Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2476 word(s) 2476 2022-01-06 07:43:41 |
2 Format correct Meta information modification 2476 2022-01-10 02:29:41 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Garcia-Vallve, S. SARS-CoV-2 Main Protease Inhibitors. Encyclopedia. Available online: https://encyclopedia.pub/entry/17881 (accessed on 06 July 2024).
Garcia-Vallve S. SARS-CoV-2 Main Protease Inhibitors. Encyclopedia. Available at: https://encyclopedia.pub/entry/17881. Accessed July 06, 2024.
Garcia-Vallve, Santiago. "SARS-CoV-2 Main Protease Inhibitors" Encyclopedia, https://encyclopedia.pub/entry/17881 (accessed July 06, 2024).
Garcia-Vallve, S. (2022, January 07). SARS-CoV-2 Main Protease Inhibitors. In Encyclopedia. https://encyclopedia.pub/entry/17881
Garcia-Vallve, Santiago. "SARS-CoV-2 Main Protease Inhibitors." Encyclopedia. Web. 07 January, 2022.
SARS-CoV-2 Main Protease Inhibitors
Edit

The main protease (M-pro) or 3C-like protease of coronaviruses plays an essential role in virus replication. This protease contributes to the cleavage of the ppa1a and pp1ab polyproteins to produce several non-structural proteins, including M-pro itself. Since the beginning of the COVID-19 pandemic, the SARS-CoV-2 M-pro enzyme has been extensively studied, and its inhibitors are promising effective drugs for fighting against SARS-CoV-2. The first attempts to discover SARS-CoV-2 M-pro inhibitors used previously developed protease inhibitors or tried to repurpose drugs from other diseases. Covalent inhibitors form a covalent bond, usually with catalytic Cys145. Non-covalent inhibitors bind by non-covalent interactions at the active site of the enzyme, inhibiting its function.

COVID-19 M-pro inhibitors 3CL-pro inhibitors computational chemistry protease inhibitors virtual screening

1. Introduction

Since the onset of the COVID-19 pandemic, the scientific community has focused on studying the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) virus that causes the disease and on developing therapies and vaccines, several of which have been developed in record time. In the pharmacological field, no drugs have yet been definitively approved to inhibit the replication of SARS-CoV-2 and stop the development of COVID-19. Several targets are being studied, including the main protease (M-pro), which plays an essential role in virus replication [1]. This protease and the papain-like protease cleave the pp1a and pp1ab polyproteins to produce several nonstructural proteins, including M-pro itself, required for virus replication and transcription [1]. The high conservation of M-pro among related viruses, the importance of M-pro in the replication of the virus and the fact that M-pro only exists in coronaviruses and not in humans makes it an attractive target for the development of antiviral drugs [2][3][4]. SARS-CoV-2 M-pro has 306 amino acids that form three domains (I, II and III) [4]. The M-pro binding site is located between domains I and II, and domain III is involved in dimerization, which is essential for M-pro activity [4]. Similar to the M-pro from SARS-CoV-1 and other coronaviruses, SARS-CoV-2 M-pro has two catalytic amino acids, His41 and Cys145 (Figure 1). A catalytic water molecule is also important and makes a strong hydrogen bond with His41 [5]. Although some allosteric binding sites have been identified for the SARS-CoV-2 M-pro [6][7][8][9], most of the inhibitors crystallized within the M-pro bind to the active site [10]. One strategy used to find SARS-CoV-2 M-pro inhibitors, especially at the beginning of the pandemic, was drug repositioning [2][11][12][13]. This strategy is based on looking for drugs approved for one disease (therefore, its safety and possible adverse effects are known) that can be used to treat another—in this case, COVID-19. One of the most widely used computational tools for repositioning drugs, or looking for compounds with new activities, is what is known as protein-ligand docking. This tool predicts whether a particular molecule can bind (and, if it can, how) to a particular target (for example, the SARS-CoV-2 M-pro [14]). However, protein-ligand docking has several limitations, such as the consideration of the protein as a rigid body and the lack of confidence in the ability of scoring functions to give accurate binding energies [15][16]. In addition, the flexibility of the SARS-CoV-2 M-pro makes it a challenging target for small-molecule inhibitor design [17]. Using two different SARS-CoV-2 M-pro structures and five protein-ligand docking methods, we have recently shown that docking scores or the Gibbs free energy (∆G) calculated with an MM-GBSA method [18] do not correlate with bioactivity [19], probably because of the inability of common docking programs to correctly reproduce the binding modes of SARS-CoV-2 M-pro inhibitors [20]. This reinforces the idea that it is essential to validate the results obtained by protein-ligand docking or any other computational tool, especially when analyzing SARS-CoV-2 M-pro inhibitors [19][21][22][23]. The results of protein-ligand docking can be computationally validated by re-docking, cross-docking and applying the same protocol to a set of known active compounds and a set of decoy or inactive compounds [19]. Protein-ligand docking is expected to discriminate decoys from active compounds. If docking scores are used to rank the potency of a set of compounds, it must first be demonstrated that there is a correlation between docking scores and activity or potency, for example, expressed as IC50 values [19]. However, the best way to validate the predictions of protein-ligand docking is to experimentally test the predicted bioactivity of selected hits.
Ijms 23 00259 g001
Figure 1. SARS-CoV-2 M-pro structure. (A) Biological assembly of the M-pro in its dimeric form. The left protomer is shown in cartoon representation, colored by protein secondary structure, and the right protomer is displayed as a surface. (B) Detailed snapshot of the catalytic water, Cys145 and His41.
Since the beginning of the COVID-19 pandemic, developing SARS-CoV-2 M-pro inhibitors has been an active area of research. However, it did not have to start from scratch. Previous research about protease inhibitors, especially from SARS-CoV and MERS-CoV, proved to be useful [24][25]. Known inhibitors of proteases from HIV and Hepatitis C virus, in addition to calpain and caspase-3 inhibitors, were systematically analyzed to test their capacity to inhibit the SARS-CoV-2 M-pro [26]. Compounds developed against the M-pro of other coronaviruses were also tested, and some were found to be potent SARS-CoV-2 M-pro inhibitors [24][27][28]. The complete genome sequence of SARS-CoV-2 [29] and the first crystallized structure of the SARS-CoV-2 M-pro [4] were two important milestones in the development of new SARS-CoV-2 M-pro inhibitors. The article describing the first crystallized SARS-CoV-2 M-pro structure (the 6LU7 structure) also presented the first SARS-CoV-2 M-pro inhibitors [4]. These first inhibitors included the N3 compound, which had previously been developed as a protease inhibitor for multiple coronaviruses, including SARS-CoV and MERS-CoV, approved drugs (such as disulfiram and carmofur) and preclinical or clinical-trial drug candidates (ebselen, shikonin, tideglusib, PX-12 and TDZD-8) [4]. Since then, thousands of compounds have been suggested as SARS-CoV-2 M-pro inhibitors through computational methods such as protein-ligand docking, high-throughput screening experiments, computer-aided design and synthesis of new compounds. Several articles have reviewed the SARS-CoV-2 M-pro inhibitors discovered to date [25][28][30][31][32][33][34][35][36][37][38].

2. SARS-CoV-2 M-Pro Inhibitors

Table 1 shows the origin of the non-redundant set of 1765 SARS-CoV-2 M-pro inhibitors collected between January 2020 and August 2021 (see supplementary file S1). This set of inhibitors includes only those compounds whose inhibitory capacity, mainly expressed as the IC50 value, against M-pro from SARS-CoV-2 has been determined. A total of 758 compounds were extracted from peer-reviewed articles published between January 2020 and August 2021. When multiple IC50 values were found for the same compound, the mean value was calculated. From a set of 1037 M-pro inhibitors with an IC50 value downloaded from the COVID Moonshot project [39][40] on 1st October 2021, the compounds that had already been collected from the bibliographic search were discarded. In the end, 999 compounds were collected from COVID Moonshot. The IC50 values of these compounds were estimated as the mean value of the IC50 values from two biochemical assays: a fluorescence-based assay and a RapidFire Mass Spectrometry assay. Finally, 8 compounds were collected from the ChEMBL database [41], which contained more SARS-CoV-2 M-pro inhibitors, but most of them had already been collected from the bibliography. The SMILES of the 1765 SARS-CoV-2 M-pro inhibitors were standardized with the Standardizer 21.15.0 program from ChemAxon (http://www.chemaxon.com, accessed on 4 September 2021). The pIC50 values of the SARS-CoV-2 M-pro inhibitors collected range from 2.5 to 9.0 (Table 1). Putative covalent inhibitors were identified by the presence of typical covalent warheads (Table 2). When one of these warheads is in the appropriate position within the M-pro binding site, it can form a covalent bond, usually with the catalytic Cys145 [25]. There are twice as many non-covalent inhibitors as putative covalent inhibitors (Table 1), although pIC50 values are highest in some putative covalent inhibitors (Table 1 and Figure 2). However, conventional IC50 measurements are of limited value for characterizing the potency of irreversible covalent inhibitors, because incubation for different periods of time would provide different IC50 values [42]. Other parameters, such as molecular weight, LogP, number of hydrogen bond donors and hydrogen bond acceptors were similar between the covalent and non-covalent sets.
Ijms 23 00259 g002
Figure 2. Violin plots of the pIC50 values from 552 putative covalent and 1213 non-covalent SARS-CoV-2 M-pro inhibitors.
Table 1. Number of SARS-CoV-2 M-pro inhibitors collected.
SARS-CoV-2 M-Pro Inhibitor Set Number of Compounds (Covalent/Non-Covalent) 1 pIC50 Range pIC50 Range Covalent pIC50 Range Non-Covalent
From the bibliography 758 (346/412) 2.5–9.0 3.4–9.0 2.5–8.3
From COVID Moonshot 999 (205/794) 4.0–7.8 4.0–7.8 4.0–7.4
From ChEMBL 8 (1/7) 5.4–6.1 5.4 5.5–6.1
All 1765 (552/1213) 2.5–9.0 3.4–9.0 2.5–8.3
1 Putative covalent and non-covalent inhibitors were identified by the presence or absence of the covalent warheads shown in Table 2.
Table 2. Covalent warheads that can be used to identify putative covalent inhibitors. It shows the SMARTS that can be used to identify each warhead and some examples of SARS-CoV-2 inhibitors that contain each warhead. These covalent warheads were used to identify putative covalent inhibitors among the known SARS-CoV-2 M-pro inhibitors.
Warhead SMARTS Examples
Acrylamide [C;H2:1]=[C;H1]C(N)=O CVD-0004255
Chloroacetamide Cl[C;H2:1]C(N)=O BFC204
Vinylsulfonamide NS(=O)([C;H1]=[C;H2:1])=O  
Nitrile N#[C:1]-[*] Isavuconazole
Michael acceptors C=!@CC=[O,S] Cinanserin, MPI2, MPI9, N3
Alpha-ketoamide C(=O)(C=O)N Boceprevir, narlaprevir, telaprevir, UAWJ248
Aldehyde [CX3H1](=O) GC373, MI-05, MI-06, MI-09, MI-11, MI-13, MI-14, MI-21, MI-23, MI-28
Bisulfite adduct of aldehyde C(O)S(=[OX1])([O])(=[OX1]) GC376
Urea carbonyl [NX3][CX3](=[OX1])([NX3,nX3]) Carmofur
Bis(dialkylaminethiocarbonyl)disulfide [CX3](=[SX1])SS[CX3](=[SX1]) Disulfiram
Carbamoylsulfanyl [NX3,nX3][C,c](=[OX1])([SX2,sx2]) Tideglusib
Disulfide [SX2][SX2] PX-12
Hydroxymethyl ketone [CX3H0](=[OX1])[CH2][OH] PF-00835231
Alkoxymethyl ketone [CX3H0](=[OX1])[CH2][OX2H0] 2683066-41-1, 2683066-42-2, 2683066-47-7
Acyloxymethyl ketone [CX3H0](=[OX1])[CH2][OX2H0][CX3H0](=O) 2683066-41-1, 2683066-42-2, 2683066-47-7
Fluoro, Chloro-methyl ketone [CX3H0](=[OX1])[CH2][Cl,F] Z-AVLD-FMK
Ebselen related [Se]n(c=O) Ebselen
Figure 3 shows the t-Distributed Stochastic Neighbor Embedding (t-SNE) visualization of the chemical space of the set of SARS-CoV-2 M-pro inhibitors extracted from the bibliography. In this representation, more similar compounds are closer together. Peptidomimetic compounds, such as alpha-acyloxymethylketones, telaprevir, boceprevir, GC373 and their derivatives, which mimic natural peptide substrates, are closer together at the top left of the figure. Other clusters of compounds represent derivative compounds that have been synthesized from a lead compound to increase its bioactivity. Thus, derivatives from perampanel, ML300, ML188, ebsulfur, ebselen and myricetin form well-defined clusters. Perampanel derivatives are an example of a very successful increase in activity. Perampanel was first predicted as a SARS-CoV-2 M-pro inhibitor by consensus docking [2]. This prediction was confirmed by Jorgensen and coworkers, although perampanel showed only an approximate IC50 of 100–250 μM [43]. The same authors also optimized this compound and synthesized several derivative compounds [44][45][46]. Some of these perampanel derivatives have IC50 values in the low nanomolar range and are some of the most potent non-covalent SARS-CoV-2 M-pro inhibitors found to date.
Ijms 23 00259 g003
Figure 3. t-Distributed Stochastic Neighbor Embedding (t-SNE) visualization of the chemical space of a set of SARS-CoV-2 M-pro inhibitors extracted from the bibliography. Embedding is based on the 2048-bit Morgan fingerprint. Markers are colored according to several manually attributed chemotypes.
ML300 and ML188 are non-covalent inhibitors that were developed against the M-pro from SARS-CoV-1 [47][48]. Both compounds have been used to obtain more potent SARS-CoV-2 M-pro inhibitors that can inhibit SARS-CoV-2 replication in infected cells [49][50]. Boceprevir and telaprevir are approved protease inhibitors for treating hepatitis caused by the hepatitis C virus. Both compounds have been identified several times as covalent inhibitors of the SARS-CoV-2 M-pro [43][51][52][53][54][55][56][57][58]. New bicycloproline derivatives have been designed and synthesized from them both [59]. All compounds inhibited SARS-CoV-2 M-pro in vitro, with IC50 values ranging from 7.6 to 748.5 nM [59]. In addition, two of them, MI-09 and MI-30, showed excellent antiviral activity in a cell-based assay and significantly reduced lung viral loads and lesions in a transgenic mouse model of SARS-CoV-2 infection [59]. GC376 is a covalent M-pro inhibitor that was developed as an inhibitor of the main protease of the feline coronavirus FCoV [60] that also showed activity against the M-pro from MERS and SARS-CoV viruses [61]. Its IC50 activity against SARS-CoV-2 M-pro ranges between 0.026 and 0.89 μM [51][52][61][62][63][64][65][66][67]. GC376 is a prodrug, and its bisulphite adduct is converted to an aldehyde to form GC373. This aldehyde forms a covalent interaction with the catalytic Cys145 of the SARS-CoV-2 M-pro [61]. Several GC373 and GC376 derivative compounds have been designed and assayed [63][68][69]. Some of them, such as UAWJ248 [70], are more potent than GC376. A group of peptidomimetic compounds with an alpha-acyloxymethyl ketone warhead designed to form an irreversible covalent bond with Cys145 showed IC50 values against the SARS-CoV-2 M-pro in the nM range [71]. They also inhibited SARS-CoV-2 replication and presented low cytotoxicity and good stability [71]. Ebselen is a covalent inhibitor of the SARS-CoV-2 M-pro, although its specificity has been questioned [72][73]. Several derivative compounds of ebselen and its sulfur derivative ebsulfur have been analyzed [74][75]. Some of the derivative compounds displayed more potent M-pro inhibition than ebselen and ebsulfur [74][75]. However, the promiscuous behavior of ebselen and ebsulfur and their lack of cellular antiviral activity [74][75] may also be applied to their derivatives. Myricetin is a flavonoid that acts as a non-peptidomimetic and covalent inhibitor of SARS-CoV-2 [76][77]. Its covalent behavior was unexpected and caused by the pyrogallol moiety that formed a covalent bond with Cys145 [76]. Myricetin and its derivatives inhibit SARS-CoV-2 M-pro and SARS-CoV-2 replication in cells [76][77][78][79], and form a cluster at the bottom of Figure 3, near quercetin and other flavonoids.
Several in vitro and in cellulo (using live cells) methods have been developed to measure the inhibitory potency of a compound against the SARS-CoV-2 M-pro. In vitro methods need to express and purify SARS-CoV-2 M-pro, so some tags are sometimes added. However, especially if they are located at the N-terminus, these tags can interfere with the binding of M-pro to its ligands. The activity values obtained by different laboratories or with different methods or conditions must be compared with great care. The presence of DTT has been reported to affect the inhibitory activity of covalent M-pro inhibitors. If the inhibitory effect of an M-pro inhibitor is eliminated or greatly reduced by the presence of DTT, the inhibition is not specific. Therefore, the potency of inhibition measured in the absence of DTT should not be used by itself. The potency of a compound to inhibit SARS-CoV-2 replication in cells cannot always be inferred from the potency to inhibit M-pro, determined in vitro. An antiviral assay that uses cells infected with SARS-CoV-2 provides a better estimate of the potency of a compound to inhibit virus replication. However, if it is to be ruled out that the toxicity of the compounds is responsible for the antiviral activity, the cytotoxicity of the compounds needs to be determined.

3. Conclusions

Although we have not yet hit the bullseye and no drug has yet been approved to inhibit SARS-CoV-2 M-pro, we may be close. Improving derivatives of a leading compound has proven to be a very successful strategy for finding potent SARS-CoV-2 M-pro inhibitors. Some derivative compounds designed in less than two years since the start of the COVID-19 pandemic represent an important step toward the development of new anti–SARS-CoV-2 drugs. Currently, there are several compounds with low nanomolar IC50 values against SARS-CoV-2 M-pro and high anti-SARS-CoV-2 efficacy in cell models, with values comparable to those of the FDA-approved RNA polymerase inhibitor remdesivir.
 

References

  1. Ziebuhr, J.; Herold, J.; Siddell, S.G. Characterization of a human coronavirus (strain 229E) 3C-like proteinase activity. J. Virol. 1995, 69, 4331–4338.
  2. Gimeno, A.; Mestres-Truyol, J.; Ojeda-Montes, M.J.; Macip, G.; Saldivar-Espinoza, B.; Cereto-Massagué, A.; Pujadas, G.; Garcia-Vallvé, S. Prediction of Novel Inhibitors of the Main Protease (M-pro) of SARS-CoV-2 through Consensus Docking and Drug Reposition. Int. J. Mol. Sci. 2020, 21, 3793.
  3. Anand, K.; Ziebuhr, J.; Wadhwani, P.; Mesters, J.R.; Hilgenfeld, R. Coronavirus main proteinase (3CLpro) structure: Basis for design of anti-SARS drugs. Science 2003, 300, 1763–1767.
  4. Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C.; et al. Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors. Nature 2020, 582, 289–293.
  5. Kneller, D.W.; Phillips, G.; O’Neill, H.M.; Jedrzejczak, R.; Stols, L.; Langan, P.; Joachimiak, A.; Coates, L.; Kovalevsky, A. Structural plasticity of SARS-CoV-2 3CL Mpro active site cavity revealed by room temperature X-ray crystallography. Nat. Commun. 2020, 11, 3202.
  6. El-Baba, T.J.; Lutomski, C.A.; Kantsadi, A.L.; Malla, T.R.; John, T.; Mikhailov, V.; Bolla, J.R.; Schofield, C.J.; Zitzmann, N.; Vakonakis, I.; et al. Allosteric Inhibition of the SARS-CoV-2 Main Protease: Insights from Mass Spectrometry Based Assays. Angew. Chem. Int. Ed. Engl. 2020, 59, 23544–23548.
  7. Günther, S.; Reinke, P.Y.A.; Fernández-García, Y.; Lieske, J.; Lane, T.J.; Ginn, H.M.; Koua, F.H.M.; Ehrt, C.; Ewert, W.; Oberthuer, D.; et al. X-ray screening identifies active site and allosteric inhibitors of SARS-CoV-2 main protease. Science 2021, 372, 642–646.
  8. Du, R.; Cooper, L.; Chen, Z.; Lee, H.; Rong, L.; Cui, Q. Discovery of chebulagic acid and punicalagin as novel allosteric inhibitors of SARS-CoV-2 3CLpro. Antivir. Res. 2021, 190, 105075.
  9. Eberle, R.J.; Olivier, D.S.; Amaral, M.S.; Gering, I.; Willbold, D.; Arni, R.K.; Coronado, M.A. The Repurposed Drugs Suramin and Quinacrine Cooperatively Inhibit SARS-CoV-2 3CLpro In Vitro. Viruses 2021, 13, 873.
  10. Nguyen, D.D.; Gao, K.; Chen, J.; Wang, R.; Wei, G.-W. Unveiling the molecular mechanism of SARS-CoV-2 main protease inhibition from 137 crystal structures using algebraic topology and deep learning. Chem. Sci. 2020, 11, 12036–12046.
  11. Liu, Z.; Fang, H.; Reagan, K.; Xu, X.; Mendrick, D.L.; Slikker, W.; Tong, W. In silico drug repositioning: What we need to know. Drug Discov. Today 2013, 18, 110–115.
  12. Chakraborti, S.; Bheemireddy, S.; Srinivasan, N. Repurposing drugs against the main protease of SARS-CoV-2: Mechanism-based insights supported by available laboratory and clinical data. Mol. Omics 2020, 16, 474–491.
  13. Wang, X.; Guan, Y. COVID-19 drug repurposing: A review of computational screening methods, clinical trials, and protein interaction assays. Med. Res. Rev. 2021, 41, 5–28.
  14. Gimeno, A.; Ojeda-Montes, M.; Tomás-Hernández, S.; Cereto-Massagué, A.; Beltrán-Debón, R.; Mulero, M.; Pujadas, G.; Garcia-Vallvé, S. The Light and Dark Sides of Virtual Screening: What Is There to Know? Int. J. Mol. Sci. 2019, 20, 1375.
  15. Warren, G.L.; Andrews, C.W.; Capelli, A.-M.; Clarke, B.; LaLonde, J.; Lambert, M.H.; Lindvall, M.; Nevins, N.; Semus, S.F.; Senger, S.; et al. A Critical Assessment of Docking Programs and Scoring Functions. J. Med. Chem. 2006, 49, 5912–5931.
  16. Pantsar, T.; Poso, A. Binding Affinity via Docking: Fact and Fiction. Molecules 2018, 23, 1899.
  17. Bzówka, M.; Mitusińska, K.; Raczyńska, A.; Samol, A.; Tuszyński, J.A.; Góra, A. Structural and Evolutionary Analysis Indicate That the SARS-CoV-2 Mpro Is a Challenging Target for Small-Molecule Inhibitor Design. Int. J. Mol. Sci. 2020, 21, 3099.
  18. Genheden, S.; Ryde, U. The MM/PBSA and MM/GBSA methods to estimate ligand-binding affinities. Expert Opin. Drug Discov. 2015, 10, 449–461.
  19. Macip, G.; Garcia-Segura, P.; Mestres-Truyol, J.; Saldivar-Espinoza, B.; Ojeda-Montes, M.J.; Gimeno, A.; Cereto-Massagué, A.; Garcia-Vallvé, S.; Pujadas, G. Haste makes waste: A critical review of docking-based virtual screening in drug repurposing for SARS-CoV-2 main protease (M-pro) inhibition. Med. Res. Rev. 2021.
  20. Zev, S.; Raz, K.; Schwartz, R.; Tarabeh, R.; Gupta, P.K.; Major, D.T. Benchmarking the Ability of Common Docking Programs to Correctly Reproduce and Score Binding Modes in SARS-CoV-2 Protease Mpro. J. Chem. Inf. Model. 2021, 61, 2957–2966.
  21. Dotolo, S.; Marabotti, A.; Facchiano, A.; Tagliaferri, R. A review on drug repurposing applicable to COVID-19. Brief. Bioinform. 2021, 22, 726–741.
  22. Bellera, C.L.; Llanos, M.; Gantner, M.E.; Rodriguez, S.; Gavernet, L.; Comini, M.; Talevi, A. Can drug repurposing strategies be the solution to the COVID-19 crisis? Expert Opin. Drug Discov. 2021, 16, 605–612.
  23. Llanos, M.A.; Gantner, M.E.; Rodriguez, S.; Alberca, L.N.; Bellera, C.L.; Talevi, A.; Gavernet, L. Strengths and Weaknesses of Docking Simulations in the SARS-CoV-2 Era: The Main Protease (Mpro) Case Study. J. Chem. Inf. Model. 2021, 61, 3758–3770.
  24. Amin, S.A.; Banerjee, S.; Gayen, S.; Jha, T. Protease targeted COVID-19 drug discovery: What we have learned from the past SARS-CoV inhibitors? Eur. J. Med. Chem. 2021, 215, 113294.
  25. Yang, H.; Yang, J. A review of the latest research on Mpro targeting SARS-COV inhibitors. RSC Med. Chem. 2021, 12, 1026–1036.
  26. Cui, W.; Yang, K.; Yang, H. Recent Progress in the Drug Development Targeting SARS-CoV-2 Main Protease as Treatment for COVID-19. Front. Mol. Biosci. 2020, 7, 616341.
  27. Pillaiyar, T.; Wendt, L.L.; Manickam, M.; Easwaran, M. The recent outbreaks of human coronaviruses: A medicinal chemistry perspective. Med. Res. Rev. 2021, 41, 72–135.
  28. Banerjee, R.; Perera, L.; Tillekeratne, L.M.V. Potential SARS-CoV-2 main protease inhibitors. Drug Discov. Today 2021, 26, 804–816.
  29. Wu, F.; Zhao, S.; Yu, B.; Chen, Y.-M.; Wang, W.; Song, Z.-G.; Hu, Y.; Tao, Z.-W.; Tian, J.-H.; Pei, Y.-Y.; et al. A new coronavirus associated with human respiratory disease in China. Nature 2020, 579, 265–269.
  30. Gao, K.; Wang, R.; Chen, J.; Tepe, J.J.; Huang, F.; Wei, G.-W. Perspectives on SARS-CoV-2 Main Protease Inhibitors. J. Med. Chem. 2021, 64, 16922–16955.
  31. Xu, J.; Xue, Y.; Zhou, R.; Shi, P.; Li, H.; Zhou, J. Drug repurposing approach to combating coronavirus: Potential drugs and drug targets. Med. Res. Rev. 2021, 41, 1375–1426.
  32. Cannalire, R.; Cerchia, C.; Beccari, A.R.; Di Leva, F.S.; Summa, V. Targeting SARS-CoV-2 Proteases and Polymerase for COVID-19 Treatment: State of the Art and Future Opportunities. J. Med. Chem. 2020.
  33. Chen, C.; Yu, X.; Kuo, C.; Min, J.; Chen, S.; Ma, L.; Liu, K.; Guo, R. Overview of antiviral drug candidates targeting coronaviral 3C-like main proteases. FEBS J. 2021, 288, 5089–5121.
  34. Xiong, M.; Su, H.; Zhao, W.; Xie, H.; Shao, Q.; Xu, Y. What coronavirus 3C-like protease tells us: From structure, substrate selectivity, to inhibitor design. Med. Res. Rev. 2021, 41, 1965–1998.
  35. Mengist, H.M.; Dilnessa, T.; Jin, T. Structural Basis of Potential Inhibitors Targeting SARS-CoV-2 Main Protease. Front. Chem. 2021, 9, 622898.
  36. Citarella, A.; Scala, A.; Piperno, A.; Micale, N. SARS-CoV-2 Mpro: A Potential Target for Peptidomimetics and Small-Molecule Inhibitors. Biomolecules 2021, 11, 607.
  37. Chia, C.S.B.; Xu, W.; Shuyi Ng, P. A Patent Review on SARS Coronavirus Main Protease (3CL pro) Inhibitors. ChemMedChem 2021.
  38. Vandyck, K.; Deval, J. Considerations for the discovery and development of 3-chymotrypsin-like cysteine protease inhibitors targeting SARS-CoV-2 infection. Curr. Opin. Virol. 2021, 49, 36–40.
  39. COVID Moonshot. Available online: https://covid.postera.ai/covid (accessed on 1 October 2021).
  40. The COVID Moonshot Consortium; Chodera, J.; Lee, A.; London, N.; von Delft, F. Open Science Discovery of Oral Non-Covalent SARS-CoV-2 Main Protease Inhibitors. ChemRxiv 2021.
  41. Mendez, D.; Gaulton, A.; Bento, A.P.; Chambers, J.; De Veij, M.; Félix, E.; Magariños, M.P.; Mosquera, J.F.; Mutowo, P.; Nowotka, M.; et al. ChEMBL: Towards direct deposition of bioassay data. Nucleic Acids Res. 2019, 47, D930–D940.
  42. Singh, J.; Petter, R.C.; Baillie, T.A.; Whitty, A. The resurgence of covalent drugs. Nat. Rev. Drug Discov. 2011, 10, 307–317.
  43. Ghahremanpour, M.M.; Tirado-Rives, J.; Deshmukh, M.; Ippolito, J.A.; Zhang, C.-H.; Cabeza de Vaca, I.; Liosi, M.-E.; Anderson, K.S.; Jorgensen, W.L. Identification of 14 Known Drugs as Inhibitors of the Main Protease of SARS-CoV-2. ACS Med. Chem. Lett. 2020, 11, 2526–2533.
  44. Zhang, C.-H.; Stone, E.A.; Deshmukh, M.; Ippolito, J.A.; Ghahremanpour, M.M.; Tirado-Rives, J.; Spasov, K.A.; Zhang, S.; Takeo, Y.; Kudalkar, S.N.; et al. Potent Noncovalent Inhibitors of the Main Protease of SARS-CoV-2 from Molecular Sculpting of the Drug Perampanel Guided by Free Energy Perturbation Calculations. ACS Cent. Sci. 2021, 7, 467–475.
  45. Deshmukh, M.G.; Ippolito, J.A.; Zhang, C.-H.; Stone, E.A.; Reilly, R.A.; Miller, S.J.; Jorgensen, W.L.; Anderson, K.S. Structure-guided design of a perampanel-derived pharmacophore targeting the SARS-CoV-2 main protease. Structure 2021, 29, 823–833.e5.
  46. Zhang, C.-H.; Spasov, K.A.; Reilly, R.A.; Hollander, K.; Stone, E.A.; Ippolito, J.A.; Liosi, M.-E.; Deshmukh, M.G.; Tirado-Rives, J.; Zhang, S.; et al. Optimization of Triarylpyridinone Inhibitors of the Main Protease of SARS-CoV-2 to Low-Nanomolar Antiviral Potency. ACS Med. Chem. Lett. 2021, 12, 1325–1332.
  47. Jacobs, J.; Grum-Tokars, V.; Zhou, Y.; Turlington, M.; Saldanha, S.A.; Chase, P.; Eggler, A.; Dawson, E.S.; Baez-Santos, Y.M.; Tomar, S.; et al. Discovery, synthesis, and structure-based optimization of a series of N-(tert-butyl)-2-(N-arylamido)-2-(pyridin-3-yl) acetamides (ML188) as potent noncovalent small molecule inhibitors of the severe acute respiratory syndrome coronavirus (SARS-CoV) 3CL pr. J. Med. Chem. 2013, 56, 534–546.
  48. Turlington, M.; Chun, A.; Tomar, S.; Eggler, A.; Grum-Tokars, V.; Jacobs, J.; Daniels, J.S.; Dawson, E.; Saldanha, A.; Chase, P.; et al. Discovery of N-(benzotriazol-1-yl)-N-(benzyl)acetamido)phenyl) carboxamides as severe acute respiratory syndrome coronavirus (SARS-CoV) 3CLpro inhibitors: Identification of ML300 and noncovalent nanomolar inhibitors with an induced-fit binding. Bioorg. Med. Chem. Lett. 2013, 23, 6172–6177.
  49. Han, S.H.; Goins, C.M.; Arya, T.; Shin, W.-J.; Maw, J.; Hooper, A.; Sonawane, D.P.; Porter, M.R.; Bannister, B.E.; Crouch, R.D.; et al. Structure-Based Optimization of ML300-Derived, Noncovalent Inhibitors Targeting the Severe Acute Respiratory Syndrome Coronavirus 3CL Protease (SARS-CoV-2 3CL pro). J. Med. Chem. 2021.
  50. Kitamura, N.; Sacco, M.D.; Ma, C.; Hu, Y.; Townsend, J.A.; Meng, X.; Zhang, F.; Zhang, X.; Ba, M.; Szeto, T.; et al. Expedited Approach toward the Rational Design of Noncovalent SARS-CoV-2 Main Protease Inhibitors. J. Med. Chem. 2021.
  51. Ma, C.; Sacco, M.D.; Hurst, B.; Townsend, J.A.; Hu, Y.; Szeto, T.; Zhang, X.; Tarbet, B.; Marty, M.T.; Chen, Y.; et al. Boceprevir, GC-376, and calpain inhibitors II, XII inhibit SARS-CoV-2 viral replication by targeting the viral main protease. Cell Res. 2020, 30, 678–692.
  52. Fu, L.; Ye, F.; Feng, Y.; Yu, F.; Wang, Q.; Wu, Y.; Zhao, C.; Sun, H.; Huang, B.; Niu, P.; et al. Both Boceprevir and GC376 efficaciously inhibit SARS-CoV-2 by targeting its main protease. Nat. Commun. 2020, 11, 4417.
  53. Kneller, D.W.; Galanie, S.; Phillips, G.; O’Neill, H.M.; Coates, L.; Kovalevsky, A. Malleability of the SARS-CoV-2 3CL Mpro Active-Site Cavity Facilitates Binding of Clinical Antivirals. Structure 2020, 28, 1313–1320.e3.
  54. Pathak, N.; Chen, Y.-T.; Hsu, Y.-C.; Hsu, N.-Y.; Kuo, C.-J.; Tsai, H.P.; Kang, J.-J.; Huang, C.-H.; Chang, S.-Y.; Chang, Y.-H.; et al. Uncovering Flexible Active Site Conformations of SARS-CoV-2 3CL Proteases through Protease Pharmacophore Clusters and COVID-19 Drug Repurposing. ACS Nano 2021, 15, 857–872.
  55. Jan, J.-T.; Cheng, T.-J.R.; Juang, Y.-P.; Ma, H.-H.; Wu, Y.-T.; Yang, W.-B.; Cheng, C.-W.; Chen, X.; Chou, T.-H.; Shie, J.-J.; et al. Identification of existing pharmaceuticals and herbal medicines as inhibitors of SARS-CoV-2 infection. Proc. Natl. Acad. Sci. USA 2021, 118, e2021579118.
  56. Mody, V.; Ho, J.; Wills, S.; Mawri, A.; Lawson, L.; Ebert, M.C.C.J.C.; Fortin, G.M.; Rayalam, S.; Taval, S. Identification of 3-chymotrypsin like protease (3CLPro) inhibitors as potential anti-SARS-CoV-2 agents. Commun. Biol. 2021, 4, 93.
  57. Baker, J.D.; Uhrich, R.L.; Kraemer, G.C.; Love, J.E.; Kraemer, B.C. A drug repurposing screen identifies hepatitis C antivirals as inhibitors of the SARS-CoV2 main protease. PLoS ONE 2021, 16, e0245962.
  58. Manandhar, A.; Blass, B.E.; Colussi, D.J.; Almi, I.; Abou-Gharbia, M.; Klein, M.L.; Elokely, K.M. Targeting SARS-CoV-2 M3CLpro by HCV NS3/4a Inhibitors: In Silico Modeling and In Vitro Screening. J. Chem. Inf. Model. 2021, 61, 1020–1032.
  59. Qiao, J.; Li, Y.-S.; Zeng, R.; Liu, F.-L.; Luo, R.-H.; Huang, C.; Wang, Y.-F.; Zhang, J.; Quan, B.; Shen, C.; et al. SARS-CoV-2 Mpro inhibitors with antiviral activity in a transgenic mouse model. Science 2021, 371, 1374–1378.
  60. Kim, Y.; Shivanna, V.; Narayanan, S.; Prior, A.M.; Weerasekara, S.; Hua, D.H.; Kankanamalage, A.C.G.; Groutas, W.C.; Chang, K.-O. Broad-spectrum inhibitors against 3C-like proteases of feline coronaviruses and feline caliciviruses. J. Virol. 2015, 89, 4942–4950.
  61. Vuong, W.; Khan, M.B.; Fischer, C.; Arutyunova, E.; Lamer, T.; Shields, J.; Saffran, H.A.; McKay, R.T.; van Belkum, M.J.; Joyce, M.A.; et al. Feline coronavirus drug inhibits the main protease of SARS-CoV-2 and blocks virus replication. Nat. Commun. 2020, 11, 4282.
  62. Hung, H.; Ke, Y.; Huang, S.Y.; Huang, P.-N.; Kung, Y.; Chang, T.-Y.; Yen, K.; Peng, T.-T.; Chang, S.-E.; Huang, C.-T.; et al. Discovery of M Protease Inhibitors Encoded by SARS-CoV-2. Antimicrob. Agents Chemother. 2020, 64, e00872-20.
  63. Rathnayake, A.D.; Zheng, J.; Kim, Y.; Perera, K.D.; Mackin, S.; Meyerholz, D.K.; Kashipathy, M.M.; Battaile, K.P.; Lovell, S.; Perlman, S.; et al. 3C-like protease inhibitors block coronavirus replication in vitro and improve survival in MERS-CoV–infected mice. Sci. Transl. Med. 2020, 12, eabc5332.
  64. Gurard-Levin, Z.A.; Liu, C.; Jekle, A.; Jaisinghani, R.; Ren, S.; Vandyck, K.; Jochmans, D.; Leyssen, P.; Neyts, J.; Blatt, L.M.; et al. Evaluation of SARS-CoV-2 3C-like protease inhibitors using self-assembled monolayer desorption ionization mass spectrometry. Antivir. Res. 2020, 182, 104924.
  65. Wang, Y.-C.; Yang, W.-H.; Yang, C.-S.; Hou, M.-H.; Tsai, C.-L.; Chou, Y.-Z.; Hung, M.-C.; Chen, Y. Structural basis of SARS-CoV-2 main protease inhibition by a broad-spectrum anti-coronaviral drug. Am. J. Cancer Res. 2020, 10, 2535–2545.
  66. Zhu, W.; Xu, M.; Chen, C.Z.; Guo, H.; Shen, M.; Hu, X.; Shinn, P.; Klumpp-Thomas, C.; Michael, S.G.; Zheng, W. Identification of SARS-CoV-2 3CL Protease Inhibitors by a Quantitative High-Throughput Screening. ACS Pharmacol. Transl. Sci. 2020, 3, 1008–1016.
  67. Iketani, S.; Forouhar, F.; Liu, H.; Hong, S.J.; Lin, F.-Y.; Nair, M.S.; Zask, A.; Huang, Y.; Xing, L.; Stockwell, B.R.; et al. Lead compounds for the development of SARS-CoV-2 3CL protease inhibitors. Nat. Commun. 2021, 12, 2016.
  68. Yang, K.S.; Ma, X.R.; Ma, Y.; Alugubelli, Y.R.; Scott, D.A.; Vatansever, E.C.; Drelich, A.K.; Sankaran, B.; Geng, Z.Z.; Blankenship, L.R.; et al. A Quick Route to Multiple Highly Potent SARS-CoV-2 Main Protease Inhibitors. ChemMedChem 2021, 16, 942–948.
  69. Vuong, W.; Fischer, C.; Khan, M.B.; van Belkum, M.J.; Lamer, T.; Willoughby, K.D.; Lu, J.; Arutyunova, E.; Joyce, M.A.; Saffran, H.A.; et al. Improved SARS-CoV-2 Mpro inhibitors based on feline antiviral drug GC376: Structural enhancements, increased solubility, and micellar studies. Eur. J. Med. Chem. 2021, 222, 113584.
  70. Sacco, M.D.; Ma, C.; Lagarias, P.; Gao, A.; Townsend, J.A.; Meng, X.; Dube, P.; Zhang, X.; Hu, Y.; Kitamura, N.; et al. Structure and inhibition of the SARS-CoV-2 main protease reveal strategy for developing dual inhibitors against Mpro and cathepsin L. Sci. Adv. 2020, 6, eabe0751.
  71. Bai, B.; Belovodskiy, A.; Hena, M.; Kandadai, A.S.; Joyce, M.A.; Saffran, H.A.; Shields, J.A.; Khan, M.B.; Arutyunova, E.; Lu, J.; et al. Peptidomimetic α-Acyloxymethylketone Warheads with Six-Membered Lactam P1 Glutamine Mimic: SARS-CoV-2 3CL Protease Inhibition, Coronavirus Antiviral Activity, and in Vitro Biological Stability. J. Med. Chem. 2021.
  72. Ma, C.; Hu, Y.; Townsend, J.A.; Lagarias, P.I.; Marty, M.T.; Kolocouris, A.; Wang, J. Ebselen, Disulfiram, Carmofur, PX-12, Tideglusib, and Shikonin Are Nonspecific Promiscuous SARS-CoV-2 Main Protease Inhibitors. ACS Pharmacol. Transl. Sci. 2020, 3, 1265–1277.
  73. Ma, C.; Tan, H.; Choza, J.; Wang, Y.; Wang, J. Validation and invalidation of SARS-CoV-2 main protease inhibitors using the Flip-GFP and Protease-Glo luciferase assays. Acta Pharm. Sin. B 2021.
  74. Sun, L.-Y.; Chen, C.; Su, J.; Li, J.-Q.; Jiang, Z.; Gao, H.; Chigan, J.-Z.; Ding, H.-H.; Zhai, L.; Yang, K.-W. Ebsulfur and Ebselen as highly potent scaffolds for the development of potential SARS-CoV-2 antivirals. Bioorg. Chem. 2021, 112, 104889.
  75. Amporndanai, K.; Meng, X.; Shang, W.; Jin, Z.; Rogers, M.; Zhao, Y.; Rao, Z.; Liu, Z.-J.; Yang, H.; Zhang, L.; et al. Inhibition mechanism of SARS-CoV-2 main protease by ebselen and its derivatives. Nat. Commun. 2021, 12, 3061.
  76. Su, H.; Yao, S.; Zhao, W.; Zhang, Y.; Liu, J.; Shao, Q.; Wang, Q.; Li, M.; Xie, H.; Shang, W.; et al. Identification of pyrogallol as a warhead in design of covalent inhibitors for the SARS-CoV-2 3CL protease. Nat. Commun. 2021, 12, 3623.
  77. Kuzikov, M.; Costanzi, E.; Reinshagen, J.; Esposito, F.; Vangeel, L.; Wolf, M.; Ellinger, B.; Claussen, C.; Geisslinger, G.; Corona, A.; et al. Identification of Inhibitors of SARS-CoV-2 3CL-Pro Enzymatic Activity Using a Small Molecule in Vitro Repurposing Screen. ACS Pharmacol. Transl. Sci. 2021, 4, 1096–1110.
  78. Liu, H.; Ye, F.; Sun, Q.; Liang, H.; Li, C.; Li, S.; Lu, R.; Huang, B.; Tan, W.; Lai, L. Scutellaria baicalensis extract and baicalein inhibit replication of SARS-CoV-2 and its 3C-like protease in vitro. J. Enzyme Inhib. Med. Chem. 2021, 36, 497–503.
  79. Nguyen, T.T.H.; Jung, J.-H.; Kim, M.-K.; Lim, S.; Choi, J.-M.; Chung, B.; Kim, D.-W.; Kim, D. The Inhibitory Effects of Plant Derivate Polyphenols on the Main Protease of SARS Coronavirus 2 and Their Structure–Activity Relationship. Molecules 2021, 26, 1924.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 534
Entry Collection: COVID-19
Revisions: 2 times (View History)
Update Date: 10 Jan 2022
1000/1000
Video Production Service