A SARS-CoV-2 protein interaction map reveals targets for drug repurposing

David E Gordon, Gwendolyn M Jang, Mehdi Bouhaddou, Jiewei Xu, Kirsten Obernier, Kris M White, Matthew J O'Meara, Veronica V Rezelj, Jeffrey Z Guo, Danielle L Swaney, Tia A Tummino, Ruth Hüttenhain, Robyn M Kaake, Alicia L Richards, Beril Tutuncuoglu, Helene Foussard, Jyoti Batra, Kelsey Haas, Maya Modak, Minkyu Kim, Paige Haas, Benjamin J Polacco, Hannes Braberg, Jacqueline M Fabius, Manon Eckhardt, Margaret Soucheray, Melanie J Bennett, Merve Cakir, Michael J McGregor, Qiongyu Li, Bjoern Meyer, Ferdinand Roesch, Thomas Vallet, Alice Mac Kain, Lisa Miorin, Elena Moreno, Zun Zar Chi Naing, Yuan Zhou, Shiming Peng, Ying Shi, Ziyang Zhang, Wenqi Shen, Ilsa T Kirby, James E Melnyk, John S Chorba, Kevin Lou, Shizhong A Dai, Inigo Barrio-Hernandez, Danish Memon, Claudia Hernandez-Armenta, Jiankun Lyu, Christopher J P Mathy, Tina Perica, Kala Bharath Pilla, Sai J Ganesan, Daniel J Saltzberg, Ramachandran Rakesh, Xi Liu, Sara B Rosenthal, Lorenzo Calviello, Srivats Venkataramanan, Jose Liboy-Lugo, Yizhu Lin, Xi-Ping Huang, YongFeng Liu, Stephanie A Wankowicz, Markus Bohn, Maliheh Safari, Fatima S Ugur, Cassandra Koh, Nastaran Sadat Savar, Quang Dinh Tran, Djoshkun Shengjuler, Sabrina J Fletcher, Michael C O'Neal, Yiming Cai, Jason C J Chang, David J Broadhurst, Saker Klippsten, Phillip P Sharp, Nicole A Wenzell, Duygu Kuzuoglu-Ozturk, Hao-Yuan Wang, Raphael Trenker, Janet M Young, Devin A Cavero, Joseph Hiatt, Theodore L Roth, Ujjwal Rathore, Advait Subramanian, Julia Noack, Mathieu Hubert, Robert M Stroud, Alan D Frankel, Oren S Rosenberg, Kliment A Verba, David A Agard, Melanie Ott, Michael Emerman, Natalia Jura, Mark von Zastrow, Eric Verdin, Alan Ashworth, Olivier Schwartz, Christophe d'Enfert, Shaeri Mukherjee, Matt Jacobson, Harmit S Malik, Danica G Fujimori, Trey Ideker, Charles S Craik, Stephen N Floor, James S Fraser, John D Gross, Andrej Sali, Bryan L Roth, Davide Ruggero, Jack Taunton, Tanja Kortemme, Pedro Beltrao, Marco Vignuzzi, Adolfo García-Sastre, Kevan M Shokat, Brian K Shoichet, Nevan J Krogan, David E Gordon, Gwendolyn M Jang, Mehdi Bouhaddou, Jiewei Xu, Kirsten Obernier, Kris M White, Matthew J O'Meara, Veronica V Rezelj, Jeffrey Z Guo, Danielle L Swaney, Tia A Tummino, Ruth Hüttenhain, Robyn M Kaake, Alicia L Richards, Beril Tutuncuoglu, Helene Foussard, Jyoti Batra, Kelsey Haas, Maya Modak, Minkyu Kim, Paige Haas, Benjamin J Polacco, Hannes Braberg, Jacqueline M Fabius, Manon Eckhardt, Margaret Soucheray, Melanie J Bennett, Merve Cakir, Michael J McGregor, Qiongyu Li, Bjoern Meyer, Ferdinand Roesch, Thomas Vallet, Alice Mac Kain, Lisa Miorin, Elena Moreno, Zun Zar Chi Naing, Yuan Zhou, Shiming Peng, Ying Shi, Ziyang Zhang, Wenqi Shen, Ilsa T Kirby, James E Melnyk, John S Chorba, Kevin Lou, Shizhong A Dai, Inigo Barrio-Hernandez, Danish Memon, Claudia Hernandez-Armenta, Jiankun Lyu, Christopher J P Mathy, Tina Perica, Kala Bharath Pilla, Sai J Ganesan, Daniel J Saltzberg, Ramachandran Rakesh, Xi Liu, Sara B Rosenthal, Lorenzo Calviello, Srivats Venkataramanan, Jose Liboy-Lugo, Yizhu Lin, Xi-Ping Huang, YongFeng Liu, Stephanie A Wankowicz, Markus Bohn, Maliheh Safari, Fatima S Ugur, Cassandra Koh, Nastaran Sadat Savar, Quang Dinh Tran, Djoshkun Shengjuler, Sabrina J Fletcher, Michael C O'Neal, Yiming Cai, Jason C J Chang, David J Broadhurst, Saker Klippsten, Phillip P Sharp, Nicole A Wenzell, Duygu Kuzuoglu-Ozturk, Hao-Yuan Wang, Raphael Trenker, Janet M Young, Devin A Cavero, Joseph Hiatt, Theodore L Roth, Ujjwal Rathore, Advait Subramanian, Julia Noack, Mathieu Hubert, Robert M Stroud, Alan D Frankel, Oren S Rosenberg, Kliment A Verba, David A Agard, Melanie Ott, Michael Emerman, Natalia Jura, Mark von Zastrow, Eric Verdin, Alan Ashworth, Olivier Schwartz, Christophe d'Enfert, Shaeri Mukherjee, Matt Jacobson, Harmit S Malik, Danica G Fujimori, Trey Ideker, Charles S Craik, Stephen N Floor, James S Fraser, John D Gross, Andrej Sali, Bryan L Roth, Davide Ruggero, Jack Taunton, Tanja Kortemme, Pedro Beltrao, Marco Vignuzzi, Adolfo García-Sastre, Kevan M Shokat, Brian K Shoichet, Nevan J Krogan

Abstract

A newly described coronavirus named severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), which is the causative agent of coronavirus disease 2019 (COVID-19), has infected over 2.3 million people, led to the death of more than 160,000 individuals and caused worldwide social and economic disruption1,2. There are no antiviral drugs with proven clinical efficacy for the treatment of COVID-19, nor are there any vaccines that prevent infection with SARS-CoV-2, and efforts to develop drugs and vaccines are hampered by the limited knowledge of the molecular details of how SARS-CoV-2 infects cells. Here we cloned, tagged and expressed 26 of the 29 SARS-CoV-2 proteins in human cells and identified the human proteins that physically associated with each of the SARS-CoV-2 proteins using affinity-purification mass spectrometry, identifying 332 high-confidence protein-protein interactions between SARS-CoV-2 and human proteins. Among these, we identify 66 druggable human proteins or host factors targeted by 69 compounds (of which, 29 drugs are approved by the US Food and Drug Administration, 12 are in clinical trials and 28 are preclinical compounds). We screened a subset of these in multiple viral assays and found two sets of pharmacological agents that displayed antiviral activity: inhibitors of mRNA translation and predicted regulators of the sigma-1 and sigma-2 receptors. Further studies of these host-factor-targeting agents, including their combination with drugs that directly target viral enzymes, could lead to a therapeutic regimen to treat COVID-19.

Figures

Extended Data Figure 1.. Mutations in overlapping…
Extended Data Figure 1.. Mutations in overlapping coding regions result in premature termination of Orf3a and Orf9c.
(a) Table of the SARS-CoV-2 proteins, including molecular weight, sequence similarity with the SARS-CoV homolog, and inferred function based on the SARS-CoV homolog. (b) Immunoblot detection of 2xStrep tag demonstrates expression of each bait in input samples, as indicated by red arrowhead. For each bait, input from one of the three replicates prepared and affinity purified for mass spectrometry was used for western blot (n=1). For gel source data, see Supplementary Figure 1. (c) Schematic representation of Orf3a (light green) and Orf3b (dark green) overlapping regions. A premature stop codon in Orf3b at position 14 (E14*) corresponds to a Q57H mutation in Orf3a. (d) Schematic of the N gene (red), Orf9b (green) and Orf9c (green) overlapping regions. Two mutations in the N protein (S194L and S197L) correspond to premature stop codons at positions 41 and 44 in Orf9c. The analysis is based on 2,784 sequences obtained from GISAID on April 4, 2020.
Extended Data Figure 2.. Clustering analysis of…
Extended Data Figure 2.. Clustering analysis of AP-MS dataset reveals biological replicates of individual baits are well correlated.
All MS runs (n=3 biologically independent samples, run in replicates) were compared and clustered using artMS (David Jimenez-Morales, Alexandre Rosa Campos, and John Von Dollen, and Nevan Krogan. (2019). artMS: Analytical R tools for Mass Spectrometry. R package version 1.3.9. https://github.com/biodavidjm/artMShttps://github.com/bio-davidjm/artMS). This figure depicts all Pearson’s pairwise correlations between MS runs, and is clustered according to similar correlation patterns. Correlation between replicates for individual baits ranges from 0.46–0.72, and in most cases the experiments corresponding to each bait cluster together, with the exception of a couple of baits with lower numbers of specific host interactions (e.g. E, Nsp2, Orf6, Orf3a, and Orf3b).
Extended Data Figure 3.. Gene ontology biological…
Extended Data Figure 3.. Gene ontology biological process enrichments for SARS-CoV-2 host factors.
We performed GO biological process enrichments (see Methods) for the host factors identified as binding to each SARS-CoV-2 viral protein and represent here the top 5 most significant terms for each viral protein. The p-values were calculated by a hypergeometric test and a false discovery rate was used to account for multiple hypothesis testing.
Extended Data Figure 4.. Pfam protein families…
Extended Data Figure 4.. Pfam protein families enrichments for SARS-CoV-2 host factors.
The enrichment of individual PFAM domains was calculated using a hypergeometric test where success is defined as the number of domains, and the number of trials is the number of individual preys affinity purified with each viral bait. The population values were the numbers of individual PFAM domains in the human proteome. The p-values were not adjusted for multiple testing. To make sure that the p-values that signify enrichment were meaningful, we only considered PFAM domains that have been affinity purified at least three times with any SARS-CoV-2 protein, and which occur in the human proteome at least five times. Here, we show PFAM domains with the lowest p-value for a given viral bait protein.
Extended Data Figure 5.. Lung mRNA expression…
Extended Data Figure 5.. Lung mRNA expression and specificity of SARS-CoV-2-interacting human proteins relative to other proteins.
(a) Scatterplot of the lung mRNA expression (TPM) versus enrichment of lung mRNA expression (lung TPM/median all tissue TPM) for human interacting proteins. Red points denote drug targets that are labeled with their gene names. Points above the horizontal blue line represent interacting proteins that are enriched in lung expression and show how most SARS-CoV-2 interacting proteins tend to be enriched in the lung. (b) Gene expression in the lung of the high-confidence human interacting proteins was observed to be higher when compared to all other proteins (blue=interacting proteins; n=332; median=25.52 TPM; grey=all other proteins; n=13583; median=3.198 TPM, p=.0007 using a t-test).
Extended Data Figure 6.. Candidate targets for…
Extended Data Figure 6.. Candidate targets for the viral Nsp5 protease.
(a) Nsp5 WT and Nsp5 C145A (catalytic dead mutant) interactome. (b) Domain maps of HDAC2 and TRMT1 illustrating predicted cleavage sites (using NetCorona 1.0). HDAC: Histone Deacetylase Domain, NLS: Nuclear Localization Sequence, MTS: Mitochondrial Targeting Sequence, SAM-MT: S-adenosylmethionine-Dependent Methyltransferase Domain. (c) Peptide docking of predicted cleavage peptides into the crystal structure of SARS-CoV Nsp5. (d) Nsp5 cleavage consensus site for SARS-CoV (left) and SARS-CoV-2 (right).
Extended Data Figure 7.. Consensus analysis of…
Extended Data Figure 7.. Consensus analysis of SARS-CoV-2 Orf6 homologs.
(a) Sequence logo of SARS-CoV-2 Orf6 homologs, showing sequence conservation at each position computed from a multiple sequence alignment of 35 sequences. The key methionine M58, and the acidic residues E55, E59, and D61 of the putative NUP98-RAE1 binding motif are shown to be highly conserved. Homology determined from alignments to full length sequences. Colors indicated chemical properties of amino acids: polar (G, S, T, Y, C, green), neutral (Q, N, purple), basic (K, R, H, blue), acidic (D, E, red), and hydrophobic (A, V, L, I, P, W, F, M, black). (b) Multiple sequence alignment of SARS-CoV-2 Orf6 homologs. Query sequence shown at top (sequence 1 ref|YP_009724394.1). Sequence coverage (cov) and percent identity (pid) shown for each homologous sequence.
Extended Data Figure 8.. Viral growth and…
Extended Data Figure 8.. Viral growth and cytotoxicity for compounds tested in New York.
Viral growth (percent infection; red) and cytotoxicity (black) results for compounds tested at Mount Sinai in New York. Zotatifin, hydroxychloroquine, and PB28 were also tested in Median Tissue Culture Infectious Dose assay (TCID50; green). Zotafitin and Midostaurin were tested in two independent experiments and data are shown in two individual panels. Data=mean±SD; all n=3 biologically independent samples.
Extended Data Figure 9.. Virus plaque assays,…
Extended Data Figure 9.. Virus plaque assays, qRT-PCR, and cell viability for compounds tested in Paris.
Plaque assay (viral titer; red), qRT-PCR (viral RNA; blue) and cell viability (Alamar Blue; black) results for compounds tested at the Pasteur Institute in Paris. PF-846 was tested in two independent experiments and data are shown in two individual panels. Data=mean±SD; n=3 biologically independent samples for drug-treated cells and n=5 for PS3061, n=6 for DMSO controls.
Extended Data Figure 10.. Activity of Sigma…
Extended Data Figure 10.. Activity of Sigma ligands.
(a) The drugs cloperastine and clemastine can be readily fit into the agonist-bound structure of the Sigma1 receptor. (b) Compounds tested for antiviral activity with annotated Sigma 1 Receptor (SIGMAR1) and/or Sigma 2 Receptor (SIGMAR2/TMEM97) activity are scatter-plotted. Inhibition pIC50 values of SARS-CoV-2 infection is shown from blue to yellow, mode of functional activity at SIGMA1R is shown by mark shape (upwards triangle: agonist, downwards triangle: antagonist, circle: binding), and pKi values for SIGMA1R and SIGMAR2 are shown along the x- and y-axes. We have not yet tested chloroquine for antiviral activity. E-52862 binding at SIGMAR2/TMEM97 is only reported to be greater than 1 μM. Activities of pimozide and olanzapine at SIGMAR2/TMEM97 have not been reported. Activity of olanzapine at SIGMAR1 is reported to be greater than 5 μM.
Extended Data Figure 11.. Astemizole is a…
Extended Data Figure 11.. Astemizole is a potent Sigma2 Receptor Ligand.
Concentration response curves of astemizole from radio-ligand displacement assays for (a) the Sigma2 (95 nM IC50) and (b) the Sigma1 (1.3 uM IC50) receptors are shown. Data=mean±SEM; n= 4 independent assays on each receptor.
Extended Data Figure 12.. Off-target activities for…
Extended Data Figure 12.. Off-target activities for characteristic Sigma receptor ligands.
Dose response curves against a panel of eight targets that can confer adverse cardiac, respiratory, and dry-mouth effects for chloroquine, hydroxychloroquine, PB28, PD-144418, and clemastine. These results are not meant to represent or replace a comprehensive test against off-target panels, as might commonly be assayed in drug progression for clinical use. The 8 targets include the Alpha-2A adrenergic receptors: Alpha 2A (ADRA2A), Alpha 2B (ADRA2B), and Alpha 2C (ADRA2C); as well as the Muscarinic acetylcholine receptors: M1 (CHRM1), M2, (CHRM2), M3 (CHRM3), M4 (CHRM4) and M5 (CHRM5). Data=mean±SD; all n= 3 biologically independent samples.
Figure 1:. AP-MS workflow for identification of…
Figure 1:. AP-MS workflow for identification of SARS-CoV-2 host protein-protein interactions.
(a) SARS-CoV-2 genome annotation, color intensity is proportional to protein sequence similarity with SARS-CoV homologs (when homologs exist). (b) Experimental workflow for AP-MS studies.
Figure 2:. Global analysis of SARS-CoV-2 protein…
Figure 2:. Global analysis of SARS-CoV-2 protein interactions.
(a) Overview of global analyses performed. (b) Gene Ontology (GO) enrichment analysis was performed on the human interacting proteins of each viral protein, p-values calculated by hypergeometric test and a false discovery rate was used to account for multiple hypothesis testing (Methods). The top GO term of each viral protein was selected for visualization. (c) Degree of differential protein expression for the human interacting proteins (n=332) across human tissues. We obtained protein abundance values for the proteome in 29 human tissues and calculated the median level of abundance for the human interacting proteins (top 16 tissues shown). This was then compared with the abundance values for the full proteome in each tissue and summarized as a Z-score from which a p-value was calculated and false discovery rate was used to account for multiple hypothesis testing. (d) The distribution of correlation of protein level changes during SARS-CoV-2 infection for pairs of viral-human proteins (median is shown) is higher than non-interacting pairs of viral-human proteins (p-value=4.8e-05, Kolmogorov–Smirnov test) The violin plots show each viral to human protein correlation for preys (n=210, min=−0.986, max=0.999, Q1=−0.468, Q2=0.396, Q3=0.850) and non-preys (n=54765, min= −0.999, max=0.999, Q1=−0.599, Q2=0.006, Q3=0.700). (e) Significance of the overlap of human interacting proteins between SARS-CoV-2 and other pathogens using a hypergeometric test (unadjusted for multiple testing). The background gene set for the test consisted of all unique proteins detected by mass spectrometry across all pathogens (n=10,181 proteins).
Figure 3:. SARS-CoV-2 protein-protein interaction network.
Figure 3:. SARS-CoV-2 protein-protein interaction network.
332 high-confidence interactions between 26 SARS-CoV-2 proteins (red diamonds) and human proteins (circles; drug targets: orange; protein complexes: yellow; proteins in the same biological process: blue). Edge color proportional to MiST score; edge thickness proportional to spectral counts. Physical interactions among host proteins (thin black lines) were curated from CORUM, IntAct, and Reactome. An interactive PPI map can be found at kroganlab.ucsf.edu/network-maps. n=3 biologically independent samples.
Figure 4:. The SARS-CoV-2 interactome reveals novel…
Figure 4:. The SARS-CoV-2 interactome reveals novel aspects of SARS-CoV-2 biology and pharmacological targets.
(a) Orf6 interacts with an mRNA nuclear export complex that (i) can be targeted by Selinexor. (ii) Carboxy-terminal peptide of SARS-CoV-2 Orf6 (dark purple) modeled into the binding site of the VSV M protein (yellow)-NUP98 (green)-RAE1 (light purple) complex (PDB ID: 4OWR). Orf6 and M protein residues labeled. (iii) C-terminal sequence of SARS-CoV-2 Orf6, highlighting described trafficking motifs and putative NUP98-RAE1 binding sequence. Chemical properties of amino acids: polar (green), neutral (purple), basic (blue), acidic (red), and hydrophobic (black). (iv) Putative NUP98-RAE1 interaction motifs (negatively charged residues (red) surrounding a conserved methionine (yellow)) from several viral species. (b) Protein N targets stress granule (SG) proteins (i). (ii) Inhibition of Casein kinase II (silmitasertib or TMCB) disrupting SGs. (iii) Translation initiation inhibition: MNK inhibitor (tomivosertib) prevents phosphorylation of eIF4E; 4ER1Cat blocks the interaction of eIF4E with eIF4G. Inhibition of eIF4A (zotatifin) may prevent unwinding of the viral 5’ UTR to thwart its translation. Targeting translation elongation factor-1A ternary complex (ternatin-4) or (iv) Sec61 translocon (PS3061) can prevent viral protein production and membrane insertion, respectively. (c) Orf10 interacts with the CUL2ZYG11B complex (i). (ii) Orf10 predicted secondary structure. (iii) Orf10 might hijack CUL2ZYG11B for ubiquitination of host proteins which can be inhibited by pevonedistat. (d) Envelope (E) protein interacts with bromodomain proteins (i). (ii) Alignment of proteins E of SARS-CoV-2, SARS-CoV and bat CoV with histone H3 and NS1 protein of Influenza A H3N2. Identical and similar amino acids are highlighted. (iii) Bromodomain inhibitors (iBETs) might disrupt the interaction between protein E and BRDs. Figure shows FDA approved drugs (green), clinical candidates (yellow), and preclinical candidates (purple).
Figure 5:. Drug-human target network.
Figure 5:. Drug-human target network.
PPIs of SARS-CoV-2 baits with approved drugs (green), clinical candidates (yellow), and preclinical candidates (purple) with experimental activities against the host proteins (white background) or previously known host factors (grey background) are shown.
Figure 6:. The anti-viral activity of the…
Figure 6:. The anti-viral activity of the translation inhibitors and Sigma receptor ligands.
(a) Schema of viral infectivity assays. (b) The mRNA translation inhibitors (zotatifin, ternatin 4) reduce viral infectivity in a concentration-dependent matter (viral infectivity: red, Anti-NP or Plaque assay, blue, qRT-PCR; cell viability: black, with the initial decline likely reflecting cytostatic and not cytotoxic effects). Data=mean±SD; n=6 biologically independent samples for cell viability data and DMSO controls from Paris; all others n=3. (c) Sigma drugs and preclinical molecules inhibit viral infectivity (colored as in b). Data=mean±SD; n=3 biologically independent samples. (d) TCID50 assays using zotatifin, PB28 and hydroxychloroquine (e) Drugs added before or after high titer virus (MOI=2) had similar antiviral effects (viral infectivity: Anti-NP). Data=mean±SD; n=3 biologically independent samples. (f) SigmaR1 and SigmaR2 are the common targets of the Sigma ligands at the 1 μM activity threshold. (g) Dextromethorphan increases viral titers (viral titer TCID50: red; cell viability: black). Data=mean±SD; n=3 biologically independent samples. (h) SigmaR1/R2 on-target Kd values vs. those for the hERG ion channel. PB28 and PD-144418 show 500 to 5000-fold, while chloroquine and hydroxychloroquine ~30-fold selectivity between these targets. pKi values for hERG vs. SigmaR1 vs. SigmaR2 are: chloroquine (5.5±0.1; 7.1±0.1; 6.3±0.1); hydroxychloroquine (5.6±0.2; 6.9±0.2; 6.0±0.1); PB28 (6.0±0.1; 8.7±0.1; 8.6±0.1); PD-144418 (5.0±0.2; 8.7±0.1; 6.1±0.1); clemastine (6.8±0.2; 8.0±0.1; 7.6±0.1). All data are shown as mean±SD; PB28, clemastine, PD-144418 n=9 biologically independent samples for SigmaR1/R2 and hERG; chloroquine, hydroxychloroquine n=6 for SigmaR1/R2 and n=4 for hERG.

References

    1. Wu F et al. A new coronavirus associated with human respiratory disease in China. Nature 579, 265–269 (2020).
    1. Novel Coronavirus (2019-nCoV) situation reports. .
    1. Wang C, Horby PW, Hayden FG & Gao GF A novel coronavirus outbreak of global health concern. The Lancet vol. 395 470–473 (2020).
    1. Zhu N et al. China Novel Coronavirus Investigating and Research Team. A novel coronavirus from patients with pneumonia in China, 2019. N. Engl. J. Med. 382, 727–733 (2020).
    1. Su S et al. Epidemiology, Genetic Recombination, and Pathogenesis of Coronaviruses. Trends Microbiol. 24, 490–502 (2016).
    1. Gates B Responding to Covid-19 — A Once-in-a-Century Pandemic? New England Journal of Medicine (2020) doi:10.1056/nejmp2003762.
    1. Sheahan TP et al. Comparative therapeutic efficacy of remdesivir and combination lopinavir, ritonavir, and interferon beta against MERS-CoV. Nat. Commun. 11, 222 (2020).
    1. Sheahan TP et al. An orally bioavailable broad-spectrum antiviral inhibits SARS-CoV-2 and multiple endemic, epidemic and bat coronavirus. bioRxiv 2020.03.19.997890 (2020) doi:10.1101/2020.03.19.997890.
    1. Paton J Moderna’s Coronavirus Vaccine Trial Set to Begin This Month. Bloomberg News (2020).
    1. Hoffmann M et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell (2020) doi:10.1016/j.cell.2020.02.052.
    1. Prussia A, Thepchatri P, Snyder JP & Plemper RK Systematic approaches towards the development of host-directed antiviral therapeutics. Int. J. Mol. Sci. 12, 4027–4052 (2011).
    1. Chan JF-W et al. Genomic characterization of the 2019 novel human-pathogenic coronavirus isolated from a patient with atypical pneumonia after visiting Wuhan. Emerg. Microbes Infect. 9, 221–236 (2020).
    1. Fehr AR & Perlman S Coronaviruses: an overview of their replication and pathogenesis. Methods Mol. Biol. 1282, 1–23 (2015).
    1. Teo G et al. SAINTexpress: improvements and additional features in Significance Analysis of INTeractome software. J. Proteomics 100, 37–43 (2014).
    1. Jäger S et al. Global landscape of HIV-human protein complexes. Nature 481, 365–370 (2012).
    1. Bojkova D et al. SARS-CoV-2 infected host cell proteomics reveal potential therapy targets. In Review (2020).
    1. Eckhardt M, Hultquist JF, Kaake RM, Hüttenhain R & Krogan NJ A systems approach to infectious disease. Nat. Rev. Genet. (2020) doi:10.1038/s41576-020-0212-5.
    1. Harcourt J et al. Severe Acute Respiratory Syndrome Coronavirus 2 from Patient with 2019 Novel Coronavirus Disease, United States. Emerg. Infect. Dis. 26, (2020).
    1. Wang D et al. A deep proteome and transcriptome abundance atlas of 29 healthy human tissues. Mol. Syst. Biol. 15, e8503 (2019).
    1. Li M et al. Identification of antiviral roles for the exon-junction complex and nonsense-mediated decay in flaviviral infection. Nat Microbiol (2019) doi:10.1038/s41564-019-0375-z.
    1. Penn BH et al. An Mtb-Human Protein-Protein Interaction Map Identifies a Switch between Host Antiviral and Antibacterial Responses. Mol. Cell 71, 637–648.e5 (2018).
    1. Barnes PJ Role of HDAC2 in the pathophysiology of COPD. Annu. Rev. Physiol. 71, 451–464 (2009).
    1. Xu P et al. NOS1 inhibits the interferon response of cancer cells by S-nitrosylation of HDAC2. J. Exp. Clin. Cancer Res. 38, 483 (2019).
    1. Dewe JM, Fuller BL, Lentini JM, Kellner SM & Fu D TRMT1-Catalyzed tRNA Modifications Are Required for Redox Homeostasis To Ensure Proper Cellular Proliferation and Oxidative Stress Survival. Mol. Cell. Biol. 37, (2017).
    1. Kondo T, Watanabe M & Hatakeyama S TRIM59 interacts with ECSIT and negatively regulates NF-κB and IRF-3/7-mediated signal pathways. Biochem. Biophys. Res. Commun. 422, 501–507 (2012).
    1. Li S, Wang L, Berman M, Kong Y-Y & Dorf ME Mapping a dynamic innate immunity protein interaction network regulating type I interferon production. Immunity 35, 426–440 (2011).
    1. Faria PA et al. VSV disrupts the Rae1/mrnp41 mRNA nuclear export pathway. Mol. Cell 17, 93–102 (2005).
    1. Slaine PD, Kleer M, Smith NK, Khaperskyy DA & McCormick C Stress Granule-Inducing Eukaryotic Translation Initiation Factor 4A Inhibitors Block Influenza A Virus Replication. Viruses 9, (2017).
    1. Reineke LC et al. Casein Kinase 2 Is Linked to Stress Granule Dynamics through Phosphorylation of the Stress Granule Nucleating Protein G3BP1. Mol. Cell. Biol. 37, (2017).
    1. Kindrachuk J et al. Antiviral potential of ERK/MAPK and PI3K/AKT/mTOR signaling modulation for Middle East respiratory syndrome coronavirus infection as identified by temporal kinome analysis. Antimicrob. Agents Chemother. 59, 1088–1099 (2015).
    1. Timms RT et al. A glycine-specific N-degron pathway mediates the quality control of protein N-myristoylation. Science 365, (2019).
    1. Quan B, Seo H-S, Blobel G & Ren Y Vesiculoviral matrix (M) protein occupies nucleic acid binding site at nucleoporin pair (Rae1 • Nup98). Proc. Natl. Acad. Sci. U. S. A. 111, 9127–9132 (2014).
    1. Frieman M et al. Severe acute respiratory syndrome coronavirus ORF6 antagonizes STAT1 function by sequestering nuclear import factors on the rough endoplasmic reticulum/Golgi membrane. J. Virol. 81, 9812–9824 (2007).
    1. Nakagawa K, Narayanan K, Wada M & Makino S Inhibition of Stress Granule Formation by Middle East Respiratory Syndrome Coronavirus 4a Accessory Protein Facilitates Viral Translation, Leading to Efficient Virus Replication. J. Virol. 92, (2018).
    1. Raaben M, Groot Koerkamp MJA, Rottier PJM & de Haan CAM Mouse hepatitis coronavirus replication induces host translational shutoff and mRNA decay, with concomitant formation of stress granules and processing bodies. Cell. Microbiol. 9, 2218–2229 (2007).
    1. Ivanov P, Kedersha N & Anderson P Stress Granules and Processing Bodies in Translational Control. Cold Spring Harb. Perspect. Biol. 11, (2019).
    1. Thompson PA et al. Abstract 2698: eFT226, a potent and selective inhibitor of eIF4A, is efficacious in preclinical models of lymphoma. Experimental and Molecular Therapeutics (2019) doi:10.1158/1538-7445.am2019-2698.
    1. Nakagawa K, Lokugamage KG & Makino S Viral and Cellular mRNA Translation in Coronavirus-Infected Cells. Adv. Virus Res. 96, 165–192 (2016).
    1. Müller C et al. Broad-spectrum antiviral activity of the eIF4A inhibitor silvestrol against corona- and picornaviruses. Antiviral Res. 150, 123–129 (2018).
    1. Cencic R et al. Blocking eIF4E-eIF4G interaction as a strategy to impair coronavirus replication. J. Virol. 85, 6381–6389 (2011).
    1. Knoops K et al. SARS-coronavirus replication is supported by a reticulovesicular network of modified endoplasmic reticulum. PLoS Biol. 6, e226 (2008).
    1. Shah PS et al. Comparative Flavivirus-Host Protein Interaction Mapping Reveals Mechanisms of Dengue and Zika Virus Pathogenesis. Cell 175, 1931–1945.e18 (2018).
    1. Heaton NS et al. Targeting Viral Proteostasis Limits Influenza Virus, HIV, and Dengue Virus Infection. Immunity 44, 46–58 (2016).
    1. Mahon C, Krogan NJ, Craik CS & Pick E Cullin E3 ligases and their rewiring by viral factors. Biomolecules 4, 897–930 (2014).
    1. Soucy TA et al. An inhibitor of NEDD8-activating enzyme as a new approach to treat cancer. Nature 458, 732–736 (2009).
    1. Faivre EJ et al. Selective inhibition of the BD2 bromodomain of BET proteins in prostate cancer. Nature 578, 306–310 (2020).
    1. Filippakopoulos P et al. Histone recognition and large-scale structural analysis of the human bromodomain family. Cell 149, 214–231 (2012).
    1. Marazzi I et al. Suppression of the antiviral response by an influenza histone mimic. Nature 483, 428–433 (2012).
    1. Outcomes Related to COVID-19 Treated With Hydroxychloroquine Among In-patients With Symptomatic Disease - Full Text View - . .
    1. Carelli JD et al. Ternatin and improved synthetic variants kill cancer cells by targeting the elongation factor-1A ternary complex. Elife 4, (2015).
    1. Spicka I et al. Randomized phase III study (ADMYRE) of plitidepsin in combination with dexamethasone vs. dexamethasone alone in patients with relapsed/refractory multiple myeloma. Ann. Hematol. 98, 2139–2150 (2019).
    1. Mitsuda T et al. Sigma-1Rs are upregulated via PERK/eIF2α/ATF4 pathway and execute protective function in ER stress. Biochem. Biophys. Res. Commun. 415, 519–525 (2011).
    1. Si L et al. Human organs-on-chips as tools for repurposing approved drugs as potential influenza and COVID19 therapeutics in viral pandemics. bioRxiv 2020.04.13.039917 (2020) doi:10.1101/2020.04.13.039917.
    1. Riva L et al. A Large-scale Drug Repositioning Survey for SARS-CoV-2 Antivirals. bioRxiv 2020.04.16.044016 (2020) doi:10.1101/2020.04.16.044016.
    1. White NJ Cardiotoxicity of antimalarial drugs. Lancet Infect. Dis. 7, 549–558 (2007).
    1. Armstrong JF et al. The IUPHAR/BPS Guide to PHARMACOLOGY in 2020: extending immunopharmacology content and introducing the IUPHAR/MMV Guide to MALARIA PHARMACOLOGY. Nucleic Acids Res. 48, D1006–D1021 (2020).
    1. Yang D & Leibowitz JL The structure and functions of coronavirus genomic 3’ and 5’ ends. Virus Res. 206, 120–133 (2015).
    1. Yang H et al. The crystal structures of severe acute respiratory syndrome virus main protease and its complex with an inhibitor. Proc. Natl. Acad. Sci. U. S. A. 100, 13190–13195 (2003).
    1. Thiel V et al. Mechanisms and enzymes involved in SARS coronavirus genome expression. J. Gen. Virol. 84, 2305–2315 (2003).
    1. Xie Y et al. GPS-Lipid: a robust tool for the prediction of multiple lipid modification sites. Sci. Rep. 6, 28249 (2016).
    1. Ren J et al. CSS-Palm 2.0: an updated software for palmitoylation sites prediction. Protein Eng. Des. Sel. 21, 639–644 (2008).
    1. Krogh A, Larsson B, von Heijne G & Sonnhammer EL Predicting transmembrane protein topology with a hidden Markov model: application to complete genomes. J. Mol. Biol. 305, 567–580 (2001).
    1. Almagro Armenteros JJ et al. SignalP 5.0 improves signal peptide predictions using deep neural networks. Nat. Biotechnol. 37, 420–423 (2019).
    1. Chiva C et al. QCloud: A cloud-based quality control system for mass spectrometry-based proteomics laboratories. PLoS One 13, e0189209 (2018).
    1. Cox J & Mann M MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372 (2008).
    1. Cox J et al. Accurate Proteome-wide Label-free Quantification by Delayed Normalization and Maximal Peptide Ratio Extraction, Termed MaxLFQ. Mol. Cell. Proteomics 13, 2513–2526 (2014).
    1. Verschueren E et al. Scoring Large-Scale Affinity Purification Mass Spectrometry Datasets with MiST. Curr. Protoc. Bioinformatics 49, 8.19.1–16 (2015).
    1. Giurgiu M et al. CORUM: the comprehensive resource of mammalian protein complexes-2019. Nucleic Acids Res. 47, D559–D563 (2019).
    1. Shannon P et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).
    1. Huttlin EL et al. The BioPlex Network: A Systematic Exploration of the Human Interactome. Cell 162, 425–440 (2015).
    1. Vizcaíno JA et al. ProteomeXchange provides globally coordinated proteomics data submission and dissemination. Nat. Biotechnol. 32, 223–226 (2014).
    1. Deutsch EW et al. The ProteomeXchange consortium in 2017: supporting the cultural change in proteomics public data deposition. Nucleic Acids Res. 45, D1100–D1106 (2017).
    1. Chaudhury S, Lyskov S & Gray JJ PyRosetta: a script-based interface for implementing molecular modeling algorithms using Rosetta. Bioinformatics 26, 689–691 (2010).
    1. Joosten RP, Long F, Murshudov GN & Perrakis A The PDB_REDO server for macromolecular structure model optimization. IUCrJ 1, 213–220 (2014).
    1. Drozdetskiy A, Cole C, Procter J & Barton GJ JPred4: a protein secondary structure prediction server. Nucleic Acids Res. 43, W389–94 (2015).
    1. Sievers F et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539 (2011).
    1. Gaulton A et al. The ChEMBL database in 2017. Nucleic Acids Res. 45, D945–D954 (2017).
    1. Sterling T & Irwin JJ ZINC 15--Ligand Discovery for Everyone. J. Chem. Inf. Model. 55, 2324–2337 (2015).
    1. Schmidt HR, Betz RM, Dror RO & Kruse AC Structural basis for σ receptor ligand recognition. Nat. Struct. Mol. Biol. 25, 981–987 (2018).
    1. Mysinger MM & Shoichet BK Rapid context-dependent ligand desolvation in molecular docking. J. Chem. Inf. Model. 50, 1561–1573 (2010).
    1. Chu DKW et al. Molecular Diagnosis of a Novel Coronavirus (2019-nCoV) Causing an Outbreak of Pneumonia. Clin. Chem. 66, 549–555 (2020).
    1. Huang X-P, Mangano T, Hufeisen S, Setola V & Roth BL Identification of human Ether-à-go-go related gene modulators by three screening platforms in an academic drug-discovery setting. Assay Drug Dev. Technol. 8, 727–742 (2010).
    1. Besnard J et al. Automated design of ligands to polypharmacological profiles. Nature 492, 215–220 (2012).

Source: PubMed

3
Tilaa