Next Article in Journal
The Application of Ultrasound in 3D Bio-Printing
Previous Article in Journal
Nanocellulose Derivative/Silica Hybrid Core-Shell Chiral Stationary Phase: Preparation and Enantioseparation Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Silico Exploration of 1,7-Diazacarbazole Analogs as Checkpoint Kinase 1 Inhibitors by Using 3D QSAR, Molecular Docking Study, and Molecular Dynamics Simulations

School of Chemistry and Environmental Engineering, Shanghai Institute of Technology, Shanghai 201418, China
*
Author to whom correspondence should be addressed.
Molecules 2016, 21(5), 591; https://doi.org/10.3390/molecules21050591
Submission received: 30 March 2016 / Revised: 11 April 2016 / Accepted: 28 April 2016 / Published: 5 May 2016
(This article belongs to the Section Medicinal Chemistry)

Abstract

:
Checkpoint kinase 1 (Chk1) is an important serine/threonine kinase with a self-protection function. The combination of Chk1 inhibitors and anti-cancer drugs can enhance the selectivity of tumor therapy. In this work, a set of 1,7-diazacarbazole analogs were identified as potent Chk1 inhibitors through a series of computer-aided drug design processes, including three-dimensional quantitative structure–activity relationship (3D-QSAR) modeling, molecular docking, and molecular dynamics simulations. The optimal QSAR models showed significant cross-validated correlation q2 values (0.531, 0.726), fitted correlation r2 coefficients (higher than 0.90), and standard error of prediction (less than 0.250). These results suggested that the developed models possess good predictive ability. Moreover, molecular docking and molecular dynamics simulations were applied to highlight the important interactions between the ligand and the Chk1 receptor protein. This study shows that hydrogen bonding and electrostatic forces are key interactions that confer bioactivity.

1. Introduction

Cancer refers to uncontrolled abnormal cell division with propensity for tissue invasion and is a highly unique genomic disease [1]. Everyone is born with a unique set of genomes, which determine the risk of cancer. After its onset, the genetic disease quickly causes other genomic mutations that lead to cancer. Treatments for cancer are mainly categorized as chemotherapy, radiotherapy, and surgery. DNA-damaging agents are introduced in chemical treatments for cancer to disturb the chromosome structural integrity or disrupt DNA metabolism, synthesis, and transcription. This method affects both tumor and normal cells; hence, anti-cancer drugs show the lowest therapeutic index among other anti-cancer therapies. However, DNA-damaging agents exert a mutual inhibiting effect on normal cells. The effectiveness of treatment for DNA damage may be improved by inhibiting cell-cycle checkpoint kinases, which can facilitate cell-cycle arrest and provide time for lesion repair.
Checkpoint kinase 1 (Chk1) is an important serine/threonine kinase with a self-protection function [2]. DNA damage can activate the Chk1 protein through the ATR/Chk1 signal transduction pathway. Consequently, the activated Chk1 causes cell-cycle arrest and repairs gene transcription, which ensures the integrity and stability of the genome [3,4]. Cancer cells are accompanied by genetic deficiencies in the P53 gene, leading to the lack of the G1 cell-cycle checkpoint [5]. Therefore, most cancer cells repair themselves through the S and G2–M checkpoint. Chk1 inhibition can disable the function of the S and G2 checkpoint, thereby impeding cancer cell repair and resulting in mitosis disorders, and even cell death or apoptosis. Normal cells possess a relatively complete cycle checkpoint function and are relatively insensitive to Chk1 inhibitors [6]. Currently, clinical studies evaluated several Chk1 inhibitors (Figure 1), such as UCN-01 [7], PF-477736 [8], and AZD7762 [9,10]. These inhibitors, in combination with other anti-cancer drugs, can improve the effect of cancer treatment. For example, AZD7762 can enhance the selectivity of gemcitabine for P53-deficient cells. Thus, the development of Chk1 inhibitors has been a highly active area of research in cancer treatment. Recent works reported a series of potential candidates, especially GNE-783 [11,12,13]. This inhibitor shows a significant half maximal inhibitory concentration (IC50) value of 1.3 nM against the Chk1 protein. Furthermore, GNE-783 demonstrates good oral availability. However, in vivo studies revealed that GNE-783 potently binds to acetylcholine esterase (AChE) and produces muscle fasciculation, thereby interrupting the development of this compound as a therapeutic agent. Gazzard et al. [14] synthesized a novel series of GNE-783 analogs with oral availability to obtain superior bioactive compounds. The influence regularity of AChE bioactivity in AChE binding mode was described. This report discussed that low binding forces in the complex between the AChE protein and its analogs achieve low AChE inhibitor activity. Meanwhile, biological evaluation obtained satisfactory results in the structure modification of GNE-783 analogs. GNE-145 (compound 17, Table 1) shows significant IC50 values of 2.5 nM and 2.42 μM against the Chk1 protein and AChE, respectively. These results indicate that this series of compounds include potent Chk1 inhibitors with low AChE bioactivity.
In silico modeling technology is widely used in drug discovery [15,16,17,18] and chemical field. The design of novel drugs [19] is difficult to achieve without computational chemistry tools because experimentation procedures are expensive and complicated. These computational tools include molecular docking [20], 3D-QSAR, and molecular dynamics simulations, which can be used to understand the relationship between chemical structure and inhibitory activity and develop novel drug candidates. For example, Veselinovića et al. [21] used Monte Carlo QSAR models for predicting the organophosphate inhibition of AChE. Caballero et al. [22] used docking and QSAR models to study the quantitative structure–activity relationships of imidazo[1,2-a]pyrazines derivatives as Chk1 inhibitors. These studies demonstrated the potential feasibility of theoretical calculations. In this work, linear regression analysis methods, such as comparative molecular field analysis (CoMFA) and comparative molecular similarity index analysis CoMSIA), while considering molecular interaction fields were used for in silico identification of 1,7-diazacarbazole analogs as Chk1 inhibitors. The developed models enable detailed examination of molecular structural factors that affect bioactivity. Moreover, these models can predict the bioactivities of new analogs. Molecular docking and dynamics simulations illustrate the possible binding modes of a certain structure and its receptor protein. These binding modes describe that hydrogen bonding and electrostatic forces significantly contribute to bioactivity.

2. Materials and Methods

2.1. Dataset

The dataset used for molecular modeling studies contains 40 compounds which were designed and biological evaluation by Gazzard [14] to explore new 1, 7-diazacarbazole analogs as potent Chk1 inhibitors. The structures of the analogues as well as the pIC50 values (pIC50 = −logIC50) are described in Table 1. The experimental data obtained are randomly divided into a training set (35 structures) for QSAR model generation, and the remaining five molecules constituted the test set for model validation. A previous study [23] enumerated feasible and effective verification methods, and the random test set is an important component for ensuring the accuracy of the method.

2.2. Energy Minimization and Modeling Alignment

All the structures were constructed using the 2D sketcher module in Sybyl-X 2.0 molecular modeling package. Minimum energy calculation of all structures was performed using the Tripos force field [24], followed by 10,000 iterations. The atomic point charges were calculated using the Gasteiger-Hückel [25] method. The root mean square (RMS) of the gradient was set to 0.005 kcal/(mol·Å) [26].
The minimum energy conformation selection and the alignment rule are two crucial factors to build an ideal model. In general, two alignment methods were used to derive the reliable model, including the maximum common substructure (MCS) alignment and the docking-based alignment. In this study, the MCS alignment rule was used to complete the molecular alignment. CoMFA and CoMSIA approaches aligned the structures to compound 28, which is assumed to be the highest bioactive conformation. The common structure (red) was used to position the rest of the compounds and the alignment of the training structures were shown in Figure 2.

2.3. Generation of the QSAR Model

In this study, CoMFA and CoMSIA methods were used to construct 3D-QSAR models. Both CoMFA and CoMSIA methods were based on the field concepts which were around the aligned molecules. The CoMFA model calculated the steric and electrostatic fields [27], and the CoMSIA method calculated five different similarity fields, including steric (S), electrostatic (E), hydrophobic (H), H-bond donor (D), and H-bond acceptor (A) fields [28]. The pIC50 values were used as dependent variables to characterize the molecular structure, and the other parameters were set by default.

2.4. Partial Least Squares (PLS) Analysis and Validation of the QSAR Models

The quantitative relationships between the molecular descriptors and bioactivities were established by partial least squares (PLS) regression analysis [29]. There are many methods used for cross-validation analysis, including leave-one-out (LOO), leave-many-out (LMO), or leave-N-out (LNO) cross-validation, y-randomized validation, and bootstrapping method. The LOO cross-validation method was used in this study to measure the quality of the model. Then, the cross-validation correlation coefficient (q2) value, the correlation coefficient (r2) value, the optimum number of components (ONC) values, the standard error of estimate (SEE), and the F-statistic values were obtained. These statistics were the results of cross-validation, which given enough information about the predictive abilities of the model. The q2 values were usually used to measure the impression of how predictive the model is. The test set was used to evaluate the capacity of external validation. Different statistics and methods can evaluate the predictive power of the model. For example, the external predictive correlation coefficients (r2pred values) [30,31,32] and four criteria proposed by Golbraikh [33]. Then, the statistics’ r2pred values were applied in this study, and the r2pred values are calculated using the following Equation (1):
r pred 2 = 1 PRESS SD
The SD value is the sum of squared deviation between pIC50 values of test set compounds and the mean pIC50 of the training set structure. The PRESS value is the sum of squared deviations between the actual and the predicted bioactivities of the test compounds.

2.5. Molecular Docking Simulations

Surflex-Dock module implemented in Sybyl-X 2.0 was used for the molecular docking studies in this work. The crystal structures of ChK1 kinase domain were downloaded from the Protein Data Bank (PDB ID: 4RVK) [14]. The hydrogen atoms were added, as well as the water molecules, and the ligands had been deleted. The other parameters were default in the software. Subsequently, each structure was docked into the activity pockets for further analysis. An item from the docking results obtained at least 20 ratings, and the highest scoring conformation was studied.

3. Results and Discussions

3.1. Statistical Analysis and Validation

CoMFA and CoMSIA approaches were carried out using the 40 1,7-diazacarbazole derivatives chosen in the Chk1 inhibition experiment. The statistical summaries are shown in Table 2. In the CoMFA model, the PLS regression analysis yielded a q2 of 0.726 with SEE of 0.215. Then, the non-cross-validated method gave an r2 value of 0.918, an F-values of 115.292 and three optimal components.
In the CoMSIA model, a q2 value of 0.531 and r2 value of 0.950 were obtained when five field descriptors were considered. The F-value, SEE value, and ONC values were 141.412, 0.171, and 4, respectively. From the field contribution results, we found that the electrostatic field played a major contribution among the present fields. These data also showed the reliability of the CoMSIA model.
A test set of 40 compounds were used to validate the accuracy and predictability of the model. The values of r2pred, shown in the Table 2, were 0.878 for CoMFA and 0.846 for the CoMSIA model. These statistical indices indicated a good external predictive capacity of the models. The graphs showing the experimental and predicted pIC50 values for the total set used in the CoMFA and CoMSIA approaches are described in Figure 3. The good linear relationships illustrated that the bioactivities predicted by the derived models were in agreement with the experimental data, indicating that these models had satisfactory predictive capacity.

3.2. CoMFA/CoMSIA Contour Map Analysis

The steady 3D QSAR models were generally applied to drug discovery processes to predict the biological activities of unknown derivatives. Moreover, the effects of the field descriptors contributing to activities can be partitioned and visualized through 3D contour plots. The field type of the contour plot was set to StDev × Coeff to aid visualization.
The CoMFA contour maps are illustrated in Figure 4. These contours plots demonstrate regions where the steric and electrostatic variations in the different molecular features lead to either increased or reduced bioactivity. The most structure 28 was chosen as a reference to aid visualization. Two green regions and two yellow places exist around the compound zones represented the steric favorable and unfavorable areas, respectively. The green maps around the methyl group of pyrazole ring indicated that bulk groups were favored there. The methyl group of pyrazole ring at this position may be favorable to the interaction between the compound and its receptor. It can be explained by the fact that the bioactivity of compound 28 (pIC50 = 9.51) was better than those of compound 27 (pIC50 = 8.56). A yellow contour around the R1 substituent indicated that the small bulky group can enhance bioactivities. This phenomenon could be concluded by comparing the activities between compounds 25 (pIC50 = 9.34) and 30 (pIC50 = 8.64).
In the CoMFA electrostatic contour maps (Figure 4), the regions in red implied where electronegative groups improved activities, whereas the positions in blue purported where electropositive groups enhanced activities. A blue contour around the methyl group of pyrazole ring, indicating a positive atomic charge group in this position had a positive effect on the inhibitory activity, such as compounds 28 and 27. Moreover, a large blue tetrahedron around the para-position of the piperidine ring revealed the importance of positive atomic groups. For example, compounds 11, 13, 14, and 15 shown moderate activity probably because the para-position of piperidine ring was replaced by electronegative groups. Simultaneously, two small red cubes near the nitrogen atoms of the pyrazole and piperidine rings indicated that the negatively-charged groups were helpful for inhibitory activity, which was consistent with the experimental results.
The CoMSIA steric and electrostatic contour plots were described in Figure 5. Not surprisingly, most contours were similar to those of CoMFA model and, hence, were not discussed. The hydrophobic contour plot was constructed through CoMSIA model, presented in Figure 6. The yellow and gray colors represented favorable and unfavorable hydrophobic areas, respectively. The hydrophobic contour maps were exclusively located around R1 and R2 substituents. A small yellow cube surrounding the C-1 position of pyrazole ring (R1 group) indicated hydrophobic groups were advantageous for inhibitory activity. Compounds 2529 showed better inhibitory activities than compounds 915, probably because –CN group was replaced by hydrophobic substituents. A yellow cone around the meta-position of benzene ring revealed that hydrophobic substituents in this region produced increased bioactivity. The gray contours around the para-position of the piperidine ring (R2 substituent) suggested that the hydrophilic group in this region could increase activity. It can be demonstrated by the fact that the biological activity of compound 11 (pIC50 = 8.745) was slightly higher than that of compound 13 (pIC50 = 8.301).
The results of the statistical analysis revealed that H-bond acceptor field had a significant contribution on the contour maps. The H-bond acceptor contour plots of the CoMSIA model containing the most active compound 28 were illustrated in Figure 7A. The magenta contour around the N atom of the pyrazole ring indicated that the hydrogen bond acceptor group was conducive to the improvement of bioactivity. The nitrogen atom in this position may be beneficial to the generation of the hydrogen bond between the compound and its receptor protein. It can also be proved by the fact that compounds 2528, which displayed higher activities than other compounds. Thus, the H-bond acceptor group was believed to have a strong bioactivity on Chk1. As shown in Figure 7A, the regions in red implied where hydrogen bond donor group decrease activity. A large red tetrahedron surrounding the para-position of piperidine ring indicated that the hydrogen bond donor group was adverse to inhibitory activity, which can be demonstrated by the fact that compounds 11 and 13 showed a satisfactory activity. The other red contours were away from the most potent compound 28 and, hence, are not discussed. A prominent purple contour around the piperidine ring (R1 substitute) indicated that an H-bond donor was adverse to bioactivity (Figure 7B). For example, the order of many compounds’ bioactivities was: 28 > 24 > 21. Compounds 21 and 24 with hydrogen bond donor groups showed low activities, which was in agreement with the contour map. The cyan color around the N-1 position of piperidine ring (R2 substitute) indicated that the hydrogen bond donor was favorable to activity. This result was the same to that of the hydrophobic contour plot.
According to the conclusions above, six new compounds (2a2g) with satisfactorily predicted pIC50 values have been designed shown in Table 3. These designed compounds exhibited satisfactory predictive values also indicated the correctness of CoMFA and CoMSIA models.

3.3. Molecular Docking

Molecular docking protocols were widely used to investigate the possible binding modes between the target derivatives and the receptor protein, which aided in the understanding the QSAR revealed by CoMFA/CoMSIA models. Prior to docking, a re-docking simulation was applied to validate the accuracy of molecular docking. The target ligand taken from the crystal structure was re-docked into the active site with the root mean square deviation (RMSD) value of 0.225 Å for 4RVK. As shown in the Figure 8, the region of re-docking ligand (blue) was same to that of the original ligand (red). Three hydrogen bonds which were formed between the ligand and the protein appeared on the horizon. The H-bond distances were observed to be 1.69 Å (C=O···H-Cys87), 1.86 Å(C=N···H-Tyr86), 2.18 Å (CN···H-Lys38), respectively. In short, the surflex-dock program could successfully reproduce the original conformation.
Subsequently, the most compound 28 was docked into the ligand-binding pocket of Chk1 protein. As described in Figure 9, the docking results demonstrated three hydrogen bonds between compound 28 and the key residues (including Lys38, Glu85 and Cys87) in the Chk1 active pocket. The nitrogen atom of 1,7-diazacarbazole provided a hydrogen-bonding with Tyr86 (2.09 Å), the N–H group was hydrogen bonded to residues Glu85 (2.06 Å), and the 4-N of pyrazole ring formed a hydrogen-bonding with Lys38 (1.85 Å). These residues could also interact with the target compound through electrostatic interactions. Moreover, the other residues, such as Tyr86, Glu91, Phe149, Asp148, and Glu55, could also stabilize ligand through electrostatic interaction. The CoMFA/CoMSIA electrostatic-favorable blue contour around the pyrazole and the hetero atom further supported the structure-based analysis. In addition, Van der Waals forces were formed between the compound and important residues (for example: Ser147, Val68, Leu84, Asn59, Gly150, Leu82, Gly90, Ala36, Leu15, Leu137, Val23, Asp94, Phe93, Gly16, and Glu17). These van der Walls interaction was involved in the activities of stabilizing compound in the active pocket. The docked model revealed that the hydrogen-bonding/electrostatic interactions played an important role in the interaction between the inhibitor and the protein, and the hydrogen bonds were similar to those in literature [11]. Moreover, the results obtained by the docking had been compared with the QSAR results to verify mutually. These interactions match well with the results of H-bond acceptor/electrostatic contour maps.
Compound 17 as a potential candidate was docked into the same active site to further analyze the impact of residues on the inhibitor activity. From the Figure 10, three hydrogen bonds were in sight. The H-bond distances were observed to be 1.88 Å (C–N∙∙∙H–Cys87), 2.03 Å (N-H···O-Glu85), and 2.18 Å (CN···H-N-Lys38), respectively. These H-bonds were the same to that of compound 28. Furthermore, electrostatic interaction was formed between this compound and the important residues (e.g., Tyr20, Asp148, Lys38, Glu85, Tyr86, Cys87, Thr14, and Leu15). These residues were mainly around R1 substituent. It could also confirm the importance of the electrostatic fields to bioactivity. Of course, the van der Waals force was essential. The amino residues, such as Ser147, Leu37, Val23, Glu91, Gly16, Asp94, Leu84, Val68, Ala36, Gly90, Ser88 and Gln13, could form van der Waals interaction with the target compound. These results also suggested that hydrogen bonding and electrostatic forces were the key interaction that confer bioactivity.

3.4. Molecular Dynamics Simulations

In a subsequent step, the molecular dynamics process was performed on compounds 28 and 17 to further explore the probable binding modes between the compounds and the receptor protein. Complex Chk1 structure of the compound as generated by the previous docking modelling was used as the initial coordinates for molecular dynamics study. The molecular dynamics simulations were performed using the dynamics module of SYBYL-X 2.0 at the vacuum environment [34,35]. Then, the energy minimization was performed for the complex molecule with Gasteiger-Hückel charge and Tripos force field without water using Boltzmann initial velocity. The simulations were executed using normal temperature and volume (NTV) [36] ensemble 300 K with coupling 100 fs. Additionally, we perform a 5 ns simulation with a time step of 1 fs and snapshot the conformation every 1000 fs.
A 5 ns simulation of the complex between the compound 28 and the protein was run to energy balance at 2 ns to obtain the stable conformation. The total energy of compound ranging from 4057 to 3620 KJ/mol was illustrated in Figure 11A. After 2 ns, the total energy of the complex dropped to 3620 KJ/mol, and tended to stability. This result suggested that the ligand-protein complex could reach the metastable conformation after 2 ns of simulation. The alignment of original and molecular dynamics simulated ligand were shown in Figure 12A to indicate a high similarity among these two ligand. From the Figure 12A, some residues (such as: Tyr86, Ser147, Gly90, Asp94, Asp148, Glu17, Leu15, Leu137, Leu84, and Gly16) were still important to the interaction between compound 28 and Chk1 protein. However, the number of hydrogen bonds and the amino acid residues decreased. One hydrogen bond was formed between Gly89 and compound 28, with the distance of 1.67 Å. This reside formed not only hydrogen-bonding to ligand, but also formed the electrostatic force. Moreover, most residues (such as: Asp148, Glu17, Leu15, et al.) were bound to the target compound through the electrostatic interaction. Therefore, we predict that the electrostatic interaction has a greater impact on activity of compound 28. These results match well with the field contribution of CoMSIA model.
The docked complex of compound 17 and the Chk1 protein was studied subsequently through a 5 ns simulation. The total energy of the docked complex was shown in Figure 11B. From Figure 11B, the total energy tends to be stable after 1 ns. The alignment of molecular dynamics simulated and the original ligand was described in Figure 12B. Meanwhile, the key residues which interacted with the target compound were in sight. Some key residues (Glu85, Cys87, and Lys38) still interacted with compounds through H-bonding and electrostatic force. The H-bond distances were 1.87 Å, 1.95 Å, and 1.86 Å, respectively. Combining the docking results, the key residues (Glu85, Cys87, and Lys38) were predicted to be the important factor influencing the bioactivity of Chk1.

4. Conclusions

In this work, several 1,7-diazacarbazole analogs were identified as potentially effective oral Chk1 inhibitors through a series of computer-aided drug design processes, such as 3D-QSAR modeling, molecular docking, and molecular dynamics simulations. The CoMFA/CoMSIA models with statistical capacity showed good internal and external validation abilities and can be used to predict new and potential molecules. Moreover, the obtained contour maps can be used to guide the design of new compounds with high Chk1 inhibitory activity. Meanwhile, molecular docking and molecular dynamics process were established to study the possible binding modes of inhibitors at the active pocket of Chk1. Some key residues, such as Glu85, Cys87, Lys38, and Gly90, were found. Hydrogen bonding and electrostatic forces were predicted to be the key interactions that confer bioactivity. Overall, these results show that the optimal CoMFA/CoMSIA models can be used to predict novel Chk1 inhibitors and guide the development of new potential oral analogs.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/21/5/591/s1.

Acknowledgments

This work was supported by the “Chemical Engineering and Technology (Perfume and Aroma Technology) plateau discipline of Shanghai Municipality” and “industry-university research institutions” collaborative innovation fund (XYCX2015-14) of Shanghai Institute of Technology.

Author Contributions

X.G.: Responsible for all of the computational experimental work and data analyses as well as was responsible for the entire manuscript write up from the cover page to the references; L.H.: was only responsible for the 3D-QSAR computational aspect which included analysis of results and write up; Y.J.R.: principal investigator and corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dan, L.L. Tumor heterogeneity and personalized medicine. N. Engl. J. Med. 2012, 366, 956–957. [Google Scholar]
  2. Bucher, N.; Britten, C.D. G2 checkpoint abrogation and checkpoint kinase-1 targeting in the treatment of cancer. Br. J. Cancer 2008, 98, 523–528. [Google Scholar] [CrossRef] [PubMed]
  3. Musk, S.R.; Steel, G.G. Override of the Radiation-induced Mitotic Block in Human Tumour Cells by Methylxanthines and Its Relationship to the Potentiation of Cytotoxicity. Int. J. Radiat. Biol. 1990, 57, 1105–1112. [Google Scholar] [CrossRef] [PubMed]
  4. Yan, L.; Leverson, J.D. New opportunities in chemosensitization and radiosensitization: Modulating the DNA-damage response. Expert. Rev. Anticancer 2005, 5, 333–342. [Google Scholar]
  5. Anderson, D.F.; Chapman, S.; Inman, F.; Smith, W.W. Targeting Chk1 in p53-deficient triple-negative breast cancer is therapeutically beneficial in human-in-mouse tumor models. J. Clin. Investig. 2012, 122, 1541–1552. [Google Scholar]
  6. Wang, J.; Han, X.; Zhang, Y. Autoregulatory mechanisms of phosphorylation of checkpoint kinase 1. Cancer Res. 2012, 72, 3786–3794. [Google Scholar] [CrossRef] [PubMed]
  7. Li, T.; Christensen, S.D.; Frankel, P.H.; Margolin, K.A.; Agarwala, S.S.; Luu, T.; Mack, P.C.; Lara, P.N., Jr.; Gandara, D.R. A phase II study of cell cycle inhibitor UCN-01 in patients with metastatic melanoma: A California Cancer Consortium trial. Investig. New Drugs 2010, 30, 741–748. [Google Scholar] [CrossRef] [PubMed]
  8. Blasina, A.; Hallin, J.; Chen, E.; Arango, M.E.; Kraynov, E.; Register, J.; Grant, S.; Ninkovic, S.; Chen, P.; Nichols, T.; et al. Breaching the DNA damage checkpoint via PF-00477736, a novel small-molecule inhibitor of checkpoint kinase 1. Mol. Cancer Ther. 2008, 7, 2394–2404. [Google Scholar] [CrossRef] [PubMed]
  9. Ma, Z.; Yao, G.; Zhou, B. The Chk1 inhibitor AZD7762 sensitises p53 mutant breast cancer cells to radiation in vitro and in vivo. Mol. Med. Rep. 2012, 6, 897–903. [Google Scholar] [PubMed]
  10. Mitchell, J.B.; Choudhuri, R.; Fabre, K. In vitro and in vivo radiation sensitization of human tumor cells by a novel checkpoint kinase inhibitor, AZD7762. Clin. Cancer Res. 2010, 16, 2076–2084. [Google Scholar] [CrossRef] [PubMed]
  11. Gazzard, L.; Appleton, B.; Chapman, K. Discovery of the 1, 7-diazacarbazole class of inhibitors of checkpoint kinase 1. Bioorg. Med. Chem. Lett. 2014, 24, 5704–5709. [Google Scholar] [CrossRef] [PubMed]
  12. Nagro, C.J.D.; Jonathan, C.; Yang, X.; Linda, R.; Sankar, M.; Ajay, P.; Zha, J.P.; Jackson, P.K.; O’Brien, T. Chk1 inhibition in p53-deficient cell lines drives rapid chromosome fragmentation followed by caspase-independent cell death. Cell Cycle 2014, 13, 303–314. [Google Scholar]
  13. Xiao, Y.; Ramiscal, J.; Kowanetz, K.; Del Nagro, C.; Malek, S.; Evangelista, M.; Blackwood, E.; Jackson, P.K.; O’Brien, T. Identification of Preferred Chemotherapeutics for Combining with a CHK1 Inhibitor. Mol. Cancer Ther. 2013, 12, 2285–2295. [Google Scholar] [CrossRef] [PubMed]
  14. Gazzard, L.; Williams, K.; Chen, H. Mitigation of Acetylcholine Esterase Activity in the 1,7-Diazacarbazole Series of Inhibitors of Checkpoint Kinase 1. J. Med. Chem. 2015, 58, 5053–5074. [Google Scholar] [CrossRef] [PubMed]
  15. Zhong, Z.; Liu, L.J.; Dong, Z.Q.; Lu, L.; Wang, M.; Leung, C.H.; Ma, D.L.; Wang, Y. Structure-based discovery of an immunomodulatory inhibitor of TLR1-TLR2 heterodimerization from a natural product-like database. Chem. Commun. 2015, 51, 11178–11181. [Google Scholar] [CrossRef] [PubMed]
  16. Ma, D.L.; Chan, S.H.; Wei, G.; Zhong, H.J.; Yang, H.; Leung, L.T.; Gullen, E.A.; Chiu, P.; Cheng, Y.C.; Leung, C.H. Virtual screening and optimization of Type II inhibitors of JAK2 from a natural product library. Chem. Commun. 2014, 50, 13885–13888. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, L.J.; Leung, K.H.; Chan, S.H.; Wang, Y.T.; Ma, D.L.; Leung, C.H. Identification of a natural product-like STAT3 dimerization inhibitor by structure-based virtual screening. Cell Death Dis. 2015, 5, 96–104. [Google Scholar] [CrossRef] [PubMed]
  18. Ma, D.L.; Chan, S.H.; Leung, C.H. Drug repositioning by structure-based virtual screening. Chem. Soc. Rev. 2013, 42, 2130–2141. [Google Scholar] [CrossRef] [PubMed]
  19. Yang, H.; Zhong, H.J.; Leung, K.H.; Chan, S.H.; Ma, P.Y.; Fu, W.C.; Nanjunda, R.; Wilson, W.D.; Ma, D.L.; Leung, C.H. Structure-based design of flavone derivatives as c- myc oncogene down-regulators. Eur. J. Pharm. Sci. 2013, 48, 130–141. [Google Scholar] [CrossRef] [PubMed]
  20. Ma, D.L.; Chan, S.H.; Leung, C.H. Molecular docking for virtual screening of natural product databases. Chem. Sci. 2011, 2, 1656–1665. [Google Scholar] [CrossRef]
  21. Veselinović, J.B.; Nikolić, G.M.; Trutić, N.V. Monte Carlo QSAR models for predicting organophosphate inhibition of acetylcholinesterase. SAR QSAR Environ. Res. 2015, 26, 449–460. [Google Scholar] [CrossRef] [PubMed]
  22. Caballero, J.; Zilocchi, S.; Tiznado, W. Docking and quantitative structure–activity relationship studies for imidazo[1,2-a]pyrazines as inhibitors of checkpoint kinase-1. Med. Chem. Res. 2012, 21, 1912–1920. [Google Scholar] [CrossRef]
  23. Cruzz, V.L.; Martinez, S.; Ramos, J. 3D-QSAR as a Tool for Understanding and Improving Single-Site Polymerization Catalysts. A Review. Organometallics 2014, 33, 2944–2959. [Google Scholar] [CrossRef]
  24. Gasteiger, J.; Marsili, M. Iterative partial equalization of orbital electronegativity: A rapid access to atomic charges. J. Tetrahedron. 1980, 36, 3219–3228. [Google Scholar] [CrossRef]
  25. Chebib, M.; Johnston, G.A.R. GABA-activated ligand gated ion channels: medicinal chemistry and molecular biology. J. Med. Chem. 2000, 43, 1427–1447. [Google Scholar] [CrossRef] [PubMed]
  26. Head, J.D.; Zerner, M.C. A Broyden-Fletcher-Goldfarb-Shanno optimization procedure for molecular geometries. J. Chem. Phys. Lett. 1985, 3, 264–270. [Google Scholar] [CrossRef]
  27. 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] [PubMed]
  28. Klebe, G.; Abraham, U. Comparative molecular similarity index analysis (CoMSIA) to study hydrogen-bonding properties and to score combinatorial libraries. Comput. Aid. Mol. Des. 1999, 13, 1–10. [Google Scholar] [CrossRef]
  29. Wold, S.; Ruhe, A.; Wold, H. The collinearity problem in linear regression. The partial least squares (PLS) approach to generalized inverses. SIAM J. Sci. Comput. 1984, 5, 735–743. [Google Scholar] [CrossRef]
  30. Aparoy, P.; Suresh, G.K.; Reddy, K.K. CoMFA and CoMSIA studies on 5-hydroxyindole-3-carboxylate derivatives as 5-lipoxygenase inhibitors: Generation of homology model and docking studies. Bioorg. Med. Chem. Lett. 2011, 21, 456–462. [Google Scholar] [CrossRef] [PubMed]
  31. Khd, S.; Naveena, Q.; Karthikeyan, M. Jak2 inhibitor-a jackpot for pharmaceutical industries: A comprehensive computational method in the discovery of new potent Jak2 inhibitors. Mol. Biosyst. 2014, 10, 2146–2159. [Google Scholar]
  32. Yu, Z.; Li, X.-C.; Ge, C.-Z.; Si, H.-Z.; Cui, L.-H.; Gao, H.; Duan, Y.-B.; Zhai, H.-B. 3D-QSAR modeling and molecular docking study on Mer kinase inhibitors of pyridine-substituted pyrimidines. Mol. Divers. 2015, 19, 135–147. [Google Scholar] [CrossRef] [PubMed]
  33. Golbraikh, A.; Tropsha, A. Beware of Q2. J. Mol. Graph. Model. 2002, 20, 269–276. [Google Scholar] [CrossRef]
  34. Wang, Z.; Cheng, L.; Kai, Z.; Wu, F.; Liu, Z.; Cai, M. Molecular modeling studies of atorvastatin analogues as HMGR inhibitors using 3D-QSAR, molecular docking and molecular dynamics simulations. Bioorg. Med. Chem. Lett. 2014, 24, 3869–3876. [Google Scholar] [CrossRef] [PubMed]
  35. Xu, C.; Ren, Y. Molecular modeling studies of [6,6,5] Tricyclic Fused Oxazolidinones as FXa inhibitors using 3D-QSAR, Topomer CoMFA, molecular docking and molecular dynamics simulations. Bioorg. Med. Chem. Lett. 2015, 25, 4522–4528. [Google Scholar] [CrossRef] [PubMed]
  36. Cheng, L.P.; Huang, X.Y.; Wang, Z.; Kai, Z.P.; Wu, F.H. Combined 3D-QSAR, molecular docking, and molecular dynamics study on potent cyclohexene-based influenza neuraminidase inhibitors. Monatshefte Chem. 2014, 145, 9–22. [Google Scholar] [CrossRef]
  • Sample Availability: Not Available.
Figure 1. The protein Chk1 inhibitors.
Figure 1. The protein Chk1 inhibitors.
Molecules 21 00591 g001
Figure 2. Common substructure (red) used in alignment, and the alignment of training structures.
Figure 2. Common substructure (red) used in alignment, and the alignment of training structures.
Molecules 21 00591 g002
Figure 3. Plots of experimental vs. predicted pIC50 values for the total set in the CoMFA (A) and CoMSIA (B) models.
Figure 3. Plots of experimental vs. predicted pIC50 values for the total set in the CoMFA (A) and CoMSIA (B) models.
Molecules 21 00591 g003
Figure 4. Steric (A) and electrostatic (B) contours of the CoMFA model. The green color shows the favored steric area and the yellow color show steric area. The red color shows the favored negative electrostatic area and the blue color shows the favored positive electrostatic area.
Figure 4. Steric (A) and electrostatic (B) contours of the CoMFA model. The green color shows the favored steric area and the yellow color show steric area. The red color shows the favored negative electrostatic area and the blue color shows the favored positive electrostatic area.
Molecules 21 00591 g004
Figure 5. Steric (A) and electrostatic (B) contours of the CoMSIA model.
Figure 5. Steric (A) and electrostatic (B) contours of the CoMSIA model.
Molecules 21 00591 g005
Figure 6. Hydrophobic contours of the CoMSIA model. The yellow color shows the favored hydrophobic area, the white color shows the disfavored hydrophobic area.
Figure 6. Hydrophobic contours of the CoMSIA model. The yellow color shows the favored hydrophobic area, the white color shows the disfavored hydrophobic area.
Molecules 21 00591 g006
Figure 7. Hydrogen bonding contour (include H-bond acceptor (A) and H-bond donor (B) contour plots) of CoMSIA model. The magenta color shows the favored H-acceptor area, the red color shows the disfavored H-acceptor area, the cyan color shows the favored H-donor area, and the purple color represents the disfavored H-donor area.
Figure 7. Hydrogen bonding contour (include H-bond acceptor (A) and H-bond donor (B) contour plots) of CoMSIA model. The magenta color shows the favored H-acceptor area, the red color shows the disfavored H-acceptor area, the cyan color shows the favored H-donor area, and the purple color represents the disfavored H-donor area.
Molecules 21 00591 g007
Figure 8. Re-docking results of the compound into the binding site of Chk1 protein and the crystal structure (blue ribbon) for the compound complex. The re-docked ligand (blue) and the original ligand (red) shown as a stick model. The hydrogen-bonds are shown as yellow lines, with distance unit of Å.
Figure 8. Re-docking results of the compound into the binding site of Chk1 protein and the crystal structure (blue ribbon) for the compound complex. The re-docked ligand (blue) and the original ligand (red) shown as a stick model. The hydrogen-bonds are shown as yellow lines, with distance unit of Å.
Molecules 21 00591 g008
Figure 9. Docking result of compound 28 in the activity site of Chk1 protein. Ligand (yellow) and the key residues colored white for binding interaction were shown as stick model. The hydrogen-bonds were described as yellow lines, with distance unit of Å.
Figure 9. Docking result of compound 28 in the activity site of Chk1 protein. Ligand (yellow) and the key residues colored white for binding interaction were shown as stick model. The hydrogen-bonds were described as yellow lines, with distance unit of Å.
Molecules 21 00591 g009
Figure 10. Docking result of compound 17 in the activity site of Chk1 protein. Ligands (yellow) and the key residues (white) for binding interactions were shown as a stick model. The hydrogen bonds were described as yellow lines, with distance measured in Å.
Figure 10. Docking result of compound 17 in the activity site of Chk1 protein. Ligands (yellow) and the key residues (white) for binding interactions were shown as a stick model. The hydrogen bonds were described as yellow lines, with distance measured in Å.
Molecules 21 00591 g010
Figure 11. Plot of the total-energy docked complex (compound 28 (A) and compound 17 (B)) and the molecular dynamics simulation time (ps) in the molecular dynamics simulated structure.
Figure 11. Plot of the total-energy docked complex (compound 28 (A) and compound 17 (B)) and the molecular dynamics simulation time (ps) in the molecular dynamics simulated structure.
Molecules 21 00591 g011
Figure 12. The alignment of molecular dynamics simulated and original ligand, as well as the docking results of compound 28 (A) and 17 (B).
Figure 12. The alignment of molecular dynamics simulated and original ligand, as well as the docking results of compound 28 (A) and 17 (B).
Molecules 21 00591 g012
Table 1. Chemical structural formulas of all structures. Statistical parameters of the actual and predicted bioactivity by CoMFA and CoMSIA, as well as the residual between the actual and predicted pIC50 values. All the aligned molecular dataset used for the 3D QSAR studies were shown in Table S1 in the supplementary materials.
Molecules 21 00591 i001
Table 1. Chemical structural formulas of all structures. Statistical parameters of the actual and predicted bioactivity by CoMFA and CoMSIA, as well as the residual between the actual and predicted pIC50 values. All the aligned molecular dataset used for the 3D QSAR studies were shown in Table S1 in the supplementary materials.
Molecules 21 00591 i001
NO.R1R2Actual pIC50CoMFACoMSIA
pIC50ResidualpIC50Residual
1 Molecules 21 00591 i003 Molecules 21 00591 i0167.5677.706−0.1397.614−0.047
2 Molecules 21 00591 i003 Molecules 21 00591 i0177.6957.4370.2587.2660.429
3 Molecules 21 00591 i003 Molecules 21 00591 i0187.4717.726−0.2557.509−0.038
4 Molecules 21 00591 i003 Molecules 21 00591 i0197.8337.7050.1287.7320.101
5 Molecules 21 00591 i003 Molecules 21 00591 i0206.8967.160−0.2647.046-0.150
6 Molecules 21 00591 i003 Molecules 21 00591 i0457.1167.219-0.1037.280−0.164
7 Molecules 21 00591 i003 Molecules 21 00591 i0467.3707.3230.0477.2920.078
8 Molecules 21 00591 i003H6.3836.600−0.2176.532−0.149
9 Molecules 21 00591 i003 Molecules 21 00591 i0218.0278.330−0.3038.158−0.131
10 Molecules 21 00591 i003 Molecules 21 00591 i0227.9918.072−0.0818.055−0.064
11 Molecules 21 00591 i003 Molecules 21 00591 i0238.7458.3100.4358.799−0.054
12 Molecules 21 00591 i003 Molecules 21 00591 i0248.7218.4970.2248.769−0.048
13 Molecules 21 00591 i003 Molecules 21 00591 i0258.3018.1210.1808.1590.142
14 Molecules 21 00591 i003 Molecules 21 00591 i0267.9368.014−0.0788.012−0.076
15 Molecules 21 00591 i003 Molecules 21 00591 i0277.8578.136−0.2797.962−0.105
16 Molecules 21 00591 i003Br6.8546.7480.1066.7620.092
17 Molecules 21 00591 i003 Molecules 21 00591 i0288.6028.4140.1888.729−0.127
18 Molecules 21 00591 i003 Molecules 21 00591 i0298.2768.1100.1668.2440.032
19 Molecules 21 00591 i003 Molecules 21 00591 i0307.8667.8430.0238.019−0.153
20 Molecules 21 00591 i003 Molecules 21 00591 i0318.6388.3900.2488.5020.136
21H Molecules 21 00591 i0327.7307.5000.2307.7030.027
22 Molecules 21 00591 i004 Molecules 21 00591 i0328.3678.2830.0848.537−0.170
23 Molecules 21 00591 i005 Molecules 21 00591 i0327.7037.802−0.0997.847−0.144
24 Molecules 21 00591 i006 Molecules 21 00591 i0327.7387.818−0.0807.5830.155
25 Molecules 21 00591 i007 Molecules 21 00591 i0329.3379.2410.0969.1160.221
26 Molecules 21 00591 i008 Molecules 21 00591 i0329.0229.189−0.1679.295−0.273
27 Molecules 21 00591 i009 Molecules 21 00591 i0328.9598.8830.0768.8000.159
28 Molecules 21 00591 i010 Molecules 21 00591 i0329.5099.585−0.0769.4710.038
29 Molecules 21 00591 i011 Molecules 21 00591 i0329.0228.7060.3168.6150.407
30 Molecules 21 00591 i012 Molecules 21 00591 i0328.6388.974−0.3368.4520.186
31 Molecules 21 00591 i013 Molecules 21 00591 i0328.5388.565−0.0278.672−0.134
32 Molecules 21 00591 i014 Molecules 21 00591 i0328.2928.713−0.4218.468−0.176
33 Molecules 21 00591 i003 Molecules 21 00591 i0337.3807.3110.0697.3740.006
34 Molecules 21 00591 i003 Molecules 21 00591 i0347.8707.963−0.0937.882−0.012
35 Molecules 21 00591 i003 Molecules 21 00591 i0357.9437.8020.1417.9370.006
Test1 Molecules 21 00591 i015 Molecules 21 00591 i0328.8869.278−0.3929.180−0.294
Test2 Molecules 21 00591 i003 Molecules 21 00591 i0368.6788.4250.2538.3930.285
Test3 Molecules 21 00591 i003 Molecules 21 00591 i0378.3778.3040.0738.2660.111
Test4 Molecules 21 00591 i003 Molecules 21 00591 i0388.3578.364−0.0078.2510.106
Test5 Molecules 21 00591 i003NHEt7.2477.1510.0966.9270.320
Table 2. The PLS statistical results of CoMFA and CoMSIA model.
Table 2. The PLS statistical results of CoMFA and CoMSIA model.
PLS StatisticsCoMFACoMSIA
q2 a0.7260.531
NOC b34
r2 c0.9180.950
SEE d0.2150.171
F e115.292141.412
r2pred f0.8780.846
Steric 0.5090.199
Electrostatic0.4910.283
H-acceptor-0.238
H-donor-0.099
Hydrophobic-0.182
a Leave-one-out cross-validated correlation coefficient; b the optimum number of components; c Non-cross-validated correlation coefficient; d Standard error estimate of non-cross-validated correlation coefficient; e F-test value; f the external predictive correlation coefficients.
Table 3. Designed molecules and predicted inhibit activities values of Chk 1 through CoMFA and CoMSIA.
Molecules 21 00591 i002
Table 3. Designed molecules and predicted inhibit activities values of Chk 1 through CoMFA and CoMSIA.
Molecules 21 00591 i002
No.R1Predicted pIC50 by CoMFAPredicted pIC50 by CoMSIA
2a Molecules 21 00591 i0399.5709.682
2b Molecules 21 00591 i0409.6549.504
2c Molecules 21 00591 i0419.4129.483
2d Molecules 21 00591 i0429.4119.400
2f Molecules 21 00591 i0439.6199.297
2g Molecules 21 00591 i0449.4249.201

Share and Cite

MDPI and ACS Style

Gao, X.; Han, L.; Ren, Y. In Silico Exploration of 1,7-Diazacarbazole Analogs as Checkpoint Kinase 1 Inhibitors by Using 3D QSAR, Molecular Docking Study, and Molecular Dynamics Simulations. Molecules 2016, 21, 591. https://doi.org/10.3390/molecules21050591

AMA Style

Gao X, Han L, Ren Y. In Silico Exploration of 1,7-Diazacarbazole Analogs as Checkpoint Kinase 1 Inhibitors by Using 3D QSAR, Molecular Docking Study, and Molecular Dynamics Simulations. Molecules. 2016; 21(5):591. https://doi.org/10.3390/molecules21050591

Chicago/Turabian Style

Gao, Xiaodong, Liping Han, and Yujie Ren. 2016. "In Silico Exploration of 1,7-Diazacarbazole Analogs as Checkpoint Kinase 1 Inhibitors by Using 3D QSAR, Molecular Docking Study, and Molecular Dynamics Simulations" Molecules 21, no. 5: 591. https://doi.org/10.3390/molecules21050591

Article Metrics

Back to TopTop