SARS-CoV-2 Receptor ACE2 Is an Interferon-Stimulated Gene in Human Airway Epithelial Cells and Is Detected in Specific Cell Subsets across Tissues

Carly G K Ziegler, Samuel J Allon, Sarah K Nyquist, Ian M Mbano, Vincent N Miao, Constantine N Tzouanas, Yuming Cao, Ashraf S Yousif, Julia Bals, Blake M Hauser, Jared Feldman, Christoph Muus, Marc H Wadsworth 2nd, Samuel W Kazer, Travis K Hughes, Benjamin Doran, G James Gatter, Marko Vukovic, Faith Taliaferro, Benjamin E Mead, Zhiru Guo, Jennifer P Wang, Delphine Gras, Magali Plaisant, Meshal Ansari, Ilias Angelidis, Heiko Adler, Jennifer M S Sucre, Chase J Taylor, Brian Lin, Avinash Waghray, Vanessa Mitsialis, Daniel F Dwyer, Kathleen M Buchheit, Joshua A Boyce, Nora A Barrett, Tanya M Laidlaw, Shaina L Carroll, Lucrezia Colonna, Victor Tkachev, Christopher W Peterson, Alison Yu, Hengqi Betty Zheng, Hannah P Gideon, Caylin G Winchell, Philana Ling Lin, Colin D Bingle, Scott B Snapper, Jonathan A Kropski, Fabian J Theis, Herbert B Schiller, Laure-Emmanuelle Zaragosi, Pascal Barbry, Alasdair Leslie, Hans-Peter Kiem, JoAnne L Flynn, Sarah M Fortune, Bonnie Berger, Robert W Finberg, Leslie S Kean, Manuel Garber, Aaron G Schmidt, Daniel Lingwood, Alex K Shalek, Jose Ordovas-Montanes, HCA Lung Biological Network. Electronic address: lung-network@humancellatlas.org, HCA Lung Biological Network, Nicholas Banovich, Pascal Barbry, Alvis Brazma, Tushar Desai, Thu Elizabeth Duong, Oliver Eickelberg, Christine Falk, Michael Farzan, Ian Glass, Muzlifah Haniffa, Peter Horvath, Deborah Hung, Naftali Kaminski, Mark Krasnow, Jonathan A Kropski, Malte Kuhnemund, Robert Lafyatis, Haeock Lee, Sylvie Leroy, Sten Linnarson, Joakim Lundeberg, Kerstin Meyer, Alexander Misharin, Martijn Nawijn, Marko Z Nikolic, Jose Ordovas-Montanes, Dana Pe'er, Joseph Powell, Stephen Quake, Jay Rajagopal, Purushothama Rao Tata, Emma L Rawlins, Aviv Regev, Paul A Reyfman, Mauricio Rojas, Orit Rosen, Kourosh Saeb-Parsy, Christos Samakovlis, Herbert Schiller, Joachim L Schultze, Max A Seibold, Alex K Shalek, Douglas Shepherd, Jason Spence, Avrum Spira, Xin Sun, Sarah Teichmann, Fabian Theis, Alexander Tsankov, Maarten van den Berge, Michael von Papen, Jeffrey Whitsett, Ramnik Xavier, Yan Xu, Laure-Emmanuelle Zaragosi, Kun Zhang, Carly G K Ziegler, Samuel J Allon, Sarah K Nyquist, Ian M Mbano, Vincent N Miao, Constantine N Tzouanas, Yuming Cao, Ashraf S Yousif, Julia Bals, Blake M Hauser, Jared Feldman, Christoph Muus, Marc H Wadsworth 2nd, Samuel W Kazer, Travis K Hughes, Benjamin Doran, G James Gatter, Marko Vukovic, Faith Taliaferro, Benjamin E Mead, Zhiru Guo, Jennifer P Wang, Delphine Gras, Magali Plaisant, Meshal Ansari, Ilias Angelidis, Heiko Adler, Jennifer M S Sucre, Chase J Taylor, Brian Lin, Avinash Waghray, Vanessa Mitsialis, Daniel F Dwyer, Kathleen M Buchheit, Joshua A Boyce, Nora A Barrett, Tanya M Laidlaw, Shaina L Carroll, Lucrezia Colonna, Victor Tkachev, Christopher W Peterson, Alison Yu, Hengqi Betty Zheng, Hannah P Gideon, Caylin G Winchell, Philana Ling Lin, Colin D Bingle, Scott B Snapper, Jonathan A Kropski, Fabian J Theis, Herbert B Schiller, Laure-Emmanuelle Zaragosi, Pascal Barbry, Alasdair Leslie, Hans-Peter Kiem, JoAnne L Flynn, Sarah M Fortune, Bonnie Berger, Robert W Finberg, Leslie S Kean, Manuel Garber, Aaron G Schmidt, Daniel Lingwood, Alex K Shalek, Jose Ordovas-Montanes, HCA Lung Biological Network. Electronic address: lung-network@humancellatlas.org, HCA Lung Biological Network, Nicholas Banovich, Pascal Barbry, Alvis Brazma, Tushar Desai, Thu Elizabeth Duong, Oliver Eickelberg, Christine Falk, Michael Farzan, Ian Glass, Muzlifah Haniffa, Peter Horvath, Deborah Hung, Naftali Kaminski, Mark Krasnow, Jonathan A Kropski, Malte Kuhnemund, Robert Lafyatis, Haeock Lee, Sylvie Leroy, Sten Linnarson, Joakim Lundeberg, Kerstin Meyer, Alexander Misharin, Martijn Nawijn, Marko Z Nikolic, Jose Ordovas-Montanes, Dana Pe'er, Joseph Powell, Stephen Quake, Jay Rajagopal, Purushothama Rao Tata, Emma L Rawlins, Aviv Regev, Paul A Reyfman, Mauricio Rojas, Orit Rosen, Kourosh Saeb-Parsy, Christos Samakovlis, Herbert Schiller, Joachim L Schultze, Max A Seibold, Alex K Shalek, Douglas Shepherd, Jason Spence, Avrum Spira, Xin Sun, Sarah Teichmann, Fabian Theis, Alexander Tsankov, Maarten van den Berge, Michael von Papen, Jeffrey Whitsett, Ramnik Xavier, Yan Xu, Laure-Emmanuelle Zaragosi, Kun Zhang

Abstract

There is pressing urgency to understand the pathogenesis of the severe acute respiratory syndrome coronavirus clade 2 (SARS-CoV-2), which causes the disease COVID-19. SARS-CoV-2 spike (S) protein binds angiotensin-converting enzyme 2 (ACE2), and in concert with host proteases, principally transmembrane serine protease 2 (TMPRSS2), promotes cellular entry. The cell subsets targeted by SARS-CoV-2 in host tissues and the factors that regulate ACE2 expression remain unknown. Here, we leverage human, non-human primate, and mouse single-cell RNA-sequencing (scRNA-seq) datasets across health and disease to uncover putative targets of SARS-CoV-2 among tissue-resident cell subsets. We identify ACE2 and TMPRSS2 co-expressing cells within lung type II pneumocytes, ileal absorptive enterocytes, and nasal goblet secretory cells. Strikingly, we discovered that ACE2 is a human interferon-stimulated gene (ISG) in vitro using airway epithelial cells and extend our findings to in vivo viral infections. Our data suggest that SARS-CoV-2 could exploit species-specific interferon-driven upregulation of ACE2, a tissue-protective mediator during lung injury, to enhance infection.

Keywords: ACE2; COVID-19; ISG; SARS-CoV-2; human; influenza; interferon; mouse; non-human primate; scRNA-seq.

Conflict of interest statement

Declaration of Interests A.R. is an SAB member of ThermoFisher Scientific, Neogene Therapeutics, Asimov, and Syros Pharmaceuticals; a co-founder of and equity holder in Celsius Therapeutics; and an equity holder in Immunitas Therapeutics. A.K.S. reports compensation for consulting and/or SAB membership from Merck, Honeycomb Biotechnologies, Cellarity, Cogen Therapeutics, Orche Bio, and Dahlia Biosciences. L.S.K. is on the SAB for HiFiBio; she reports research funding from Kymab Limited, Bristol Meyers Squibb, Magenta Therapeutics, BlueBird Bio, and Regeneron Pharmaceuticals and consulting fees from Equillium, FortySeven, Inc, Novartis, Inc, EMD Serono, Gilead Sciences, and Takeda Pharmaceuticals. A.S. is an employee of Johnson and Johnson. N.K. is an inventor on a patent using thyroid hormone mimetics in acute lung injury that is now being considered for intervention in COVID-19 patients. J.L. is a scientific consultant for 10X Genomics, Inc. O.R.R, is a co-inventor on patent applications filed by the Broad Institute to inventions relating to single-cell genomics applications, such as in PCT/US2018/060860 and US Provisional Application No. 62/745,259. S.T. in the last three years was a consultant at Genentech, Biogen, and Roche and is a member of the SAB of Foresite Labs. M.H.W. is now an employee of Pfizer. F.J.T. reports receiving consulting fees from Roche Diagnostics GmbH and ownership interest in Cellarity, Inc. P.H. is a co-inventor on a patent using artificial intelligence and high-resolution microscopy for COVID-19 infection testing based on serology.

Copyright © 2020 The Authors. Published by Elsevier Inc. All rights reserved.

Figures

Graphical abstract
Graphical abstract
Figure 1
Figure 1
Expression of ACE2 in Type II Pneumocytes in Healthy Lungs of Non-human Primates (A) Schematic of protocol for isolation of lung tissue at necropsy from healthy non-human primates (M. mulatta, n = 3), creation of scRNA-seq libraries by using Seq-Well v1, and computational analysis to identify cell types by using unbiased methods. UMAP projection of 3,793 single cells, points colored by cell identity (see STAR Methods). (B) Uniform manifold approximation and projection (UMAP) as in (A), points colored by detection of ACE2 (coronavirus receptor, top) or TMPRSS2 (coronavirus S protein priming for entry, bottom). Color coding is as follows: black, RNA positive; blue, RNA negative. (C) Dot plot of 2 defining genes for each cell type (Table S1) (Bonferroni-adjusted p ACE2 and TMPRSS2. Dot size represents fraction of cells within that type expressing a given gene, and color intensity represents binned count-based expression amount (log(scaled UMI+1)) among expressing cells. ACE2 is enriched in type II pneumocytes (6.7% expressing, Bonferroni-adjusted p = 8.62E−33), as is TMPRSS2 (29.5% expressing, Bonferroni-adjusted p = 8.73E−153). Of all type II pneumocytes, 3.8% co-express ACE2 and TMPRSS2 (Table S9). Red arrow indicates cell type with largest proportion of ACE2+TMPRSS2+ cells. (D) Genes differentially expressed among ACE2+ and ACE2− type II pneumocytes. (SCDE package, FDR-adjusted p < 0.05 for IFNGR2, NT5DC1, ARL6IP1, and TRIM27; full results can be found in Table S1). See also Table S1.
Figure 2
Figure 2
Select Lung Epithelial Cells from Control, HIV-1-Infected, and Mycobacterium-tuberculosis-Infected Human Donors Co-Express ACE2 and TMPRSS2 (A) Schematic of protocol for isolation of human lung tissue from surgical excess, creation of scRNA-seq libraries by using Seq-Well S3, and computational analysis to identify cell types by using unbiased methods. Shown on the right is a UMAP projection of 18,915 cells across 8 donors (n = 3 TB+HIV+; n = 3 TB+; n = 2 non-infected patients). Cells represented by points, colored according to cell type (see STAR Methods). (B) UMAP projection as in (A), points colored by detection of ACE2 (top) or TMPRSS2 (bottom). Color coding is as follows: black, RNA positive; blue, RNA negative. (C) Dot plot of 2 defining genes for each cell type (FDR-adjusted p ACE2 and TMPRSS2; dot size represents fraction of cells within cell type expressing a given gene, and color intensity represents binned count-based expression amount (log(scaled UMI+1)) among expressing cells. All cluster-defining genes are provided in Table S2. Red arrow indicates cell types with largest proportion of ACE2+TMPRSS2+ cells. (D) Volcano plot identifying significantly upregulated genes in ACE2+TMPRSS2+ pneumocytes compared with all remaining pneumocytes. Red points represent genes with a FDR-adjusted p < 0.05, and log2(fold change) >1.5. Text highlighting specific genes; the full list is available in Table S2. (E) Expression of ACE2 across human donors by HIV and TB status (p = 0.009 by likelihood-ratio test). See also Table S2.
Figure S1
Figure S1
NHP Tuberculosis Infected Lung and Granuloma, Related to Figures 1 and 2 (A). UMAP projection of epithelial cells (1,099 cells) colored by annotated cell type, tissue source, and gating as ACE2+TMPRSS2+ cells. ACE2+TMPRSS2+ cells comprise 11% of ciliated cells, 16% of club cells, 10% type I pneumocytes, and 22% type II pneumocytes. Data generated using Seq-Well S3 (Table S3). (B). Number of cells (left) and % (right) ACE2+TMPRSS2+ cells by tissue source (granuloma versus uninvolved lung) and cell type. Ciliated cells and club cells were omitted from this analysis as we detected too few cells (< 7 total cells) belonging to these clusters in the granulomas. Statistical significance assessed by Fisher Exact Test (Table S3). (C). Dot plot of top cluster defining genes for each epithelial cell type and ACE2 and TMPRSS2. Dot size represents fraction of cells expressing, and color intensity represents average log(normalized UMI + 1) among all cells in each group scaled between 0 and 1 by gene. ACE2 expression is enriched in club cells (Bimodal test, Bonferroni-corrected p < 0.001), ciliated cells (p < 0.005), and type I pneumocytes (p < 0.001). TMPRSS2 expression is enriched in type I pneumocytes (p < 0.001) and ciliated cells (p < 0.001) (Table S3). (D). Dot plot of genes differentially expressed between ACE2+TMPRSS2+ epithelial cells versus rest (Bimodal test, Bonferroni-corrected p < 0.01, log fold change > 0.5). (Table S3, c = number of cells, n = number of animals).
Figure 3
Figure 3
NHP and Human Ileal Absorptive Enterocytes Co-Express ACE2 and TMPRSS2 (A) Expression ACE2 across diverse tissues in healthy NHPs (n = 3 animals; 52,858 cells). (B) Schematic of protocol for isolation of NHP ileum (n = 5) at necropsy for scRNA-seq using Seq-Well v1, and computational pipeline to identify cell types by using unbiased methods. Shown on the right is a UMAP projection of 4,515 cells colored by cell type. (C) Dot plot of 2 defining genes for each cell type, with ACE2 and TMPRSS2. Dot size represents fraction of cells within cell type expressing a given gene, and color intensity represents binned count-based expression amounts (log(scaled UMI+1)) among expressing cells. All cluster defining genes are provided in Table S4. Red arrow indicates cell type with largest proportion of ACE2+TMPRSS2+ cells. (D) Schematic of protocol for isolation of human ileal cells from endoscopic pinch biopsies in non-inflamed regions (n = 13). Shown on the right is a tSNE plot of 13,689 epithelial cells selected from original dataset generated by 10x 3′ v2 (see Figure S2), colored by cellular subsets. (E). Dot plot of 2 defining genes for each cell type, with ACE2 and TMPRSS2. Dot size represents fraction of cells within cell type expressing a given gene, and color intensity represents binned count-based expression amounts (log(scaled UMI+1)) among expressing cells. All cluster defining genes are provided in Table S5. Red arrow indicates cell type with largest proportion of ACE2+TMPRSS2+ cells. (F). Expression of ACE2 (left) and TMPRSS2 (right) among all epithelial subsets from human donors. See also Figure S2 and Tables S4 and S5.
Figure S2
Figure S2
Human and NHP Ileum, Related to Figure 3 (A). Top: tSNE projection of all cells from healthy pediatric human ileum within a previously-unpublished 10x 3′ v2 dataset (115,569 cells). Black: higher expression of ACE2 (left), TMPRSS2 (right). Bottom: Corresponding violin plots of expression values for ACE2 (left) and TMPRSS2 (right). Solid line: epithelial cells. (B). Co-expression of ACE2 and TMPRSS2 by epithelial cell subset. Number indicates % of ACE2+TMPRSS2+ cells by cell subset. (C). tSNE projection of 13,689 cells as in Figure 3D, cells colored by co-expression of ACE2 and TMPRSS2 (black). (D). Expression of ACE2 and canonical interferon-responsive genes among absorptive enterocytes from Healthy (n = 2) and SHIV-infected, anti-retroviral treated animals (n = 3). Bonferroni-adjusted p-values by Wilcoxon test (healthy: 510 cells, SHIV-infected: 636 cells).
Figure 4
Figure 4
Healthy and Allergic Inflamed Human Nasal Mucosa Co-Express ACE2 and TMPRSS2 in a Subset of Goblet Secretory Cells (A) Schematic for sampling of n = 12 ethmoid sinus surgical samples and n = 9 inferior turbinate nasal scrapings to generate scRNA-seq libraries by using Seq-Well v1. See Ordovas-Montanes et al., (2018). (B) Dot plot of all cell types from ethmoid-sinus-derived cells (n = 6 non-polyp CRS samples, n = 6 polyp CRS samples). Two defining genes for each cell type, in addition to CDHR3 (rhinovirus receptor), ACE2, TMPRSS2, and JAK1. Dot size represents fraction of cells within that type expressing a given gene, and color intensity represents binned count-based expression amounts (log(scaled UMI+1)) among expressing cells (see Table S6 for statistics by subset). Red arrow indicates cell types with largest proportion of ACE2+TMPRSS2+ cells. (C) Dot plot for 2 defining genes for each cell type identified from granular clustering of epithelial cells (18,325 single cells) derived from both ethmoid sinus and inferior turbinate sampling (healthy inferior turbinate [3,681 cells; n = 3 samples], polyp-bearing patient inferior turbinate [1,370 cells; n = 4 samples], non-polyp ethmoid sinus surgical samples [5,928 cells; n = 6 samples], and polyp surgical and scraping samples directly from polyp in ethmoid sinus [7,346 cells; n = 8 samples]). Red arrow indicates cell type with largest proportion of ACE2+TMPRSS2+ cells. (D) tSNE of 18,325 single epithelial cells from inferior turbinate and ethmoid sinus (omitting immune cells). Colored by cell types 3,152 basal, 3,089 differentiating, 8,840 secretory, 1,105 ciliated, and 2,139 glandular cells. (E) tSNE as in (D), identifying epithelial cells co-expressing ACE2 and TMPRSS2 (30 cells, black points). (F) tSNE as in (D), colored by detailed cell types with higher granularity, as in (C). (G) Individual differentially expressed genes between ACE2+TMPRSS2+ cells and all other secretory epithelial cells (see Table S6 for full gene list with statistics). Bonferroni-adjusted likelihood-ratio test p < 0.02 for all genes displayed. (H) Stacked bar plot of each subset of epithelial cells among all epithelial cells by donor (each bar) and sampling location (noted below graph) (unpaired t test p ACE2+TMPRSS2+ cells. See also Figure S3 and Table S6.
Figure S3
Figure S3
Nasal and Sinus Mucosa, Related to Figures 4 and 5 (A). Expression of ACE2 and TMPRSS2 across donors. (B). Enhanced capture of ACE2 mRNA with second strand synthesis protocol employed in Seq-Well S3. Dot size represents fraction of cells expressing. (C). Cultured human primary basal epithelial cells at confluence were treated with increasing doses (0.1 to 10ng/mL) of IFNα2, IFNγ, IL-4, IL-13, IL-17A, and IL-1B for 12 h and bulk RNA-seq analysis was performed (Replicate experiment using Human Donor 1 as in Figure 5) (D). ACE2 expression by stimulation condition. Wilcoxon test between each cytokine (combined doses) versus rest: IFNα Bonferroni-adjusted p = 4.1E-07; IFNγ Bonferroni-adjusted p = 9.3E-03; all else n.s. ∗∗∗ p < 0.001. (E). ACE2 expression by IFNα2 dose. Bonferroni-corrected t-test compared to 0 ng/mL condition: ∗∗∗ p < 0.001, ∗ p < 0.05. (F). ACE2 expression by IFNγ dose. Bonferroni-corrected t-test compared to 0 ng/mL condition: ∗∗∗ p < 0.001, ∗ p < 0.05. (G). IFITM1 expression by IFNα2 dose. Bonferroni-corrected t-test compared to 0 ng/mL condition: ∗∗∗ p < 0.001. (H). IFITM1 expression by IFNγ dose. Bonferroni-corrected t-test compared to 0 ng/mL condition: ∗∗∗ p < 0.001. (I). GBP5 expression among cultured human primary basal epithelial cells. Wilcoxon test: IFNα versus IFNγ Bonferroni-adjusted p = 2.94E-07; IFNγ Bonferroni-adjusted p = 9.3E-03. TP10K: transcripts per 10,000 reads. ∗∗∗ p < 0.001. (J). GBP5 expression by IFNα2 dose. Bonferroni-corrected t-test compared to 0 ng/mL condition: ∗∗∗ p < 0.001. (K). GBP5 expression by IFNγ dose. Bonferroni-corrected t-test compared to 0 ng/mL condition: ∗∗∗ p < 0.001.
Figure 5
Figure 5
ACE2 is an Interferon-Stimulated Gene in Primary Human Barrier Tissue Epithelial Cells (A–D) Basal epithelial cells from distinct sources were cultured to confluence and treated with increasing doses (0.1–10 ng/mL) of IFN-α2, IFN-γ, IL-4, IL-17A, and/or IFN-β for 12 h and bulk RNA-seq analysis was performed. Expression of ACE2 (human) or Ace2 (mouse) by cell type and stimulation condition. (A) Primary mouse basal cells from tracheal epithelium are shown. (B) BEAS-2B human bronchial cell line is shown. (C) Primary human basal cells from nasal scraping, Donor 1, is shown. (D) Primary human basal cells from nasal scraping, Donor 2. Abbreviation is as follows: TP10K, transcripts per 10,000 reads. ∗∗∗p < 0.001, ∗∗p < 0.01, ∗p < 0.05, Bonferroni-corrected t test compared with untreated condition. (E–H) Co-expression of STAT1/Stat1 and ACE2/Ace2 by cell type. (E) Primary mouse basal cells from tracheal epithelium are shown. (F) BEAS-2B human bronchial cell line is shown. (G) Primary human basal cells from nasal scraping, Donor 1, are shown. (H) Primary human basal cells from nasal scraping, Donor 2 are shown. Abbreviation is as follows: TP10K, transcripts per 10,000 reads. Statistical significance assessed by Spearman’s rank correlation. (I–L) Expression of ACE2 in primary human basal cells from nasal scrapings across a range of concentrations of IFN-γ or IFN-α2. (I) IFN-α2 dose response in Donor 1 (p < 0.001 by one-way ANOVA) is shown. (J) IFN-γ dose response in Donor 1 (p < 0.01 by one-way ANOVA) is shown. (K) IFN-α2 dose response in Donor 2 (p < 0.001 by one-way ANOVA) is shown. (L) IFN-γ dose response in Donor 2 (p < 0.001 by one-way ANOVA). Abbreviation is as follows: TP10K, transcripts per 10,000 reads. ∗∗∗p < 0.001, ∗∗p < 0.01, ∗p < 0.05, Bonferroni-corrected post hoc testing compared with 0 ng/mL condition. See also Figures S3 and S4 and Table S7.
Figure S4
Figure S4
Published Studies of Epithelial Cells Following Interferon Treatment Related to Figure 5 (A). Fold change of ACE2 expression among human or mouse datasets following Type I or Type II interferon treatment compared to untreated control. Generated from publicly available microarray data curated at interferome.org. Includes all studies with abs(fold-change) > 1. (B). Location of transcription factors binding regions spanning −1500 bp to +500 bp from the transcription start site of ACE2 (human, top) or Ace2 (mouse, bottom). Generated from TRANSFAC data using the interferome.org database (Matys et al., 2003, Rusinova et al., 2013).
Figure S5
Figure S5
Mouse Nasal Epithelium Following Interferon-α Exposure Related to Figure 6 (A). Schematic: mice were exposed to 10,000 units of IFN-α or saline by intranasal application (n = 2 per group). After 12 h, animals were sacrificed and nasal epithelium was dissected and dissociated for scRNA-seq using Seq-Well S3. (B). Dot plot of 2 defining genes for each cell type, with Ace2, Tmprss2, and Cdhr3. Dot size represents fraction of cells within cell type expressing, and color intensity binned count-based expression level (log(scaled UMI+1)) among expressing cells. All cluster defining genes are provided in Table S8. Red arrows: cell types with largest proportion of Ace2+ cells. Dendrogram (left) by person correlation over differentially expressed genes with Ward clustering. (C). UMAP of Basal Epithelial Cells (380 cells) across 4 mice. Black: Saline-treated mouse; red: IFN-α treated. (D). UMAP of Basal Epithelial Cells as in C, points colored by detection of Ace2. Black: RNA positive, blue: RNA negative (6.6% Ace2+, Bonferroni-adjusted p = 1.1E-10 for Basal Epithelial Cell expression versus all other cells). (E). Schematic: wildtype (WT) and IFNγ-receptor knockout (IFNγR−/−) mice were infected intranasally with murine gamma-herpesvirus-68 (MHV68). Cells from whole lung were digested for scRNA-seq using Drop-seq (yielding 5,558 Epcam+ cells). (F). Expression of Ace2 by epithelial cell type, wild type (WT) mice. Statistical significance by Wilcoxon rank sum test with Bonferroni correction. (G). Expression of Ace2 among type II pneumocytes binned by infection status in WT mice. All pairwise comparisons non-significant (p > 0.05) by Wilcoxon rank sum test. (H). Percent of Ace2+ cells by infection condition (uninfected, bystander cells in MHV68-infected mouse, MHV68 RNA+ cells) and mouse genotype (WT, IFNγR −/−). Black bars: Ace2+ positive cells; white bars: Ace2- cells. (I). Schematic of RNA-Seq data from (Matos et al., 2019) of human lung explants (n = 5 donors) exposed to influenza A virus (IAV, H3N2) at 24 h post infection. (J). Expression of SFTPC (surfactant protein C, a marker of type II pneumocytes) versus ACE2 among mock-infected lung explants. Statistical significance assessed by Pearson’s correlation, r = 0.93, p = 0.021. TPM: transcripts per million. (K). SFTPC expression among matched donors following mock or IAV infection for 24 h. Statistical significance assessed by ratio paired t test, p = 0.86. (L). ACE2 expression among matched donors following mock or IAV infection for 24 h. Statistical significance assessed by ratio paired t test, p = 0.0054. (M). Western blot of fully-differentiated air-liquid interface cultures from bronchial cells derived from 4 human donors with asthma. Cells from each donor were treated with 10 ng/mL IFNγ for 24 h, and compared to a matched untreated condition. ACE2 protein: AF933 (R&D). Fold changes quantified for IFNγ treated versus untreated for each patient donor following normalization to GAPDH.
Figure 6
Figure 6
In Vivo Administration of Interferons in Mice Does Not Induce Ace2, and ACE2 Is Induced in Goblet Secretory Cells during Human Influenza Infection (A) UMAP of 11,358 single cells from mouse nasal epithelium (n = 4). (B) UMAP projection as in (A), points colored by detection of Ace2 (SARS-CoV-2 receptor homolog). Color coding is as follows: black, RNA positive; blue, RNA negative. (C) Percent of Ace2+ cells by treatment condition (n = 4 arrays per condition; n = 2 arrays per mouse). Black bars indicate Ace2+ cells; white bars indicate Ace2 cells. p = 0.4 by Student’s t test. (D) Heatmap of cell-type-defining genes (Trp63 and Krt17), interferon-induced genes (Irf7, Stat1, Irf9, and Oasl2), and Ace2 among basal epithelial cells, separated by cells derived from saline-treated mice (left) and IFN-α-treated mice (right). Statistical significance by likelihood-ratio test with Bonferroni correction is shown. A full list of differentially expressed genes can be found in Table S8. (E) Schematic for sampling cells derived from nasal washes of n = 18 human donors with and without current influenza A or B infection for Seq-Well v1 (35,840 single cells). See Cao et al., (2020). (F and G) ACE2 expression among goblet cells (F) and squamous cells (G) by infection status. Shown are Healthy Donor cells from influenza-negative donors (white); Bystander Cells from influenza A (IAV)- or influenza B (IBV)-infected donors, no intracellular viral RNA detected (black); Flu Viral RNA+ Cells with detectable intracellular influenza A or B viral RNA (red). Statistical significance by Wilcoxon test with Bonferroni correction, n.s. for Bystander versus Flu Viral RNA+. See also Figure S5 and Tables S6 and S8.
Figure S6
Figure S6
Power Calculations and Statistical Modeling of ACE2 Capture and Dropout Related to STAR Methods (A). Probability of capturing and transcribing at least 1 ACE2 cDNA molecule, as a function of the capture/reverse transcription efficiency for a single molecule and the number of ACE2 molecules expressed in an individual cell. Note that Drop-Seq provides a capture/transcription efficiency of approximately 11-13%, setting a floor on this parameter, and the experimental platforms used in this study are either equivalent or superior (Macosko et al., 2015). (B). Distribution of ACE2 fractional abundance within individual cells’ cDNA libraries (i.e., ACE2 UMIs / total number of reads), across non-human primate lung and ileum cell populations (see Figures 1 and 3). Mean fractional abundance among ACE2+ lung cells = 5.0E-5; mean fractional abundance among ACE2+ ileum cells = 2.7E-4. (C). Distribution of the number of reads within non-human primate lung and ileum cell populations (see Figures 1 and 3). Mean ± SEM reads among all lung cells = 28,512 ± 344; ACE2+ lung cells = 28,553 ± 2,988; all ileum cells = 14,864 ± 288; ACE2+ ileum cells = 10,591 ± 441. (D). Probability of observing at least one transcript for a gene of interest (e.g., ACE2) within an individual cell, as a function of sequencing depth and the gene’s fractional abundance (i.e., ACE2 reads / all reads) within the cell’s cDNA library. Fractional abundance provides the probability that a single read corresponds to the gene of interest, and presented heatmap indicates the probability that at least one read in the total number of reads allocated to the cell (i.e., from 103 to 106) originates from the gene of interest. Mean read depths and ACE2 fractional abundances for each tissue produce a 93.7% probability of detecting at least 1 ACE2 read in ileum cells, and a 76.0% chance for lung cells. Outlined rectangles highlight the regimes where cells from lung (turquoise) and ileum (pink) samples typically lie. (E). Number of ACE2+ cells within each cluster, as a function of average read depth for all cells in that cluster. Number of cells detected as ACE2+ is not correlated with read depth, even across relatively wide ranges of average read depths (Pearson’s r = −0.31, n.s.). (F). Probability of observing a particular number of cells positive for a gene of interest within a cluster, as a function of number of cells in the cluster. Probabilities were calculated under a negative binomial distribution with parameter p = 0.063 (the proportion of ACE2+ cells among type II pneumocytes presented in Figure 1; STAR Methods). The horizontal gray line indicates the arbitrary cut-off value of p = 0.05. (G). Given a population of cells with a known proportion that are positive for a gene of interest, probability of observing no positive cells (i.e., false negative identification of the cluster; solid lines) and probability of observing at least one positive cell as a function of cluster size.

References

    1. Adler H., Messerle M., Wagner M., Koszinowski U.H. Cloning and mutagenesis of the murine gammaherpesvirus 68 genome as an infectious bacterial artificial chromosome. J. Virol. 2000;74:6964–6974.
    1. Amanat F., Nguyen T., Chromikova V., Strohmeier S., Stadlbauer D., Javier A., Jiang K., Asthagiri-Arunkumar G., Polanco J., Bermudez-Gonzalez M. A serological assay to detect SARS-CoV-2 seroconversion in humans. medRxiv. 2020 doi: 10.1101/2020.03.17.20037713.
    1. Ardain A., Domingo-Gonzalez R., Das S., Kazer S.W., Howard N.C., Singh A., Ahmed M., Nhamoyebonde S., Rangel-Moreno J., Ogongo P. Group 3 innate lymphoid cells mediate early protective immunity against tuberculosis. Nature. 2019;570:528–532.
    1. Bailey C.C., Zhong G., Huang I.C., Farzan M. IFITM-Family Proteins: The Cell’s First Line of Antiviral Defense. Annu. Rev. Virol. 2014;1:261–283.
    1. Bao L., Deng W., Gao H., Xiao C., Liu J., Xue J., Lv Q., Liu J., Yu P., Xu Y. Reinfection could not occur in SARS-CoV-2 infected rhesus macaques. bioRxiv. 2020 doi: 10.1101/2020.03.13.990226.
    1. Blanco-Melo D., Nilsson-Payant B.E., Liu W.-C., Møller R., Panis M., Sachs D., Albrecht R.A., tenOever B.R. SARS-CoV-2 launches a unique transcriptional signature from in vitro, ex vivo, and in vivo systems. bioRxiv. 2020 doi: 10.1101/2020.03.24.004655.
    1. Böttcher-Friebertshäuser E., Klenk H.D., Garten W. Activation of influenza viruses by proteases from host cells and bacteria in the human airway epithelium. Pathog. Dis. 2013;69:87–100.
    1. Brann D., Tsukahara T., Weinreb C., Logan D.W., Datta S.R. Non-neural expression of SARS-CoV-2 entry genes in the olfactory epithelium suggests mechanisms underlying anosmia in COVID-19 patients. bioRxiv. 2020 doi: 10.1101/2020.03.25.009084.
    1. Braun E., Sauter D. Furin-mediated protein processing in infectious diseases and cancer. Clin. Transl. Immunology. 2019;8:e1073.
    1. Broggi A., Granucci F., Zanoni I. Type III interferons: Balancing tissue tolerance and resistance to pathogen invasion. J. Exp. Med. 2020;217:e20190295.
    1. Bugge T.H., Antalis T.M., Wu Q. Type II transmembrane serine proteases. J. Biol. Chem. 2009;284:23177–23181.
    1. Butler A., Hoffman P., Smibert P., Papalexi E., Satija R. Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nat. Biotechnol. 2018;36:411–420.
    1. Cao Y., Guo Z., Vangala P., Donnard E., Liu P., McDonel P., Ordovas-Montanes J., Shalek A.K., Finberg R.W., Wang J.P. Single-cell analysis of upper airway cells reveals host-viral dynamics in influenza infected adults. bioRxiv. 2020 doi: 10.1101/2020.04.15.042978.
    1. Channappanavar R., Perlman S. Pathogenic human coronavirus infections: causes and consequences of cytokine storm and immunopathology. Semin. Immunopathol. 2017;39:529–539.
    1. Channappanavar R., Fehr A.R., Vijay R., Mack M., Zhao J., Meyerholz D.K., Perlman S. Dysregulated Type I Interferon and Inflammatory Monocyte-Macrophage Responses Cause Lethal Pneumonia in SARS-CoV-Infected Mice. Cell Host Microbe. 2016;19:181–193.
    1. Channappanavar R., Fehr A.R., Zheng J., Wohlford-Lenane C., Abrahante J.E., Mack M., Sompallae R., McCray P.B., Jr., Meyerholz D.K., Perlman S. IFN-I response timing relative to virus replication determines MERS coronavirus infection outcomes. J. Clin. Invest. 2019;130:3625–3639.
    1. Chu H.Y., Boeckh M., Englund J.A., Famulare M., Lutz B.R., Nickerson D.A., Rieder M.J., Starita L.M., Adler A., Brandstetter E. The Seattle Flu Study: a multi-arm community-based prospective study protocol for assessing influenza prevalence, transmission, and genomic epidemiology. medRxiv. 2020 doi: 10.1101/2020.03.02.20029595.
    1. Colonna L., Peterson C.W., Schell J.B., Carlson J.M., Tkachev V., Brown M., Yu A., Reddy S., Obenza W.M., Nelson V. Evidence for persistence of the SHIV reservoir early after MHC haploidentical hematopoietic stem cell transplantation. Nat. Commun. 2018;9:4438.
    1. Coronaviridae Study Group of the International Committee on Taxonomy of V. The species Severe acute respiratory syndrome-related coronavirus: classifying 2019-nCoV and naming it SARS-CoV-2. Nat. Microbiol. 2020 doi: 10.1038/s41564-020-0695-z.
    1. Coutard B., Valle C., de Lamballerie X., Canard B., Seidah N.G., Decroly E. The spike glycoprotein of the new coronavirus 2019-nCoV contains a furin-like cleavage site absent in CoV of the same clade. Antiviral Res. 2020;176:104742.
    1. Davidson D.J., Gray M.A., Kilanowski F.M., Tarran R., Randell S.H., Sheppard D.N., Argent B.E., Dorin J.R. Murine epithelial cells: isolation and culture. J. Cyst. Fibros. 2004;3(Suppl 2):59–62.
    1. Davidson S., Maini M.K., Wack A. Disease-promoting effects of type I interferons in viral, bacterial, and coinfections. J. Interferon Cytokine Res. 2015;35:252–264.
    1. de Lang A., Osterhaus A.D., Haagmans B.L. Interferon-gamma and interleukin-4 downregulate expression of the SARS coronavirus receptor ACE2 in Vero E6 cells. Virology. 2006;353:474–481.
    1. Dear T.N., Boehm T., Keverne E.B., Rabbitts T.H. Novel genes for potential ligand-binding proteins in subregions of the olfactory mucosa. EMBO J. 1991;10:2813–2819.
    1. Deeks S.G., Odorizzi P.M., Sekaly R.P. The interferon paradox: can inhibiting an antiviral mechanism advance an HIV cure? J. Clin. Invest. 2017;127:103–105.
    1. Derr A., Yang C., Zilionis R., Sergushichev A., Blodgett D.M., Redick S., Bortell R., Luban J., Harlan D.M., Kadener S. End Sequence Analysis Toolkit (ESAT) expands the extractable information from single-cell RNA-seq data. Genome Res. 2016;26:1397–1410.
    1. Ding Y., He L., Zhang Q., Huang Z., Che X., Hou J., Wang H., Shen H., Qiu L., Li Z. Organ distribution of severe acute respiratory syndrome (SARS) associated coronavirus (SARS-CoV) in SARS patients: implications for pathogenesis and virus transmission pathways. J. Pathol. 2004;203:622–630.
    1. Dong E., Du H., Gardner L. An interactive web-based dashboard to track COVID-19 in real time. Lancet Infect. Dis. 2020 S1473-3099(20)30120-1.
    1. Dong L., Hu S., Gao J. Discovering drugs to treat coronavirus disease 2019 (COVID-19) Drug Discov. Ther. 2020;14:58–60.
    1. Dunston D., Ashby S., Krosnowski K., Ogura T., Lin W. An effective manual deboning method to prepare intact mouse nasal tissue with preserved anatomical organization. J. Vis. Exp. 2013;(78) doi: 10.3791/50538.
    1. Dupuis S., Jouanguy E., Al-Hajjar S., Fieschi C., Al-Mohsen I.Z., Al-Jumaah S., Yang K., Chapgier A., Eidenschenk C., Eid P. Impaired response to interferon-alpha/beta and lethal viral disease in human STAT1 deficiency. Nat. Genet. 2003;33:388–391.
    1. Everitt A.R., Clare S., Pertel T., John S.P., Wash R.S., Smith S.E., Chin C.R., Feeley E.M., Sims J.S., Adams D.J., GenISIS Investigators. MOSAIC Investigators IFITM3 restricts the morbidity and mortality associated with influenza. Nature. 2012;484:519–523.
    1. Fleming F.E., Böhm R., Dang V.T., Holloway G., Haselhorst T., Madge P.D., Deveryshetty J., Yu X., Blanchard H., von Itzstein M., Coulson B.S. Relative roles of GM1 ganglioside, N-acylneuraminic acids, and α2β1 integrin in mediating rotavirus infection. J. Virol. 2014;88:4558–4571.
    1. Fung T.S., Liu D.X. Human Coronavirus: Host-Pathogen Interaction. Annu. Rev. Microbiol. 2019;73:529–557.
    1. Genshaft A.S., Li S., Gallant C.J., Darmanis S., Prakadan S.M., Ziegler C.G., Lundberg M., Fredriksson S., Hong J., Regev A. Multiplexed, targeted profiling of single-cell proteomes and transcriptomes in a single reaction. Genome Biol. 2016;17:188.
    1. Gerstein M.B., Kundaje A., Hariharan M., Landt S.G., Yan K.K., Cheng C., Mu X.J., Khurana E., Rozowsky J., Alexander R. Architecture of the human regulatory network derived from ENCODE data. Nature. 2012;489:91–100.
    1. Gierahn T.M., Wadsworth M.H., 2nd, Hughes T.K., Bryson B.D., Butler A., Satija R., Fortune S., Love J.C., Shalek A.K. Seq-Well: portable, low-cost RNA sequencing of single cells at high throughput. Nat. Methods. 2017;14:395–398.
    1. Giovannini-Chami L., Marcet B., Moreilhon C., Chevalier B., Illie M.I., Lebrigand K., Robbe-Sermesant K., Bourrier T., Michiels J.F., Mari B. Distinct epithelial gene expression phenotypes in childhood respiratory allergy. Eur. Respir. J. 2012;39:1197–1205.
    1. Glowacka I., Bertram S., Müller M.A., Allen P., Soilleux E., Pfefferle S., Steffen I., Tsegaye T.S., He Y., Gnirss K. Evidence that TMPRSS2 activates the severe acute respiratory syndrome coronavirus spike protein for membrane fusion and reduces viral control by the humoral immune response. J. Virol. 2011;85:4122–4134.
    1. Good C., Wells A.I., Coyne C.B. Type III interferon signaling restricts enterovirus 71 infection of goblet cells. Sci Adv. 2019 doi: 10.1126/sciadv.aau4255.
    1. Griggs T.F., Bochkov Y.A., Basnet S., Pasic T.R., Brockman-Schneider R.A., Palmenberg A.C., Gern J.E. Rhinovirus C targets ciliated airway epithelial cells. Respir. Res. 2017;18:84.
    1. Grove J., Marsh M. The cell biology of receptor-mediated virus entry. J. Cell Biol. 2011;195:1071–1082.
    1. Guan W.J., Ni Z.Y., Hu Y., Liang W.H., Ou C.Q., He J.X., Liu L., Shan H., Lei C.L., Hui D.S.C., China Medical Treatment Expert Group for Covid-19 Clinical Characteristics of Coronavirus Disease 2019 in China. N. Engl. J. Med. 2020 doi: 10.1056/NEJMoa2002032.
    1. Guerrero-Plata A., Baron S., Poast J.S., Adegboyega P.A., Casola A., Garofalo R.P. Activity and regulation of alpha interferon in respiratory syncytial virus and human metapneumovirus experimental infections. J. Virol. 2005;79:10190–10199.
    1. Hadfield J., Megill C., Bell S.M., Huddleston J., Potter B., Callender C., Sagulenko P., Bedford T., Neher R.A. Nextstrain: real-time tracking of pathogen evolution. Bioinformatics. 2018;34:4121–4123.
    1. Hamming I., Timens W., Bulthuis M.L., Lely A.T., Navis G., van Goor H. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus. A first step in understanding SARS pathogenesis. J. Pathol. 2004;203:631–637.
    1. Harmer D., Gilbert M., Borman R., Clark K.L. Quantitative mRNA expression profiling of ACE 2, a novel homologue of angiotensin converting enzyme. FEBS Lett. 2002;532:107–110.
    1. Hoffmann M., Kleine-Weber H., Schroeder S., Krüger N., Herrler T., Erichsen S., Schiergens T.S., Herrler G., Wu N.-H., Nitsche A. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell. 2020;181:271–280.e8.
    1. Hofmann H., Geier M., Marzi A., Krumbiegel M., Peipp M., Fey G.H., Gramberg T., Pöhlmann S. Susceptibility to SARS coronavirus S protein-driven infection correlates with expression of angiotensin converting enzyme 2 and infection can be blocked by soluble receptor. Biochem. Biophys. Res. Commun. 2004;319:1216–1221.
    1. Holly M.K., Smith J.G. Adenovirus Infection of Human Enteroids Reveals Interferon Sensitivity and Preferential Infection of Goblet Cells. J. Virol. 2018;92:e00250-18.
    1. Holshue M.L., DeBolt C., Lindquist S., Lofy K.H., Wiesman J., Bruce H., Spitters C., Ericson K., Wilkerson S., Tural A., Washington State 2019-nCoV Case Investigation Team First Case of 2019 Novel Coronavirus in the United States. N. Engl. J. Med. 2020;382:929–936.
    1. Huang F., Guo J., Zou Z., Liu J., Cao B., Zhang S., Li H., Wang W., Sheng M., Liu S. Angiotensin II plasma levels are linked to disease severity and predict fatal outcomes in H7N9-infected patients. Nat. Commun. 2014;5:3595.
    1. Huang C., Wang Y., Li X., Ren L., Zhao J., Hu Y., Zhang L., Fan G., Xu J., Gu X. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet. 2020;395:497–506.
    1. Hughes T.K., Wadsworth M.H., Gierahn T.M., Do T., Weiss D., Andrade P.R., Ma F., de Andrade Silva B.J., Shao S., Tsoi L.C. Highly Efficient, Massively-Parallel Single-Cell RNA-Seq Reveals Cellular States and Molecular Features of Human Skin Pathology. bioRxiv. 2019 doi: 10.1101/689273.
    1. Imai Y., Kuba K., Rao S., Huan Y., Guo F., Guan B., Yang P., Sarao R., Wada T., Leong-Poi H. Angiotensin-converting enzyme 2 protects from severe acute lung failure. Nature. 2005;436:112–116.
    1. Iwasaki A., Pillai P.S. Innate immunity to influenza virus infection. Nat. Rev. Immunol. 2014;14:315–328.
    1. Iwasaki A., Foxman E.F., Molony R.D. Early local immune defences in the respiratory tract. Nat. Rev. Immunol. 2017;17:7–20.
    1. Iwata-Yoshikawa N., Okamura T., Shimizu Y., Hasegawa H., Takeda M., Nagata N. TMPRSS2 Contributes to Virus Spread and Immunopathology in the Airways of Murine Models after Coronavirus Infection. J. Virol. 2019;93:e01815-18.
    1. Jewell N.A., Cline T., Mertz S.E., Smirnov S.V., Flaño E., Schindler C., Grieves J.L., Durbin R.K., Kotenko S.V., Durbin J.E. Lambda interferon is the predominant interferon induced by influenza A virus infection in vivo. J. Virol. 2010;84:11515–11522.
    1. Jonsdottir H.R., Dijkman R. Coronaviruses and the human airway: a universal system for virus-host interaction studies. Virol. J. 2016;13:24.
    1. Jovanovic M., Rooney M.S., Mertins P., Przybylski D., Chevrier N., Satija R., Rodriguez E.H., Fields A.P., Schwartz S., Raychowdhury R. Immunogenetics. Dynamic profiling of the protein life cycle in response to pathogens. Science. 2015;347:1259038.
    1. Kazer S.W., Aicher T.P., Muema D.M., Carroll S.L., Ordovas-Montanes J., Miao V.N., Tu A.A., Ziegler C.G.K., Nyquist S.K., Wong E.B. Integrated single-cell analysis of multicellular immune dynamics during hyperacute HIV-1 infection. Nat. Med. 2020;26:511–518.
    1. Kharchenko P.V., Silberstein L., Scadden D.T. Bayesian approach to single-cell differential expression analysis. Nat. Methods. 2014;11:740–742.
    1. Krischuns T., Günl F., Henschel L., Binder M., Willemsen J., Schloer S., Rescher U., Gerlt V., Zimmer G., Nordhoff C. Phosphorylation of TRIM28 Enhances the Expression of IFN-β and Proinflammatory Cytokines During HPAIV Infection of Human Lung Epithelial Cells. Front. Immunol. 2018;9:2229. doi: 10.3389/fimmu.2018.02229.
    1. Kuba K., Imai Y., Rao S., Gao H., Guo F., Guan B., Huan Y., Yang P., Zhang Y., Deng W. A crucial role of angiotensin converting enzyme 2 (ACE2) in SARS coronavirus-induced lung injury. Nat. Med. 2005;11:875–879.
    1. Kucharski A.J., Russell T.W., Diamond C., Liu Y., Edmunds J., Funk S., Eggo R.M., Centre for Mathematical Modelling of Infectious Diseases COVID-19 working group Early dynamics of transmission and control of COVID-19: a mathematical modelling study. Lancet Infect. Dis. 2020 doi: 10.1016/S1473-3099(20)30144-4.
    1. Lechien J.R., Chiesa-Estomba C.M., De Siati D.R., Horoi M., Le Bon S.D., Rodriguez A., Dequanter D., Blecic S., El Afia F., Distinguin L. Olfactory and gustatory dysfunctions as a clinical presentation of mild-to-moderate forms of the coronavirus disease (COVID-19): a multicenter European study. Eur. Arch. Otorhinolaryngol. 2020 doi: 10.1007/s00405-020-05965-1.
    1. Lei J., Li J., Li X., Qi X. CT Imaging of the 2019 Novel Coronavirus (2019-nCoV) Pneumonia. Radiology. 2020;295:18.
    1. Letko M., Marzi A., Munster V. Functional assessment of cell entry and receptor usage for SARS-CoV-2 and other lineage B betacoronaviruses. Nat. Microbiol. 2020;5:562–569.
    1. Li G., De Clercq E. Therapeutic options for the 2019 novel coronavirus (2019-nCoV) Nat. Rev. Drug Discov. 2020;19:149–150.
    1. Li W., Moore M.J., Vasilieva N., Sui J., Wong S.K., Berne M.A., Somasundaran M., Sullivan J.L., Luzuriaga K., Greenough T.C. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature. 2003;426:450–454.
    1. Liao M., Liu Y., Yuan J., Wen Y., Xu G., Zhao J., Chen L., Li J., Wang X., Wang F. The landscape of lung bronchoalveolar immune cells in COVID-19 revealed by single-cell RNA sequencing. medRxiv. 2020 doi: 10.1101/2020.02.23.20026690.
    1. Lu R., Zhao X., Li J., Niu P., Yang B., Wu H., Wang W., Song H., Huang B., Zhu N. Genomic characterisation and epidemiology of 2019 novel coronavirus: implications for virus origins and receptor binding. Lancet. 2020;395:565–574.
    1. Lukassen S., Chua R.L., Trefzer T., Kahn N.C., Schneider M.A., Muley T., Winter H., Meister M., Veith C., Boots A.W. SARS-CoV-2 receptor ACE2 and TMPRSS2 are predominantly expressed in a transient secretory cell type in subsegmental bronchial branches. bioRxiv. 2020 doi: 10.1101/2020.03.13.991455.
    1. Lun A.T., McCarthy D.J., Marioni J.C. A step-by-step workflow for low-level analysis of single-cell RNA-seq data with Bioconductor. F1000Res. 2016;5:2122. doi: 10.12688/f1000research.9501.2.
    1. Macosko E.Z., Basu A., Satija R., Nemesh J., Shekhar K., Goldman M., Tirosh I., Bialas A.R., Kamitaki N., Martersteck E.M. Highly Parallel Genome-wide Expression Profiling of Individual Cells Using Nanoliter Droplets. Cell. 2015;161:1202–1214.
    1. Mar K.B., Rinkenberger N.R., Boys I.N., Eitson J.L., McDougal M.B., Richardson R.B., Schoggins J.W. LY6E mediates an evolutionarily conserved enhancement of virus infection by targeting a late entry step. Nat. Commun. 2018;9:3603. doi: 10.1038/s41467-018-06000-y.
    1. Martin C.J., Cadena A.M., Leung V.W., Lin P.L., Maiello P., Hicks N., Chase M.R., Flynn J.L., Fortune S.M. Digitally Barcoding Mycobacterium tuberculosis Reveals In Vivo Infection Dynamics in the Macaque Model of Tuberculosis. MBio. 2017;8:e00312-17.
    1. Matos A.D.R., Wunderlich K., Schloer S., Schughart K., Geffers R., Seders M., Witt M., Christersson A., Wiewrodt R., Wiebe K. Antiviral potential of human IFN-α subtypes against influenza A H3N2 infection in human lung explants reveals subtype-specific activities. Emerg. Microbes Infect. 2019;8:1763–1776.
    1. Matsuyama S., Nagata N., Shirato K., Kawase M., Takeda M., Taguchi F. Efficient activation of the severe acute respiratory syndrome coronavirus spike protein by the transmembrane protease TMPRSS2. J. Virol. 2010;84:12658–12664.
    1. Matys V., Fricke E., Geffers R., Gössling E., Haubrock M., Hehl R., Hornischer K., Karas D., Kel A.E., Kel-Margoulis O.V. TRANSFAC: transcriptional regulation, from patterns to profiles. Nucleic Acids Res. 2003;31:374–378.
    1. McCray P.B., Jr., Pewe L., Wohlford-Lenane C., Hickey M., Manzel L., Shi L., Netland J., Jia H.P., Halabi C., Sigmund C.D. Lethal infection of K18-hACE2 mice infected with severe acute respiratory syndrome coronavirus. J. Virol. 2007;81:813–821.
    1. Mead B.E., Karp J.M. All models are wrong, but some organoids may be useful. Genome Biol. 2019;20:66. doi: 10.1186/s13059-019-1677-4.
    1. Mead B.E., Ordovas-Montanes J., Braun A.P., Levy L.E., Bhargava P., Szucs M.J., Ammendolia D.A., MacMullan M.A., Yin X., Hughes T.K. Harnessing single-cell genomics to improve the physiological fidelity of organoid-derived cell types. BMC Biol. 2018;16:62. doi: 10.1186/s12915-018-0527-2.
    1. Medzhitov R., Schneider D.S., Soares M.P. Disease tolerance as a defense strategy. Science. 2012;335:936–941.
    1. Monteil V., Kwon H., Prado P., Hagelkryus A., Wimmer R.A., al e. Inhibition of SARS-CoV-2 infections in engineered human tissues using clinical-grade soluble human ACE2. Cell. 2020 doi: 10.1016/j.cell.2020.04.004.
    1. Montoro D.T., Haber A.L., Biton M., Vinarsky V., Lin B., Birket S.E., Yuan F., Chen S., Leung H.M., Villoria J. A revised airway epithelial hierarchy includes CFTR-expressing ionocytes. Nature. 2018;560:319–324.
    1. Müller U., Steinhoff U., Reis L.F., Hemmi S., Pavlovic J., Zinkernagel R.M., Aguet M. Functional role of type I and type II interferons in antiviral defense. Science. 1994;264:1918–1921.
    1. Munster V.J., Feldmann F., Williamson B.N., van Doremalen N., Pérez-Pérez L., Schulz J., Meade-White K., Okumura A., Callison J., Brumbaugh B. Respiratory disease and virus shedding in rhesus macaques inoculated with SARS-CoV-2. bioRxiv. 2020 doi: 10.1101/2020.03.21.001628.
    1. Murphy T.L., Tussiwand R., Murphy K.M. Specificity through cooperation: BATF-IRF interactions control immune-regulatory networks. Nat. Rev. Immunol. 2013;13:499–509.
    1. Ordovas-Montanes J., Dwyer D.F., Nyquist S.K., Buchheit K.M., Vukovic M., Deb C., Wadsworth M.H., 2nd, Hughes T.K., Kazer S.W., Yoshimoto E. Allergic inflammatory memory in human respiratory epithelial progenitor cells. Nature. 2018;560:649–654.
    1. Ordovas-Montanes J., Beyaz S., Rakoff-Nahoum S., Shalek A.K. Distribution and storage of inflammatory memory in barrier tissues. Nat. Rev. Immunol. 2020 doi: 10.1038/s41577-019-0263-z.
    1. Paules C.I., Marston H.D., Fauci A.S. Coronavirus Infections-More Than Just the Common Cold. JAMA. 2020 doi: 10.1001/jama.2020.0757.
    1. Pinto B.G., Oliveira A.E., Singh Y., Jimenez L., Goncalves A.N., Ogava R.L., Creighton R., Peron J.P., Nakaya H.I. ACE2 Expression is Increased in the Lungs of Patients with Comorbidities Associated with Severe COVID-19. medRxiv. 2020 doi: 10.1101/2020.03.21.20040261.
    1. Prokunina-Olsson L., Alphonse N., Dickenson R.E., Durbin J.E., Glenn J.S., Hartmann R., Kotenko S.V., Lazear H.M., O’Brien T.R., Odendall C. COVID-19 and emerging viral infections: The case for interferon lambda. J. Exp. Med. 2020;217:e20200653. doi: 10.1084/jem.20200653.
    1. Qi F., Qian S., Zhang S., Zhang Z. Single cell RNA sequencing of 13 human tissues identify cell types and receptors of human coronaviruses. bioRxiv. 2020 doi: 10.1101/2020.02.16.951913.
    1. Qian Z., Travanty E.A., Oko L., Edeen K., Berglund A., Wang J., Ito Y., Holmes K.V., Mason R.J. Innate immune response of human alveolar type II cells infected with severe acute respiratory syndrome-coronavirus. Am. J. Respir. Cell Mol. Biol. 2013;48:742–748.
    1. Rabani M., Levin J.Z., Fan L., Adiconis X., Raychowdhury R., Garber M., Gnirke A., Nusbaum C., Hacohen N., Friedman N. Metabolic labeling of RNA uncovers principles of RNA production and degradation dynamics in mammalian cells. Nat. Biotechnol. 2011;29:436–442.
    1. Regev A., Teichmann S.A., Lander E.S., Amit I., Benoist C., Birney E., Bodenmiller B., Campbell P., Carninci P., Clatworthy M., Human Cell Atlas Meeting Participants The Human Cell Atlas. eLife. 2017;6:e27041.
    1. Robinson M.D., McCarthy D.J., Smyth G.K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26:139–140.
    1. Rockx B., Kuiken T., Herfst S., Bestebroer T., Lamers M.M., Oude Munnink B.B., de Meulder D., van Amerongen G., van den Brand J., Okba N.M.A. Comparative pathogenesis of COVID-19, MERS, and SARS in a nonhuman primate model. Science. 2020 doi: 10.1126/science.abb7314.
    1. Rodriguez A., Laio A. Machine learning. Clustering by fast search and find of density peaks. Science. 2014;344:1492–1496.
    1. Ruiz García S., Deprez M., Lebrigand K., Cavard A., Paquet A., Arguel M.J., Magnone V., Truchi M., Caballero I., Leroy S. Novel dynamics of human mucociliary differentiation revealed by single-cell RNA sequencing of nasal epithelial cultures. Development. 2019;146:dev177428. doi: 10.1242/dev.177428.
    1. Rusinova I., Forster S., Yu S., Kannan A., Masse M., Cumming H., Chapman R., Hertzog P.J. Interferome v2.0: an updated database of annotated interferon-regulated genes. Nucleic Acids Res. 2013;41:D1040–D1046.
    1. Russell A.B., Trapnell C., Bloom J.D. Extreme heterogeneity of influenza virus infection in single cells. eLife. 2018;7:e32303. doi: 10.7554/eLife.32303.
    1. Sainz B., Jr., Mossel E.C., Peters C.J., Garry R.F. Interferon-beta and interferon-gamma synergistically inhibit the replication of severe acute respiratory syndrome-associated coronavirus (SARS-CoV) Virology. 2004;329:11–17.
    1. Sajuthi S.P., DeFord P., Jackson N.D., Montgomery M.T., Everman J.L., Rios C.L., Pruesse E., Nolin J.D., Plender E.G., Wechsler M.E. Type 2 and interferon inflammation strongly regulate SARS-CoV-2 related gene expression in the airway epithelium. bioRxiv. 2020 doi: 10.1101/2020.04.09.034454.
    1. Sanda C., Weitzel P., Tsukahara T., Schaley J., Edenberg H.J., Stephens M.A., McClintick J.N., Blatt L.M., Li L., Brodsky L., Taylor M.W. Differential gene induction by type I and type II interferons and their combination. J. Interferon Cytokine Res. 2006;26:462–472.
    1. Satija R., Farrell J.A., Gennert D., Schier A.F., Regev A. Spatial reconstruction of single-cell gene expression data. Nat. Biotechnol. 2015;33:495–502.
    1. Schiller H.B., Montoro D.T., Simon L.M., Rawlins E.L., Meyer K.B., Strunz M., Vieira Braga F.A., Timens W., Koppelman G.H., Budinger G.R.S. The Human Lung Cell Atlas: A High-Resolution Reference Map of the Human Lung in Health and Disease. Am. J. Respir. Cell Mol. Biol. 2019;61:31–41.
    1. Schneider D.S., Ayres J.S. Two ways to survive infection: what resistance and tolerance can teach us about treating infectious diseases. Nat. Rev. Immunol. 2008;8:889–895.
    1. Schneider W.M., Chevillotte M.D., Rice C.M. Interferon-stimulated genes: a complex web of host defenses. Annu. Rev. Immunol. 2014;32:513–545.
    1. Shalek A.K., Satija R., Adiconis X., Gertner R.S., Gaublomme J.T., Raychowdhury R., Schwartz S., Yosef N., Malboeuf C., Lu D. Single-cell transcriptomics reveals bimodality in expression and splicing in immune cells. Nature. 2013;498:236–240.
    1. Shulla A., Heald-Sargent T., Subramanya G., Zhao J., Perlman S., Gallagher T. A transmembrane serine protease is linked to the severe acute respiratory syndrome coronavirus receptor and activates virus entry. J. Virol. 2011;85:873–882.
    1. Skeggs L.T., Dorer F.E., Levine M., Lentz K.E., Kahn J.R. The biochemistry of the renin-angiotensin system. Adv. Exp. Med. Biol. 1980;130:1–27.
    1. Smillie C.S., Biton M., Ordovas-Montanes J., Sullivan K.M., Burgin G., Graham D.B., Herbst R.H., Rogel N., Slyper M., Waldman J. Intra- and Inter-cellular Rewiring of the Human Colon during Ulcerative Colitis. Cell. 2019 doi: 10.1016/j.cell.2019.06.029.
    1. Smits S.L., van den Brand J.M., de Lang A., Leijten L.M., van Ijcken W.F., van Amerongen G., Osterhaus A.D., Andeweg A.C., Haagmans B.L. Distinct severe acute respiratory syndrome coronavirus-induced acute lung injury pathways in two different nonhuman primate species. J. Virol. 2011;85:4234–4245.
    1. Stetson D.B., Medzhitov R. Type I interferons in host defense. Immunity. 2006;25:373–381.
    1. Steuerman Y., Cohen M., Peshes-Yaloz N., Valadarsky L., Cohn O., David E., Frishberg A., Mayo L., Bacharach E., Amit I. Dissection of Influenza Infection In Vivo by Single-Cell RNA Sequencing. Cell Syst. 2018 doi: 10.1016/j.cels.2018.05.008.
    1. Sungnak W., Huang N., Becavin C., Berg M., Network H.L.B. SARS-CoV-2 Entry Genes Are Most Highly Expressed in Nasal Goblet and Ciliated Cells within Human Airways. Nat. Med. 2020 doi: 10.1038/s41591-020-0868-6.
    1. Svensson V. Droplet scRNA-seq is not zero-inflated. Nat. Biotechnol. 2020;38:147–150.
    1. Tepe B., Hill M.C., Pekarek B.T., Hunt P.J., Martin T.J., Martin J.F., Arenkiel B.R. Single-Cell RNA-Seq of Mouse Olfactory Bulb Reveals Cellular Heterogeneity and Activity-Dependent Molecular Census of Adult-Born Neurons. Cell Rep. 2018 doi: 10.1016/j.celrep.2018.11.034.
    1. Trombetta J.J., Gennert D., Lu D., Satija R., Shalek A.K., Regev A. Preparation of Single-Cell RNA-Seq Libraries for Next Generation Sequencing. Curr Protoc Mol Biol. 2014 doi: 10.1002/0471142727.mb0422s107.
    1. Utay N.S., Douek D.C. Interferons and HIV Infection: The Good, the Bad, and the Ugly. Pathog. Immun. 2016;1:107–116.
    1. Vaduganathan M., Vardeny O., Michel T., McMurray J.J.V., Pfeffer M.A., Solomon S.D. Renin-Angiotensin-Aldosterone System Inhibitors in Patients with Covid-19. N. Engl. J. Med. 2020 doi: 10.1056/NEJMsr2005760.
    1. Vieira Braga F.A., Kar G., Berg M., Carpaij O.A., Polanski K., Simon L.M., Brouwer S., Gomes T., Hesse L., Jiang J. A cellular census of human lungs identifies novel cell states in health and in asthma. Nat. Med. 2019;25:1153–1163.
    1. Walls A.C., Park Y.J., Tortorici M.A., Wall A., McGuire A.T., Veesler D. Structure, Function, and Antigenicity of the SARS-CoV-2 Spike Glycoprotein. Cell. 2020;181:281–292.e6.
    1. Wang P.-H., Cheng Y. Increasing Host Cellular Receptor—Angiotensin-Converting Enzyme 2 (ACE2) Expression by Coronavirus may Facilitate 2019-nCoV Infection. bioRxiv. 2020 doi: 10.1101/2020.02.24.963348.
    1. Wang J., Zhuang J., Iyer S., Lin X., Whitfield T.W., Greven M.C., Pierce B.G., Dong X., Kundaje A., Cheng Y. Sequence features and chromatin structure around the genomic regions bound by 119 human transcription factors. Genome Res. 2012;22:1798–1812.
    1. Wang J., Zhuang J., Iyer S., Lin X.Y., Greven M.C., Kim B.H., Moore J., Pierce B.G., Dong X., Virgil D. : a Wiki-based database for transcription factor-binding data generated by the ENCODE consortium. Nucleic Acids Res. 2013;41:D171–D176.
    1. Wang W., Xu Y., Gao R., Lu R., Han K., Wu G., Tan W. Detection of SARS-CoV-2 in Different Types of Clinical Specimens. JAMA. 2020 doi: 10.1001/jama.2020.3786.
    1. Weinheimer V.K., Becher A., Tönnies M., Holland G., Knepper J., Bauer T.T., Schneider P., Neudecker J., Rückert J.C., Szymanski K. Influenza A viruses target type II pneumocytes in the human lung. J. Infect. Dis. 2012;206:1685–1694.
    1. Wolf F.A., Angerer P., Theis F.J. SCANPY: large-scale single-cell gene expression data analysis. Genome Biol. 2018;19:15. doi: 10.1186/s13059-017-1382-0.
    1. Wölfel R., Corman V.M., Guggemos W., Seilmaier M., Zange S., Müller M.A., Niemeyer D., Jones T.C., Vollmar P., Rothe C. Virological assessment of hospitalized patients with COVID-2019. Nature. 2020 doi: 10.1038/s41586-020-2196-x.
    1. Wrapp D., Wang N., Corbett K.S., Goldsmith J.A., Hsieh C.L., Abiona O., Graham B.S., McLellan J.S. Cryo-EM structure of the 2019-nCoV spike in the prefusion conformation. Science. 2020;367:1260–1263.
    1. Wu C., Zheng S., Chen Y., Zheng M. Single-cell RNA expression profiling of ACE2, the putative receptor of Wuhan 2019-nCoV, in the nasal tissue. medRxiv. 2020 doi: 10.1101/2020.02.11.20022228.
    1. Xiao F., Tang M., Zheng X., Liu Y., Li X., Shan H. Evidence for gastrointestinal infection of SARS-CoV-2. Gastroenterology. 2020 doi: 10.1053/j.gastro.2020.02.055.
    1. Xu Y., Li X., Zhu B., Liang H., Fang C., Gong Y., Guo Q., Sun X., Zhao D., Shen J. Characteristics of pediatric SARS-CoV-2 infection and potential evidence for persistent fecal viral shedding. Nat. Med. 2020;26:502–505.
    1. Young M.D., Behjati S. SoupX removes ambient RNA contamination from droplet based single-cell RNA sequencing data. bioRxiv. 2020 doi: 10.1101/303727.
    1. Zhang H., Kang Z., Gong H., Xu D., Wang J., Li Z., Cui X., Xiao J., Meng T., Zhou W. The digestive system is a potential route of 2019-nCov infection: a bioinformatics analysis based on single-cell transcriptomes. bioRxiv. 2020 doi: 10.1101/2020.01.30.927806.
    1. Zhao X., Guo F., Liu F., Cuconati A., Chang J., Block T.M., Guo J.T. Interferon induction of IFITM proteins promotes infection by human coronavirus OC43. Proc. Natl. Acad. Sci. USA. 2014;111:6756–6761.
    1. Zheng B., He M.L., Wong K.L., Lum C.T., Poon L.L., Peng Y., Guan Y., Lin M.C., Kung H.F. Potent inhibition of SARS-associated coronavirus (SCOV) infection and replication by type I interferons (IFN-alpha/beta) but not by type II interferon (IFN-gamma) J. Interferon Cytokine Res. 2004;24:388–390.
    1. Zou Z., Yan Y., Shu Y., Gao R., Sun Y., Li X., Ju X., Liang Z., Liu Q., Zhao Y. Angiotensin-converting enzyme 2 protects from lethal avian influenza A H5N1 infections. Nat. Commun. 2014;5:3594.
    1. Zou L., Ruan F., Huang M., Liang L., Huang H., Hong Z., Yu J., Kang M., Song Y., Xia J. SARS-CoV-2 Viral Load in Upper Respiratory Specimens of Infected Patients. N. Engl. J. Med. 2020;382:1177–1179.

Source: PubMed

3
Subscribe