• Aucun résultat trouvé

General Discussion

Dans le document Sendai virus and the innate antiviral defence (Page 105-154)

The ability of the host cell to fight viral infection requires recognition of the invading viral pathogen as such and subsequent induction of signaling cascades that lead to the production of innate and adaptive antiviral responses. IFNβ plays a crucial role in priming neighbouring cells for possible infection and is also involved (with other IFN-stimulated genes) in recruiting cells of the host immune system to the site of infection, providing a way to eliminate infected cells. In the present study, we investigated the mechanisms used by the host cell to recognize SeV infection, and those used by SeV to counteract the cellular innate immune system.

By using different SeV stocks carrying mutations either in the promoter region or in the C proteins, we demonstrated that the presence of the leader RNA transcript (or the overexpression of the trailer) as well as the expression of the viral C and V proteins have an important role in counteracting the IFNβ activation and the expression of the chemokine IL-8 (c.f. Paper 1). As opposed to the V protein and the leader RNA, a number of investigations have been made on the SeV C protein. As mentioned in the first paper, this particular protein is in charge of many functions during infection: (i) it stimulates viral RNA synthesis early in infection (Latorre et al., 1998a); (ii) it inhibits viral RNA synthesis late in infection by binding to the viral polymerase (Cadd et al., 1996b); (iii) it has a role in the virus particle budding, which is facilitated by its binding to Alix, and in virion assembly, possibly by interaction with the matrix (M) protein (Mottet et al., 1996; Sakaguchi et al., 2005); (iv) The C protein interacts with Stat1 in two ways: counteracting IFN signaling and inducing Stat1 instability (Garcin et al., 2003). (v) Finally, the C protein inhibits the IRF-3-dependent activation of IFNβ as well as the activation of IL-8 expression (Strahle et al., 2003). These multiple functions of SeV C protein during infection fit well with its property to interact with various viral and cellular proteins, such as Stat1, Alix and Rig-I (unpublished). However, these interactions remain to be closely analyzed. Additionally, the reason why IL-8 expression has been only recently discovered could be explained by the fact that standard SeV induces very little IL-8. Indeed IL-8 is only expressed upon infection of SeV bearing mutations in the C and V proteins, and in the leader region. Moreover, it has been already reported that SeV and MeV infections induce an other CC-chemokine RANTES via the activation of IRF-3 (Genin et al., 2000; Tenoever et al., 2002). These data confirm the results suggesting that SeV targets the inflammatory and adaptive immune responses (IL-6 and IL-8) as well as the IFN-induced intracellular antiviral state (IFNβ and STAT1).

SeV is commonly used by many laboratories for its capacity to strongly induce IFN. This ability was long known to be associated with the presence of DI genomes in the SeV stock.

The third paper provides evidence that the strong induction of IFNβ activation upon SeV infection (SeV stock containing DI genomes) is mainly due to the presence of copyback DI genomes. Moreover, the level of IFNβ activation was found to be proportional to that of DI genome replication. Apparently, the ability of DI-H4 to induce IFNβ is not only related to its strong interference with the helper, which leads to lower levels of V and C proteins intracellularly but also to the relative content of copyback DI genomes, which can presumably form dsRNA. (cf. paper 2).

It is important to emphasize the difference between a standard SeV-WT infection from an infection of SeV that contains DI genomes. In the case of a SeV-WT infection, the RNA genomes and antigenomes are tightly assembled into NCs, meaning that they are never free to anneal (like other NNV). Furthermore, Weber et al. have shown recently that no detectable amounts of dsRNA produced by NNV have been found in infected cells (Weber et al., 2006).

This result can be explained by the fact that the amount of dsRNA produced by NNV are below their detection limit, since one molecule of dsRNA per cell can be effective in triggering an antiviral response. If it is the case, we can imagine two different ways for SeV-WT (devoided of DI genomes) to produce dsRNA: 1) During SeV replication, the polymerase ignores occasionally the stop signal at the end of the trailer template, which results in the extension of the trailer beyond the trailer/L gene junction. This event will produce a long trailer whose extended 3’sequences can anneal to those of the L gene (Vidal and Kolakofsky, 1989). 2) dsRNA can also be the result of a readthrough by the transcriptase of the L gene-end site until the end of the genome (an event that can happen at 5% of the time at all gene junctions). In this case the genome 5’end can anneal to the trailer RNAs (Le et al., 2002).

Finally, if we consider that there is a small amount of dsRNA that is generated in SeV infection, it is evident that the presence of dsRNA will be dampen by the SeV C and V proteins expressed during ND genome replication and that the expression of IFNβ will consequently be blocked.

In the case of copyback DI infections, it is a different story. Because DI-H4 contains two antigenomic promoters located at both the genomes and antigenomes ends, it has strong replicative advantage over the standard virus. Consequently, a huge amount of DI genomes is rapidly generated and the level of viral proteins (including the level of the nucleocapsid protein) is also strongly attenuated. This rapid generation of DI genomes and the low level of

NC proteins could explain why the genomes and antigenomes may not assemble into nucleocapsids and would consequently form dsRNA. If DI-H4 genomes and antigenomes are not encapsidated during infection, two forms of dsRNA can be found (in addition to those mentioned for a SeV infection). First, the extremities of the DI genome (complementary on 110 nucleotides) can self anneal in a concentration independent manner and form dsRNA, resulting in the shape of panhandles. Secondly, the genome and the antigenome can anneal together, which remains unlikely since this would require the melting of intramolecular secondary structure.

Recent evidence suggests that NNV infections could activate the innate immune response by producing viral signatures other than dsRNA, namely 5’tri-phosphorylated single-stranded RNA transcripts (5’pppRNAs). This discovery suggests that dsRNA may not always be implicated in the IFNβ induction upon viral infection and that there is probably more than one factor able to induce this primary cellular induction. Two cytoplasmic helicases, namely RIG-I and Mda-5, have been recently found to respond to dsRNA and, at least for RIG-I, to 5’pppRNAs. These RNAs are generated in the cytoplasm during RNA virus replication. Upon detection of specific viral RNAs RIG-I and Mda-5 interact with Cardif which is present in the mitochondrial membrane, and this interaction is thought to lead to the recruitment and activation of the TBK1, IKKε and other IKK kinases that activate NF-kB and IRF3, thereby activating the IFNβ promoter.

In the third paper, we investigated different potential inducers of IFNβ, including the synthetic dsRNA (polyI/C) and 5’pppRNAs and the involvement of RIG-I in these inductions.

We showed evidences that i) transfection of poly-I/C and ii) a SeV coinfection that can produce GFP dsRNA, as well as iii) transfections of small 5’pppRNAs can all induce IFNβ.

In all cases, the IFNβ induction was dependent on RIG-I. Not all 5’pppRNAs generated by SeV can act as PAMPs: 1) SeV mRNAs carry 5’ ppp ends, but because they are capped, these mRNAs cannot be potential targets. 2) SeV genomes and antigenomes also carry triphosphorylated 5’ends, but because the RNA is tightly and entirely encapsidated, it cannot be seen. 3) Finally, leader and trailer RNA transcript are the most susceptible to induce IFNβ, because they are short unencapsidated 5’pppRNAs that are independent on on-going (N) protein synthesis. For this reason, we suspect that, in the case of DI-H4 infection, there is a real overproduction of pppRNAs (trailer) resulting in the induction of the IFNβ activation.

Alternatively, other viral structures may take the role of RNA as a danger signal for the host cell. Indeed, it was previously shown that the NCs from VSV and MeV are capable of

triggering IFN induction. It was interesting to observe in our experiments that NCs purified from SeV infection could strongly induce IFNβ. However, RNase treatment revealed that it was probably not the NCs that were responsible for this activation, but more likely RNA products that have contaminated the CsCl banded NCs. Because the NC is the first viral element that enters the cell, one can imagine that viruses have evolved to avoid its detection.

Thus, it is reasonable to think that NCs do not act as PAMPs. Concerning MeV, it was shown recently that the NCs were not responsible for the IFNβ activation after all and that it was the leader transcript of MeV that was responsible for the activation of IFNβ via the RIG-I pathway (Helin et al., 2001; Plumet et al., 2007). This again comforts the idea that SeV leader and trailer RNAs are more likely to be the main PAMPs responsible for triggering the antiviral responses upon SeV infection.

Regarding the C and V proteins, we presented evidence that for SeV infection of MEFs, it is the C protein (and not V) that is primarily responsible for this effect, and that C acts by countering RIG-I dependent signaling to IFNβ. For example, independent expression of either the C or the V proteins inhibited IFNβ activation due to RIG-I over-expression. In addition to this, both proteins inhibited IFNβ activation due to DI-H4 infection, although C was always more effective than V (in our experiments). However, only C expression effectively inhibited IFNβ activation due to GFP(+/-) infection, or transfected poly-I/C or pppRNAs (cf paper 3).

These new findings, suggesting that the 5’pppRNA is a PAMP specific to the NNV infections, lead us to reconsider the way the cell could detect the presence of viruses.

However, no proof has been found concerning the role of 5’pppRNA products in the activation of IFNβ in SeV infection. Further analyses of the precise RNAs involved in the IFNβ activation need to be done. Additionally, more investigations are necessary to determine how these PAMPs are detected and the mechanisms that drive cellular responses. Finally, learning more about the protein targets of SeV proteins also appear important in order to understand the details of the IFN suppressive activities of SeV and virus infection in general.

VIII.

Reference List

Ali,A. and Nayak,D.P. (2000). Assembly of Sendai virus: M protein interacts with F and HN proteins and with the cytoplasmic tail and transmembrane domain of F protein. Virology 276, 289-303.

Andrejeva,J., Childs,K.S., Young,D.F., Carlos,T.S., Stock,N., Goodbourn,S., and Randall,R.E. (2004). The V proteins of paramyxoviruses bind the IFN-inducible RNA helicase, mda-5, and inhibit its activation of the IFN-beta promoter. Proc. Natl. Acad. Sci. U.

S. A 101, 17264-17269.

Baltimore,D. and Huang,A. (1975). Proceedings: Defective interfering particles. Neurology 25, 492.

Baron,M.D. and Barrett,T. (1997). Rescue of rinderpest virus from cloned cDNA. J Virol 71, 1265-1271.

Barrett,A.D. and Dimmock,N.J. (1986). Defective interfering viruses and infections of animals. Curr. Top. Microbiol. Immunol. 128, 55-84.

Basler,C.F. and Garcia-Sastre,A. (2007). Sensing RNA virus infections. Nat. Chem. Biol. 3, 20-21.

Basler,C.F., Mikulasova,A., Martinez-Sobrido,L., Paragas,J., Muhlberger,E., Bray,M., Klenk,H.D., Palese,P., and Garcia-Sastre,A. (2003). The Ebola virus VP35 protein inhibits activation of interferon regulatory factor 3. J Virol 77, 7945-7956.

Baudin,F., Bach,C., Cusack,S., and Ruigrok,R.W.H. (1994). Structure of influenza virus RNP. I. Influenza virus nucleoprotein melts secondary structure in panandle RNA and exposes the bases to the solvent. EMBO J. 13, 3158-3165.

Bontron,S., Ucla,C., Mach,B., and Steimle,V. (1997). Efficient repression of endogenous major histocompatibility complex class II expression through dominant negative CIITA mutants isolated by a functional selection strategy. Mol. Cell Biol. 17, 4249-4258.

Bowie,A.G. and Fitzgerald,K.A. (2007). RIG-I: tri-ing to discriminate between self and non-self RNA. Trends Immunol. 28, 147-150.

Brzozka,K., Finke,S., and Conzelmann,K.K. (2005). Identification of the rabies virus alpha/beta interferon antagonist: phosphoprotein P interferes with phosphorylation of interferon regulatory factor 3. J Virol 79, 7673-7681.

Buchholz,C.J., Spehner,D., Drillien,R., Neubert,W.J., and Homann,H.E. (1993). The conserved N-terminal region of Sendai virus nucleocapsid protein NP is required for nucleocapsid assembly. J Virol 67, 5803-5812.

Buchholz,U.J., Finke,S., and Conzelmann,K.K. (1999). Generation of bovine respiratory syncytial virus (BRSV) from cDNA: BRSV NS2 is not essential for virus replication in tissue

culture, and the human RSV leader region acts as a functional BRSV genome promoter. J Virol 73, 251-259.

Cadd,T., Garcin,D., Tapparel,C., Itoh,M., Homma,M., Roux,L., Curran,J., and Kolakofsky,D.

(1996b). The Sendai paramyxovirus accessory C proteins inhibit viral genome amplification in a promoter-specific fashion. J Virol 70, 5067-5074.

Cadd,T., Garcin,D., Tapparel,C., Itoh,M., Homma,M., Roux,L., Curran,J., and Kolakofsky,D.

(1996a). The Sendai paramyxovirus accessory C proteins inhibit viral genome amplification in a promoter-specific fashion. J Virol 70, 5067-5074.

Calain,P., Curran,J., Kolakofsky,D., and Roux,L. (1992). Molecular cloning of natural paramyxovirus copy-back defective interfering RNAs and their expression from DNA.

Virology 191, 62-71.

Calain,P. and Roux,L. (1993). The rule of six, a basic feature for efficient replication of Sendai virus defective interfering RNA. J Virol 67, 4822-4830.

Calain,P. and Roux,L. (1995). Functional characterisation of the genomic and antigenomic promoter of the Sendai virus. Virology 211, 163-173.

Casola,A., Burger,N., Liu,T., Jamaluddin,M., Brasier,A.R., and Garofalo,R.P. (2001).

Oxidant tone regulates RANTES gene expression in airway epithelial cells infected with respiratory syncytial virus. Role in viral-induced interferon regulatory factor activation. J.

Biol. Chem. 276, 19715-19722.

Chen,Z., Li,Y., and Krug,R.M. (1999). Influenza A virus NS1 protein targets poly(A)-binding protein II of the cellular 3'-end processing machinery. EMBO J. 18, 2273-2283.

Childs,K., Stock,N., Ross,C., Andrejeva,J., Hilton,L., Skinner,M., Randall,R., and Goodbourn,S. (2007). mda-5, but not RIG-I, is a common target for paramyxovirus V proteins. Virology 359, 190-200.

Choppin,P.W., Scheid,A., and Mountcastle,W.E. (1975). Proceedings: Paramyxoviruses, membranes, and persistent infections. Neurology 25, 494.

Collins,P.L., Mink,M.A., and Stec,D.S. (1991). Rescue of synthetic analogs of respiratory syncytial genomic RNA and effect of truncations and mutations on the expression of a foreign reporter gene. Proc. Natl. Acad. Sci. U. S. A. 88, 9663-9667.

Conzelmann,K.-K. and Schnell,M.J. (1994). Rescue of Synthetic Genomic RNA Analogs of Rabies Virus by Plasmid-Encoded Proteins. J Virol 68, 713-719.

Creagh,E.M. and O'Neill,L.A. (2006). TLRs, NLRs and RLRs: a trinity of pathogen sensors that co-operate in innate immunity. Trends Immunol. 27, 352-357.

Curran,J. and Kolakofsky,D. (1990). Sendai virus P gene produces multiple proteins from overlapping open reading frames. Enzyme 44, 244-249.

Curran,J. and Kolakofsky,D. (1988). Ribosomal initiation from an ACG codon in the Sendai virus P/C mRNA. EMBO J. 7, 245-251.

Curran,J., Marq,J.-B., and Kolakofsky,D. (1995). An N-terminal domain of the Sendai paramyxovirus P protein actsw as a chaperone for the NP protein during the nascent chain assembly step of genome replication. J Virol 69, 849-855.

Curran,J.A. and Kolakofsky,D. (1991). Rescue of a Sendai virus DI genome by other parainfluenza viruses: implications for genome replication. Virology 182, 168-176.

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.

Das,S.C. and Pattnaik,A.K. (2005). Role of the hypervariable hinge region of phosphoprotein P of vesicular stomatitis virus in viral RNA synthesis and assembly of infectious virus

particles. J Virol 79, 8101-8112.

Decker,T., Stockinger,S., Karaghiosoff,M., Muller,M., and Kovarik,P. (2002). IFNs and STATs in innate immunity to microorganisms. J. Clin. Invest 109, 1271-1277.

Deshpande,K.L. and Portner,A. (1984). Structural and functional analysis of Sendai virus nucleocapsid protein NP with monoclonal antibodies. Virology 139, 32-42.

Donelan,N.R., Basler,C.F., and Garcia-Sastre,A. (2003). A recombinant influenza A virus expressing an RNA-binding-defective NS1 protein induces high levels of beta interferon and is attenuated in mice. J Virol 77, 13257-13266.

Enami,M. and Palese,P. (1991). High-efficiency formation of influenza virus transfectants. J Virol 65, 2711-2713.

Errington,W. and Emmerson,P.T. (1997). Assembly of recombinant Newcastle disease virus nucleocapsid protein into nucleocapsid-like structures is inhibited by the phosphoprotein. J.

Gen. Virol. 78 ( Pt 9), 2335-2339.

Fischer,S.F., Ludwig,H., Holzapfel,J., Kvansakul,M., Chen,L., Huang,D.C., Sutter,G., Knese,M., and Hacker,G. (2006). Modified vaccinia virus Ankara protein F1L is a novel BH3-domain-binding protein and acts together with the early viral protein E3L to block virus-associated apoptosis. Cell Death. Differ. 13, 109-118.

Fitzgerald,K.A., Rowe,D.C., Barnes,B.J., Caffrey,D.R., Visintin,A., Latz,E., Monks,B., Pitha,P.M., and Golenbock,D.T. (2003). LPS-TLR4 signaling to IRF-3/7 and NF-kappaB involves the toll adapters TRAM and TRIF. J. Exp. Med. 198, 1043-1055.

Fortes,P., Beloso,A., and Ortin,J. (1994). Influenza virus NS1 protein inhibits pre-mRNA splicing and blocks mRNA nucleocytoplasmic transport. EMBO J. 13, 704-712.

Fouillot-Coriou,N. and Roux,L. (2000). Structure-function analysis of the Sendai virus F and HN cytoplasmic domain: different role for the two proteins in the production of virus particle.

Virology 270, 464-475.

Fuerst,T.R., Niles,E.G., Studier,F.W., and Moss,B. (1986). Eukaryotic transient-expression system based on recombinant vaccinia virus that synthesizes bacteriophage T7 RNA polymerase. Proc. Natl. Acad. Sci. U. S. A 83, 8122-8126.

Fujii,Y., Sakaguchi,T., Kiyotani,K., Huang,C., Fukuhara,N., Egi,Y., and Yoshida,T. (2002).

Involvement of the leader sequence in Sendai virus pathogenesis revealed by recovery of a pathogenic field isolate from cDNA. J Virol 76, 8540-8547.

Garcin,D., Marq,J.B., Goodbourn,S., and Kolakofsky,D. (2003). The amino-terminal

extensions of the longer Sendai virus C proteins modulate pY701-Stat1 and bulk Stat1 levels independently of interferon signaling. J Virol 77, 2321-2329.

Garcin,D., Marq,J.B., Strahle,L., Le,M.P., and Kolakofsky,D. (2002). All four Sendai Virus C proteins bind Stat1, but only the larger forms also induce its mono-ubiquitination and

degradation. Virology 295, 256-265.

Garcin,D., Pelet,T., Calain,P., Roux,L., Curran,J., and Kolakofsky,D. (1995). A highly recombinogenic system for the recovery of infectious Sendai paramyxovirus from cDNA:

generation of a novel copy-back nondefective interfering virus. EMBO J. 14, 6087-6094.

Gassen,U., Collins,F.M., Duprex,W.P., and Rima,B.K. (2000). Establishment of a rescue system for canine distemper virus. J Virol 74, 10737-10744.

Genin,P., Algarte,M., Roof,P., Lin,R., and Hiscott,J. (2000). Regulation of RANTES

chemokine gene expression requires cooperativity between NF-kappa B and IFN-regulatory factor transcription factors. J. Immunol. 164, 5352-5361.

Gershon,P.D., Ahn,B.Y., Garfield,M., and Moss,B. (1991). Poly(A) polymerase and a

dissociable polyadenylation stimulatory factor encoded by vaccinia virus. Cell 66, 1269-1278.

Gotoh,B., Takeuchi,K., Komatsu,T., and Yokoo,J. (2003). The STAT2 activation process is a crucial target of Sendai virus C protein for the blockade of alpha interferon signaling. J Virol 77, 3360-3370.

Gupta,K.C. and Patwardhan,S. (1988). ACG, the initiator codon for a Sendai virus protein. J.

Biol. Chem. 263, 8553-8556.

Haller,O., Kochs,G., and Weber,F. (2006). The interferon response circuit: induction and suppression by pathogenic viruses. Virology 344, 119-130.

Haller,O., Staeheli,P., and Kochs,G. (2007). Interferon-induced Mx proteins in antiviral host defense. Biochimie 89, 812-818.

Hatada,E. and Fukuda,R. (1992). Binding of influenza A virus NS1 protein to dsRNA in vitro.

J. Gen. Virol. 73 ( Pt 12), 3325-3329.

Hausmann,S., Garcin,D., Delenda,C., and Kolakofsky,D. (1999). The versatility of paramyxovirus RNA polymerase stuttering. J Virol 73, 5568-5576.

Helin,E., Vainionpaa,R., Hyypia,T., Julkunen,I., and Matikainen,S. (2001). Measles virus activates NF-kappa B and STAT transcription factors and production of IFN-alpha/beta and IL-6 in the human lung epithelial cell line A549. Virology 290, 1-10.

Hiscott,J., Pitha,P., Genin,P., Nguyen,H., Heylbroeck,C., Mamane,Y., Algarte,M., and Lin,R.

(1999). Triggering the interferon response: the role of IRF-3 transcription factor. J. Interferon Cytokine Res. 19, 1-13.

Holland,J.J., Grabau,E.A., Jones,C.L., and Semler,B.L. (1979). Evolution of multiple genome mutations during long-term persistent infection by vesicular stomatitis virus. Cell 16, 495-504.

Holland,J.J., Kennedy,I.T., Semler,B.L., Jones,C.L., Roux,L., and Grabau,E.A. (1980).

Defective interfering RNA viruses and the host-cell response. In Comprehensive Virology, H.Fraenkel-Conrat and R.R.Wagner, eds. (New York and London: Plenum Press), pp. 137-192.

Horikami,S.M., Hector,R.E., Smallwood,S., and Moyer,S.A. (1997). The Sendai virus C protein binds the L polymerase protein to inhibit viral RNA synthesis. Virology 235, 261-270.

Horikami,S.M. and Moyer,S.A. (1982). Host range mutants of vesicular stomatitis virus defective in in vitro RNA methylation. Proc. Natl. Acad. Sci. U. S. A 79, 7694-7698.

Hornung,V., Ellegast,J., Kim,S., Brzozka,K., Jung,A., Kato,H., Poeck,H., Akira,S.,

Conzelmann,K.K., Schlee,M., Endres,S., and Hartmann,G. (2006). 5'-Triphosphate RNA Is the Ligand for RIG-I. Science.

Horodyski,F.M. and Holland,J.J. (1980). Viruses isolated from cells persistently infected with vesicular stomatitis virus show altered interactions with defective interfering particles. J Virol 36, 627-631.

Horvath,C.M. (2004). Weapons of STAT destruction. Interferon evasion by paramyxovirus V protein. Eur. J. Biochem. 271, 4621-4628.

Houben,K., Marion,D., Tarbouriech,N., Ruigrok,R.W., and Blanchard,L. (2007).

INTERACTION OF THE C-TERMINAL DOMAINS OF SENDAI VIRUS N AND P PROTEINS: Comparison of polymerase-nucleocapsid interaction within the paramyxovirus family. J Virol.

Huang,A.S. (1977). Viral pathogenesis and molecular biology. Bacteriol. Rev. 41, 811-821.

Huang,A.S. and Baltimore,D. (1970). Defective viral particles and viral disease processes.

Huang,A.S. and Baltimore,D. (1970). Defective viral particles and viral disease processes.

Dans le document Sendai virus and the innate antiviral defence (Page 105-154)