Next Article in Journal
Epidemiology and Genetic Variability of HHV-8/KSHV among Rural Populations and Kaposi’s Sarcoma Patients in Gabon, Central Africa. Review of the Geographical Distribution of HHV-8 K1 Genotypes in Africa
Next Article in Special Issue
Quinine Inhibits Infection of Human Cell Lines with SARS-CoV-2
Previous Article in Journal
Genomic Evidence for Sequestration of Influenza A Virus Lineages in Sea Duck Host Species
Previous Article in Special Issue
Neuroinvasion and Encephalitis Following Intranasal Inoculation of SARS-CoV-2 in K18-hACE2 Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of a Cell-Based Luciferase Complementation Assay for Identification of SARS-CoV-2 3CLpro Inhibitors

1
Viral Recombination Section, HIV Dynamics and Replication Program, National Cancer Institute, Frederick, MD 21702, USA
2
Viral Mutation Section, HIV Dynamics and Replication Program, National Cancer Institute, Frederick, MD 21702, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2021, 13(2), 173; https://doi.org/10.3390/v13020173
Submission received: 17 December 2020 / Revised: 13 January 2021 / Accepted: 20 January 2021 / Published: 24 January 2021
(This article belongs to the Special Issue Pathogenesis of Human and Animal Coronaviruses)

Abstract

:
The 3C-like protease (3CLpro) of SARS-CoV-2 is considered an excellent target for COVID-19 antiviral drug development because it is essential for viral replication and has a cleavage specificity distinct from human proteases. However, drug development for 3CLpro has been hindered by a lack of cell-based reporter assays that can be performed in a BSL-2 setting. Current efforts to identify 3CLpro inhibitors largely rely upon in vitro screening, which fails to account for cell permeability and cytotoxicity of compounds, or assays involving replication-competent virus, which must be performed in a BSL-3 facility. To address these limitations, we have developed a novel cell-based luciferase complementation reporter assay to identify inhibitors of SARS-CoV-2 3CLpro in a BSL-2 setting. The assay is based on a lentiviral vector that co-expresses 3CLpro and two luciferase fragments linked together by a 3CLpro cleavage site. 3CLpro-mediated cleavage results in a loss of complementation and low luciferase activity, whereas inhibition of 3CLpro results in 10-fold higher levels of luciferase activity. The luciferase reporter assay can easily distinguish true 3CLpro inhibition from cytotoxicity, a powerful feature that should reduce false positives during screening. Using the assay, we screened 32 small molecules for activity against SARS-CoV-2 3CLpro, including HIV protease inhibitors, HCV protease inhibitors, and various other compounds that have been reported to inhibit SARS-CoV-2 3CLpro. Of these, only five exhibited significant inhibition of 3CLpro in cells: GC376, boceprevir, Z-FA-FMK, calpain inhibitor XII, and GRL-0496. This assay should greatly facilitate efforts to identify more potent inhibitors of SARS-CoV-2 3CLpro.

1. Introduction

In December 2019, a novel human coronavirus, now known as severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), emerged in Wuhan, China [1,2]. This virus is the causative agent of the global COVID-19 pandemic, which has resulted in over 70 million infections and 1.5 million deaths as of 11 December 2020. While promising results have been reported for several candidate vaccines in clinical trials [3,4,5], the widespread distribution of any vaccine will require a significant amount of time, leading to a pressing need for antiviral therapeutics. Recently, two monoclonal antibody therapies received emergency use authorization from the FDA for treatment of mild-to-moderate COVID-19 in patients at high risk for severe disease [6]. In addition, remdesivir, a nucleotide analog that inhibits the SARS-CoV-2 RNA-dependent RNA polymerase [7], has been approved by the FDA for COVID-19 in hospitalized patients. However, clinical trials of remdesivir have produced mixed results [8,9,10,11], highlighting an urgent need for additional antiviral drugs.
Following entry and uncoating of the SARS-CoV-2 genome into the cytoplasm, two overlapping polyproteins, pp1a and pp1ab, are expressed by translation of ORF1a and 1b, with the latter resulting from −1 programmed ribosomal frameshifting [12]. These polyproteins contain two cysteine proteases: papain-like protease (PLpro, Nsp3) and 3C-like protease (3CLpro, Mpro, Nsp5), which are responsible for 3 and 11 cleavage events, respectively [12]. SARS-CoV-2 3CLpro consists of three domains: Domain I (residues 10–99) and II (residues 100–182) resemble picornavirus 3C proteases and chymotrypsin [13]. The active site, which contains a Cys-His catalytic dyad (H41, C145), is situated between these domains. Domain III (residues 198–303) regulates 3CLpro homodimerization, which is required for enzymatic activity [14]. 3CLpro is widely regarded as an excellent target for COVID-19 drug development because it is essential for viral replication and has a cleavage specificity distinct from human cellular proteases, reducing the likelihood of off-target effects [13,15,16]. The preferred sequence at cleavage sites is LQ ↓ (S, A, G), with the arrow indicating the cleavage site [13]. Furthermore, the structure of 3CLpro is highly conserved across coronaviruses, suggesting it may be possible to develop pan-coronavirus inhibitors that target 3CLpro [17].
Unfortunately, drug development for SARS-CoV-2 3CLpro has been greatly hindered by a lack of tractable, cell-based assays that can be performed in a biosafety level 2 (BSL-2) setting. Some of the existing experimental methods that have been used to identify 3CLpro inhibitors are described below. Many groups have performed in vitro screens using purified 3CLpro protein [18,19,20,21,22,23,24], but these systems do not account for cell permeability, metabolization, or cytotoxicity of compounds. In addition, the recombinant 3CLpro protein used in these assays may differ from 3CLpro expressed in human cells. Consequently, initial hits often have high toxicity and/or poor activity in cell-based assays. Alternatively, inhibitors can be identified by directly assessing SARS-CoV-2 replication in cell culture, but this requires biosafety level 3 (BSL-3) containment and is not amenable to high-throughput screening. Additionally, in these assays, virus replication is often quantified by a plaque assay or quantitative RT-PCR, which can lead to difficulties in distinguishing antiviral activity from cytotoxicity, as cytotoxicity also results in reduced virus production. Recently, a cell-based BSL-2 reporter system for identification of SARS-CoV-2 3CLpro inhibitors was developed [25]. However, this system, based on a fluorescent reporter called Flip-GFP, represents a loss-of-function approach, which may lead to difficulty in discriminating cytotoxicity from 3CLpro inhibition. Lastly, virtual screening and molecular dynamics simulations have widely been used to identify candidate inhibitors for experimental follow-up [26,27,28].
Here, we describe a novel luciferase complementation reporter assay to identify inhibitors of SARS-CoV-2 3CLpro in living cells. The assay can be performed in a BSL-2 setting and is based on a single lentiviral vector that co-expresses 3CLpro and a split reporter in which two luciferase fragments are connected by a cleavage site. When 3CLpro cuts the cleavage site, luciferase complementation is lost, resulting in low luciferase activity. In contrast, when a compound is added that inhibits 3CLpro, cleavage is blocked and complementation is maintained, resulting in significantly higher luciferase activity. Using the assay, we screened 32 small molecules for activity against SARS-CoV-2 3CLpro, leading to the identification of five active compounds: GC376, boceprevir, Z-FA-FMK, calpain inhibitor XII, and GRL-0496. The luciferase reporter assay is sensitive, rapid, easy to perform, and can readily differentiate cytotoxicity from true inhibition of 3CLpro, a feature that should be advantageous during antiviral drug screening.

2. Materials and Methods

2.1. Plasmids, Cell Lines, and Reagents

Lentiviral plasmids encoding human codon-optimized SARS-CoV-2 3CLpro (wild-type [WT] or C145A) were generous gifts from Nevan Korgan (pLVX-EF1α-SARS-CoV-2-nsp5-2xStrep-IRES-Puro, Addgene # 141370 [WT] and # 141371 [C145A]) [29]. NanoBiT plasmids encoding SmBiT (S) or LgBiT (L) were purchased from Promega (Madison, WI, USA). The 2xStrep tag on the C-terminal end of 3CLpro was replaced with P2A from porcine teschovirus-1 (GSGATNFSLLKQAGDVEENPGP), S, L, and green fluorescent protein (GFP) using NdeI and BamHI restriction sites. S and L were separated by two flexible linkers (each with the sequence GGGSGGGSGGGS) surrounding a SARS-CoV-2 nsp4-nsp5 cleavage site (TSAVLQ↓SGFRKM, with the arrow indicating the cleavage site). 293T human embryonic kidney cells were obtained from American Type Culture Collection (Manassas, VA, USA; catalog # CRL-3216) and maintained in Dulbecco’s modified Eagle’s medium supplemented with 10% fetal bovine serum (FBS), 100 U/mL penicillin, and 100 µg/mL streptomycin. GC376, boceprevir, telaprevir, simeprevir, asunaprevir, grazoprevir, ebselen, Z-FA-FMK, TBB, masitinib, and doxorubicin were obtained from Selleck Chemicals (Houston, TX, USA). Calpain inhibitor XII, walrycin B, DA-3003-1, MG-115, suramin, and quinacrine were obtained from Cayman Chemical (Ann Arbor, MI, USA). Calpain inhibitor II, penta-O-galloyl-beta-D-glucose hydrate, baicalein, and tafenoquine were obtained from Sigma (St. Louis, MO, USA). GRL-0496 was obtained from Focus Biomolecules (Plymouth Meeting, PA, USA). MK 0893 was obtained from APExBIO (Houston, TX, USA), and MAC 5576 was obtained from Thermo Fisher Scientific (Waltham, MA, USA). HIV protease inhibitors were obtained from the NIH AIDS Reagent Program, Division of AIDS, NIAID, NIH (Bethesda, MD, USA).

2.2. Transfections and Luciferase Assays

For comparison of luciferase reporter constructs, 293T cells were seeded into solid white 96-well plates, using 10,000 cells in 0.1 mL medium per well. Cells were transfected 24 h later with 100 ng or 20 ng 3CLpro plasmid, using 0.25 µL TransIT-LT1/well (Mirus Bio; Madison, WI, USA). For the transfections of 20 ng plasmid, 80 ng carrier DNA (Promega; Madison, WI, USA) was included to normalize the total amount of DNA. Dimethyl sulfoxide (DMSO) or 100 µM GC376 was added 4 h post-transfection. Luciferase activity was measured 30 h post-transfection using the Nano-Glo Live Cell Assay System (Promega), following the manufacturer’s instructions. Briefly, the 96-well plates, Opti-Mem I reduced serum medium (Life Technologies; Carlsbad, CA, USA), Nano-Glo dilution buffer, and Nano-Glo live cell substrate were equilibrated to room temperature for 30 min. Next, the medium on the cells was aspirated and replaced with 100 µL Opti-Mem I, 23.75 µL Nano-Glo dilution buffer, and 1.25 µL Nano-Glo live cell substrate/well. The plate was mixed on an orbital shaker for 15 s and immediately read using a SpectraMax iD3 (Molecular Devices; San Jose, CA, USA) and an integration time of 400 ms. Compounds were tested for inhibition of 3CLpro following a similar protocol with several small modifications. The 293T cells were seeded into 6-well plates, using 330,000 cells in 2 mL medium per well. Cells were transfected 24 h later with 2 µg 3CLpro plasmid and 5 µL TransIT-LT1 per well. Cells were seeded 4 h post-transfection into solid white 96-well plates, using 10,000 cells in 0.1 mL medium per well. Compounds (or DMSO) were immediately added to final concentrations ranging from 0.1 to 100 µM. Untransfected cells with DMSO, transfected cells with DMSO, and transfected cells with 100 µM GC376 were included as controls on every plate. Luciferase activity was measured 30 h post-transfection as described. The average luciferase activity of transfected cells with DMSO was subtracted from the values obtained for the other samples, and the data were further normalized to the value obtained for transfected cells with 100 µM GC376, which was set to 100%. NanoBiT data for the DMSO control were graphed at a drug concentration of 10 nM so that they could be displayed on a log axis.

2.3. Cell Viability Analysis

Cell viability was measured using the CellTiter-Glo 2.0 cell viability assay (Promega), following the manufacturer’s instructions. Briefly, 293T cells were seeded, transfected, and treated with compounds as described for the NanoBiT luciferase assay. Cell viability was measured 30 h post-transfection by equilibrating the 96-well plate and CellTiter-Glo 2.0 reagent to room temperature for 30 min and adding 100 µL reagent per well. The plate was mixed for 2 min on an orbital shaker, and firefly luciferase activity was measured 10 min later using a SpectraMax iD3 (Molecular Devices) and an integration time of 400 ms. The average luciferase activity in wells containing only media was subtracted from the values obtained for the other samples, and the data were further normalized to the value obtained for transfected cells with DMSO, which was set to 100%.

2.4. Western Blotting

The 293T cells were seeded into 6-well plates, using 330,000 cells in 2 mL medium per well. Cells were transfected 24 h later with 2 or 0.4 µg 3CLpro plasmid and 5 µL TransIT-LT1/well (Mirus Bio). For the transfections of 0.4 µg plasmid, 1.6 µg carrier DNA (Promega) was included to normalize the total amount of DNA. DMSO or 100 µM GC376 was added 4 h post-transfection. At 30 h post-transfection, the cells were washed with phosphate-buffered saline and lysed on plate using CelLytic solution (Sigma) containing cOmplete EDTA-free protease inhibitor cocktail tablets (Roche; Basel, Switzerland). Cell lysates were centrifuged at 16,000× g for 10 min at 4 °C, and supernatants were transferred to new tubes. Lastly, an appropriate volume of 4× Laemmli sample buffer (Bio-Rad; Hercules, CA, USA) containing 10% beta-mercaptoethanol was added to cell lysates, and the samples were incubated at 99 °C for 5 min. SDS-PAGE was performed using 4–20% Criterion TGX precast gels (Bio-Rad). After transfer to PVDF membranes, blots were probed with 1:5000 rabbit anti-SARS-CoV 3CLpro (Rockland Immunochemicals; Limerick, PA, USA; catalog #200-401-A51), mouse anti-P2A (Novus Biologicals; Littleton, CO, USA; catalog #NBP2-59627), mouse anti-L (Promega; catalog #N710A), or rabbit anti-GFP antibodies (Invitrogen; Carlsbad, CA, USA; catalog #A6455). These were followed with incubation of either goat anti-mouse IgG (IRDye 680CW; LI-COR; Lincoln, NE, USA) or goat anti-rabbit IgG (IRDye 800RD, LI-COR) secondary antibodies at 1:10,000 dilutions. Blots were also probed with mouse anti-HSP90 antibody (Santa Cruz Biotechnology; Dallas, TX, USA; catalog #SC-69703) at a dilution of 1:10,000 as a loading control. Western blots were imaged and quantified using the Odyssey CLx infrared imaging system and Image Studio Lite v5.2 (LI-COR). To calculate the percentage of uncleaved S-L-GFP reporter, background subtracted intensities of S-L-GFP bands were divided by the sum intensities of S-L-GFP and L-GFP bands. For each experiment, at least three independent blots were performed.

2.5. Statistical Analysis

Statistical analysis was performed in GraphPad Prism v8.4.3 (GraphPad Software, San Diego, CA, USA). In the first figure, data were analyzed by mixed-effects analysis, followed by Dunnett’s post-test and correction for multiple comparisons. The treatment and vector type were considered fixed effects, whereas the biological replicate was considered a random effect. In the remaining figures, data were analyzed by one-way ANOVA followed by Dunnett’s post-test and correction for multiple comparisons. For all experiments, at least three independent biological replicates were performed.

3. Results

3.1. Development of Cell-Based Luciferase Complementation Reporters to Detect Inhibition of SARS-CoV-2 3CLpro Activity

To develop a reporter for inhibition of SARS-CoV-2 3CLpro activity, we chose to use NanoBiT, a complementation reporter engineered from NanoLuc luciferase [30]. NanoBiT is comprised of two protein fragments: Large BiT (L, 17.6 kDa) and Small BiT (S,1.3 kDa). When the L and S fragments are in close proximity, they can complement each other to form a functional luciferase protein. However, when the two fragments are not in close proximity, they do not associate, leading to low luciferase activity. Importantly, this system was engineered to be fully reversible. We connected the L and S fragments with a linker containing the SARS-CoV-2 3CLpro nsp4-nsp5 cleavage site (TSAVLQ↓SGFRKM, with the arrow indicating the cleavage site) (Figure 1A). We hypothesized that when 3CLpro is expressed, it would cleave the linker, causing the two NanoBiT fragments to dissociate and resulting in low luciferase activity. In contrast, when 3CLpro is inhibited (e.g., by a small molecule) or inactivated (e.g., by mutation of its catalytic site), NanoBiT would remain intact, leading to high luciferase activity. Thus, 3CLpro inhibition should result in a gain, rather than a loss, of luciferase reporter signal. In principle, this feature would also allow the discrimination of 3CLpro inhibition from cytotoxicity, thereby avoiding false positives from cytotoxic compounds.
To test this strategy, we constructed lentiviral vectors that encode both SARS-CoV-2 3CLpro and reporter cassettes (Figure 1B). Four different reporter cassettes were designed by fusing green fluorescent protein (GFP), L, and S in various orders. GFP was included as a marker for transfection and expression; previous studies have indicated that both GFP and NanoLuc luciferase remain functional when fused together [31,32]. L and S were separated by a nsp4-nsp5 cleavage site surrounded by two flexible linkers. 3CLpro and the reporter cassette were separated by P2A, a small “self-cleaving” peptide from porcine teschovirus-1 [33]. After cleavage, most of the P2A sequence should remain on the C-terminus of 3CLpro; a previous report indicated that SARS-CoV-2 3CLpro maintains high activity when tagged on the C-terminus [20]. These lentiviral vectors should express equimolar amounts of 3CLpro and reporter. In addition to vectors encoding wild-type (WT) 3CLpro, we also constructed four control vectors encoding 3CLpro C145A, a catalytically inactive mutant [29].
To analyze expression and cleavage, the lentiviral vectors were transfected into 293T cells, followed by the addition of DMSO (control) or 100 µM GC376, a small molecule inhibitor of SARS-CoV-2 3CLpro [18,19,20,34]. GFP was clearly detectable by fluorescence microscopy and flow cytometry in cells transfected with each vector (data not shown), indicating that GFP remained functional when fused to L and S. Western blotting was performed on cell lysates collected 30 h post-transfection (Figure 1C). Expression of 3CLpro and reporters was similar between vectors with different reporter cassettes. However, inhibition or inactivation of 3CLpro with GC376 or the C145A mutation, respectively, resulted in increased expression of both 3CLpro and reporters. Others have also observed that 3CLpro C145A is expressed better than 3CLpro WT [29]. In the absence of GC376, 3CLpro WT cleaved each GFP-NanoBiT reporter, generating products of the expected sizes (arrows in Figure 1C). As expected, these cleavage products were not detected with 3CLpro C145A. In the presence of GC376, cleavage by 3CLpro WT was partially prevented, generating an intermediate phenotype.
To determine if 3CLpro inhibition led to increased luciferase activity, we measured NanoBiT luciferase activity in transfected 293T cells (Figure 1D). For 3 out of 4 reporter cassettes (all except L-S-GFP), compared to the 3CLpro WT control, we observed significant increases in luciferase activity for both 3CLpro WT with GC376 and 3CLpro C145A. The lentiviral vector with the S-L-GFP reporter cassette had the greatest enhancement of luciferase activity for both 3CLpro WT with GC376 (9.7-fold difference) and 3CLpro C145A (10.6-fold difference). To determine if increased luciferase activity was due to higher levels of protein expression, we also transfected five-fold less of the 3CLpro C145A vector DNAs, which resulted in comparable protein expression between 3CLpro WT and C145A vectors (Figure 1C). For three out of four reporter cassettes (all except L-S-GFP), we observed significantly higher luciferase activity for 3CLpro C145A than for 3CLpro WT. Thus, 3CLpro inhibition led to enhanced luciferase activity due to the combined effects of increased expression and reduced 3CLpro-mediated cleavage. As the lentiviral vector with the S-L-GFP reporter cassette had the best signal-to-background ratio, this vector was used for all further analysis.

3.2. The Luciferase Complementation Assay Accurately Reflects Concentration-Dependent Inhibition of SARS-CoV-2 3CLpro-Mediated Cleavage

To evaluate whether the luciferase complementation assay reflects the concentration-dependent inhibition of 3CLpro-mediated cleavage, we treated cells transfected with the S-L-GFP lentiviral vector with varying concentrations of GC376 (ranging from 0.1 to 200 µM). By Western blotting, we found that only 15% of the S-L-GFP reporter remained uncleaved in the absence of GC376 (Figure 2A, lane 2). In the presence of GC376, we observed concentration-dependent inhibition of 3CLpro-mediated cleavage, with half maximal effective concentration (EC50) values of 12.6 µM as detected by anti-GFP antibody or 23.2 µM as detected by anti-L antibody (Figure 2B). Maximally (i.e., at 200 µM), GC376 prevented cleavage of ~70% of the S-L-GFP reporter. We next determined the effects of varying concentrations of GC376 on luciferase reporter activity and cell viability (Figure 2C). GC376 resulted in a concentration-dependent increase in luciferase activity, with an EC50 of 23.8 µM, which is close to the values obtained by Western blotting. The increased luciferase activity observed with GC376 was statistically significant (relative to no drug) at all concentrations tested except for the lowest concentration (0.1 µM). We also found that GC376 had no significant effect on cell viability up to 100 µM (Figure 2C). These data further confirm the validity of the luciferase complementation assay for detecting inhibition of SARS-CoV-2 3CLpro.

3.3. The Luciferase Complementation Assay Can Easily Distinguish 3CLpro Inhibition from Cytotoxicity

In loss-of-function assays, cytotoxic compounds could be identified as hits, leading to false positives. In contrast, our luciferase complementation assay is a gain-of-function assay that should readily distinguish cytotoxic compounds from true inhibitors of SARS-CoV-2 3CLpro. To determine whether cytotoxicity could lead to false positives in our assay, we examined the impact of doxorubicin, a DNA topoisomerase II inhibitor and anti-cancer drug, on luciferase reporter activity and cell viability in 293T cells transfected with the S-L-GFP lentiviral vector (Figure 3). We found that doxorubicin resulted in a concentration-dependent decrease in cell viability, with a 50% cytotoxic concentration (CC50) of 3.7 µM. However, doxorubicin did not result in any significant change in luciferase reporter activity. These results clearly demonstrate that the luciferase complementation reporter assay can differentiate between 3CLpro inhibition and cytotoxicity. This advantageous feature should eliminate many false positive hits caused by cytotoxic compounds when screening small molecules for activity against 3CLpro.

3.4. HIV Protease Inhibitors Do Not Block SARS-CoV-2 3CLpro Activity in a Cell-Based Luciferase Complementation Assay

Human immunodeficiency virus (HIV) protease inhibitors, particularly lopinavir and nelfinavir, have been reported to inhibit the replication of SARS-CoV-2 in cell culture [35,36,37]. However, other groups have reported that HIV protease inhibitors do not block SARS-CoV-2 3CLpro in vitro, even at the maximal concentration tested (20 or 100 µM) [18,19,20]. In one recent report, the apparent anti-SARS-CoV-2 activity of lopinavir and nelfinavir in cell culture was found to be an artifact due to cytotoxicity [38]. Using our luciferase complementation reporter assay, we determined the effects of nine FDA-approved HIV protease inhibitors, including lopinavir and nelfinavir, on SARS-CoV-2 3CLpro activity in cells. We transfected 293T cells with the S-L-GFP lentiviral vector, added DMSO (control) or varying concentrations of inhibitors, and measured luciferase activity and cell viability 30 h post-transfection. Results from three independent experiments showed that none of the HIV protease inhibitors resulted in significant enhancement of luciferase reporter activity (Figure 4, Table 1). Instead, for most of the inhibitors, there was a statistically significant decrease in luciferase activity at the highest concentration tested (100 µM), presumably due to cell death and loss of background luminescence. For indinavir and amprenavir, the least toxic inhibitors, there was no significant increase or decrease in luciferase activity. These results agree with previous findings from in vitro assays [18,19,20] and suggest that the possible inhibition of SARS-CoV-2 replication in cell culture by HIV protease inhibitors is due to cytotoxicity and/or off-target effects (see Discussion). These results also further demonstrate that our luciferase complementation assay can easily distinguish 3CLpro inhibition from cytotoxicity.

3.5. Four Additional Compounds Inhibit SARS-CoV-2 3CLpro in a Cell-Based Luciferase Complementation Assay: Boceprevir, Z-FA-FMK, Calpain Inhibitor XII, and GRL-0496

Although many inhibitors of SARS-CoV-2 3CLpro have been reported in the literature, their ability to inhibit 3CLpro in cells is uncertain due to issues such as cell permeability, cytotoxicity, and off-target effects. Using our newly developed luciferase complementation reporter assay, we determined the effects of 22 additional compounds on SARS-CoV-2 3CLpro activity and cell viability. These compounds included five hepatitis C virus (HCV) protease inhibitors, a SARS-CoV 3CLpro inhibitor (GRL-0496), three host/cellular cysteine protease inhibitors, and compounds of various other classes (Table 1). Most of these compounds have been reported to inhibit SARS-CoV-2 3CLpro in vitro and possibly SARS-CoV-2 virus replication in cell culture. However, we also included several compounds that do not inhibit SARS-CoV-2 3CLpro in vitro (telaprevir, asunaprevir, grazoprevir) [18,20] or that inhibit SARS-CoV-2 3CLpro in vitro but not virus replication in cell culture (MAC 5576) [39]. Results from three independent experiments showed that, of the 22 compounds tested, only four clearly inhibited SARS-CoV-2 3CLpro in our luciferase complementation reporter assay: boceprevir, Z-FA-FMK, calpain inhibitor XII, and GRL-0496 (Figure 5, Table 1). Of these, boceprevir, an FDA-approved HCV protease inhibitor, and Z-FA-FMK, originally developed as an inhibitor of cellular cysteine proteases and recently identified by high-throughput screening as an inhibitor of SARS-CoV-2 3CLpro [24,40], had the greatest effect on 3CLpro activity and little cytotoxicity. Calpain inhibitor XII, but not II, also inhibited SARS-CoV-2 3CLpro in the cell-based luciferase assay, but to a lesser extent (Figure 5, Table 1). As the names imply, these compounds were originally developed as inhibitors of the calpain family of cellular cysteine proteases [41,42]. Calpain inhibitors II and XII have been proposed to have a dual mechanism of action against SARS-CoV-2 in cell culture: inhibition of 3CLpro and inhibition of host proteases (such as calpains and cathepsin L) that are important for SARS-CoV-2 entry in certain cell types (see Discussion) [43]. GRL-0496 also inhibited SARS-CoV-2 3CLpro in cells, but its activity and cytotoxicity were poorly separated, such that luciferase reporter activity actually decreased from 25 to 100 µM (Figure 5, Table 1). GRL-0496 was originally developed as an inhibitor of SARS-CoV 3CLpro [44] but has also been found to be active against SARS-CoV-2 [45]. None of the remaining compounds had significant activity in our assay (Figure S1, Table 1), despite reported inhibition against SARS-CoV-2 3CLpro in vitro and/or virus replication in cell culture. For these compounds, there are several possible explanations for the observed discrepancies, including poor cell permeability, lack of specificity, alternative mechanisms of action (i.e., off-target effects), and excessive cytotoxicity (see Discussion).

4. Discussion

Although SARS-CoV-2 3CLpro is widely considered an attractive target for antiviral drugs, the development of inhibitors has been impeded by a lack of cell-based assays that can be performed in a BSL-2 setting and are amenable to high-throughput screening. While many groups have performed in vitro screens using purified protein, prospective compounds often lack activity in cells due to poor permeability or high cytotoxicity. Alternatively, compounds can be assessed by examining their impact on SARS-CoV-2 replication in cell culture, but this approach requires a BSL-3 facility. Recently, a BSL-2 reporter system for identification of SARS-CoV-2 inhibitors in cells was developed [25]. In this assay, called Flip-GFP, 3CLpro-mediated cleavage leads to the formation of functional GFP, whereas inhibition of 3CLpro leads to a loss of GFP fluorescence. One potential limitation of this approach is that it may be difficult to distinguish 3CLpro inhibition from cytotoxicity. Two additional cell-based reporter assays have been described in pre-print manuscripts. Resnick and co-workers have developed a system in which inhibitors are identified by their ability to prevent 3CLpro-mediated cytotoxicity, as measured by crystal violet staining [45]. While simple in principle, approaches based on measuring cytotoxicity could potentially identify compounds that inhibit 3CLpro-mediated toxicity through indirect effects, such as effects on gene expression or protein stability. Lastly, a gain-of-function approach has been described in which 3CLpro and GFP expression are suppressed unless 3CLpro is inhibited [55]. However, the mechanism of the assay has not been defined.
Here, we have developed a novel luciferase complementation reporter assay that circumvents many of the limitations of previously described approaches. The assay is simple, rapid, and has a high signal-to-background ratio. As it represents a gain-of-function approach, the assay can easily distinguish inhibition of 3CLpro from cytotoxicity. Using our assay, none of the FDA-approved HIV protease inhibitors had any effect on SARS-CoV-2 3CLpro activity. These results agree with previous reports in which HIV protease inhibitors lacked detectable activity against 3CLpro in vitro [18,19,20]. In contrast, other early studies found that some HIV protease inhibitors block SARS-CoV-2 replication in cell culture [35,36,37]. However, Hattori and co-workers recently demonstrated that these findings were likely false positives due to compound cytotoxicity [38]. In their study, lopinavir and nelfinavir initially appeared to have antiviral activity when SARS-CoV-2 replication, measured by qRT-PCR, was examined in Vero E6 cells. However, lopinavir and nelfinavir did not reduce the proportion of infected cells as determined by immunofluorescence microscopy, leading the authors to conclude that these compounds have no detectable antiviral activity. The combination of lopinavir and ritonavir has also been tested for COVID-19 in clinical trials, but no significant benefit has been observed [9,56,57]. Our findings reinforce the conclusions of Hattori and co-workers [38] and demonstrate the benefit of using a gain-of-function reporter assay to evaluate the antiviral activity of compounds.
Using the luciferase complementation reporter assay, we confirmed the ability of the small molecule GC376 to inhibit SARS-CoV-2 3CLpro, as reported by others [18,19,20,34]. GC376 was developed as an inhibitor of another coronavirus, feline infectious peritonitis virus, but has broad activity against 3CLpro from many coronaviruses [58]. In our assay, the EC50 of GC376 was slightly higher than those reported by others (Table 1). The cause of this is unclear, but it may be due to differences in experimental conditions (e.g., cell type, time of drug addition, or duration of drug treatment). We screened 22 additional compounds for anti-3CLpro activity, including HCV protease inhibitors, cellular protease inhibitors, and various other compounds reported to inhibit SARS-CoV-2 3CLpro in vitro and/or virus replication in cells (Table 1). Of these, only four compounds exhibited significant activity against 3CLpro in cells: boceprevir, Z-FA-FMK, calpain inhibitor XII, and GRL-0496. Boceprevir is an FDA-approved HCV protease inhibitor reported by two groups to be active against SARS-CoV-2 3CLpro [18,20]. Z-FA-FMK is a preclinical compound originally developed as an inhibitor of cellular cysteine proteases but recently identified as an inhibitor of SARS-CoV-2 3CLpro by high-throughput screening [24,40]. Calpain inhibitor XII is another preclinical compound originally developed as an inhibitor of cellular cysteine proteases [41]. This compound has been proposed to have a dual antiviral mechanism of action: inhibition of 3CLpro and inhibition of the host protease cathepsin L, which is important for SARS-CoV-2 entry in some cell types [43]. Lastly, GRL-0496 is a preclinical compound originally developed as an inhibitor of SARS-CoV 3CLpro [44]. None of these compounds was able to inhibit SARS-CoV-2 3CLpro to the same extent as GC376, however. All four compounds have been previously reported to block SARS-CoV-2 replication in cell culture [18,20,24,45].
For the 18 remaining compounds, we did not detect significant activity against SARS-CoV-2 3CLpro in cells (Figure S1, Table 1). For the other HCV protease inhibitors, we did not expect to observe activity based on in vitro studies [18,20]. For the other compounds, there are several possible explanations for their lack of activity. First, the compounds may be too toxic, which would result in low luciferase activity regardless of whether they inhibited 3CLpro in the luciferase reporter assay. For many of these compounds, their effect on cytotoxicity was not previously determined, or their cytotoxicity was poorly separated from their antiviral activity. Thus, it is not clear whether these compounds truly had activity against SARS-CoV-2 in cell culture. Our findings suggest that they do not, and that their possible antiviral activity may have been an artifact due to cytotoxicity, as described for lopinavir and nelfinavir [38]. Although most of these compounds were reported to inhibit 3CLpro in vitro, they may have done so through non-specific mechanisms, as recently reported for ebselen [59]. Second, the compounds may not be toxic but still fail to inhibit 3CLpro in cells due to poor permeability, metabolization into an inactive form, or a lack of specificity for 3CLpro. For example, MAC 5576 was reported to inhibit SARS-CoV-2 3CLpro in vitro but had no effect on virus replication [39]. Third, some compounds may inhibit SARS-CoV-2 virus replication through a different mechanism of action, such as inhibition of PLpro or cellular proteases. For example, calpain inhibitors II and XII have been reported to have a dual mechanism of action: inhibition of 3CLpro and inhibition of the host protease cathepsin L [43]. Our findings suggest that the primary mechanism of action for calpain inhibitor II is blocking cathepsin L, whereas calpain inhibitor XII may primarily target 3CLpro. Alternative antiviral mechanisms of action have also been proposed for suramin and baicalein [51,52].
In sum, we have developed a novel luciferase complementation reporter assay for identification of SARS-CoV-2 3CLpro inhibitors in living cells. The assay is sensitive, rapid, easy to perform, and can readily differentiate cytotoxicity from 3CLpro inhibition, a powerful feature that should eliminate many false positives during screening. With minor modifications, this assay could potentially be used to identify inhibitors of 3CLpro from other coronaviruses or inhibitors of other proteases that are considered targets for COVID-19 drug development. The widespread distribution of this reporter vector should significantly accelerate efforts to identify more potent 3CLpro inhibitors.

Supplementary Materials

The following are available online at https://www.mdpi.com/1999-4915/13/2/173/s1, Figure S1: Compounds that lack activity against SARS-CoV-2 3CLpro in a cell-based luciferase complementation assay.

Author Contributions

Conceptualization, J.M.O.R., A.D., V.K.P., and W.-S.H.; methodology, J.M.O.R. and A.D.; validation, J.M.O.R. and A.D.; investigation, J.M.O.R., A.D., and O.A.N.; writing—original draft preparation, J.M.O.R.; writing—review and editing, J.M.O.R., A.D., V.K.P., O.A.N., and W.-S.H.; All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Intramural Research Program of the National Institutes of Health (NIH), National Cancer Institute (NCI), Center for Cancer Research, by NIH Intramural Targeted Anti-COVID-19 program (ITAC) grant funding, and by an NIH Intramural AIDS Research Fellowship (to A.D.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or supplementary material.

Acknowledgments

We thank Ryan Burdick and Jianbo Chen for helpful feedback and discussion.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. 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. [Google Scholar]
  2. Zhou, P.; Yang, X.-L.; Wang, X.-G.; Hu, B.; Zhang, L.; Zhang, W.; Si, H.-R.; Zhu, Y.; Li, B.; Huang, C.-L.; et al. A pneumonia outbreak associated with a new coronavirus of probable bat origin. Nature 2020, 579, 270–273. [Google Scholar]
  3. Jackson, L.A.; Anderson, E.J.; Rouphael, N.G.; Roberts, P.C.; Makhene, M.; Coler, R.N.; McCullough, M.P.; Chappell, J.D.; Denison, M.R.; Stevens, L.J.; et al. An mRNA Vaccine against SARS-CoV-2—Preliminary Report. N. Engl. J. Med. 2020, 383, 1920–1931. [Google Scholar]
  4. Mulligan, M.J.; Lyke, K.E.; Kitchin, N.; Absalon, J.; Gurtman, A.; Lockhart, S.; Neuzil, K.; Raabe, V.; Bailey, R.; Swanson, K.A.; et al. Phase I/II study of COVID-19 RNA vaccine BNT162b1 in adults. Nature 2020, 586, 589–593. [Google Scholar]
  5. Widge, A.T.; Rouphael, N.G.; Jackson, L.A.; Anderson, E.J.; Roberts, P.C.; Makhene, M.; Chappell, J.D.; Denison, M.R.; Stevens, L.J.; Pruijssers, A.J.; et al. Durability of Responses after SARS-CoV-2 mRNA-1273 Vaccination. N. Engl. J. Med. 2020, 384, 80–82. [Google Scholar]
  6. Chen, P.; Nirula, A.; Heller, B.; Gottlieb, R.L.; Boscia, J.; Morris, J.; Huhn, G.; Cardona, J.; Mocherla, B.; Stosor, V.; et al. SARS-CoV-2 Neutralizing Antibody LY-CoV555 in Outpatients with Covid-19. N. Engl. J. Med. 2020. [Google Scholar] [CrossRef]
  7. Gordon, C.J.; Tchesnokov, E.P.; Woolner, E.; Perry, J.K.; Feng, J.Y.; Porter, D.P.; Gotte, M. Remdesivir is a direct-acting antiviral that inhibits RNA-dependent RNA polymerase from severe acute respiratory syndrome coronavirus 2 with high potency. J. Biol. Chem. 2020, 295, 6785–6797. [Google Scholar]
  8. Beigel, J.H.; Tomashek, K.M.; Dodd, L.E.; Mehta, A.K.; Zingman, B.S.; Kalil, A.C.; Hohmann, E.; Chu, H.Y.; Luetkemeyer, A.; Kline, S.; et al. Remdesivir for the Treatment of Covid-19 - Final Report. N. Engl. J. Med. 2020, 383, 1813–1826. [Google Scholar]
  9. Pan, H.; Peto, R.; Henao-Restrepo, A.-M.; Preziosi, M.-P.; Sathiyamoorthy, V.; Abdool Karim, Q.; Alejandria, M.M.; Hernandez Garcia, C.; Kieny, M.-P.; Malekzadeh, R.; et al. Repurposed Antiviral Drugs for Covid-19 - Interim WHO Solidarity Trial Results. N. Engl. J. Med. 2020. [Google Scholar] [CrossRef]
  10. Spinner, C.D.; Gottlieb, R.L.; Criner, G.J.; Arribas Lopez, J.R.; Cattelan, A.M.; Soriano Viladomiu, A.; Ogbuagu, O.; Malhotra, P.; Mullane, K.M.; Castagna, A.; et al. Effect of Remdesivir vs Standard Care on Clinical Status at 11 Days in Patients With Moderate COVID-19: A Randomized Clinical Trial. JAMA 2020, 324, 1048–1057. [Google Scholar]
  11. Wang, Y.; Zhang, D.; Du, G.; Du, R.; Zhao, J.; Jin, Y.; Fu, S.; Gao, L.; Cheng, Z.; Lu, Q.; et al. Remdesivir in adults with severe COVID-19: A randomised, double-blind, placebo-controlled, multicentre trial. Lancet 2020, 395, 1569–1578. [Google Scholar]
  12. V'Kovski, P.; Kratzel, A.; Steiner, S.; Stalder, H.; Thiel, V. Coronavirus biology and replication: Implications for SARS-CoV-2. Nat. Rev. Microbiol. 2020, 1–16. [Google Scholar] [CrossRef]
  13. 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 alpha-ketoamide inhibitors. Science 2020, 368, 409–412. [Google Scholar]
  14. Barrila, J.; Bacha, U.; Freire, E. Long-range cooperative interactions modulate dimerization in SARS 3CLpro. Biochemistry 2006, 45, 14908–14916. [Google Scholar]
  15. 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. [Google Scholar]
  16. Li, Q.; Kang, C. Progress in Developing Inhibitors of SARS-CoV-2 3C-Like Protease. Microorganisms 2020, 8, 1250. [Google Scholar]
  17. Ullrich, S.; Nitsche, C. The SARS-CoV-2 main protease as drug target. Bioorg. Med. Chem. Lett. 2020, 30, 127377. [Google Scholar]
  18. 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. [Google Scholar]
  19. Hung, H.-C.; Ke, Y.-Y.; Huang, S.Y.; Huang, P.-N.; Kung, Y.-A.; Chang, T.-Y.; Yen, K.-J.; 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. [Google Scholar]
  20. 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. [Google Scholar]
  21. Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C.; et al. Structure of M(pro) from SARS-CoV-2 and discovery of its inhibitors. Nature 2020, 582, 289–293. [Google Scholar]
  22. 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. [Google Scholar]
  23. 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. [Google Scholar]
  24. 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. [Google Scholar]
  25. Froggatt, H.M.; Heaton, B.E.; Heaton, N.S. Development of a Fluorescence-Based, High-Throughput SARS-CoV-2 3CL(pro) Reporter Assay. J. Virol. 2020, 94, e01265-20. [Google Scholar]
  26. Khan, A.; Ali, S.S.; Khan, M.T.; Saleem, S.; Ali, A.; Suleman, M.; Babar, Z.; Shafiq, A.; Khan, M.; Wei, D.Q. Combined drug repurposing and virtual screening strategies with molecular dynamics simulation identified potent inhibitors for SARS-CoV-2 main protease (3CLpro). J. Biomol. Struct. Dyn. 2020, 1–12. [Google Scholar] [CrossRef]
  27. Khan, M.T.; Ali, A.; Wang, Q.; Irfan, M.; Khan, A.; Zeb, M.T.; Zhang, Y.-J.; Chinnasamy, S.; Wei, D.Q. Marine natural compounds as potents inhibitors against the main protease of SARS-CoV-2-a molecular dynamic study. J. Biomol.Struct. Dyn. 2020, 1–11. [Google Scholar] [CrossRef]
  28. Wang, Q.; Zhao, Y.; Chen, X.; Hong, A. Virtual screening of approved clinic drugs with main protease (3CL(pro)) reveals potential inhibitory effects on SARS-CoV-2. J. Biomol. Struct. Dyn. 2020, 1–11. [Google Scholar] [CrossRef]
  29. Gordon, D.E.; Jang, G.M.; Bouhaddou, M.; Xu, J.; Obernier, K.; White, K.M.; O'Meara, M.J.; Rezelj, V.V.; Guo, J.Z.; Swaney, D.L. A SARS-CoV-2 protein interaction map reveals targets for drug repurposing. Nature 2020, 583, 459–468. [Google Scholar]
  30. Dixon, A.S.; Schwinn, M.K.; Hall, M.P.; Zimmerman, K.; Otto, P.; Lubben, T.H.; Butler, B.L.; Binkowski, B.F.; Machleidt, T.; Kirkland, T.A.; et al. NanoLuc Complementation Reporter Optimized for Accurate Measurement of Protein Interactions in Cells. ACS Chem. Biol. 2016, 11, 400–408. [Google Scholar]
  31. Hou, Y.J.; Okuda, K.; Edwards, C.E.; Martinez, D.R.; Asakura, T.; Dinnon, K.H., 3rd; Kato, T.; Lee, R.E.; Yount, B.L.; Mascenik, T.M.; et al. SARS-CoV-2 Reverse Genetics Reveals a Variable Infection Gradient in the Respiratory Tract. Cell 2020, 182, 429–4461.e14. [Google Scholar]
  32. Schaub, F.X.; Reza, M.S.; Flaveny, C.A.; Li, W.; Musicant, A.M.; Hoxha, S.; Guo, M.; Cleveland, J.L.; Amelio, A.L. Fluorophore-NanoLuc BRET Reporters Enable Sensitive In Vivo Optical Imaging and Flow Cytometry for Monitoring Tumorigenesis. Cancer Res. 2015, 75, 5023–5033. [Google Scholar]
  33. Kim, J.H.; Lee, S.R.; Li, L.H.; Park, H.J.; Park, J.H.; Lee, K.Y.; Kim, M.K.; Shin, B.A.; Choi, S.Y. High cleavage efficiency of a 2A peptide derived from porcine teschovirus-1 in human cell lines, zebrafish and mice. PLoS ONE 2011, 6, e18556. [Google Scholar]
  34. 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. [Google Scholar]
  35. Choy, K.-T.; Wong, A.Y.-L.; Kaewpreedee, P.; Sia, S.F.; Chen, D.; Hui, K.P.Y.; Chu, D.K.W.; Chan, M.C.W.; Cheung, P.P.-H.; Huang, X.; et al. Remdesivir, lopinavir, emetine, and homoharringtonine inhibit SARS-CoV-2 replication in vitro. Antiviral. Res. 2020, 178, 104786. [Google Scholar]
  36. Fintelman-Rodrigues, N.; Sacramento, C.Q.; Ribeiro Lima, C.; Souza da Silva, F.; Ferreira, A.C.; Mattos, M.; de Freitas, C.S.; Cardoso Soares, V.; da Silva Gomes Dias, S.; Temerozo, J.R.; et al. Atazanavir, Alone or in Combination with Ritonavir, Inhibits SARS-CoV-2 Replication and Proinflammatory Cytokine Production. Antimicrob. Agents Chemother. 2020, 64, e00825-20. [Google Scholar]
  37. Yamamoto, N.; Matsuyama, S.; Hoshino, T.; Yamamoto, N. Nelfinavir inhibits replication of severe acute respiratory syndrome coronavirus 2 in vitro. bioRxiv 2020. [Google Scholar] [CrossRef]
  38. Hattori, S.-I.; Higshi-Kuwata, N.; Raghavaiah, J.; Das, D.; Bulut, H.; Davis, D.A.; Takamatsu, Y.; Matsuda, K.; Takamune, N.; Kishimoto, N.; et al. GRL-0920, an Indole Chloropyridinyl Ester, Completely Blocks SARS-CoV-2 Infection. mBio 2020, 11, e01833-20. [Google Scholar]
  39. 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. bioRxiv 2020. [Google Scholar] [CrossRef]
  40. Rasnick, D. Synthesis of peptide fluoromethyl ketones and the inhibition of human cathepsin B. Anal. Biochem. 1985, 149, 461–465. [Google Scholar]
  41. Li, Z.; Ortega-Vilain, A.C.; Patil, G.S.; Chu, D.L.; Foreman, J.E.; Eveleth, D.D.; Powers, J.C. Novel peptidyl alpha-keto amide inhibitors of calpains and other cysteine proteases. J. Med. Chem. 1996, 39, 4089–4098. [Google Scholar]
  42. Sasaki, T.; Kishi, M.; Saito, M.; Tanaka, T.; Higuchi, N.; Kominami, E.; Katunuma, N.; Murachi, T. Inhibitory effect of di- and tripeptidyl aldehydes on calpains and cathepsins. J. Enzyme. Inhib. 1990, 3, 195–201. [Google Scholar]
  43. 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 reveals strategy for developing dual inhibitors against M(pro) and cathepsin L. Sci. Adv. 2020, 6, eabe0751. [Google Scholar] [CrossRef]
  44. Ghosh, A.K.; Gong, G.; Grum-Tokars, V.; Mulhearn, D.C.; Baker, S.C.; Coughlin, M.; Prabhakar, B.S.; Sleeman, K.; Johnson, M.E.; Mesecar, A.D. Design, synthesis and antiviral efficacy of a series of potent chloropyridyl ester-derived SARS-CoV 3CLpro inhibitors. Bioorg. Med. Chem. Lett. 2008, 18, 5684–5688. [Google Scholar]
  45. Resnick, S.J.; Iketani, S.; Hong, S.J.; Zask, A.; Liu, H.; Kim, S.; Melore, S.; Nair, M.S.; Huang, Y.; Tay, N.E.S.; et al. A simplified cell-based assay to identify coronavirus 3CL protease inhibitors. bioRxiv 2020. [Google Scholar] [CrossRef]
  46. Luan, X.; Shang, W.; Wang, Y.; Yin, W.; Jiang, Y.; Feng, S.; Wang, Y.; Liu, M.; Zhou, R.; Zhang, Z.; et al. Structure Basis for Inhibition of SARS-CoV-2 by the Feline Drug GC376. bioRxiv 2020. [Google Scholar] [CrossRef]
  47. De Meyer, S.; Bojkova, D.; Cinatl, J.; Van Damme, E.; Buyck, C.; Van Loock, M.; Woodfall, B.; Ciesek, S. Lack of antiviral activity of darunavir against SARS-CoV-2. Int. J. Infect. Dis. 2020, 97, 7–10. [Google Scholar]
  48. Drayman, N.; Jones, K.A.; Azizi, S.-A.; Froggatt, H.M.; Tan, K.; Maltseva, N.I.; Chen, S.; Nicolaescu, V.; Dvorkin, S.; Furlong, K.; et al. Drug repurposing screen identifies masitinib as a 3CLpro inhibitor that blocks replication of SARS-CoV-2 in vitro. bioRxiv 2020. [Google Scholar] [CrossRef]
  49. Chen, Y.; Yang, W.-H.; Huang, L.-M.; Wang, Y.-C.; Yang, C.-S.; Liu, Y.-L.; Hou, M.-H.; Tsai, C.-L.; Chou, Y.-Z.; Huang, B.-Y.; et al. Inhibition of Severe Acute Respiratory Syndrome Coronavirus 2 main protease by tafenoquine in vitro. bioRxiv 2020. [Google Scholar] [CrossRef]
  50. Dow, G.S.; Luttick, A.; Fenner, J.; Wesche, D.; Yeo, K.R.; Rayner, C. Tafenoquine inhibits replication of SARS-Cov-2 at pharmacologically relevant concentrations in vitro. bioRxiv 2020. [Google Scholar] [CrossRef]
  51. Salgado-Benvindo, C.; Thaler, M.; Tas, A.; Ogando, N.S.; Bredenbeek, P.J.; Ninaber, D.K.; Wang, Y.; Hiemstra, P.S.; Snijder, E.J.; van Hemert, M.J. Suramin Inhibits SARS-CoV-2 Infection in Cell Culture by Interfering with Early Steps of the Replication Cycle. Antimicrob. Agents Chemother. 2020, 64, e00900-20. [Google Scholar]
  52. Huang, S.; Liu, Y.; Zhang, Y.; Zhang, R.; Zhu, C.; Fan, L.; Pei, G.; Zhang, B.; Shi, Y. Baicalein inhibits SARS-CoV-2/VSV replication with interfering mitochondrial oxidative phosphorylation in a mPTP dependent manner. Signal Transduct. Target. Ther. 2020, 5, 266. [Google Scholar]
  53. Liu, H.; Ye, F.; Sun, Q.; Liang, H.; Li, C.; 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. bioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  54. Su, H.; Yao, S.; Zhao, W.; Li, M.; Liu, J.; Shang, W.; Xie, H.; Ke, C.; Gao, M.; Yu, K.; et al. Discovery of baicalin and baicalein as novel, natural product inhibitors of SARS-CoV-2 3CL protease in vitro. bioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  55. Moghadasi, S.A.; Becker, J.T.; Belica, C.; Wick, C.; Brown, W.L.; Harris, R.S. Gain-of-function assay for SARS-CoV-2 M (pro) inhibition in living cells. bioRxiv 2020. [Google Scholar] [CrossRef]
  56. Cao, B.; Wang, Y.; Wen, D.; Liu, W.; Wang, J.; Fan, G.; Ruan, L.; Song, B.; Cai, Y.; Wei, M.; et al. A Trial of Lopinavir-Ritonavir in Adults Hospitalized with Severe Covid-19. N. Engl. J. Med. 2020, 382, 1787–1799. [Google Scholar]
  57. Group, R.C. Lopinavir-ritonavir in patients admitted to hospital with COVID-19 (RECOVERY): A randomised, controlled, open-label, platform trial. Lancet 2020, 396, 1345–1352. [Google Scholar]
  58. Kim, Y.; Lovell, S.; Tiew, K.-C.; Mandadapu, S.-R.; Alliston, K.R.; Battaile, K.P.; Groutas, W.C.; Chang, K.-O. Broad-spectrum antivirals against 3C or 3C-like proteases of picornaviruses, noroviruses, and coronaviruses. J. Virol. 2012, 86, 11754–11762. [Google Scholar]
  59. 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 non-specific promiscuous SARS-CoV-2 main protease inhibitors. bioRxiv 2020. [Google Scholar] [CrossRef]
Figure 1. Development of cell-based luciferase complementation reporters to detect inhibition of SARS-CoV-2 3CLpro activity. (A) Strategy for detection of 3CLpro inhibition using NanoBiT, a luciferase complementation reporter comprised of Large BiT (L) and Small BiT (S). When connected by a linker containing a 3CLpro cleavage site, 3CLpro cleavage should result in a loss of complementation and low luciferase activity. Conversely, 3CLpro inhibitors should prevent cleavage, resulting in high luciferase activity. (B) Lentiviral vectors that express 3CLpro and reporters. P2A, a self-cleaving peptide, was inserted between 3CLpro and reporters. Reporters consisted of GFP linked to L and S in various orders. L and S were separated by the SARS-CoV-2 nsp4-nsp5 cleavage site. LTR, long terminal repeat; PEF1α, EF1α promoter; IRES, internal ribosome entry site; puro, puromycin resistance gene; WPRE, woodchuck hepatitis virus post-transcriptional regulatory element. (C) Western blotting to examine reporter expression and cleavage. Lentiviral vectors expressing 3CLpro WT or C145A (a catalytically inactive mutant) were transfected into 293T cells, and DMSO (control) or the 3CLpro inhibitor GC376 (100 µM) was added. Cell lysates were collected 30 h post-transfection. 3CLpro was detected using anti-SARS-CoV 3CLpro or anti-P2A antibodies, whereas reporters were detected using anti-GFP or anti-L antibodies. Arrows denote cleavage products of expected sizes. For 3CLpro C145A, two different plasmid amounts were transfected: 2 or 0.4 µg. HSP90 was included as a loading control. (D) NanoBiT luciferase activity. 293T cells were transfected (same sample order as in C), and NanoBiT activity (expressed as relative luciferase units, or RLU) was measured 30 h post-transfection. The data represent the mean ± standard deviation of four independent experiments; ns: not significant; * p < 0.05, ** p < 0.01, *** p < 0.001 (relative to 3CLpro WT without drug; mixed effects analysis with Dunnett’s post-test).
Figure 1. Development of cell-based luciferase complementation reporters to detect inhibition of SARS-CoV-2 3CLpro activity. (A) Strategy for detection of 3CLpro inhibition using NanoBiT, a luciferase complementation reporter comprised of Large BiT (L) and Small BiT (S). When connected by a linker containing a 3CLpro cleavage site, 3CLpro cleavage should result in a loss of complementation and low luciferase activity. Conversely, 3CLpro inhibitors should prevent cleavage, resulting in high luciferase activity. (B) Lentiviral vectors that express 3CLpro and reporters. P2A, a self-cleaving peptide, was inserted between 3CLpro and reporters. Reporters consisted of GFP linked to L and S in various orders. L and S were separated by the SARS-CoV-2 nsp4-nsp5 cleavage site. LTR, long terminal repeat; PEF1α, EF1α promoter; IRES, internal ribosome entry site; puro, puromycin resistance gene; WPRE, woodchuck hepatitis virus post-transcriptional regulatory element. (C) Western blotting to examine reporter expression and cleavage. Lentiviral vectors expressing 3CLpro WT or C145A (a catalytically inactive mutant) were transfected into 293T cells, and DMSO (control) or the 3CLpro inhibitor GC376 (100 µM) was added. Cell lysates were collected 30 h post-transfection. 3CLpro was detected using anti-SARS-CoV 3CLpro or anti-P2A antibodies, whereas reporters were detected using anti-GFP or anti-L antibodies. Arrows denote cleavage products of expected sizes. For 3CLpro C145A, two different plasmid amounts were transfected: 2 or 0.4 µg. HSP90 was included as a loading control. (D) NanoBiT luciferase activity. 293T cells were transfected (same sample order as in C), and NanoBiT activity (expressed as relative luciferase units, or RLU) was measured 30 h post-transfection. The data represent the mean ± standard deviation of four independent experiments; ns: not significant; * p < 0.05, ** p < 0.01, *** p < 0.001 (relative to 3CLpro WT without drug; mixed effects analysis with Dunnett’s post-test).
Viruses 13 00173 g001
Figure 2. The luciferase complementation assay accurately reflects concentration-dependent inhibition of SARS-CoV-2 3CLpro-mediated cleavage. (A) Western blotting to examine reporter cleavage. 293T cells were transfected with the lentiviral vector containing 3CLpro WT and the S-L-GFP reporter cassette, and DMSO (control) or GC376 (0.1, 0.4, 1.6, 6.3, 25, 100, or 200 µM) was added. Mock transfection (MT) and 3CLpro C145A S-L-GFP were included as controls. On the GFP and LgBiT (L) panels, the bottom band corresponds to cleaved reporter (L-GFP, 46 kDa, indicated by arrows), while the top band corresponds to uncleaved reporter (S-L-GFP, 49 kDa). (B) Quantification of the percentage of uncleaved reporter from three independent western blots based on GFP (circle) or L (triangle) detection. (C) NanoBiT luciferase activity and cell viability. 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or GC376 (0.1, 0.4, 1.6, 6.3, 25, or 100 µM) was added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of four independent experiments; ns: not significant; * p < 0.05, ** p < 0.01, *** p < 0.001 (relative to no-drug; one-way ANOVA with Dunnett’s post-test).
Figure 2. The luciferase complementation assay accurately reflects concentration-dependent inhibition of SARS-CoV-2 3CLpro-mediated cleavage. (A) Western blotting to examine reporter cleavage. 293T cells were transfected with the lentiviral vector containing 3CLpro WT and the S-L-GFP reporter cassette, and DMSO (control) or GC376 (0.1, 0.4, 1.6, 6.3, 25, 100, or 200 µM) was added. Mock transfection (MT) and 3CLpro C145A S-L-GFP were included as controls. On the GFP and LgBiT (L) panels, the bottom band corresponds to cleaved reporter (L-GFP, 46 kDa, indicated by arrows), while the top band corresponds to uncleaved reporter (S-L-GFP, 49 kDa). (B) Quantification of the percentage of uncleaved reporter from three independent western blots based on GFP (circle) or L (triangle) detection. (C) NanoBiT luciferase activity and cell viability. 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or GC376 (0.1, 0.4, 1.6, 6.3, 25, or 100 µM) was added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of four independent experiments; ns: not significant; * p < 0.05, ** p < 0.01, *** p < 0.001 (relative to no-drug; one-way ANOVA with Dunnett’s post-test).
Viruses 13 00173 g002
Figure 3. The luciferase complementation assay can easily distinguish 3CLpro inhibition from cytotoxicity. To determine if the luciferase complementation reporter assay can differentiate 3CLpro inhibition from cytotoxicity, 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or the anti-cancer drug doxorubicin (0.1–100 µM) was added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of three independent experiments.
Figure 3. The luciferase complementation assay can easily distinguish 3CLpro inhibition from cytotoxicity. To determine if the luciferase complementation reporter assay can differentiate 3CLpro inhibition from cytotoxicity, 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or the anti-cancer drug doxorubicin (0.1–100 µM) was added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of three independent experiments.
Viruses 13 00173 g003
Figure 4. HIV protease inhibitors do not block SARS-CoV-2 3CLpro activity in a cell-based luciferase complementation assay. To determine if HIV protease inhibitors are active against SARS-CoV-2 3CLpro in cells, 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or inhibitors (0.1, 0.4, 1.6, 6.3, 25, or 100 µM) were added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of three independent experiments; * p < 0.05, ** p < 0.01, *** p < 0.001 (relative to no-drug; one-way ANOVA with Dunnett’s post-test).
Figure 4. HIV protease inhibitors do not block SARS-CoV-2 3CLpro activity in a cell-based luciferase complementation assay. To determine if HIV protease inhibitors are active against SARS-CoV-2 3CLpro in cells, 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or inhibitors (0.1, 0.4, 1.6, 6.3, 25, or 100 µM) were added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of three independent experiments; * p < 0.05, ** p < 0.01, *** p < 0.001 (relative to no-drug; one-way ANOVA with Dunnett’s post-test).
Viruses 13 00173 g004
Figure 5. Boceprevir, Z-FA-FMK, calpain inhibitor XII, and GRL-0496, inhibit SARS-CoV-2 3CLpro in a cell-based luciferase complementation assay. To determine if other compounds are active against SARS-CoV-2 3CLpro in cells, 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or compounds (0.1, 0.4, 1.6, 6.3, 25, or 100 µM) were added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of three independent experiments; ** p < 0.01, *** p < 0.001 (relative to no-drug; one-way ANOVA with Dunnett’s post-test).
Figure 5. Boceprevir, Z-FA-FMK, calpain inhibitor XII, and GRL-0496, inhibit SARS-CoV-2 3CLpro in a cell-based luciferase complementation assay. To determine if other compounds are active against SARS-CoV-2 3CLpro in cells, 293T cells were transfected with the S-L-GFP lentiviral vector, and DMSO (control) or compounds (0.1, 0.4, 1.6, 6.3, 25, or 100 µM) were added. NanoBiT activity (circle, left axis) and cell viability (measured with the CellTiter-Glo 2.0 assay; triangle, right axis) were analyzed 30 h post-transfection and are expressed as relative luciferase units (RLU). Cell viability data were normalized to the DMSO sample, which was set to 100%. NanoBiT data were normalized to the DMSO and 100 µM GC376 samples, which were set to 0 and 100%, respectively. The data represent the mean ± standard deviation of three independent experiments; ** p < 0.01, *** p < 0.001 (relative to no-drug; one-way ANOVA with Dunnett’s post-test).
Viruses 13 00173 g005
Table 1. Compounds screened for inhibition of SARS-CoV-2 3CLpro in cells.
Table 1. Compounds screened for inhibition of SARS-CoV-2 3CLpro in cells.
CompoundClass 1Max Inhibition 2EC50 (µM) 3CC50 (µM) 4Reported EC50 (µM) 5Reported CC50 (µM) 5References 5
GC376FIPV PI100%23.8>1000.2–4.5>100–>200[18,19,20,34,39,43,46]
SaquinavirHIV PINANA27.38.844.4[37]
RitonavirHIV PINANA36.18.674.1[37]
IndinavirHIV PINANA>10059.1>81[37]
NelfinavirHIV PINANA11.21.1–3.124.3–53.0[37,38]
AmprenavirHIV PINANA>10031.3>81[37]
LopinavirHIV PINANA29.15.7–19.051.0–74.4[37,38]
AtazanavirHIV PINANA>1002.0–9.4>81, 312[36,37]
TipranavirHIV PINANA85.113.376.8[37]
DarunavirHIV PINANA>10046.4, NA>81–>100[37,47]
BoceprevirHCV PI46%38.6>1001.3–15.6>100–>200[18,20]
TelaprevirHCV PINANA71.5NDND-
SimeprevirHCV PINANA19.6NDND-
AsunaprevirHCV PINANA31.5NDND-
GrazoprevirHCV PINANA57.4NDND-
GRL-0496SARS-CoV PI24%ND53.29.181.0[45]
Z-FA-FMKHost PI68%26.3>1000.13>20[24]
Calpain Inhibitor XIIHost PI26%ND>1000.49>100[20]
Calpain Inhibitor IIHost PINANA>1002.1>100[20]
MG-115Proteasome InNANA11.00.021.1[24]
TBBKinase InNANA10014.18.9[24]
MasitinibKinase InNANA3.03.2ND[48]
DA-3003-1Phosphatase InNANA7.74.57.7[24]
TafenoquineAntiparasiticNANA2.02.5–15.737.2[49,50]
SuraminAntiparasiticNANA>10020.0, NA>20–>5000[24,51]
QuinacrineAntiparasiticNANA2.5NDND-
Walrycin BAntibioticNANA29.03.64.3[24]
MK 0893MiscellaneousNANA42.33.212.6[24]
penta-O-galloyl-beta-D-glucose hydrateMiscellaneousNANA30.412.611.2[24]
MAC 5576MiscellaneousNANA>100NA>100[39]
EbselenMiscellaneousNANA8.04.7ND[21]
BaicaleinMiscellaneousNANA>1001.7–17.6>50–>200[52,53,54]
1 Compound class: FIPV PI, feline infectious peritonitis virus protease inhibitor, HIV PI, human immunodeficiency virus protease inhibitor; HCV PI, hepatitis C virus protease inhibitor, SARS PI, severe acute respiratory syndrome coronavirus protease inhibitor; Host PI, host protease inhibitor; In, inhibitor. 2 Maximum inhibition of SARS-CoV-2 3CLpro in cell-based luciferase complementation assay relative to 100 µM GC376, which was set to 100%; NA, no activity. 3 Half maximal effective concentration (EC50) values, as determined by the SARS-CoV-2 3CLpro luciferase complementation assay. ND, not determined; EC50 values could not be calculated for two compounds due to weak activity (calpain inhibitor XII) or overlapping activity and cytotoxicity (GRL-0496); NA, no activity. 4 50% cytotoxic concentration (CC50) values, as determined by the CellTiter-Glo 2.0 assay. 5 Previously reported EC50 (and corresponding CC50) values for inhibition of SARS-CoV-2 replication in cell culture. Note that, in some cases, EC50 values may be affected by compound toxicity. ND, not determined; NA, no activity.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rawson, J.M.O.; Duchon, A.; Nikolaitchik, O.A.; Pathak, V.K.; Hu, W.-S. Development of a Cell-Based Luciferase Complementation Assay for Identification of SARS-CoV-2 3CLpro Inhibitors. Viruses 2021, 13, 173. https://doi.org/10.3390/v13020173

AMA Style

Rawson JMO, Duchon A, Nikolaitchik OA, Pathak VK, Hu W-S. Development of a Cell-Based Luciferase Complementation Assay for Identification of SARS-CoV-2 3CLpro Inhibitors. Viruses. 2021; 13(2):173. https://doi.org/10.3390/v13020173

Chicago/Turabian Style

Rawson, Jonathan M. O., Alice Duchon, Olga A. Nikolaitchik, Vinay K. Pathak, and Wei-Shau Hu. 2021. "Development of a Cell-Based Luciferase Complementation Assay for Identification of SARS-CoV-2 3CLpro Inhibitors" Viruses 13, no. 2: 173. https://doi.org/10.3390/v13020173

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop