• Non ci sono risultati.

Restriction factors against human CMV

N/A
N/A
Protected

Academic year: 2021

Condividi "Restriction factors against human CMV"

Copied!
32
0
0

Testo completo

(1)

Restriction Factors against Human Cytomegalovirus

Santo Landolfo1*, Marco De Andrea1,2, Marisa Gariglio2

1Viral Pathogenesis Unit, Department of Public Health and Pediatric Sciences, Medical School, University of Turin, Italy.

2Virology Unit, Department of Translational Medicine, Medical School of Novara, Italy

*Author for correspondence: Viral Pathogenesis Unit,Department of Public Health and Pediatric Sciences, Medical School, University of Turin. Via Santena, 9 – 10126 Turin, Italy - Tel.: +39 011 670 5636, Fax: +30 011 670 5648, email: santo.landolfo@unito.it

Keywords: Human Cytomegalovirus (HCMV), restriction factor (RF), Nuclear Domain 10 (ND10), Interferon, IFI16, Viperin.

(2)

Summary

Cellular proteins called “restriction factors” (RFs) form an important component of the innate immune response to viral replication. However, over time viruses have learned how to antagonize RFs through mechanisms that are specific for each virus. Here, we summarize the general hallmarks of RFs before going on to discuss the specific strategies recruited by some key RFs that strive to hold Human Cytomegalovirus (HCMV) infection back, as well as the counter-restriction mechanisms employed by the virus to overcome this innate defense. Such RFs include the cellular proteins PML, hDaxx, and Sp100 (constituents of Nuclear Domain 10: ND10) and IFI16, a nuclear protein and member of the PYHIN protein family. Viral regulatory proteins, such as IE1 or pp71, try to oppose the ND10-induced blockade of virus replication by either modifying or disrupting this RF. IFI16, on the other hand, inhibits virus DNA synthesis by down-regulating the transcription of viral gene UL54; the intruding virus attempts to antagonize IFI16 by mislocalizing it from the nucleus to the cytoplasm via the action of viral protein UL97. Finally, we consider how Viperin, a RF initially thought to inhibit HCMV maturation late during infection, has actually been demonstrated to enhance virus maturation by increasing lipid metabolism and enhancing virus envelopment. All in all, detailed knowledge of these HCMV restriction and counter-restriction mechanisms will provide new insights into how we might control HCMV infection and, in turn, contribute to the development of novel therapeutic interventions.

(3)

Introduction

The mammalian immune system has developed an arsenal of defense mechanisms, including innate and adaptive immune responses, primarily responsible for protecting cells against viral pathogens [1]. Classical innate immunity against viruses is mediated by specialized cells, including natural killer cells, dendritic cells, and macrophages. However, recent work has also discovered the existence of cellular-based defense mechanisms, now known as intrinsic immunity, giving rise to a third branch of the traditionally bipartite immune system [2]. The cellular proteins directing this form of antiviral activity are often termed “restriction factors” (RFs), as they inhibit viral replication [3]. However, viruses have themselves evolved mechanisms to evade and defend themselves against this form of host intrinsic immunity [4,5]. Indeed, the defense and counter-defense measures generating this antagonistic conflict have molded both viral and host cell functions and genomes. The term “restriction factor” was historically adopted by laboratories studying retroviruses following the observation that the mouse Friend virus susceptibility factor-1 (Fv1 locus) conferred resistance to murine retroviruses [6,7]. In the mouse genome, there are at least two Fv1 alleles (Fv1n, Fv1b) conferring resistance to B-tropic murine leukemia virus (B-MLV) and N-B-tropic MLV (N-MLV) infection. The B-MLV strains efficiently infect Fv1b/b homozygous BALB/c mice, but not the Fv1n/n homozygous NIH/Swiss mice, whereas the N-MLV strains exhibit the opposite tropism [6]. Fv1 blocks the nuclear import of reverse transcribed retroviral pre-integration complex [8], but the precise mechanism is still unknown. Retroviruses have come to represent a model system that has been instrumental in expanding our understanding of how viruses and hosts interact. Indeed, the concept of “intrinsic immunity” arose from the discovery that prototype human antiretroviral restriction factors, such as APOBEC3G, TRIM5, and Tetherin, are constitutively expressed in cell types targeted by primate lentiviruses, where target cells are resistant to infection in the apparent absence of any signaling event [9–14]. Although it is now apparent that the expression of most restriction factors can be increased by interferon (IFN),

(4)

IFN-induction is not required for constitutive antiretroviral activity in target cells. Thus, there is a clear and critical biological distinction between intrinsic (pre-existing) immunity and IFN-dependent innate immunity.

Innate Immunity

Typically, in the first step of the innate immune response, viral nucleic acids and proteins are recognized as pathogen-associated molecular patterns (PAMPs), hallmarks of infection usually absent from healthy host cells. Endosomal Toll-like receptors (TLR3, TLR7/8 and TLR9) and cytosolic pattern recognition receptors (PRRs) (reviewed by [15,16]) can detect viral DNA and RNA species, including double-stranded (ds) RNA, single-stranded (ss) RNA, and DNA containing CpG motifs [17–19]. A variety of cell types including macrophages, dendritic cells and keratinocytes can detect PAMPs via TLRs located in the endosomes. However, it has recently emerged that many cells can also detect nucleic acids in their own cytosol during viral infection, and this occurs in a TLR-independent manner. The cytosolic RNA receptors MDA5 (Melanoma Differentiation-Associated factor 5) and RIG-I (Retinoic acid-Inducible Gene 1) detect dsRNA and ssRNA with a 5’-triphosphate group in the cytoplasm of cells infected with RNA viruses (reviewed by [20]).

Pathogen DNA constitutes one of the most important PAMPs signaling infection by intracellular pathogens, such as bacteria and viruses [21–23]. The predominant sensor of cytosolic DNA is the IFN-inducible PYHIN family protein AIM2 (Absent in Melanoma 2), which binds DNA via its HIN domain and then activates multiple signaling cascades leading to inflammasome complex formation and IFN-b production [24–28]. While AIM2 acts as a cytoplasmic sensor, another member of the PYHIN family, designated IFI16, senses pathogenic DNA both in the cytoplasm and the nucleus [29–34]. The detection of exogenous DNA by IFI16 then initiates two distinct innate immunity signaling cascades: the first involves the activation of transcription factors Interferon Regulatory Factor 3 (IRF3) and nuclear factor kappa B (NF-B), which result in the transcription of IFN-, chemokines, and

(5)

pro-inflammatory cytokines [32]. The second signaling pathway results in the formation of an inflammasome complex, which activates caspase-1, a protease that processes pro-interleukin (IL)-1 and pro-IL-18 to generate active cytokines ready for secretion [35,36]. Thus, viral infections are detected by cytoplasmic, membrane-bound or nuclear PRRs, which trigger an response that induces the expression of IFN-stimulated genes, many of which are RFs that specifically inhibit viral growth within infected cells.

General Restriction Factor Characteristics

Restriction Factors are germline-encoded factors that mediate a cell-intrinsic immune response. They can also be classified as PRRs because they directly recognize and bind to viral components [3,37]. However, unlike TLRs and RLRs, which activate signaling cascades that inhibit virus infection indirectly, RFs inhibit viral replication by directly interfering with the activity of essential viral genes, often before the activation of the IFN response. Thus, RFs are characterized by properties that differentiate them from pathogen sensors. Briefly, RFs: i) autonomously exhibit antiviral activity in culture-based assays and inhibit specific processes in viral replication; ii) are constitutively expressed in various cell types, but can be induced by IFNs; iii) are often antagonized by viral proteins; and iv) are subject to positive genetic selection driven by host-pathogen coevolution. Indeed, the notion of “intrinsic immunity” emerges from the early observation that prototype human antiretroviral factors, such as APOBEC3G and TRIM5, were found constitutively expressed in cell types targeted by primate lentiviruses in the complete absence of any signaling event [2,38]. Thus, the concept of intrinsic immunity served as a basis to explain why defined cell lines are “restrictive” or “permissive”, depending on whether wild-type or mutant viruses could efficiently replicate therein. Although intrinsic immune mechanisms were discovered as being active against retroviruses [3,39,40], increasing evidence suggests that such mechanisms also counteract several other viruses, including Herpesviruses [41–43]. In present review, we therefore concentrate on the most recent discoveries regarding human cytomegalovirus

(6)

(HCMV) and describe how the actions of different restriction factors affect different steps of its replication. Interestingly, evidence also implicates HCMV as having evolved mechanisms to counteract these restriction factors, which ultimately leads to the successfully replication of this virus in target cells.

Nuclear Domain 10 (ND10)

Nuclear domain 10 (ND10; also known as PML nuclear bodies or PODS) are discrete bodies of accumulated proteins located within the nuclear matrix, an ill-defined structure within the nucleus proposed to anchor and regulate various nuclear functions [44,45]. PML protein (a member of the TRIM/RBCC family of proteins), hDaxx, and Sp100 are the key constituents of ND10 and they organize the nuclear matrix by recruiting proteins characterized by their ability to be SUMOylated [46,47]. The recruited proteins include an ubiquitin-like protein named SUMO, and the conjugation of PML to SUMO plays a critical role in the recruitment of other partners [48–50]. The ND10 body is a sphere with a diameter ranging between 0.1-1 m, and which may present a granular center. These proteinaceous bodies are typically found in clusters of 5 to 15 and do not contain RNA or DNA [51]. PML protein is located in the outer shell of the structure, with its associated partners residing inside. ND10 are distributed within the interchromosomal space, frequently adjacent to other bodies. Functionally, ND10 have been implicated in the regulation of diverse cellular key processes, like oncogenesis [52–54], DNA damage repair [55], apoptosis [56–58], senescence [59–61], and the regulation of gene expression [44,49]. Interestingly, ND10 undergo profound modification during virus infection [55,62–65]. They appear to accumulate viral genome at their periphery or within their central core during infection of quiescent cells. Together with the knowledge that IFNs increase the gene transcription of PML, hDaxx, and Sp100 [66,67], these findings point to a role of ND10 in the regulation of the antiviral response.

(7)

Initially regarded as the optimal cellular site for the initiation of HCMV IE gene expression, whose transcripts are juxtaposed to these domains, it soon became apparent that the virus targets ND10 early during infection in order to destroy them [68–72]. This activity, executed by the viral regulatory protein IE1, correlated with efficient lytic replication cycle. Important advances in the comprehension of the biological role of ND10 in the prevention of viral replication came from the finding that depletion of PML from ND10 bodies by specific small interfering RNA (siRNA) rendered human primary foreskin fibroblasts (HFF) more susceptible to HCMV infection [73]. Together, these observations support the notion that ND10 bodies regulate an intrinsic immune response of the cell that exerts its antiviral activity by down-regulating viral IE gene expression. Interestingly, HCMV infection of PML-null HFF induced de novo formation of two other protein constituents of ND10 bodies, namely hDaxx and Sp100 [73]; hence, it emerged the concept that HCMV triggers an active recruitment of ND10 components to the site of viral nucleoprotein complex. Moreover, it has been suggested that other ND10 associated proteins, like hDaxx and Sp100, could indeed form part of the intrinsic antiviral response instituted by ND10 bodies [74,75].

The scenario that emerges from studies performed by different group shows that the three major ND10 proteins – PML, hDaxx, and Sp100 – repress viral replication by silencing viral IE gene expression [72,74–77]. Furthermore, in infection experiments exploiting all possible double knock-down combinations of the ND10 constituent proteins, the initiation of HCMV gene expression is enhanced compared to that in the respective single-knock down cells [75]. These finding clearly argue for an independent role of these factors in the suppression of HCMV replication. Furthermore, some authors predict that epigenetic mechanisms are involved, capable of preventing viral genome transcription and lytic replication [78,79]. During early infection, viral DNA has been found to exist in a repressive chromatin state, modulated by posttranslational histone modification [80]. The ND10 protein responsible for inducing a transcriptionally inactive chromatin state of the

(8)

immediate-early enhancer/promoter (MIEP) has been identified as being hDaxx [81,82]. The state is induced via recruitment of the chromatin remodeling protein ATRX (alpha thalassemia/mental retardation syndrome X-linked) or of chromatin modifying enzymes, like histone deacetylases (HDACs), to the viral DNA [79,83]. Similarly, PML and Sp100 have also been demonstrated to interact with chromatin modifying enzymes, suggesting the involvement of epigenetic mechanisms in the repression of viral IE gene expression [71,84].

Despite the restriction activity orchestrated by the ND10 constituent proteins, the observation that HCMV is still able to replicate successfully indicates that the virus has co-evolved to counteract such activity. Soon after virus entry, the viral tegument protein and transactivator pp71 translocates and interacts with ND10 constituents [85]. The viral pp71 binds to hDaxx protein, which is responsible for silencing the MIEP and down-regulating IE gene expression [86]. Following pp71 binding, hDaxx undergoes proteasome degradation and relieves MIEP repression. However, recent studies by different groups have demonstrated that the scenario is more complex. Degradation of hDaxx is preceded by the release of the chromatin-remodeling protein ATRX from ND10, which is stimulated by pp71. As a consequence, the displacement of ATRX alleviates the repression of viral IE gene expression [83].

While pp71 appears to be responsible for counteracting both hDaxx- and ATRX-mediated gene silencing, the IE1 protein appears to abolish the repressive effects of PML. The IE1 protein synergizes with IE2 to promote the stimulation of several viral early and host gene promoters [87–90]. It has since been demonstrated that IE1, but not IE2, is able to stimulate the dispersal of the ND10 bodies with the displacement of their constituents proteins, such as PML and Sp100. IE1-mediated disruption of PML does not proceed via proteasomal degradation, but rather via de-SUMOylation that then leads to the inhibition of PML oligomerization [91].

In conclusion, it appears that the frontline defense against HCMV replication occurring soon after infection is accomplished by ND10 proteins. Moreover, it is now

(9)

well established that viral proteins, including pp71 and IE1 (FIGURE 1), are

responsible for the viral countermeasures put into effect in order to overwhelm these host restriction factors. Present research efforts are being directed at gaining a more detailed comprehension of the molecular mechanisms underlying the virus-host interactions mediated by ND10.

The PYHIN-200 family IFI16 protein

The IFN-inducible PYHIN-200 protein family includes the structurally related human AIM2, IFI16, IFIX, and MNDA [92]; murine homologues include, for example, p202, p203, and p204. All the proteins in the family share a conserved domain of 200-amino acid residues at the C-terminus, present in single or tandem copies. Three IFI16 protein isoforms exist, derived from alternative splicing [93]; all of which contain two 200-amino acid domains, A and B, separated by a spacer region of variable length. Most proteins of this family also contain a homotypic protein-protein interaction PYRIN domain (PYD) in the N-terminus [94]. The predominant B isoform can be detected in cells of different histological type, such as immune cells, endothelial cells, and epithelial cells [95,96]. The N-terminus of IFI16 protein contains a bi-partite nuclear localization signal (NLS). Accordingly, the subcellular localization of IFI16 is normally nuclear in resting cells, including: fibroblasts [29], endothelial cells, and keratinocytes (reviewed in [95,97]). Notably, within the nuclear compartment, IFI16 protein is also detected in the nucleolus [98]. Upon infection of macrophages or fibroblasts by Herpesvirus [34,99], or the exposure of keratinocytes to UV-B light, IFI16 is redistributed from the nucleus to the cytoplasm [100]. This redistribution of IFI16 protein was associated to inflammasome assembly in the case of Herpesvirus infection, and apoptosis in the case of UV-B exposure. Protein-protein interactions have been recognized as influencing the subcellular localization of proteins, and IFI16 protein can form homo- or heterodimers with several other proteins, including: p53, Rb, BRCA1, ASC, and STING [30,31,101–104]. However, the molecular mechanisms that regulate IFI16 redistribution between the nuclear and

(10)

the cytoplasmic compartment remain unknown. Finally, IFI16 has been shown to bind to DNA both in vitro and in vivo [30,105–107]. Consistent with this DNA binding ability, IFI16 has also been recognized to act as a viral DNA sensor in the cytoplasm and to play an important role in tuning the innate immune response by enhancing IFN-production [34,97]. It therefore seems conceivable that IFI16 redistribution within the cell depends, not only on the type of protein interactions occurring, but also the binding of IFI16 protein to pathogenic DNA.

Restriction of HCMV replication by IFI16 protein and virus evasion

Two members of the PYHIN family, namely AIM2 and IFI16, have been shown to function as PRRs of virus-derived intracellular DNA [28,31,35,108,109]. In particular, IFI16 has been demonstrated to interact with the adaptor molecule ASC and procaspase-1 to form a functional inflammasome during Kaposi Sarcoma Associated Herpesvirus (KSVH) infection [31]. Moreover, reducing the expression of IFI16, or its mouse ortholog p204, by means of RNA-mediated interference inhibited gene induction and the activation of transcription factors IRF3 and NF-κB normally seen after DNA transfection and herpes simplex virus type 1 (HSV-1) infection [32,110,111]. In addition its role as a PRR, many different functions had been ascribed to IFI16 protein, with the exception of any form of antiviral activity (reviewed in [92]). This view of IFI16 has changed over recent years thanks to the findings generated from two different experimental approaches. The first involved the silencing of IFI16 using specific siRNA or its inactivation by introducing a lentivirus expressing a dominant negative mutant form [109]. In both cases, the lack of functional IFI16 protein in human embryo lung fibroblasts (HELF) significantly enhanced herpesvirus replication, including that of HCMV. Consistent with these results, when HELF overexpressing IFI16 were infected with HCMV, a 2.5 log decrease in viral production was observed. To investigate the molecular basis of the antiviral activity of IFI16, the effects of IFI16 overexpression on the different phases of the HCMV replication cycle were examined. Transfection experiments with the

(11)

luciferase reporter gene driven by deleted or site-specific mutated forms of the HCMV polymerase (UL54) or UL44 promoters demonstrated that the inverted repeat element 1 (IR-1) is the target DNA motif of the IFI16 suppression. Suppression of the UL54 promoter was mediated by the IFI16-induced blocking of Sp1-like factors, as shown by electrophoretic mobility shift assays and chromatin immunoprecipitation. Accordingly, deletion of the Sp1 responsive element from the promoter of the UL44 protein, associated with UL54 in HCMV DNA synthesis, also relieved IFI16 suppression. Together, these results demonstrate that in addition to the previously recognized activity of IFI16 as DNA sensor, it also acts as restriction factor of herpesvirus replication [109] (FIGURE 2).

The fact that HCMV is still able to accomplish its replication cycle despite IFI16 restriction activity suggests that the virus has evolved evasion strategies to antagonize IFI16 activity. Insights into a possible HCMV evasion strategy have come from experiments using HCMV-infected fibroblasts (unpublished results). Early on during infection, IFI16 binds to virus DNA, however, at a later time point during viral DNA synthesis, IFI16 is mislocalized from the nucleus to the cytoplasm. The molecular mechanisms underlying this virus-induced nuclear egression of IFI16 require the binding of viral protein kinase UL97 to IFI16 (FIGURE 2). Upon binding, IFI16

undergoes phosphorylation, which in turn promotes its nucleo-cytoplasmic relocalization. The ESCRT (Endosomal Sorting Complex Required for Transport) machinery regulates thereafter the translocation of IFI16 into the virus assembly complex (AC). Finally, during the virus maturation step, IFI16 becomes incorporated into the new virions budding from the cells. Altogether, these results demonstrate that the strategy HCMV has evolved to evade IFI16 restriction activity consists of the nuclear egression of the protein followed by its incorporation into newly formed virions. It is likely that this is the event that leads to the inactivation of IFI16 antiviral activity.

(12)

Normal expression levels of the cellular protein Viperin are low. Its expression, however, is strongly induced by type I IFN, various viruses, lipopolysaccharide (LPS), and synthetic double-stranded RNA (poly I:C) [21,112,113]. It was initially demonstrated that infection of fibroblasts with HCMV triggered the expression of a subset of type I IFN-inducible genes. Two cDNA fragments, termed cig5 and cig33 (cytomegalovirus inducible gene 5 and 33, respectively) spanned the 5’-end and the 3’-end of the viperin mRNA, respectively [114]. Chin and Cresswell showed that these fragments contained the sequence of a HCMV-inducible mRNA and named the encoded protein “viperin” (standing for Virus Inhibitory Protein, Endoplasmic Reticulum-Associated, Interferon Inducible) [112]. In addition to HCMV, many viruses induce viperin expression including vesicular stomatitis virus and pseudorabies virus [115], Japanese encephalitis virus [116], sendai virus [113], West Nile virus [117], Chikungunya virus [118,119], hepatitis C virus [120], yellow fever virus [121], and human rhinovirus [122]. The viral induction of viperin may occur in a direct IFN-independent manner, or indirectly through the induction of IFN.

Several lines of evidence point to an antiviral function of viperin. For instance, viperin displays antiviral activity against influenza A virus by inhibiting virus release from infected cells. Through binding to and inhibiting farnesyl diphosphate synthase (FPPS; an enzyme essential for the biosynthesis of isoprenoid-derived lipids), viperin inhibits the release of influenza A virus from the plasma membrane by altering lipid raft organization and, in turn, budding [123]. In hepatic cells infected with hepatitis C virus, on the other hand, viperin acts by blocking the formation of the HCV RNA replication complex on the lipid raft – a process that requires the binding of host protein hVAP-33 to viral proteins NS5A and NS5B. By binding to hVAP-33, viperin disrupts the interaction with NS5A and NS5B, and thereby inhibits HCV replication complex formation and down-regulates virus replication [124].

However, while it was first thought that viperin acted as a restriction factor against HCMV infection, more recent evidence has brought to light a different scenario where it seems that HCMV actually exploits viperin to enhance its

(13)

expression. Stable expression of the protein before HCMV infection reduced the redistribution of viperin from the endoplasmic reticulum to the Golgi apparatus, thus preventing its vacuolar localization and inhibiting productive HCMV infection. Now it has been demonstrated that HCMV triggers an increase in lipogenesis in infected cells via a viperin-dependent pathway for its proper envelopment, mainly consisting of glycolipids [125]. The increase in lipogenesis follows the redistribution of viperin to mitochondria, where the protein interacts with and blocks the function of the mitochondrial trifunctional protein (TFP), the enzyme that mediates fatty acid- -peroxidation. This leads to a decrease of intracellular ATP, which in turn activates AMP-activated protein kinase (AMPK) and, consequently, the expression of the glucose transporter GLUT4; this is accompanied by an increase of glucose import and nuclear translocation of the glucose-regulated transcription factor ChREBP. As a consequence, the increase in the expression of enzymes involved in lipogenesis leads to an increased lipid synthesis and enhanced viral envelopment. Altogether, these findings demonstrate that HCMV maturation relies on its ability to exploit viperin activity in order to up-regulate lipid metabolism, needed for its proper glycolipid envelopment (FIGURE 3).

Nevertheless, since viperin also exhibits antiviral activity, it is in the interests of the virus to silence the protein. To prevent viperin antiviral activity, different viruses have adopted different strategies. For instance, the Chikungunya virus, which induces viperin expression via IRF3-mediated pathway, triggers a widespread cellular translation block, which down-regulates the synthesis of IFN-inducible genes including viperin [118]. Japanese encephalitis virus, on the other hand, although stimulating viperin expression, drives the protein to degradation by proteasomes [116].

Future perspectives

Significant progress has now been made into unraveling mechanisms RFs relying on to limit replication of HCMV. However, there is still significant scope for

(14)

future discovery and exploitation of RFs to cope HCMV infections. The evidence reviewed in this chapter supports the view that the innate antiviral immunity respond robustly following primary HCMV infection. This response is very likely protective in healthy immunocompetent individuals resulting in asymptomatic infections. However, despite this HCMV is able to establish latency and the virus is never eliminated from the infected individuals. A complex interplay between the innate antiviral and the adaptive immunity must exist in order to antagonize the HCMV-mediated immune evasion strategies that allow viral reactivation, extensive viral dissemination accompanied by clinical disease. A major challenge, therefore, will be to unveil the molecular mechanisms exploited by RFs to activate the adaptive immune response, that leads to the control of HCMV replication. A second major challenge will be the therapeutic exploitation of RFs biology that up to now has not received widespread attention. Indeed, the pharmacologic manipulation of RFs, aimed at blocking interactions between a viral antagonist and a host restriction factor, would appear as an attractive approach for the development of novel antivirals. However, consistent with their critical role in innate antiviral immunity, RFs such as IFI16 have been implicated in the pathogenesis of autoimmune diseases [126]. Therefore, it can be envisaged that such uncontrolled manipulation could possibly unbalance the immune response, favoring the onset of an autoimmune process.

(15)

Executive summary

 Restriction factors (RFs) form an important component of the innate immune response to viral replication.

 RFs so far described include: i) the cellular proteins PML, hDaxx, and Sp100, constituents of Nuclear Domain 10 (ND10); ii) IFI16, a nuclear protein and member of the PYHIN protein family; iii) Viperin (Virus Inhibitory Protein, Endoplasmic Reticulum-Associated, Interferon Inducible)

 Human Cytomegalovirus (HCMV) and other viruses have developed specific strategies in order to overcome this innate defense system

 The recently discovered property of IFI16 as RF for HCMV lies in its ability to inhibit virus DNA synthesis by down-regulating the transcription of viral gene UL54. The virus attempts to antagonize IFI16 by mislocalizing it from the nucleus to the cytoplasm via the action of viral protein UL97

 Detailed information of these HCMV restriction and counter-restriction mechanisms will provide new insights into how we might contribute to the development of novel therapeutic interventions

(16)

References

1. Takeuchi O, Akira S. Recognition of viruses by innate immunity. Immunol. Rev. 220, 214–224 (2007).

2. Bieniasz PD. Intrinsic immunity: a front-line defense against viral attack. Nat. Immunol. 5(11), 1109–1115 (2004).

●● The first description of the third branch of the traditionally bipartite immune system as a natural defense against viral infections

3. Bieniasz PD. Restriction factors: a defense against retroviral infection. Trends Microbiol. 11(6), 286–291 (2003).

4. Duggal NK, Emerman M. Evolutionary conflicts between viruses and restriction factors shape immunity. Nat. Rev. Immunol. 12(10), 687–695 (2012). 5. Johnson WE. Rapid adversarial co-evolution of viruses and cellular restriction

factors. Curr. Top. Microbiol. Immunol. 371, 123–151 (2013).

6. Hartley JW, Rowe WP, Huebner RJ. Host-range restrictions of murine leukemia viruses in mouse embryo cell cultures. J. Virol. 5(2), 221–225 (1970).

● The first time the term “restriction factor” has been adopted to describe a cellular component naturally conferring resistance to murine retroviruses. 7. Pincus T, Hartley JW, Rowe WP. A major genetic locus affecting resistance to infection with murine leukemia viruses. I. Tissue culture studies of naturally occurring viruses. J. Exp. Med. 133(6), 1219–1233 (1971).

8. Jolicoeur P, Rassart E. Effect of Fv-1 gene product on synthesis of linear and supercoiled viral DNA in cells infected with murine leukemia virus. J. Virol. 33(1), 183–195 (1980).

(17)

9. Harris RS, Bishop KN, Sheehy AM, et al. DNA deamination mediates innate immunity to retroviral infection. Cell. 113(6), 803–809 (2003).

10. Mangeat B, Turelli P, Caron G, Friedli M, Perrin L, Trono D. Broad antiretroviral defence by human APOBEC3G through lethal editing of nascent reverse transcripts. Nature. 424(6944), 99–103 (2003).

● In this study the mechanisms triggered by APOBEC3G to induce lethal editing of HIV-1 nascent reverse transcripts are described.

11. Keckesova Z, Ylinen LMJ, Towers GJ. The human and African green monkey TRIM5alpha genes encode Ref1 and Lv1 retroviral restriction factor activities. Proc. Natl. Acad. Sci. U. S. A. 101(29), 10780–10785 (2004).

12. Stremlau M, Owens CM, Perron MJ, Kiessling M, Autissier P, Sodroski J. The cytoplasmic body component TRIM5alpha restricts HIV-1 infection in Old World monkeys. Nature. 427(6977), 848–853 (2004).

● In this study the authors describe for the first how TRIM5alpha inhibits HIV-1 infection in Old World monkeys.

13. Neil SJD, Zang T, Bieniasz PD. Tetherin inhibits retrovirus release and is antagonized by HIV-1 Vpu. Nature. 451(7177), 425–430 (2008).

● In this study the authors describe the mechanisms tetherin relies on to inhibit HIV release and how the virus counteracts its activity.

14. Tokarev A, Skasko M, Fitzpatrick K, Guatelli J. Antiviral activity of the interferon-induced cellular protein BST-2/tetherin. AIDS Res. Hum. Retroviruses. 25(12), 1197–1210 (2009).

15. Akira S, Hemmi H. Recognition of pathogen-associated molecular patterns by TLR family. Immunol. Lett. 85(2), 85–95 (2003).

(18)

16. Barbalat R, Ewald SE, Mouchess ML, Barton GM. Nucleic acid recognition by the innate immune system. Annu. Rev. Immunol. 29, 185–214 (2011).

17. Hemmi H, Takeuchi O, Kawai T, et al. A Toll-like receptor recognizes bacterial DNA. Nature. 408(6813), 740–745 (2000).

18. Alexopoulou L, Holt AC, Medzhitov R, Flavell RA. Recognition of double-stranded RNA and activation of NF-kappaB by Toll-like receptor 3. Nature. 413(6857), 732–738 (2001).

19. Heil F, Hemmi H, Hochrein H, et al. Species-specific recognition of single-stranded RNA via toll-like receptor 7 and 8. Science. 303(5663), 1526–1529 (2004).

20. Loo Y-M, Gale M Jr. Immune signaling by RIG-I-like receptors. Immunity. 34(5), 680–692 (2011).

21. Ishii KJ, Coban C, Kato H, et al. A Toll-like receptor-independent antiviral response induced by double-stranded B-form DNA. Nat. Immunol. 7(1), 40–48 (2006).

22. Stetson DB, Medzhitov R. Recognition of cytosolic DNA activates an IRF3-dependent innate immune response. Immunity. 24(1), 93–103 (2006).

23. Nie Y, Wang Y-Y. Innate immune responses to DNA viruses. Protein Cell. 4(1), 1–7 (2013).

24. Bürckstümmer T, Baumann C, Blüml S, et al. An orthogonal proteomic-genomic screen identifies AIM2 as a cytoplasmic DNA sensor for the inflammasome. Nat. Immunol. 10(3), 266–272 (2009).

25. Fernandes-Alnemri T, Yu J-W, Datta P, Wu J, Alnemri ES. AIM2 activates the inflammasome and cell death in response to cytoplasmic DNA. Nature. 458(7237), 509–513 (2009).

(19)

● One of the first description for the IFN-inducible PYHIN family protein AIM2 (Absent in Melanoma 2) as a cytoplasmic DNA sensor.

26. Hornung V, Ablasser A, Charrel-Dennis M, et al. AIM2 recognizes cytosolic dsDNA and forms a caspase-1-activating inflammasome with ASC. Nature. 458(7237), 514–518 (2009).

● One of the first description for the IFN-inducible PYHIN family protein AIM2 (Absent in Melanoma 2) as a cytoplasmic DNA sensor.

27. Schroder K, Muruve DA, Tschopp J. Innate immunity: cytoplasmic DNA sensing by the AIM2 inflammasome. Curr. Biol. CB. 19(6), R262–265 (2009). 28. Rathinam VAK, Jiang Z, Waggoner SN, et al. The AIM2 inflammasome is

essential for host defense against cytosolic bacteria and DNA viruses. Nat. Immunol. 11(5), 395–402 (2010).

29. Cristea IM, Moorman NJ, Terhune SS, et al. Human cytomegalovirus pUL83 stimulates activity of the viral immediate-early promoter through its interaction with the cellular IFI16 protein. J. Virol. 84(15), 7803–7814 (2010).

● In this paper the authors identify a previously unknown role of viral pUL83 in the initiation of the human cytomegalovirus gene expression cascade through its interaction with the IFN-inducible protein IFI16.

30. Unterholzner L, Keating SE, Baran M, et al. IFI16 is an innate immune sensor for intracellular DNA. Nat. Immunol. 11(11), 997–1004 (2010).

●● Seminal paper describing for the first time the role of the PYHIN proteins IFI16 and AIM2 as components of the new family of innate DNA sensors 'AIM2-like receptors' (ALRs).

(20)

31. Kerur N, Veettil MV, Sharma-Walia N, et al. IFI16 acts as a nuclear pathogen sensor to induce the inflammasome in response to Kaposi Sarcoma-associated herpesvirus infection. Cell Host Microbe. 9(5), 363–375 (2011).

32. Orzalli MH, DeLuca NA, Knipe DM. Nuclear IFI16 induction of IRF-3 signaling during herpesviral infection and degradation of IFI16 by the viral ICP0 protein. Proc. Natl. Acad. Sci. U. S. A. 109(44), E3008–E3017 (2012). ● In this study the authors describe how HSV-1 ICP0 protein degrades IFI16

and inhibits IRF3 induction.

33. Ansari MA, Singh VV, Dutta S, et al. Constitutive interferon-inducible protein 16-inflammasome activation during Epstein-Barr virus latency I, II, and III in B and epithelial cells. J. Virol. 87(15), 8606–8623 (2013).

34. Horan KA, Hansen K, Jakobsen MR, et al. Proteasomal degradation of herpes simplex virus capsids in macrophages releases DNA to the cytosol for recognition by DNA sensors. J. Immunol. Baltim. Md 1950. 190(5), 2311–2319 (2013).

35. Johnson KE, Chikoti L, Chandran B. Herpes simplex virus 1 infection induces activation and subsequent inhibition of the IFI16 and NLRP3 inflammasomes. J. Virol. 87(9), 5005–5018 (2013).

36. Singh VV, Kerur N, Bottero V, et al. Kaposi’s sarcoma-associated herpesvirus latency in endothelial and B cells activates gamma interferon-inducible protein 16-mediated inflammasomes. J. Virol. 87(8), 4417–4431 (2013).

37. Baumann JG. Intracellular restriction factors in mammalian cells--An ancient defense system finds a modern foe. Curr. HIV Res. 4(2), 141–168 (2006).

38. Reymond A, Meroni G, Fantozzi A, et al. The tripartite motif family identifies cell compartments. EMBO J. 20(9), 2140–2151 (2001).

(21)

39. Greene WC. Redistricting the retroviral restriction factors. Nat. Med. 10(8), 778–780 (2004).

40. Harris RS, Hultquist JF, Evans DT. The restriction factors of human immunodeficiency virus. J. Biol. Chem. 287(49), 40875–40883 (2012).

41. Lilley CE, Chaurushiya MS, Boutell C, Everett RD, Weitzman MD. The intrinsic antiviral defense to incoming HSV-1 genomes includes specific DNA repair proteins and is counteracted by the viral protein ICP0. PLoS Pathog. 7(6), e1002084 (2011).

42. Tavalai N, Adler M, Scherer M, Riedl Y, Stamminger T. Evidence for a Dual Antiviral Role of the Major Nuclear Domain 10 Component Sp100 during the Immediate-Early and Late Phases of the Human Cytomegalovirus Replication Cycle. J. Virol. 85(18), 9447–9458 (2011).

43. Boutell C, Everett RD. Regulation of alphaherpesvirus infections by the ICP0 family of proteins. J. Gen. Virol. 94(Pt 3), 465–481 (2013).

44. Stuurman N, Meijne AM, van der Pol AJ, de Jong L, van Driel R, van Renswoude J. The nuclear matrix from cells of different origin. Evidence for a common set of matrix proteins. J. Biol. Chem. 265(10), 5460–5465 (1990). 45. Lallemand-Breitenbach V, de The H. PML Nuclear Bodies. Cold Spring Harb.

Perspect. Biol. [Internet]. 2(5) (2010). Available from: http://www.ncbi.nlm.nih.gov/pmc/articles/PMC2857171/.

46. Lallemand-Breitenbach V, Zhu J, Puvion F, et al. Role of promyelocytic leukemia (PML) sumolation in nuclear body formation, 11S proteasome recruitment, and As2O3-induced PML or PML/retinoic acid receptor alpha degradation. J. Exp. Med. 193(12), 1361–1371 (2001).

(22)

47. Lindsay CR, Morozov VM, Ishov AM. PML NBs (ND10) and Daxx: from nuclear structure to protein function. Front. Biosci. J. Virtual Libr. 13, 7132– 7142 (2008).

48. Shen TH, Lin H-K, Scaglioni PP, Yung TM, Pandolfi PP. The mechanisms of PML-nuclear body formation. Mol. Cell. 24(3), 331–339 (2006).

49. Bernardi R, Pandolfi PP. Structure, dynamics and functions of promyelocytic leukaemia nuclear bodies. Nat. Rev. Mol. Cell Biol. 8(12), 1006–1016 (2007). 50. Russell RA, Adams NM, Stephens DA, Batty E, Jensen K, Freemont PS.

Segmentation of fluorescence microscopy images for quantitative analysis of cell nuclear architecture. Biophys. J. 96(8), 3379–3389 (2009).

51. Boisvert FM, Hendzel MJ, Bazett-Jones DP. Promyelocytic leukemia (PML) nuclear bodies are protein structures that do not accumulate RNA. J. Cell Biol. 148(2), 283–292 (2000).

52. Koken MH, Linares-Cruz G, Quignon F, et al. The PML growth-suppressor has an altered expression in human oncogenesis. Oncogene. 10(7), 1315–1324 (1995).

53. Terris B, Baldin V, Dubois S, et al. PML nuclear bodies are general targets for inflammation and cell proliferation. Cancer Res. 55(7), 1590–1597 (1995). 54. Gurrieri C, Capodieci P, Bernardi R, et al. Loss of the tumor suppressor PML in

human cancers of multiple histologic origins. J. Natl. Cancer Inst. 96(4), 269– 279 (2004).

55. Everett RD. Interactions between DNA viruses, ND10 and the DNA damage response. Cell. Microbiol. 8(3), 365–374 (2006).

56. Guo A, Salomoni P, Luo J, et al. The function of PML in p53-dependent apoptosis. Nat. Cell Biol. 2(10), 730–736 (2000).

(23)

57. Bernardi R, Pandolfi PP. Role of PML and the PML-nuclear body in the control of programmed cell death. Oncogene. 22(56), 9048–9057 (2003).

58. Bernardi R, Papa A, Pandolfi PP. Regulation of apoptosis by PML and the PML-NBs. Oncogene. 27(48), 6299–6312 (2008).

59. Bischof O, Kirsh O, Pearson M, Itahana K, Pelicci PG, Dejean A. Deconstructing PML-induced premature senescence. EMBO J. 21(13), 3358– 3369 (2002).

60. Salomoni P, Pandolfi PP. The role of PML in tumor suppression. Cell. 108(2), 165–170 (2002).

61. Ito K, Bernardi R, Morotti A, et al. PML targeting eradicates quiescent leukaemia-initiating cells. Nature. 453(7198), 1072–1078 (2008).

62. Maul GG, Guldner HH, Spivack JG. Modification of discrete nuclear domains induced by herpes simplex virus type 1 immediate early gene 1 product (ICP0). J. Gen. Virol. 74 ( Pt 12), 2679–2690 (1993).

63. Carvalho T, Seeler JS, Ohman K, et al. Targeting of adenovirus E1A and E4-ORF3 proteins to nuclear matrix-associated PML bodies. J. Cell Biol. 131(1), 45–56 (1995).

64. Szekely L, Pokrovskaja K, Jiang WQ, de The H, Ringertz N, Klein G. The Epstein-Barr virus-encoded nuclear antigen EBNA-5 accumulates in PML-containing bodies. J. Virol. 70(4), 2562–2568 (1996).

65. Everett RD. DNA viruses and viral proteins that interact with PML nuclear bodies. Oncogene. 20(49), 7266–7273 (2001).

66. Regad T, Chelbi-Alix MK. Role and fate of PML nuclear bodies in response to interferon and viral infections. Oncogene. 20(49), 7274–7286 (2001).

(24)

67. Negorev DG, Vladimirova OV, Maul GG. Differential functions of interferon-upregulated Sp100 isoforms: herpes simplex virus type 1 promoter-based immediate-early gene suppression and PML protection from ICP0-mediated degradation. J. Virol. 83(10), 5168–5180 (2009).

68. Kelly C, Van Driel R, Wilkinson GW. Disruption of PML-associated nuclear bodies during human cytomegalovirus infection. J. Gen. Virol. 76 ( Pt 11), 2887–2893 (1995).

69. Korioth F, Maul GG, Plachter B, Stamminger T, Frey J. The nuclear domain 10 (ND10) is disrupted by the human cytomegalovirus gene product IE1. Exp. Cell Res. 229(1), 155–158 (1996).

70. Ahn JH, Hayward GS. The major immediate-early proteins IE1 and IE2 of human cytomegalovirus colocalize with and disrupt PML-associated nuclear bodies at very early times in infected permissive cells. J. Virol. 71(6), 4599– 4613 (1997).

71. Kim Y-E, Lee J-H, Kim ET, et al. Human cytomegalovirus infection causes degradation of Sp100 proteins that suppress viral gene expression. J. Virol. 85(22), 11928–11937 (2011).

72. Tavalai N, Stamminger T. Intrinsic cellular defense mechanisms targeting human cytomegalovirus. Virus Res. 157(2), 128–133 (2011).

73. Tavalai N, Papior P, Rechter S, Leis M, Stamminger T. Evidence for a Role of the Cellular ND10 Protein PML in Mediating Intrinsic Immunity against Human Cytomegalovirus Infections. J. Virol. 80(16), 8006–8018 (2006).

●● The first evidence that ND10 protein PML acts as restriction factor against HCMV by inhibiting viral IE gene expression.

(25)

74. Tavalai N, Stamminger T. New insights into the role of the subnuclear structure ND10 for viral infection. Biochim. Biophys. Acta. 1783(11), 2207–2221 (2008). 75. Tavalai N, Papior P, Rechter S, Stamminger T. Nuclear domain 10 components

promyelocytic leukemia protein and hDaxx independently contribute to an intrinsic antiviral defense against human cytomegalovirus infection. J. Virol. 82(1), 126–137 (2008).

76. Everett RD, Chelbi-Alix MK. PML and PML nuclear bodies: implications in antiviral defence. Biochimie. 89(6-7), 819–830 (2007).

77. Adler M, Tavalai N, Müller R, Stamminger T. Human cytomegalovirus immediate-early gene expression is restricted by the nuclear domain 10 component Sp100. J. Gen. Virol. 92(Pt 7), 1532–1538 (2011).

78. Woodhall DL, Groves IJ, Reeves MB, Wilkinson G, Sinclair JH. Human Daxx-mediated Repression of Human Cytomegalovirus Gene Expression Correlates with a Repressive Chromatin Structure around the Major Immediate Early Promoter. J. Biol. Chem. 281(49), 37652–37660 (2006).

79. Lukashchuk V, McFarlane S, Everett RD, Preston CM. Human Cytomegalovirus Protein pp71 Displaces the Chromatin-Associated Factor ATRX from Nuclear Domain 10 at Early Stages of Infection. J. Virol. 82(24), 12543–12554 (2008).

80. Nitzsche A, Steinhäusser C, Mücke K, Paulus C, Nevels M. Histone H3 lysine 4 methylation marks postreplicative human cytomegalovirus chromatin. J. Virol. 86(18), 9817–9827 (2012).

81. Preston CM, Nicholl MJ. Role of the cellular protein hDaxx in human cytomegalovirus immediate-early gene expression. J. Gen. Virol. 87(Pt 5), 1113–1121 (2006).

(26)

82. Groves IJ, Sinclair JH. Knockdown of hDaxx in normally non-permissive undifferentiated cells does not permit human cytomegalovirus immediate-early gene expression. J. Gen. Virol. 88(Pt 11), 2935–2940 (2007).

83. Reeves M, Woodhall D, Compton T, Sinclair J. Human cytomegalovirus IE72 protein interacts with the transcriptional repressor hDaxx to regulate LUNA gene expression during lytic infection. J. Virol. 84(14), 7185–7194 (2010). 84. Shin HJ, Kim Y-E, Kim ET, Ahn J-H. The chromatin-tethering domain of

human cytomegalovirus immediate-early (IE) 1 mediates associations of IE1, PML and STAT2 with mitotic chromosomes, but is not essential for viral replication. J. Gen. Virol. 93(Pt 4), 716–721 (2012).

85. Hofmann H, Sindre H, Stamminger T. Functional Interaction between the pp71 Protein of Human Cytomegalovirus and the PML-Interacting Protein Human Daxx. J. Virol. 76(11), 5769–5783 (2002).

86. Cantrell SR, Bresnahan WA. Interaction between the Human Cytomegalovirus UL82 Gene Product (pp71) and hDaxx Regulates Immediate-Early Gene Expression and Viral Replication. J. Virol. 79(12), 7792–7802 (2005).

87. Greaves RF, Mocarski ES. Defective growth correlates with reduced accumulation of a viral DNA replication protein after low-multiplicity infection by a human cytomegalovirus ie1 mutant. J. Virol. 72(1), 366–379 (1998).

88. Gawn JM, Greaves RF. Absence of IE1 p72 protein function during low-multiplicity infection by human cytomegalovirus results in a broad block to viral delayed-early gene expression. J. Virol. 76(9), 4441–4455 (2002).

89. Stinski MF, Meier JL. Immediate–early viral gene regulation and function [Internet]. In: Human Herpesviruses: Biology, Therapy, and Immunoprophylaxis. Arvin A, Campadelli-Fiume G, Mocarski E, et al. (Eds.).

(27)

Cambridge University Press, Cambridge (2007) [cited 2013 Dec 9]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK47383/.

90. White EA, Spector DH. Early viral gene expression and function [Internet]. In: Human Herpesviruses: Biology, Therapy, and Immunoprophylaxis. Arvin A, Campadelli-Fiume G, Mocarski E, et al. (Eds.). Cambridge University Press, Cambridge (2007) [cited 2013 Dec 9]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK47437/.

91. Xu Y, Ahn JH, Cheng M, et al. Proteasome-independent disruption of PML oncogenic domains (PODs), but not covalent modification by SUMO-1, is required for human cytomegalovirus immediate-early protein IE1 to inhibit PML-mediated transcriptional repression. J. Virol. 75(22), 10683–10695 (2001).

92. Gariglio M, Mondini M, De Andrea M, Landolfo S. The multifaceted interferon-inducible p200 family proteins: from cell biology to human pathology. J. Interferon Cytokine Res. Off. J. Int. Soc. Interferon Cytokine Res. 31(1), 159–172 (2011).

93. Johnstone RW, Kershaw MH, Trapani JA. Isotypic variants of the interferon-inducible transcriptional repressor IFI 16 arise through differential mRNA splicing. Biochemistry (Mosc.). 37(34), 11924–11931 (1998).

94. Cridland JA, Curley EZ, Wykes MN, et al. The mammalian PYHIN gene family: phylogeny, evolution and expression. BMC Evol. Biol. 12, 140 (2012). 95. Gariglio M, Azzimonti B, Pagano M, et al. Immunohistochemical expression

analysis of the human interferon-inducible gene IFI16, a member of the HIN200 family, not restricted to hematopoietic cells. J. Interferon Cytokine Res. Off. J. Int. Soc. Interferon Cytokine Res. 22(7), 815–821 (2002).

(28)

96. Wei W, Clarke CJP, Somers GR, et al. Expression of IFI 16 in epithelial cells and lymphoid tissues. Histochem. Cell Biol. 119(1), 45–54 (2003).

97. Veeranki S, Choubey D. Interferon-inducible p200-family protein IFI16, an innate immune sensor for cytosolic and nuclear double-stranded DNA: Regulation of subcellular localization. Mol. Immunol. 49(4), 567–571 (2012). 98. Dawson MJ, Trapani JA. The interferon-inducible autoantigen, IFI 16:

localization to the nucleolus and identification of a DNA-binding domain. Biochem. Biophys. Res. Commun. 214(1), 152–162 (1995).

99. Li T, Diner BA, Chen J, Cristea IM. Acetylation modulates cellular distribution and DNA sensing ability of interferon-inducible protein IFI16. Proc. Natl. Acad. Sci. U. S. A. 109(26), 10558–10563 (2012).

100. Costa S, Borgogna C, Mondini M, et al. Redistribution of the nuclear protein IFI16 into the cytoplasm of ultraviolet B-exposed keratinocytes as a mechanism of autoantigen processing. Br. J. Dermatol. 164(2), 282–290 (2011).

101. Johnstone RW, Wei W, Greenway A, Trapani JA. Functional interaction between p53 and the interferon-inducible nucleoprotein IFI 16. Oncogene. 19(52), 6033–6042 (2000).

102. Aglipay JA, Lee SW, Okada S, et al. A member of the Pyrin family, IFI16, is a novel BRCA1-associated protein involved in the p53-mediated apoptosis pathway. Oncogene. 22(55), 8931–8938 (2003).

103. Gugliesi F, Mondini M, Ravera R, et al. Up-regulation of the interferon-inducible IFI16 gene by oxidative stress triggers p53 transcriptional activity in endothelial cells. J. Leukoc. Biol. 77(5), 820–829 (2005).

(29)

104. Liao JCC, Lam R, Brazda V, et al. Interferon-inducible protein 16: insight into the interaction with tumor suppressor p53. Struct. Lond. Engl. 1993. 19(3), 418–429 (2011).

105. Sharma S, DeOliveira RB, Kalantari P, et al. Innate immune recognition of an AT-rich stem-loop DNA motif in the Plasmodium falciparum genome. Immunity. 35(2), 194–207 (2011).

106. Brázda V, Coufal J, Liao JCC, Arrowsmith CH. Preferential binding of IFI16 protein to cruciform structure and superhelical DNA. Biochem. Biophys. Res. Commun. 422(4), 716–720 (2012).

107. Jin T, Perry A, Jiang J, et al. Structures of the HIN domain:DNA complexes reveal ligand binding and activation mechanisms of the AIM2 inflammasome and IFI16 receptor. Immunity. 36(4), 561–571 (2012).

108. Nour AM, Reichelt M, Ku C-C, Ho M-Y, Heineman TC, Arvin AM. Varicella-zoster virus infection triggers formation of an interleukin-1β (IL-1β)-processing inflammasome complex. J. Biol. Chem. 286(20), 17921–17933 (2011).

109. Gariano GR, Dell’Oste V, Bronzini M, et al. The intracellular DNA sensor IFI16 gene acts as restriction factor for human cytomegalovirus replication. PLoS Pathog. 8(1), e1002498 (2012).

●● Implication of IFI16 as a novel restriction factor against HCMV replication. The data presented provide new insight into the physiological functions of the IFN-inducible gene IFI16 as a viral restriction factor.

110. Conrady CD, Zheng M, Fitzgerald KA, Liu C, Carr DJJ. Resistance to HSV-1 infection in the epithelium resides with the novel innate sensor, IFI-16. Mucosal Immunol. 5(2), 173–183 (2012).

(30)

111. Søby S, Laursen RR, Ostergaard L, Melchjorsen J. HSV-1-induced chemokine expression via IFI16-dependent and IFI16-independent pathways in human monocyte-derived macrophages. Herpesviridae. 3(1), 6 (2012).

112. Chin KC, Cresswell P. Viperin (cig5), an IFN-inducible antiviral protein directly induced by human cytomegalovirus. Proc. Natl. Acad. Sci. U. S. A. 98(26), 15125–15130 (2001).

113. Severa M, Coccia EM, Fitzgerald KA. Toll-like receptor-dependent and -independent viperin gene expression and counter-regulation by PRDI-binding factor-1/BLIMP1. J. Biol. Chem. 281(36), 26188–26195 (2006).

114. Zhu H, Cong JP, Shenk T. Use of differential display analysis to assess the effect of human cytomegalovirus infection on the accumulation of cellular RNAs: induction of interferon-responsive RNAs. Proc. Natl. Acad. Sci. U. S. A. 94(25), 13985–13990 (1997).

115. Boudinot P, Riffault S, Salhi S, et al. Vesicular stomatitis virus and pseudorabies virus induce a vig1/cig5 homologue in mouse dendritic cells via different pathways. J. Gen. Virol. 81(11), 2675–2682 (2000).

116. Chan Y-L, Chang T-H, Liao C-L, Lin Y-L. The Cellular Antiviral Protein Viperin Is Attenuated by Proteasome-Mediated Protein Degradation in Japanese Encephalitis Virus-Infected Cells. J. Virol. 82(21), 10455–10464 (2008).

117. Szretter KJ, Brien JD, Thackray LB, Virgin HW, Cresswell P, Diamond MS. The interferon-inducible gene viperin restricts West Nile virus pathogenesis. J. Virol. 85(22), 11557–11566 (2011).

118. White LK, Sali T, Alvarado D, et al. Chikungunya virus induces IPS-1-dependent innate immune activation and protein kinase R-inIPS-1-dependent translational shutoff. J. Virol. 85(1), 606–620 (2011).

(31)

119. Teng T-S, Foo S-S, Simamarta D, et al. Viperin restricts chikungunya virus replication and pathology. J. Clin. Invest. 122(12), 4447–4460 (2012).

120. Helbig KJ, Lau DT-Y, Semendric L, Harley HAJ, Beard MR. Analysis of ISG expression in chronic hepatitis C identifies viperin as a potential antiviral effector. Hepatol. Baltim. Md. 42(3), 702–710 (2005).

121. Khaiboullina SF, Rizvanov AA, Holbrook MR, St Jeor S. Yellow fever virus strains Asibi and 17D-204 infect human umbilical cord endothelial cells and induce novel changes in gene expression. Virology. 342(2), 167–176 (2005). 122. Proud D, Turner RB, Winther B, et al. Gene expression profiles during in vivo

human rhinovirus infection: insights into the host response. Am. J. Respir. Crit. Care Med. 178(9), 962–968 (2008).

123. Wang X, Hinson ER, Cresswell P. The interferon-inducible protein viperin inhibits influenza virus release by perturbing lipid rafts. Cell Host Microbe. 2(2), 96–105 (2007).

124. Wang S, Wu X, Pan T, et al. Viperin inhibits hepatitis C virus replication by interfering with binding of NS5A to host protein hVAP-33. J. Gen. Virol. 93(Pt 1), 83–92 (2012).

125. Seo J-Y, Cresswell P. Viperin regulates cellular lipid metabolism during human cytomegalovirus infection. PLoS Pathog. 9(8), e1003497 (2013).

126. Mondini M, Costa S, Sponza S, Gugliesi F, Gariglio M, Landolfo S. The interferon-inducible HIN-200 gene family in apoptosis and inflammation: implication for autoimmunity. Autoimmunity. 43(3), 226–231 (2010).

(32)

Legends to Figures

Figure 1. Restriction of HCMV replication by ND10 and virus evasion. Schematic representation of the mechanisms exploited by HCMV to overwhelm the biological role of ND10 proteins as restriction factors in the prevention of viral replication.

Figure 2. Restriction of HCMV replication by IFI16 protein and virus evasion. Nuclear IFI16 protein inhibits HCMV replication by blocking the activity of Sp1-like transcription factors on viral UL54 promoter. Conversely, HCMV mislocalizes IFI16 into the cytoplasm through its phosphorylation by the viral protein kinase UL97. Figure 3. HCMV exploits viperin to enhance its expression. For its proper envelopment, HCMV triggers an increase in lipogenesis in infected cells via a viperin-dependent pathway, involving cellular enzymes committed to lipid synthesis. In particular, Viperin interaction with the viral protein vMIA (viral mitochondrial inhibitor of apoptosis) results in Viperin relocalization from the endoplasmic reticulum to the mitochondria, and consequent reduced cellular ATP generation, which resulted in actin cytoskeleton disruption.

Riferimenti

Documenti correlati

Many would instinctively cite Rawls’ opening statement in A Theory of Justice; the idea that “Justice is the first virtue of social institutions” (1971/99). However, we would argue

Alteration of the vaginal microbial milieu and changes in the genital tract mucosal immune environment could exert a primary role in both pregnancy establishment and

To systematically identify host cell factors that modulate the different steps of AAV transduction, we performed a high-throughput RNAi screening using a genome-wide siRNA

Neither IFN-I-Mediated Phosphorylation nor Nuclear Translocation of STAT2 Are Required for STAT2 Interaction with YFV NS5 (A) Confocal microscopy images showing subcellular

Sino a pochi anni fa le strategie d’indagine per l’analisi dell’aDNA a disposizione dei ricercatori si limitavano all’amplificazione mediante PCR delle porzioni

Representative images of the three major histopathological growth patterns (HGPs) in colorectal liver metastases identified on H&E-stained tissue sections: (A) replacement

Nel 1847, poco prima che venisse istituita la Prefettura, occupavano gli uffi- ci della Camera di soprintendenza grossetana il provveditore Giacomo Grandoni, cavaliere di Santo

In searching for the molecular mechanisms at the basis of CIITA and TRIM22-mediated inhibition of HIV-1 transcription, we demonstrated that CIITA and TRIM22 inhibit HIV-1