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:

Cut site preference allows influenza A virus PA-X to discriminate between host and viral mRNAs

Abstract

Many viruses block host gene expression to take over the infected cell. This process, termed host shutoff, is thought to promote viral replication by preventing antiviral responses and redirecting cellular resources to viral processes. Several viruses from divergent families accomplish host shutoff through RNA degradation by endoribonucleases. However, viruses also need to ensure expression of their own genes. The influenza A virus endoribonuclease PA-X solves this problem by sparing viral mRNAs and some host RNAs necessary for viral replication. To understand how PA-X distinguishes between RNAs, we characterized PA-X cut sites transcriptome-wide using 5′ rapid amplification of complementary DNA ends coupled to high-throughput sequencing. This analysis, along with RNA structure predictions and validation experiments using reporters, shows that PA-Xs from multiple influenza strains preferentially cleave RNAs at GCUG tetramers in hairpin loops. Importantly, GCUG tetramers are enriched in the human but not the influenza transcriptome. Moreover, optimal PA-X cut sites inserted in the influenza A virus genome are quickly selected against during viral replication in cells. This finding suggests that PA-X evolved these cleavage characteristics to preferentially target host over viral mRNAs in a manner reminiscent of cellular self versus non-self discrimination.

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: 5′ RACE-seq with PyDegradome analysis identifies PA-X cut sites transcriptome-wide.
Fig. 2: Sequences around the cut sites drive cleavage by PA-X.
Fig. 3: PA-X preferentially cleaves RNAs at GCUG tetramers.
Fig. 4: PA-X preferentially cleaves RNA within hairpin loop structures.
Fig. 5: PA-X likely cleaves GCUG sequences to preferentially target host over viral mRNAs.

Similar content being viewed by others

Data availability

All primary data (including uncropped gel and blot images) except for 5′ RACE-seq data are publicly available as a Figshare collection (https://doi.org/10.6084/m9.figshare.c.6616663)68. The raw 5′ RACE-seq sequences and identified cut sites are deposited in the NCBI Gene Expression Omnibus database (GSE207253).

Code availability

All Python scripts used, including the PyDegradome 2.0 code, can be found in our laboratory’s GitHub page, https://github.com/mgaglia81/PyDegradome (ref. 69). The workflow explaining the use of the PyDegradome 2.0 code can be found in the Methods section (‘PyDegradome and other Python analyses’). The method of Reed and Muench44 was used to calculate virus titres using the Bloom Laboratory Python script at https://github.com/jbloomlab/reedmuenchcalculator (ref. 45). The RNA structure prediction software LinearFold52 is now available at https://linearfold.eecs.oregonstate.edu/. WebLogo56 is made available by the Computational Genomics Research Group (University of California, Berkeley) at https://weblogo.berkeley.edu/logo.cgi.

References

  1. Gaucherand, L. & Gaglia, M. M. The role of viral RNA degrading factors in shutoff of host gene expression. Annu. Rev. Virol. 9, 213–238 (2022).

    Article  PubMed  Google Scholar 

  2. Bercovich-Kinori, A. et al. A systematic view on influenza induced host shutoff. eLife 5, e18311 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Jagger, B. W. et al. An overlapping protein-coding region in influenza A virus segment 3 modulates the host response. Science 337, 199–204 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Gaucherand, L. et al. The influenza A virus endoribonuclease PA-X usurps host mRNA processing machinery to limit host gene expression. Cell Rep. 27, 776–792.e7 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Gao, H. et al. The contribution of PA-X to the virulence of pandemic 2009 H1N1 and highly pathogenic H5N1 avian influenza viruses. Sci. Rep. 5, 8262 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Gong, X. Q. et al. PA-X protein decreases replication and pathogenicity of swine influenza virus in cultured cells and mouse models. Vet. Microbiol. 205, 66–70 (2017).

    Article  CAS  PubMed  Google Scholar 

  7. Hayashi, T., MacDonald, L. A. & Takimoto, T. Influenza A virus protein PA-X contributes to viral growth and suppression of the host antiviral and immune responses. J. Virol. 89, 6442–6452 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Hu, J. et al. PA-X decreases the pathogenicity of highly pathogenic H5N1 influenza A virus in avian species by inhibiting virus replication and host response. J. Virol. 89, 6442–6452 (2015).

    Article  Google Scholar 

  9. Firth, A. E. et al. Ribosomal frameshifting used in influenza A virus expression occurs within the sequence UCC-UUU-CGU and is in the +1 direction. Open Biol. 2, 120109 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Dias, A. et al. The cap-snatching endonuclease of influenza virus polymerase resides in the PA subunit. Nature 458, 914–918 (2009).

    Article  CAS  PubMed  Google Scholar 

  11. Yuan, P. et al. Crystal structure of an avian influenza polymerase PAN reveals an endonuclease active site. Nature 458, 909–913 (2009).

    Article  CAS  PubMed  Google Scholar 

  12. Khaperskyy, D. A., Schmaling, S., Larkins-Ford, J., McCormick, C. & Gaglia, M. M. Selective degradation of host RNA polymerase II transcripts by influenza A virus PA-X host shutoff protein. PLoS Pathog. 12, e1005427 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Chaimayo, C., Dunagan, M., Hayashi, T., Santoso, N. & Takimoto, T. Specificity and functional interplay between influenza virus PA-X and NS1 shutoff activity. PLoS Pathog. 14, e1007465 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Chao, Y. et al. In vivo cleavage map illuminates the central role of RNase E in coding and non-coding RNA pathways. Mol. Cell 65, 39–51 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Gaglia, M. M., Rycroft, C. H. & Glaunsinger, B. A. Transcriptome-wide cleavage site mapping on cellular mRNAs reveals features underlying sequence-specific cleavage by the viral Ribonuclease SOX. PLoS Pathog. 11, e1005305 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Abernathy, E., Gilbertson, S., Alla, R. & Glaunsinger, B. Viral nucleases induce an mRNA degradation-transcription feedback loop in mammalian cells. Cell Host Microbe 18, 243–253 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Liu, R. & Moss, B. Opposing roles of double-stranded RNA effector pathways and viral defense proteins revealed with CRISPR–Cas9 knockout cell lines and vaccinia virus mutants. J. Virol. 90, 7864–7879 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Younis, I. et al. Rapid-response splicing reporter screens identify differential regulators of constitutive and alternative splicing. Mol. Cell. Biol. 30, 1718–1728 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Levene, R. E., Shrestha, S. D. & Gaglia, M. M. The influenza A virus host shutoff factor PA-X is rapidly turned over in a strain-specific manner. J. Virol. 95, e02312–e02320 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bavagnoli, L. et al. The novel influenza A virus protein PA-X and its naturally deleted variant show different enzymatic properties in comparison to the viral endonuclease PA. Nucleic Acids Res. 43, 9405–9417 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Plotch, S. J., Bouloy, M., Ulmanen, I. & Krug, R. M. A unique cap(m7GpppXm)-dependent influenza virion endonuclease cleaves capped RNAs to generate the primers that initiate viral RNA transcription. Cell 23, 847–858 (1981).

    Article  CAS  PubMed  Google Scholar 

  22. Datta, K., Wolkerstorfer, A., Szolar, O. H. J., Cusack, S. & Klumpp, K. Characterization of PA-N terminal domain of Influenza A polymerase reveals sequence specific RNA cleavage. Nucleic Acids Res. 41, 8289–8299 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Sikora, D., Rocheleau, L., Brown, E. G. & Pelchat, M. Deep sequencing reveals the eight facets of the influenza A/HongKong/1/1968 (H3N2) virus cap-snatching process. Sci. Rep. 4, 6181 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Gu, W. et al. Influenza A virus preferentially snatches noncoding RNA caps. RNA 21, 2067–2075 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Brannan, K. et al. mRNA decapping factors and the exonuclease Xrn2 function in widespread premature termination of RNA polymerase II transcription. Mol. Cell 46, 311–324 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Wang, X.-H. et al. The role of PA-X C-terminal 20 residues of classical swine influenza virus in its replication and pathogenicity. Vet. Microbiol. 251, 108916 (2020).

    Article  CAS  PubMed  Google Scholar 

  27. Oishi, K., Yamayoshi, S. & Kawaoka, Y. Identification of amino acid residues in influenza A virus PA-X that contribute to enhanced shutoff activity. Front. Microbiol. 10, 432 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Feng, K. H., Sun, M., Iketani, S., Holmes, E. C. & Parrish, C. R. Comparing the functions of equine and canine influenza H3N8 virus PA-X proteins: suppression of reporter gene expression and modulation of global host gene expression. Virology 496, 138–146 (2016).

    Article  CAS  PubMed  Google Scholar 

  29. Nogales, A. et al. Natural selection of H5N1 avian influenza A viruses with increased PA-X and NS1 shutoff activity. Viruses 13, 1760 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Wilamowski, M., Gorecki, A., Dziedzicka-Wasylewska, M. & Jura, J. Substrate specificity of human MCPIP1 endoribonuclease. Sci. Rep. 8, 7381 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Takata, M. A. et al. CG dinucleotide suppression enables antiviral defence targeting non-self RNA. Nature 550, 124–127 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Pezda, A. C., Penn, A., Barton, G. M. & Coscoy, L. Suppression of TLR9 immunostimulatory motifs in the genome of a gammaherpesvirus. J. Immunol. 187, 887–896 (2011).

    Article  CAS  PubMed  Google Scholar 

  33. Hartenian, E. & Glaunsinger, B. A. Feedback to the central dogma: cytoplasmic mRNA decay and transcription are interdependent processes. Crit. Rev. Biochem. Mol. Biol. 54, 385–398 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Duncan-Lewis, C., Hartenian, E., King, V. & Glaunsinger, B. A. Cytoplasmic mRNA decay represses RNA polymerase II transcription during early apoptosis. eLife 10, e58342 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Zhao, N. et al. Influenza virus infection causes global RNAPII termination defects. Nat. Struct. Mol. Biol. 25, 885–893 (2018).

    Article  CAS  PubMed  Google Scholar 

  36. Bauer, D. L. V. et al. Influenza virus mounts a two-pronged attack on host RNA polymerase II transcription. Cell Rep. 23, 2119–2129.e3 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Khaperskyy, D. A. et al. Influenza A virus host shutoff disables antiviral stress-induced translation arrest. PLoS Pathog. 10, e1004217 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Glaunsinger, B. & Ganem, D. Lytic KSHV infection inhibits host gene expression by accelerating global mRNA turnover. Mol. Cell 13, 713–723 (2004).

    Article  CAS  PubMed  Google Scholar 

  39. Jones, F. E., Smibert, C. A. & Smiley, J. R. Mutational analysis of the herpes simplex virus virion host shutoff protein: evidence that vhs functions in the absence of other viral proteins. J. Virol. 69, 4863–4871 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Le Sage, V. et al. Cell-culture adaptation of H3N2 influenza virus impacts acid stability and reduces airborne transmission in ferret model. Viruses 13, 719 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Hoffmann, E., Neumann, G., Kawaoka, Y., Hobom, G. & Webster, R. G. A DNA transfection system for generation of influenza A virus from eight plasmids. Proc. Natl Acad. Sci. USA 97, 6108–6113 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Khaperskyy, D. A., Hatchette, T. F. & McCormick, C. Influenza A virus inhibits cytoplasmic stress granule formation. FASEB J. 26, 1629–1639 (2012).

    Article  CAS  PubMed  Google Scholar 

  43. Matrosovich, M., Matrosovich, T., Garten, W. & Klenk, H. D. New low-viscosity overlay medium for viral plaque assays. Virol. J. 3, 63 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Reed, L. J. & Muench, H. A simple method of estimating fifty per cent endPOINTS12. Am. J. Epidemiol. 27, 493–497 (1938).

    Article  Google Scholar 

  45. Gaglia, M. & Gaucherand, L. Purification of influenza A virus RNA from cell supernatant to check sequences. Protocols.io https://doi.org/10.17504/protocols.io.36wgqj8wxvk5/v1 (2023).

  46. Gaglia, M. 5′ RACE for RNA fragments with 5′ phosphate. Protocols.io https://doi.org/10.17504/protocols.io.261ge3mzwl47/v1 (2023).

  47. German, M. A., Luo, S., Schroth, G., Meyers, B. C. & Green, P. J. Construction of parallel analysis of RNA ends (PARE) libraries for the study of cleaved miRNA targets and the RNA degradome. Nat. Protoc. 4, 356–362 (2009).

    Article  CAS  PubMed  Google Scholar 

  48. Zhai, J., Arikit, S., Simon, S. A., Kingham, B. F. & Meyers, B. C. Rapid construction of parallel analysis of RNA end (PARE) libraries for Illumina sequencing. Methods 67, 84–90 (2014).

    Article  CAS  PubMed  Google Scholar 

  49. Gaglia, M. & Gaucherand, L. 5′ RACE-seq for RNA fragments with 5′ phosphate—library prep protocol. Protocols.io https://doi.org/10.17504/protocols.io.n2bvj847wgk5/v1 (2023).

  50. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet. J. 17, 10–12 (2011).

    Article  Google Scholar 

  51. Kim, D., Paggi, J. M., Park, C., Bennett, C. & Salzberg, S. L. Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype. Nat. Biotechnol. 37, 907–915 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Huang, L. et al. LinearFold: linear-time approximate RNA folding by 5′-to-3′ dynamic programming and beam search. Bioinformatics 35, i295–i304 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Do, C. B., Woods, D. A. & Batzoglou, S. CONTRAfold: RNA secondary structure prediction without physics-based models. Bioinformatics 22, e90–e98 (2006).

    Article  CAS  PubMed  Google Scholar 

  54. Lorenz, R. et al. ViennaRNA Package 2.0. Algorithms Mol. Biol. 6, 26 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  55. Mathews, D. H. et al. Incorporating chemical modification constraints into a dynamic programming algorithm for prediction of RNA secondary structure. Proc. Natl Acad. Sci. USA 101, 7287–7292 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Crooks, G. E., Hon, G., Chandonia, J. M. & Brenner, S. E. WebLogo: a sequence logo generator. Genome Res. 14, 1188–1190 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Gao, H. et al. PA-X is a virulence factor in avian H9N2 influenza virus. J. Gen. Virol. 96, 2587–2594 (2015).

    Article  CAS  PubMed  Google Scholar 

  58. Rigby, R. E., Wise, H. M., Smith, N., Digard, P. & Rehwinkel, J. PA-X antagonises MAVS-dependent accumulation of early type I interferon messenger RNAs during influenza A virus infection. Sci. Rep. 9, 7216 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Cunningham, F. et al. Ensembl 2022. Nucleic Acids Res. 50, D988–D995 (2022).

    Article  CAS  PubMed  Google Scholar 

  60. Warr, A. et al. An improved pig reference genome sequence to enable pig genetics and genomics research. Gigascience 9, giaa051 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Li, M. et al. Comprehensive variation discovery and recovery of missing sequence in the pig genome using multiple de novo assemblies. Genome Res. 27, 865–874 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Zhang, Y. et al. Influenza Research Database: an integrated bioinformatics resource for influenza virus research. Nucleic Acids Res. 45, D466–D474 (2017).

    Article  CAS  PubMed  Google Scholar 

  63. Liang, H. et al. Expansion of genotypic diversity and establishment of 2009 H1N1 pandemic-origin internal genes in pigs in China. J. Virol. 88, 10864–10874 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Iqbal, M., Yaqub, T., Reddy, K. & McCauley, J. W. Novel genotypes of H9N2 influenza A viruses isolated from poultry in Pakistan containing NS genes similar to highly pathogenic H7N3 and H5N1 viruses. PLoS ONE 4, e5788 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Liu, Q. et al. Characterization of a highly pathogenic avian influenza H5N1 clade 2.3.4 virus isolated from a tree sparrow. Virus Res. 147, 25–29 (2010).

    Article  CAS  PubMed  Google Scholar 

  66. Li, Y. et al. Continued evolution of H5N1 influenza viruses in wild birds, domestic poultry, and humans in China from 2004 to 2009. J. Virol. 84, 8389–8397 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Tran, V. et al. Influenza virus repurposes the antiviral protein IFIT2 to promote translation of viral mRNAs. Nat. Microbiol. 5, 1490–1503 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Gaglia, M. Cut site preference allows influenza A virus PA-X to discriminate between host and viral mRNAs. Figshare https://doi.org/10.6084/m9.figshare.c.6616663 (2023).

  69. Sievers, F. et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank A. Tai and the personnel of the Tufts University Core Facility–Genomics Core for help with the sequencing; C. McCormick and D. Khaperskyy (Dalhousie University, Halifax, NS, Canada) and members of their laboratories for their advice and input;. C. McCormick (Dalhousie University, Halifax, NS, Canada), B. Glaunsinger (University of California, Berkeley, CA, USA), R. Webby (St Jude’s Children Research Hospital, Memphis, TN, USA), G. Dreyfuss (University of Pennsylvania, Philadelphia, PA, USA), J. Bloom (Fred Hutchinson Cancer Center, Seattle, WA, USA) and S. Lakdawala (Emory University School of Medicine, Atlanta, GA, USA) for constructs; M. G. Blanco (University of Virginia, Charlottesville, VA, USA) for providing the northern blot protocol; B. Moss (National Institute of Health, Bethesda, MD, USA) for cell lines; and J. Coffin, C. Moore, K. Munger (Tufts University, Boston, MA, USA) and members of the Gaglia Laboratory for critical reading of the manuscript. This work was supported by NIH grant R01 AI137358 (to M.M.G.). L.G. was supported by NIH F31 AI154587 and T32 GM007310. C.H.R. was partially supported by the Applied Mathematics Program of the US DOE Office of Science Advanced Scientific Computing Research under contract number DE-AC02-05CH11231. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: L.G. and M.M.G. Methodology: L.G., C.H.R. and M.M.G. Investigation: L.G., A.I. and I.G. Writing original draft: L.G. and M.M.G. Writing, review and editing: A.I., I.G. and C.H.R. Funding acquisition: M.M.G. Supervision: M.M.G.

Corresponding author

Correspondence to Marta M. Gaglia.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks the anonymous reviewers for their contribution to the peer review of this work.

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 Characteristics of the system used to identify PA-X cut sites transcriptome-wide.

(a) Strategy used to engineer a virus that lacks PA-X, PR8-PA(∆X), compared to WT PR8. (b) Protein lysates of XRN1 knock out A549 cells infected with WT PR8 or PR8-PA(∆X), or mock infected, were probed with antibodies against PR8 PA or β-tubulin as a loading control. Images are representative of 2 experiments. (c) Protein lysates of WT or XRN1 knock out (ko) A549 cells were probed with antibodies against Xrn1 or β-tubulin as a loading control to check for loss of XRN1. Images are representative of 2 experiments. (d, e) Representation of individual chromosomal positions in the 5′ RACE-seq data. For each sample, reads with their 5′ end mapping to the same nucleotide were counted and plotted to compare different datasets: (d) mock samples from replicate 1 vs. replicate 2, (e) replicate 1 WT PR8 vs. PR8-PA(∆X) (top), WT PR8 vs. mock infected (middle) and PR8-PA(∆X) vs. mock infected (bottom). For each plot, light blue dots correspond to locations that are unique to one sample, while red and black dots correspond to locations that are common in the two samples. Red dots represent locations that have two-fold or more reads in the sample on the x axis vs. than in the sample on the y axis. Similar plots were obtained when comparing other replicates. (f) PyDegradome was run on concatenated read files comparing WT PR8 vs. PR8-PA(∆X) infections using different parameters as indicated on the x axis (see Methods for parameter explanation). cl = confidence levels, mf = multiplicative factor. The window of analysis was 4 nucleotides in all cases. The total number of cut sites identified by PyDegradome (red circles, left y axis) and the percentage of shared cut sites also found as concatenated cut sites (blue diamonds, right y axis) were plotted for each set of parameters. Panel a adapted with permission from 4, Elsevier.

Extended Data Fig. 2 Further validation of PA-X cut sites identified by PyDegradome.

(a, b) XRN1 knock out A549 cells were infected with WT PR8 or PR8-PA(∆X), or mock infected. (a) RNA was isolated, 5′ RACE was performed using primers specific for BCAP31, TUBA1B, INSIG1 or SLC7A5, and the PCR products were run on an agarose gel. Primers were positioned ~ 200–300 nucleotides downstream of the predicted cut sites. The predicted size of DNA bands coming from cut sites identified by PyDegradome are indicated by red dotted arrows. (b) PR8 HA RNA levels were quantified by qRT-PCR, normalized to 18S and plotted as mean ± standard deviation. AU, arbitrary units; n.d. not defined; ns, not significant, One-way ANOVA with Dunnett’s multiple comparison test. n = 3 independent experiments. (c) HEK293T ishXRN1 cells were treated with no drug or doxycycline for 3-4 days to induce shRNAs against XRN1, then protein lysates were collected and probed with antibodies against XRN1, or β-tubulin as a loading control to check for efficient XRN1 knock down. Images are representative of 3 experiments. (d, e) HEK293T ishXRN1 cells were treated with doxycycline for 3-4 days to induce knock down of XRN1, then transfected with luciferase reporters containing 99 bp insertions from the indicated genes, and where indicated, with PR8 PA-X. (d) RNA was extracted and luciferase mRNA levels were quantified by qRT-PCR, normalized to 18S and plotted as mean ± standard deviation. AU, arbitrary units. luc + INSIG1/SLC7A5 n = 2, luc + STOML2/YKT6/TUBA1B n = 3, luc + BCAP31 n = 4, luc + repeat n = 8, luc n = 9 independent experiments. (e) The RNA was also used to run 5′ RACE. Expected sizes of DNA bands coming from cut sites in the introduced target sequences are indicated by red arrows, while the blue arrow in D indicates the size of the original luciferase cut site fragment. For all gels, the DNA bands were purified and sequenced to confirm their identities. Images are representative of 3 experiments or 2 experiments for luciferase + INSIG1/SLC7A5. The SLC7A5 sequence was only consistently cut when inside the luciferase reporter.

Extended Data Fig. 3 The cut sites identified by PyDegradome are specific to PA-X and conserved across multiple influenza strains.

(a–c) HEK293T ishXRN1 cells were treated with doxycycline for 3-4 days to induce knock down of XRN1, then transfected with luciferase reporters containing 99 bp insertions from the indicated genes, and where indicated, with WT PA-X from the PR8 strain (a–c), the catalytic mutant PR8 PA-X D108A (a), PR8 PA with a mutation to reduce frameshifting and prevent PA-X production (PA(fs)) (a), herpes simplex virus 1 (HSV-1) vhs or Kaposi’s sarcoma-associated herpes virus (KSHV) SOX (b), or WT PA-X from the Perth influenza strain (c). RNA was extracted and used to run 5′ RACE. Expected sizes of DNA bands coming from cut sites in the introduced target sequences are indicated by the red dotted arrows. (d) XRN1 knock out A549 cells were infected with WT PR8, Perth or pH1N1pdm09 or the corresponding PA(∆X) mutants, or mock infected. 5′ RACE was then performed using primers specific for STOML2 or YKT6 ~ 250–300 nt downstream of the predicted cut sites. The PCR products were run on an agarose gel. The predicted size of DNA bands coming from cut sites identified by PyDegradome are indicated by the red dotted arrows. For all gels, the DNA bands were purified and sequenced to confirm their identities, and images are representative of 3 experiments. (e) Protein alignment of PA-X from the three different influenza strains PR8, H1N1pdm09 and Perth, generated using Clustal Omega69.

Extended Data Fig. 4 PA-X preferentially cleaves RNA at GCUG tetramers based on the concatenated cut site analysis.

(a, b) WebLogo56 representation of base enrichment around PA-X cut sites (WT PR8 vs. PR8-PA(∆X), a) or around control sites enriched in the PR8-PA(∆X) sample (PR8-PA(∆X) vs. WT PR8, b) for sites predicted by PyDegradome using the concatenated approach. (c) Percentage of PA-X cut sites containing GCUG or a tetramer with one nucleotide difference from GCUG, for sites identified by PyDegradome using the concatenated approach. // indicates the location of the cut, that is GCU//G indicates that PA-X cuts between the U and the G. (d) Further breakdown of the PA-X cut sites containing a tetramer with one nucleotide difference from GCUG around the cut site (marked by an x). (e, f) HEK293T ishXRN1 cells were treated with doxycycline for 3-4 days to induce knock down of XRN1, then transfected with luciferase reporters containing 99 bp insertions from the indicated genes, with or without PR8 PA-X. (e) RNA was extracted and luciferase mRNA levels were quantified by qRT-PCR, normalized to 18S and plotted as mean ± standard deviation. Left plot shows mRNA levels of each reporter in the absence of PA-X. Right plot shows mRNA levels of each reporter in the presence vs. absence of PA-X. AU, arbitrary units. n = 3 independent experiments. (f) The RNA was also used to run 5′ RACE. Expected sizes of DNA bands coming from cut sites in the introduced target sequences are indicated by the red dotted arrows. For both gels, DNA bands were purified and sequenced to confirm their identities, and images are representative of 3 experiments.

Extended Data Fig. 5 PA-X preferentially cleaves RNA within a hairpin loop structure in transfected and infected cells.

(a) HEK293T ishXRN1 cells were treated with doxycycline for 3-4 days to induce knock down of XRN1, then transfected with luciferase reporters containing insertions of the indicated lengths from the BCAP31 and SLC7A5 genes, with or without PR8 PA-X. RNA was extracted and used to run 5′ RACE. Expected sizes of DNA bands coming from cut sites in the introduced target sequences are indicated by the red dotted arrows (~250 bp for 99 bp constructs, ~ 220 bp for 51 bp constructs, ~ 210 bp for 27 bp constructs and ~ 205 bp for the 15 bp constructs). (b) HEK293T ishXRN1 cells were treated with doxycycline for 3-4 days to induce knock down of XRN1, then transfected with luciferase reporters containing 15 bp insertions from the STOML2 gene with or without the indicated mutations. 24 hours post transfection, cells were infected with WT PR8 or PR8-PA(∆X), or mock infected overnight. RNA was then extracted and used to run 5′ RACE. Expected sizes of DNA bands coming from cut sites in the introduced target sequences are indicated by the red dotted arrow, while the blue arrow indicates the size of the original luciferase cut site fragment. For all gels, the DNA bands were purified and sequenced to confirm their identities, and images are representative of 3 experiments (b) or 2 experiments (a).

Extended Data Fig. 6 PA-X preferentially cleaves RNAs within exons.

(a–c) Percentage of PA-X cut sites found within introns or exons for sites identified by PyDegradome using the shared cut sites approach (a and c, left) or the concatenated approach (A and C, right), compared to the percentage of reads found in exons vs. introns (b). The reads used in B are from WT PR8 replicate 1 as an example, but other samples have similar read distribution. The difference between reads and PA-X cut site distribution in introns vs. exons is statistically significant (P < 0.001), but the difference between reads and ∆X specific sites distribution in the concatenated analysis is not (P > 0.2), Chi-Square Test for Goodness of Fit with degree of freedom 1. While only 3% of PR8-PA(∆X) fragments from the shared analysis mapped to introns (0.025 < P < 0.05 compared to reads distribution, Chi-Square Test for Goodness of Fit with degree of freedom 1), this may stem from the very low number of sites, 37, found with this method. (d–f) XRN1 knock out A549 cells were infected with WT PR8 or PR8-PA(∆X), or mock infected. 5′RACE was performed and PCR products were run on an agarose gel. The boxes show diagrams of the BCAP31, STOML2 or SLC7A5 genes and the positions of the reverse primers for 5′ RACE. Top gels are the same gels as Fig. 2b and Extended Data Fig. 3a, and are included for comparison. Red dotted arrows indicate the size of fragments originating from previously validated cut sites. Bottom gels show products obtained using reverse primers in the intron (light purple primers), and red dotted arrows indicate the predicted size of PCR products that would appear if PA-X cleaved unspliced pre-mRNAs. White arrowhead indicates fragments that map to exon/intron junctions. New gel images are representative of 3 experiments.

Extended Data Fig. 7 PA-X only cleaves STOML2 and YKT6 preferred sequences if they are located within exons.

(a) Diagram of the luciferase reporters tested. Green and magenta arrows indicate positions of 5′ RACE PCR reverse primers, vertical arrows indicate location of predicted cut sites. (b–d) Testing of luciferase reporters with sequence insertions in the introns, as shown in diagram in A. HEK293T ishXRN1 cells were treated with doxycycline for 3-4 days to induce knock down of XRN1, then transfected with luciferase reporters containing the 99 bp insertions in the indicated genes either in an exon or an intron, with or without PR8 PA-X. (b) RNA was extracted and luciferase mRNA levels were quantified by qRT-PCR, normalized to 18S and plotted as mean ± standard deviation. AU, arbitrary units; n = 3 except for luc + STOML2/YKT6 in exon n = 2, each independent experiments. The top graph shows levels of the processed mRNA (obtained using two primers that bind in exons). The bottom graph shows levels of the unspliced pre-mRNA (obtained using one primer in an exon and one primer in the intron). (c, d) RNA was also used to run 5′ RACE. PCR products were separated on an agarose gel. The top gel in C and the gel in D represent products obtained using the green arrow primer from A, and the bottom gel in C using the magenta arrow primer from A. Blue dotted arrow indicates size of PCR products originating from the original luciferase cut site, red dotted arrow indicates the predicted sizes of PCR products that would originate from the sequence inserted in the introns. Gel images are representative of 3 experiments.

Extended Data Fig. 8 GCUG tetramers are enriched in the human transcriptome.

The percentage of tetramers in the human and viral genomes and transcriptomes was calculated by counting the number of each tetramer and dividing it by the total number of tetramers in the sequence (that is length of the sequence minus 3). Percentages for each tetramer are plotted to visualize which tetramers are more abundant in the human genome vs. transcriptome (a) and the human transcriptome vs. the transcriptome of 3 influenza A virus strains (c-e). Each dot represents a specific tetramer, with the red dot representing the GCUG tetramer. (b) The percentage of GCUG tetramers is plotted for each influenza mRNA (that is the positive strand, purple), and for each influenza genomic RNA (that is negative strand, blue). Data are plotted as mean ± standard deviation, with each symbol representing a different influenza strain.

Extended Data Fig. 9 Viral mRNAs are not efficiently cleaved by PA-X.

(a, b) XRN1 knock out A549 cells were infected with WT PR8 or PR8-PA(∆X) (a), or WT H1N1pdm09 or H1N1pdm09-PA(∆X) (b), or mock infected. RNA was extracted to run 5′ RACE using primers ~ 150–200 nt downstream of GCUG sites that are predicted to be in a hairpin loop in the indicated viral mRNAs. Red dotted arrows indicate the predicted sizes of PCR products that would originate from cleavage at these GCUG sites. Gel images are representative of 3 experiments. (c) MDCK cells were infected at MOI 0.05 with the indicated viruses (see Fig. 5f for details on virus construction). Supernatants were collected at 1, 8, 24 and 48 hours post infection and viral titers were quantified by TCID50. n = 2 independent experiments.

Extended Data Fig. 10 Analysis of PyDegradome output characteristics.

(a) PyDegradome was run for each replicate on WT PR8 vs. PR8-PA(∆X) infection samples using different parameters as indicated on the x axis (see Methods). cl = confidence levels, mf = multiplicative factor. The window of analysis was 4 nucleotides in all cases. The percentage of cut sites shared between all three replicates was plotted for each set of parameters. (b) Histogram of the number of reads at the cut site for PA-X sites (PyDegradome comparison: WT PR8 vs. PR8-PA(∆X)) found in one (light grey), two (dark grey), or three replicates (‘all’, magenta), or through the concatenated analysis (‘concat’, orange). (c) Number of PA-X cut sites (orange squares) or RNA fragments enriched in the PR8-PA(∆X) infected cells (purple circles) identified by PyDegradome relative to the number of input reads. (d) Percentage of sites identified by PyDegradome in at least one replicate that are located within the first 20 nucleotides of a gene when comparing WT PR8 vs.PR8-PA(∆X), PR8-PA(∆X) vs. WT PR8, WT PR8 vs. mock and PR8-PA(∆X) vs. mock.

Supplementary information

Supplementary Information

Supplementary Tables 1 and 3.

Reporting Summary

Supplementary Table 2

List of PA-X cut site positions obtained from 5′ RACE-seq.

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

Gaucherand, L., Iyer, A., Gilabert, I. et al. Cut site preference allows influenza A virus PA-X to discriminate between host and viral mRNAs. Nat Microbiol 8, 1304–1317 (2023). https://doi.org/10.1038/s41564-023-01409-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-023-01409-8

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