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:

Mitochondria-localised ZNFX1 functions as a dsRNA sensor to initiate antiviral responses through MAVS

Abstract

In the past two decades, emerging studies have suggested that DExD/H box helicases belonging to helicase superfamily 2 (SF2) play essential roles in antiviral innate immunity. However, the antiviral functions of helicase SF1, which shares a conserved helicase core with SF2, are little understood. Here we demonstrate that zinc finger NFX1-type containing 1 (ZNFX1), a helicase SF1, is an interferon (IFN)-stimulated, mitochondrial-localised dsRNA sensor that specifically restricts the replication of RNA viruses. Upon virus infection, ZNFX1 immediately recognizes viral RNA through its Armadillo-type fold and P-loop domain and then interacts with mitochondrial antiviral signalling protein to initiate the type I IFN response without depending on retinoic acid-inducible gene I-like receptors (RLRs). In short, as is the case with interferon-stimulated genes (ISGs) alone, ZNFX1 can induce IFN and ISG expression at an early stage of RNA virus infection to form a positively regulated loop of the well-known RLR signalling. This provides another layer of understanding of the complexity of antiviral immunity.

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: Viral infection induces Znfx1 expression, an IFN-stimulated gene.
Fig. 2: ZNFX1 deficiency impairs cellular antiviral RNA response.
Fig. 3: ZNFX1 is essential for host defence against RNA virus in mice.
Fig. 4: ZNFX1 initiates IFN-based antiviral responses.
Fig. 5: ZNFX1 binds to viral RNA.
Fig. 6: ZNFX1 localises to mitochondria and interacts with MAVS.
Fig. 7: ZNFX1 induces IFN signalling in an RLR-independent manner.

Similar content being viewed by others

Data availability

RNA sequencing data that support the findings of this study have been deposited in the Gene Expression Omnibus (GEO) under accession code GSE132979. Source data for Figs. 17 and Extended Data Figs. 17 have been provided as Statistics Source Data. All other data supporting the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Ablasser, A. et al. cGAS produces a 2′–5′-linked cyclic dinucleotide second messenger that activates STING. Nature 498, 380–384 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Wu, J. et al. Cyclic GMP-AMP is an endogenous second messenger in innate immune signaling by cytosolic DNA. Science 339, 826–830 (2013).

    Article  CAS  PubMed  Google Scholar 

  3. Yoneyama, M. et al. The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses. Nat. Immunol. 5, 730–737 (2004).

    Article  CAS  PubMed  Google Scholar 

  4. Pichlmair, A. et al. RIG-I-mediated antiviral responses to single-stranded RNA bearing 5′-phosphates. Science 314, 997–1001 (2006).

    Article  CAS  PubMed  Google Scholar 

  5. Kato, H. et al. Differential roles of MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature 441, 101–105 (2006).

    Article  CAS  PubMed  Google Scholar 

  6. Goubau, D. et al. Antiviral immunity via RIG-I-mediated recognition of RNA bearing 5′-diphosphates. Nature 514, 372–375 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Ahmad, S. & Hur, S. Helicases in antiviral immunity: dual properties as sensors and effectors. Trends Biochem. Sci. 40, 576–585 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Oshiumi, H. et al. DDX60 is involved in RIG-I-dependent and independent antiviral responses, and its function is attenuated by virus-induced EGFR activation. Cell Rep. 11, 1193–1207 (2015).

    Article  CAS  PubMed  Google Scholar 

  9. Zhang, Z. et al. DDX1, DDX21 and DHX36 helicases form a complex with the adaptor molecule TRIF to sense dsRNA in dendritic cells. Immunity 34, 866–878 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Zhang, Z., Yuan, B., Lu, N., Facchinetti, V. & Liu, Y. J. DHX9 pairs with IPS-1 to sense double-stranded RNA in myeloid dendritic cells. J. Immunol. 187, 4501–4508 (2011).

    Article  CAS  PubMed  Google Scholar 

  11. Wang, P. H. et al. Nlrp6 regulates intestinal antiviral innate immunity. Science 350, 826–830 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Singleton, M. R., Dillingham, M. S. & Wigley, D. B. Structure and mechanism of helicases and nucleic acid translocases. Annu. Rev. Biochem. 76, 23–50 (2007).

    Article  CAS  PubMed  Google Scholar 

  13. Fairman-Williams, M. E., Guenther, U. P. & Jankowsky, E. SF1 and SF2 helicases: family matters. Curr. Opin. Struct. Biol. 20, 313–324 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Linder, P. & Jankowsky, E. From unwinding to clamping—the DEAD box RNA helicase family. Nat. Rev. Mol. Cell. Bio. 12, 505–516 (2011).

    Article  CAS  Google Scholar 

  15. Seth, R. B., Sun, L., Ea, C. K. & Chen, Z. J. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-κB and IRF 3. Cell 122, 669–682 (2005).

    Article  CAS  PubMed  Google Scholar 

  16. Ishikawa, H. & Barber, G. N. STING is an endoplasmic reticulum adaptor that facilitates innate immune signalling. Nature 455, 674–678 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Chen, Q., Sun, L. J. & Chen, Z. J. Regulation and function of the cGAS-STING pathway of cytosolic DNA sensing. Nat. Immunol. 17, 1142–1149 (2016).

    Article  CAS  PubMed  Google Scholar 

  18. Munoz-Wolf, N. & Lavelle, E. C. Hippo interferes with antiviral defences. Nat. Cell Biol. 19, 267–269 (2017).

    Article  CAS  PubMed  Google Scholar 

  19. Zhang, Q. et al. Hippo signalling governs cytosolic nucleic acid sensing through YAP/TAZ-mediated TBK1 blockade. Nat. Cell Biol. 19, 362–374 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. O’Shea, J. J. & Plenge, R. JAK and STAT signaling molecules in immunoregulation and immune-mediated disease. Immunity 36, 542–550 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Schoggins, J. W. et al. Pan-viral specificity of IFN-induced genes reveals new roles for cGAS in innate immunity. Nature 505, 691–695 (2014).

    Article  CAS  PubMed  Google Scholar 

  22. Chan, Y. K. & Gack, M. U. Viral evasion of intracellular DNA and RNA sensing. Nat. Rev. Microbiol. 14, 360–373 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Schneider, W. M., Chevillotte, M. D. & Rice, C. M. Interferon-stimulated genes: a complex web of host defenses. Annu. Rev. Immunol. 32, 513–545 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Schoggins, J. W. et al. A diverse range of gene products are effectors of the type I interferon antiviral response. Nature 472, 481–485 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Pillai, P. S. et al. Mx1 reveals innate pathways to antiviral resistance and lethal influenza disease. Science 352, 463–466 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bailey, C. C. et al. IFITM-family proteins: the cell’s first line of antiviral defense. Annu. Rev. Virol. 1, 261–283 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Munir, M. & Berg, M. The multiple faces of proteinkinase R in antiviral defense. Virulence 4, 85–89 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Neil, S. J., Zeng, T. & Bieniasz, P. D. Tetherin inhibits retrovirus release and is antagonized by HIV-1 Vpu. Nature 451, 425–430 (2008).

    Article  CAS  PubMed  Google Scholar 

  29. Yan, N. & Chen, Z. J. Intrinsic antiviral immunity. Nat. Immunol. 13, 214–222 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Yoh, S. M. et al. PQBP1 is a proximal sensor of the cGAS-dependent innate response to HIV-1. Cell 161, 1293–1305 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Song, G. et al. E3 ubiquitin ligase RNF128 promotes innate antiviral immunity through K63-linked ubiquitination of TBK1. Nat. Immunol. 17, 1342–1351 (2016).

    Article  CAS  PubMed  Google Scholar 

  32. Zhu, J. et al. Antiviral activity of human OASL protein is mediated by enhancing signaling of the RIG-I RNA sensor. Immunity 40, 936–948 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Jia, X. et al. The role of alternative polyadenylation in the antiviral innate immune response. Nat. Commun. 8, 14605 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Goodier, J. L., Cheung, L. E. & Kazazian, H. H. Jr. MOV10 RNA helicase is a potent inhibitor of retrotransposition in cells. PLoS Genet. 8, e1002941 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Serquina, A. K. et al. UPF1 is crucial for the infectivity of human immunodeficiency virus type 1 progeny virions. J. Virol. 87, 8853–8861 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Miller, M. S. et al. Senataxin suppresses the antiviral transcriptional response and controls viral biogenesis. Nat. Immunol. 16, 485–494 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. López, D. G. et al. The Eutherian Armcx genes regulate mitochondrial trafficking in neurons and interact with Miro and Trak2. Nat. Commun. 3, 814 (2012).

    Article  CAS  Google Scholar 

  38. Mukherjee, P. et al. Activation of the innate signaling molecule MAVS by bunyavirus infection upregulates the adaptor protein SARM1, leading to neuronal death. Immunity 38, 705–716 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. West, A. P., Shadel, G. S. & Ghosh, S. Mitochondria in innate immune responses. Nat. Rev. Immunol. 11, 389–402 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Yoshizumi, T. et al. Influenza A virus protein PB1-F2 translocates into mitochondria via Tom40 channels and impairs innate immunity. Nat. Commun. 5, 4713 (2014).

    Article  CAS  PubMed  Google Scholar 

  41. Loo, Y. M. et al. Distinct RIG-I and MDA5 signaling by RNA viruses in innate immunity. J. Virol. 82, 335–345 (2008).

    Article  CAS  PubMed  Google Scholar 

  42. Lu, H. B. et al. DHX15 senses double-stranded RNA in myeloid dendritic cells. J. Immunol. 193, 1364–1372 (2014).

    Article  CAS  PubMed  Google Scholar 

  43. Lian, H. et al. The zinc-finger protein ZCCHC3 binds RNA and facilitates viral RNA sensing and activation of the RIG-I-like receptors. Immunity 18, 49 (2018).

    Google Scholar 

  44. Serrat, R. et al. The non-canonical Wnt/PKC pathway regulates mitochondrial dynamics through degradation of the arm-like domain-containing protein Alex3. PLoS One 8, e67773 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Panneerselvam, P. et al. T-cell death following immune activation is mediated by mitochondria-localized SARM. Cell Death Differ. 20, 478–489 (2013).

    Article  CAS  PubMed  Google Scholar 

  46. Yuan, S. et al. Amphioxus SARM involved in neural development may function as a suppressor of TLR signaling. J. Immunol. 184, 6874–6881 (2010).

    Article  CAS  PubMed  Google Scholar 

  47. Hou, F. et al. MAVS forms functional prion-like aggregates to activate and propagate antiviral innate immune response. Cell 146, 448–461 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Gack, M. U. et al. TRIM25 RING-finger E3 ubiquitin ligase is essential for RIG-I-mediated antiviral activity. Nature 446, 916–920 (2007).

    Article  CAS  PubMed  Google Scholar 

  49. Jiang, X. et al. Ubiquitin-induced oligomerization of the RNA sensors RIG-I and MDA5 activates antiviral innate immune response. Immunity 36, 959–973 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Ishidate, T. et al. ZNFX-1 functions within perinuclear nuage to balance epigenetic signals. Mol. Cell 70, 639–649 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Wan, G. et al. Spatiotemporal regulation of liquid-like condensates in epigenetic inheritance. Nature 557, 679–683 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation (NNSF) of China under grants 31770943, 81430099 and 31900661 and by the Natural Science Foundation of Guangdong Province of China under grants 2015A030306043, 2018A030313924 and 2018A030313051.

Author information

Authors and Affiliations

Authors

Contributions

Y.W., S.Y. and A.X. conceived the ideas and designed the experiments. Y.W., X.J., Y.G., T.L., M.N. and X.L. performed the experiments. Y.W., S.Y. and X.J. analysed the data. S.Y., Y.W., X.J. and S.C. contributed to editing the manuscript. S.Y. and A.X. supervised the research and wrote the paper. S.Y. and X.J. are joint co first authors.

Corresponding author

Correspondence to Anlong Xu.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 Bioinformatics analysis of in vitro transcription-sequencing APA sites (IVT-SAPAS) data of VSV-infected macrophages and previously collected viral infection microarray data.

a, IVT-SAPAS revealed the genes with transcriptional changes in MDMs infected with VSV for 0, 2, 4, 8, 16, 24 hrs. b-c, The percentage of GFP+ cells of FACS analysis of A549 cells transfected with control siRNA or the indicated siRNAs followed by VSV-eGFP infection for another 6 hrs (n = 3 independent experiments). d, qRT-PCR revealed the mRNA expression of target gene in A549 cells transfected with indicated siRNAs for 48 hrs (n = 3 independent experiments). e-f, RNA levels of Znfx1 and Rig-I are significantly increased after different virial infection in different cell types. g, Schematic representation of ZNFX1 promoter containing core region bound by STAT1, STAT2, IRF1 and IRF9. For e, n = 3 independent experiments. Data in f, n = 4 wells for SeV infected epithelial cells and n = 6 wells for uninfected cells; n = 5 samples for IVA infected pDCs; n = 4 independent experiments for SeV infected monocytoid cells; n = 2 independent experiments for SeV or HIV infected Macrophages or mDCs. All data are shown as the mean ± s.d. Statistical differences were detected using two-tailed unpaired Student’s t-tests.

Source data

Extended Data Fig. 2 ZNFX1 deficiency impairs antiviral immune response in vitro and in vivo.

a, Quantitative RT-PCR (qRT-PCR) analysis of Znfx1 mRNA expression in A549 and L929 cells transfected with control siRNA or ZNFX1 siRNA 1#, 2# or 3# for 48 hrs (n = 3 independent experiments). b, Western blot analysis of ZNFX1 protein expression in A549 cells transfected with control siRNA or human ZNFX1 siRNA 1# for 48 hrs. c, ELISA of IFN-α or IFN-β production in cell supernatants from A549 cells with target gene knockdown for 48 hrs followed by VSV infection or poly I:C stimulation for another 12 hrs (n = 4 independent experiments). d, qRT-PCR analysis of VSV mRNA expression (left panel) and plaque assay analysis of VSV titer (right panel) of A549 cells transfected with RIG-I, ZNFX1 expressing plasmids or empty vector plasmid for 24 hrs and then infected with VSV at an MOI of 2 for 16 hrs (n = 3 independent experiments). e, FACS analysis of Znfx1+/+ and Znfx1-/- 293T cells followed by VSV-eGFP infection at 0.5 MOI for the indicated time points (n = 3 independent experiments). f, Znfx1-/- 293T and A549 clones were generated by the CRISPR-Cas9 method. Deficiency of target genes in the KO clones were confirmed by immunoblotting analysis. g, qRT-PCR analysis of viral mRNA transcripts in VSV, EMCV, H1N1 and HSV-1 infected A549 cells with control siRNA (si Control) or Znfx1-specific siRNA (si ZNFX1) (n = 3 independent experiments). h, ELISA of IFN-α in supernatants of BMDMs from WT and Znfx1-/- mice infected with VSV or HSV-1 for 16 hrs (n = 5 independent experiments). All data are shown as the mean ± s.d. P values were calculated using two-tailed unpaired Student’s t-test. For b, f, the experiments were repeated three times, independently, with similar results obtained.

Source data

Source data

Extended Data Fig. 3 ZNFX1 positively regulates IFN-β signaling.

a, Illustration of the CRISPR-Cas9 strategy to generate Znfx1-deficient mice and primer design used in (b). b, Genotyping of the ZNFX1 mutant pups. c, Immunoblot analysis of ZNFX1 protein levels in Znfx1+/+ and Znfx1-/- MEFs cells. d, qRT-PCR analysis of Ifnb1 and ISGs mRNA levels in A549 cells transfected with siControl (si Ctrl) or si ZNFX1 for 48 hrs and then infected with VSV for the indicated time points (n = 3 independent experiments). e, qRT-PCR analysis of Ifnb1 and ISGs mRNA expression in Znfx1+/+ and Znfx1-/- 293T cells followed by VSV infected with increasing MOI (0.5 and 1) for 0, 8 and 16 hrs (n = 3 independent experiments). For b, c, the experiments were repeated three times, independently, with similar results obtained. Data in d, e are the mean ± s.d. P values were calculated using two-tailed unpaired Student’s t-test.

Source data

Source data

Extended Data Fig. 4 ZNFX1 localizes to mitochondria and interacts with MAVS.

a, FACS analysis of Znfx1-/- A549 cells transfected with ZNFX1 and its mutants expressing plasmids or empty vector (EV) for 24 hrs followed by VSV-eGFP infection at an MOI of 2 for 6 hrs. b, Endogenous level of ZNFX1 protein in mitochondrial fractions in WT and Mavs-/- 293T with or without VSV infection at 1 MOI for 6 hrs. COX-IV was used as the loading control. Data are representative of at least three independent experiments.

Source data

Extended Data Fig. 5 The expression of Znfx1 in different tissues and cell types, and the phylogenetic tree of Znfx1.

a-c, The expression of Znfx1, Rig-I and Mda5 in different tissues and cell types as per BioGPS. d, Phylogenetic tree of ZNFX1 and RIG-I using an amino acid sequence alignment among different species.

Extended Data Fig. 6 Alignment of ZNFX1 amino acid sequences in human, mouse, rat and zebrafish.

Shading indicates sequence conservation, with darker gray indicating a higher degree of conservation.

Extended Data Fig. 7 Work model of mitochondria-localized ZNFX1 functions as a dsRNA sensor to initiate antiviral responses through MAVS.

Upon RNA virus infection, ZNFX1 induces type I interferon response by interacting with MAVS in the early stage, thus primes the expression of a number of ISGs, including RIG-I and MDA5. The induced sensors further enhance the antiviral immune response by amplifying ISGs expression.

Supplementary information

Reporting Summary

Supplementary Tables 1

Conservative analysis of ZNFX1 and DDX58 in different species.

Supplementary Tables 2

Information about the primers used in the study.

Source data

Source Data Fig. 1

Statistical Source Data

Source Data Fig. 1

Unprocessed Western blots and/or gels

Source Data Fig. 2

Statistical Source Data

Source Data Fig. 2

Unprocessed Western blots and/or gels

Source Data Fig. 3

Statistical Source Data

Source Data Fig. 4

Statistical Source Data

Source Data Fig. 4

Unprocessed Western blots and/or gels

Source Data Fig. 5

Statistical Source Data

Source Data Fig. 5

Unprocessed Western blots and/or gels

Source Data Fig. 6

Statistical Source Data

Source Data Fig. 6

Unprocessed Western blots and/or gels

Source Data Fig. 7

Statistical Source Data

Source Data Extended Data Fig. 1

Statistical Source Data

Source Data Extended Data Fig. 2

Statistical Source Data

Source Data Extended Data Fig. 2

Unprocessed Western blots and/or gels

Source Extended Data Fig. 3

Statistical Source Data

Source Data Extended Data Fig. 3

Unprocessed Western blots and/or gels

Source Data Extended Data Fig. 4

Unprocessed Western blots and/or gels

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, Y., Yuan, S., Jia, X. et al. Mitochondria-localised ZNFX1 functions as a dsRNA sensor to initiate antiviral responses through MAVS. Nat Cell Biol 21, 1346–1356 (2019). https://doi.org/10.1038/s41556-019-0416-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-019-0416-0

This article is cited by

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