the last 5 ns) of the MD simulations was extracted to calculate the binding free energy using the MM-PBSA method. The binding free energy was calculated as follows:
Δ G b i n d = G c o m p l e x G f r e e p r o t e i n G f r e e l i g a n d
in which Gcomplex, Gfree-protein, and Gfree-ligand represent the free energy of protein–ligand complex, protein, and ligand, respectively [33].

4. Conclusions

In this work, an integrated computational study, including 3D-QSAR models, molecular dockings, pharmacophore models, and MD simulations, was performed on 46 novel ODC XOIs. 3D-QSAR models with good statistical parameters were constructed to provide significant insights into the SARs of these ODCs. Molecular docking results indicated that residues Glu802, Arg880, Thr1010, and Asn768 could form hydrogen bonds with these XOIs, and residues Phe914 and Phe1009 could also form π-π interactions with them. These interactions might be essential for their affinity with the XO protein. The best pharmacophore model with eight features was inconsistent with the 3D-QSAR and docking results. Four hit compounds (VS12, VS16, VS19, and VS26) with ideal compatibility for the pharmacophore model, relatively high docking scores, and good ADME characteristics were retrieved by the multi-round virtual screening. The MD simulations also indicated hits VS12, VS16, VS19, and VS26 could bind well with the XO protein during the dynamic environment. We expect that these results will be helpful for the design and development of novel XOIs, and the screened compounds could provide more alternatives and ideas for XOIs.

Supplementary Materials

Supplementary materials are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ijms22158122/s1.

Author Contributions

All authors contributed to the study conception and design. G.L. and N.Z. proposed the research idea and designed the experiment. N.Z. performed the experiment. N.Z., C.W., L.X., and F.W. analyzed the data. N.Z., L.X., X.L., X.J., and G.L. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (21807082), the Outstanding Young and Middle-aged Scientific Innovation Team of Colleges and Universities of Hubei Province: “Biomass chemical technologies and materials” (T201908), Special Projects of the Central Government in Guidance of Local Science and Technology Development in Hubei Province (2020ZYYD040), and the second batch of the Key Research and Development Project of Hubei Province (2020BAB073).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Chen, Y.; Gao, Y.; Wu, F.; Luo, X.; Ju, X.; Liu, G. Computationally exploring novel xanthine oxidase inhibitors using docking-based 3D-QSAR, molecular dynamics, and virtual screening. N. J. Chem. 2020, 44, 19276–19287. [Google Scholar] [CrossRef]
  2. Burns, C.M.; Wortmann, R.L. Gout therapeutics: New drugs for an old disease. Lancet 2011, 377, 165–177. [Google Scholar] [CrossRef]
  3. Zhang, B.; Dai, X.; Bao, Z.; Mao, Q.; Duan, Y.; Yang, Y.; Wang, S. Targeting the subpocket in xanthine oxidase: Design, synthesis, and biological evaluation of 2-[4-alkoxy-3-(1H-tetrazol-1-yl)phenyl]-6-oxo-1,6-dihydropyrimidine-5-carboxylic acid derivatives. Eur. J. Med. Chem. 2019, 181, 111559. [Google Scholar] [CrossRef]
  4. Mao, Q.; Dai, X.; Xu, G.; Su, Y.; Zhang, B.; Liu, D.; Wang, S. Design, synthesis and biological evaluation of 2-(4-alkoxy-3-cyano)phenyl-6-oxo-1,6-dihydropyrimidine-5-carboxylic acid derivatives as novel xanthine oxidase inhibitors. Eur. J. Med. Chem. 2019, 181, 111558. [Google Scholar] [CrossRef]
  5. Guan, Q.; Cheng, Z.; Ma, X.; Wang, L.; Feng, D.; Cui, Y.; Bao, K.; Wu, L.; Zhang, W. Synthesis and bioevaluation of 2-phenyl-4-methyl-1,3-selenazole-5-carboxylic acids as potent xanthine oxidase inhibitors. Eur. J. Med. Chem. 2014, 85, 508–516. [Google Scholar] [CrossRef]
  6. Tang, H.; Zhao, D. Studies of febuxostat analogues as xanthine oxidase inhibitors through 3D-QSAR, Topomer CoMFA and molecular modeling. J. Iran. Chem. Soc. 2019, 16, 2659–2671. [Google Scholar] [CrossRef]
  7. Luna, G.; Dolzhenko, A.V.; Mancera, R.L. Inhibitors of xanthine oxidase: Scaffold diversity and structure-based drug design. ChemMedChem 2019, 14, 714–743. [Google Scholar] [CrossRef] [Green Version]
  8. Mehmood, A.; Ishaq, M.; Zhao, L.; Safdar, B.; Rehman, A.U.; Munir, M.; Raza, A.; Nadeem, M.; Iqbal, W.; Wang, C. Natural compounds with xanthine oxidase inhibitory activity: A review. Chem. Biol. Drug Des. 2019, 93, 387–418. [Google Scholar] [CrossRef] [PubMed]
  9. Li, S.; Zhang, T.; Wu, Q.; Olounfeh, K.M.; Zhang, Y.; Meng, F. Synthesis and biological evaluation of 5-benzyl-3-pyridyl-1H-1,2,4-triazole derivatives as xanthine oxidase inhibitors. Med. Chem. 2020, 16, 119–127. [Google Scholar] [CrossRef]
  10. Casas, A.I.; Nogales, C.; Mucke, H.A.M.; Petraina, A.; Cuadrado, A.; Rojo, A.I.; Ghezzi, P.; Jaquet, V.; Augsburger, F.; Dufrasne, F.; et al. On the clinical pharmacology of reactive oxygen species. Pharm. Rev. 2020, 72, 801–828. [Google Scholar] [CrossRef] [PubMed]
  11. Dhiman, R.; Sharma, S.; Singh, G.; Nepali, K.; Singh Bedi, P.M. Design and synthesis of aza-flavones as a new class of xanthine oxidase inhibitors. Arch. Pharm. 2013, 346, 7–16. [Google Scholar] [CrossRef] [PubMed]
  12. Smelcerovic, A.; Tomovic, K.; Smelcerovic, Z.; Petronijevic, Z.; Kocic, G.; Tomasic, T.; Jakopin, Z.; Anderluh, M. Xanthine oxidase inhibitors beyond allopurinol and febuxostat; an overview and selection of potential leads based on in silico calculated physico-chemical properties, predicted pharmacokinetics and toxicity. Eur. J. Med. Chem. 2017, 135, 491–516. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, T.; Zhang, Y.; Tu, S.; Wu, Y.; Zhang, Z.; Meng, F. Design, synthesis and biological evaluation of N-(3-(1H-tetrazol-1-yl)phenyl)isonicotinamide derivatives as novel xanthine oxidase inhibitors. Eur. J. Med. Chem. 2019, 183, 111717. [Google Scholar] [CrossRef] [PubMed]
  14. Li, X.; Zhou, H.; Mo, X.; Zhang, L.; Li, J. In silico study of febuxostat analogs as inhibitors of xanthine oxidoreductase: A combined 3D-QSAR and molecular docking study. J. Mol. Struct. 2019, 1181, 428–435. [Google Scholar] [CrossRef]
  15. Okamoto, K.; Eger, B.T.; Nishino, T.; Kondo, S.; Pai, E.F.; Nishino, T. An extremely potent inhibitor of xanthine oxidoreductase. Crystal structure of the enzyme-inhibitor complex and mechanism of inhibition. J. Biol. Chem. 2003, 278, 1848–1855. [Google Scholar] [CrossRef] [Green Version]
  16. Okafor, O.N.; Farrington, K.; Gorog, D.A. Allopurinol as a therapeutic option in cardiovascular disease. Pharmacol. Ther. 2017, 172, 139–150. [Google Scholar] [CrossRef] [PubMed]
  17. Malik, N.; Dhiman, P.; Khatkar, A. In silico and 3D QSAR studies of natural based derivatives as xanthine oxidase inhibitors. Curr. Top. Med. Chem. 2019, 19, 123–138. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, T.; Lv, Y.; Lei, Y.; Liu, D.; Feng, Y.; Zhao, J.; Chen, S.; Meng, F.; Wang, S. Design, synthesis and biological evaluation of 1-hydroxy-2-phenyl-4-pyridyl-1H-imidazole derivatives as xanthine oxidase inhibitors. Eur. J. Med. Chem. 2018, 146, 668–677. [Google Scholar] [CrossRef]
  19. Xu, X.; Deng, L.; Nie, L.; Chen, Y.; Liu, Y.; Xie, R.; Li, Z. Discovery of 2-phenylthiazole-4-carboxylic acid, a novel and potent scaffold as xanthine oxidase inhibitors. Bioorg. Med. Chem. Lett. 2019, 29, 525–528. [Google Scholar] [CrossRef] [PubMed]
  20. Singh, J.V.; Mal, G.; Kaur, G.; Gupta, M.K.; Singh, A.; Nepali, K.; Singh, H.; Sharma, S.; PM, S.B. Benzoflavone derivatives as potent antihyperuricemic agents. Medchemcomm 2019, 10, 128–147. [Google Scholar] [CrossRef]
  21. Yang, Y.; Zhang, L.; Tian, J.; Ye, F.; Xiao, Z. Identification of xanthine oxidase inhibitors through hierarchical virtual screening. RSC Adv. 2020, 10, 27752–27763. [Google Scholar] [CrossRef]
  22. Singh, J.V.; Bedi, P.M.S.; Singh, H.; Sharma, S. Xanthine oxidase inhibitors: Patent landscape and clinical development (2015–2020). Expert Opin. Ther. Pat. 2020, 30, 769–780. [Google Scholar] [CrossRef]
  23. Peng, J.; Li, Y.; Zhou, Y.; Zhang, L.; Liu, X.; Zuo, Z. Pharmacophore modeling, molecular docking and molecular dynamics studies on natural products database to discover novel skeleton as non-purine xanthine oxidase inhibitors. J. Recept. Signal Transduct. Res. 2018, 38, 246–255. [Google Scholar] [CrossRef] [PubMed]
  24. Li, X.; Liu, J.; Ma, L.; Fu, P. Pharmacological urate-lowering approaches in chronic kidney disease. Eur. J. Med. Chem. 2019, 166, 186–196. [Google Scholar] [CrossRef]
  25. Kayikci, M.; Venkatakrishnan, A.J.; Scott-Brown, J.; Ravarani, C.N.J.; Flock, T.; Babu, M.M. Visualization and analysis of non-covalent contacts using the Protein Contacts Atlas. Nat. Struct. Mol. Biol. 2018, 25, 185–194. [Google Scholar] [CrossRef]
  26. Gao, Y.; Wang, H.; Wang, J.; Cheng, M. In silico studies on p21-activated kinase 4 inhibitors: Comprehensive application of 3D-QSAR analysis, molecular docking, molecular dynamics simulations, and MM-GBSA calculation. J. Biomol. Struct. Dyn. 2020, 38, 4119–4133. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, G.; Wan, Y.; Wang, W.; Fang, S.; Gu, S.; Ju, X. Docking-based 3D-QSAR and pharmacophore studies on diarylpyrimidines as non-nucleoside inhibitors of HIV-1 reverse transcriptase. Mol. Divers. 2019, 23, 107–121. [Google Scholar] [CrossRef]
  28. Gao, Y.; Chen, Y.; Tian, Y.; Zhao, Y.; Wu, F.; Luo, X.; Ju, X.; Liu, G. In silico study of 3-hydroxypyrimidine-2,4-diones as inhibitors of HIV RT-associated RNase H using molecular docking, molecular dynamics, 3D-QSAR, and pharmacophore models. N. J. Chem. 2019, 43, 17004–17017. [Google Scholar] [CrossRef]
  29. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A free web tool to evaluate pharmacokinetics, drug-likeness and medicinal chemistry friendliness of small molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef] [Green Version]
  30. Shaaban, S.; Vervandier-Fasseur, D.; Andreoletti, P.; Zarrouk, A.; Richard, P.; Negm, A.; Manolikakes, G.; Jacob, C.; Cherkaoui-Malki, M. Cytoprotective and antioxidant properties of organic selenides for the myelin-forming cells, oligodendrocytes. Bioorg. Chem. 2018, 80, 43–56. [Google Scholar] [CrossRef]
  31. Li, X.; Xu, Y.; Lai, L.; Pei, J. Prediction of human cytochrome P450 inhibition using a multitask deep autoencoder neural network. Mol. Pharm. 2018, 15, 4336–4345. [Google Scholar] [CrossRef]
  32. Gao, Y.; Zhang, Y.; Wu, F.; Pei, J.; Luo, X.; Ju, X.; Zhao, C.; Liu, G. Exploring the interaction mechanism of desmethyl-broflanilide in insect GABA receptors and screening potential antagonists by in silico simulations. J. Agric. Food Chem. 2020, 68, 14768–14780. [Google Scholar] [CrossRef]
  33. Liu, W.; Wang, R.; Sun, Y.; Li, W.; Li, H.; Liu, C.; Ma, Y.; Wang, R. Exploring the effect of inhibitor AKB-9778 on VE-PTP by molecular docking and molecular dynamics simulation. J. Cell Biochem. 2019, 120, 17015–17029. [Google Scholar] [CrossRef]
  34. Gerlt, J.A.; Kreevoy, M.M.; Cleland, W.W.; Frey, P.A. Understanding enzymic catalysis: The importance of short, strong hydrogen bonds. Chem. Biol. 1997, 4, 259–267. [Google Scholar] [CrossRef] [Green Version]
  35. Tian, Y.; Gao, Y.; Chen, Y.; Liu, G.; Ju, X. Identification of the fipronil resistance associated mutations in Nilaparvata lugens GABA receptors by molecular modeling. Molecules 2019, 24, 4116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Oh, Y.; Jung, H.; Kim, H.; Baek, J.; Jun, J.; Cho, H.; Im, D.; Hah, J. Design and synthesis of a novel PLK1 inhibitor scaffold using a hybridized 3D-QSAR model. Int. J. Mol. Sci. 2021, 22, 3865. [Google Scholar] [CrossRef] [PubMed]
  37. Cramer, R.D.; Patterson, D.E.; Bunce, J.D. Comparative molecular field analysis (CoMFA). 1. Effect of shape on binding of steroids to carrier proteins. J. Am. Chem. Soc. 1988, 110, 5959–5967. [Google Scholar] [CrossRef]
  38. Klebe, G. Comparative molecular similarity indices: CoMSIA. Perspect. Drug Discov. Des. 1998, 12, 87–104. [Google Scholar] [CrossRef]
  39. Wang, W.; Tian, Y.; Wan, Y.; Gu, S.; Ju, X.; Luo, X.; Liu, G. Insights into the key structural features of N1-ary-benzimidazols as HIV-1 NNRTIs using molecular docking, molecular dynamics, 3D-QSAR, and pharmacophore modeling. Struct. Chem. 2018, 30, 385–397. [Google Scholar] [CrossRef]
  40. Liu, G.; Wang, W.; Wan, Y.; Ju, X.; Gu, S. Application of 3D-QSAR, pharmacophore, and molecular docking in the molecular design of diarylpyrimidine derivatives as HIV-1 nonnucleoside reverse transcriptase inhibitors. Int. J. Mol. Sci. 2018, 19, 1436. [Google Scholar] [CrossRef] [Green Version]
  41. Wan, Y.; Tian, Y.; Wang, W.; Gu, S.; Ju, X.; Liu, G. In silico studies of diarylpyridine derivatives as novel HIV-1 NNRTIs using docking-based 3D-QSAR, molecular dynamics, and pharmacophore modeling approaches. RSC Adv. 2018, 8, 40529–40543. [Google Scholar] [CrossRef] [Green Version]
  42. Zhou, L.; Peng, J.; Wang, J.; Geng, Y.; Zuo, Z.; Hua, Y. Structure-activity relationship of xanthones as inhibitors of xanthine oxidase. Molecules 2018, 23, 365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Chen, Y.; Tian, Y.; Gao, Y.; Wu, F.; Luo, X.; Ju, X.; Liu, G. In silico design of novel HIV-1 NNRTIs based on combined modeling studies of dihydrofuro[3,4-d]pyrimidines. Front. Chem. 2020, 8, 164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Sterling, T.; Irwin, J.J. ZINC 15—Ligand Discovery for Everyone. J. Chem. Inf. Model. 2015, 55, 2324–2337. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of different XOIs. Febuxostat contains an aromatic moiety (blue) and a nitrogen heterocyclic ring (red). Other febuxostat analogues have different nitrogen-containing heterocyclic rings (magenta).
Figure 1. Chemical structures of different XOIs. Febuxostat contains an aromatic moiety (blue) and a nitrogen heterocyclic ring (red). Other febuxostat analogues have different nitrogen-containing heterocyclic rings (magenta).
Ijms 22 08122 g001
Figure 2. Scatter plots of actual versus predicted pIC50 values for the used XOIs based on the CoMFA (a) and CoMSIA-SEHDA (b) models.
Figure 2. Scatter plots of actual versus predicted pIC50 values for the used XOIs based on the CoMFA (a) and CoMSIA-SEHDA (b) models.
Ijms 22 08122 g002
Figure 3. Contour maps of the CoMFA and CoMSIA-SEHDA models using compound 44 as a reference. (a) The steric contours of the CoMFA model. (b) The electrostatic contours of the CoMFA model. (c) The steric contours of the CoMSIA model. (d) The electrostatic contours of the CoMSIA model.
Figure 3. Contour maps of the CoMFA and CoMSIA-SEHDA models using compound 44 as a reference. (a) The steric contours of the CoMFA model. (b) The electrostatic contours of the CoMFA model. (c) The steric contours of the CoMSIA model. (d) The electrostatic contours of the CoMSIA model.
Ijms 22 08122 g003
Figure 4. Contour maps of the CoMSIA-SEHDA model using compound 44 as a reference. (a) The hydrophobic field. (b) The hydrogen-bond donor field. (c) The hydrogen-bond acceptor field.
Figure 4. Contour maps of the CoMSIA-SEHDA model using compound 44 as a reference. (a) The hydrophobic field. (b) The hydrogen-bond donor field. (c) The hydrogen-bond acceptor field.
Ijms 22 08122 g004
Figure 5. The binding pocket of the XO protein (PDB ID: 1N5X). (a) The binding site of the XO protein (chain A, cyan cartoon; chain B, violet cartoon; binding pocket, gray surface). (b) The comparisons of febuxostat (orange sticks), compound 44 (cyan sticks), and compound 26 (yellow sticks) in the docking position of the binding pocket (gray surface).
Figure 5. The binding pocket of the XO protein (PDB ID: 1N5X). (a) The binding site of the XO protein (chain A, cyan cartoon; chain B, violet cartoon; binding pocket, gray surface). (b) The comparisons of febuxostat (orange sticks), compound 44 (cyan sticks), and compound 26 (yellow sticks) in the docking position of the binding pocket (gray surface).
Ijms 22 08122 g005
Figure 6. The docking results of febuxostat (a), compound 44 (b), and compound 26 (c) in the binding site of protein (PDB ID: 1N5X). The crystal ligand, redocked ligand, important residues, and hydrogen bond were shown in orange sticks, blue sticks, green sticks, and yellow dashes, respectively.
Figure 6. The docking results of febuxostat (a), compound 44 (b), and compound 26 (c) in the binding site of protein (PDB ID: 1N5X). The crystal ligand, redocked ligand, important residues, and hydrogen bond were shown in orange sticks, blue sticks, green sticks, and yellow dashes, respectively.
Ijms 22 08122 g006
Figure 7. The optimal pharmacophore model. Green, magenta, cyan, and blue spheres represent hydrogen-bond acceptor atoms (HAs), hydrogen-bond donor atoms (DAs), hydrophobes (HYs), and negative centers (NCs), respectively.
Figure 7. The optimal pharmacophore model. Green, magenta, cyan, and blue spheres represent hydrogen-bond acceptor atoms (HAs), hydrogen-bond donor atoms (DAs), hydrophobes (HYs), and negative centers (NCs), respectively.
Ijms 22 08122 g007
Figure 8. The SARs of 1,6-dihydropyrimidine-5-carboxylic acid XOIs based on the 3D-QSAR models, molecular docking results, and the pharmacophore model.
Figure 8. The SARs of 1,6-dihydropyrimidine-5-carboxylic acid XOIs based on the 3D-QSAR models, molecular docking results, and the pharmacophore model.
Ijms 22 08122 g008
Figure 9. The docking results of the screened compounds VS12 (a), VS16 (b), VS19 (c), and VS26 (d) in the XO protein (PDB ID: 1N5X). The ligands, important residues, and hydrogen bonds were shown in blue sticks, green sticks, and yellow dashes, respectively.
Figure 9. The docking results of the screened compounds VS12 (a), VS16 (b), VS19 (c), and VS26 (d) in the XO protein (PDB ID: 1N5X). The ligands, important residues, and hydrogen bonds were shown in blue sticks, green sticks, and yellow dashes, respectively.
Ijms 22 08122 g009
Figure 10. The results of the 50 ns MD simulations of complexes XO-febuxostat (black), XO-44 (red), XO-VS12 (green), XO-VS16 (cyan), XO-VS19 (magenta), and XO-VS26 (violet). (a) The RMSDs of the XO backbone atoms. (b) The RMSDs of ligands (febuxostat, 44, VS12, VS16, VS19, and VS26). (c) The residue RMSFs of XO. (d) The Rg of XO.
Figure 10. The results of the 50 ns MD simulations of complexes XO-febuxostat (black), XO-44 (red), XO-VS12 (green), XO-VS16 (cyan), XO-VS19 (magenta), and XO-VS26 (violet). (a) The RMSDs of the XO backbone atoms. (b) The RMSDs of ligands (febuxostat, 44, VS12, VS16, VS19, and VS26). (c) The residue RMSFs of XO. (d) The Rg of XO.
Ijms 22 08122 g010
Figure 11. Molecular alignment for the 3D-QSAR models. (a) The common scaffold (magenta) used for the alignment. (b) The alignment result of all used XOIs.
Figure 11. Molecular alignment for the 3D-QSAR models. (a) The common scaffold (magenta) used for the alignment. (b) The alignment result of all used XOIs.
Ijms 22 08122 g011
Table 1. Chemical structures of the used non-purine XOIs and their actual and predicted pIC50 values.
Table 1. Chemical structures of the used non-purine XOIs and their actual and predicted pIC50 values.
Ijms 22 08122 i001
No.R1R2R3IC50 (µM)pIC50CoMFACoMSIA
Predicted pIC50ResidualsPredicted pIC50Residuals
01 *methylHO0.09207.03626.9840.05227.0750.0388
02iso-propylHO0.07377.13257.1670.03457.1320.0005
03iso-butylHO0.06447.19117.2040.01297.1920.0009
04iso-pentylHO0.05417.26687.250.01687.2550.0118
05 *,#allylHO0.04377.35957.4070.04757.3290.0305
06iso-butenylHO0.05697.24497.2460.00117.2530.0081
07iso-pentenylHO0.06927.15997.1630.00317.1460.0139
08#propinylHO0.05007.3017.2630.0387.3020.001
09#methylene
cyclopropane
HO0.04617.33637.3240.01237.3280.0083
10cyclopentylHO0.05857.23287.2750.04227.2350.0022
11methylene
cyclohexane
HO0.06837.16567.1430.02267.1440.0216
12benzylHO0.09457.02467.1020.07747.0180.0066
13p-methylbenzylHO0.08947.04877.0740.02537.1060.0573
14 *p-tert-butylbenzylHO0.14906.82686.8040.02286.8010.0258
15p-methoxylbenzylHO0.05077.2957.3300.0357.3010.006
16p-fluorobenzylHO0.05317.27497.2260.04897.2110.0639
17p-chlorobenzylHO0.06917.16057.1950.03457.2010.0405
18p-bromobenzylHO0.05527.25817.1940.06417.2330.0251
19m-methoxylbenzylHO0.05167.28747.2730.01447.3010.0136
20m-fluorobenzylHO0.04777.32157.2980.02357.3190.0025
21 *,#m-chlorobenzylHO0.02887.54067.5930.05247.5350.0056
22m-bromobenzylHO0.04507.34687.3290.01787.3420.0048
23o-chlorobenzylHO0.09177.03767.0620.02447.0040.0336
242,5-dichlorobenzylHO0.06397.19457.1200.07457.2150.0205
252,4-dichlorobenzylHO0.08387.07687.1290.05227.1160.0392
26hydrogenHO0.62906.20136.2050.03136.2010.0003
27iso-propylHO0.09167.03817.0130.02227.0420.0039
28 *iso-butylHO0.06097.21547.2380.02267.2700.0546
29 *,#iso-pentylHO0.02507.60217.5610.04117.7050.1029
30 *allylHO0.08117.0917.1070.0167.0260.065
31 #iso-butenylHO0.03367.47377.5050.00617.5070.0333
32iso-pentenylHO0.03887.41127.3890.10317.4290.0178
33 *benzylHO0.03877.41237.2970.11537.4050.0073
34p-fluorobenzylHO0.03827.41797.4240.09987.3980.0199
35p-chlorobenzylHO0.04997.30197.4050.067.4100.1081
36 #p-bromobenzylHO0.02987.52587.4260.01167.4620.0638
37 *,#p-tert-butylbenzylHO0.19706.70556.8740.16856.7870.0815
38p-methylbenzylHO0.03547.4517.3910.067.3800.071
39iso-pentylCH3O0.54006.26766.2560.01166.2480.0196
40iso-butenylCH3O0.56776.24596.2530.00716.2630.0171
41 *p-bromobenzylCH3O0.18546.73196.8220.09016.5120.2199
42 *p-methylbenzylCH3O0.15906.79866.8070.00846.4690.3296
43 #iso-pentylHNH0.02407.61987.6640.04427.6420.0222
44 #iso-butenylHNH0.01817.74237.6840.05837.7080.0343
45 #p-bromobenzylHNH0.02717.5677.580.0137.5580.009
46p-methylbenzylHNH0.03397.46987.5280.05827.4870.0172
Febuxostat #---0.02367.6271----
* The test set compounds used for the 3D-QSAR models. # The compounds used for the pharmacophore models.
Table 2. Internal statistical parameters of the CoMFA and CoMSIA models.
Table 2. Internal statistical parameters of the CoMFA and CoMSIA models.
Modelq2ONCSEER2Frpred2Field Contribution (%)
SEHDA
CoMFAS + E0.89770.0500.983229.500.9480.7730.227
CoMSIAS + E + H + D + A0.922110.0410.990212.260.8400.1050.2480.3720.1930.082
H + D0.907120.0480.987141.510.670 0.7630.237
S + E + D0.93170.0550.980189.710.6650.1750.404 0.421
E + H + A0.875110.0430.989195.520.871 0.3310.528 0.141
S + E + H + D0.926100.0410.990232.650.8250.1070.2570.4220.214
E + H + D + A0.921110.0420.990205.7110.848 0.2710.4360.2040.089
q2: cross-validated correlation coefficient; ONC: optimal number of components; SEE: standard error of estimate; R2: non-cross-validated correlation coefficient; F: F-statistic values; rpred2: predictive correlation coefficient; S: steric fields; E: electrostatic fields; H: hydrophobic fields; D: hydrogen bond donor fields; A: hydrogen bond acceptor fields.
Table 3. External validation parameters of the CoMFA and CoMSIA models.
Table 3. External validation parameters of the CoMFA and CoMSIA models.
Validation ParametersRMSEr2r02r’02(r2 − r’02)/r2kk’rm2r’m2Δrm2 r m 2 ¯
CoMFA0.0740.9500.9460.9360.01470.9981.0020.8900.8380.0520.864
CoMSIA (S + E + H + D + A)0.1300.9220.8400.8970.0271.0050.9950.6580.7760.1180.717
RMSE: root mean square error for the test set compounds; r2: the regression line coefficient of correlation for the test set compounds; r02 (predicted vs. observed activities) and r02 (observed vs. predicted activities): the correlation coefficient of regression lines with a zero intercept; k (predicted vs. observed activities) and k′ (observed vs. predicted activities): the slope of regression lines with a zero intercept; rm2: calculated by [r2 (1−(r2 − r02)0.5)]; r’m2: calculated by [r2 (1−(r2 − r02)0.5)]; Δrm2 and r m 2 ¯ : the difference and average values between rm2 and rm2.
Table 4. Statistical results of the pharmacophore models.
Table 4. Statistical results of the pharmacophore models.
NameSPECIFICITYN_HITSFEATSPARETOENERGYSTERICSHBONDMOL_QRY
Model_14.62911808.762052.10497.6084.69
Model_24.63011809.532064.40498.4081.58
Model_32.33091008.472023.00499.9065.32
Model_44.651118011.391986.30494.5087.28
Model_54.629118010.101911.90492.0091.15
Model_65.70911809.411864.20486.00104.79
Model_73.4401210011.681834.50495.80117.25
Model_84.630118010.801840.80503.1085.90
Model_93.4401210012.211835.30494.70117.25
Model_104.629118022.582149.70496.1085.45
Model_114.628118010.471965.60490.2085.45
Model_124.649118012.181996.40487.4093.3
Model_134.63011801240.812059.40497.3085.90
Model_144.630118010.801786.20499.5085.90
Model_153.440121009.891780.70484.10141.11
Model_163.4401210014.361750.50492.50152.73
Model_173.4391210010.261929.00470.00129.80
Model_184.637118010.821830.30487.9097.00
Model_193.4411210010.191723.60485.60130.83
Model_204.63611801385.322189.60481.2097.00
Table 5. The ADME prediction results of the virtual-screened hits (VS12, VS16, VS19, and VS26), compound 44, and febuxostat.
Table 5. The ADME prediction results of the virtual-screened hits (VS12, VS16, VS19, and VS26), compound 44, and febuxostat.
ParameterCompound
VS12VS16VS19VS26Febuxostat44
MW (g/mol)375.4417.39401.44429.4316.37310.31
Fraction Csp30.30.090.220.350.310.12
Rotatable bonds478955
TPSA (Å2)113.4105.798.6130.3111.5122.9
GI absorptionHighHighHighHighHighHigh
BBB permeantNoNoNoNoNoNo
CYP1A2 inhibitorYesNoYesNoYesNo
CYP2C19 inhibitorNoYesNoYesYesNo
CYP2C9 inhibitorNoNoNoNoYesNo
CYP2D6 inhibitorNoNoNoNoNoNo
CYP3A4 inhibitorNoNoNoNoNoNo
Lipinski violations000000
SA score3.43.323.253.343.122.64
Table 6. Chemical structures and docking scores of the virtual-screened hits obtained by the third-round screening.
Table 6. Chemical structures and docking scores of the virtual-screened hits obtained by the third-round screening.
Hit
Compound
VS12VS16VS19VS26
Structure Ijms 22 08122 i002 Ijms 22 08122 i003 Ijms 22 08122 i004 Ijms 22 08122 i005
Docking score10.6211.8710.7611.03
Table 7. The free binding energies (kJ/mol) of febuxostat, compound 44, and the virtual-screened hits (VS12, VS16, VS19, and VS26) in the XO protein.
Table 7. The free binding energies (kJ/mol) of febuxostat, compound 44, and the virtual-screened hits (VS12, VS16, VS19, and VS26) in the XO protein.
ComplexΔEvdWΔEeleΔGPBΔGSAΔGbinding
XO-febuxostat−175.32 ± 8.81−39.95 ± 6.37136.43 ± 11.05−18.20 ± 0.73−97.04 ± 9.45
XO-44−160.03 ± 10.29−107.25 ± 10.15189.25 ± 11.17−17.27 ± 0.78−95.30 ± 8.32
XO-VS12−195.31 ± 9.21−69.99 ± 13.91197.26 ± 13.45−20.27 ± 0.81−88.31 ± 10.40
XO-VS16−220.50 ± 9.93−28.10 ± 8.40109.67 ± 10.71−20.78 ± 0.86−159.71 ± 11.41
XO-VS19−152.19 ± 8.97−49.01 ± 6.76124.32 ± 11.20−19.03 ± 0.86−95.91 ± 10.52
XO-VS26−216.89 ± 9.57−40.34 ± 8.46127.33 ± 14.15−20.94 ± 0.91−150.84 ± 9.61
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhai, N.; Wang, C.; Wu, F.; Xiong, L.; Luo, X.; Ju, X.; Liu, G. Exploration of Novel Xanthine Oxidase Inhibitors Based on 1,6-Dihydropyrimidine-5-Carboxylic Acids by an Integrated in Silico Study. Int. J. Mol. Sci. 2021, 22, 8122. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22158122

AMA Style

Zhai N, Wang C, Wu F, Xiong L, Luo X, Ju X, Liu G. Exploration of Novel Xanthine Oxidase Inhibitors Based on 1,6-Dihydropyrimidine-5-Carboxylic Acids by an Integrated in Silico Study. International Journal of Molecular Sciences. 2021; 22(15):8122. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22158122

Chicago/Turabian Style

Zhai, Na, Chenchen Wang, Fengshou Wu, Liwei Xiong, Xiaogang Luo, Xiulian Ju, and Genyan Liu. 2021. "Exploration of Novel Xanthine Oxidase Inhibitors Based on 1,6-Dihydropyrimidine-5-Carboxylic Acids by an Integrated in Silico Study" International Journal of Molecular Sciences 22, no. 15: 8122. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22158122

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop