Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Seeking antiviral drugs to inhibit SARS-CoV-2 RNA dependent RNA polymerase: A molecular docking analysis

Abstract

COVID-19 outbreak associated with the severe acute respiratory syndrome coronavirus (SARS-CoV-2) raised health concerns across the globe and has been considered highly transmissible between people. In attempts for finding therapeutic treatment for the new disease, this work has focused on examining the polymerase inhibitors against the SARS-CoV-2 nsp12 and co-factors nsp8 and nsp7. Several polymerase inhibitors were examined against PDB ID: 6M71 using computational analysis evaluating the ligand’s binding affinity to replicating groove to the active site. The findings of this analysis showed Cytarabine of -5.65 Kcal/mol with the highest binding probability (70%) to replicating groove of 6M71. The complex stability was then examined over 19 ns molecular dynamics simulation suggesting that Cytarabine might be possible potent inhibitor for the SARS-CoV-2 RNA Dependent RNA Polymerase.

1. Introduction

The 2019 coronavirus disease (COVID-19) pandemic associated with the severe acute respiratory syndrome coronavirus (SARS-CoV-2) has become a humanitarian crisis [1]. Similar epidemics associated with viral infections of the coronaviruses were reported over the past two decades including the pandemics of the previous severe acute respiratory syndrome coronavirus (SARS-CoV) identified in late year of 2002 [24] and the influenza pandemic (H1N1) identified in the year of 2009 [5, 6], in addition to the Middle East Respiratory Syndrome coronavirus (MERS-CoV) that was found in Saudi Arabia during the year of 2012 [710].

Human coronaviruses (HCoVs) target the human respiratory system, mainly the lungs. Past reported pandemics related to such coronavirus infection belonged to the Alphacoronaviruses family of 229E and NL63 strains, and the Betacoronaviruses family of OC43, HKU1, SARS, and MERS strains [11]. The most aggressive coronavirus infection was associated with both SARS and MERS strains. The newly emerged SARS-CoV-2 virus is highly contagious and transmissible across the nations, especially with the new variant surges. Analyzing the genomic sequence of the newly emerged SARS-CoV-2 demonstrated approximately sequence identity of 88% to that of SARS genomic sequence validating SARS-CoV-2 as a new member of Betacoronaviruses [11], [1214]. Epidemiological studies reported the symptoms of the SARS-CoV-2 similar to those symptoms caused by other Betacoronaviruses [13, 1518].

The HCoVs are described as lengthy positive single-stranded RNA viruses of about 30K bp [19] and carrying structural and non-structural proteins. There are four structural proteins that characterize all coronaviruses: the spike protein (S), the nucleocapsid protein (N), the membrane protein (M), and the envelope protein (E), as well as non-structural proteins like proteases (nsp3 and nsp5) and the RNA-dependent-RNA polymerase RdRp (nsp12) [2023]. SARS-CoV-2 is a positive strand RNA virus with numerous component replication-and-transcription complexes of viral nonstructural proteins (nsps) that control its replication [24, 25]. Nsp12 function is dependent on accessory proteins like as nsp7 and nsp8 [22, 26]. Nsp12 contains N-terminal domain (NiRAN), an interface domain, and C-terminal RdRp domain [27]. Fingers, palm, and thumb subdomains make up the RNA dependent RNA polymerase (RdRp) domain, where nsp7 and nsp8 subunits attach to the thumb and an additional copy of nsp8 attach itself to the fingers [22, 26, 28].

The conserved polymerase motifs A-G in the palm domain form the active site of the RdRp domain. The traditional divalent-cation-binding residue D618, which is conserved in most viral polymerases, is found in motif A. In the turn between two β-strands, motif C contains the catalytic residues (759-SDD-761), these catalytic residues are also conserved in most viral RdRps, with the first residue being either serine or glycine. Motif D stabilizes the core structure while motif E controls the flexibility of the thumb. Motif F contains K545, K551 and R553 which are responsible for rNTP binding and positioning. Motif G predicted to be involved in positioning of template overhang [20, 23, 2931].

Developing a therapeutic antiviral treatment that is safe and effective would take few years, therefore, in February 2020, the World Health Organization (WHO) research forum on the coronavirus disease 2019 recommended the evaluation of the commonly used approved antiviral regimens against COVID-19 [32, 33]. The regimen recommended searching all previously approved antiviral drugs against the coronavirus disease, this screening process would speed the process of finding a quick antiviral drug for COVID-19. Comprehensive computational studies repurposed approved antiviral drugs against SARS-CoV-2 [3438], where commonly used approved antiviral drugs were examined against SARS-CoV-2 protein structures including the RNA-dependent-RNA polymerase [3944], papin-like protease [4548], and the main protease [4954], using in-silico molecular docking to seek potential SARS-CoV-2 inhibitors by analyzing the binding probabilities [5562]. The findings of those studies suggested that the top scoring drugs could be used as lead compounds for further experimental validation for the development of effective antiviral treatment against SARS-CoV-2.

A large-scale analysis of regularly used antiviral medications may provide therapeutic possibilities that may be positioned to speed up experimental and clinical testing. In this study, we looked through the drug library for authorized antiviral medications to investigate possible antiviral activity against SARS-CoV-2. The current research work is an in-silico analysis seeking authorized antiviral treatments that inhibit SARS-CoV-2 RNA-dependent-RNA polymerase (SARS-CoV-2 RdRp). We used an in-silico approach to shortlist polymerase inhibitor candidate drugs, and we analyzed published studies. Our analysis identified a few candidate drugs, some of which are already being investigated for COVID-19 treatment and can serve as a basis for prioritizing additional viable COVID-19 candidate drugs.

2. Materials and methods

2.1. SARS-CoV-2 RdRp Structure

The SARS-CoV-2 RdRp (nsp12) complexed with its cofactors nsp8 and nsp7 in apo form was obtained from the Protein Data Bank (PDB ID: 6M71). The 6M71 structure consists of four chains: the A chain (nsp12), the B and the D chains (nsp8) and the C chain (nsp7). Proteins were prepared for molecular docking analysis using the AutoDock Vina protocol [63].

2.2. Polymerase inhibitors optimization and molecular docking

The structures of the polymerase inhibitors were downloaded from the DrugBank [64]. The MMFF94 force field function of Avogadro software was used to optimize the geometry of all inhibitors [65]. Molecular Docking analysis was performed using the AutoDock Vina protocol [63]. The docking was performed against the entire protein to evaluate the free natural affinity of the binding ligand for the replicating groove without pushing the ligand to dock selectively to the active region. The docking was repeated 10 times for each ligand, and the affinity of docking was assessed using the docking scores and the likelihood of binding to the replicating groove.

2.3. Analysis of interactions between inhibitors and RdRp

The fully automated protein–ligand interaction profiler (PLIP) web tool was used. PLIP detects and visualizes protein–ligand interaction patterns in 3D structures, either directly from the PDB or from user-supplied structures [66]. Results are presented in 3D interaction diagrams for manual examination, either online using JSmol or offline using PyMOL, as well as XML and text files for additional processing for each binding site [66]. The PLIP web tool was used to examine the interactions established between the inhibitors and the SARS-CoV-2 RdRp to evaluate the docking results. All interactions are described, down to the atom level, allowing for detailed analysis of specific binding properties. Ligand efficiency is the binding affinity divided by a measure of the size of a ligand [67]. Compounds that can provide the desired binding affinity with fewer atoms are considered efficient [6870].

2.4. Molecular dynamics simulations

CHARMM-GUI was used to create the protein topologies and the parameter files [7173]. GROMACS-2019 software package [74] and CHARMM36 force field [75] were used for the molecular dynamics simulation. The system was solvated with TIP3P water in the add solvation box [76] and the entire complexes were neutralized by using the Monte-Carlo ion-placing approach to add appropriate amounts of K+ and Cl ions. The system was energy-minimized for 5000 steps using the steepest descent approach before simulations [77] and equilibrated for 125 ps at constant number of molecules, volume, and temperature (NVT). Finally, the molecular dynamics simulations were performed for 1900 ps (19 ns) at constant temperature (310 K), pressure (1 atm), and number of molecules (NPT ensemble), and was good enough for RMSD straight line [78], [79]. Ramachandran plot analysis was carried out for validating the docked complex structure. The root mean square deviation (RMSD) of the protein atom backbone, the radius of gyration (Rg) and the number of hydrogen bonds and solvent accessible surface area (SASA) were plotted as a function of time [80]. The average root mean square fluctuation (RMSF) was then plotted as a function of residues number. Compressed coordinates were measured every 10 ps (1900 frames).

3. Results and discussion

Molecular docking approach was employed on 6M71 and the mean values of the docking scores and the probabilities of binding to the replicating groove are shown in Table 1. Cytarabine (-5.65 Kcal/mol) had the best likelihood (70%) of binding to the replicating groove of 6M71 based on the docking results.

thumbnail
Table 1. List of the molecular docking scores in Kcal/mol calculated using AutoDock Vina against SARS-CoV-2 (PDB ID: 6M71, nsp12-nsp8-nsp7).

The docking procedure was done ten times for each ligand, and the likelihood of binding to the replicating groove were determined. The highest probabilities of binding to PDB ID: 6M71 are shown in bold red color.

https://doi.org/10.1371/journal.pone.0268909.t001

To investigate the likely reasons for the binding energy differences, we examined the formed complexes using PLIP web server. Ligands are shown in licorice color, while the protein residues are labeled with one-letter code. The H-bonds are represented in solid yellow lines. Fig 1 illustrates the formed interactions between Cytarabine and 6M71 after the molecular docking. The three H-bonds formed between Cytarabine molecules and 6m71, two of them formed with D761 and one was formed with D760 (catalytic residues in C motif). The catalytic residues’ function was limited by hydrogen bonds formed with the active site pocket, which prevented them from being shared in virus replication.

thumbnail
Fig 1. Cytarabine, with a -5.65 Kcal/mol, has the highest likelihood (70%) of binding to RdRp’s replicating groove (6M71).

Ligands are labeled with a three-letter code, whereas protein residues are tagged with a licorice color. H-bonds are represented by solid yellow lines.

https://doi.org/10.1371/journal.pone.0268909.g001

The docking mechanism created fast connections with the legends and the protein, which could be unstable [81]. The molecular dynamics simulations provide information on the stability of the generated complexes’ molecular interactions. Based on the binding energy, Cytarabine expressed their highest probability to bind to the 6M71 protein. The stability of the complex was assessed using the RMSD for the backbone atoms of 6M71 protein in comparison to the initial structures [82]. Fig 2 shows a graph of the RMSD values of the 6M71-Cytarabine complex after 1900 ps stabilization (19 ns). Furthermore, the stability of the complex was assessed by graphing Rg [82]. Fig 2 shows the computed Rg values along the simulation time scale, showing that the parameter is stable for the 6M71-Cytarabine complex over time. Fig 3 depicts the number of hydrogen bonds that exist between 6M71 and Cytarabine. During the simulation, the number of hydrogen bonds in the complex varies from 0 to 5. Similar findings were made using SASA analysis, which represented the solvent-defined protein surface and its orientation during the folding process, resulting in changes in the exposed and buried areas of the protein surface area. Fig 4 depicts the results of SASA plotted over the simulation time. Fig 4 as well displays a convincing SASA value for the 6M71-Cytarabine solvation profile, indicating a stable structure and robust binding contact with the Cytarabine. Fig 5 shows the average RMSF for 6M71 over 19 ns per residue. The variations of the 6M71 catalytic residues ASP-60 and ASP-61, which form hydrogen bonds with Cytarabine, are less than 1.5 Ao, indicating that the contact is robust and stable. In summary, the approved drugs (Cytarabine, Streptolydigin, Ribavirin, Sofosbuvir, Deacetylcolchiceine, Mithramycin, Avigan, Remdesivir, IDX-184) can bind to SARS-CoV-2 RdRp, with different binding energies.

thumbnail
Fig 2. Graphs of RMSD for the backbone atoms and Rg as a function of time are shown over the course of a 19-ns simulation.

https://doi.org/10.1371/journal.pone.0268909.g002

thumbnail
Fig 3. Graphs depicting the number of hydrogen bonds formed between RdRp and Cytarabine as a function of time during a 19-ns simulation.

https://doi.org/10.1371/journal.pone.0268909.g003

thumbnail
Fig 4. Graphs of SASA of RdRp as a function of time for a 19 ns simulation.

https://doi.org/10.1371/journal.pone.0268909.g004

thumbnail
Fig 5. Graphs depicting the average RMSF per residue throughout the course of a 19-ns RdRp simulation.

https://doi.org/10.1371/journal.pone.0268909.g005

4. Conclusion

Since the start of the COVID-19 pandemic, a large and impressive number of research studies and clinical trials have been launched attempting to find treatments for the rapidly spreading coronavirus infection. The prospective approach of medication repurposing research employing in-silico screening techniques has been shown to be successful in identifying active compounds against SARS-CoV-2-targeted proteins. The goal of this work was to find prospective candidates among the licensed antiviral medications that can bind and interact with SARS-CoV-2 RdRp using a drug repurposing approach. The study examined a variety of polymerase inhibitors that are currently on the market to inhibit the SARS-CoV-2 RNA-dependent-RNA polymerase. Because Cytarabine showed the highest likelihood of binding to the active site pocket of the SARS-CoV-2, the results of the current in-silico molecular docking analysis employing binding affinity and interactions may support the use of Cytarabine as a possible candidate inhibitor for the treatment of COVID-19. However, major concern about such treatment method is its side effect inside the human body, considering some of Cytarabine side effect when used as a chemotherapeutic agent for Leukemia, further studies are needed to evaluate its biological significance within humans to justify its overall significance.

References

  1. 1. Acuti C, Elena M, Cappadona R, Bravi F, Mantovani L, Manzoli L. Advances in Biological Regulation SARS-CoV-2 pandemic: An overview. 2020;77: 1–11.
  2. 2. Drosten C, Preiser W, Günther S, Schmitz H, Doerr HW. Severe acute respiratory syndrome: identification of the etiological agent. Trends Mol Med. 2003;9: 325–327. pmid:12928032
  3. 3. Zhang W, Zheng Q, Yan M, Chen X, Yang H, Zhou W, et al. Structural characterization of the HCoV-229E fusion core. Biochem Biophys Res Commun. 2018;497: 705–712. pmid:29458023
  4. 4. Belouzard S, Millet JK, Licitra BN, Whittaker GR. Mechanisms of coronavirus cell entry mediated by the viral spike protein. Viruses. 2012;4: 1011–1033. pmid:22816037
  5. 5. Morens DM, Taubenberger JK, Harvey HA, Memoli MJ. The 1918 influenza pandemic: lessons for 2009 and the future. Crit Care Med. 2010;38: e10–e20. pmid:20048675
  6. 6. Al-Muharrmi Z. Understanding the Influenza A H1N1 2009 Pandemic. Sultan Qaboos Univ Med J. 2010/07/19. 2010;10: 187–195. Available: https://pubmed.ncbi.nlm.nih.gov/21509228 pmid:21509228
  7. 7. Cascella M, Rajnik M, Cuomo A, Dulebohn SC, Di Napoli R. Features, Evaluation and Treatment Coronavirus (COVID-19). StatPearls. 2020.
  8. 8. Tyrovolas S, El Bcheraoui C, Alghnam SA, Alhabib KF, Almadi MAH, Al-Raddadi RM, et al. The burden of disease in Saudi Arabia 1990–2017: results from the Global Burden of Disease Study 2017. Lancet Planet Heal. 2020;4: e195–e208. pmid:32442495
  9. 9. Al-Dorzi HM, Van Kerkhove MD, Peiris JSM, Arabi YM. Middle east respiratory syndrome coronavirus. ERS Monogr. 2016;2016: 21–34.
  10. 10. Ashour HM, Elkhatib WF, Rahman MM, Elshabrawy HA. Insights into the Recent 2019 Novel Coronavirus (SARS-CoV-2) in Light of Past Human Coronavirus Outbreaks. Pathog (Basel, Switzerland). 2020;9. pmid:32143502
  11. 11. Hui DS, I Azhar E, Madani TA, Ntoumi F, Kock R, Dar O, et al. The continuing 2019-nCoV epidemic threat of novel coronaviruses to global health—The latest 2019 novel coronavirus outbreak in Wuhan, China. Int J Infect Dis. 2020;91: 264–266. pmid:31953166
  12. 12. Lu R, Zhao X, Li J, Niu P, Yang B, Wu H, et al. Genomic characterisation and epidemiology of 2019 novel coronavirus: implications for virus origins and receptor binding. Lancet. 2020;395: 565–574. pmid:32007145
  13. 13. Wang W, Tan J, Wei F. Updated understanding of the outbreak of 2019 novel coronavirus (2019-nCoV) in Wuhan, China. J Med Virol. 2020;92: 441–447. pmid:31994742
  14. 14. Bogoch II, Watts A, Thomas-Bachli A, Huber C, Kraemer MUG, Khan K. Pneumonia of Unknown Etiology in Wuhan, China: Potential for International Spread Via Commercial Air Travel. J Travel Med. 2020; 1–3. pmid:31943059
  15. 15. Huang C, Wang Y, Li X, Ren L, Zhao J, Hu Y, et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet. 2020;395: 497–506. pmid:31986264
  16. 16. Ren L-L, Wang Y-M, Wu Z-Q, Xiang Z-C, Guo L, Xu T, et al. Identification of a novel coronavirus causing severe pneumonia in human. Chin Med J (Engl). 2020; 1. pmid:32004165
  17. 17. Graham Carlos W, Dela Cruz CS, Cao B, Pasnick S, Jamil S. Novel Wuhan (2019-NCoV) coronavirus. Am J Respir Crit Care Med. 2020;201: P7–P8. pmid:32004066
  18. 18. Li Q, Guan X, Wu P, Wang X, Zhou L, Tong Y, et al. Early Transmission Dynamics in Wuhan, China, of Novel Coronavirus–Infected Pneumonia. N Engl J Med. 2020; 1199–1207. pmid:31995857
  19. 19. Naqvi AAT, Fatima K, Mohammad T, Fatima U, Singh IK, Singh A, et al. Insights into SARS-CoV-2 genome, structure, evolution, pathogenesis and therapies: Structural genomics approach. Biochim Biophys acta Mol basis Dis. 2020;1866: 165878. pmid:32544429
  20. 20. Gao Y, Yan L, Huang Y, Liu F, Zhao Y, Cao L, et al. Structure of the RNA-dependent RNA polymerase from COVID-19 virus. Science (80-). 2020;368: 779–782. pmid:32277040
  21. 21. Lehmann KC, Gulyaeva A, Zevenhoven-Dobbe JC, Janssen GMC, Ruben M, Overkleeft HS, et al. Discovery of an essential nucleotidylating activity associated with a newly delineated conserved domain in the RNA polymerase-containing protein of all nidoviruses. Nucleic Acids Res. 2015;43: 8416–8434. pmid:26304538
  22. 22. Kirchdoerfer RN, Ward AB. Structure of the SARS-CoV nsp12 polymerase bound to nsp7 and nsp8 co-factors. Nat Commun. 2019;10: 1–9.
  23. 23. Elfiky AA. SARS-CoV-2 RNA dependent RNA polymerase (RdRp) targeting: an in silico perspective. J Biomol Struct Dyn. 2020. pmid:32338164
  24. 24. Ziebuhr J. The coronavirus replicase. Curr Top Microbiol Immunol. 2005;287: 57–94. pmid:15609509
  25. 25. Compans RW, Cooper MD, Honjo T, Melchers F, Olsnes S, Vogt PK. The coronavirus replicase. Current Topics in Microbiology and Immunology. 2005. pp. 287, 57–94. Available: pmid:15609509
  26. 26. Subissi L, Posthuma CC, Collet A, Zevenhoven-dobbe JC, Gorbalenya AE. One severe acute respiratory syndrome coronavirus protein complex integrates processive RNA polymerase and exonuclease activities. 2014.
  27. 27. Hillen HS, Kokic G, Farnung L, Dienemann C, Tegunov D, Cramer P. Structure of replicating SARS-CoV-2 polymerase. Nature. 2020;584. pmid:32438371
  28. 28. Yin W, Yin W, Mao C, Luan X, Shen D, Shen Q, et al. Structural basis for inhibition of the RNA-dependent RNA polymerase from SARS-CoV-2 by remdesivir. 2020;1560: 1–10. Available: https://science.sciencemag.org/content/sci/early/2020/04/30/science.abc1560.full.pdf pmid:32358203
  29. 29. Xu X, Liu Y, Weiss S, Arnold E, Sara SG. Molecular model of SARS coronavirus polymerase: implications for biochemical functions and drug design. 2003;31.
  30. 30. Gong P, Peersen OB. Structural basis for active site closure by the poliovirus RNA-dependent RNA polymerase. 2010;107: 22505–22510.
  31. 31. Appleby TC, Perry JK, Murakami E, Barauskas O, Feng J, Cho A, et al. Structural basis for RNA replication by the hepatitis C virus polymerase. 2015;347.
  32. 32. Repurposed Antiviral Drugs for Covid-19—Interim WHO Solidarity Trial Results. N Engl J Med. 2020;384: 497–511. pmid:33264556
  33. 33. WHO. A coordinated global research roadmap: 2019 novel coronavirus. March 2020. In: World Health Organization [Internet]. 2020 [cited 8 Jan 2021]. Available: https://www.who.int/publications/m/item/a-coordinated-global-research-roadmap
  34. 34. Adhami M, Sadeghi B, Rezapour A, Haghdoost AA, MotieGhader H. Repurposing novel therapeutic candidate drugs for coronavirus disease-19 based on protein-protein interaction network analysis. BMC Biotechnol. 2021;21: 22. pmid:33711981
  35. 35. Barh D, Tiwari S, Weener ME, Azevedo V, Góes-Neto A, Gromiha MM, et al. Multi-omics-based identification of SARS-CoV-2 infection biology and candidate drugs against COVID-19. Comput Biol Med. 2020;126: 104051. pmid:33131530
  36. 36. Das S, Camphausen K, Shankavaram U. In silico Drug Repurposing to combat COVID-19 based on Pharmacogenomics of Patient Transcriptomic Data. Research square. 2020. pmid:32702730
  37. 37. Gordon DE, Jang GM, Bouhaddou M, Xu J, Obernier K, White KM, et al. A SARS-CoV-2 protein interaction map reveals targets for drug repurposing. Nature. 2020;583: 459–468. pmid:32353859
  38. 38. Yi D, Li Q, Pang L, Wang Y, Zhang Y, Duan Z, et al. Identification of a Broad-Spectrum Viral Inhibitor Targeting a Novel Allosteric Site in the RNA-Dependent RNA Polymerases of Dengue Virus and Norovirus. Front Microbiol. 2020;11: 1440. pmid:32670253
  39. 39. Elfiky AA. Ribavirin, Remdesivir, Sofosbuvir, Galidesivir, and Tenofovir against SARS-CoV-2 RNA dependent RNA polymerase (RdRp): A molecular docking study. Life Sci. 2020;253: 117592. pmid:32222463
  40. 40. Aftab SO, Ghouri MZ, Masood MU, Haider Z, Khan Z, Ahmad A, et al. Analysis of SARS-CoV-2 RNA-dependent RNA polymerase as a potential therapeutic drug target using a computational approach. J Transl Med. 2020;18: 275. pmid:32635935
  41. 41. Saha S, Nandi R, Vishwakarma P, Prakash A, Kumar D. Discovering Potential RNA Dependent RNA Polymerase Inhibitors as Prospective Drugs Against COVID-19: An in silico Approach. Front Pharmacol. 2021;12: 267. pmid:33716752
  42. 42. Muhammed Y, Yusuf Nadabo A, Pius M, Sani B, Usman J, Anka Garba N, et al. SARS-CoV-2 spike protein and RNA dependent RNA polymerase as targets for drug and vaccine development: A review. Biosaf Heal. 2021. pmid:34396086
  43. 43. Molavi Z, Razi S, Mirmotalebisohi SA, Adibi A, Sameni M, Karami F, et al. Identification of FDA approved drugs against SARS-CoV-2 RNA dependent RNA polymerase (RdRp) and 3-chymotrypsin-like protease (3CLpro), drug repurposing approach. Biomed Pharmacother. 2021/03/31. 2021;138: 111544. pmid:34311539
  44. 44. Pirzada RH, Haseeb M, Batool M, Kim M, Choi S. Remdesivir and ledipasvir among the fda-approved antiviral drugs have potential to inhibit sars-cov-2 replication. Cells. 2021;10. pmid:33946869
  45. 45. Khater I, Nassar A. Repurposing Antiviral Drugs to Inhibit SARS-CoV-2 Papin-Like Protease Activity. Int J Intell Comput Inf Sci. 2021;21: 149–164.
  46. 46. Wu C, Liu Y, Yang Y, Zhang P, Zhong W, Wang Y, et al. Analysis of therapeutic targets for SARS-CoV-2 and discovery of potential drugs by computational methods. Acta Pharm Sin B. 2020;10: 766–788. pmid:32292689
  47. 47. Fu Z, Huang B, Tang J, Liu S, Liu M, Ye Y, et al. The complex structure of GRL0617 and SARS-CoV-2 PLpro reveals a hot spot for antiviral drug discovery. Nat Commun. 2021;12: 488. pmid:33473130
  48. 48. Shin D, Mukherjee R, Grewe D, Bojkova D, Baek K, Bhattacharya A, et al. Papain-like protease regulates SARS-CoV-2 viral spread and innate immunity. Nature. 2020;587: 657–662. pmid:32726803
  49. 49. Khater I, Nassar A. In silico molecular docking analysis for repurposing approved antiviral drugs against SARS-CoV-2 main protease. Biochem Biophys Reports. 2021;27: 101032. pmid:34099985
  50. 50. Liu X, Zhang X, Lu Z, Zhu Y, Wang T. Potential molecular targets of nonstructural proteins for the development of antiviral drugs against SARS-CoV-2 infection. Biomed Pharmacother. 2021;133: 111035. pmid:33254013
  51. 51. Qiao J, Li Y-S, Zeng R, Liu F-L, Luo R-H, Huang C, et al. SARS-CoV-2 Mpro inhibitors with antiviral activity in a transgenic mouse model. Science (80-). 2021;371: 1374–1378. pmid:33602867
  52. 52. Dai W, Zhang B, Jiang X-M, Su H, Li J, Zhao Y, et al. Structure-based design of antiviral drug candidates targeting the SARS-CoV-2 main protease. Science (80-). 2020;368: 1331–1335. pmid:32321856
  53. 53. Bhowmick S, Saha A, Osman SM, Alasmary FA, Almutairi TM, Islam MA. Structure-based identification of SARS-CoV-2 main protease inhibitors from anti-viral specific chemical libraries: an exhaustive computational screening approach. Mol Divers. 2021;25: 1979–1997. pmid:33844135
  54. 54. Jin Z, Du X, Xu Y, Deng Y, Liu M, Zhao Y, et al. Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors. Nature. 2020;582: 289–293. pmid:32272481
  55. 55. Deshpande RR, Tiwari AP, Nyayanit N, Modak M. In silico molecular docking analysis for repurposing therapeutics against multiple proteins from SARS-CoV-2. Eur J Pharmacol. 2020;886: 173430. pmid:32758569
  56. 56. Cavasotto CN, Di Filippo JI. In silico Drug Repurposing for COVID-19: Targeting SARS-CoV-2 Proteins through Docking and Consensus Ranking. Mol Inform. 2021;40: e2000115. pmid:32722864
  57. 57. Kumar V, Liu H, Wu C. Drug repurposing against SARS-CoV-2 receptor binding domain using ensemble-based virtual screening and molecular dynamics simulations. Comput Biol Med. 2021;135: 104634. pmid:34256255
  58. 58. Prajapat M, Shekhar N, Sarma P, Avti P, Singh S, Kaur H, et al. Virtual screening and molecular dynamics study of approved drugs as inhibitors of spike protein S1 domain and ACE2 interaction in SARS-CoV-2. J Mol Graph Model. 2020;101: 107716. pmid:32866780
  59. 59. Braz HLB, Silveira JA de M, Marinho AD, de Moraes MEA, Moraes Filho MO de, Monteiro HSA, et al. In silico study of azithromycin, chloroquine and hydroxychloroquine and their potential mechanisms of action against SARS-CoV-2 infection. Int J Antimicrob Agents. 2020;56: 106119. pmid:32738306
  60. 60. Maffucci I, Contini A. In Silico Drug Repurposing for SARS-CoV-2 Main Proteinase and Spike Proteins. J Proteome Res. 2020;19: 4637–4648. pmid:32893632
  61. 61. Riva L, Yuan S, Yin X, Martin-Sancho L, Matsunaga N, Pache L, et al. Discovery of SARS-CoV-2 antiviral drugs through large-scale compound repurposing. Nature. 2020;586: 113–119. pmid:32707573
  62. 62. Jang WD, Jeon S, Kim S, Lee SY. Drugs repurposed for COVID-19 by virtual screening of 6,218 drugs and cell-based assay. Proc Natl Acad Sci. 2021;118. pmid:34234012
  63. 63. Trott AJ. Autodock vina: improving the speed and accuracy of docking. J Comput Chem. 2019;31: 455–461.
  64. 64. Wishart DS, Wang WW, Tang J, Wei F, Salentin S, Schreiber S, et al. Pneumonia of Unknown Etiology in Wuhan, China: Potential for International Spread Via Commercial Air Travel. Nucleic Acids Res. 2020;395: 441–447.
  65. 65. Hanwell MD, Curtis DE, Lonie DC, Vandermeersch T, Zurek E, Hutchison GR. Avogadro: an advanced semantic chemical editor, visualization, and analysis platform. 2012; 1–17.
  66. 66. Salentin S, Schreiber S, Haupt VJ, Adasme MF, Schroeder M. PLIP: Fully automated protein-ligand interaction profiler. Nucleic Acids Res. 2015;43: W443–W447. pmid:25873628
  67. 67. Eldridge MD, Murray CW, Auton TR, Paolini G V, Mee RP. Empirical scoring functions: I. The development of a fast empirical scoring function to estimate the binding affinity of ligands in receptor complexes. J Comput Aided Mol Des. 1997;11: 425–445. pmid:9385547
  68. 68. Reynolds CH, Bembenek SD, Tounge BA. The role of molecular size in ligand efficiency. Bioorg Med Chem Lett. 2007;17: 4258–4261. pmid:17532632
  69. 69. Reynolds CH, Tounge BA, Bembenek SD. Ligand Binding Efficiency: Trends, Physical Basis, and Implications. J Med Chem. 2008;51: 2432–2438. pmid:18380424
  70. 70. Kuntz ID, Chen K, Sharp KA, Kollman PA. The maximal affinity of ligands. Proc Natl Acad Sci. 1999;96: 9997–10002. pmid:10468550
  71. 71. Jo S, Kim T, Iyer VG, Im W. CHARMM-GUI: a web-based graphical user interface for CHARMM. Journal of computational chemistry. United States; 2008. pp. 1859–1865. pmid:18351591
  72. 72. Kim S, Lee J, Jo S, Brooks 3rd CL, Lee HS, Im W. CHARMM-GUI ligand reader and modeler for CHARMM force field generation of small molecules. J Comput Chem. 2017/05/11. 2017;38: 1879–1886. pmid:28497616
  73. 73. Jo S, Cheng X, Islam SM, Huang L, Rui H, Zhu A, et al. CHARMM-GUI PDB manipulator for advanced modeling and simulations of proteins containing nonstandard residues. Adv Protein Chem Struct Biol. 2014/08/24. 2014;96: 235–265. pmid:25443960
  74. 74. Pronk S, Páll S, Schulz R, Larsson P, Bjelkmar P, Apostolov R, et al. GROMACS 4.5: a high-throughput and highly parallel open source molecular simulation toolkit. Bioinformatics. 2013;29: 845–854. pmid:23407358
  75. 75. Huang J, MacKerell AD Jr. CHARMM36 all-atom additive protein force field: Validation based on comparison to NMR data. J Comput Chem. 2013;34: 2135–2145. pmid:23832629
  76. 76. Mark P, Nilsson L. Structure and Dynamics of the TIP3P, SPC, and SPC/E Water Models at 298 K. J Phys Chem A. 2001;105: 9954–9960.
  77. 77. Piche SW. Steepest descent algorithms for neural network controllers and filters. IEEE Trans neural networks. 1994;5: 198–212. pmid:18267791
  78. 78. Gapsys V, de Groot BL. On the importance of statistics in molecular simulations for thermodynamics, kinetics and simulation box size. Faraldo-Gómez JD, Grossfield A, editors. Elife. 2020;9: e57589. pmid:32812868
  79. 79. Ode H, Nakashima M, Kitamura S, Sugiura W, Sato H. Molecular dynamics simulation in virus research. Front Microbiol. 2012;3: 258. pmid:22833741
  80. 80. Knapp B, Frantal S, Cibena M, Schreiner W, Bauer P. Is an Intuitive Convergence Definition of Molecular Dynamics Simulations Solely Based on the Root Mean Square Deviation Possible? J Comput Biol. 2011;18: 997–1005. pmid:21702691
  81. 81. Hollingsworth SA, Dror RO. Molecular Dynamics Simulation for All. Neuron. 2018;99: 1129–1143. pmid:30236283
  82. 82. Reva BA, Finkelstein A V., Skolnick J. What is the probability of a chance prediction of a protein structure with an rmsd of 6 Å? Fold Des. 1998;3: 141–147. pmid:9565758