Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

ACE2-dependent and -independent SARS-CoV-2 entries dictate viral replication and inflammatory response during infection

Abstract

Excessive inflammation is the primary cause of mortality in patients with severe COVID-19, yet the underlying mechanisms remain poorly understood. Our study reveals that ACE2-dependent and -independent entries of SARS-CoV-2 in epithelial cells versus myeloid cells dictate viral replication and inflammatory responses. Mechanistically, SARS-CoV-2 NSP14 potently enhances NF-κB signalling by promoting IKK phosphorylation, while SARS-CoV-2 ORF6 exerts an opposing effect. In epithelial cells, ACE2-dependent SARS-CoV-2 entry enables viral replication, with translated ORF6 suppressing NF-κB signalling. In contrast, in myeloid cells, ACE2-independent entry blocks the translation of ORF6 and other viral structural proteins due to inefficient subgenomic RNA transcription, but NSP14 could be directly translated from genomic RNA, resulting in an abortive replication but hyperactivation of the NF-κB signalling pathway for proinflammatory cytokine production. Importantly, we identified TLR1 as a critical factor responsible for viral entry and subsequent inflammatory response through interaction with E and M proteins, which could be blocked by the small-molecule inhibitor Cu-CPT22. Collectively, our findings provide molecular insights into the mechanisms by which strong viral replication but scarce inflammatory response during the early (ACE2-dependent) infection stage, followed by low viral replication and potent inflammatory response in the late (ACE2-independent) infection stage, may contribute to COVID-19 progression.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: SARS-CoV-2 NSP14 facilitates the activation of NF-κB signalling and inflammatory response.
Fig. 2: SARS-CoV-2 NSP14 upregulates NF-κB signalling by promoting phosphorylation of the IKK complex.
Fig. 3: SARS-CoV-2 ORF6 shuts down NF-κB signalling by blocking the nuclear translocation of p65.
Fig. 4: The expression of ACE2 is distinctly associated with productive viral replication and the inflammatory response in epithelial and myeloid cells.
Fig. 5: ACE2-dependent and -independent SARS-CoV-2 entries dictate the transcription of viral subgenomic RNA in epithelial and myeloid cells.
Fig. 6: Ectopic expression of ACE2 in myeloid cells confers SARS-CoV-2 productive replication but hampers the virus-induced inflammatory response.
Fig. 7: TLR1 is a critical entry factor responsible for SARS-CoV-2 entry and virus-induced inflammatory response in myeloid cells.
Fig. 8: TLR1-mediated SARS-CoV-2 entry in myeloid cells through the recognition and binding of E and M proteins.

Similar content being viewed by others

Data availability

RNA-seq data that support the findings of this study have been deposited in the Gene Expression Omnibus under accession code GSE249500. All of the other data supporting the findings of this study are available from the corresponding author upon reasonable request. Source data are provided with this paper.

References

  1. Guan, W.-J. et al. Clinical characteristics of coronavirus disease 2019 in China. N. Engl. J. Med. 382, 1708–1720 (2020).

    Article  CAS  PubMed  Google Scholar 

  2. Hoehl, S. et al. Evidence of SARS-CoV-2 infection in returning travelers from Wuhan. China 382, 1278–1280 (2020).

    Google Scholar 

  3. Wu, F. et al. A new coronavirus associated with human respiratory disease in China. Nature 579, 265–269 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Zhou, P. et al. A pneumonia outbreak associated with a new coronavirus of probable bat origin. Nature 579, 270–273 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Zhu, N. et al. A novel coronavirus from patients with pneumonia in China, 2019. N. Engl. J. Med. 382, 727–733 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Huang, C. et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet 395, 497–506 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Diamond, M. S. & Kanneganti, T.-D. Innate immunity: the first line of defense against SARS-CoV-2. Nat. Immunol. 23, 165–176 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Duan, T., Du, Y., Xing, C., Wang, H. Y. & Wang, R.-F. Toll-like receptor signaling and its role in cell-mediated immunity. Front. Immunol. 13, 812774 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Janssen, N. A. F. et al. Dysregulated innate and adaptive immune responses discriminate disease severity in COVID-19. J. Infect. Dis. 223, 1322–1333 (2021).

    Article  CAS  PubMed  Google Scholar 

  10. Xing, C. et al. Interaction between microbiota and immunity and its implication in colorectal cancer. Front. Immunol. 13, 963819 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Blot, M. et al. The dysregulated innate immune response in severe COVID-19 pneumonia that could drive poorer outcome. J. Transl. Med. 18, 457 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Ren, L. L. et al. Identification of a novel coronavirus causing severe pneumonia in human: a descriptive study. Chin. Med. J. 133, 1015–1024 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Liu, B., Li, M., Zhou, Z., Guan, X. & Xiang, Y. Can we use interleukin-6 (IL-6) blockade for coronavirus disease 2019 (COVID-19)-induced cytokine release syndrome (CRS)? J. Autoimmun. 111, 102452 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Long, Q.-X. et al. Clinical and immunological assessment of asymptomatic SARS-CoV-2 infections. Nat. Med. 26, 1200–1204 (2020).

    Article  CAS  PubMed  Google Scholar 

  15. Del Valle, D. M. et al. An inflammatory cytokine signature predicts COVID-19 severity and survival. Nat. Med. 26, 1636–1643 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Zhang, X. et al. Viral and host factors related to the clinical outcome of COVID-19. Nature 583, 437–440 (2020).

    Article  CAS  PubMed  Google Scholar 

  17. Mehta, P. et al. COVID-19: consider cytokine storm syndromes and immunosuppression. Lancet 395, 1033–1034 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Moore, J. B. & June, C. H. Cytokine release syndrome in severe COVID-19. Science 368, 473–474 (2020).

    Article  CAS  PubMed  Google Scholar 

  19. Schett, G., Sticherling, M. & Neurath, M. F. COVID-19: risk for cytokine targeting in chronic inflammatory diseases? Nat. Rev. Immunol. 20, 271–272 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Zhang, C., Wu, Z., Li, J.-W., Zhao, H. & Wang, G.-Q. Cytokine release syndrome in severe COVID-19: interleukin-6 receptor antagonist tocilizumab may be the key to reduce mortality. Int. J. Antimicrob. Agents 55, 105954 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Liao, M. F. et al. Single-cell landscape of bronchoalveolar immune cells in patients with COVID-19. Nat. Med. 26, 842–844 (2020).

    Article  CAS  PubMed  Google Scholar 

  22. Giamarellos-Bourboulis, E. J. et al. Complex immune dysregulation in COVID-19 patients with severe respiratory failure. Cell Host Microbe 27, 992–1000 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Gong, J. et al. Correlation analysis between disease severity and inflammation-related parameters in patients with COVID-19: a retrospective study. BMC Infect. Dis. 20, 963 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Vabret, N. et al. Immunology of COVID-19: current state of the science. Immunity 52, 910–941 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Hoffmann, M. et al. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell 181, 271–280 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bost, P. et al. Host-viral infection maps reveal signatures of severe COVID-19 patients. Cell 181, 1475–1488 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Grant, R. A. et al. Circuits between infected macrophages and T cells in SARS-CoV-2 pneumonia. Nature 590, 635–641 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Boumaza, A. et al. Monocytes and macrophages, targets of severe acute respiratory syndrome coronavirus 2: the clue for coronavirus disease 2019 immunoparalysis. J. Infect. Dis. 224, 395–406 (2021).

    Article  CAS  PubMed  Google Scholar 

  29. Yang, D. et al. Attenuated interferon and proinflammatory response in SARS-CoV-2-infected human dendritic cells is associated with viral antagonism of STAT1 phosphorylation. J. Infect. Dis. 222, 734–745 (2020).

    Article  CAS  PubMed  Google Scholar 

  30. Junqueira, C. et al. FcγR-mediated SARS-CoV-2 infection of monocytes activates inflammation. Nature 606, 576–584 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jalloh, S. et al. CD169-mediated restrictive SARS-CoV-2 infection of macrophages induces pro-inflammatory responses. PLoS Pathog. 18, e1010479 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Hui, K. P. et al. Tropism, replication competence, and innate immune responses of the coronavirus SARS-CoV-2 in human respiratory tract and conjunctiva: an analysis in ex-vivo and in-vitro cultures. Lancet Resp. Med. 8, 687–695 (2020).

    Article  CAS  Google Scholar 

  33. Martínez-Colón, G. J. et al. SARS-CoV-2 infection drives an inflammatory response in human adipose tissue through infection of adipocytes and macrophages. Sci. Transl. Med. 14, eabm9151 (2022).

    Article  PubMed  Google Scholar 

  34. Gordon, D. E. et al. A SARS-CoV-2 protein interaction map reveals targets for drug repurposing. Nature 583, 459–468 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Hsu, J. C.-C., Laurent-Rolle, M., Pawlak, J. B., Wilen, C. B. & Cresswell, P. Translational shutdown and evasion of the innate immune response by SARS-CoV-2 NSP14 protein. Proc. Natl Acad. Sci. USA 118, e2101161118 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Hayden, M. S. & Ghosh, S. Shared principles in NF-κB signaling. Cell 132, 344–362 (2008).

    Article  CAS  PubMed  Google Scholar 

  37. Gao, X. et al. Structural basis for Sarbecovirus ORF6 mediated blockage of nucleocytoplasmic transport. Nat. Commun. 13, 4782 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Li, T. et al. Molecular mechanism of SARS-CoVs Orf6 targeting the Rae1–Nup98 complex to compete with mRNA nuclear export. Front. Mol. Biosci. 8, 813248 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Addetia, A. et al. SARS-CoV-2 ORF6 disrupts bidirectional nucleocytoplasmic transport through interactions with Rae1 and Nup98. mBio 12, e00065-21 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  40. Liang, P. et al. KPNB1, XPO7 and IPO8 mediate the translocation of NF-κB/p65 into the nucleus. Traffic 14, 1132–1143 (2013).

    Article  CAS  PubMed  Google Scholar 

  41. Radu, A., Moore, M. S. & Blobel, G. The peptide repeat domain of nucleoporin Nup98 functions as a docking site in transport across the nuclear pore complex. Cell 81, 215–222 (1995).

    Article  CAS  PubMed  Google Scholar 

  42. Bayliss, R., Littlewood, T. & Stewart, M.Structural basis for the interaction between FxFG nucleoporin repeats and importin-β in nuclear trafficking. Cell 102, 99–108 (2000).

    Article  CAS  PubMed  Google Scholar 

  43. Shen, Q., Wang, Y. E. & Palazzo, A. F. Crosstalk between nucleocytoplasmic trafficking and the innate immune response to viral infection. J. Biol. Chem. 297, 100856 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Bonifaci, N., Moroianu, J., Radu, A. & Blobel, G. Karyopherin β2 mediates nuclear import of a mRNA binding protein. Proc. Natl Acad. Sci. USA 94, 5055–5060 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Fontoura, B. M., Blobel, G. N. & Yaseen, N. R. The nucleoporin Nup98 is a site for GDP/GTP exchange on ran and termination of karyopherin β2-mediated nuclear import. J. Biol. Chem. 275, 31289–31296 (2000).

    Article  CAS  PubMed  Google Scholar 

  46. Chook, Y. & Blobel, G. Karyopherins and nuclear import. Curr. Opin. Struct. Biol. 11, 703–715 (2001).

    Article  CAS  PubMed  Google Scholar 

  47. Lott, K. & Cingolani, G. The importin β binding domain as a master regulator of nucleocytoplasmic transport. Biochim. Biophys. Acta 1813, 1578–1592 (2011).

    Article  CAS  PubMed  Google Scholar 

  48. Xie, X. et al. An infectious cDNA clone of SARS-CoV-2. Cell Host Microbe 27, 841–848 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. V’kovski, P., Kratzel, A., Steiner, S., Stalder, H. & Thiel, V. Coronavirus biology and replication: implications for SARS-CoV-2. Nat. Rev. Microbiol. 19, 155–170 (2021).

    Article  PubMed  Google Scholar 

  50. Baggen, J., Vanstreels, E., Jansen, S. & Daelemans, D. Cellular host factors for SARS-CoV-2 infection. Nat. Microbiol. 6, 1219–1232 (2021).

    Article  CAS  PubMed  Google Scholar 

  51. Mulay, A. et al. SARS-CoV-2 infection of primary human lung epithelium for COVID-19 modeling and drug discovery. Cell Rep. 35, 109055 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Hartenian, E. et al. The molecular virology of coronaviruses. J. Biol. Chem. 295, 12910–12934 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Irigoyen, N. et al. High-resolution analysis of coronavirus gene expression by RNA sequencing and ribosome profiling. PLoS Pathog. 12, e1005473 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Zarember, K. A. & Godowski, P. J. Tissue expression of human Toll-like receptors and differential regulation of Toll-like receptor mRNAs in leukocytes in response to microbes, their products, and cytokines. J. Immunol. 168, 554–561 (2002).

    Article  CAS  PubMed  Google Scholar 

  55. Kawasaki, T. & Kawai, T. Toll-like receptor signaling pathways. Front Immunol. 5, 461 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  56. Li, S. et al. Porcine epidemic diarrhea virus nsp14 inhibits NF-κB pathway activation by targeting the IKK complex and p65. Anim. Dis. 1, 24 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Queromes, G. et al. Characterization of SARS-CoV-2 ORF6 deletion variants detected in a nosocomial cluster during routine genomic surveillance, Lyon, France. Emerg. Microbes Infect. 10, 167–177 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Miyamoto, Y. et al. SARS-CoV-2 ORF6 disrupts nucleocytoplasmic trafficking to advance viral replication. Commun. Biol. 5, 483 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Gori Savellini, G., Anichini, G., Gandolfo, C. & Cusi, M. G. Nucleopore traffic is hindered by SARS-CoV-2 ORF6 protein to efficiently suppress IFN-β and IL-6 secretion. Viruses 14, 1273 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Kato, K. et al. Overexpression of SARS-CoV-2 protein ORF6 dislocates RAE1 and NUP98 from the nuclear pore complex. Biochem. Biophys. Res. Commun. 536, 59–66 (2021).

    Article  CAS  PubMed  Google Scholar 

  61. Schultze, J. L. & Aschenbrenner, A. C. COVID-19 and the human innate immune system. Cell 184, 1671–1692 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Blanco-Melo, D. et al. Imbalanced host response to SARS-CoV-2 drives development of COVID-19. Cell 181, 1036 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Merad, M. & Martin, J. C. Pathological inflammation in patients with COVID-19: a key role for monocytes and macrophages. Nat. Rev. Immunol. 20, 355–362 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Writing Committee for the REMAP-CAP Investigators et al. Long-term (180-day) outcomes in critically ill patients with COVID-19 in the REMAP-CAP randomized clinical trial. J. Am. Med. Assoc. 329, 39–51 (2023).

    Article  Google Scholar 

  65. Ziegler, C. G. K. et al. SARS-CoV-2 receptor ACE2 is an interferon-stimulated gene in human airway epithelial cells and is detected in specific cell subsets across tissues. Cell 181, 1016–1035 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Wang, K. et al. CD147-spike protein is a novel route for SARS-CoV-2 infection to host cells. Signal Transduct. Target. Ther. 5, 283 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Morniroli, D., Gianni, M. L., Consales, A., Pietrasanta, C. & Mosca, F. Human sialome and coronavirus disease-2019 (COVID-19) pandemic: an understated correlation? Front. Immunol. 11, 1480 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Daly, J. L. et al. Neuropilin-1 is a host factor for SARS-CoV-2 infection. Science 370, 861–865 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Cantuti-Castelvetri, L. et al. Neuropilin-1 facilitates SARS-CoV-2 cell entry and infectivity. Science 370, 856–860 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Clausen, T. M. et al. SARS-CoV-2 infection depends on cellular heparan sulfate and ACE2. Cell 183, 1043–1057 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Lu, Q. et al. SARS-CoV-2 exacerbates proinflammatory responses in myeloid cells through C-type lectin receptors and Tweety family member 2. Immunity 54, 1304–1319 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. McClure, R. & Massari, P. TLR-dependent human mucosal epithelial cell responses to microbial pathogens. Front. Immunol. 5, 386 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  73. Bir, S. C., Chernyshev, O. Y. & Minagar, A. in Neuroinflammation 2nd edn (ed. Minagar, A.) 541–562 (Academic Press, 2018).

  74. Juarez, E. et al. Differential expression of Toll-like receptors on human alveolar macrophages and autologous peripheral monocytes. Resp. Res. 11, 2 (2010).

    Article  Google Scholar 

  75. Cheung, C. Y. et al. Cytokine responses in severe acute respiratory syndrome coronavirus-infected macrophages in vitro: possible relevance to pathogenesis. J. Virol. 79, 7819–7826 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Yilla, M. et al. SARS-coronavirus replication in human peripheral monocytes/macrophages. Virus Res. 107, 93–101 (2005).

    Article  CAS  PubMed  Google Scholar 

  77. Tang, T., Bidon, M., Jaimes, J. A., Whittaker, G. R. & Daniel, S. Coronavirus membrane fusion mechanism offers a potential target for antiviral development. Antivir. Res. 178, 104792 (2020).

    Article  CAS  PubMed  Google Scholar 

  78. Imprachim, N., Yosaatmadja, Y. & Newman, J. A. Crystal structures and fragment screening of SARS-CoV-2 NSP14 reveal details of exoribonuclease activation and mRNA capping and provide starting points for antiviral drug development. Nucleic Acids Res. 51, 475–487 (2022).

    Article  PubMed Central  Google Scholar 

  79. Du, Y. et al. Activation of cGAS-STING by lethal malaria N67C dictates immunity and mortality through induction of CD11b+ Ly6Chi proinflammatory monocytes. Adv. Sci. 9, e2103701 (2022).

    Article  Google Scholar 

  80. Feng, Y. et al. LRRC25 functions as an inhibitor of NF-κB signaling pathway by promoting p65/RelA for autophagic degradation. Sci. Rep. 7, 13448 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  81. Lu, H. et al. Potent NKT cell ligands overcome SARS-CoV-2 immune evasion to mitigate viral pathogenesis in mouse models. PLoS Pathog. 19, e1011240 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Batéjat, C., Grassin, Q., Manuguerra, J. C. & Leclercq, I. Heat inactivation of the severe acute respiratory syndrome coronavirus 2. J. Biosaf. Biosec. 3, 1–3 (2021).

    Article  Google Scholar 

  83. Chu, J. et al. Pharmacological inhibition of fatty acid synthesis blocks SARS-CoV-2 replication. Nat. Metab. 3, 1466–1475 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Du, Y. et al. LRRC25 inhibits type I IFN signaling by targeting ISG15-associated RIG-I for autophagic degradation. EMBO J. 37, 351–366 (2018).

    Article  CAS  PubMed  Google Scholar 

  85. Duan, T. et al. USP3 plays a critical role in the induction of innate immune tolerance. EMBO Rep. 24, e57828 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Liu, X. et al. Development of a TCR-like antibody and chimeric antigen receptor against NY-ESO-1/HLA-A2 for cancer immunotherapy. J. Immunother. Cancer 10, e004035 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  87. Jin, S. et al. Tetherin suppresses type I interferon signaling by targeting MAVS for NDP52-mediated selective autophagic degradation in human cells. Mol. Cell 68, 308–322 (2017).

    Article  CAS  PubMed  Google Scholar 

  88. Xing, C. et al. Microbiota regulate innate immune signaling and protective immunity against cancer. Cell Host Microbe 29, 959–974 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank Y. Chen and M. Li from the USC Libraries Bioinformatics Service for help with the bioinformatics analysis of the RNA-seq data. We thank Q. Zhu and J. Xu from Tsinghua University for help with the bioinformatics analysis of the bronchoalveolar lavage fluid macrophage bulk RNA-seq data. We thank W. D. Wallace and M.-L. Chen from the Translational Pathology Core of the USC for providing the autopsy specimens from patients with lung cancer and COVID-19. We thank N. J. Krogan, D. E. Gordon, J. Z. Guo and M. Soucheray from the University of California, San Francisco for providing the plasmids encoding 27 SARS-CoV-2 viral proteins. We thank W. Yuan from the USC for providing the plasmids encoding E proteins of other human coronaviruses. We also thank: P.-Y. Shi from the University of Texas Medical Branch for the generous gift of the SARS-CoV-2-mNeoGreen virus; L. Comai and J. Henley from the USC BSL-3 Core for support and resources; and K. O’Brien from USC Environmental Health and Safety for biosafety support. This work was in part supported by the USC Startup Fund, as well as grants from the NCI/NIH (R01CA101795 and R01CA246547) and Department of Defense Congressionally Directed Medical Research Programs (Breast Cancer Research Program (BC151081) and Lung Cancer Research Program (LC200368)) to R.-F.W.

Author information

Authors and Affiliations

Authors

Contributions

R.-F.W. supervised the entire study. T.D., C.X., J.C. and R.-F.W. conceived of and designed the study. T.D. performed all of the experiments, with assistance from C.X., J.C., X.D., Y.D., X.L., Y.H., C.Q. and B.Y. in some of the experiments. C.X., J.C. and X.D. performed the SARS-CoV-2 infection studies in the BSL-3 and Animal BSL-3 facilities. T.D., C.X., J.C., H.Y.W. and R.-F.W. collected and analysed the data and wrote the manuscript with input from all of the authors.

Corresponding author

Correspondence to Rong-Fu Wang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Xiangxi Wang and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 SARS-CoV-2 NSP14 promotes the virus-/TLR-induced inflammatory response.

(a) Quantitative PCR with reverse transcription (RT-PCR) analysis of proinflammatory relative cytokines genes (IL6, IL1B, TNF, IL8) of THP-1 cells with SARS-CoV-2 or VSV-GFP infection at an MOI = 1 for indicated time points. (b) Sixteen nonstructural proteins, four structural proteins, and nine accessory proteins of SARS-CoV-2 were delineated. (c) RT-PCR analysis of SARS-CoV-2 NSP14 in THP-1 or MDM transfected with NSP14-specific siRNAs or control scramble siRNA, followed by SARS-CoV-2 infection at an MOI = 1 for 48 hrs. (d) Expression of proinflammatory genes in THP-1 transfected with NSP14-specific or control scramble siRNA, followed by SARS-CoV-2 infection at an MOI = 1 for 48 hrs. (e) Expression of proinflammatory genes in doxycycline-inducible SARS-CoV-2 NSP14 expression THP-1 cells was measured after being treated with doxycycline (1 μg/ml) or PBS for 24 h, followed by LPS (1 μg/ml) stimulation at indicated periods. Data in (a, c-e) are plotted as mean ± SD (n = 3 independent samples). Statistical analyses were performed using one-way ANOVA followed by Dunnett’s post-test (c, d) or two-way ANOVA followed by Sidak post-test (e).

Source data

Extended Data Fig. 2 SARS-CoV-2 NSP14 promotes the phosphorylation of IKKs in an N7-MTase activity-independent manner.

(a) Co-immunoprecipitation and immunoassay of extracts of HEK293T cells transfected with the plasmids for Flag-tagged MyD88, TRAF6, TAK1 + TAB2, ΝΕΜΟ (inhibitor of NF-κB kinase subunit gamma (IKKγ)) or p65, together with NSP14. (b-c) The recombinant NEMO protein was directly incubated with recombinant NSP14 protein pre-bound on the anti-Flag-beads (b) or Strep-beads (c) overnight at 4 °C, followed by immunoprecipitation. (d, e) HEK293T cells were transfected with an NF-κB-luc, together with the vector for IKKα (d) and IKKβ (e), along with an empty vector or expression vector for NSP14 or its mutants. (f) HEK293T cells were transfected with an NF-κB-luc, together with vector for MyD88 and NSP14, along with DMSO or increasing amounts (wedge) of Nitazoxanide (1 nM, 10 nM, 100 nM) and Sinefungin (100 nM, 1 μM, 10 μM) treatment. (g) Expression of the proinflammatory genes in MDMs treated with DMSO (no wedge) or Nitazoxanide (100 nM) or sinefungin (10 μM), followed by SARS-CoV-2 infection at an MOI = 1 for 48 h. (h) Sequence alignment of NSP14 from seven pathogenic CoVs (NL63, 299E, OC43, HKU1, MERS-CoV, SARS-CoV, and SARS-CoV-2). (i) HEK293T cells were transfected with an NF-κB-luc, together with the vector for IKKα and the adjusted amount of expression vector for SARS-CoV-2 NSP14 or its mutants. (j) Confocal microscopy is shown in HEK293T cells co-transfected with BFP-tagged IKKα and GFP-tagged SARS-CoV-2 NSP14 or its point mutants (Scale bar: 10 μm). (k) The recombinant protein of GFP-IKKβ, along with or without NSP14 WT or its mutant, was directly incubated with recombinant Flag-IKKβ pre-bound on the anti-Flag beads overnight at 4 °C, followed by immunoprecipitation. (l) HEK293T cells were transfected with an NF-κB-luc, together with a vector for IKKα and increasing amounts (wedge) of an expression vector for SARS-CoV-2 ancestral NSP14 or Omicron variant NSP14. Data in (a, b, c, j, k) are the representative of multiple independent experiments, data in (d-g, i, l) are plotted as the mean ± SD (n = 3 independent samples). Statistical analyses were performed using one-way ANOVA followed by Dunnett’s post-test (d, e, l) or two-tailed Student’s t-test (f, g, i). ns, not significant.

Source data

Extended Data Fig. 3 SARS-CoV-2 ORF6 attenuates the TLR-induced inflammatory response through its C-terminal tail.

(a) Western blot to show the expression of GFP-ORF6 in the THP-1 cells. (b) Production of cytokines TNFα and IL-6 in the culture medium of control or SARS-CoV-2 ORF6 THP-1 cells were measured after indicated TLRs ligands treatment for 24 h. (c) Co-immunoprecipitation and immunoassay of extracts of HEK293T cells transfected with the plasmids for Flag-tagged MyD88, TRAF6, TAK1, IKKα, IKKβ, NEMO (IKKγ) or p65, together without or with an expression vector for SARS-CoV-2 ORF6. (d) The fluorescence intensity distribution of GFP signaling and Hochest signaling on the defined red lines in Fig. 3c (GFP-ORF6) were quantified using the NIS-Elements Imaging Software. The red dotted line box marked the fluorescence intensity located on the nuclear membrane. (e) Alignment of representative sequences across different sarbecoviruses species. The C-terminal tail of ORF6 is highly conserved, as illustrated here by WebLogo (http://weblogo.berkeley.edu). (f) HEK293T cells were transfected with an NF-κB-luc, together with the vector for p65, along with an empty vector or adjusted amount of expression vector for SARS-CoV-2 ORF6 or its mutants. (g) The fluorescence intensity distribution of GFP signaling and Hochest signaling on the defined red lines in Fig. 3c (GFP-ORF6 M58A) were quantified using the NIS-Elements Imaging Software. The red dotted line box marked the fluorescence intensity located on the nuclear membrane. (h) HEK293T cells were transfected with plasmids for p65, Myc-KPNA2, HA-KPNB1 and Flag-Nup98 in the indicated combinations, together with or without an expression vector for GFP-tagged SARS-CoV-2 ORF6. After transfection, whole-cell lysates were collected and used for immunoprecipitation with anti-Flag beads. Subsequently, immunoblot analysis was carried out using the specified antibodies. Data in (a, c) are the representative of multiple independent experiments, data in (b) and (f) are plotted as the mean ± SD (n = 3 independent samples). Statistical analyses were performed using two-way ANOVA followed by Sidak post-test (b) or one-way ANOVA followed by Dunnett’s post-test (f).

Source data

Extended Data Fig. 4 SARS-CoV-2 ORF6 downregulates genes associated with the TLR signaling pathway and the NF-κB signaling pathway in myeloid cells.

(a) KEGG pathway enrichment analysis shows that the differentially expressed genes (DEGs) were enriched in multiple innate immune signaling pathways. (b) Gene-set enrichment analysis (GSEA) of genes with differential expression in ORF6-electroporated THP-1 cells compared to EV-electroporated control cells 4 hours after Pam3CSK4 stimulation. (c) The volcano plot displays DEGs when comparing ORF6-electroporated THP-1 cells to EV-electroporated control cells 4 hours after Pam3CSK4 stimulation. The p-value and fold change are calculated using gene-specific analysis (GSA) with the default settings of Partek Flow. (d) The heatmap shows some downregulated DEGs in the TLR and NF-κB signaling pathway.

Source data

Extended Data Fig. 5 Myeloid cells are not efficiently ‘infected’ but are highly activated by SARS-CoV-2.

(a) Quantitative PCR with reverse transcription analysis of SARS-CoV-2 S proteins in Calu-3, Caco-2, NCI-H1355, NHBE, Macrophage or MDM with SARS-CoV-2 infection at an MOI = 0.05 (Caco-2 and NCI-H1355), or an MOI = 0.5 (Calu-3), or an MOI = 1 (NHBE, Macrophage, MDM) for indicated time points. (b) Quantitative PCR with reverse transcription analysis of SARS-CoV-2 S proteins in mouse BMDM (Bone marrow-derived macrophage) with SARS-CoV-2 infection at MOI = 1 for indicated time points. (c, d) Expression of proinflammatory genes in mouse BMDM was measured after SARS-CoV-2 infection at MOI = 1 for indicated time points. (e) Expression of the proinflammatory genes in Calu-3, NCI-H1355, Caco-2, NHBE, Macrophage or MDM with SARS-CoV-2 infection at an MOI = 0.05 (NCI-H1355 and Caco-2), or an MOI = 0.5 (Calu-3), or an MOI = 1 (NHBE, Macrophage and MDM) for indicated time points. Data in (a-e) are plotted as mean ± SD (n = 3 independent samples).

Source data

Extended Data Fig. 6 The subgenomic RNA of SARS-CoV-2 could only be highly transcribed in ACE2-positive epithelial cells but not in ACE2-negative myeloid cells.

(a) Quantitative PCR with reverse transcription (RT-PCR) analysis of RNA of SARS-CoV-2 NSP1, NSP14, ORF6, S protein, and N protein or subgenomic RNA of ORF6, S, and N proteins in epithelial cells (Caco-2 and NCI-H1355) or myeloid cells (Macrophage and MDM) after SARS-CoV-2 infection at an MOI = 0.05 (epithelial cells) or an MOI = 1 (myeloid cells) for 24 h. (b) RT-PCR analysis of subgenomic RNA of SARS-CoV-2 S protein, N protein, and ORF6 in Calu-3, NHBE, Macrophage, and MDM after SARS-CoV-2 infection at an MOI = 0.5 (Calu-3) or an MOI = 1 (NHBE, Macrophage, MDM) for indicated time points. (c, d) Super-resolution microscopic (Airyscan) analysis of NSP12 (c) and S (d) proteins, along with CD68 in the lung autopsy sections from the COVID-19 patients or in normal lung sections from uninfected lung cancer patients (Ctrl). (Scale bars 20 μm). (e) Quantitative PCR with reverse transcription analysis of SARS-CoV-2 ORF6 subgenomic RNA in NCI-H1355 cells transduced with Ctrl-sgRNA or ACE2-sgRNA, followed with SARS-CoV-2 infection at an MOI = 1 for the indicated lengths of time. (f) Quantitative PCR with reverse transcription analysis of SARS-CoV-2 ORF6 subgenomic RNA in Caco-2 cells transduced with Ctrl-sgRNA or ACE2-sgRNA, followed with SARS-CoV-2 infection at an MOI = 1 for the indicated lengths of time. (g) Expression of the proinflammatory genes in NCI-H1355 transduced with Ctrl-sgRNA or ACE2-sgRNA was measured after being treated with SARS-CoV-2 infection at an MOI = 1 for the indicated lengths of time. (h) Expression of the proinflammatory genes in Caco-2 cells transfected with Ctrl-sgRNA or ACE2-sgRNA was measured after being treated with SARS-CoV-2 infection at an MOI = 1 for the indicated lengths of time. Data in (c, d) are the representative of multiple independent experiments, data in (a, b, e-h) are plotted as the mean ± SD (n = 3 independent samples). Statistical analyses were performed using two-way ANOVA followed by Sidak post-test (g, h).

Source data

Extended Data Fig. 7 SARS-CoV-2 enters myeloid cells in a Spike- and ACE2-independent manner.

(a) Quantitative PCR with reverse transcription (RT-PCR) analysis of RNA of SARS-CoV-2 S, RdRp (NSP12), and NSP1 proteins in MDMs, THP-1WT or THP-1ACE2 cells pretreated with Human BD Fc Block™ (Fc Block), along with control IgG or anti-ACE2 blocking antibody (2 μg/ml or 20 μg/ml), followed by SARS-CoV-2 infection at an MOI = 1 for 48 h. (b) RT-PCR analysis of the RNA of SARS-CoV-2 S, RdRp (NSP12), and NSP1 proteins in MDM, THP-1WT or THP-1ACE2 cells pretreated with Fc blocker, along with control IgG or anti-S protein neutralizing monoclonal antibody (2 μg/ml or 20 μg/ml), followed by SARS-CoV-2 infection at an MOI = 1 for 48 h. (c) RT-PCR analysis of RNA of SARS-CoV-2 S protein in THP-1 WT cells, Macrophages, MDMs or THP-1ACE2 cells pretreated with DMSO, MM3122 (2 μM or 10 μM), E64d (2 μM or 10 μM), Chloroquine (2 μM or 10 μM) and EK1C4 (20 nM or 100 nM), followed by SARS-CoV-2 infection at an MOI = 1 for 48 h. (d) RT-PCR analysis of RNA of SARS-CoV-2 RdRp in THP-1 WT cells, Macrophages, MDMs or THP-1ACE2 cells pretreated with DMSO, MM3122 (2 μM or 10 μM), E64d (2 μM or 10 μM), Chloroquine (2 μM or 10 μM) and EK1C4 (20 nM or 100 nM), followed by SARS-CoV-2 infection at an MOI = 1 for 48 h. (e) RT-PCR analysis of RNA of SARS-CoV-2 S protein in THP-1 WT cells, or THP-1ACE2 cells pretreated with Fc blocker, along with normal serum or pooled convalescent serum (0.2 % or 1 %) from monkey (BEI Resources, NR-52401), followed by SARS-CoV-2 infection at an MOI = 1 for 48 h. (f) RT-PCR analysis of RNA of SARS-CoV-2 RdRp in THP-1 WT cells, or THP-1ACE2 cells pretreated with Fc blocker, along with normal serum or pooled convalescent serum (0.2 % or 1 %) from monkey, followed by SARS-CoV-2 infection at an MOI = 1 for 48 h. Data in (a-f) are plotted as the mean ± SD (n = 3 independent samples).

Source data

Extended Data Fig. 8 ACE2-independent SARS-CoV-2 entry in myeloid cells may rely on its interaction with TLR1.

(a) Quantitative PCR with reverse transcription (RT-PCR) analysis of SARS-CoV-2 S and RdRp (NSP12) proteins in MDM cells with live SARS-CoV-2 or heat-inactivated SARS-CoV-2 infection at an MOI = 1 for indicated time points. (b) RT-PCR analysis of SARS-CoV-2 S and RdRp (NSP12) proteins in THP-1 cells with live SARS-CoV-2 or heat-inactivated SARS-CoV-2 infection at an MOI = 1 for indicated time points. (c) RT-PCR analysis of expression of inflammatory genes in THP-1 cells with live SARS-CoV-2 or heat-inactivated SARS-CoV-2 infection at an MOI = 1 for indicated time points. (d) RT-PCR analysis of SARS-CoV-2 S, RdRp (NSP12), and NSP1 proteins in mouse BMDM (Bone marrow-derived macrophage) with SARS-CoV-2 infection at an MOI = 1 for indicated time points. (e) HEK293T ACE2-/- cells were transfected with plasmids containing either a control empty vector (EV) or TLR1, followed by SARS-CoV-2 infection at MOI = 1 for indicated lengths of time and temperature. RT-PCR analysis was then performed to analyze the RNA of SARS-CoV-2 S and RdRp. (f) Lysates of HEK293T cells transfected with the plasmid for TLR1-Flag, together with the empty vector or expression vector of Strep-tagged SARS-CoV-2 structural proteins (E, M, N), were subjected to immunoprecipitation with anti-Strep beads and immunoblot analysis with anti-Flag. (g) Cell proliferation of THP-1 WT or TLR1-/- cells were continuously assessed by counting the cells using flow cytometry for five days. (h) Phase contrast micrographs show the morphology of THP-1 WT or TLR1-/- cells (Scale bars 100 μm). Data in (f) is the representative of multiple independent experiments, data in (a-e, g) are plotted as the mean ± SD (n = 3 independent samples). Statistical analyses were performed using two-way ANOVA followed by Sidak post-test (a-c).

Source data

Extended Data Fig. 9 The entry of SARS-CoV-2 into myeloid cells relies on TLR1 rather than NF-κB signaling.

(a) Quantitative PCR with reverse transcription (RT-PCR) analysis of SARS-CoV-2 S and RdRp (NSP12) in MDMs transfected with Ctrl-siRNA or TLR1-siRNA, followed with SARS-CoV-2 infection at an MOI = 1 for the indicated lengths of time. (b) Ctrl-siRNA or TLR1-siRNA transfected MDMs were infected with SARS-CoV-2 at an MOI = 1 for the indicated lengths of time, and the lysates were harvested for immunoblot analysis with indicated antibodies. (c) Expression of the proinflammatory genes in MDMs transfected with Ctrl-siRNA or TLR1-siRNA was measured after being treated with SARS-CoV-2 infection at an MOI = 1 for the indicated lengths of time. (d) Expression of the proinflammatory genes in MDMs with ethanol or C25-140 (30 μM) treatment, followed by SARS-CoV-2 infection at an MOI = 1 for indicated time points. (e) RT-PCR analysis of SARS-CoV-2 S and RdRp (NSP12) proteins in MDMs pretreated with ethanol or C25-140 (30 μM), followed by SARS-CoV-2 infection at an MOI = 1 for indicated time points. (f) RT-PCR analysis of SARS-CoV-2 S and RdRp (NSP12) proteins in WT or Myd88-/- BMDM infected with SARS-CoV-2 infection at multiple MOIs for 24 h. Data in (a, c-f) are plotted as mean ± SD (n = 3 independent samples). Statistical analyses were performed using two-way ANOVA followed by Sidak post-test (a, c-f). ns, not significant.

Source data

Extended Data Fig. 10 The Omicron variant and SARS-CoV-1 may trigger a hyperinflammatory response in myeloid cells via a mechanism similar to that of SARS-CoV-2.

(a) The recombinant TLR1 protein was directly incubated with Flag-Envelope (E) protein or Flag-Membrane (M) protein pre-bound on the anti-Flag beads overnight at 4 °C, followed by immunoprecipitation. (b) Expression of the proinflammatory genes in alveolar macrophage isolated from mouse bronchoalveolar lavage fluid (BALF) was measured after being stimulated with purified recombinant His-E protein for 4 hours. (c) Expression of the proinflammatory genes in WT or TLR1-/- THP-1 cells was measured after being stimulated with four different sources of purified recombinant E protein of SARS-CoV-2 (1 μg/ml or 2 μg/ml) or E protein of SARS-CoV-1 (equivalent molar quantity as SARS-CoV-2) for 4 hours. (d) Quantitative PCR with reverse transcription (RT-PCR) analysis of proinflammatory relative cytokines genes (IL6, IL1B, TNFA, IL8) of THP-1 cells with HCoV-OC43, HCoV-NL63, and HCoV-229E infection at an MOI = 1 for the indicated lengths of time. (e) The recombinant TLR1 protein was directly incubated with recombinant Flag-M proteins of indicated human coronavirus pre-bound on the anti-Flag beads overnight at 4 °C, followed by immunoprecipitation. (f, g) Expression of the proinflammatory genes in THP-1 (f) or MDM (g) after SARS-CoV-2 omicron variant (BA. 2) infection at multiple MOIs for 24 h. (h, i) RT-PCR analysis of total RNA of SARS-CoV-2 ORF6, S protein, and N protein or subgenomic RNA of ORF6, S, and N proteins in THP-1 cells (h) or MDMs (i) after SARS-CoV-2 omicron variant (BA. 2) infection at multiple MOIs for 24 h. Data in (a, e) are the representative of multiple independent experiments, data in (b-d, f-i) are plotted as mean ± SD (n = 3 independent samples). Statistical analyses were performed using two-way ANOVA followed by Sidak post-test (c).

Source data

Supplementary information

Reporting Summary

Peer Review File

Supplementary Tables

Supplementary Table 1. Primer sequences. Supplementary Table 2. siRNA sequences.

Source data

Source_data for all Figures

Unprocessed western Blots and Statistical Source Data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Duan, T., Xing, C., Chu, J. et al. ACE2-dependent and -independent SARS-CoV-2 entries dictate viral replication and inflammatory response during infection. Nat Cell Biol 26, 628–644 (2024). https://doi.org/10.1038/s41556-024-01388-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-024-01388-w

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing