Life-threatening influenza pneumonitis in a child with inherited IRF9 deficiency

Nicholas Hernandez, Isabelle Melki, Huie Jing, Tanwir Habib, Susie S Y Huang, Jeffrey Danielson, Tomasz Kula, Scott Drutman, Serkan Belkaya, Vimel Rattina, Lazaro Lorenzo-Diaz, Anais Boulai, Yoann Rose, Naoki Kitabayashi, Mathieu P Rodero, Cecile Dumaine, Stéphane Blanche, Marie-Noëlle Lebras, Man Chun Leung, Lisa Sara Mathew, Bertrand Boisson, Shen-Ying Zhang, Stephanie Boisson-Dupuis, Silvia Giliani, Damien Chaussabel, Luigi D Notarangelo, Stephen J Elledge, Michael J Ciancanelli, Laurent Abel, Qian Zhang, Nico Marr, Yanick J Crow, Helen C Su, Jean-Laurent Casanova, Nicholas Hernandez, Isabelle Melki, Huie Jing, Tanwir Habib, Susie S Y Huang, Jeffrey Danielson, Tomasz Kula, Scott Drutman, Serkan Belkaya, Vimel Rattina, Lazaro Lorenzo-Diaz, Anais Boulai, Yoann Rose, Naoki Kitabayashi, Mathieu P Rodero, Cecile Dumaine, Stéphane Blanche, Marie-Noëlle Lebras, Man Chun Leung, Lisa Sara Mathew, Bertrand Boisson, Shen-Ying Zhang, Stephanie Boisson-Dupuis, Silvia Giliani, Damien Chaussabel, Luigi D Notarangelo, Stephen J Elledge, Michael J Ciancanelli, Laurent Abel, Qian Zhang, Nico Marr, Yanick J Crow, Helen C Su, Jean-Laurent Casanova

Abstract

Life-threatening pulmonary influenza can be caused by inborn errors of type I and III IFN immunity. We report a 5-yr-old child with severe pulmonary influenza at 2 yr. She is homozygous for a loss-of-function IRF9 allele. Her cells activate gamma-activated factor (GAF) STAT1 homodimers but not IFN-stimulated gene factor 3 (ISGF3) trimers (STAT1/STAT2/IRF9) in response to IFN-α2b. The transcriptome induced by IFN-α2b in the patient's cells is much narrower than that of control cells; however, induction of a subset of IFN-stimulated gene transcripts remains detectable. In vitro, the patient's cells do not control three respiratory viruses, influenza A virus (IAV), parainfluenza virus (PIV), and respiratory syncytial virus (RSV). These phenotypes are rescued by wild-type IRF9, whereas silencing IRF9 expression in control cells increases viral replication. However, the child has controlled various common viruses in vivo, including respiratory viruses other than IAV. Our findings show that human IRF9- and ISGF3-dependent type I and III IFN responsive pathways are essential for controlling IAV.

© 2018 Hernandez et al.

Figures

Figure 1.
Figure 1.
A private IRF9 variant alters mRNA splicing in a child with severe influenza pneumonitis. (A) Pedigrees of the IRF9-deficient family. The double lines connecting the parents indicate consanguinity. The proband is indicated by an arrow. Filled shapes indicate affected individuals while open shapes identify unaffected individuals. (B) Chromatograms demonstrating c.991G>A mutation in patient PBMC-derived DNA (red arrow). (C) Population genetics of homozygous coding missense and predicted loss-of-function IRF9 mutations taken from GnomAD and in-house cohorts. The patient’s variant is private and shown in red, while two other variants, shown in blue, were also identified in our cohort. (D) Schematic illustration of the IRF9 gene. The exons are numbered 1–9, and regions corresponding to functionally significant domains are colored brown (for the DNA-binding domain, DBD), gray (nuclear localization sequence, NLS), or purple (IAD). Patient mutation indicated in red; other mutations indicated in blue. (E)IRF9 transcripts (left panel) and relative frequencies (right panel) produced during exon trapping in U2A cells. The results are representative of two independent experiments. (F) cDNA sequencing to detect the splicing of IRF9 mRNA from F-SV40 cells. Numbers of total and abnormal clones sequenced are indicated. Results representative of two experiments.
Figure 2.
Figure 2.
Impact of IRF9 Δex7 on IFN receptor-proximal signaling. (A) qRT-PCR measuring of IRF9 mRNA levels in PBMCs from the patient, her mother, and a healthy control with two probes—one probe spanning intron 7, and a second probe spanning intron 1. Representative results of four independent experiments are shown. (B) Top: WB of endogenous IRF9 in patient F-SV40 cells; GAPDH was used as a loading control. Bottom: STAT and phospho-STAT (pSTAT) levels were also assessed following stimulation with 1,000 U/ml of either IFN–α2b or –γ for 0.5 h on F-SV40 cells from two healthy controls (C1 and C2), the IRF9-deficient patient (IRF9−/−), her mother (IRF9+/−), a STAT1-deficient patient (STAT1−/−), a STAT2-deficient patient (STAT2−/−), an IFNGR2-deficient patient (IFNGR2−/−), and an IRF7-deficient patient (IRF7−/−). Representative results of five independent experiments are shown. (C) WB of IRF9 in IRF9-deficient U2A cells stably transfected with indicated variants (green: variants reported to be loss-of-function in in vitro assays, blue: variants found in-house, red: patient). GAPDH was used as loading control. Representative results of four independent experiments are shown. (D) WB of IRF9 in patient F-SV40 cells stably transfected with indicated variants. GAPDH was used as loading control. Representative results of four independent experiments are shown. (E) WB analysis of IRF9 localization in F-SV40 cells from two healthy controls (C1 and C2), the IRF9-deficient patient (IRF9−/−), her mother (IRF9+/−), a STAT1-deficient patient (STAT1−/−), a STAT2-deficient patient (STAT2−/−), an IFNGR2-deficient patient (IFNGR2−/−), and an IRF7-deficient patient (IRF7−/−). GAPDH and LaminA/C were used as loading controls. Representative results of three independent experiments are shown. (F) Reporter assays of ISRE or GAS-dependent firefly luciferase tested in U2A cells stimulated with 1,000 U/ml of either IFN-α2b or -γ for 16 h after being stably transfected with indicated variants (green: variants reported to be loss-of-function in in vitro assays, blue: variants found in-house, red: patient). The specific response to IFN stimulation was calculated by the ratio of firefly luciferase reporter gene activity to constitutively expressed renilla luciferase activity (RLU, relative luciferase ratio). Representative results of three independent experiments are shown. (G) EMSA analysis of ISRE and GAS binding by IFN-stimulated B-LCLs from three healthy controls (C1, C2, and C3), the IRF9-deficient patient (IRF9−/−), her mother (IRF9+/−), a STAT1-deficient patient (STAT1−/−), a STAT2-deficient patient (STAT2−/−), and an IRF7-deficient patient (IRF7−/−). Representative results of three independent experiments are shown.
Figure 3.
Figure 3.
Impaired ISG induction in IRF9-deficient cells. (A) Transcription levels of MX1, IFIT1, IFIT3, and CXCL9 assessed by qRT-PCR on F-SV40 cells treated with 1,000 U/ml of IFN–α2b, -β, or –γ for 2 h. Cells were from three healthy controls (C1, C2, and C3), an IRF9-deficient patient (IRF9−/−), her mother (IRF9+/−), and STAT1-deficient (STAT1−/−), STAT2-deficient (STAT2−/−), IRF7-deficient (IRF7−/−), and IFNGR1-deficient (IFNGR1−/−) patients. Representative results of four independent experiments are shown. (B and C) WB of MX1 and IFIT3 on F-SV40 (B) or B-LCL (C) cells treated with 1,000 U/ml of IFN–α2b for various time points. GAPDH was used as a loading control. Representative results of three independent experiments are shown. (D) Transcription levels of MX1, IFIT1, IFIT3, and CXCL9 assessed by qRT-PCR of B-LCL cells treated with 1,000 U/ml of IFN–α2b, -β, or –γ for 2 h. Cells were from three healthy controls (T1, T2, and T3), an IRF9-deficient patient (IRF9−/−), her mother (IRF9+/−), and STAT1-deficient (STAT1−/−), STAT2-deficient (STAT2−/−), IRF7-deficient (IRF7−/−), and IFNGR2-deficient (IFNGR2−/−) patients. Representative results four independent experiments are shown. (E) Transcription levels of MX1, IFIT1, and CXCL9 assessed by qRT-PCR in F-SV40 cells from a healthy control (C1), P’s mother (IRF9+/−), and P (IRF9−/−) stably transfected with luciferase as a control (Luc) or indicated IRF9 variants (WT: WT IRF9, green: reported loss-of-function variants, blue: variants found in-house, red: patient variant). Cells were stimulated with 1,000 U/ml of IFN-α2b, -β, or -γ for 2 or 8 h. Representative results of four independent experiments are shown. (F) Similar to E, qRT-PCR analysis of MX1, IFIT1, and CXCL9 expression in parental HT1080 cells and U2A cells. Cells were stimulated with 1,000 U/ml of IFN-α2b, -β, or -γ for 2 or 8 h. Representative results of three independent experiments are shown.
Figure 4.
Figure 4.
Transcriptomic analysis of ISGs in IRF9-deficient cells. mRNA-seq analysis of primary fibroblasts (A and C) and B-LCLs (B and D) from three healthy controls (C1, C2, and C3), the IRF9-deficient patient (IRF9−/−). Cells were treated with 1,000 U/ml IFN–α2b for 2 h. Heatmaps (A and B) show log2 FC values of all ISGs that were found to be differentially regulated (≥1.5-fold) in all three control subjects relative to unstimulated cells. Bar graphs (C and D) quantify the number of ISGs that were differentially regulated (≥1.5-fold) compared with unstimulated cells in healthy controls or the IRF9-deficient patient. (E) Shown are Δ log2 fold change values of a subset of ISGs that were found to be induced ≥1.5-fold (linear scale) in B-LCL cells (upper panels) or primary fibroblasts (lower panels) of the IRF9-deficient patient upon in vitro stimulation with IFN-α. To select this subset of ISGs, the IFN-α2b–induced genes in the healthy controls identified in the mRNA-seq analysis were used. In the IRF9-deficient patient, these genes were first passed through a filter by querying the gene identifiers against the interferome database and by retaining genes that were responsive to in vitro IFN stimulation. ISGs that failed to be induced at least 1.5-fold (linear scale) in patient cells were excluded. The retained ISGs were stratified in three groups of less (Δ less than −0.585), similar (−0.585 < Δ < 0.585), and higher (Δ > 0.585) induced genes relative to the average responses in the healthy control subjects. The numbers of genes in each group are shown in brackets. ***, significant differences at P < 0.0001 by the Kruskal-Wallis test. (F) Log2 FC of induced ISGs in IRF9-deficient B-LCLs (upper panels) or primary fibroblasts (lower panels) and their corresponding values in healthy donors. (G) Network analysis of a subset of highly inducible (> fivefold linear scale) ISGs among control subjects’ B-LCLs and their responsiveness in the IRF9-deficient patient. Biological pathway and physical interactions are depicted as blue and red lines, respectively. 1.5 FC was used as the cut-off to distinguish responsive (red circle) and nonresponsive (blue circle) ISGs. The highly inducible ISGs that were used for query are shaded in yellow.
Figure 5.
Figure 5.
Crippled control of IAV and other viral infections in IRF9-deficient cells. (A) IAV titers in F-SV40 unstimulated (left) or pretreated (right) with 1,000 U/ml IFN-α2 for 16 h, followed by infection with (A/H1N1/CA/2009) IAV at MOI = 1. Mean ± SD (n = 3) is shown. Cells from three healthy controls were included (C1, C2, and C3), as well as those from the IRF9-deficient patient (IRF9−/−), her mother (IRF9+/−), and STAT1-deficient (STAT1−/−), STAT2-deficient (STAT2−/−), and IRF7-deficient (IRF7−/−) patients. Four independent experiments (mean ± SD) are shown. (B) VSV titers in F-SV40 cells unstimulated (left) or pretreated (right) with 1,000 U/ml IFN-α2 for 16 h, followed by infection with VSV at MOI = 3. Four independent experiments (mean ± SD) are shown. (C) IAV titers in stably transfected F-SV40 cells unstimulated (left) or pretreated (right) with 1,000 U/ml IFN-α2 for 16 h, followed by infection with IAV at MOI = 1. Cells were from a healthy control (C1), a STAT1-deficient patient (STAT1−/−), P (IRF9−/−), and P’s cells stably transfected with luciferase or WT IRF9 (gray), variants reported to be loss-of-function in in vitro assays (green), variants found in-house (blue), or the patient’s variant (red). Three independent experiments (mean ± SD) are shown. (D) VSV titers in stably transfected F-SV40 cells unstimulated (left) or pretreated (right) with 1,000 U/ml IFN-α2 for 16 h, followed by infection with VSV at MOI = 3. Four independent experiments (mean ± SD) are shown. (E) Percentage of RSV-infected (GFP+) F-SV40 cells at 24 and 48 h after infection. Cells from three healthy controls were included (C1, C4, and C5, black), as well as those from the IRF9-deficient patient (IRF9−/−, red), and cells from STAT1-deficient (STAT1−/−), STAT2-deficient (STAT2−/−), and IRF7-deficient (IRF7−/−) patients. Three independent experiments (mean ± SD) are shown. (F) Mean fluorescence intensity (MFI) of RSV-infected (GFP+) F-SV40 cells at 24 and 48 h after infection. Three independent experiments (mean ± SD) are shown. (G) Percentage of PIV-infected (GFP+) F-SV40 cells at 24 and 48 h after infection. Three independent experiments (mean ± SD) are shown. (H) MFI of PIV-infected (GFP+) F-SV40 cells at 24 and 48 h after infection. Three independent experiments (mean ± SD) are shown. MFI of GFP+ cells in individual samples were normalized to the averaged MFI of the three healthy controls at 24 h after infection in F and H.
Figure 6.
Figure 6.
IRF9 is required for optimal control of viral infections. (A) WB confirms the efficiency of RNAi of IRF9 or MAVS in primary dermal fibroblasts. (B–F) Primary human dermal fibroblasts previously transfected with the indicated siRNA (negative control, IRF9, MAVS) were tested for control of HRV, RSV, and PIV. Cells were infected with HRV-A16 at MOI of 10 (B), RSV at MOI of 0.5 (C and D), or PIV3 at MOI of 0.1 (E and F). Relative HRV transcripts (B) were measured by qRT-PCR, and values were normalized to the siNeg control. Percentage of infected cells (C and E) and relative virus per infected cell (D and F) were measured by flow cytometric analysis of GFP+ cells. MFI of GFP+ cells in individual samples were normalized to negative control at 24 h (D and F). Shown are the mean ± SD of six (B–F) experiments. *, P < 0.05; **, P < 0.01, by Kruskal-Wallis test.

References

    1. Abdul-Sater A.A., Majoros A., Plumlee C.R., Perry S., Gu A.D., Lee C., Shresta S., Decker T., and Schindler C.. 2015. Different STAT Transcription Complexes Drive Early and Delayed Responses to Type I IFNs. J. Immunol. 195:210–216. 10.4049/jimmunol.1401139
    1. Adzhubei I.A., Schmidt S., Peshkin L., Ramensky V.E., Gerasimova A., Bork P., Kondrashov A.S., and Sunyaev S.R.. 2010. A method and server for predicting damaging missense mutations. Nat. Methods. 7:248–249. 10.1038/nmeth0410-248
    1. Anders S., Pyl P.T., and Huber W.. 2015. HTSeq--a Python framework to work with high-throughput sequencing data. Bioinformatics. 31:166–169. 10.1093/bioinformatics/btu638
    1. Andrews S.2010. FastQC: a quality control tool for high throughput sequence data. Available at: .
    1. Arimoto K.I., Löchte S., Stoner S.A., Burkart C., Zhang Y., Miyauchi S., Wilmes S., Fan J.B., Heinisch J.J., Li Z., et al. . 2017. STAT2 is an essential adaptor in USP18-mediated suppression of type I interferon signaling. Nat. Struct. Mol. Biol. 24:279–289. 10.1038/nsmb.3378
    1. Asgari S., Schlapbach L.J., Anchisi S., Hammer C., Bartha I., Junier T., Mottet-Osman G., Posfay-Barbe K.M., Longchamp D., Stocker M., et al. . 2017. Severe viral respiratory infections in children with IFIH1 loss-of-function mutations. Proc. Natl. Acad. Sci. USA. 114:8342–8347. 10.1073/pnas.1704259114
    1. Asselin-Paturel C., Brizard G., Chemin K., Boonstra A., O’Garra A., Vicari A., and Trinchieri G.. 2005. Type I interferon dependence of plasmacytoid dendritic cell activation and migration. J. Exp. Med. 201:1157–1167. 10.1084/jem.20041930
    1. Begue B., Verdier J., Rieux-Laucat F., Goulet O., Morali A., Canioni D., Hugot J.P., Daussy C., Verkarre V., Pigneur B., et al. . 2011. Defective IL10 signaling defining a subgroup of patients with inflammatory bowel disease. Am. J. Gastroenterol. 106:1544–1555. 10.1038/ajg.2011.112
    1. Belkadi A., Pedergnana V., Cobat A., Itan Y., Vincent Q.B., Abhyankar A., Shang L., El Baghdadi J., Bousfiha A., Alcais A., et al. Exome/Array Consortium . 2016. Whole-exome sequencing to analyze population structure, parental inbreeding, and familial linkage. Proc. Natl. Acad. Sci. USA. 113:6713–6718. 10.1073/pnas.1606460113
    1. Boisson B., Laplantine E., Prando C., Giliani S., Israelsson E., Xu Z., Abhyankar A., Israël L., Trevejo-Nunez G., Bogunovic D., et al. . 2012. Immunodeficiency, autoinflammation and amylopectinosis in humans with inherited HOIL-1 and LUBAC deficiency. Nat. Immunol. 13:1178–1186. 10.1038/ni.2457
    1. Boisson-Dupuis S., Kong X.F., Okada S., Cypowyj S., Puel A., Abel L., and Casanova J.L.. 2012. Inborn errors of human STAT1: allelic heterogeneity governs the diversity of immunological and infectious phenotypes. Curr. Opin. Immunol. 24:364–378. 10.1016/j.coi.2012.04.011
    1. Bousfiha A., Jeddane L., Picard C., Ailal F., Bobby Gaspar H., Al-Herz W., Chatila T., Crow Y.J., Cunningham-Rundles C., Etzioni A., et al. . 2018. The 2017 IUIS Phenotypic Classification for Primary Immunodeficiencies. J. Clin. Immunol. 38:129–143. 10.1007/s10875-017-0465-8
    1. Burns C., Cheung A., Stark Z., Choo S., Downie L., White S., Conyers R., and Cole T.. 2016. A novel presentation of homozygous loss-of-function STAT-1 mutation in an infant with hyperinflammation-A case report and review of the literature. J. Allergy Clin. Immunol. Pract. 4:777–779. 10.1016/j.jaip.2016.02.015
    1. Casanova J.-L. 2015a Human genetic basis of interindividual variability in the course of infection. Proc. Natl. Acad. Sci. USA. 112:E7118–E7127.
    1. Casanova J.-L. 2015b Severe infectious diseases of childhood as monogenic inborn errors of immunity. Proc. Natl. Acad. Sci. USA. 112:E7128–E7137.
    1. Casanova J.L., Conley M.E., Seligman S.J., Abel L., and Notarangelo L.D.. 2014. Guidelines for genetic studies in single patients: lessons from primary immunodeficiencies. J. Exp. Med. 211:2137–2149. 10.1084/jem.20140520
    1. Centers for Disease Control and Prevention 2013. FluView: Influenza Associated Hospitalization Surveillance Network. Available at: .
    1. Chapgier A., Boisson-Dupuis S., Jouanguy E., Vogt G., Feinberg J., Prochnicka-Chalufour A., Casrouge A., Yang K., Soudais C., Fieschi C., et al. . 2006. Novel STAT1 alleles in otherwise healthy patients with mycobacterial disease. PLoS Genet. 2:e131 10.1371/journal.pgen.0020131
    1. Ciancanelli M.J., Huang S.X.L., Luthra P., Garner H., Itan Y., Volpi S., Lafaille F.G., Trouillet C., Schmolke M., Albrecht R.A., et al. . 2015. Infectious disease. Life-threatening influenza and impaired interferon amplification in human IRF7 deficiency. Science. 348:448–453. 10.1126/science.aaa1578
    1. Ciancanelli M.J., Abel L., Zhang S.-Y., and Casanova J.-L.. 2016. Host genetics of severe influenza: from mouse Mx1 to human IRF7. Curr. Opin. Immunol. 38:109–120. 10.1016/j.coi.2015.12.002
    1. Coccia E.M., Severa M., Giacomini E., Monneron D., Remoli M.E., Julkunen I., Cella M., Lande R., and Uzé G.. 2004. Viral infection and Toll-like receptor agonists induce a differential expression of type I and lambda interferons in human plasmacytoid and monocyte-derived dendritic cells. Eur. J. Immunol. 34:796–805. 10.1002/eji.200324610
    1. Collin M., Dickinson R., and Bigley V.. 2015. Haematopoietic and immune defects associated with GATA2 mutation. Br. J. Haematol. 169:173–187. 10.1111/bjh.13317
    1. Darnell J.E. Jr., Kerr I.M., and Stark G.R.. 1994. Jak-STAT pathways and transcriptional activation in response to IFNs and other extracellular signaling proteins. Science. 264:1415–1421. 10.1126/science.8197455
    1. Dawood F.S., Kamimoto L., D’Mello T.A., Reingold A., Gershman K., Meek J., Arnold K.E., Farley M., Ryan P., Lynfield R., et al. Emerging Infections Program Network . 2011. Children with asthma hospitalized with seasonal or pandemic influenza, 2003-2009. Pediatrics. 128:e27–e32. 10.1542/peds.2010-3343
    1. de Weerd N.A., Vivian J.P., Nguyen T.K., Mangan N.E., Gould J.A., Braniff S.J., Zaker-Tabrizi L., Fung K.Y., Forster S.C., Beddoe T., et al. . 2013. Structural basis of a unique interferon-β signaling axis mediated via the receptor IFNAR1. Nat. Immunol. 14:901–907. 10.1038/ni.2667
    1. Dobin A., Davis C.A., Schlesinger F., Drenkow J., Zaleski C., Jha S., Batut P., Chaisson M., and Gingeras T.R.. 2013. STAR: ultrafast universal RNA-seq aligner. Bioinformatics. 29:15–21. 10.1093/bioinformatics/bts635
    1. Duncan C.J., Mohamad S.M., Young D.F., Skelton A.J., Leahy T.R., Munday D.C., Butler K.M., Morfopoulou S., Brown J.R., Hubank M., et al. . 2015. Human IFNAR2 deficiency: Lessons for antiviral immunity. Sci. Transl. Med. 7:307ra154 10.1126/scitranslmed.aac4227
    1. Eletto D., Burns S.O., Angulo I., Plagnol V., Gilmour K.C., Henriquez F., Curtis J., Gaspar M., Nowak K., Daza-Cajigal V., et al. . 2016. Biallelic JAK1 mutations in immunodeficient patient with mycobacterial infection. Nat. Commun. 7:13992 10.1038/ncomms13992
    1. Engelhardt K.R., and Grimbacher B.. 2014. IL-10 in humans: lessons from the gut, IL-10/IL-10 receptor deficiencies, and IL-10 polymorphisms. Curr. Top. Microbiol. Immunol. 380:1–18.
    1. Fu X.Y., Kessler D.S., Veals S.A., Levy D.E., and Darnell J.E. Jr. 1990. ISGF3, the transcriptional activator induced by interferon alpha, consists of multiple interacting polypeptide chains. Proc. Natl. Acad. Sci. USA. 87:8555–8559. 10.1073/pnas.87.21.8555
    1. Glocker E.O., Kotlarz D., Boztug K., Gertz E.M., Schäffer A.A., Noyan F., Perro M., Diestelhorst J., Allroth A., Murugan D., et al. . 2009. Inflammatory bowel disease and mutations affecting the interleukin-10 receptor. N. Engl. J. Med. 361:2033–2045. 10.1056/NEJMoa0907206
    1. Hahm B., Trifilo M.J., Zuniga E.I., and Oldstone M.B.. 2005. Viruses evade the immune system through type I interferon-mediated STAT2-dependent, but STAT1-independent, signaling. Immunity. 22:247–257. 10.1016/j.immuni.2005.01.005
    1. Hambleton S., Goodbourn S., Young D.F., Dickinson P., Mohamad S.M.B., Valappil M., McGovern N., Cant A.J., Hackett S.J., Ghazal P., et al. . 2013. STAT2 deficiency and susceptibility to viral illness in humans. Proc. Natl. Acad. Sci. USA. 110:3053–3058. 10.1073/pnas.1220098110
    1. Honda K., Yanai H., Negishi H., Asagiri M., Sato M., Mizutani T., Shimada N., Ohba Y., Takaoka A., Yoshida N., and Taniguchi T.. 2005. IRF-7 is the master regulator of type-I interferon-dependent immune responses. Nature. 434:772–777. 10.1038/nature03464
    1. Hoyos-Bachiloglu R., Chou J., Sodroski C.N., Beano A., Bainter W., Angelova M., Al Idrissi E., Habazi M.K., Alghamdi H.A., Almanjomi F., et al. . 2017. A digenic human immunodeficiency characterized by IFNAR1 and IFNGR2 mutations. J. Clin. Invest. 127:4415–4420. 10.1172/JCI93486
    1. Itan Y., Zhang S.Y., Vogt G., Abhyankar A., Herman M., Nitschke P., Fried D., Quintana-Murci L., Abel L., and Casanova J.L.. 2013. The human gene connectome as a map of short cuts for morbid allele discovery. Proc. Natl. Acad. Sci. USA. 110:5558–5563. 10.1073/pnas.1218167110
    1. Itan Y., Shang L., Boisson B., Patin E., Bolze A., Moncada-Vélez M., Scott E., Ciancanelli M.J., Lafaille F.G., Markle J.G., et al. . 2015. The human gene damage index as a gene-level approach to prioritizing exome variants. Proc. Natl. Acad. Sci. USA. 112:13615–13620. 10.1073/pnas.1518646112
    1. Itan Y., Shang L., Boisson B., Ciancanelli M.J., Markle J.G., Martinez-Barricarte R., Scott E., Shah I., Stenson P.D., Gleeson J., et al. . 2016. The mutation significance cutoff: gene-level thresholds for variant predictions. Nat. Methods. 13:109–110. 10.1038/nmeth.3739
    1. Iwasaki A., and Pillai P.S.. 2014. Innate immunity to influenza virus infection. Nat. Rev. Immunol. 14:315–328. 10.1038/nri3665
    1. Jaworska J., Gravel A., and Flamand L.. 2010. Divergent susceptibilities of human herpesvirus 6 variants to type I interferons. Proc. Natl. Acad. Sci. USA. 107:8369–8374. 10.1073/pnas.0909951107
    1. John J., McKendry R., Pellegrini S., Flavell D., Kerr I.M., and Stark G.R.. 1991. Isolation and characterization of a new mutant human cell line unresponsive to alpha and beta interferons. Mol. Cell. Biol. 11:4189–4195. 10.1128/MCB.11.8.4189
    1. Johnson K.D., Hsu A.P., Ryu M.J., Wang J., Gao X., Boyer M.E., Liu Y., Lee Y., Calvo K.R., Keles S., et al. . 2012. Cis-element mutated in GATA2-dependent immunodeficiency governs hematopoiesis and vascular integrity. J. Clin. Invest. 122:3692–3704. 10.1172/JCI61623
    1. Kaminski M.M., Ohnemus A., Cornitescu M., and Staeheli P.. 2012. Plasmacytoid dendritic cells and Toll-like receptor 7-dependent signalling promote efficient protection of mice against highly virulent influenza A virus. J. Gen. Virol. 93:555–559. 10.1099/vir.0.039065-0
    1. Kimura T., Kadokawa Y., Harada H., Matsumoto M., Sato M., Kashiwazaki Y., Tarutani M., Tan R.S., Takasugi T., Matsuyama T., et al. . 1996. Essential and non-redundant roles of p48 (ISGF3 gamma) and IRF-1 in both type I and type II interferon responses, as revealed by gene targeting studies. Genes Cells. 1:115–124. 10.1046/j.1365-2443.1996.08008.x
    1. Kircher M., Witten D.M., Jain P., O’Roak B.J., Cooper G.M., and Shendure J.. 2014. A general framework for estimating the relative pathogenicity of human genetic variants. Nat. Genet. 46:310–315. 10.1038/ng.2892
    1. Koerner I., Kochs G., Kalinke U., Weiss S., and Staeheli P.. 2007. Protective role of beta interferon in host defense against influenza A virus. J. Virol. 81:2025–2030. 10.1128/JVI.01718-06
    1. Kreins A.Y., Ciancanelli M.J., Okada S., Kong X.F., Ramírez-Alejo N., Kilic S.S., El Baghdadi J., Nonoyama S., Mahdaviani S.A., Ailal F., et al. . 2015. Human TYK2 deficiency: Mycobacterial and viral infections without hyper-IgE syndrome. J. Exp. Med. 212:1641–1662. 10.1084/jem.20140280
    1. Lamborn I.T., Jing H., Zhang Y., Drutman S.B., Abbott J.K., Munir S., Bade S., Murdock H.M., Santos C.P., Brock L.G., et al. . 2017. Recurrent rhinovirus infections in a child with inherited MDA5 deficiency. J. Exp. Med. 214:1949–1972. 10.1084/jem.20161759
    1. Lauterbach H., Bathke B., Gilles S., Traidl-Hoffmann C., Luber C.A., Fejer G., Freudenberg M.A., Davey G.M., Vremec D., Kallies A., et al. . 2010. Mouse CD8alpha+ DCs and human BDCA3+ DCs are major producers of IFN-lambda in response to poly IC. J. Exp. Med. 207:2703–2717. 10.1084/jem.20092720
    1. Li H., and Durbin R.. 2009. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. 25:1754–1760. 10.1093/bioinformatics/btp324
    1. Li H., Handsaker B., Wysoker A., Fennell T., Ruan J., Homer N., Marth G., Abecasis G., and Durbin R.. 1000 Genome Project Data Processing Subgroup . 2009. The Sequence Alignment/Map format and SAMtools. Bioinformatics. 25:2078–2079. 10.1093/bioinformatics/btp352
    1. Li W., Hofer M.J., Songkhunawej P., Jung S.R., Hancock D., Denyer G., and Campbell I.L.. 2017. Type I interferon-regulated gene expression and signaling in murine mixed glial cells lacking signal transducers and activators of transcription 1 or 2 or interferon regulatory factor 9. J. Biol. Chem. 292:5845–5859. 10.1074/jbc.M116.756510
    1. Lou Y.J., Pan X.R., Jia P.M., Li D., Xiao S., Zhang Z.L., Chen S.J., Chen Z., and Tong J.H.. 2009. IRF-9/STAT2 [corrected] functional interaction drives retinoic acid-induced gene G expression independently of STAT1. Cancer Res. 69:3673–3680. 10.1158/0008-5472.CAN-08-4922
    1. McKenna A., Hanna M., Banks E., Sivachenko A., Cibulskis K., Kernytsky A., Garimella K., Altshuler D., Gabriel S., Daly M., and DePristo M.A.. 2010. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20:1297–1303. 10.1101/gr.107524.110
    1. McQuillan R., Leutenegger A.L., Abdel-Rahman R., Franklin C.S., Pericic M., Barac-Lauc L., Smolej-Narancic N., Janicijevic B., Polasek O., Tenesa A., et al. . 2008. Runs of homozygosity in European populations. Am. J. Hum. Genet. 83:359–372. 10.1016/j.ajhg.2008.08.007
    1. Moens L., Van Eyck L., Jochmans D., Mitera T., Frans G., Bossuyt X., Matthys P., Neyts J., Ciancanelli M., Zhang S.Y., et al. . 2017. A novel kindred with inherited STAT2 deficiency and severe viral illness. J. Allergy Clin. Immunol. 139:1995–1997.e9. 10.1016/j.jaci.2016.10.033
    1. Mordstein M., Kochs G., Dumoutier L., Renauld J.C., Paludan S.R., Klucher K., and Staeheli P.. 2008. Interferon-lambda contributes to innate immunity of mice against influenza A virus but not against hepatotropic viruses. PLoS Pathog. 4:e1000151 10.1371/journal.ppat.1000151
    1. Mostafavi S., Yoshida H., Moodley D., LeBoité H., Rothamel K., Raj T., Ye C.J., Chevrier N., Zhang S.Y., Feng T., et al. Immunological Genome Project Consortium . 2016. Parsing the Interferon Transcriptional Network and Its Disease Associations. Cell. 164:564–578. 10.1016/j.cell.2015.12.032
    1. Ng P.C., and Henikoff S.. 2001. Predicting deleterious amino acid substitutions. Genome Res. 11:863–874. 10.1101/gr.176601
    1. Ning S., Pagano J.S., and Barber G.N.. 2011. IRF7: activation, regulation, modification and function. Genes Immun. 12:399–414. 10.1038/gene.2011.21
    1. O’Donnell L.A., Conway S., Rose R.W., Nicolas E., Slifker M., Balachandran S., and Rall G.F.. 2012. STAT1-independent control of a neurotropic measles virus challenge in primary neurons and infected mice. J. Immunol. 188:1915–1923. 10.4049/jimmunol.1101356
    1. Onodera K., Fujiwara T., Onishi Y., Itoh-Nakadai A., Okitsu Y., Fukuhara N., Ishizawa K., Shimizu R., Yamamoto M., and Harigae H.. 2016. GATA2 regulates dendritic cell differentiation. Blood. 128:508–518. 10.1182/blood-2016-02-698118
    1. Osterlund P.I., Pietilä T.E., Veckman V., Kotenko S.V., and Julkunen I.. 2007. IFN regulatory factor family members differentially regulate the expression of type III IFN (IFN-lambda) genes. J. Immunol. 179:3434–3442. 10.4049/jimmunol.179.6.3434
    1. Palese P., and Shaw M.L.. 2007. Fields Virology. Lippincott Williams & Wilkins, Philadephia.
    1. Pasquet M., Bellanné-Chantelot C., Tavitian S., Prade N., Beaupain B., Larochelle O., Petit A., Rohrlich P., Ferrand C., Van Den Neste E., et al. . 2013. High frequency of GATA2 mutations in patients with mild chronic neutropenia evolving to MonoMac syndrome, myelodysplasia, and acute myeloid leukemia. Blood. 121:822–829. 10.1182/blood-2012-08-447367
    1. Picard C., Bobby Gaspar H., Al-Herz W., Bousfiha A., Casanova J.L., Chatila T., Crow Y.J., Cunningham-Rundles C., Etzioni A., Franco J.L., et al. . 2018. International Union of Immunological Societies: 2017 Primary Immunodeficiency Diseases Committee Report on Inborn Errors of Immunity. J. Clin. Immunol. 38:96–128. 10.1007/s10875-017-0464-9
    1. Purcell S., Neale B., Todd-Brown K., Thomas L., Ferreira M.A., Bender D., Maller J., Sklar P., de Bakker P.I., Daly M.J., and Sham P.C.. 2007. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 81:559–575. 10.1086/519795
    1. Rengachari S., Groiss S., Devos J.M., Caron E., Grandvaux N., and Panne D.. 2018. Structural basis of STAT2 recognition by IRF9 reveals molecular insights into ISGF3 function. Proc. Natl. Acad. Sci. USA. 115:E601–E609. 10.1073/pnas.1718426115
    1. Rusinova I., Forster S., Yu S., Kannan A., Masse M., Cumming H., Chapman R., and Hertzog P.J.. 2013. Interferome v2.0: an updated database of annotated interferon-regulated genes. Nucleic Acids Res. 41(Database issue, D1):D1040–D1046. 10.1093/nar/gks1215
    1. Sato M., Hata N., Asagiri M., Nakaya T., Taniguchi T., and Tanaka N.. 1998. Positive feedback regulation of type I IFN genes by the IFN-inducible transcription factor IRF-7. FEBS Lett. 441:106–110. 10.1016/S0014-5793(98)01514-2
    1. Shahni R., Cale C.M., Anderson G., Osellame L.D., Hambleton S., Jacques T.S., Wedatilake Y., Taanman J.W., Chan E., Qasim W., et al. . 2015. Signal transducer and activator of transcription 2 deficiency is a novel disorder of mitochondrial fission. Brain. 138:2834–2846. 10.1093/brain/awv182
    1. Shieh W.J., Blau D.M., Denison A.M., Deleon-Carnes M., Adem P., Bhatnagar J., Sumner J., Liu L., Patel M., Batten B., et al. . 2010. 2009 pandemic influenza A (H1N1): pathology and pathogenesis of 100 fatal cases in the United States. Am. J. Pathol. 177:166–175.
    1. Short K.R., Kroeze E.J.B.V., Fouchier R.A.M., and Kuiken T.. 2014. Pathogenesis of influenza-induced acute respiratory distress syndrome. Lancet Infect. Dis. 14:57–69. 10.1016/S1473-3099(13)70286-X
    1. Sole-Violan J., Sologuren I., Betancor E., Zhang S., Pérez C., Herrera-Ramos E., Martínez-Saavedra M., López-Rodriguez M., Pestano J., Ruiz-Hernández J., et al. . 2013. Lethal influenza virus A H1N1 infection in two relatives with autosomal dominant GATA-2 deficiency. Crit. Care. 17(Suppl 2):15 10.1186/cc11953
    1. Sologuren I., Martinez-Saavedra M.T., Sole-Violan J., de Borges de Oliveira E. Jr., Betancor E., Casas I., Oleaga-Quintas C., Martinez-Gallo M., Zhang S.Y., Pestano J., et al. . 2018. Lethal Influenza in Two Related Adults with Inherited GATA2 Deficiency. J. Clin. Immunol.
    1. Staeheli P., Grob R., Meier E., Sutcliffe J.G., and Haller O.. 1988. Influenza virus-susceptible mice carry Mx genes with a large deletion or a nonsense mutation. Mol. Cell. Biol. 8:4518–4523. 10.1128/MCB.8.10.4518
    1. Stark G.R., and Darnell J.E. Jr. 2012. The JAK-STAT pathway at twenty. Immunity. 36:503–514. 10.1016/j.immuni.2012.03.013
    1. Stoletzki N., and Eyre-Walker A.. 2011. Estimation of the neutrality index. Mol. Biol. Evol. 28:63–70. 10.1093/molbev/msq249
    1. Tang X., Gao J.S., Guan Y.J., McLane K.E., Yuan Z.L., Ramratnam B., and Chin Y.E.. 2007. Acetylation-dependent signal transduction for type I interferon receptor. Cell. 131:93–105. 10.1016/j.cell.2007.07.034
    1. Taniguchi T., Ogasawara K., Takaoka A., and Tanaka N.. 2001. IRF family of transcription factors as regulators of host defense. Annu. Rev. Immunol. 19:623–655. 10.1146/annurev.immunol.19.1.623
    1. Taubenberger J.K., and Morens D.M.. 2006. 1918 Influenza: the mother of all pandemics. Emerg. Infect. Dis. 12:15–22. 10.3201/eid1209.05-0979
    1. Taubenberger J.K., and Morens D.M.. 2008. The pathology of influenza virus infections. Annu. Rev. Pathol. 3:499–522. 10.1146/annurev.pathmechdis.3.121806.154316
    1. Team R.D.C. 2008. R: A language and environment for statistical computing. R Foundation for Statistical Comuting, Vienna, Austria.
    1. Veals S.A., Schindler C., Leonard D., Fu X.Y., Aebersold R., Darnell J.E. Jr., and Levy D.E.. 1992. Subunit of an alpha-interferon-responsive transcription factor is related to interferon regulatory factor and Myb families of DNA-binding proteins. Mol. Cell. Biol. 12:3315–3324. 10.1128/MCB.12.8.3315
    1. Veals S.A., Santa Maria T., and Levy D.E.. 1993. Two domains of ISGF3 gamma that mediate protein-DNA and protein-protein interactions during transcription factor assembly contribute to DNA-binding specificity. Mol. Cell. Biol. 13:196–206. 10.1128/MCB.13.1.196
    1. Warde-Farley D., Donaldson S.L., Comes O., Zuberi K., Badrawi R., Chao P., Franz M., Grouios C., Kazi F., Lopes C.T., et al. . 2010. The GeneMANIA prediction server: biological network integration for gene prioritization and predicting gene function. Nucleic Acids Res. 38(Web Server issue, suppl_2):W214-20 10.1093/nar/gkq537
    1. Wilk E., Pandey A.K., Leist S.R., Hatesuer B., Preusse M., Pommerenke C., Wang J., and Schughart K.. 2015. RNAseq expression analysis of resistant and susceptible mice after influenza A virus infection identifies novel genes associated with virus replication and important for host resistance to infection. BMC Genomics. 16:655 10.1186/s12864-015-1867-8
    1. Writing Committee of the WHO Consultation on Clinical Aspects of Pandemic (H1N1) 2009 Influenza, Bautista E., Chotpitayasunondh T., Gao Z., Harper S.A., Shaw M., Uyeki T.M., Zaki S.R., Hayden F.G., Hui D.S., Kettner J.D., et al. . 2010. Clinical aspects of pandemic 2009 influenza A (H1N1) virus infection. N. Engl. J. Med. 362:1708–1719.
    1. Xu G.J., Kula T., Xu Q., Li M.Z., Vernon S.D., Ndung’u T., Ruxrungtham K., Sanchez J., Brander C., Chung R.T., et al. . 2015. Viral immunology. Comprehensive serological profiling of human populations using a synthetic human virome. Science. 348:aaa0698 10.1126/science.aaa0698
    1. Zaki M., Thoenes M., Kawalia A., Nürnberg P., Kaiser R., Heller R., and Bolz H.J.. 2017. Recurrent and Prolonged Infections in a Child with a Homozygous IFIH1 Nonsense Mutation. Front. Genet. 8:130.
    1. Zhang C., Cai Z., Kim Y.C., Kumar R., Yuan F., Shi P.Y., Kao C., and Luo G.. 2005. Stimulation of hepatitis C virus (HCV) nonstructural protein 3 (NS3) helicase activity by the NS3 protease domain and by HCV RNA-dependent RNA polymerase. J. Virol. 79:8687–8697. 10.1128/JVI.79.14.8687-8697.2005
    1. Zhang S.Y., Jouanguy E., Ugolini S., Smahi A., Elain G., Romero P., Segal D., Sancho-Shimizu V., Lorenzo L., Puel A., et al. . 2007. TLR3 deficiency in patients with herpes simplex encephalitis. Science. 317:1522–1527. 10.1126/science.1139522

Source: PubMed

3
Prenumerera