Next Article in Journal
PPI Modulators of E6 as Potential Targeted Therapeutics for Cervical Cancer: Progress and Challenges in Targeting E6
Next Article in Special Issue
Pharmacophore-Based Virtual Screening of Novel Competitive Inhibitors of the Neurodegenerative Disease Target Kynurenine-3-Monooxygenase
Previous Article in Journal
New O-Aryl-Carbamoyl-Oxymino-Fluorene Derivatives with MI-Crobicidal and Antibiofilm Activity Enhanced by Combination with Iron Oxide Nanoparticles
Previous Article in Special Issue
Discovery of Monoacylglycerol Lipase (MAGL) Inhibitors Based on a Pharmacophore-Guided Virtual Screening Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Prioritisation of Compounds for 3CLpro Inhibitor Development on SARS-CoV-2 Variants

1
Laboratory of Physical Chemistry and Chemical Thermodynamics, Faculty of Chemistry and Chemical Engineering, University of Maribor, Smetanova ulica 17, SI-2000 Maribor, Slovenia
2
Faculty of Mathematics, Natural Sciences and Information Technologies, University of Primorska, Glagoljaška 8, SI-6000 Koper, Slovenia
3
Institute Jožef Stefan, Jamova cesta 39, SI-1000 Ljubljana, Slovenia
4
Independent Researcher, Cesta Cirila Kosmača 66, SI-1000 Ljubljana, Slovenia
5
Erudio, Litostrojska Cesta 40, SI-1000 Ljubljana, Slovenia
6
Faculty of Chemistry and Chemical Technology, University of Ljubljana, Večna pot 113, SI-1000 Ljubljana, Slovenia
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(10), 3003; https://doi.org/10.3390/molecules26103003
Submission received: 23 March 2021 / Revised: 12 May 2021 / Accepted: 14 May 2021 / Published: 18 May 2021
(This article belongs to the Special Issue Virtual Screening)

Abstract

:
COVID-19 represents a new potentially life-threatening illness caused by severe acute respiratory syndrome coronavirus 2 or SARS-CoV-2 pathogen. In 2021, new variants of the virus with multiple key mutations have emerged, such as B.1.1.7, B.1.351, P.1 and B.1.617, and are threatening to render available vaccines or potential drugs ineffective. In this regard, we highlight 3CLpro, the main viral protease, as a valuable therapeutic target that possesses no mutations in the described pandemically relevant variants. 3CLpro could therefore provide trans-variant effectiveness that is supported by structural studies and possesses readily available biological evaluation experiments. With this in mind, we performed a high throughput virtual screening experiment using CmDock and the “In-Stock” chemical library to prepare prioritisation lists of compounds for further studies. We coupled the virtual screening experiment to a machine learning-supported classification and activity regression study to bring maximal enrichment and available structural data on known 3CLpro inhibitors to the prepared focused libraries. All virtual screening hits are classified according to 3CLpro inhibitor, viral cysteine protease or remaining chemical space based on the calculated set of 208 chemical descriptors. Last but not least, we analysed if the current set of 3CLpro inhibitors could be used in activity prediction and observed that the field of 3CLpro inhibitors is drastically under-represented compared to the chemical space of viral cysteine protease inhibitors. We postulate that this methodology of 3CLpro inhibitor library preparation and compound prioritisation far surpass the selection of compounds from available commercial “corona focused libraries”.

Graphical Abstract

1. Introduction

Coronavirus disease (COVID-19) is an infectious disease caused by a novel severe acute respiratory syndrome coronavirus 2 or SARS-CoV-2. COVID-19 was initially reported in Wuhan province in China and was declared as a global pandemic [1]. COVID-19 is a severe illness similar to flu, with major symptoms being cough, fever and breathing difficulty. Furthermore, the illness can cause systemic inflammation [2,3]. The pathogen SARS-CoV-2 belongs to the Coronaviridae family, an enveloped positive-sense single-stranded (+ssRNA) RNA virus, and is closely related to the previously described SARS-CoV and MERS-CoV coronaviruses [4]. The SARS-CoV-2 genome shares 82% sequence identity with SARS-CoV and 90% identity with MERS-CoV and shares common pathogenesis mechanisms [5].
Currently, there are registered vaccines available to fight this global crisis, and multiple vaccine development programs are underway [6,7,8,9]. However, there are only a handful of therapeutic options for COVID-19 treatment and no registered antiviral drugs against SARS-CoV-2 at the moment [10,11,12,13,14]. Furthermore, research suggests a minimal variation in the genome sequence of SARS-CoV-2 pathogen may translate to changes in the structures of viral proteins rendering available vaccines or even medicines ineffective [15]. In late 2020, early 2021, the emergence of the new SARS-CoV-2 variants was reported; namely the B.1.1.7 variant, dubbed the UK variant, the B.1.351 variant or South African variant and B.1.617, known as the Indian variant [16,17,18]. Both variants are reported to possess N501Y mutation in the RBD (receptor binding domain) of the Sprot (spike protein) that is associated with increased viral transmission [19]. The South African variant also possesses K417N and E484K mutations in the Sprot that are potentially responsible for the diminished binding of viral Sprot to host antibodies [20]. In Brazil, the P.1 variant with known N501Y, E484K and novel K417T mutation at the Sprot was identified [21]. A SARS-CoV-2 variant summary is presented in Table 1.
It is of key interest that no mutations have been observed on SARS-CoV-2 3CLpro main protease (location 3292 → 3582 on ORF1ab polyprotein; YP_009724389.1) and only one mutation on SARS-CoV-2 PLpro protease (location 1564 → 1882 on ORF1ab polyprotein; YP_009724389.1) for each variant [22]. It should be stressed, however, that this does not mean that these proteins cannot mutate in the future. In this context, 3CLpro is an attractive target for novel antiviral discovery with a potential for trans-variant activity [23,24,25,26]
3CLpro (Picornain 3C-like protease, also referred to as Mpro for main protease) is a homodimeric cysteine protease (EC 3.4.22.69) and is 96% sequence identical with the SARS-CoV Mpro [27]. The enzyme belongs to the family C30 peptidases in the PA peptidases clan. It consists of two 306 residues long polypeptide chains, which fold into three domains (I, II and III). Domains I and II have an antiparallel β-barrel structure, while domain III is composed of 5 α-helices, which connect to domain II by a long loop region. Judging by its dimer interface, it seems the dimer (comprised of protomers A and B) is an active form which is considerably less efficient when isolated in its monomer form. The active site is readily accessible to the solvent and is located distal to the dimer interface [28]. The substrate binding site is comprised of pockets P1, P1’, P2 and P3. The P1 subsite is formed with Phe140, Asn142, Glu166, His163 and His172 residues (Figure 1) and two conserved water molecules. P2 is a deep pocket comprised of His41, Met49, Tyr54, Met165 and Asp187 residues while P3 is defined by Leu168 and flanked by Glu166, Pro168 and Gly170 [29]. Proteolysis occurs via a catalytic dyad defined by Cys145 and His41 [30]. The enzyme is responsible for cleavage on no less than 11 sites on the large viral polyprotein 1ab. Cleavage generally follows the pattern Leu/Phe/Met-Gln ↓ Gly/Ser/Ala (↓ denotes the cleavage site). Glutamine at the P1 position is crucial for proteolysis to occur. As there are no known native human enzymes with such cleavage sites, Mpro looks to be an ideal drug target, since there is a low risk for toxic effects on host cells [31,32,33]. Structural data on the enzyme is available and reporter assays developed making this target suitable for novel antiviral design [34,35].
In this study, we perform a high-throughput virtual screening (HTVS) experiment coupled with machine learning classification to offer a prioritisation approach for compounds with potential activity on SARS-CoV-2 3CLpro. We employ fast methodologies to cover a comprehensive chemical space based on molecular docking scores in order to offer prioritisation lists of compounds for further free energy calculations and suitable for biological evaluation [36]. We focus on identifying novel potential non-covalent protease inhibitors, as we firmly believe they offer the flexibility of optimisation and synthetic or commercial availability [37,38].

2. Results and Discussion

2.1. Database Preparation

For our HTVS (high-throughput virtual screening) experiment, we employed the ZINC 15 library [39]. To produce robust results that would enable downstream experimental support and biological evaluation, we specifically selected the "In-Stock" library subset that includes commercially readily-available compounds. The subset was trenched to exclude small fragments below the molecular weight of 200 g/mol and included 9,232,022 compounds in total (Figure 2). The next step was compound ionisation at the pH of 7.4, the calculation of initial 3D conformations for the whole database, the enumeration of undefined chiral centres, and the removal of structural faults. Smiles strings were syntactically validated (all rings/branches closed, no illegal atom types) with RDKit software (MolFromSmiles). Ionisation was performed using OpenEye QUACPAC 2.1.1.0 software (OpenEye Scientific Software, Inc., Santa Fe, NM, USA; www.eyesopen.com; accessed on 8 May 2021) with FixpKa module and ionise 7.4 parameters. The initial 3D conformation was calculated with OpenEye OMEGA 2.5.1.4 software (OpenEye Scientific Software, Inc., Santa Fe, NM, USA; www.eyesopen.com). The following parameters were used: maxconfs 1 and flipper: true [40].

2.2. Binding Site Identification

Upon examination of available SARS-CoV-2 3CLpro crystal structures at the time, we selected the complex (PDB ID: 6Y7M) with an excellent resolution of 1.9 Å deposited by Zhang L. et al. [41]. This complex contained a relatively large peptidomimetic inhibitor OEW (MW = 585.69 g/mol) occupying major pockets at the enzyme active site, essentially keeping the S1 pocket in the correct shape and the enzyme in the active conformation. After superimposition to a reference structure (PDB ID:6LU7) by Yang et al., the catalytic binding pocket was defined around Cys145 as published previously by Jukic et al. [29,42]. Finally, the target was prepared as a sphere of 7 Å around the ligand for docking volume calculation using the CmDock docking package (https://gitlab.com/Jukic/cmdock/; accessed on 8 May 2021; Figure 3; details are in Supplementary Materials).

2.3. HTVS

For the HTVS campaign, the complete pre-prepared database was docked using CmDock into the prepared receptor binding site to afford 1‰ of top-scoring compounds with the Z-score cutoff of –7.7 (Figure 4).
In order to inspect all chemical space, filtering was done post-docking, where subsequently PAINS [43,44,45] and REOS structures were filtered out [46,47]. KNIME software with RDKit nodes was applied to compare all structures in the library to the selection of SMARTS-formatted PAINS and to remove flagged hits from the database followed by REOS filtering with Schrödinger SMD software (Schrödinger LLC., New York, NY, USA). In this way, approximately 40% of the pan-assay interfering and reactive compounds (with labile functional groups) were filtered out, and the top 200 scoring compounds were examined by cluster analysis and classified in the current 3CLpro-actives space. Hierarchical clustering was performed using Schrödinger SMD (Schrödinger LLC., New York, NY, USA) using Molprint2D hashed fingerprints, Tanimoto similarity metric and average cluster linkage method; the 17 clusters have been estimated by the Kelley criterion [48]. The final compound selection was performed with top-scoring hits with consideration of QuickProp QPlogS (Schrödinger LLC., New York, NY, USA) descriptor (>−6.5) in order to focus on compounds that have a greater potential to be soluble. Complete QuickProp descriptor set was calculated to flag all un-/desired properties and enable further custom prioritisation of compounds, and the whole set of top 200 hits is supplied in the Supplementary Materials for the convenience of future research. Upon examination of all top-scoring compounds, we identified that they all conform to the classic P1-P2-P3 binding pockets, as described previously [42]. Predicted bound conformations for the top-scoring hit as well as for the first ten hit compounds are analogous (and in accordance to the crystal ligand OEW), and compounds interact with key Thr25, Leu27, Gly143, Ser144, Cys145, His163, His164, Met165, Glu166, Asp187, Thr190, Gln189 and Gln192 residues at the 3CLpro active site (Table 2, Figure 5).

2.4. Chemical Space Prediction and Classification Supported by Machine Learning

We also investigated whether state-of-the-art machine learning algorithms could be of use for compound prioritisation. In this work, we implemented four different machine learning algorithms capable of detecting various complexities and evaluated their performance on the training data. The training data for the regression experiment consisted of all compounds with activity on the 3CLpro enzyme (ChEMBL standard value IC50 in nM, 133 compounds with 1k decoy set, available in Supplementary Materials) present in the ChEMBL database, while the training data for classification consisted of chemical libraries active on 3CLpro as well as actives on all viral cysteine proteases (3145 compounds with 1k decoy set, available in Supplementary Materials) [49]. For all compounds, we used RDKit.Chem package. RDKit.Chem.Descriptors module calculated a set of 208 chemical descriptors in order to capture the respective chemical space. In this manner, we could effectively classify compounds in the context of their representative chemical space: similar to known 3CLpro inhibitors, similar to known viral cysteine protease inhibitors, or belonging to a different chemical space to help in future design directions for individual scaffolds. The learning algorithms were (from simplest to more complex ones): a majority classifier (prediction of the most common class), a logistic regression classifier (linear classifier), and two non-linear classifiers, namely a deep feedforward neural network, designed specifically for this task, as well as the extreme gradient boosting machines (XGB), which are considered a strong learner in contemporary chemoinformatics [50,51,52,53].

3. Materials and Methods

3.1. Target Preparation

Complex (PDB ID: 6Y7M) contains {tert}-butyl ~{N}-[1-[(2~{S})-3-cyclohexyl-1-[[(2~{S},3~{R})-4-(cyclopropylamino)-3-oxidanyl-4-oxidanylidene-1-[(3~{R})-2-oxidanylidene-3,4-dihydropyrrol-3-yl]butan-2-yl]amino]-1-oxidanylidene-propan-2-yl]-2-oxidanylidene-pyridin-3-yl]carbamate (OEW), a peptide-like covalent inhibitor where thiohemiketal is formed by the nucleophilic attack of the catalytic cysteine (Cys145) onto the carbonyl group of the inhibitor. Residue conformation at the active site was checked by superposition with PDB ID: 6M2N and 6XMK with an all atom RMSD of 0.330 and 0.662 Å, respectively [54,55]. For the docking receptor preparation, the covalent bond was cleaved, a small molecule removed, and the Cys145 amino-acid residue regenerated (Open Source PyMOL, release 2.1). The target was prepared using Schrödinger Small-Molecule Discovery Suite (Schrödinger LLC., New York), protein preparation module. Missing hydrogens were added, h-bonding network was optimised using PROPKA tool at pH 7.4, waters were removed, and restrained minimisation was performed with a convergence of heavy atoms towards 0.3 Å. Finally, using the cmcavity program from CmDock docking package (https://gitlab.com/Jukic/cmdock/; accessed on 8 May 2021), we generated a docking receptor [56,57,58]. The reference ligand method was employed for cavity calculation (receptor definition), where we used the OEW-cleaved regenerated ligand as a reference and a sphere of 7 Å around the ligand for the docking volume (cavity volume) calculation. We calculated a total cavity volume of 3106.25 A3 and included the calculated cavity (Cavity #1) in the docking receptor definition. Cavity #1 parameters were the size of 24850 points; min = (−33.5,−53.5,−8.5); max = (−13,−26.5,12); centre = (−24.4138,−38.9632,−0.179235); coordinates and extent = (20.5,27,20.5) Å (Figure 3, details are in Supplementary Materials).

3.2. Virtual Screening Experiment Design

For the virtual screening experiment (HTVS), a docking approach with a robust CmDock software, the prepared compound database, and the docking receptor (Cavity #1), as described in the previous section, were employed [56,57]. Firstly, we conducted a redocking experiment. The reference-regenerated OEW peptidomimetic ligand (PDB ID: 6Y7M) was prepared as a SMILES string and energy-minimised in Ligprep tool from Schrödinger SMD using the OPLS 3e forcefield. The minimised structure was subsequently used as an input for the redocking experiment into the prepared receptor in a non-covalent manner. Applied parameters for the CmDock software (v 0.1.1) were standard docking protocol (dock.prm) with 100 runs, no constraints, and no score filters. We successfully retrieved the crystal-complex binding conformation of the OEW ligand with an RMSD of 1.34 Å. Furthermore, we calculated the receiver operating characteristic (ROC) curve to validate the performance of the classifier docking method. We selected a set of known SARS-CoV 3CLpro inhibitors from the ChEMBL database with experimental IC50 < 100 μM values and created a testing database by the addition of negative control compounds that were calculated decoys based on employed actives using DUD-E: A Database of Useful (Docking) Decoys [59]. Upon using 1% and 10% of actives in the test database, we obtained a ROC AUC of 0.80 and 0.79, respectively. We also employed the activity data from the PostEra Covid Moonshot project (https://covid.postera.ai/covid/activity_data; accessed on 8 May 2021). We selected compounds with pIC50 above 7 as true actives and compounds with pIC50 up to 4.00436 as inactives or experimental decoys (compounds with no data were left out). When using 2% and 10% of actives in the test set, we obtained ROC AUCs of 0.61, indicating that our docking protocol can indeed identify active compounds and produce enriched libraries.
In order to effectively utilise CPU-time in downstream calculations, we analysed the chemical library performance in HTVS. We sampled a random 10% population of the designed library and performed an exhaustive docking experiment on 977,600 molecules (dock.prm protocol, 100 runs per molecule, no constraints, and no score filter). We then analysed the docking results using the sdreport script (part of CmDock package) and KNIME software, where the mean docking score was −12.317, and the standard deviation was 4.273. Upon passthrough analysis using rbhtfinder script (part of CmDock package, parameters: −15 and −20), the HTVS analysis afforded an optimal HTVS experiment workflow where up to 5 runs were performed for molecules that possessed docking score of −18 and up to 15 runs for molecules that were found with docking scores of −25. In the case of docking scores lower than −25, 50 runs were performed. The first filter was passed by 25.7% of molecules, the second filter was passed by 1.71% of molecules, and the run was on average 7.8% CPU-time compared to the exhaustive docking run-time, thereby effectively improving the HTVS efficiency.

3.3. Machine Learning

The deep neural network consisted of six hidden layers, where each layer was activated with the ReLU activation, followed by dropout-based regularisation, set to 0.3 [60]. The final layer of the neural network was activated with a softmax function to obtain a distribution across the space of possible classes. During prediction, the most probable class is selected (argmax across class probabilities). The extreme gradient boosting machines are a type of tree-based ensemble classifier capable of fast learning and robust and accurate predictions. The evaluation regime was designed as follows: we randomly sampled ten different stratified splits, where the training data was used to learn the models, followed by the evaluation of their predictions on the test data. The splits were in ratio 8:2. We report the average performance with a standard deviation for the macro F1 score (the task considered is a multiclass classification). Apart from the classification task, the point was to differentiate between chemical spaces occupied by known 3Clpro inhibitors reported in ChEMBL, known viral cysteine protease inhibitors in the ChEMBL database, and other chemical space. We also conducted a similar series of experiments to assess to what extent machine learning can predict the inhibition of 3CLpro directly (on the basis of the ChEMBL standard value IC50 in nM), which is a regression task. The regression variants of the algorithms mentioned above were considered, namely the XGB with the “req:squarederror” loss, the neural network’s classification head was replaced with a regression one, and instead of logistic regression, we used a simple linear regressor. The neural network that performed best consisted of seven (regularized) hidden layers; other hyperparameters were the same as in the classification scenario. As the neural network’s expected output is a positive real number, the final activation used was a ReLU as well, as we observed faster convergence compared to using no activation at all. Mean squared errors with standard deviation across ten random splits (see the previous section) are reported (Table 3).
The neural network and XGB performed, as expected, the best, followed by a simpler linear learner. Compared to the majority baseline, e.g., the neural network’s performance is substantially higher, indicating that it learned to differentiate between the classes. On the contrary, the average-training baseline (averaged train predictions) performed better than the linear classifier, indicating that the regression problem at hand is relatively hard and potentially not linearly separable. The XGB and neural networks performed better than this baseline, albeit not by a large margin, indicating the actives 3CLpro dataset was insufficient for effective training. However, the regression task similarly indicates that it is possible to use non-linear models for direct activity prediction necessitating good data collection on 3CLpro actives in the future. Furthermore, the authors are aware of the generalisation of the idea on other well-represented datasets and postulate its great value in future in silico drug design.

4. Conclusions

Herein, we presented a novel in silico approach towards compound prioritisation in the design of novel 3CLpro inhibitors. Rather than relying on commercial “corona-focused libraries”, we performed a robust and efficient HTVS experiment to obtain virtual hit compounds. We then coupled the method to a machine learning classification experiment where each compound is classified into the chemical space of 3CLpro inhibitors, into general viral cysteine protease inhibitors, or into a completely novel unrelated chemical space. In this way, medicinal chemists can turn their attention towards compounds that would otherwise escape and derive insight into the possible mechanism of action. Therefore, we employed a complete library of 3CLpro inhibitors and viral cysteine protease inhibitors obtained from the ChEMBL library using the calculated ensemble of 208 descriptors for each compound in the library. We reported ten top-scoring compounds as viable binders supported by CmDock docking calculations considering the QuickProp QPlogS descriptor that indicates possible soluble compounds. Namely, solubility forms one of the serious problems when transitioning from in silico enrichment lists towards physical samples for in vitro evaluation. We supply full lists of hit compounds in the Supplementary Materials for the reader’s benefit, especially if alternative compound selection criteria are desired, all with the aim of providing the medicinal chemistry community with a viable prioritisation library of potential 3CLpro inhibitors tailored for further molecular dynamics (MD) studies [61] as well as experimental design and development. Namely, reported enriched lists of compounds can serve as starting points, whereas compounds can be purchased commercially or be a subject of a synthetic campaign. Compounds can thus serve as experimental decoys or, if successful, progress as leads or probes.

Supplementary Materials

Details on target preparation and virtual screening. References [62] and [63] are cited in the supplementary materials.

Author Contributions

Conceptualization, M.J., Č.P. and U.B.; data curation, M.J. and Č.P.; formal analysis, M.J., B.Š. and Č.P.; funding acquisition, Č.P. and U.B.; investigation, M.J., B.Š., G.T., S.P., Č.P. and U.B.; methodology, M.J. and B.Š.; project administration, M.J., U.B. and Č.P.; supervision, Č.P. and U.B.; validation, M.J., Č.P. and U.B.; writing—original draft, M.J.; writing—review and editing, all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Slovenian Ministry of Science and Education infrastructure project grant HPC-RIVR, by the Slovenian Research Agency (ARRS) programme and project grants P2-0046 and J1-2471, the Physical Chemistry programme grant P1-0201, the Knowledge Technologies programme group (P2-0103 and P6-0411) as well as Slovenian Ministry of Education, Science and Sports programme grant OP20.04342. The work of the second author was funded via the ARRS junior research grant.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or supplementary materials.

Acknowledgments

We gratefully acknowledge NVIDIA Corporation’s support with the donation of GPU hardware used in this research. We thank OpenEye for the academic licencing of their software and their support. We thank Rita Podzuna and support from Schrödinger LLC. Thanks to Boštjan Laba for support in the community project. Last but not least: a heartfelt thanks from all authors to all participants in the COVID.SI community (www.covid.si and www.sidock.si) for supporting our work. This represents a wonderful example of citizen science, and we are grateful for all joint scientific and community work. Thank You All!

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available (compounds are commercially available).

Abbreviations

3CLpro3C-like protease
MDMolecular Dynamics
VSVirtual Screening
HTVSHigh-Throughput Virtual Screening
LIELinear Interaction Energy
XgbGradient boosting framework library

References

  1. Huang, C.; Wang, Y.; Li, X.; Ren, L.; Zhao, J.; Hu, Y.; Cao, B. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet 2020, 395, 497–506. [Google Scholar] [CrossRef] [Green Version]
  2. Wang, D.; Hu, B.; Hu, C.; Zhu, F.; Liu, X.; Zhang, J.; Peng, Z. Clinical characteristics of 138 hospitalized patients with 2019 novel coronavirus–infected pneumonia in Wuhan, China. JAMA 2020, 323, 1061–1069. [Google Scholar] [CrossRef]
  3. Wang, C.; Horby, P.W.; Hayden, F.G.; Gao, G.F. A novel coronavirus outbreak of global health concern. Lancet 2020, 395, 470–473. [Google Scholar] [CrossRef] [Green Version]
  4. Chen, Y.; Liu, Q.; Guo, D. Emerging coronaviruses: Genome structure, replication, and pathogenesis. J. Med. Virol. 2020, 92, 418–423. [Google Scholar] [CrossRef]
  5. Lu, R.; Zhao, X.; Li, J.; Niu, P.; Yang, B.; Wu, H.; Tan, W. Genomic characterisation and epidemiology of 2019 novel coronavirus: Implications for virus origins and receptor binding. Lancet 2020, 395, 565–574. [Google Scholar] [CrossRef] [Green Version]
  6. Widge, A.T.; Rouphael, N.G.; Jackson, L.A.; Anderson, E.J.; Roberts, P.C.; Makhene, M.; Beigel, J.H. Durability of responses after SARS-CoV-2 mRNA-1273 vaccination. N. Engl. J. Med. 2021, 384, 80–82. [Google Scholar] [CrossRef] [PubMed]
  7. Amanat, F.; Krammer, F. SARS-CoV-2 vaccines: Status report. Immunity 2020, 52, 583–589. [Google Scholar] [CrossRef] [PubMed]
  8. Krammer, F. SARS-CoV-2 vaccines in development. Nature 2020, 586, 516–527. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, W.H.; Strych, U.; Hotez, P.J.; Bottazzi, M.E. The SARS-CoV-2 vaccine pipeline: An overview. Curr. Trop. Med. Rep. 2020, 7, 61–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Wu, C.; Liu, Y.; Yang, Y.; Zhang, P.; Zhong, W.; Wang, Y.; Li, H. Analysis of therapeutic targets for SARS-CoV-2 and discovery of potential drugs by computational methods. Acta Pharm. Sin. B 2020, 10, 766–788. [Google Scholar] [CrossRef] [PubMed]
  11. McKee, D.L.; Sternberg, A.; Stange, U.; Laufer, S.; Naujokat, C. Candidate drugs against SARS-CoV-2 and COVID-19. Pharmacol. Res. 2020, 157, 104859. [Google Scholar] [CrossRef] [PubMed]
  12. Li, H.; Zhou, Y.; Zhang, M.; Wang, H.; Zhao, Q.; Liu, J. Updated approaches against SARS-CoV-2. Antimicrob. Agents Chemother. 2020, 64, e00483-20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Gordon, D.E.; Jang, G.M.; Bouhaddou, M.; Xu, J.; Obernier, K.; White, K.M.; Krogan, N.J. A SARS-CoV-2 protein interaction map reveals targets for drug repurposing. Nature 2020, 583, 459–468. [Google Scholar] [CrossRef] [PubMed]
  14. Riva, L.; Yuan, S.; Yin, X.; Martin-Sancho, L.; Matsunaga, N.; Pache, L.; Chanda, S.K. Discovery of SARS-CoV-2 antiviral drugs through large-scale compound repurposing. Nature 2020, 586, 113–119. [Google Scholar] [CrossRef]
  15. Naqvi, A.A.T.; Fatima, K.; Mohammad, T.; Fatima, U.; Singh, I.K.; Singh, A.; Hassan, M.I. Insights into SARS-CoV-2 genome, structure, evolution, pathogenesis and therapies: Structural genomics approach. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2020, 10, 165878. [Google Scholar] [CrossRef]
  16. Volz, E.; Swapnil, M.; Meera, C.; Barrett, J.C.; Johnson, R.; Geidelberg, L.; Hinsley, W.R.; Laydon, D.J.; Dabrera, G.; O’Toole, Á.; et al. Transmission of SARS-CoV-2 Lineage, B. 1.1. 7 in England: Insights from linking epidemiological and genetic data. MedRxiv 2021, 2020-12. [Google Scholar] [CrossRef]
  17. Tegally, H.; Wilkinson, E.; Giovanetti, M.; Iranzadeh, A.; Fonseca, V.; Giandhari, J.; Doolabh, D.; Pillay, S.; San, E.J.; Msomi, N.; et al. Emergence and rapid spread of a new severe acute respiratory syndrome-related coronavirus 2 (SARS-CoV-2) lineage with multiple spike mutations in South Africa. MedRxiv 2020. [Google Scholar] [CrossRef]
  18. Gupta, R.K. Will SARS-CoV-2 variants of concern affect the promise of vaccines? Nat. Rev. Immunol. 2021, 1–2. [Google Scholar] [CrossRef]
  19. Hongjing, G.; Chen, Q.; Yang, G.; He, L.; Fan, H.; Deng, Y.-Q.; Wang, Y.; Teng, Y.; Zhao, Z.; Cui, Y.; et al. Adaptation of SARS-CoV-2 in BALB/c mice for testing vaccine efficacy. Science 2020, 369, 1603–1607. [Google Scholar]
  20. Wibmer, C.K.; Ayres, F.; Hermanus, T.; Madzivhandila, M.; Kgagudi, P.; Lambson, B.E.; Vermeulen, M.; van den Berg, K.; Rossouw, T.; Boswell, M.; et al. SARS-CoV-2 501Y. V2 escapes neutralization by South African COVID-19 donor plasma. BioRxiv 2021, 27, 622–625. [Google Scholar]
  21. Nuno, R.F.; Morales Claro, I.; Candido, D.; Franco, L.A.M.; Andrade, P.S.; Coletti, T.M.; Silva, C.A.M.; Sales, F.C.; Manuli, E.R.; Aguiar, R.S. Genomic characterisation of an emergent SARS-CoV-2 lineage in Manaus: Preliminary findings. Virological 2021. [Google Scholar]
  22. Klemm, T.; Ebert, G.; Calleja, D.J.; Allison, C.C.; Richardson, L.W.; Bernardini, J.P.; Lu, B.G.; Kuchel, N.W.; Grohmann, C.; Shibata, Y.; et al. Mechanism and inhibition of SARS-CoV-2 PLpro. Biorxiv 2020. [Google Scholar] [CrossRef]
  23. Sumit, K.; Sharma, P.P.; Shankar, U.; Kumar, D.; Joshi, S.K.; Pena, L.; Durvasula, R.; Kumar, A.; Kempaiah, P.; Rathi, P.; et al. Discovery of new hydroxyethylamine analogs against 3CLpro protein target of SARS-CoV-2: Molecular docking, molecular dynamics simulation, and structure–activity relationship studies. J. Chem. Inf. Model. 2020, 60, 5754–5770. [Google Scholar]
  24. ul Qamar, M.T.; Alqahtani, S.M.; Alamri, M.A.; Chen, L.L. Structural basis of SARS-CoV-2 3CLpro and anti-COVID-19 drug discovery from medicinal plants. J. Pharm. Anal. 2020, 10, 313–319. [Google Scholar] [CrossRef] [PubMed]
  25. Koulgi, S.; Jani, V.; Uppuladinne, M.; Sonavane, U.; Nath, A.K.; Darbari, H.; Joshi, R. Drug repurposing studies targeting SARS-CoV-2: An ensemble docking approach on drug target 3C-like protease (3CLpro). J. Biomol. Struct. Dyn. 2020, 1–21. [Google Scholar] [CrossRef]
  26. Macchiagodena, M.; Pagliai, M.; Procacci, P. Inhibition of the main protease 3cl-pro of the coronavirus disease 19 via structure-based ligand design and molecular modeling. arXiv 2020, arXiv:2002.09937. [Google Scholar]
  27. Malcolm, B.A. The picornaviral 3C proteinases: Cysteine nucleophiles in serine proteinase folds. Prot. Sci. 1995, 4, 1439–1445. [Google Scholar] [CrossRef] [Green Version]
  28. Shi, J.; Sivaraman, J.; Song, J. Mechanism for Controlling the Dimer-Monomer Switch and Coupling Dimerization to Catalysis of the Severe Acute Respiratory Syndrome Coronavirus 3C-Like Protease. J. Virol. 2008, 82, 4620–4629. [Google Scholar] [CrossRef] [Green Version]
  29. 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 COVID-19 virus and discovery of its inhibitors. Nature 2020, 582, 289–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. 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. [Google Scholar] [CrossRef] [Green Version]
  31. Hilgenfeld, R. From SARS to MERS: Crystallographic studies on coronaviral proteases enable antiviral drug design. FEBS 2014, 281, 4085–4096. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, L.; Lin, D.; Kusov, Y.; Nian, Y.; Ma, Q.; Wang, J.; von Brunn, A.; Leyssen, P.; Lanko, K.; Neyts, J.; et al. α-Ketoamides as Broad-Spectrum Inhibitors of Coronavirus and Enterovirus Replication: Structure-Based Design, Synthesis, and Activity Assessment. J. Med. Chem. 2020, 63, 4562–4578. [Google Scholar] [CrossRef]
  33. 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 central science 2021, 7, 467–475. [Google Scholar] [CrossRef] [PubMed]
  34. Froggatt, H.M.; Heaton, B.E.; Heaton, N.S. Development of a fluorescence-based, high-throughput SARS-CoV-2 3CLpro reporter assay. J. Virol. 2020, 94, e01265-20. [Google Scholar] [CrossRef] [PubMed]
  35. Mariusz, J.; Dauter, Z.; Shabalin, I.G.; Gilski, M.; Brzezinski, D.; Kowiel, M.; Rupp, B.; Wlodawer, A. Crystallographic models of SARS-CoV-2 3CLpro: In-depth assessment of structure quality and validation. IUCrJ 2021, 8, 238–256. [Google Scholar]
  36. Macchiagodena, M.; Pagliai, M.; Karrenbrock, M.; Guarnieri, G.; Iannone, F.; Procacci, P. Virtual Double-System Single-Box: A Nonequilibrium Alchemical Technique for Absolute Binding Free Energy Calculations: Application to Ligands of the SARS-CoV-2 Main Protease. J. Chem. Theory Comput. 2020, 16, 7160–7172. [Google Scholar] [CrossRef] [PubMed]
  37. Bauer, R.A. Covalent inhibitors in drug discovery: From accidental discoveries to avoided liabilities and designed therapies. Drug Disc. Today 2015, 20, 1061–1073. [Google Scholar] [CrossRef]
  38. Baillie, T.A. Targeted covalent inhibitors for drug design. Angew. Chem. Int. Ed. 2016, 55, 13408–13421. [Google Scholar] [CrossRef]
  39. Sterling, T.; Irwin, J.J. ZINC 15–ligand discovery for everyone. J. Chem. Inf. Model. 2015, 55, 2324–2337. [Google Scholar] [CrossRef]
  40. Hawkins, P.C.D.; Skillman, A.G.; Warren, G.L.; Ellingson, B.A.; Stahl, M.T. Conformer Generation with OMEGA: Algorithm and Validation Using High Quality Structures from the Protein Databank and the Cambridge Structural Database. J. Chem. Inf. Model. 2010, 50, 572–584. [Google Scholar] [CrossRef]
  41. Zhang, L.; Lin, D.; Sun, X.; Curth, U.; Drosten, C.; Sauerhering, L.; Becker, S.; Rox, K.; Hilgenfeld, R. Crystal structure of SARS-CoV-2 main protease provides a basis for design of improved α-ketoamide inhibitors. Science 2020, 368, 409–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Jukič, M.; Janežič, D.; Bren, U. Ensemble Docking Coupled to Linear Interaction Energy Calculations for Identification of Coronavirus Main Protease (3CLpro) Non-Covalent Small-Molecule Inhibitors. Molecules 2020, 25, 5808. [Google Scholar] [CrossRef] [PubMed]
  43. Shoichet, B.K. Interpreting steep dose-response curves in early inhibitor discovery. J. Med. Chem. 2006, 49, 7274–7277. [Google Scholar] [CrossRef] [PubMed]
  44. Baell, J.B.; Holloway, G.A. New substructure filters for removal of pan assay interference compounds (PAINS) from screening libraries and for their exclusion in bioassays. J. Med. Chem. 2010, 53, 2719–2740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Saubern, S.; Guha, R.; Baell, J.B. KNIME workflow to assess PAINS filters in SMARTS format. Comparison of RDKit and Indigo cheminformatics libraries. Mol. Inform. 2011, 30, 847–850. [Google Scholar] [CrossRef] [PubMed]
  46. Walters, W.P.; Stahl, M.T.; Murcko, M.A. Virtual screening—An overview. Drug Disc. Today 1998, 3, 160–178. [Google Scholar] [CrossRef]
  47. Zhu, T.; Cao, S.; Su, P.C.; Patel, R.; Shah, D.; Chokshi, H.B.; Hevener, K.E. Hit identification and optimization in virtual screening: Practical recommendations based on a critical literature analysis: Miniperspective. J. Med. Chem. 2013, 56, 6560–6572. [Google Scholar] [CrossRef] [Green Version]
  48. Kelley, L.A.; Gardner, S.P.; Sutcliffe, M.J. An automated approach for clustering an ensemble of NMR-derived protein structures into conformationally-related subfamilies. Protein Eng. 1996, 9, 1063–1065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Gaulton, A.; Bellis, L.J.; Bento, A.P.; Chambers, J.; Davies, M.; Hersey, A.; Overington, J.P. ChEMBL: A large-scale bioactivity database for drug discovery. Nucleic Acids Res. 2012, 40, D1100–D1107. [Google Scholar] [CrossRef] [Green Version]
  50. Goodfellow, I.; Bengio, Y.; Courville, A.; Bengio, Y. Deep Learning; MIT Press: Cambridge, MA, USA, 2016; Volume 1, No. 2. [Google Scholar]
  51. Pedregosa, F.; Varoquaux, G.; Gramfort, A.; Michel, V.; Thirion, B.; Grisel, O.; Duchesnay, E. Scikit-learn: Machine learning in Python. J. Mach. Learn. Res. 2011, 12, 2825–2830. [Google Scholar]
  52. Chen, T.; Guestrin, C. Xgboost: A Scalable Tree Boosting System. In Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, San Francisco, CA, USA, 13–17 August 2016; pp. 785–794. [Google Scholar]
  53. Chollet, F. Keras. 2015. Available online: https://keras.io (accessed on 8 May 2021).
  54. Su, H.X.; Yao, S.; Zhao, W.F.; Li, M.J.; Liu, J.; Shang, W.J.; Xie, H.; Ke, C.; Hu, H.; Gao, M.; et al. Anti-SARS-CoV-2 activities in vitro of Shuanghuanglian preparations and bioactive ingredients. Acta Pharmacol. Sin. 2020, 41, 1167–1177. [Google Scholar] [CrossRef] [PubMed]
  55. 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, 1–11. [Google Scholar] [CrossRef] [PubMed]
  56. Morley, S.D.; Afshar, M. Validation of an empirical RNA-ligand scoring function for fast flexible docking using RiboDock. J. Comput. Aided Mol. Des. 2004, 18, 189–208. [Google Scholar] [CrossRef] [PubMed]
  57. Ruiz-Carmona, S.; Alvarez-Garcia, D.; Foloppe, N.; Garmendia-Doval, A.B.; Juhos, S.; Schmidtke, B.; Barril, X.; Hubbard, R.E.; Morley, S.D. rDock: A Fast, Versatile and Open Source Program for Docking Ligands to Proteins and Nucleic Acids. PLoS Comput. Biol. 2014, 10, e1003571. [Google Scholar] [CrossRef] [Green Version]
  58. DeLano, W.L. Pymol: An open-source molecular graphics tool. CCP4 Newsl. Protein Crystallogr. 2002, 40, 82–92. [Google Scholar]
  59. Mysinger, M.M.; Carchia, M.; Irwin, J.J.; Shoichet, B.K. Directory of useful decoys, enhanced (DUD-E): Better ligands and decoys for better benchmarking. J. Med. Chem. 2012, 55, 6582–6594. [Google Scholar] [CrossRef]
  60. Srivastava, N.; Hinton, G.; Krizhevsky, A.; Sutskever, I.; Salakhutdinov, R. Dropout: A simple way to prevent neural networks from overfitting. J. Mach. Learn. Res. 2014, 15, 1929–1958. [Google Scholar]
  61. Graf, M.M.; Bren, U.; Haltrich, D.; Oostenbrink, C. Molecular dynamics simulations give insight into D-glucose dioxidation at C2 and C3 by Agaricus meleagris pyranose dehydrogenase. Comput. Aided Mol. Des. 2013, 27, 295–304. [Google Scholar] [CrossRef] [Green Version]
  62. Jukic, M.; Ilc, N.; Sluga, D.; Tomšič, G.; Podlipnik, Č. CmDock. 2021. Available online: https://gitlab.com/Jukic/cmdock/ (accessed on 8 May 2021).
  63. Tosco, P.; Stiefl, N.; Landrum, G. Bringing the MMFF force field to the RDKit: Implementation and validation. J. Cheminformatics 2014, 6, 1–4. [Google Scholar] [CrossRef]
Figure 1. Active site of SARS-CoV-2 3Clpro or Mpro enzyme (PDB ID: 6Y7M) with emphasised small molecule and pocket designation on the right. Active site residues are depicted in green coloured line model with emphasized transparent blue-white surface 6 Å around the small molecule inhibitor depicted in blue coloured stick model.
Figure 1. Active site of SARS-CoV-2 3Clpro or Mpro enzyme (PDB ID: 6Y7M) with emphasised small molecule and pocket designation on the right. Active site residues are depicted in green coloured line model with emphasized transparent blue-white surface 6 Å around the small molecule inhibitor depicted in blue coloured stick model.
Molecules 26 03003 g001
Figure 2. Detailed ZINC 15 database “In-Stock” subset tranche description with a total of 9,322,002 compounds used for further calculations (As obtained from the https://zinc15.docking.org; accessed on 8 May 2021; at the time of smiles compound download; the database on the master server is continuously updating).
Figure 2. Detailed ZINC 15 database “In-Stock” subset tranche description with a total of 9,322,002 compounds used for further calculations (As obtained from the https://zinc15.docking.org; accessed on 8 May 2021; at the time of smiles compound download; the database on the master server is continuously updating).
Molecules 26 03003 g002
Figure 3. Generated receptor volume with the 3CLpro PDB ID: 6Y7M, the active site near the Cys145 residue and the docking volume defined as a sphere of 7 Å around the reference ligand OEW. Protein is depicted as a pink-magenta-coloured cartoon model with residues emphasised in the line model and coloured atoms (carbon in green, oxygen in red, nitrogen in blue and hydrogen in white colour) with a blue-white transparent active site surface. Docking volume boundary mesh is depicted in blue colour. Isomesh (0.99) was constructed using PyMol 2.1.0 using a grid calculated with cmgrid software (v 0.1.1).
Figure 3. Generated receptor volume with the 3CLpro PDB ID: 6Y7M, the active site near the Cys145 residue and the docking volume defined as a sphere of 7 Å around the reference ligand OEW. Protein is depicted as a pink-magenta-coloured cartoon model with residues emphasised in the line model and coloured atoms (carbon in green, oxygen in red, nitrogen in blue and hydrogen in white colour) with a blue-white transparent active site surface. Docking volume boundary mesh is depicted in blue colour. Isomesh (0.99) was constructed using PyMol 2.1.0 using a grid calculated with cmgrid software (v 0.1.1).
Molecules 26 03003 g003
Figure 4. HTVS workflow with post-docking filtering, cluster analysis, and compound classification according to the chemical space of 3CLpro inhibitors collected in the ChEMBL database.
Figure 4. HTVS workflow with post-docking filtering, cluster analysis, and compound classification according to the chemical space of 3CLpro inhibitors collected in the ChEMBL database.
Molecules 26 03003 g004
Figure 5. Calculated bound conformations in the 3CLpro active site of the top-scoring hit compounds. Protein is presented in a cartoon model coloured pink with an emphasised molecular surface in light-blue colour. (A); the reference OEW ligand is presented in stick model cored magenta, while the top-scoring hit is coloured grey. (B) the reference OEW ligand is presented in stick model cored magenta while the top 10 scoring compounds are depicted in red-coloured line representations to emphasise their analogous binding mode in the P1-P2-P3 pockets of the active site.
Figure 5. Calculated bound conformations in the 3CLpro active site of the top-scoring hit compounds. Protein is presented in a cartoon model coloured pink with an emphasised molecular surface in light-blue colour. (A); the reference OEW ligand is presented in stick model cored magenta, while the top-scoring hit is coloured grey. (B) the reference OEW ligand is presented in stick model cored magenta while the top 10 scoring compounds are depicted in red-coloured line representations to emphasise their analogous binding mode in the P1-P2-P3 pockets of the active site.
Molecules 26 03003 g005
Table 1. Summary of dominant SARS-CoV-2 variants and relevant mutations.
Table 1. Summary of dominant SARS-CoV-2 variants and relevant mutations.
Variant 1Alternative NameSprot/
All Mutations
Key MutationsComment3CLpro/PLpro
Mutations 2
B.1.1.7UK Variant8/23E69/70 del
144Y del
N501Y (RBD interface)
A570D
P681H
higher transmissibilitynone/A1708D
B.1.351South African Variant9/21K417N (RBD)
E484K (RBD)
N501Y (RBD)
orf1b del
escape host immune responsenone/K1655N
P.1Brasil Variant10/17K417N/T (RBD)
E484K (RBD)
N501Y (RBD)
orf1b del
under researchnone/K1795Q
B.1.617Indian Variant7/23G142D
E154K
L452R (RBD)
E484Q (RBD)
D614G
P681R
Q1071H
under researchnone/under research
1 Other known variants are COH.20G, S Q677H (Midwest variant) and L452R, B1429; 2 The mutations on PLpro are located far outside the enzyme’s active site.
Table 2. Identified top-scoring compounds in the HTVS on the SARS-CoV-2 main protease for further compound prioritisation in biological evaluation experiments.
Table 2. Identified top-scoring compounds in the HTVS on the SARS-CoV-2 main protease for further compound prioritisation in biological evaluation experiments.
noStructureMr (g/mol)Cluster/QPlogS 1CmDock Docking Score 2Classification 3
1 Molecules 26 03003 i001451.545/−6.42−32.51general
2 Molecules 26 03003 i002465.355/−6.22−29.02general
3 Molecules 26 03003 i003459.495/−3.89−26.80viral_cys_prot
4 Molecules 26 03003 i004400.454/−4.54−25.58viral_cys_prot
5 Molecules 26 03003 i005325.375/−3.59−25.53viral_cys_prot
6 Molecules 26 03003 i006396.855/−6.47−25.05general
7 Molecules 26 03003 i007399.832/−3.44−24.76general
8 Molecules 26 03003 i008425.504/−5.38−24.51general
9 Molecules 26 03003 i009353.445/−4.40−24.17general
10 Molecules 26 03003 i010494.555/−5.60−24.01general
11 Molecules 26 03003 i011490.605/−2.87−23.98general
12 Molecules 26 03003 i012337.375/−4.75−23.61viral_cys_prot
13 Molecules 26 03003 i013425.525/−5.66−23.53general
14 Molecules 26 03003 i014335.345/−3.90−23.26general
15 Molecules 26 03003 i015401.444/−5.00−23.18general
1 QPlogS Predicted aqueous solubility, log S. S in mol dm−3 is the solute concentration in a saturated solution in equilibrium with the crystalline solid (recommended value range by QuickProp is between –6.5 and –0.5); 3 CmDock INTER-molecular docking score.; 3 As per NeuralNet model.
Table 3. Machine learning model accuracy comparison table.
Table 3. Machine learning model accuracy comparison table.
Macro F1/mse 1NeuralNet 2XGB 2Linear 2Majority/Average 3
Classification0.895 ± 0.050.889 ± 0.0140.667 ± 0.0150.283 ± 0.001
Regression0.002 ± 0.0010.003 ±0.0010.012 ± 0.0020.005 ± 0.001
1 macro F1 for classification and mse for regression; 2 Learner; 3 majority class classifier in classification and average of the training target space in a regression model.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jukič, M.; Škrlj, B.; Tomšič, G.; Pleško, S.; Podlipnik, Č.; Bren, U. Prioritisation of Compounds for 3CLpro Inhibitor Development on SARS-CoV-2 Variants. Molecules 2021, 26, 3003. https://doi.org/10.3390/molecules26103003

AMA Style

Jukič M, Škrlj B, Tomšič G, Pleško S, Podlipnik Č, Bren U. Prioritisation of Compounds for 3CLpro Inhibitor Development on SARS-CoV-2 Variants. Molecules. 2021; 26(10):3003. https://doi.org/10.3390/molecules26103003

Chicago/Turabian Style

Jukič, Marko, Blaž Škrlj, Gašper Tomšič, Sebastian Pleško, Črtomir Podlipnik, and Urban Bren. 2021. "Prioritisation of Compounds for 3CLpro Inhibitor Development on SARS-CoV-2 Variants" Molecules 26, no. 10: 3003. https://doi.org/10.3390/molecules26103003

Article Metrics

Back to TopTop