Recurrent rhinovirus infections in a child with inherited MDA5 deficiency

Ian T Lamborn, Huie Jing, Yu Zhang, Scott B Drutman, Jordan K Abbott, Shirin Munir, Sangeeta Bade, Heardley M Murdock, Celia P Santos, Linda G Brock, Evan Masutani, Emmanuel Y Fordjour, Joshua J McElwee, Jason D Hughes, Dave P Nichols, Aziz Belkadi, Andrew J Oler, Corinne S Happel, Helen F Matthews, Laurent Abel, Peter L Collins, Kanta Subbarao, Erwin W Gelfand, Michael J Ciancanelli, Jean-Laurent Casanova, Helen C Su, Ian T Lamborn, Huie Jing, Yu Zhang, Scott B Drutman, Jordan K Abbott, Shirin Munir, Sangeeta Bade, Heardley M Murdock, Celia P Santos, Linda G Brock, Evan Masutani, Emmanuel Y Fordjour, Joshua J McElwee, Jason D Hughes, Dave P Nichols, Aziz Belkadi, Andrew J Oler, Corinne S Happel, Helen F Matthews, Laurent Abel, Peter L Collins, Kanta Subbarao, Erwin W Gelfand, Michael J Ciancanelli, Jean-Laurent Casanova, Helen C Su

Abstract

MDA5 is a cytosolic sensor of double-stranded RNA (ds)RNA including viral byproducts and intermediates. We studied a child with life-threatening, recurrent respiratory tract infections, caused by viruses including human rhinovirus (HRV), influenza virus, and respiratory syncytial virus (RSV). We identified in her a homozygous missense mutation in IFIH1 that encodes MDA5. Mutant MDA5 was expressed but did not recognize the synthetic MDA5 agonist/(ds)RNA mimic polyinosinic-polycytidylic acid. When overexpressed, mutant MDA5 failed to drive luciferase activity from the IFNB1 promoter or promoters containing ISRE or NF-κB sequence motifs. In respiratory epithelial cells or fibroblasts, wild-type but not knockdown of MDA5 restricted HRV infection while increasing IFN-stimulated gene expression and IFN-β/λ. However, wild-type MDA5 did not restrict influenza virus or RSV replication. Moreover, nasal epithelial cells from the patient, or fibroblasts gene-edited to express mutant MDA5, showed increased replication of HRV but not influenza or RSV. Thus, human MDA5 deficiency is a novel inborn error of innate and/or intrinsic immunity that causes impaired (ds)RNA sensing, reduced IFN induction, and susceptibility to the common cold virus.

This is a work of the U.S. Government and is not subject to copyright protection in the United States. Foreign copyrights may apply.

Figures

Figure 1.
Figure 1.
Infection history in a human with recurrent respiratory tract infections. (A) Timeline of pathogens recovered from the respiratory tract. These were classified as positive single-stranded RNA virus, negative single-stranded RNA virus, double-stranded DNA virus, or bacteria as indicated in the adjacent symbol key. (B) RT-PCR molecular typing of RNA isolated from nasopharyngeal samples, using primer sets that preferentially amplify the indicated HRV species. Patient’s samples a to h, as indicated in A, were collected between 2 and 4 yr of age. Sample a was negative for respiratory pathogens and h positive for influenza A. Purified HRV-B14 and -A16 virus stocks, and a sample from a different subject having respiratory symptoms but negative for HRV/enterovirus, were also tested. Results are representative of three repeats. (C) Phylogenetic tree based on nucleotide sequences of 5′UTR of HRV isolates showing evolutionary relationship of the patient’s samples to closest serotypes and representative HRV species.
Figure 2.
Figure 2.
Autosomal recessive, homozygous IFIH1 mutation in the proband. (A) Pedigree indicating genotypes. (B) Confirmatory Sanger sequencing. Schematic below showing mutation location relative to MDA5 protein domains. (C) Ribbon diagram of the MDA5 structure with close-up showing lysine 365 interaction with ribose in RNA, and effects of glutamic acid substitution on distances (Å) between charged groups (yellow sphere, positive; red spheres, negative). (D) Immunoblot of MDA5, RIG-I, and MAVS proteins, relative to β-actin, in cycling T cells, either untreated or treated for 20 h with IFN-α. NC, healthy normal control. Pt, patient. (E) Immunoblot of overexpressed MDA5 (WT or K365E) after affinity precipitation with biotinylated-poly(I:C). D and E are representative of four repeats.
Figure 3.
Figure 3.
Loss-of-function IFIH1 mutation. (A, C, and D) Relative increase in normalized luciferase activity driven by the IFNΒ1 promoter (A), ISRE (C), or NF-κB (D) reporter constructs. Cells were cotransfected with WT or mutant MDA5, and with (filled) or without (open) transfected poly(I:C) as indicated. R337G and R779H are gain-of-function mutants. (B) Immunoblot for MDA5 proteins after transient transfection of 20 ng WT or mutant K365E MDA5, both under CMV promoters. (E) Similar to A, except that cells were cotransfected with 20 ng WT MDA5 (under CMV promoter) and either K365E MDA5 or empty vector (EV; under MSCV promoter). (F) Immunoblot for MDA5 proteins in lysates from E. Data show means ± SD from four (A and E), three (C), and five experiments (D). Data in B and F are examples from experiments shown in A and C–E, respectively. 293T cells lack endogenous MDA5 expression (not depicted). Equivalent lysates from ∼30,000 cells were run across lanes. **, P < 0.01; ***, P < 0.001; ****, P < 0.0001, by one-way ANOVA.
Figure 4.
Figure 4.
Loss of MDA5 function results in increased replication of HRV in respiratory epithelial cells. (A) HRV transcripts, normalized to nonspecific siRNA negative (siNeg) control at 20 h. HRV-B14–infected (MOI: 1) A549 cells were previously transfected with the indicated siRNA. (B) Immunoblot showing efficiencies of MDA5 and RIG-I knockdown in A. Transfected cells were left uninfected or infected with HRV-B14 for 48 h. 120 µg of lysates were run per lane. (C) Similar to A, at 48 h after infection and two rounds of transient transfection with the indicated siRNA. (D) Immunoblot showing efficiencies of MAVS and MDA5 knockdown in C. (E) HRV transcripts and HRV simultaneously quantitated by infectious plaque assay, at 60 h after HRV-B14 infection. (F) Number of HRV reads in RNA-seq data during HRV-B14 infection. A549 cells were previously transfected with MDA5 or nonspecific negative control siRNA (siNeg). Mean of triplicate expression ratios from triplicate infections were shown for each time point. (G) HRV transcripts, normalized to nonspecific siRNA negative (siNeg) control at 48 h, HRV-A16–infected (MOI: 10). Human primary fibroblasts were previously transfected with the indicated siRNA. (H) Immunoblot showing efficiencies of MDA5 and MAVS knockdown in G, except that cells were treated overnight with IFN-α. Data show means ± SD from six (A); eight (C), four (E), and seven experiments (G). Representative experiments are shown in B, D, and H. *, P < 0.05; **, P < 0.01; ***, P < 0.001; ****, P < 0.0001, by Kruskal-Wallis test (A, C, and G); Mann-Whitney U test in (E); and Kolmogorov-Smirnov test (F); other comparisons in A and F were nonsignificant.
Figure 5.
Figure 5.
MDA5 effects on IFN response during HRV infection. (A) “Response to type I interferon” genes (GO term GO0034340) versus all genes. (B) Heatmap representation of RNA-seq data showing HRV-induced IFN-regulated gene transcripts (RPKM) over the course of infection with HRV-B14 (MOI: 1). Data in A show box-and-whisker plots from triplicate infections. A549 cells were previously transfected with MDA5 or nonspecific negative control siRNA (siNeg) and correspond to Fig. 4 F. Mean of triplicate expression ratios from triplicate infections were shown for each time point in B. *, P < 0.05, by Kolmogorov-Smirnov test in A and B; other comparisons were nonsignificant. (C) IFNB1 transcripts were first normalized to β-actin expression and then shown relative to uninfected cells for each siRNA. HRV-A16-infected (MOI: 10) primary human fibroblasts were previously transfected with the indicated siRNA. (D) Similar to C, except for IFNL3 (IL-28) transcripts. (E) Immunoblot showing efficiencies of MAVS and MDA5 knockdown in C and D, except that cells were treated overnight with IFN-α. Data show means ± SD from four independent experiments in C and D, and a representative experiment in E. *, P < 0.05, by Kruskal-Wallis test comparing areas under the curves in C and D.
Figure 6.
Figure 6.
Loss of MDA5 function results in increased replication of HRV in respiratory epithelial cells and fibroblasts. (A) HRV transcripts, normalized to father control at 48 h. Primary nasal epithelial cells from the patient, parents, and six normal healthy controls were infected with HRV-B14 (MOI: 1) as indicated. (B) HRV transcripts in maternal fibroblasts, gene-edited to have IFIH1 genotypes WT/- (2 clones), WT/K365E (1 clone), K365E/- (2 clones), −/− (3 clones), after infection with HRV-B14 (MOI: 1). (C) HRV transcripts in A549 cells previously transduced with empty vector (EV), WT, or K365E MDA5, at 72 h after infection with HRV-B14 (MOI:1). (D) Immunoblot showing MDA5 overexpression and comparable RIG-Iexpression in C. 10.5 µg of lysates were run per lane. Data show means ± SD from two (A), six (B), and five experiments (C). *, P < 0.05; ***, P < 0.001 by Kruskal-Wallis test; comparisons in B and C were nonsignificant.
Figure 7.
Figure 7.
Loss of MDA5 function does not affect replication of influenza virus or production of proinflammatory cytokines in respiratory epithelial cells. (A) Influenza (MOI: 0.1) transcripts, normalized to siNeg control at 24 h. Influenza-infected A549 cells were previously transfected with the indicated siRNA. (B) Immunoblotting showing efficiencies of MDA5, RIG-I, and MAVS protein expression, relative to HSP90 loading control, after transient transfection with the indicated siRNA into A549 cells. Transfected cells were left uninfected or infected with influenza strain A/Victoria/361/2011 (MOI: 0.5) for 24 h or treated with IFN-β (10 IU/ml) for 24 h. 20 µg of lysates were run per lane. Shown is a representative experiment corresponding to A. (C–F) Proinflammatory gene transcripts quantitated by qRT-PCR from A. Levels of IL-1α (C), IL-6 (D), IL-8 (E), and TNF (F) were normalized to β-actin and are shown relative to normalized levels at 8 h after infection. Data show means ± SD from six to seven experiments (A) and three independent experiments in (C–F). *, P < 0.05, by Kruskal-Wallis test (A); all other comparisons were nonsignificant.
Figure 8.
Figure 8.
Loss of MDA5 function does not affect influenza virus replication, or influenza-induced IFN production or cytotoxicity, in respiratory epithelial cells or fibroblasts. (A) Influenza transcripts, normalized to father control at 24 h. Primary nasal epithelial cells from the patient, parents, and two normal healthy controls were infected with influenza (MOI: 0.02). (B) Influenza virus quantitated by infectious plaque assay, after infection (MOI: 1) of SV40-transformed fibroblasts having the indicated genotypes: empty squares, IFIH1−/−; empty inverted triangles, IFIH1K365E/-; solid triangles, IFIH1K365E/WT; empty circles, STAT1−/−; solid circles or solid squares, healthy controls. (C) IFN-β released into supernatants as measured by ELISA after infection with influenza strain A/Puerto Rico/8/1934 (MOI: 0.37–54). Decreasing MOI of influenza virus correspond to the bars proceeding from left to right for each cell line. Sendai Virus (SeV) was included as positive control of IFN induction as the right-most bar for each cell line. Genotypes of SV40-transformed fibroblasts are as indicated, with each cell line separated by vertical dotted lines. (D) LDH released into samples from C. Data show means ± SD from four experiments (A) and three independent experiments (B) that are representative of 11 experiments with varying MOIs (0.1–30) and three different influenza strains (A/Netherlands/602/2009, A/California/4/2009, and A/Puerto Rico/8/1934), and three independent experiments (C and D) that are representative of 6 and 7 experiments, respectively, in which the MOI were varied.
Figure 9.
Figure 9.
Loss of MDA5 function does not affect RSV replication, whereas affecting RSV-induced IFN-regulated transcripts. (A) Number of RSV reads in RNA-seq data during RSV infection. RSV-infected (MOI: 1) A549 cells were previously transfected with MDA5 or nonspecific negative control siRNA (siNeg). (B) Immunoblotting showing efficiencies of MDA5 protein expression, after transient transfection of A549 cells with the indicated siRNA, or without transfection (MOCK). Cells were either left uninfected or infected with RSV, as indicated. 20 µg of lysates were run per lane. Shown is a representative experiment corresponding to A. (C) “Response to type I interferon” genes (GO term GO0034340) versus all genes, from A. (D) Heatmap representation of RNA-seq data showing the expression change between MDA5 siRNA and nonspecific siRNA control of IFN-regulated gene over the course of RSV infection. Mean of triplicate expression ratios from triplicate infections were shown for each time point in A, C, and D. (E) RSV transcripts in primary nasal epithelial cells from the patient (open bar), parents (hatched bars), and two normal healthy controls (solid bars), normalized to father at 6 h, after RSV-GFP infection (MOI: 0.2). (F) Percent GFP+ of gated live RSV-infected cells from (E). Data show means ± SD from four experiments (E and F). *, P < 0.05, by Kruskal-Wallis test (C and D); other comparisons were nonsignificant.
Figure 10.
Figure 10.
Functional activity of population-level missense variants in IFIH1. (A) IFNB1-driven luciferase activity of overexpressed IFIH1 variants, measured as means % of the wild-type from at least three independent experiments, plotted against CADD score (A) or MAF (B). Black, missense variants. Orange, patient K365E variant. Red, loss-of-function (nonsense, splice acceptor/donor, frameshift) variants. Blue, gain-of-function variants. MSC, mutation significance cutoff. Variant information with corresponding plotted values (100 ng of construct) are presented in Table S3.

References

    1. Arora R., Kaplan M., and Nelson M.. 2011. Enterovirus-specific IgG in intravenous immunoglobulin preparations. Ann. Allergy Asthma Immunol. 106:544–545. 10.1016/j.anai.2011.03.007
    1. Balish A.L., Katz J.M., and Klimov A.I.. 2013. Influenza: propagation, quantification, and storage. Curr Protoc Microbiol Chapter 15:Unit 15G 11.
    1. Baños-Lara M.R., Ghosh A., and Guerrero-Plata A.. 2013. Critical role of MDA5 in the interferon response induced by human metapneumovirus infection in dendritic cells and in vivo. J. Virol. 87:1242–1251. 10.1128/JVI.01213-12
    1. Benitez A.A., Panis M., Xue J., Varble A., Shim J.V., Frick A.L., López C.B., Sachs D., and tenOever B.R.. 2015. In Vivo RNAi Screening Identifies MDA5 as a Significant Contributor to the Cellular Defense against Influenza A Virus. Cell Reports. 11:1714–1726. 10.1016/j.celrep.2015.05.032
    1. Bochkov Y.A., Grindle K., Vang F., Evans M.D., and Gern J.E.. 2014. Improved molecular typing assay for rhinovirus species A, B, and C. J. Clin. Microbiol. 52:2461–2471. 10.1128/JCM.00075-14
    1. Byington C.L., Ampofo K., Stockmann C., Adler F.R., Herbener A., Miller T., Sheng X., Blaschke A.J., Crisp R., and Pavia A.T.. 2015. Community Surveillance of Respiratory Viruses Among Families in the Utah Better Identification of Germs-Longitudinal Viral Epidemiology (BIG-LoVE) Study. Clin. Infect. Dis. 61:1217–1224. 10.1093/cid/civ486
    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. Case D.A., Berryman J.T., Betz R.M., Cerutti D.S., and Cheatham I. T.E., T.A. Darden, R.E. Duke, T.J. Giese, H. Gohlke, A.W. Goetz, N. Homeyer, S. Izadi, P. Janowski, J. Kaus, A. Kovalenko, T.S. Lee, S. LeGrand, P. Li, T. Luchko, R. Luo, B. Madej, K.M. Merz, G. Monard, P. Needham, H. Nguyen, H.T. Nguyen, I. Omelyan, A. Onufriev, D.R. Roe, A. Roitberg, R. Salomon-Ferrer, C.L. Simmerling, W. Smith, J. Swails, R.C. Walker, J. Wang, R.M. Wolf, X. Wu, D.M. York, and P.A. Kollman. 2015. AMBER 2015. In University of California, San Francisco.
    1. Chapgier A., Wynn R.F., Jouanguy E., Filipe-Santos O., Zhang S., Feinberg J., Hawkins K., Casanova J.L., and Arkwright P.D.. 2006. Human complete Stat-1 deficiency is associated with defective type I and II IFN responses in vitro but immunity to some low virulence viruses in vivo. J. Immunol. 176:5078–5083. 10.4049/jimmunol.176.8.5078
    1. Ciancanelli M.J., Huang S.X., 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. de Chasseval R., and de Villartay J.P.. 1992. High level transient gene expression in human lymphoid cells by SV40 large T antigen boost. Nucleic Acids Res. 20:245–250. 10.1093/nar/20.2.245
    1. Delaloye J., Roger T., Steiner-Tardivel Q.G., Le Roy D., Knaup Reymond M., Akira S., Petrilli V., Gomez C.E., Perdiguero B., Tschopp J., et al. . 2009. Innate immune sensing of modified vaccinia virus Ankara (MVA) is mediated by TLR2-TLR6, MDA-5 and the NALP3 inflammasome. PLoS Pathog. 5:e1000480 10.1371/journal.ppat.1000480
    1. Deschamps M., Laval G., Fagny M., Itan Y., Abel L., Casanova J.L., Patin E., and Quintana-Murci L.. 2016. Genomic Signatures of Selective Pressures and Introgression from Archaic Hominins at Human Innate Immunity Genes. Am. J. Hum. Genet. 98:5–21. 10.1016/j.ajhg.2015.11.014
    1. Dittmann M., Hoffmann H.H., Scull M.A., Gilmore R.H., Bell K.L., Ciancanelli M., Wilson S.J., Crotta S., Yu Y., Flatley B., et al. . 2015. A serpin shapes the extracellular environment to prevent influenza A virus maturation. Cell. 160:631–643. 10.1016/j.cell.2015.01.040
    1. Eilertson K.E., Booth J.G., and Bustamante C.D.. 2012. SnIPRE: selection inference using a Poisson random effects model. PLOS Comput. Biol. 8:e1002806 10.1371/journal.pcbi.1002806
    1. Errett J.S., Suthar M.S., McMillan A., Diamond M.S., and Gale M. Jr. 2013. The essential, nonredundant roles of RIG-I and MDA5 in detecting and controlling West Nile virus infection. J. Virol. 87:11416–11425. 10.1128/JVI.01488-13
    1. Galama J.M., Gielen M., and Weemaes C.M.. 2000. Enterovirus antibody titers after IVIG replacement in agammaglobulinemic children. Clin. Microbiol. Infect. 6:630–632. 10.1046/j.1469-0691.2000.00173.x
    1. Gaunt E.R., Harvala H., McIntyre C., Templeton K.E., and Simmonds P.. 2011. Disease burden of the most commonly detected respiratory viruses in hospitalized patients calculated using the disability adjusted life year (DALY) model. J. Clin. Virol. 52:215–221. 10.1016/j.jcv.2011.07.017
    1. Gitlin L., Barchet W., Gilfillan S., Cella M., Beutler B., Flavell R.A., Diamond M.S., and Colonna M.. 2006. Essential role of mda-5 in type I IFN responses to polyriboinosinic:polyribocytidylic acid and encephalomyocarditis picornavirus. Proc. Natl. Acad. Sci. USA. 103:8459–8464. 10.1073/pnas.0603082103
    1. Gitlin L., Benoit L., Song C., Cella M., Gilfillan S., Holtzman M.J., and Colonna M.. 2010. Melanoma differentiation-associated gene 5 (MDA5) is involved in the innate immune response to Paramyxoviridae infection in vivo. PLoS Pathog. 6:e1000734 10.1371/journal.ppat.1000734
    1. Global Burden of Disease Study 2013 Collaborators 2015. Global, regional, and national incidence, prevalence, and years lived with disability for 301 acute and chronic diseases and injuries in 188 countries, 1990-2013: a systematic analysis for the Global Burden of Disease Study 2013. Lancet. 386:743–800. 10.1016/S0140-6736(15)60692-4
    1. Grandvaux N., Guan X., Yoboua F., Zucchini N., Fink K., Doyon P., Martin L., Servant M.J., and Chartier S.. 2014. Sustained activation of interferon regulatory factor 3 during infection by paramyxoviruses requires MDA5. J. Innate Immun. 6:650–662. 10.1159/000360764
    1. Greenberg S.B. 2011. Update on rhinovirus and coronavirus infections. Semin. Respir. Crit. Care Med. 32:433–446. 10.1055/s-0031-1283283
    1. Hanawa H., Hematti P., Keyvanfar K., Metzger M.E., Krouse A., Donahue R.E., Kepes S., Gray J., Dunbar C.E., Persons D.A., and Nienhuis A.W.. 2004. Efficient gene transfer into rhesus repopulating hematopoietic stem cells using a simian immunodeficiency virus-based lentiviral vector system. Blood. 103:4062–4069. 10.1182/blood-2004-01-0045
    1. Hasegawa K., Mansbach J.M., and Camargo C.A. Jr. 2014. Infectious pathogens and bronchiolitis outcomes. Expert Rev. Anti Infect. Ther. 12:817–828. 10.1586/14787210.2014.906901
    1. Heikkinen T., and Järvinen A.. 2003. The common cold. Lancet. 361:51–59. 10.1016/S0140-6736(03)12162-9
    1. Hornak V., Abel R., Okur A., Strockbine B., Roitberg A., and Simmerling C.. 2006. Comparison of multiple Amber force fields and development of improved protein backbone parameters. Proteins. 65:712–725. 10.1002/prot.21123
    1. Israel L., Wang Y., Bulek K., Della Mina E., Zhang Z., Pedergnana V., Chrabieh M., Lemmens N.A., Sancho-Shimizu V., Descatoire M., et al. . 2017. Human Adaptive Immunity Rescues an Inborn Error of Innate Immunity. Cell. 168:789–800.e10. 10.1016/j.cell.2017.01.039
    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. Jaïdane H., Sauter P., Sane F., Goffard A., Gharbi J., and Hober D.. 2010. Enteroviruses and type 1 diabetes: towards a better understanding of the relationship. Rev. Med. Virol. 20:265–280. 10.1002/rmv.647
    1. Jain S., Williams D.J., Arnold S.R., Ampofo K., Bramley A.M., Reed C., Stockmann C., Anderson E.J., Grijalva C.G., Self W.H., et al. CDC EPIC Study Team . 2015. Community-acquired pneumonia requiring hospitalization among U.S. children. N. Engl. J. Med. 372:835–845. 10.1056/NEJMoa1405870
    1. Jin Y.H., Kim S.J., So E.Y., Meng L., Colonna M., and Kim B.S.. 2012. Melanoma differentiation-associated gene 5 is critical for protection against Theiler’s virus-induced demyelinating disease. J. Virol. 86:1531–1543. 10.1128/JVI.06457-11
    1. Jing H., Zhang Q., Zhang Y., Hill B.J., Dove C.G., Gelfand E.W., Atkinson T.P., Uzel G., Matthews H.F., Mustillo P.J., et al. . 2014. Somatic reversion in dedicator of cytokinesis 8 immunodeficiency modulates disease phenotype. J. Allergy Clin. Immunol. 133:1667–1675. 10.1016/j.jaci.2014.03.025
    1. Kato H., Takeuchi O., Sato S., Yoneyama M., Yamamoto M., Matsui K., Uematsu S., Jung A., Kawai T., Ishii K.J., et al. . 2006. Differential roles of MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature. 441:101–105. 10.1038/nature04734
    1. Kim W.K., Jain D., Sánchez M.D., Koziol-White C.J., Matthews K., Ge M.Q., Haczku A., Panettieri R.A. Jr., Frieman M.B., and López C.B.. 2014. Deficiency of melanoma differentiation-associated protein 5 results in exacerbated chronic postviral lung inflammation. Am. J. Respir. Crit. Care Med. 189:437–448. 10.1164/rccm.201307-1338OC
    1. Kingston R.E., Chen C.A., and Okayama H.. 2003. Calcium phosphate transfection. Curr Protoc Cell Biol Chapter 20:Unit 20 23.
    1. Lee W.M., Chen Y., Wang W., and Mosser A.. 2015a Growth of human rhinovirus in H1-HeLa cell suspension culture and purification of virions. Methods Mol. Biol. 1221:49–61. 10.1007/978-1-4939-1571-2_5
    1. Lee W.M., Chen Y., Wang W., and Mosser A.. 2015b Infectivity assays of human rhinovirus-A and -B serotypes. Methods Mol. Biol. 1221:71–81. 10.1007/978-1-4939-1571-2_7
    1. Lee W.M., Grindle K., Vrtis R., Pappas T., Vang F., Lee I., and Gern J.E.. 2015c Molecular identification and quantification of human rhinoviruses in respiratory samples. Methods Mol. Biol. 1221:25–38. 10.1007/978-1-4939-1571-2_3
    1. Li J., Liu Y., and Zhang X.. 2010. Murine coronavirus induces type I interferon in oligodendrocytes through recognition by RIG-I and MDA5. J. Virol. 84:6472–6482. 10.1128/JVI.00016-10
    1. Li P., Roberts B.P., Chakravorty D.K., and Merz K.M. Jr. 2013. Rational Design of Particle Mesh Ewald Compatible Lennard-Jones Parameters for +2 Metal Cations in Explicit Solvent. J. Chem. Theory Comput. 9:2733–2748. 10.1021/ct400146w
    1. Lin R., Génin P., Mamane Y., and Hiscott J.. 2000. Selective DNA binding and association with the CREB binding protein coactivator contribute to differential activation of alpha/beta interferon genes by interferon regulatory factors 3 and 7. Mol. Cell. Biol. 20:6342–6353. 10.1128/MCB.20.17.6342-6353.2000
    1. Loo Y.M., Fornek J., Crochet N., Bajwa G., Perwitasari O., Martinez-Sobrido L., Akira S., Gill M.A., García-Sastre A., Katze M.G., and Gale M. Jr. 2008. Distinct RIG-I and MDA5 signaling by RNA viruses in innate immunity. J. Virol. 82:335–345. 10.1128/JVI.01080-07
    1. Mäkelä M.J., Puhakka T., Ruuskanen O., Leinonen M., Saikku P., Kimpimäki M., Blomqvist S., Hyypiä T., and Arstila P.. 1998. Viruses and bacteria in the etiology of the common cold. J. Clin. Microbiol. 36:539–542.
    1. McCartney S.A., Thackray L.B., Gitlin L., Gilfillan S., Virgin H.W., and Colonna M.. 2008. MDA-5 recognition of a murine norovirus. PLoS Pathog. 4:e1000108 10.1371/journal.ppat.1000108
    1. McCartney S.A., Vermi W., Lonardi S., Rossini C., Otero K., Calderon B., Gilfillan S., Diamond M.S., Unanue E.R., and Colonna M.. 2011. RNA sensor-induced type I IFN prevents diabetes caused by a β cell-tropic virus in mice. J. Clin. Invest. 121:1497–1507. 10.1172/JCI44005
    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. Meagher K.L., Redman L.T., and Carlson H.A.. 2003. Development of polyphosphate parameters for use with the AMBER force field. J. Comput. Chem. 24:1016–1025. 10.1002/jcc.10262
    1. Munir S., Le Nouen C., Luongo C., Buchholz U.J., Collins P.L., and Bukreyev A.. 2008. Nonstructural proteins 1 and 2 of respiratory syncytial virus suppress maturation of human dendritic cells. J. Virol. 82:8780–8796. 10.1128/JVI.00630-08
    1. Nejentsev S., Walker N., Riches D., Egholm M., and Todd J.A.. 2009. Rare variants of IFIH1, a gene implicated in antiviral responses, protect against type 1 diabetes. Science. 324:387–389. 10.1126/science.1167728
    1. Oda H., Nakagawa K., Abe J., Awaya T., Funabiki M., Hijikata A., Nishikomori R., Funatsuka M., Ohshima Y., Sugawara Y., et al. . 2014. Aicardi-Goutières syndrome is caused by IFIH1 mutations. Am. J. Hum. Genet. 95:121–125. 10.1016/j.ajhg.2014.06.007
    1. Palmenberg A.C., Spiro D., Kuzmickas R., Wang S., Djikeng A., Rathe J.A., Fraser-Liggett C.M., and Liggett S.B.. 2009. Sequencing and analyses of all known human rhinovirus genomes reveal structure and evolution. Science. 324:55–59. 10.1126/science.1165557
    1. Pavia A.T. 2011. Viral infections of the lower respiratory tract: old viruses, new viruses, and the role of diagnosis. Clin. Infect. Dis. 52(Suppl 4):S284–S289. 10.1093/cid/cir043
    1. Pichlmair A., Schulz O., Tan C.P., Rehwinkel J., Kato H., Takeuchi O., Akira S., Way M., Schiavo G., and Reis e Sousa C.. 2009. Activation of MDA5 requires higher-order RNA structures generated during virus infection. J. Virol. 83:10761–10769. 10.1128/JVI.00770-09
    1. Poole A., Urbanek C., Eng C., Schageman J., Jacobson S., O’Connor B.P., Galanter J.M., Gignoux C.R., Roth L.A., Kumar R., et al. . 2014. Dissecting childhood asthma with nasal transcriptomics distinguishes subphenotypes of disease. J. Allergy Clin. Immunol. 133:670–8.e12. 10.1016/j.jaci.2013.11.025
    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. Retterer K., Juusola J., Cho M.T., Vitazka P., Millan F., Gibellini F., Vertino-Bell A., Smaoui N., Neidich J., Monaghan K.G., et al. . 2016. Clinical application of whole-exome sequencing across clinical indications. Genet. Med. 18:696–704. 10.1038/gim.2015.148
    1. Reynolds S.D., Rios C., Wesolowska-Andersen A., Zhuang Y., Pinter M., Happoldt C., Hill C.L., Lallier S.W., Cosgrove G.P., Solomon G.M., et al. . 2016. Airway Progenitor Clone Formation Is Enhanced by Y-27632-Dependent Changes in the Transcriptome. Am. J. Respir. Cell Mol. Biol. 55:323–336. 10.1165/rcmb.2015-0274MA
    1. Rice G.I., Forte G.M., Szynkiewicz M., Chase D.S., Aeby A., Abdel-Hamid M.S., Ackroyd S., Allcock R., Bailey K.M., Balottin U., et al. . 2013. Assessment of interferon-related biomarkers in Aicardi-Goutières syndrome associated with mutations in TREX1, RNASEH2A, RNASEH2B, RNASEH2C, SAMHD1, and ADAR: a case-control study. Lancet Neurol. 12:1159–1169. 10.1016/S1474-4422(13)70258-8
    1. Rice G.I., del Toro Duany Y., Jenkinson E.M., Forte G.M., Anderson B.H., Ariaudo G., Bader-Meunier B., Baildam E.M., Battini R., Beresford M.W., et al. . 2014. Gain-of-function mutations in IFIH1 cause a spectrum of human disease phenotypes associated with upregulated type I interferon signaling. Nat. Genet. 46:503–509. 10.1038/ng.2933
    1. Rodero M.P., and Crow Y.J.. 2016. Type I interferon-mediated monogenic autoinflammation: The type I interferonopathies, a conceptual overview. J. Exp. Med. 213:2527–2538. 10.1084/jem.20161596
    1. Rodriguez-Calvo T., and von Herrath M.G.. 2015. Enterovirus infection and type 1 diabetes: closing in on a link? Diabetes. 64:1503–1505. 10.2337/db14-1931
    1. Roth-Cross J.K., Bender S.J., and Weiss S.R.. 2008. Murine coronavirus mouse hepatitis virus is recognized by MDA5 and induces type I interferon in brain macrophages/microglia. J. Virol. 82:9829–9838. 10.1128/JVI.01199-08
    1. Rutsch F., MacDougall M., Lu C., Buers I., Mamaeva O., Nitschke Y., Rice G.I., Erlandsen H., Kehl H.G., Thiele H., et al. . 2015. A specific IFIH1 gain-of-function mutation causes Singleton-Merten syndrome. Am. J. Hum. Genet. 96:275–282. 10.1016/j.ajhg.2014.12.014
    1. Salomon-Ferrer R., Goetz A.W., Poole D., Le Grand S., and Walker R.C.. 2013. Routine microsecond molecular dynamics simulations with AMBER – Part II: Particle Mesh Ewald. J. Chem. Theory Comput. 9:3878–3888. 10.1021/ct400314y
    1. Shingai M., Ebihara T., Begum N.A., Kato A., Honma T., Matsumoto K., Saito H., Ogura H., Matsumoto M., and Seya T.. 2007. Differential type I IFN-inducing abilities of wild-type versus vaccine strains of measles virus. J. Immunol. 179:6123–6133. 10.4049/jimmunol.179.9.6123
    1. Sirén J., Imaizumi T., Sarkar D., Pietilä T., Noah D.L., Lin R., Hiscott J., Krug R.M., Fisher P.B., Julkunen I., and Matikainen S.. 2006. Retinoic acid inducible gene-I and mda-5 are involved in influenza A virus-induced expression of antiviral cytokines. Microbes Infect. 8:2013–2020. 10.1016/j.micinf.2006.02.028
    1. Slater L., Bartlett N.W., Haas J.J., Zhu J., Message S.D., Walton R.P., Sykes A., Dahdaleh S., Clarke D.L., Belvisi M.G., et al. . 2010. Co-ordinated role of TLR3, RIG-I and MDA5 in the innate response to rhinovirus in bronchial epithelium. PLoS Pathog. 6:e1001178 10.1371/journal.ppat.1001178
    1. Smyth D.J., Cooper J.D., Bailey R., Field S., Burren O., Smink L.J., Guja C., Ionescu-Tirgoviste C., Widmer B., Dunger D.B., et al. . 2006. A genome-wide association study of nonsynonymous SNPs identifies a type 1 diabetes locus in the interferon-induced helicase (IFIH1) region. Nat. Genet. 38:617–619. 10.1038/ng1800
    1. Stewart S.A., Dykxhoorn D.M., Palliser D., Mizuno H., Yu E.Y., An D.S., Sabatini D.M., Chen I.S., Hahn W.C., Sharp P.A., et al. . 2003. Lentivirus-delivered stable gene silencing by RNAi in primary cells. RNA. 9:493–501. 10.1261/rna.2192803
    1. Sulem P., Helgason H., Oddson A., Stefansson H., Gudjonsson S.A., Zink F., Hjartarson E., Sigurdsson G.T., Jonasdottir A., Jonasdottir A., et al. . 2015. Identification of a large set of rare complete human knockouts. Nat. Genet. 47:448–452. 10.1038/ng.3243
    1. Suprynowicz F.A., Upadhyay G., Krawczyk E., Kramer S.C., Hebert J.D., Liu X., Yuan H., Cheluvaraju C., Clapp P.W., Boucher R.C. Jr., et al. . 2012. Conditionally reprogrammed cells represent a stem-like state of adult epithelial cells. Proc. Natl. Acad. Sci. USA. 109:20035–20040. 10.1073/pnas.1213241109
    1. Triantafilou K., Vakakis E., Richer E.A., Evans G.L., Villiers J.P., and Triantafilou M.. 2011. Human rhinovirus recognition in non-immune cells is mediated by Toll-like receptors and MDA-5, which trigger a synergetic pro-inflammatory immune response. Virulence. 2:22–29. 10.4161/viru.2.1.13807
    1. Van Eyck L., De Somer L., Pombal D., Bornschein S., Frans G., Humblet-Baron S., Moens L., de Zegher F., Bossuyt X., Wouters C., and Liston A.. 2015. IFIH1 mutation causes systemic lupus erythematosus with selective IgA-deficiency. Arthritis Rheumatol. 67:1592–1597. 10.1002/art.39110
    1. Wang J.P., Cerny A., Asher D.R., Kurt-Jones E.A., Bronson R.T., and Finberg R.W.. 2010. MDA5 and MAVS mediate type I interferon responses to coxsackie B virus. J. Virol. 84:254–260. 10.1128/JVI.00631-09
    1. Wang Q., Nagarkar D.R., Bowman E.R., Schneider D., Gosangi B., Lei J., Zhao Y., McHenry C.L., Burgens R.V., Miller D.J., et al. . 2009. Role of double-stranded RNA pattern recognition receptors in rhinovirus-induced airway epithelial cell responses. J. Immunol. 183:6989–6997. 10.4049/jimmunol.0901386
    1. Wang Q., Miller D.J., Bowman E.R., Nagarkar D.R., Schneider D., Zhao Y., Linn M.J., Goldsmith A.M., Bentley J.K., Sajjan U.S., and Hershenson M.B.. 2011. MDA5 and TLR3 initiate pro-inflammatory signaling pathways leading to rhinovirus-induced airways inflammation and hyperresponsiveness. PLoS Pathog. 7:e1002070 10.1371/journal.ppat.1002070
    1. Wiszniewski W., Hunter J.V., Hanchard N.A., Willer J.R., Shaw C., Tian Q., Illner A., Wang X., Cheung S.W., Patel A., et al. . 2013. TM4SF20 ancestral deletion and susceptibility to a pediatric disorder of early language delay and cerebral white matter hyperintensities. Am. J. Hum. Genet. 93:197–210. 10.1016/j.ajhg.2013.05.027
    1. Wu B., Peisley A., Richards C., Yao H., Zeng X., Lin C., Chu F., Walz T., and Hur S.. 2013. Structural basis for dsRNA recognition, filament formation, and antiviral signal activation by MDA5. Cell. 152:276–289. 10.1016/j.cell.2012.11.048
    1. Yang Y., Muzny D.M., Reid J.G., Bainbridge M.N., Willis A., Ward P.A., Braxton A., Beuten J., Xia F., Niu Z., et al. . 2013. Clinical whole-exome sequencing for the diagnosis of mendelian disorders. N. Engl. J. Med. 369:1502–1511. 10.1056/NEJMoa1306555
    1. Yang Y., Muzny D.M., Xia F., Niu Z., Person R., Ding Y., Ward P., Braxton A., Wang M., Buhay C., et al. . 2014. Molecular findings among patients referred for clinical whole-exome sequencing. JAMA. 312:1870–1879. 10.1001/jama.2014.14601
    1. Züst R., Cervantes-Barragan L., Habjan M., Maier R., Neuman B.W., Ziebuhr J., Szretter K.J., Baker S.C., Barchet W., Diamond M.S., et al. . 2011. Ribose 2′-O-methylation provides a molecular signature for the distinction of self and non-self mRNA dependent on the RNA sensor Mda5. Nat. Immunol. 12:137–143. 10.1038/ni.1979

Source: PubMed

3
订阅