Network-based drug repurposing for novel coronavirus 2019-nCoV/SARS-CoV-2

Yadi Zhou, Yuan Hou, Jiayu Shen, Yin Huang, William Martin, Feixiong Cheng, Yadi Zhou, Yuan Hou, Jiayu Shen, Yin Huang, William Martin, Feixiong Cheng

Abstract

Human coronaviruses (HCoVs), including severe acute respiratory syndrome coronavirus (SARS-CoV) and 2019 novel coronavirus (2019-nCoV, also known as SARS-CoV-2), lead global epidemics with high morbidity and mortality. However, there are currently no effective drugs targeting 2019-nCoV/SARS-CoV-2. Drug repurposing, representing as an effective drug discovery strategy from existing drugs, could shorten the time and reduce the cost compared to de novo drug discovery. In this study, we present an integrative, antiviral drug repurposing methodology implementing a systems pharmacology-based network medicine platform, quantifying the interplay between the HCoV-host interactome and drug targets in the human protein-protein interaction network. Phylogenetic analyses of 15 HCoV whole genomes reveal that 2019-nCoV/SARS-CoV-2 shares the highest nucleotide sequence identity with SARS-CoV (79.7%). Specifically, the envelope and nucleocapsid proteins of 2019-nCoV/SARS-CoV-2 are two evolutionarily conserved regions, having the sequence identities of 96% and 89.6%, respectively, compared to SARS-CoV. Using network proximity analyses of drug targets and HCoV-host interactions in the human interactome, we prioritize 16 potential anti-HCoV repurposable drugs (e.g., melatonin, mercaptopurine, and sirolimus) that are further validated by enrichment analyses of drug-gene signatures and HCoV-induced transcriptomics data in human cell lines. We further identify three potential drug combinations (e.g., sirolimus plus dactinomycin, mercaptopurine plus melatonin, and toremifene plus emodin) captured by the "Complementary Exposure" pattern: the targets of the drugs both hit the HCoV-host subnetwork, but target separate neighborhoods in the human interactome network. In summary, this study offers powerful network-based methodologies for rapid identification of candidate repurposable drugs and potential drug combinations targeting 2019-nCoV/SARS-CoV-2.

Keywords: Bioinformatics; Comparative genomics; Proteomic analysis.

Conflict of interest statement

Conflict of interestThe authors declare that they have no conflict of interest.

© The Author(s) 2020.

Figures

Fig. 1. Overall workflow of this study.
Fig. 1. Overall workflow of this study.
Our network-based methodology combines a systems pharmacology-based network medicine platform that quantifies the interplay between the virus–host interactome and drug targets in the human PPI network. a Human coronavirus (HCoV)-associated host proteins were collected from literatures and pooled to generate a pan-HCoV protein subnetwork. b Network proximity between drug targets and HCoV-associated proteins was calculated to screen for candidate repurposable drugs for HCoVs under the human protein interactome model. c, d Gene set enrichment analysis was utilized to validate the network-based prediction. e Top candidates were further prioritized for drug combinations using network-based method captured by the “Complementary Exposure” pattern: the targets of the drugs both hit the HCoV–host subnetwork, but target separate neighborhoods in the human interactome network. f Overall hypothesis of the network-based methodology: (i) the proteins that functionally associate with HCoVs are localized in the corresponding subnetwork within the comprehensive human interactome network; and (ii) proteins that serve as drug targets for a specific disease may also be suitable drug targets for potential antiviral infection owing to common protein–protein interactions elucidated by the human interactome.
Fig. 2. Phylogenetic analysis of coronaviruses.
Fig. 2. Phylogenetic analysis of coronaviruses.
a Phylogenetic tree of coronavirus (CoV). Phylogenetic algorithm analyzed evolutionary conservation among whole genomes of 15 coronaviruses. Red color highlights the recent emergent coronavirus, 2019-nCoV/SARS-CoV-2. Numbers on the branches indicate bootstrap support values. The scale shows the evolutionary distance computed using the p-distance method. b Schematic plot for HCoV genomes. The genus and host information of viruses was labeled on the left by different colors. Empty dark gray boxes represent accessory open reading frames (ORFs). ce The 3D structures of SARS-CoV nsp12 (PDB ID: 6NUR) (c), spike (PDB ID: 6ACK) (d), and nucleocapsid (PDB ID: 2CJR) (e) shown were based on homology modeling. Genome information and phylogenetic analysis results are provided in Supplementary Tables S1 and S2.
Fig. 3. Drug-target network analysis of the…
Fig. 3. Drug-target network analysis of the HCoV–host interactome.
a A subnetwork highlighting the HCoV–host interactome. Nodes represent three types of HCoV-associated host proteins: targetgable (proteins can be targeted by approved drugs or drugs under clinical trials), non-targetable (proteins do not have any known ligands), neighbors (protein–protein interaction partners). Edge colors indicate five types of experimental evidence of the protein–protein interactions (see Materials and methods). 3D three-dimensional structure. b, c KEGG human pathway (b) and gene ontology enrichment analyses (c) for the HCoV-associated proteins.
Fig. 4. A discovered drug-HCoV network.
Fig. 4. A discovered drug-HCoV network.
a A subnetwork highlighting network-predicted drug-HCoV associations connecting 135 drugs and HCoVs. From the 2938 drugs evaluated, 135 ones achieved significant proximities between drug targets and the HCoV-associated proteins in the human interactome network. Drugs are colored by their first-level of the Anatomical Therapeutic Chemical (ATC) classification system code. b A heatmap highlighting network proximity values for SARS-CoV, MERS-CoV, IBV, and MHV, respectively. Color key denotes network proximity (Z-score) between drug targets and the HCoV-associated proteins in the human interactome network. P value was computed by permutation test.
Fig. 5. A discovered drug-protein-HCoV network for…
Fig. 5. A discovered drug-protein-HCoV network for 16 candidate repurposable drugs.
a Network-predicted evidence and gene set enrichment analysis (GSEA) scores for 16 potential repurposable drugs for HCoVs. The overall connectivity of the top drug candidates to the HCoV-associated proteins was examined. Most of these drugs indirectly target HCoV-associated proteins via the human protein–protein interaction networks. All the drug–target-HCoV-associated protein connections were examined, and those proteins with at least five connections are shown. The box heights for the proteins indicate the number of connections. GSEA scores for eight drugs were not available (NA) due to the lack of transcriptome profiles for the drugs. be Inferred mechanism-of-action networks for four selected drugs: b toremifene (first-generation nonsteroidal-selective estrogen receptor modulator), c irbesartan (an angiotensin receptor blocker), d mercaptopurine (an antimetabolite antineoplastic agent with immunosuppressant properties), and e melatonin (a biogenic amine for treating circadian rhythm sleep disorders).
Fig. 6. Network-based rational design of drug…
Fig. 6. Network-based rational design of drug combinations for 2019-nCoV/SARS-CoV-2.
a The possible exposure mode of the HCoV-associated protein module to the pairwise drug combinations. An effective drug combination will be captured by the “Complementary Exposure” pattern: the targets of the drugs both hit the HCoV–host subnetwork, but target separate neighborhoods in the human interactome network. ZCA and ZCB denote the network proximity (Z-score) between targets (Drugs A and B) and a specific HCoV. SAB denotes separation score (see Materials and methods) of targets between Drug A and Drug B. bd Inferred mechanism-of-action networks for three selected pairwise drug combinations: b sirolimus (a potent immunosuppressant with both antifungal and antineoplastic properties) plus dactinomycin (an RNA synthesis inhibitor for treatment of various tumors), c toremifene (first-generation nonsteroidal-selective estrogen receptor modulator) plus emodin (an experimental drug for the treatment of polycystic kidney), and d melatonin (a biogenic amine for treating circadian rhythm sleep disorders) plus mercaptopurine (an antimetabolite antineoplastic agent with immunosuppressant properties).

References

    1. Zumla A, Chan JF, Azhar EI, Hui DS, Yuen KY. Coronaviruses—drug discovery and therapeutic options. Nat. Rev. Drug Discov. 2016;15:327–347. doi: 10.1038/nrd.2015.37.
    1. Paules CI, Marston HD, Fauci AS. Coronavirus infections—more than just the common cold. JAMA. 2020;323:707–708. doi: 10.1001/jama.2020.0757.
    1. de Wit E, van Doremalen N, Falzarano D, Munster VJ. SARS and MERS: recent insights into emerging coronaviruses. Nat. Rev. Microbiol. 2016;14:523–534. doi: 10.1038/nrmicro.2016.81.
    1. de Wilde AH, Snijder EJ, Kikkert M, van Hemert MJ. Host factors in coronavirus replication. Curr. Top. Microbiol. Immunol. 2018;419:1–42.
    1. Chen N, et al. Epidemiological and clinical characteristics of 99 cases of 2019 novel coronavirus pneumonia in Wuhan, China: a descriptive study. Lancet. 2020;395:507–513. doi: 10.1016/S0140-6736(20)30211-7.
    1. Li, Q. et al. Early transmission dynamics in Wuhan, China, of novel coronavirus-infected pneumonia. N. Engl. J. Med. 10.1056/NEJMoa2001316 (2020) (in press).
    1. Greene JA, Loscalzo J. Putting the patient back together—social medicine, network medicine, and the limits of reductionism. N. Engl. J. Med. 2017;377:2493–2499. doi: 10.1056/NEJMms1706744.
    1. Avorn J. The $2.6 billion pill-methodologic and policy considerations. N. Engl. J. Med. 2015;372:1877–1879. doi: 10.1056/NEJMp1500848.
    1. Cheng F. In silico oncology drug repositioning and polypharmacology. Methods Mol. Biol. 2019;1878:243–261. doi: 10.1007/978-1-4939-8868-6_15.
    1. Cheng F, Hong H, Yang S, Wei Y. Individualized network-based drug repositioning infrastructure for precision oncology in the panomics era. Brief Bioinformatics. 2017;18:682–697.
    1. Cheng F, Murray JL, Rubin DH. Drug repurposing: new treatments for Zika virus infection? Trends Mol. Med. 2016;22:919–921. doi: 10.1016/j.molmed.2016.09.006.
    1. Santos R, et al. A comprehensive map of molecular drug targets. Nat. Rev. Drug Discov. 2017;16:19–34. doi: 10.1038/nrd.2016.230.
    1. Cheng F, et al. Network-based approach to prediction and population-based validation of in silico drug repurposing. Nat. Commun. 2018;9:2691. doi: 10.1038/s41467-018-05116-5.
    1. Cheng F, et al. Systems biology-based investigation of cellular antiviral drug targets identified by gene-trap insertional mutagenesis. PLoS Comput. Biol. 2016;12:e1005074. doi: 10.1371/journal.pcbi.1005074.
    1. Yang S, Fu C, Lian X, Dong X, Zhang Z. Understanding human-virus protein-protein interactions using a human protein complex-based analysis framework. mSystems. 2019;4:e00303.
    1. Liu Chuang, Ma Yifang, Zhao Jing, Nussinov Ruth, Zhang Yi-Cheng, Cheng Feixiong, Zhang Zi-Ke. Computational network biology: Data, models, and applications. Physics Reports. 2020;846:1–66. doi: 10.1016/j.physrep.2019.12.004.
    1. Dyall J, et al. Repurposing of clinically developed drugs for treatment of Middle East respiratory syndrome coronavirus infection. Antimicrob. Agents Chemother. 2014;58:4885–4893. doi: 10.1128/AAC.03036-14.
    1. Johansen LM, et al. FDA-approved selective estrogen receptor modulators inhibit Ebola virus infection. Sci. Transl. Med. 2013;5:190ra179. doi: 10.1126/scitranslmed.3005471.
    1. He S, et al. Repurposing of the antihistamine chlorcyclizine and related compounds for treatment of hepatitis C virus infection. Sci. Transl. Med. 2015;7:282ra249. doi: 10.1126/scitranslmed.3010286.
    1. Barrows NJ, et al. A screen of FDA-approved drugs for inhibitors of Zika virus infection. Cell Host. Microbe. 2016;20:259–270. doi: 10.1016/j.chom.2016.07.004.
    1. Xu M, et al. Identification of small-molecule inhibitors of Zika virus infection and induced neural cell death via a drug repurposing screen. Nat. Med. 2016;22:1101–1107. doi: 10.1038/nm.4184.
    1. Cheng F, et al. Prediction of drug-target interactions and drug repositioning via network-based inference. PLoS Comput. Biol. 2012;8:e1002503. doi: 10.1371/journal.pcbi.1002503.
    1. Cheng F, et al. A genome-wide positioning systems network algorithm for in silico drug repurposing. Nat. Commun. 2019;10:3476. doi: 10.1038/s41467-019-10744-6.
    1. Zeng X, et al. deepDR: a network-based deep learning approach to in silico drug repositioning. Bioinformatics. 2019;35:5191–5198. doi: 10.1093/bioinformatics/btz418.
    1. Zeng X, et al. Target identification among known drugs by deep learning from heterogeneous networks. Chem. Sci. 2020;11:1775–1797. doi: 10.1039/C9SC04336E.
    1. Zeng, X. et al. Network-based prediction of drug-target interactions using an arbitrary-order proximity embedded deep forest. Bioinformatics10.1093/bioinformatics/btaa010 (2020) (in press).
    1. Fang, J. et al. Network-based translation of GWAS findings to pathobiology and drug repurposing for Alzheimer’s disease. MedRxiv. 10.1101/2020.01.15.20017160 (2020).
    1. Cheng F, Kovacs IA, Barabasi AL. Network-based prediction of drug combinations. Nat. Commun. 2019;10:1197. doi: 10.1038/s41467-019-09186-x.
    1. Forni D, Cagliani R, Clerici M, Sironi M. Molecular evolution of human coronavirus genomes. Trends Microbiol. 2017;25:35–48. doi: 10.1016/j.tim.2016.09.001.
    1. Kirchdoerfer RN, Ward AB. Structure of the SARS-CoV nsp12 polymerase bound to nsp7 and nsp8 co-factors. Nat. Commun. 2019;10:2342. doi: 10.1038/s41467-019-10280-3.
    1. Li F, Li W, Farzan M, Harrison SC. Structure of SARS coronavirus spike receptor-binding domain complexed with receptor. Science. 2005;309:1864–1868. doi: 10.1126/science.1116480.
    1. Lu R, et al. Genomic characterisation and epidemiology of 2019 novel coronavirus: implications for virus origins and receptor binding. Lancet. 2020;395:565–574. doi: 10.1016/S0140-6736(20)30251-8.
    1. Zhou Peng, Yang Xing-Lou, Wang Xian-Guang, Hu Ben, Zhang Lei, Zhang Wei, Si Hao-Rui, Zhu Yan, Li Bei, Huang Chao-Lin, Chen Hui-Dong, Chen Jing, Luo Yun, Guo Hua, Jiang Ren-Di, Liu Mei-Qin, Chen Ying, Shen Xu-Rui, Wang Xi, Zheng Xiao-Shuang, Zhao Kai, Chen Quan-Jiao, Deng Fei, Liu Lin-Lin, Yan Bing, Zhan Fa-Xian, Wang Yan-Yi, Xiao Geng-Fu, Shi Zheng-Li. A pneumonia outbreak associated with a new coronavirus of probable bat origin. Nature. 2020;579(7798):270–273. doi: 10.1038/s41586-020-2012-7.
    1. Wrapp Daniel, Wang Nianshuang, Corbett Kizzmekia S., Goldsmith Jory A., Hsieh Ching-Lin, Abiona Olubukola, Graham Barney S., McLellan Jason S. Cryo-EM structure of the 2019-nCoV spike in the prefusion conformation. Science. 2020;367(6483):1260–1263. doi: 10.1126/science.abb2507.
    1. Chang CK, Chen CM, Chiang MH, Hsu YL, Huang TH. Transient oligomerization of the SARS-CoV N protein-implication for virus ribonucleoprotein packaging. PLoS ONE. 2013;8:e65045. doi: 10.1371/journal.pone.0065045.
    1. Lamb J, et al. The Connectivity Map: using gene-expression signatures to connect small molecules, genes, and disease. Science. 2006;313:1929–1935. doi: 10.1126/science.1132939.
    1. Lasso G, et al. A structure-informed atlas of human-virus interactions. Cell. 2019;178:1526–1541.e1516. doi: 10.1016/j.cell.2019.08.005.
    1. de Wilde AH, et al. Screening of an FDA-approved compound library identifies four small-molecule inhibitors of Middle East respiratory syndrome coronavirus replication in cell culture. Antimicrob. Agents Chemother. 2014;58:4875–4884. doi: 10.1128/AAC.03011-14.
    1. Zhao Y, et al. Toremifene interacts with and destabilizes the Ebola virus glycoprotein. Nature. 2016;535:169–172. doi: 10.1038/nature18615.
    1. Emmott E, et al. The cellular interactome of the coronavirus infectious bronchitis virus nucleocapsid protein and functional implications for virus biology. J. Virol. 2013;87:9486–9500. doi: 10.1128/JVI.00321-13.
    1. V’Kovski P, et al. Determination of host proteins composing the microenvironment of coronavirus replicase complexes by proximity-labeling. Elife. 2019;8:e42037. doi: 10.7554/eLife.42037.
    1. Moskowitz DW, Johnson FE. The central role of angiotensin I-converting enzyme in vertebrate pathophysiology. Curr. Top. Med. Chem. 2004;4:1433–1454. doi: 10.2174/1568026043387818.
    1. Seko Y. Effect of the angiotensin II receptor blocker olmesartan on the development of murine acute myocarditis caused by coxsackievirus B3. Clin. Sci. 2006;110:379–386. doi: 10.1042/CS20050299.
    1. Erlandson KM, et al. The impact of statin and angiotensin-converting enzyme inhibitor/angiotensin receptor blocker therapy on cognitive function in adults with human immunodeficiency virus infection. Clin. Infect. Dis. 2017;65:2042–2049. doi: 10.1093/cid/cix645.
    1. Wang XJ, et al. Irbesartan, an FDA approved drug for hypertension and diabetic nephropathy, is a potent inhibitor for hepatitis B virus entry by disturbing Na(+)-dependent taurocholate cotransporting polypeptide activity. Antivir. Res. 2015;120:140–146. doi: 10.1016/j.antiviral.2015.06.007.
    1. Ko C, et al. The FDA-approved drug irbesartan inhibits HBV-infection in HepG2 cells stably expressing sodium taurocholate co-transporting polypeptide. Antivir. Ther. 2015;20:835–842. doi: 10.3851/IMP2965.
    1. Hong M, et al. Identification of a novel transcriptional repressor (HEPIS) that interacts with nsp-10 of SARS coronavirus. Viral Immunol. 2008;21:153–162. doi: 10.1089/vim.2007.0108.
    1. McNulty S, Flint M, Nichol ST, Spiropoulou CF. Host mTORC1 signaling regulates andes virus replication. J. Virol. 2013;87:912–922. doi: 10.1128/JVI.02415-12.
    1. Stohr S, et al. Host cell mTORC1 is required for HCV RNA replication. Gut. 2016;65:2017–2028. doi: 10.1136/gutjnl-2014-308971.
    1. Wang CH, et al. Adjuvant treatment with a mammalian target of rapamycin inhibitor, sirolimus, and steroids improves outcomes in patients with severe H1N1 pneumonia and acute respiratory failure. Crit. Care Med. 2014;42:313–321. doi: 10.1097/CCM.0b013e3182a2727d.
    1. Dyall J, et al. Middle East respiratory syndrome and severe acute respiratory syndrome: current therapeutic options and potential targets for novel therapies. Drugs. 2017;77:1935–1966. doi: 10.1007/s40265-017-0830-1.
    1. Karran P, Attard N. Thiopurines in current medical practice: molecular mechanisms and contributions to therapy-related cancer. Nat. Rev. Cancer. 2008;8:24–36. doi: 10.1038/nrc2292.
    1. Chen X, Chou CY, Chang GG. Thiopurine analogue inhibitors of severe acute respiratory syndrome-coronavirus papain-like protease, a deubiquitinating and deISGylating enzyme. Antivir. Chem. Chemother. 2009;19:151–156. doi: 10.1177/095632020901900402.
    1. Cheng KW, et al. Thiopurine analogs and mycophenolic acid synergistically inhibit the papain-like protease of Middle East respiratory syndrome coronavirus. Antivir. Res. 2015;115:9–16. doi: 10.1016/j.antiviral.2014.12.011.
    1. Chen H, Wurm T, Britton P, Brooks G, Hiscox JA. Interaction of the coronavirus nucleoprotein with nucleolar antigens and the host cell. J. Virol. 2002;76:5233–5250. doi: 10.1128/JVI.76.10.5233-5250.2002.
    1. Rainsford KD. Influenza (“Bird Flu”), inflammation and anti-inflammatory/analgesic drugs. Inflammopharmacology. 2006;14:2–9. doi: 10.1007/s10787-006-0002-5.
    1. Garcia CC, Guabiraba R, Soriani FM, Teixeira MM. The development of anti-inflammatory drugs for infectious diseases. Discov. Med. 2010;10:479–488.
    1. Silvestri M, Rossi GA. Melatonin: its possible role in the management of viral infections-a brief review. Ital. J. Pediatr. 2013;39:61. doi: 10.1186/1824-7288-39-61.
    1. Srinivasan V, Mohamed M, Kato H. Melatonin in bacterial and viral infections with focus on sepsis: a review. Recent Pat. Endocr. Metab. Immune Drug Discov. 2012;6:30–39. doi: 10.2174/187221412799015317.
    1. Tan DX, Korkmaz A, Reiter RJ, Manchester LC. Ebola virus disease: potential use of melatonin as a treatment. J. Pineal Res. 2014;57:381–384. doi: 10.1111/jpi.12186.
    1. Tan DX, Manchester LC, Terron MP, Flores LJ, Reiter RJ. One molecule, many derivatives: a never-ending interaction of melatonin with reactive oxygen and nitrogen species? J. Pineal Res. 2007;42:28–42. doi: 10.1111/j.1600-079X.2006.00407.x.
    1. Galano A, Tan DX, Reiter RJ. On the free radical scavenging activities of melatonin’s metabolites, AFMK and AMK. J. Pineal Res. 2013;54:245–257. doi: 10.1111/jpi.12010.
    1. Xiao J, Shimada M, Liu W, Hu D, Matsumori A. Anti-inflammatory effects of eplerenone on viral myocarditis. Eur. J. Heart Fail. 2009;11:349–353. doi: 10.1093/eurjhf/hfp023.
    1. Wang Manli, Cao Ruiyuan, Zhang Leike, Yang Xinglou, Liu Jia, Xu Mingyue, Shi Zhengli, Hu Zhihong, Zhong Wu, Xiao Gengfu. Remdesivir and chloroquine effectively inhibit the recently emerged novel coronavirus (2019-nCoV) in vitro. Cell Research. 2020;30(3):269–271. doi: 10.1038/s41422-020-0282-0.
    1. Tan X, et al. Systematic identification of synergistic drug pairs targeting HIV. Nat. Biotechnol. 2012;30:1125–1130. doi: 10.1038/nbt.2391.
    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. 2015;59:1088–1099. doi: 10.1128/AAC.03659-14.
    1. Lewis EL, Harbour DA, Beringer JE, Grinsted J. Differential in vitro inhibition of feline enteric coronavirus and feline infectious peritonitis virus by actinomycin D. J. Gen. Virol. 1992;73:3285–3288. doi: 10.1099/0022-1317-73-12-3285.
    1. Zhou WB, Ding Q, Chen L, Liu XA, Wang S. Toremifene is an effective and safe alternative to tamoxifen in adjuvant endocrine therapy for breast cancer: results of four randomized trials. Breast Cancer Res. Treat. 2011;128:625–631. doi: 10.1007/s10549-011-1556-5.
    1. Cong Y, et al. MERS-CoV pathogenesis and antiviral efficacy of licensed drugs in human monocyte-derived antigen-presenting cells. PLoS ONE. 2018;13:e0194868. doi: 10.1371/journal.pone.0194868.
    1. Schwarz S, Wang K, Yu W, Sun B, Schwarz W. Emodin inhibits current through SARS-associated coronavirus 3a protein. Antivir. Res. 2011;90:64–69. doi: 10.1016/j.antiviral.2011.02.008.
    1. Ho TY, Wu SL, Chen JC, Li CC, Hsiang CY. Emodin blocks the SARS coronavirus spike protein and angiotensin-converting enzyme 2 interaction. Antivir. Res. 2007;74:92–101. doi: 10.1016/j.antiviral.2006.04.014.
    1. Lambert DW, Clarke NE, Hooper NM, Turner AJ. Calmodulin interacts with angiotensin-converting enzyme-2 (ACE2) and inhibits shedding of its ectodomain. FEBS Lett. 2008;582:385–390. doi: 10.1016/j.febslet.2007.11.085.
    1. Dai J, Inscho EW, Yuan L, Hill SM. Modulation of intracellular calcium and calmodulin by melatonin in MCF-7 human breast cancer cells. J. Pineal Res. 2002;32:112–119. doi: 10.1034/j.1600-079x.2002.1844.x.
    1. Fung TS, Liu DX. Activation of the c-Jun NH2-terminal kinase pathway by coronavirus infectious bronchitis virus promotes apoptosis independently of c-Jun. Cell Death Dis. 2017;8:3215. doi: 10.1038/s41419-017-0053-0.
    1. Biedenkopf N, et al. The natural compound silvestrol is a potent inhibitor of Ebola virus replication. Antivir. Res. 2017;137:76–81. doi: 10.1016/j.antiviral.2016.11.011.
    1. Muller C, et al. Broad-spectrum antiviral activity of the eIF4A inhibitor silvestrol against corona- and picornaviruses. Antivir. Res. 2018;150:123–129. doi: 10.1016/j.antiviral.2017.12.010.
    1. Halder AK, Dutta P, Kundu M, Basu S, Nasipuri M. Review of computational methods for virus-host protein interaction prediction: a case study on novel Ebola-human interactions. Brief. Funct. Genomics. 2018;17:381–391.
    1. Bedi O, Dhawan V, Sharma PL, Kumar P. Pleiotropic effects of statins: new therapeutic targets in drug design. Naunyn Schmiedebergs Arch. Pharmacol. 2016;389:695–712. doi: 10.1007/s00210-016-1252-4.
    1. Li Q, et al. Integrative functional genomics of hepatitis C virus infection identifies host dependencies in complete viral replication cycle. PLoS Pathog. 2014;10:e1004163. doi: 10.1371/journal.ppat.1004163.
    1. Gebre, M., Nomburg, J. L. & Gewurz, B. E. CRISPR-Cas9 genetic analysis of virus-host interactions. Viruses10, 55 (2018).
    1. Kim JH, et al. Acute eosinophilic pneumonia related to a mesalazine suppository. Asia Pac. Allergy. 2013;3:136–139. doi: 10.5415/apallergy.2013.3.2.136.
    1. Gupta A, Gulati S. Mesalamine induced eosinophilic pneumonia. Respir. Med. Case Rep. 2017;21:116–117.
    1. Chiang CW, et al. Translational high-dimensional drug interaction discovery and validation using health record databases and pharmacokinetics models. Clin. Pharmacol. Ther. 2018;103:287–295. doi: 10.1002/cpt.914.
    1. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018;35:1547–1549. doi: 10.1093/molbev/msy096.
    1. Kuleshov MV, et al. Enrichr: a comprehensive gene set enrichment analysis web server 2016 update. Nucleic Acids Res. 2016;44:W90–W97. doi: 10.1093/nar/gkw377.
    1. Law V, et al. DrugBank 4.0: shedding new light on drug metabolism. Nucleic Acids Res. 2014;42:D1091–D1097. doi: 10.1093/nar/gkt1068.
    1. Yang H, et al. Therapeutic target database update 2016: enriched resource for bench to clinical drug target and targeted pathway information. Nucleic Acids Res. 2016;44:D1069–D1074. doi: 10.1093/nar/gkv1230.
    1. Gaulton A, et al. ChEMBL: a large-scale bioactivity database for drug discovery. Nucleic Acids Res. 2012;40:D1100–D1107. doi: 10.1093/nar/gkr777.
    1. Liu TQ, Lin YM, Wen X, Jorissen RN, Gilson MK. BindingDB: a web-accessible database of experimentally determined protein-ligand binding affinities. Nucleic Acids Res. 2007;35:D198–D201. doi: 10.1093/nar/gkl999.
    1. Pawson AJ, et al. The IUPHAR/BPS Guide to PHARMACOLOGY: an expert-driven knowledgebase of drug targets and their ligands. Nucleic Acids Res. 2014;42:D1098–D1106. doi: 10.1093/nar/gkt1143.
    1. Apweiler R, et al. UniProt: the Universal Protein knowledgebase. Nucleic Acids Res. 2004;32:D115–D119. doi: 10.1093/nar/gkh131.
    1. Coordinators NR. Database resources of the National Center for Biotechnology Information. Nucleic Acids Res. 2016;44:D7–D19. doi: 10.1093/nar/gkv1290.
    1. Smith IN, Thacker S, Seyfi M, Cheng F, Eng C. Conformational dynamics and allosteric regulation landscapes of germline PTEN. Am. J. Hum. Genet. 2019;104:861–878. doi: 10.1016/j.ajhg.2019.03.009.
    1. Reghunathan R, et al. Expression profile of immune response genes in patients with severe acute respiratory syndrome. BMC Immunol. 2005;6:2. doi: 10.1186/1471-2172-6-2.
    1. Josset L, et al. Cell host response to infection with novel human coronavirus EMC predicts potential antivirals and important differences with SARS coronavirus. mBio. 2013;4:e00165–00113. doi: 10.1128/mBio.00165-13.
    1. Yuan S, et al. SREBP-dependent lipidomic reprogramming as a broad-spectrum antiviral target. Nat. Commun. 2019;10:120. doi: 10.1038/s41467-018-08015-x.
    1. Sirota M, et al. Discovery and preclinical validation of drug indications using compendia of public gene expression data. Sci. Transl. Med. 2011;3:96ra77. doi: 10.1126/scitranslmed.3001318.

Source: PubMed

3
Se inscrever