[BACK]
BIOCELL
DOI:10.32604/biocell.2021.017019
images
Article

Repurposing of FDA-Approved drugs to predict new inhibitors against key regulatory genes in Mycobacterium tuberculosis

XINJUN YANG1, AFTAB ALAM2, NAIYAR IQBAL3 and KHALID RAZA4,*

1Department of Pediatrics, 3201 Hospital, Hanzhong, China
2Centre for Interdisciplinary Research in Basic Sciences, Jamia Millia Islamia, New Delhi, India
3Department of Computer Science and Information Technology, Maulana Azad National Urdu University, Hyderabad, India
4Department of Computer Science, Jamia Millia Islamia, New Delhi, India
*Address correspondence to: Khalid Raza, kraza@jmi.ac.in
Received: 21 April 2021; Accepted: 25 May 2021

Abstract: Tuberculosis (TB) disease has become one of the major public health concerns globally, especially in developing countries. Numerous research studies have already been carried out for TB, but we are still struggling for a complete and quick cure for it. The progress of Mycobacterium tuberculosis (MTB) strains resistant to existing drugs makes its cure and control very complicated. Therefore, it is the need of the hour to search for newer and effective drugs that can inhibit an increasing number of putative drug targets. We applied the drug repurposing concept to identify promising FDA-approved drugs against five key-regulatory genes (FurB, IdeR, KstR, MosR, and RegX3) of the MTB. The FDA drugs were virtually screened using a structure-based approach by GOLD versions 5.2, and subjected to rigid docking followed by an induced-fit docking algorithm to enhance the accuracy and prioritize drugs for repurposing. We found 11 candidate drugs (including ZINC03871613, ZINC03871614, ZINC03871615 as top scorer candidate drugs) that were frequently present within the top 20 GoldScore ranks and showed promising results. Furthermore, molecular dynamics simulation was performed to monitor the effect of the top scorer drugs on the structural stability of all the five targets, indicating that inhibitors preferentially bind to the active site of the targets. This work suggests that these known FDA-approved drugs open new application domains in the form of anti-tuberculosis agents.

Keywords: Virtual screening; Molecular docking; MD simulation; PPI network

Introduction

Tuberculosis (TB) is one of the oldest diseases with molecular evidence that dates back to17,000 years ago. Despite advances in modalities for diagnosis and treatment of TB, unfortunately, people are still suffering, and it is among the top 10 killer communicable diseases worldwide, second only to HIV. TB is a communicable disease caused by Mycobacterium tuberculosis. It typically affects the lungs (pulmonary TB) but can affect other sites as well (extra-pulmonary TB) (Ahmad, 2011). The disease is spread in the air when people who are sick with pulmonary diseases expel bacteria during coughing. According to WHO, six countries hold 60% of all the TB burden of the entire world in which India is the leading country followed by Indonesia, China, Nigeria, Pakistan, and South Africa (shown in Fig. 1). The Central TB Division, Ministry of Health and Family Welfare, India has published a report on “India TB report 2018” which reported TB statistics for India for 2016 to give an estimated incidence figure of 2.79 million cases of TB for India and up to 4.23 lakh patients were estimated to have died during the year (India TB Report, 2018). Multidrug-resistant (MDR) tuberculosis (TB) is a serious global public health problem (Dua et al., 2018), and estimates indicate that unless the management of MDR TB changes radically, it will be one of the main drivers of antimicrobial resistance, which could kill more persons than cancer by 2050 (Ho et al., 2016). Major challenges to curb the TB in India include poor primary healthcare system in rural areas, unregulated private health care leading to the indiscriminate use of first-line and second-line anti-TB drugs, spreading HIV infection, poverty, and lack of administrative coordination among government bodies responsible for health and hygiene (Singh et al., 2017). Hence, there is an urgent need to design or developed nontoxic biocompatible drugs against TB (Peters et al., 2020; Weiner et al., 2020).

images

Figure 1: Six leading countries account for 60% of the total, with India leading the count, followed by Indonesia, China, Nigeria, Pakistan, and South Africa (WHO).

In our study, we have used the drug repurposing concept, which is a method for disease cure that accelerates the identification of new uses for existing drugs with minimum side effects (Karaman and Sippl, 2019; Lanza et al., 2018; Savoia, 2016). Previous research has shown that the drug repurposing concept (a combination of virtual screening with docking and dynamics studies) can be successfully applied to treat new diseases with existing drugs. For example, the drug sildenafil, which was initially designed for the treatment of hypertension and ischemic heart disease, was later approved for the treatment of erectile dysfunction, representing a successful history of drug repurposing (Afzal et al., 2014; Karthick and Ramanathan, 2013; Li et al., 2001; Palos et al., 2017). Here, we use a computer-guided drug repurposing method and apply it to virtually screened FDA approved drugs against five key regulatory proteins from MTB: (a) FurB–as shown by the regulatory profiles, furB gene is up-regulated in macrophages-phagocytosed bacteria, fascinating that it might be involved in the regulation of intracellular gene expressions. The furB gene is co-transcribed with its upstream gene (Rv2358), which encodes another zinc-dependent regulator and responsible for repressing at least 32 genes, several of which have been implicated with zinc homeostasis. In addition, FurB controls five genes encoding ribosomal proteins three of which containing a putative zinc-binding motif. Based on comparative genomics data these ribosomal proteins have also been suggested to be involved in zinc homeostasis (Lucarelli et al., 2007; Panina et al., 2003). (b) IdeR–It is an iron-dependent regulator that maintains a moderate level of iron inside a bacterium by controlling the transcription of genes, including the expression of its own gene involved in iron uptake, transport, and storage (Banerjee et al., 2011). In M. tuberculosis regulation of the intracellular levels of iron are performed by the transcription factor IdeR.In the iron sufficient stage in MTB, iron binds and activates intracellular IdeR, which then represses the iron acquisition machinery of the pathogen and activates the synthesis of iron storage proteins. IdeR also plays a key role in protecting the cells against the oxidative stress generated by the host (Rohilla et al., 2017). (c) KstR–It is a highly conserved TetR family transcriptional repressor (TFR) that regulates 78 genes responsible for cholesterol catabolism in MTB (Kendall et al., 2010; Sanz et al., 2011). In MTB, a highly critical enzyme known as 3-β HSO(EC 1.1.1.145) has been primarily involved in the degradation of cholesterol, which transforms cholesterol into cholestenone, In fact, this enzyme is expressed in a cholesterol-independent manner suggesting its availability in the pathogen at any time (García et al., 2012). (d) MosR–This is a highly connected protein in the transcriptional regulatory (TR) network of MTB, one of its functions is to regulate around 295 genes at the level of transcription. The main role of MosR is to up-regulate expression of rv1050 (a putative exported oxidoreductase that has not yet been characterized) in response to oxidants and propose that it is through this role that MosR contributes to the bacterium survival in the macrophage (Brugarolas et al., 2012). (e) RegX3–It is a self-regulatory protein and controls the expression of its own gene. Many RegX3-dependent genes have a role in the biosynthesis of DNA, RNA and other macromolecules (Pei et al., 2021). Whereas, a number of genes participate in the catabolism of fatty acids and are repressed by fadE6, fadE14 and accD2 genes in a direct way (Ahmad and Kumar, 2016; Carroll et al., 2011). This two component system is also responsible for cell wall biosynthesis, lipid biosynthesis and cell envelope protein synthesis (Almatar et al., 2017; Roberts et al., 2011).

Materials and Methods

Preparation of drug target and drug-library

In our study, we considered five key regulatory genes from MTB; FurB(PDB ID: 2O03) (Lucarelli et al., 2007), IdeR (PDB ID: 1B1B) (Pohl et al., 1999), KstR (PDB ID: 3MNL) (Ho et al., 2016), MosR (PDB ID: 4FX0) (Brugarolas et al., 2012) and RegX3 (PDB ID: 2OQR) (King-Scott et al., 2007), details are given in Fig. 2. All the atomic coordinates (x-ray structure with good resolution, from 1.8Å to 2.7Å) of these proteins were downloaded from Protein Data Bank (Berman et al., 2000). The proteins were optimized for docking by following standard procedures, which include removal of crystallographic water molecules, the addition of missing H-atoms, and minimizing the energy level of all the proteins using the Swiss-PDB viewer tool and this server provides an energy minimization facility using the GROMOS96 force field (Alam et al., 2016; Johansson et al., 2012). All the targets were further refined using SCWRL 4.0 tool to minimize side-chain to backbone clashes and the side-chain-to-side chain clashes (Wang et al., 2008). The drug library consists of 1985 drugs (FDA approved) which were downloaded from the ZINC-15database (Irwin et al., 2012). CORINA Classic (the classic command-line version of CORINA) (https://www.mn-am.com/products/corina) was used to prepare the drug-library because it facilitates preparation of high quality, low energy, single, 3D structure with correct chirality.

images

Figure 2: Key regulatory genes from MTB. FurB (PDB ID: 2O03), IdeR (PDB ID: 1B1B), KstR (PDB ID: 3MNL), MosR (PDB ID: 4FX0) and RegX3 (PDB ID: 2OQR).

Virtual screening

Virtual screening (VS) is a computational method used in drug designing and discovery to find out the series of (compound library) small molecules in order to identify the most potent molecules which are most likely to bind to a drug target (Krueger et al., 2016; Morrone Xavier et al., 2016; Rester, 2008; Rollinger et al., 2008). In our study, VS was performed by using GOLD versions 5.2 (Genetic Optimisation for Ligand Docking; Cambridge Crystallographic Data Centre) (Jones et al., 1995; Jones et al., 1997). GOLD uses genetic algorithms to dock ligand with the receptor. The scoring function “GoldScore” was tested individually for each protein. The molecular docking that utilizes the genetic algorithm (GA) was set to three different execution runs–(a) 50, (b) 100, and (c) 200 GA runs, so as to identify influential and high-affinity binding modes. The radius of the docking sphere was 10–12 Å centered at the position, a maximum of 100,000 operations were performed that deliver high prediction accuracy but are relatively slow (Heberlé and De Azevedo, 2011; Potemkin and Grishina, 2018). The information on the binding site for all the targets was observed from the available literature which was then used further for grid generation. For FurB and IdeR, we used its active repressor site (Lucarelli et al., 2007) for KstR and MosR, we used its DNA binding site (Brugarolas et al., 2012; Ho et al., 2016) and for RegX3, we used its conformational switch region (present ‘hotspot’ residues) (Ahmad et al., 2015; Khalid et al., 2018).

Molecular dynamics study

As we are aware of the remarkable biological functions of proteins and DNA and also their profound dynamic mechanisms which include switching between active and inactive states (Autiero et al., 2013), cooperative effects the intercalation of drugs into DNA (Chou and Mao, 1988), and allosteric transition (Ward et al., 2010) can be unveiled by studying their internal motions (Lin and Lapointe, 2013). Likewise, to understand the action mechanism of receptor-ligand binding, we should not consider only the static structures concerned, but also the dynamical information obtained by simulating their internal motions or dynamic processes (de Azevedo, Jr et al., 2001; de Azevedo, 2011; Qazi and Raza, 2021). To realize this, initially, MD Simulations were performed for proteins only, e.g., FurB,IdeR, KstR, RegX3, and MosRat 300K to investigate the stability profile. We observed that drug ZINC03871613 was best docked with the three targets (FurB, KstR, and RegX3), and drugs ZINC03871615, ZINC03871614 were docked with MosR and IdeR targets, respectively. These are the best candidate drugs that exhibit high binding affinity with the targets. To observe conformational changes due to the presence of ligands in the active site, we performed MD simulations for protein complexes too.

Molecular dynamics (MD) simulations were performed using the GROMACS version 5.1 software (van der Spoel et al., 2005) with the standard GROMOS 53A6 force field was used. The proteins were soaked in a cubic box of water molecules with a dimension of 10 Å, i.e., setting the box edge 10 Å from the molecule perimeters, using the editconf module for making boundary conditions and genbox for solvation. For the solvation of proteins, the spc216 template was used. The charges of the complete system are neutralized by the addition of Na+ and Cl ions (if the system has a positive charge (e.g., +6), add 6 Cl ions and if the system has a negative charge (e.g., −6), add 6 Na+ ions)) using the gmxgenion module to maintain neutrality. The system was then minimized using the 1500 steps of steepest descent at 300K during their equilibration period (100 ps) at a constant volume under periodic boundary conditions. Equilibration was performed in two phases: NVT ensemble (constant number of particles, volume, and temperature at 100 ps) and NPT ensemble (constant number of particles, pressure, and temperature at 100 ps). After the equilibration phase, the particle MeshEwald method (Cheatham et al., 1995) was used. Production phases were performed at 300K. The resulting trajectories were analyzed, using RMSD, RMSF, RG, and SASA by the utilities provided by GROMACS. For ligand topology, we used SWISSPARAM that provides topologies and parameters for small organic molecules, compatible with the CHARMM all atoms force field, for use with the CHARMM or GROMACS software (Bjelkmar et al., 2010; Zoete et al., 2011). The server is fully automatic, and we only need to provide the drug molecule in the mol2 format.

Protein–Protein interaction network

The protein-protein interaction (PPI) networks are much important for the system-level understanding of cellular processes (Cau et al., 2018). Such networks are used for screening and evaluation of functional genomics data and provide an instinctive platform for annotating structural, functional, and evolutionary properties of proteins. The five proteins (FurB, IdeR, KstR, RegX3, and MosR) were taken as seed proteins from which we obtained direct and indirect PPIs using theSTRING 10.5 database (Snel et al., 2000). This database provides details on both predicted and experimentally verified interactions from different sources based on their neighborhood, co-occurrence, co-expression, gene fusions, experiments, and literature mining. We constructed a PPI network based on high confidence interaction which means that only interactions with a high level of confidence are considered as potential associations in the PPI network.

Results

Virtual screening

In the first step, the library of FDA approved drugs were screened using GOLD versions 5.2 server. Dock runs of all 1985 drugs on targets namely “FurB (binding cavity: His-81, His-83, Cys-76, Glu-78, Asp-24, Asp-62, etc.), IdeR (binding cavity: Glu-20, Asp-17, Arg-80, etc.), KstR (binding cavity: Arg-174, Glu-112, Pro-108, Trp-17, etc.), RegX3 (binding cavity: Pro-102, Tyr-103, Ser-104, etc.) and MosR (binding cavity: Thr-36, His-287, Glu-244, etc.)” using GOLD suite (with 50 GA runs were performed) resulted in a few best candidate drugs evaluated based on the Gold fitness score 55 (The fitness score >50 is supposed to good binding).

We found a total of 828 drugs that were bound with all the targets like FurB, IdeR, KstR, MosR, and RegX3 that were docked with 111 , 355 , 193 , 157, and 12 drugs out of 1985, respectively. In the second step, the genetic search algorithm was initially set to 100 then 200 GA runs for 828 drugs to each target and in most cases, similar results were obtained. The docking and scoring details are shown in Fig. 3 and Suppl. Tab. S1. In the third step, we again filtered out the drugs that frequently bounded with all the targets at set threshold among the top 20 GoldScore ranks. It was found that only 11 candidate drugs (designated with Zinc Database ids as ZINC03871612, ZINC03871613, ZINC03871614, ZINC03871615, ZINC01529323, ZINC03830426, ZINC03830428, ZINC03830430, ZINC03830635, ZINC03799072, and ZINC08551107) are consistently binding with all the targets.

images

Figure 3: The plot represents the Gold Score of compounds that have scored equal, or above the threshold value of 55.0. (A) IdeR bind with 355 compounds (B) KstR bind with 193 compounds (C) MosR bind with 157 compounds (D) FurB bind with 111 compounds (E) RegX3 bind with 12 compounds.

The consistent binding idea was given preference for the selected 11 candidate drugs over top 20 (for each target) as a matter of feasibility towards in-vitro or in-vivo testing for these bindings. The 2D structures of these 11 candidate drugs can be obtained from the ZINC-15 database (https://zinc15.docking.org/). In this study, we have visualized only top-ranked candidate drugs which are best bound with the targets like FurB, KstR, and RegX3, are docked with ZINC03871613, while IdeR and MosR are docked with ZINC03871614, and ZINC03871615 with GoldScore 77.58, 75.02, 62.54, 81.69, and 79.72, respectively. While studying the docked conformation of the FurB complex, mainly the three residues, ASP-24, GLU-78, and HIS-81 were found to form H-bonds with the ligands in the docked complex along with 12 surrounding residues. As in the case of KstR and RegX3 complexes, ASP-82, PRO-102, SER-104, and LYS-157 (also forms salt bridge interaction) and the residues LEU-109, ARG-174 (also forms π–π and π-cation and salt bridge interactions) make H-bond with the ligand along with 14 and eight surrounding residues, respectively. While in the case of the MosR complex, the residues, ARG-20 (also make π–π and π-cation interaction), ASP-27, THR-36, ASN-37, THR-38, ARG-75 (also forms salt bridge interaction), and ASN-76 while in the IdeR complex, the residues ARG-13, ARG 33, ARG-80 (also make π–π interaction) were seemed to make H-bonds with the ligands along with six and 17 surrounding residues, respectively. The complete details are given in Fig. 4 and Tab. 1. All the top scorer candidate drugs are purine derivatives in the form of triphosphate and the basic difference among them is in the cyclic pyranose which are isomers of each other due to the different spatial orientation of the hydroxyl group.

MD simulation analysis

We performed MD simulation at 300K of the protein-ligand complex with top-ranked drugs to observe the effect on the structure, behavior, and flexibility of all the targets. We analyzed the RMSD, RMSF, RG, and SASA for all the selected targets, as shown in Fig. 5.

imagesimages

Figure 4: Binding poses are suggested by molecular docking. (A) Molecular docking poses of compound ZINC03871613 with FurB. (B) Compound ZINC03871614 docked with IdeR. (C) Compound ZINC03871613 docked with KstR. (D) Compound ZINC03871615 docked with MosR. (E) Compound ZINC03871613 docked with RegX3. The 2D depiction of the binding pocket and interactions are generated using Maestro with a 4.0 A cutoff. Protein residues forming hydrogen bonds are indicated with two different magenta lines (i) solid lines (backbone) (ii) dashes lines (side chain). The π-π stacking interaction between residues and the aromatic ring of the ligand is shown in the green solid line while π-Cation is a non-covalent molecular interaction is shown in the red solid line.

images

Figure 5: Molecular dynamics study. In this figure, we have shown RMSD, RMSF, RG, and SASA values each at 300k for our all targets (FurB, IdeR, KstR, MosR, and RegX3). The target protein and the complex (protein with ligands) are shown in green and red color, respectively.

images

Stability of the trajectory

a) Root-mean-square deviations (RMSD) from the native structure coordinates during the MD simulation were used to assess the stability and the reliability of the simulation. In our study, we noticed that FurB underwent structural fluctuations starting from the first 5ns of the trajectory, although the amplitude of the fluctuation was smaller for the FurB complex at 6.44ns , a slight transition of 0.325nm was spotted and the system got optimized after 8ns at 0.65nm . This significant drift in RMSD of 0.325nm is verified by the protein-ligand interactions within the complex. However, the simulation ends with a total drift of 0.28nm from the native structure. A significant drift in RMSD of FurB and FurBcomplex indicate the conformational change in the receptor that has occurred to fit the ligand in the active site of the FurB.

b) In the case of the IdeR, the trajectory (RMSD) remains stable throughout the simulation with small drifts, but the fluctuations were smaller too for the IdeRcomplex , up to 7.15ns . At the three points, a gradual transition of 0.083 , 0.096 , and 0.15nm were observed at 0.24,2.54and5.91ns , respectively and after 11ns , the trajectory was optimized. These significant drifts are verified by the protein-ligand interactions within the complex.

c) In the case of theRegX3, the trajectory (RMSD) drastically fluctuated with large drifts, but the fluctuations were smaller for the RegX3complex . A gradual transition of 1.056nm at 4.06ns was observed in the native structure, but the system is optimized after 12ns . A significant difference in RMSD of RegX3 and complex strongly indicates a conformational change in the protein due to ligand.

d) In the case of the MosR, the trajectory (RMSD) drastically fluctuated with large drifts, but the fluctuations were smaller for the MosRcomplex . A drastic transition of 0.5064nm at 1.07ns was observed in the native structure, but later system is optimized after 3.5ns . However, the simulation ends with a total drift of 0.249nm from the native structure, and this difference in RMSD indicates the conformational changes in the protein due to ligand.

e) In the case of the KstR, the trajectory (RMSD) remains stable throughout the simulation with small drifts of 0.103nm , but the fluctuation pattern was slightly different from the kstRcomplex , a drastic transition of 0.287nm is observed but the system is optimized after 8.6 ns. The average total drift in RMSD at the end of both simulations was 0.192nm . This difference in RMSD of kstR and kstRcomplex indicates a conformational change in the protein due to ligand.

The root mean square fluctuation (RMSF)

a) We have performed RMSF from its time-averaged position to set up the conformational fluctuations of the FurB and FurB complex at the residual level. The average residual fluctuation of all the residues in the case of FurBcomplex is lower than FurB. These results suggest that the α-helical structure of FurB is more specifically maintained due to the binding of the ligand.

b) RMSF analysis for the IdeR revealed that fluctuation of amino acids was largest in the region from 33–41 residues in IdeR, but in the case of IdeRcomplex , residues V56, A57, G58, and D59 have a greater RMSF score. The region 61–80 residues slightly fluctuated up and down in both cases (complex and protein), which confirms the important role being played by these residues in the ligand-binding process.

c) On analyzing the fluctuations in the RegX3 and RegX3complex complex, it was observed that fluctuation of amino acids was the largest in the two regions, 80–89 and 91–114 residues in both cases due to the presence of the active site in this region, which confirms the important role being played by these residues in ligand binding.

d) In RMSF analysis of theMosR, we observed that fluctuation of amino acids was largest in the two regions from 14–40 and 44–110 residues, in both MosR and mosRcomplex . The First fluctuation (between 14–40 residues) was started due to the presence of an active site where the ligand binds, and the second fluctuation occurred due to residual conformational changes in MosR protein.

e) The fluctuations of KstR are less than kstRcomplex, but at two points (77–92 and 149–172 residues) fluctuations are larger due to the presence of binding regions. In the presence of ligand with KstR protein, at two points, lower transitions were observed, and the system’s overall compactness shows an increase after 11ns . The difference in RMSFvalue (for both) to the endpoint of the simulation is 0.07nm . This significant difference in the RMSFvalue of the native structure indicates that the ligand fits into the binding site.

Radius of gyration (Rg)

a) The radius of gyration ( Rg ) analysis of the FurB was carried out to determine the compactness of the protein system during the simulations. In the presence of ligand, the average Rg score was drastically reduced from 1.80rg (native structure of FurB) to 1.59nm . This complex underwent two drastic transitions which were observed from 1.60nm to 1.40nm and 1.40nm to 1.55nm , then the system was optimized after 15.4ns at 1.5rg(nm) . The decrease in Rgscore from the native structure points out that ligand binds into the binding cavity of the FurB.

b) In the presence of ligand with IdeR protein, the average Rg score was 0.06rg decreased. A slight transition of 0.12nm was spotted and after 12.6ns, the system was optimized at 1.45rg and the average difference in rgscore at the end of both simulations was 0.11nm . This significant difference in Rgscore from the native structure points out that ligand binds into the binding site of the FurB.

c) In the case of RegX3, we observed structural fluctuations throughout the simulation, but in the case of RegX3complex the amplitude of the fluctuations was smaller. The decrease in the transition of 0.487nm occurred at 3.7ns , but the system was optimized after 14.6ns at 1.9nm . However, the simulation finished with a total drift of 0.69nm from the native structure and this difference indicates that the ligand binds into the RegX3 pocket.

d) In the presence of ligand with MosR, the average Rg score drastically decreases from 1.73nm (native structure) to 1.35n . This complex underwent structural fluctuations starting from first 0.17ns to 5.29ns and the compactness of the system increasing after 5.5ns at 1.37nm . This difference indicates that the ligand binds to the active site of MosR.

e) In presence of ligand with KstR protein, at two points, lower transitions were observed, and the system’s overall compactness witnessed an increase after 11n . The difference in rgscore (for both) to the endpoint of the simulation is 0.07nm . This significant difference in Rgscore from the native structure strongly indicates that the ligand binds to the active site of KstR.

Solvent accessible surface area

a) Solvent accessible surface area (SASA) remained relatively constant for the FurB protein at 73.68nm 2, but in the case of FurB complex the pattern is different, SASA decreases from 73.20nm 2 to 71.nm 2 and average fluctuation occurred near 70nm 2 upto 10ns . The system is optimized after 68.5nm 2 at the 11ns upto the endpoint of the simulation. The decrease in SASA caused by the change in protein conformation during its folding must be accompanied by the corresponding increase in the number of native contacts with the ligand, clearly indicating that ligand ( ZINC03871613 ) binds into the active site of the FurB.

b) In the case of IdeR, we observed the structural fluctuations in SASA throughout the simulation, but these fluctuations diminish later on and the overall system is optimized after 79.5 but in the case of IdeRcomplex the pattern is different, the average surfacearea is drastically decreased from 71.78nm 2 (native structure of IdeR) to 77.78nm 2. The decrease in SASA strongly indicates that inhibitor ( ZINC03871614 ) preferentially fits into the active site of IdeR.

c) In the case of RegX3, SASA remained stable up to 127.434nm 2 at 8.5ns , but the slight transition was observed at 7.183nm 2, but in the case of RegX3complex the pattern is different, the average surface area is slightly decreased from 123.75nm 2 (native structure of IdeR) to 120.014nm 2 and system is optimized after 6.68ns at 119nm 2. However, the simulation finished with a total drift off 10.209nm 2 from the native structure and this difference in SASA indicates the ligand binding into the RegX3 pocket.

d) The SASA remained relatively constant for the MosR at 70.25nm 2, but in the case of FurBcomplex the pattern is different; the SASA is decreasing from 69.59nm 2 to below 62nm 2, and therefore, systems compactness increases. The decrease in SASA strongly indicates that inhibitor ( ZINC03871615 ) preferentially fits into the active site of MosR.

e) In the case of KstR protein and complex, gradual upper transition and lower transition were observed, respectively, but SASA remained relatively constant, and the overall system is optimized after 7.5ns at 103nm 2.

Network analysis and nodes characterization

The Protein-protein interaction (PPI) network was constructed using the five seed genes from the STRING database. In the present study, we considered only the immediate interacting partners (not more than 5 interactors) that abided the medium confidence score of 0.400. We have considered occurrence across the genome (suggesting that protein has a similar function or an occurrence in the same metabolic pathway further which to be expressed together and hold similar phylogenetic profile.), neighborhood (similar genomic context in different species suggesting similar functions of the proteins), putative homologs, gene fusions (that means proteins that are fused in some genomes are the most likely to be functionally linked), co-expression. We observed that FurB is present in the neighborhood of MRA_2382 and Glys and IdeR, KstR, Regx3, and MosR are in the neighborhood of SigB , FadE34 , SenX3 , and Rv1050 respectively, while RegX3 is fused with PhoR , MtrB , and MtubH3_010100022685 . In the light, being a broad-spectrum antibiotics concept (i.e., designing multi-target drugs), it can be inferred from Tab. 2 that the interacting partners of RegX3 make it a most pliable target to any drug design for it. As homologs of RegX3‘s interacting partners are found co-expressed in other species of bacteria, it means it may be closely linked to MTB. In order to rank the targets based on their network interactions, KstR and RegX3 showcase highly influential interactions followed by MosR, IdeR, and FurB. The complete details of five seed genes and their interactions with other important genes are demonstrated in Fig. 6 and Tab. 2, which illustrates the PPI neighborhood of the 5 receptors (FurB, IdeR, KstR, RegX3, and MosR). The nodes of the graph represent proteins and the connections illustrate their known or predicted direct and indirect interactions.

images

Figure 6: Analysis of protein-protein interaction (PPI) using STRING 10.5 server. The figure illustrates the PPI neighborhood of the 5 receptors (FurB, IdeR, KstR, RegX3, and MosR), the nodes of the graph represent proteins and the connections illustrate their known or predicted direct and indirect interactions. The connection between any two proteinnodes is based on the available information mined from relevant.

images

Discussion

Till now TB has claimed the lives of millions of people across the world, despite our continuous efforts it is emerging as the next challenge for us to eradicate this ailment. Here we have tried to identify new scaffolds for drug discovery by repurposing FDA-approved drugs against M. tuberculosis. Almost all drugs repositioning concepts require at least some experimental validation of assessment for its efficacy, therefore we have adopted an in-silico approach to support experimental work in developing effective drugs in a shorter duration. Our systematic study has highlighted several important observations; molecular docking analysis of a set of 828 drugs with the best GoldScore ( 55 ) were selected and out of which 11 drugs were frequently present within the top 20 GoldScore ranks and ZINC03871613, ZINC03871614, ZINC03871615 are top scorer drugs. Hence, we consider all these 11 drugs as promising drugs as lead drugs for further research towards repurposing against TB.

The ROC curve applied to the retrospective analysis of a virtual screening experiment is a plot of the true positive fractions (TPR, y-axis) versus false positive fractions (FPR, x-axis) for all drugs in a ranked dataset. The ROC curve analysis reveals either perfect discrimination, i.e., passes through the upper left corner (100% sensitivity, 100% specificity), or defective discrimination. Therefore, the closer the ROC curve is to the upper left corner, the higher the overall accuracy of the test58. The area under the ROC curve (ROC AUC) outlines the overall performance thus higher the AUC, the higher is the accuracy. In this case, we performed former experimentin order to identify the interaction capability of the targets with other potential based on the fitness value calculated using various parameters, e.g., interaction entropy. We found fitness values give a perfect discrimination, i.e., highly accurate, while interaction entropy contributes to the larger part of this perfection in each case (for both active and inactive ligands) as shown in ROC curve analysis of active and inactive ligands against targets in Fig. 7.

The protein-ligand complex with top-ranked drugs was used for the dynamics simulation study in order to see the effect of top scorer drugs on the structural stability of all five targets. It was found that these top scorer drugs preferentially fit into the active site of the targets. RMSD of all the protein’s initial structures and protein-ligand complexes were calculated. Throughout the simulation period, significant drifts were observed in the backbone of FurB, IdeR, KstR, RegX3, and MosR and protein-ligand complexes, implying that binding of drug candidates at the active sites of the targets is not only preferentially fitting, but also disturbs the structure of proteins, i.e., the structure became more compact that is also proved by RG-score analysis (The lowest radius of gyration mean more compact structure). RMSF analysis suggests that the observed residual fluctuations were largest in active sites of proteins that confirm the important role being played by these residues in the ligand-binding process. The significant drifts in native structures and protein-ligand complexes were observed in SASA analysis. The decrease in SASA caused by the change in protein conformation during its folding must be accompanied by the corresponding increase in the number of native contacts with ligand, clearly indicating that ligands tightly fit into the active sites. We performed hydrogen bond analysis and extracted the complex at the particular interval (5000, 10,000, 15,000, 20,000 ps, respectively) in the PDB format using g_h bond utility of GROMACS. However, we have noticed that one or two residues (that formed H-bonds from ligands) from each protein are common in all 4 intervals of time (5, 10, 15, 20 ns) that indicate that ligands are tightly bounds in the active sites. Apart from virtual screening and simulation study, we analyzed the PPI network of target proteins to evaluate the functional genomics information and for providing an instinctive platform for annotating structural, functional, and evolutionary properties of proteins. For a better understanding of the effects of cellular interconnectedness on disease progression the analysis of PPI’s may lead to the detection of disease genes and disease pathways, which may offer better targets for drug development. It has been clearly shown that the method utilized in this study is successful in finding out the most promising drug candidates against these targets that play a key role in the progression of TB disease. This work can be further explored to study the receptor-ligand interactions experimentally and assessment of their biological activity would help in specifying drugs against these key targets of MTB.

Conclusion

Tuberculosis (TB), being one of the major public health concerns, still lacks a complete and quick cure due to the Mycobacterium tuberculosis (MTB) strains resistant to existing drugs. Hence, it is important to look for new and effective drugs that can inhibit several putative drug targets. In this paper, we executed a computational drug repurposing pipeline to identify promising FDA-approved drugs against five key-regulatory genes of Mtb, namely, FurB, IdeR, KstR, MosR, and RegX3. Our study found 11 drug candidates including ZINC03871613, ZINC03871614, and ZINC03871615 as the top three drug candidates. In order to validate these results, a molecular dynamics simulation study was performed which indicates that inhibitors preferentially bind to the active site of the targets. This finding opensa new application domain in the form of anti-tuberculosis agents. This study can be further extended to explore the receptor-ligand interactions experimentally and assessment of their biological activity against key targets of MTB.

imagesimages

Figure 7: ROC curve analysis of active and inactive ligands against targets.

Availability of Data and Materials: Available on request to the authors.

Author Contribution: The idea was conceived by XY and KR, an experiment was designed by AA and KR, result analysis by AA, and the manuscript written by XY, AA, and KR. All authors reviewed the results and approved the final version of the manuscript.

Conflicts of Interest: The authors declare that they have no conflicts of interest to report regarding the present study.

References

  1. Afzal, O., Kumar, S., Kumar, R., Firoz, A., Jaggi, M., & Bawa, S. (2014). Docking based virtual screening and molecular dynamics study to identify potential monoacylglycerol lipase inhibitors. Bioorganic & Medicinal Chemistry Letters, 24, 3986-3996. [Google Scholar]
  2. Ahmad, A., Cai, Y., Chen, X., Shuai, J., & Han, A. (2015). Conformational dynamics of response regulator RegX3 from Mycobacterium tuberculosis. PLoS One, 10, e0133389. [Google Scholar]
  3. Ahmad, F., & Kumar, S. (2016). Rewiring of interconnected hubs in Gene Interaction Network of and its survival in human host. Journal of Applied Computing, 1, 38-50. [Google Scholar]
  4. Ahmad, S. (2011). Pathogenesis, immunology, and diagnosis of latent Mycobacterium tuberculosis infection. Clinical and Developmental Immunology, 2011, 814943. [Google Scholar]
  5. Alam, A., Ali, S., Ahamad, S., Malik, MZ., & Ishrat, R. (2016). From ZikV genome to vaccine: approach for the epitope-based peptide vaccine against Zika virus envelope glycoprotein. Immunology, 149, 386-399. [Google Scholar]
  6. Almatar, M., Almandeal, H., Var, I., Kayar, B., & Köksal, F. (2017). New drugs for the treatment of infection. Biomedicine & Pharmacotherapy, 91, 546-558. [Google Scholar]
  7. Autiero, I., Langella, E., & Saviano, M. (2013). Insights into the mechanism of interaction between trehalose-conjugated beta-sheet breaker peptides and Aβ (1-42) fibrils by molecular dynamics simulations. Molecular BioSystems, 9, 2835-2841. [Google Scholar]
  8. Banerjee, S., Farhana, A., Ehtesham, NZ., & Hasnain, SE. (2011). Iron acquisition, assimilation and regulation in mycobacteria. Infection, Genetics and Evolution, 11, 825-838. [Google Scholar]
  9. Berman, HM., Westbrook, J., Feng, Z., Gilliland, G., & Bhat, TN. (2000). The protein data bank. Nucleic Acids Research, 28, 235-242. [Google Scholar]
  10. Bjelkmar, P., Larsson, P., Cuendet, MA., Hess, B., & Lindahl, E. (2010). Implementation of the CHARMM force field in GROMACS: Analysis of protein stability effects from correction maps, virtual interaction sites, and water models. Journal of Chemical Theory and Computation, 6, 459-466. [Google Scholar]
  11. Brugarolas, P., Movahedzadeh, F., Wang, Y., Zhang, N., & Bartek, IL. (2012). The oxidation-sensing regulator (MosR) is a new redox-dependent transcription factor in Mycobacterium tuberculosis. Journal of Biological Chemistry, 287, 37703-37712. [Google Scholar]
  12. Carroll, P., Faray-Kele, MC., & Parish, T. (2011). Identifying vulnerable pathways in by using a knockdown approach. Applied and Environmental Microbiology, 77, 5040-5043. [Google Scholar]
  13. Cau, Y., Valensin, D., Mori, M., Draghi, S., & Botta, M. (2018). Structure, function, involvement in diseases and targeting of 14-3-3 proteins: An update. Current Medicinal Chemistry, 25, 5-21. [Google Scholar]
  14. Cheatham, T., Miller, J., Fox, T., Darden, T., & Kollman, P. (1995). Molecular dynamics simulations on solvated biomolecular systems: The particle mesh Ewald method leads to stable trajectories of DNA, RNA, and proteins. Journal of the American Chemical Society, 117, 4193-4194. [Google Scholar]
  15. Chou, KC., & Mao, B. (1988). Collective motion in DNA and its role in drug intercalation. Biopolymers: Original Research on Biomolecules, 27, 1795-1815. [Google Scholar]
  16. De Azevedo Jr, WF., Canduri, F., Fadel, V., Teodoro, LG., & Hial, V. (2001). Molecular model for the binary complex of uropepsin and pepstatin. Biochemical and Biophysical Research Communications, 287, 277-281. [Google Scholar]
  17. De Azevedo, WF. (2011). Molecular dynamics simulations of protein targets identified in . Current Medicinal Chemistry, 18, 1353-1366. [Google Scholar]
  18. Dua, K., Rapalli, VK., Shukla, SD., Singhvi, G., & Shastri, MD. (2018). Multi-drug resistant Mycobacterium tuberculosis & oxidative stress complexity: Emerging need for novel drug delivery approaches. Biomedicine & Pharmacotherapy, 107, 1218-1229. [Google Scholar]
  19. García, JL., Uhía, I., & Galán, B. (2012). Catabolism and biotechnological applications of cholesterol degrading bacteria. Microbial Biotechnology, 5, 679-699. [Google Scholar]
  20. Heberlé, G., & De Azevedo, WF. (2011). Bio-inspired algorithms applied to molecular docking simulations. Current Medicinal Chemistry, 18, 1339-1352. [Google Scholar]
  21. Ho, NAT., Dawes, SS., Crowe, AM., Casabon, I., & Gao, C. (2016). The structure of the transcriptional repressor KstR in complex with CoA thioester cholesterol metabolites sheds light on the regulation of cholesterol catabolism in Mycobacterium tuberculosis. Journal of Biological Chemistry, 291, 7256-7266. [Google Scholar]
  22. India TB Report ( 2018). Ministry Of Health and Family Welfare.
  23. Irwin, JJ., Sterling, T., Mysinger, MM., Bolstad, ES., & Coleman, RG. (2012). ZINC: A free tool to discover chemistry for biology. Journal of Chemical Information and Modeling, 52, 1757-1768. [Google Scholar]
  24. Johansson, MU., Zoete, V., Michielin, O., & Guex, N. (2012). Defining and searching for structural motifs using DeepView/Swiss-PdbViewer. BMC Bioinformatics, 13, 173. [Google Scholar]
  25. Jones, G., Willett, P., & Glen, RC. (1995). Molecular recognition of receptor sites using a genetic algorithm with a description of desolvation. Journal of Molecular Biology, 245, 43-53. [Google Scholar]
  26. Jones, G., Willett, P., Glen, RC., Leach, AR., & Taylor, R. (1997). Development and validation of a genetic algorithm for flexible docking. Journal of Molecular Biology, 267, 727-748. [Google Scholar]
  27. Karaman, B., & Sippl, W. (2019). Computational drug repurposing: Current trends. Current Medicinal Chemistry, 26, 5389-5409. [Google Scholar]
  28. Karthick, V., & Ramanathan, K. (2013). Virtual screening for oseltamivir-resistant a (H5N1) influenza neuraminidase from traditional Chinese medicine database: A combined molecular docking with molecular dynamics approach. Springerplus, 2, 115. [Google Scholar]
  29. Kendall, SL., Burgess, P., Balhana, R., Withers, M., & Ten Bokum, A. (2010). Cholesterol utilization in mycobacteria is controlled by two TetR-type transcriptional regulators: kstR and kstR2. Microbiology, 156, 1362. [Google Scholar]
  30. Khalid, Z., Ahmad, S., Raza, S., & Azam, SS. (2018). Subtractive proteomics revealed plausible drug candidates in the proteome of multi-drug resistant . Meta Gene, 17, 34-42. [Google Scholar]
  31. King-Scott, J., Nowak, E., Mylonas, E., Panjikar, S., & Roessle, M. (2007). The structure of a full-length response regulator from Mycobacterium tuberculosis in a stabilized three-dimensional domain-swapped, activated state. Journal of Biological Chemistry, 282, 37717-37729. [Google Scholar]
  32. Krueger, J., Thiel, P., Merelli, I., Grunzke, R., & Gesing, S. (2016). Portals and web-based resources for virtual screening. Current Drug Targets, 17, 1649-1660. [Google Scholar]
  33. Lanza, V., Milardi, D., Di Natale, G., & Pappalardo, G. (2018). Repurposing of copper (II)-chelating drugs for the treatment of neurodegenerative diseases. Current Medicinal Chemistry, 25, 525-539. [Google Scholar]
  34. Li, MS., Monahan, IM., Waddell, SJ., Mangan, JA., & Martin, SL. (2001). cDNA-RNA subtractive hybridization reveals increased expression of mycocerosic acid synthase in intracellular Mycobacterium bovis BCG. Microbiology, 147, 2293-2305. [Google Scholar]
  35. Lin, SX., & Lapointe, J. (2013). Theoretical and experimental biology in one—A symposium in honour of Professor Kuo-Chen Chou’s 50th anniversary and Professor Richard Giegé’s 40th anniversary of their scientific careers. Journal of Biomedical Sciences and Engineering, 06, 435-442. [Google Scholar]
  36. Lucarelli, D., Russo, S., Garman, E., Milano, A., Meyer-Klaucke, W., & Pohl, E. (2007). Crystal structure and function of the zinc uptake regulator FurB from . Journal of Biological Chemistry, 282, 9914-9922. [Google Scholar]
  37. Morrone Xavier, M., Sehnem Heck, G., Boff De Avila, M., Maria Bernhardt Levin, N., & Oliveira Pintro, V. (2016). SAnDReS a computational tool for statistical analysis of docking results and development of scoring functions. Combinatorial Chemistry & High Throughput Screening, 19, 801-812. [Google Scholar]
  38. Palos, I., Lara-Ramirez, EE., Lopez-Cedillo, JC., Garcia-Perez, C., & Kashif, M. (2017). Repositioning FDA drugs as potential cruzain inhibitors from Trypanosoma cruzi: Virtual screening, and studies. Molecules, 22, 1015. [Google Scholar]
  39. Panina, EM., Mironov, AA., & Gelfand, MS. (2003). Comparative genomics of bacterial zinc regulons: Enhanced ion transport, pathogenesis, and rearrangement of ribosomal proteins. Proceedings of the National Academy of Sciences of the United States of America, 100, 9912-9917. [Google Scholar]
  40. Pei, JF., Qi, N., Li, YX., Wo, J., & Ye, BC. (2021). RegX3-mediated regulation of methylcitrate cycle in . Frontiers in Microbiology, 12, 119. [Google Scholar]
  41. Peters, JS., Ismail, N., Dippenaar, A., Ma, S., & Sherman, DR. (2020). Genetic diversity in Mycobacterium tuberculosis clinical isolates and resulting outcomes of tuberculosis infection and disease. Annual Review of Genetics, 54, 511-537. [Google Scholar]
  42. Pohl, E., Holmes, RK., & Hol, WG. (1999). Crystal structure of the iron-dependent regulator (IdeR) from shows both metal binding sites fully occupied. Journal of Molecular Biology, 285, 1145-1156. [Google Scholar]
  43. Potemkin, V., & Grishina, M. (2018). Grid-based technologies for in silico screening and drug design. Current Medicinal Chemistry, 25, 3526-3537. [Google Scholar]
  44. Qazi, S., & Raza, K. (2021). Phytochemicals from Ayurvedic plants as potential medicaments for ovarian cancer: An analysis. Journal of Molecular Modeling, 27, 114. [Google Scholar]
  45. Rester, U. (2008). From virtuality to reality-Virtual screening in lead discovery and lead optimization: A medicinal chemistry perspective. Current Opinion in Drug Discovery & Development, 11, 559-568. [Google Scholar]
  46. Roberts, G., Vadrevu, IS., Madiraju, MV., & Parish, T. (2011). Control of CydB and GltA1 expression by the SenX3 RegX3 two component regulatory system of Mycobacterium tuberculosis. PLoS One, 6, e21090. [Google Scholar]
  47. Rohilla, A., Khare, G., & Tyagi, AK. (2017). Virtual Screening, pharmacophore development and structure based similarity search to identify inhibitors against IdeR, a transcription factor of . Scientific Reports, 7, 4653. [Google Scholar]
  48. Rollinger, JM., Stuppner, H., & Langer, T. (2008). Virtual screening for the discovery of bioactive natural products. Natural Compounds as Drugs, I, 211-249. [Google Scholar]
  49. Sanz, J., Navarro, J., Arbués, A., Martín, C., & Marijuán, PC. (2011). The transcriptional regulatory network of Mycobacterium tuberculosis. PLoS One, 6, e22178. [Google Scholar]
  50. Savoia, D. (2016). New antimicrobial approaches: Reuse of old drugs. Current Drug Targets, 17, 731-738. [Google Scholar]
  51. Singh, P., Jaiyeola, B., Kerru, N., Ebenezer, O., & Bissessur, A. (2017). A review of recent advancements in anti-tubercular molecular hybrids. Current Medicinal Chemistry, 24, 4180-4212. [Google Scholar]
  52. Snel, B., Lehmann, G., Bork, P., & Huynen, MA. (2000). STRING: A web-server to retrieve and display the repeatedly occurring neighbourhood of a gene. Nucleic Acids Research, 28, 3442-3444. [Google Scholar]
  53. van der Spoel, D., Lindahl, E., Hess, B., Groenhof, G., Mark, AE., & Berendsen, HJ. (2005). GROMACS: Fast, flexible, and free. Journal of Computational Chemistry, 26, 1701-1718. [Google Scholar]
  54. Wang, Q., Canutescu, AA., & Dunbrack, RL. (2008). SCWRL and MolIDE: Computer programs for side-chain conformation prediction and homology modeling. Nature Protocols, 3, 1832. [Google Scholar]
  55. Ward, SK., Abomoelak, B., Marcus, S., & Talaat, AM. (2010). Transcriptional profiling of during infection: Lessons learned. Frontiers in Microbiology, 1, 121. [Google Scholar]
  56. Weiner, J., Domaszewska, T., Donkor, S., Kaufmann, SH., Hill, PC., & Sutherland, JS. (2020). Changes in transcript, metabolite, and antibody reactivity during the early protective immune response in humans to infection. Clinical Infectious Diseases, 71, 30-40. [Google Scholar]
  57. Zoete, V., Cuendet, MA., Grosdidier, A., & Michielin, O. (2011). SwissParam: A fast force field generation tool for small organic molecules. Journal of Computational Chemistry, 32, 2359-2368. [Google Scholar]

images

images This work is licensed under a Creative Commons Attribution 4.0 International License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.