The interplay between host and viral factors in shaping the ... - Nature

19 downloads 19 Views 199KB Size Report
Dec 5, 2006 - ... Alexandra J Corbett1,2, William D Rawlinson3, Gillian M Scott4 and ...... 19 Scalzo AA, Fitzgerald NA, Wallace CR, Gibbons AE, Smart YC, ...
Immunology and Cell Biology (2007) 85, 46–54 & 2007 Australasian Society for Immunology Inc. All rights reserved 0818-9641/07 $30.00 www.nature.com/icb

REVIEW

The interplay between host and viral factors in shaping the outcome of cytomegalovirus infection Anthony A Scalzo1,2, Alexandra J Corbett1,2, William D Rawlinson3, Gillian M Scott4 and Mariapia A Degli-Esposti1,2 Cytomegalovirus (CMV) remains a major human pathogen causing significant morbidity and mortality in immunosuppressed or immunoimmature individuals. Although significant advances have been made in dissecting out certain features of the host response to human CMV (HCMV) infection, the strict species specificity of CMVs means that most aspects of antiviral immunity are best assessed in animal models. The mouse model of murine CMV (MCMV) infection is an important tool for analysis of in vivo features of host–virus interactions and responses to antiviral drugs that are difficult to assess in humans. Important studies of the contribution of host resistance genes to infection outcome, interplays between innate and adaptive host immune responses, the contribution of virus immune evasion genes and genetic variation in these genes to the establishment of persistence and in vivo studies of resistance to antiviral drugs have benefited from the well-developed MCMV model. In this review, we discuss recent advances in the immunobiology of host–CMV interactions that provide intriguing insights into the complex interplay between host and virus that ultimately facilitates viral persistence. We also discuss recent studies of genetic responses to antiviral therapy, particularly changes in DNA polymerase and protein kinase genes of MCMV and HCMV. Immunology and Cell Biology (2007) 85, 46–54. doi:10.1038/sj.icb.7100013; published online 5 December 2006 Keywords: viral infection; genetic variability; immune evasion; cytomegalovirus; NK cells; CTL

The outcome of host–virus interactions is determined by a range of factors including inherent host resistance and environmental factors, such as nutritional status, together with viral factors such as variation in genes governing levels of infectivity, tropism and immune evasion. Some host genes that confer resistance are polymorphic between individuals within populations, and thus different individuals will vary in their relative susceptibility to infection. The outbred nature of the human population has made it difficult to assess the contribution of host resistance genes in determining disease outcome of human cytomegalovirus (HCMV) infection. However, murine CMV (MCMV) infection of inbred mouse strains has provided many important insights into host genes that regulate CMV infection. ROLE OF HOST RESISTANCE GENES IN CONTROLLING CMV INFECTION Research on the contribution of host resistance genes to the outcome of MCMV infection was initiated about three decades ago.1,2 Studies by Grundy et al.3 established that both murine major histocompatibility complex (MHC) or H2 loci, together with non-H2 loci contribute to the control of acute MCMV infection. Resistance was associated with the protective H2k haplotype. Non-H2 resistance loci

were identified as being present on the C57BL/6J (abbreviated B6), C3H and CBA genetic backgrounds. Role of H2 genes in resistance To investigate the contribution of the H2 complex to MCMV resistance, Price et al.4 isolated macrophages from mice differing in H2 genotype and assessed their susceptibility to MCMV infection in vitro. These studies showed that cells from H2k haplotype mice were considerably less sensitive to infection,4 and mapping studies using intra-H2 haplotype mice indicated that MHC class I genes contributed to this effect.5 Further evidence for a role of MHC class I molecules was demonstrated by the finding that transfection of cells that were largely resistant to MCMV infection with specific MHC class I molecules increased their susceptibility to MCMV infection.6 Altogether, these data indicated that MHC class I molecules could function as a receptor or coreceptor for MCMV entry. Subsequent investigations using MHC class I-negative, b2-microglobulin (b2m) knockout mice on the B6 background,7 or 129/SvB6 background,8 did not support a role for class I molecules as entry coreceptors. However, it is conceivable that these analyses may have missed differences between wild-type and b2m / mice. The in vivo assessments of virus replication were only performed in the spleen7 or

1Immunology and Virology Program, Centre for Ophthalmology and Visual Science, The University of Western Australia, Nedlands, Western Australia, Australia; 2Centre for Experimental Immunology, Lions Eye Institute, Nedlands, Western Australia, Australia; 3Department of Microbiology, Virology Division, SEALS, Prince of Wales Hospital, Randwick, New South Wales, Australia and 4Virology Research, POWH and UNSW Research Laboratories, Prince of Wales Hospital, Randwick, New South Wales, Australia Correspondence: Dr AA Scalzo, Centre for Experimental Immunology, Lions Eye Institute, 2 Verdun St, Nedlands, Western Australia 6009, Australia. E-mail: [email protected] Received 9 October 2006; accepted 12 October 2006; published online 5 December 2006

The interplay between host and viral factors AA Scalzo et al 47

after the peak of visceral virus replication (i.e. after 2 weeks postinfection (p.i.)).8 The in vitro demonstration of H2 effects in fibroblasts is dependent on seeding densities of fibroblasts cultures.9 Hence, further studies on the role of MHC class I molecules as coreceptors for MCMV entry are warranted using b2m / mice on the BALB/c background where the contribution of natural killer (NK) cells to resistance is not as great. More recently, a role for the protective H2k haplotype in contributing to NK cell-associated resistance in the MA/My mouse strain has been established10,11 which is discussed below. Effects of host genotype on type I interferon production Type I interferons (IFN) are among the key early cytokines that have direct antiviral as well as immunoregulatory effects (reviewed by Biron12), including the regulation of NK cell cytolytic activity. Analyses of IFN levels in sera at 6 h p.i. have revealed that mousestrain-dependent differences exist and that these are dependent on non-H2 genetic effects.13 A more recent study by Dalod et al.14 confirmed non-H2 regulation of IFN-a production in H2b-matched mice, with 129/J mice producing 47-fold more than B6 mice at 36 h p.i. Whether the strain-dependent differences in levels of type I IFNs following MCMV infection are due to different frequencies of cells that produce this cytokine, such as plasmacytoid dendritic cells (DCs), or allelic variability in genes that regulate type I IFN production, remains to be defined. Host genes that regulate NK cell control of MCMV infection The importance of NK cells in controlling MCMV was demonstrated about two decades ago in studies showing exacerbation or amelioration of infection when adult B6 mice were depleted of NK cells or when NK cells were adoptively transferred into suckling mice before MCMV infection, respectively.15,16 Analyses of NK cell activity in inbred strains following MCMV infection also revealed a general correlation between the level of NK cell cytotoxicity and resistance status to MCMV.17 Cmv1. A Mendelian analysis of the genetic basis for differences in the level of MCMV replication in the spleens of susceptible BALB/c mice and moderately resistant B6 mice led to the identification of the Cmv1 resistance locus. Cmv1 was shown to have a dominant effect and was mapped to the distal region of mouse chromosome 6.18 The effect of Cmv1 was subsequently shown to be mediated via NK cell control of viral infection and the locus was mapped to the NK cell gene complex (NKC), a region which encodes both inhibitory and activation NK cell receptors.19 Following detailed genetic and physical mapping studies, the gene mediating the Cmv1 effect was identified as Ly49h or Klra8, a member of the Ly49 family of molecules in the distal NKC region.20–22 Ly49H is an activation receptor expressed on NK cells. Although Ly49H lacks intracellular signaling motifs in its cytoplasmic domain, it physically associates with the DAP12 adaptor molecule via residues in its transmembrane domain and phosphorylation of DAP12, via the immunoreceptor tyrosine-based activation motif (ITAM) present in the cytoplasmic domain of this molecule, is responsible for NK cell signaling.23,24 Formal proof of the role of Ly49H in mediating the Cmv1 effect came from the finding that Ly49H transgenic mice on the susceptible BALB/c and FVB genetic backgrounds are resistant to MCMV infection.25 The critical role of the DAP12 adaptor in Cmv1/Ly49H-mediated resistance in B6 mice was independently confirmed by studies showing that mice carrying targeted mutations in the DAP12 ITAM signaling motif are highly susceptible to MCMV infection.26

Analyses of how Ly49H specifically recognizes MCMV-infected cells and triggers NK cell activation revealed that it binds to a MHC class I-like, MCMV-encoded molecule called m157.27,28 Engagement of m157 is sufficient to trigger NK cell cytolysis as well as the secretion of IFN-g and the chemokine ATAC/lymphotactin.28 Also it binds the Ly49H activation receptor that is expressed in B6 mice, the m157 viral glycoprotein also binds to the inhibitory Ly49I molecule expressed in the 129/J mouse strain,27 raising the possibility that, at least in some infections, m157 may function as an immune evasion molecule. Disruption of the m157 gene in the Smith strain of MCMV led to escape from NK cell control in the resistant B6 strain,29 with only a weak effect observed in the susceptible 129/J strain.29 Recent analyses have shown that the m157 molecule is a cell-surface glycophosphatidylinositol (gpi)-anchored protein that does not require b2m or transporter associated with antigen processing 1 (TAP1) for expression.30 The expression profile of m157 in the context of infection of resistant and susceptible mouse strains remains to be investigated and it is likely to provide important information about the kinetics of interaction between this molecule and its receptors. It is also worth noting that expansion of Ly49H-positive NK cells is observed during the late stages of infection in an MCMV-specific manner (i.e. not observed in other viral infections).31 The impact that the late expansion of Ly49H-positive NK cells has on the ongoing control of infection and/or the generation of adaptive antiviral immune responses remains to be evaluated. Variation in the NKC and implications for host resistance. Cmv1mediated resistance was first defined in the moderately resistant B6 mouse strain.18 Surveys of acute virus replication in the spleens of a range of inbred mouse strains indicated that resistance-like phenotypes were relatively uncommon with only the MA/My strain showing an NK cell-mediated control of viral replication that was reversed following depletion of NK cells with anti-NK1.1 monoclonal antibody (MAb).32 Analyses of the NKC regions of inbred strains of mice have revealed that conserved NKC haplotypes exist33–35; interestingly, the B6 NKC haplotype was found to be uncommon among the mouse strains tested. In addition, more recent analysis of outbred wild mice has provided independent evidence for the existence of haplotypes within the NKC region, and has confirmed that the B6-like haplotype and Cmv1-like resistance are uncommon in the outbred mouse populations studied.36 Consequently, as expected, most outbred mice in the population tested are susceptible to MCMV infection. Given the paucity of Cmv1-like alleles in outbred mouse strains, further analysis has been conducted to identify other MCMV resistance genes. Other non-H2 resistance genes – Cmv2, Cmv3 and Cmv4. The New Zealand inbred mouse strains NZW and NZB were previously found to share very similar haplotypes across the NKC region.34 Recent analyses of MCMV resistance phenotypes in these strains revealed that the NZW strain is resistant and NZB is susceptible.37 NZW resistance can be reversed by anti-NK1.1 MAb treatment, but not by anti-Ly49H (3D10) MAb treatment, despite the fact that NZW NK cells express receptors that crossreact with 3D10. Furthermore, the NZW NK cells do not bind to the viral glycoprotein m157. Altogether these findings suggest that MCMV resistance in the NZW strain is Cmv1-independent. Further analysis of the genetic control of resistance in the NZW strain indicated that a multigenic mode of resistance (collectively termed Cmv2) operates in this strain with contributions from loci that map to chromosomes 17 and X.37 The genes encoded by these loci remain to be identified. Immunology and Cell Biology

The interplay between host and viral factors AA Scalzo et al 48

Two recent studies have analyzed the NK cell-mediated MCMV resistance that operates in the MA/My inbred strain.10,11 Resistance in this strain was found to be Ly49h-independent since MA/My NK cells do not bind the anti-Ly49H MAb or soluble m157 proteins,11 and indeed the Ly49h/Klra8 gene was found not be expressed in this strain.10 Based on genetic analyses both studies found that the resistance of NK cells in this strain was H2k dependent. In the study by Desrosier et al.,10 the Cmv3 locus was defined as being linked to the Klra gene cluster in the NKC, and the Klra16MA/My (or Ly49PMA/My) NK cell activation receptor was shown to be involved in recognition of MCMV-infected cells. As recognition of MCMV-infected cells in MA/My mice could be blocked either by anti-H2Dk or anti-Ly49P antibodies, NK cell-mediated virus resistance in this strain might involve interactions between the NK cell receptor Ly49P and the H2 molecule Dk. In addition to innate H2-mediated control of MCMV infection, Dighe et al.,11 using crosses between MA/My and C57L inbred mice, identified further contributions from both H2 and nonH2 loci in mediating resistance. Collectively, these findings support an important role for NK cells in innate H2-linked protection in MCMV infection, and suggest that NK cell receptors may be rapidly selected for specific viral sensing. In a recent study a panel of wild-derived inbred strains was screened for resistance to MCMV infection and identified the PWK/Pas strain as having a resistance phenotype similar to that of B6 mice.38 That is, resistance in PWK/Pas mice was found to be NK cell-dependent and mediated by an autosomal dominant gene that is linked to the NKC region on mouse chromosome 6. The PWK/Pas strain was also found to have a genetic profile in the Ly49H region that is similar to that seen in B6 mice. However, PWK/Pas mice infected with an MCMV mutant lacking the m157 viral glycoprotein remained resistant to MCMV infection. Furthermore, PWK/Pas NK cells failed to lyse m157-expressing target cells.38 These data indicate that MCMV resistance in the PWK/ Pas mouse strain is mediated by a novel resistance locus, named Cmv4, which most likely encodes a yet to be defined activating NK cell receptor.38 Investigations of whether MCMV-encoded class I-like molecules other than m157 serve as ligands for the PWK/Pas NK cell activating receptor will provide useful insights as to how common this mechanism of NK cell recognition of virus-infected cells is. The recent analyses summarized above have provided important evidence that different mouse strains have evolved a range of strategies for NK cell-mediated control of MCMV infection to limit the severity of acute viral infection. An implication of these studies is that a spectrum of similar effector mechanisms is likely to exist in humans as a means to combat HCMV infections. MCMV host resistance genes identified by forward genetics Classical genetics approaches based on systematic mapping and positional cloning of resistance loci that control defined phenotypic traits in inbred strains of mice have provided the basis for many resistance gene identification studies. However, the forward genetics approach has recently been adopted as a powerful alternative strategy to identify host resistance genes (reviewed by Beutler et al.39). This approach is based on the generation of random germline mutants through use of the mutagen N-ethyl-N-nitrosourea (ENU). The forward genetics approach can be used to generate libraries of point mutants that represent a ‘resistome’ to particular infectious agents. For instance, an ENU-induced mutant carrying a mutation in the Tlr9 gene was used to demonstrate that the Toll-like receptor (TLR) TLR9, which signals through the MyD88 adaptor molecule, plays an important role in regulating the level of MCMV infectioninduced IFN-a/b and subsequent NK cell activation.40 A more recent Immunology and Cell Biology

MCMV infection screen of an ENU mutant library has identified eight independent mutants that show differing disease outcomes following MCMV infection.41 One of these mutants, called Domino, was identified as having a point mutation in the DNA binding domain of signal transducer and activator of transcription 1 (STAT1), which is involved in signal transduction via the Type I IFN receptor (IFNAR). These studies illustrate the utility of the forward genetics approach for the identification of key host resistance mechanisms. Despite the utility of this approach, forward genetics will not identify differences owing to allelism at the level of gene sequence or at the level of gene expression. Hence, both classical positional cloning and forward genetics approaches should be considered for defining the contribution of host resistance genes to susceptibility and/or resistance to MCMV infection. EFFECTS OF HOST GENOTYPE ON THE INTERPLAY BETWEEN IMMUNE EFFECTORS Over the last few years, it has become increasingly clear that effective immune responses require interactions between different components of the immune system and the notion that innate and adaptive immunity are independent and standalone has become clearly redundant. The murine model of CMV infection has proven ideal to dissect the complexity of antiviral immunity, and studies conducted over the last 5 years have provided important evidence of novel interactions and interplays that occur between components of the immune system and their contribution to the control of viral infection. Importantly, the availability of the well-characterized mouse models of MCMV resistance and susceptibility have also permitted the analysis of the effects of host genotype on the interplays that take place between immune effectors during the course of MCMV infection and their contribution to viral pathogenesis and disease outcome. As discussed above, NK cells play a critical role in the early control of MCMV infection and NK cell activation induced by ligation of the Ly49H receptor allows B6 mice to effectively control viral replication in the spleen and lung. Our analysis of the effectors involved in the immune response to MCMV infection also identified dendritic cells (DCs) as being important. DCs are the antigen presenting cells par excellence, critical in the initiation and regulation of antigen-specific immunity.42 More recently, it has become clear that DCs can also regulate several aspects of innate immunity.43 In MCMV infection DCs are also a principal target of the virus.44 Importantly, infection of DCs results in early activation followed by impairment of DC function. Early after MCMV infection, DCs become activated and contribute to the generation of both innate and adaptive immune responses.44,45 This phase is followed by viral-induced impairment of DC function.44,46 The early activation of DCs induced by MCMV leads to the production of the inflammatory cytokines IFN-a, IL-12 and IL-18, which contribute to the activation of NK cell responses.45 Indeed, adoptive transfer of DCs activated ex vivo by exposure to MCMV leads to improved control of infection in vivo through the activation of NK cells.45 Interestingly, we also noted further interactions between DCs and NK cells, which vary depending on host genotype. Thus, in MCMV-susceptible BALB/c mice we noted the early disappearance of CD8a+DCs,47 one of the main DC subsets described in the mouse.48 In contrast, MCMV-resistant B6 mice retained the CD8a+DC population during MCMV infection.47 CD8a+DCs have been reported to play a critical role in the generation of virus-specific T-cell responses in a number of viral infection models through their capacity to crosspresent viral antigens.49 Although in MCMV infection, the requirement for cross-presentation remains unclear, because the virus can directly infect DCs, the significance of the differences observed in the

The interplay between host and viral factors AA Scalzo et al 49

fate of CD8a+DCs in MCMV-resistant and susceptible mouse strains remains to be defined. What is clearer is the impact of CD8a+DCs on NK cell responses. Thus, in MCMV-resistant B6 mice, CD8a+DCs are required for the proliferation of the Ly49H+ NK cells that characterizes the late stage of acute infection.47 Whereas the impact of Ly49H+ NK cell proliferation in relation to the generation and/or maintenance of adaptive antiviral immunity is not clear, the current findings demonstrate that activation of NK cells requires input from DCs. Importantly, the available data suggest that host genotype influences the type of interactions that occur between immune effectors, an issue that needs careful consideration in follow up studies. As we will discuss in the following sections, MCMV possesses an armory of viral genes that confer upon the virus a survival advantage. Therefore, it is worth taking into account the possibility that the interactions taking place between immune effectors during MCMV infection may be influenced directly by the virus as an attempt to improve long-term survival and future studies should be designed to address this possibility. CONTRIBUTIONS OF DIVERSITY IN VIRAL GENES TO ENHANCED VIRAL SURVIVAL HCMV has been shown to exhibit extensive genetic variability in a range of genes that code for proteins with diverse functions, including structural envelope glycoproteins, regulatory proteins such as the major immediate early protein (ppUL123), and proteins that contribute to immune evasion. This has recently been comprehensively reviewed by Pignatelli et al.50 It is worth noting that although differences between laboratory strains and clinical isolates are an important issue in HCMV research, differences between viral strains are seldom considered in MCMV research, despite the fact that at least two laboratory strains of virus (Smith and K181) are utilized. Furthermore, with the advent of bacterial artificial chromosome (BAC) mutagenesis, a number of BACs derived from Smith and K181 have been constructed as backbones for further mutagenesis of specific viral genes.51,52 The effects of viral passage in tissue culture and/or in vivo on viral virulence have been clearly established.53,54 The role of cell source, time of harvest and host genotype has also been partly investigated.54 A comprehensive analysis of infectivity and virulence, of at least the two commonly used MCMV strains, after passage in an extended panel of cell types in vitro or in multiple host strains in vivo, remains to be undertaken. The relevance of this type of analysis is brought to notice by recent studies of the role of TLRs in the activation of antiviral immunity. Increased MCMV replication was reported in the spleens of TLR9deficient mice by some,40,55,56 but not all studies. In our studies,45 where we did not observe an effect of TLR9 in MCMV pathogenesis in the spleen, the K181 strain of MCMV was used, and the mice were infected with 104 plaque-forming units (PFU) of virus per mouse. In other studies, where TLR9 was reported to affect MCMV pathogenesis, mice were infected with the Smith strain of MCMV at doses in excess of 5104 PFU. These differences could be due to different viral strains being used, or could be accounted for by the fact that different viral doses, and therefore different numbers of viral particles, were used for infection. The principal role of TLRs is the recognition of foreign particles; it is therefore possible that the increased number of viral units administered when using MCMV-Smith (generally 4105), preferentially activates TLR9 pathways. Alternatively, different viral strains might have different access to intracellular compartments expressing TLR9. As discussed above, these studies highlight the importance of giving careful consideration to variations that may exist between viral strains, with the scope of ultimately identifying the viral gene products involved.

In this review, we will primarily focus on genetic variation in viral genes that target cellular responses of both the innate and adaptive host immune response, and how this is likely to impact the ability of the virus to evade the host immune responses and establish persistent infection. We will also review genetic diversity in viral genes that are targets for antiviral drugs. Viral genes that code for proteins that interact with NK cell receptors or their ligands NK cell-selected mutation in m157. Ly49H+ NK cells have been shown to play a major role in limiting early MCMV replication in B6 mice after activation through recognition of the m157 viral glycoprotein by the Ly49H activating receptor.27,28 The finding that NK cells specifically recognize a virally encoded glycoprotein led to the question of whether Ly49H+ NK cells could exert sufficiently strong selective pressure on the virus to lead to the emergence of viral mutants, capable of escaping Ly49H+NK cell-mediated immune surveillance. Two approaches were used to address this issue. In the first approach the K181 MCMV strain was used to sequentially infect Cmv1r/Ly49H+ mice six times.57 After these sequential infections a number of viral stocks were purified by limiting dilution. Sequence analyses of the m157 gene of viral isolates derived from sequentially passaging the virus in vivo revealed the presence of mutations in the m157 open reading frame resulting in premature stop codons or the loss of several amino-acid residues.57 These mutant m157 molecules were subsequently found not to bind to Ly49H, and the mutant viruses replicated in the spleens of resistant Cmv1r/Ly49H+ mice to the levels equivalent to those of wild-type parental virus in susceptible Cmv1s/Ly49H mice. In subsequent studies, the Smith MCMV strain was used to infect B6 severe combined immunodeficiency disease (SCID) mice.58,59 These mice lacked T and B cells, but had a fully competent NK cell repertoire. At 28 days p.i. B6 SCID mice showed raised viral titers in the spleens. Analysis of viral isolates purified from these mice at this late stage of infection revealed a spectrum of mutations in the m157 gene resulting in premature stop codons or loss/gain of codons coding for specific amino-acid residues. These viruses were also found to be functionally defective in binding to the Ly49H receptor and showed enhanced pathogenicity in vivo in B6 mice. Together, the studies in immunocompetent B6 and immunodeficient B6 SCID mice have revealed that Ly49H+ NK cells can provide sufficient selective pressure to rapidly and specifically drive the emergence of immune escape mutants in MCMV m157. Natural variation in m157. Previous analysis of m157–Ly49 interactions showed that the Smith strain m157 molecule could bind to the Ly49H NK cell activation receptor from B6 mice and to the NK cell inhibitory Ly49I molecule from 129/J mice.27 The MCMV m157 gene in wild-derived isolates showed considerable natural variation, with sequence variants falling into three groups based on the levels of amino-acid identity with the Smith and K181 laboratory strains.57 These groups were either 100% identical or shared identity levels of 80.4–87.9% and 63.9–68.2%, respectively. Apart from the two strains that shared 100% identity (K4 and K17A), only the G1F and N5 m157 molecules, which have the closest identity to the Smith sequence (86.7 and 84.6% identity, respectively) could bind to Ly49H.57 Our recent analyses (Corbett et al., unpublished data) indicate that the m157 sequence variants show distinct patterns of binding to NK cells from different inbred strains. These findings suggest that MCMV has evolved diversity in its m157 sequence to provide it with the ability to bind to NK cell receptors other than the Ly49H activation receptor, Immunology and Cell Biology

The interplay between host and viral factors AA Scalzo et al 50

and that m157 may primarily act as an immunoevasin by interacting and interfering with the functions of NK cells. The concept of host resistance to CMV infection is context dependent. Resistance to MCMV in the B6 mouse strain is governed, at least in part, by the Ly49H-mediated mechanism, which is dependent on specific recognition of infected cells that express the virally encoded m157 molecule. However, infection of the normally MCMVresistant B6 mice by m157 variants that do not bind Ly49H results in high levels of viral replication and increased susceptibility to infection caused by these viral isolates is observed in this mouse strain.57 Variation in other MCMV genes affecting NK cell recognition. Other MCMV genes that affect NK cell function include the m144 gene, which encodes a MHC class I homolog that is capable of binding b2-microglobulin, but not endogenous peptides.60,61 Disruption of the expression the m144 gene attenuated viral growth in vivo and this could be reversed by depletion of NK cells.61 The MCMV m145, m152 and m155 genes encode viral proteins that downmodulate expression of the cellular MULT-1, H60 and RAE-1 stress-related molecules, respectively; these molecules are specifically recognized by the NKG2D NK cell activation receptor.62–66 A recent analysis of variation in the protein sequence of these viral genes (m144, m145, m152 and m155) from wild MCMV isolates indicates that these genes are relatively highly conserved (generally greater than 95% amino-acid identity compared with laboratory strain sequences).67 It is feasible that the high degree of conservation observed may reflect the binding of these molecules to relatively nonpolymorphic host proteins (i.e. MULT-1, H60 and RAE-1) and therefore, the need to evolve is lost. Variation in the HCMV UL18 MHC class I homolog that binds the inhibitory receptor LIR-1. Previous studies have shown that the HCMV UL18 gene encodes a MHC class I-like homolog that binds to b2m and endogenous peptides.68–70 Whereas investigations of the role of UL18 in inhibiting NK cell function have resulted in contradictory outcomes,71–74 it has been definitively shown that UL18 specifically binds to the host cell surface receptor LIR-1 (also known as ILT2/CD85j), an Ig-like receptor expressed by monocytes, T cells, B cells, DCs and NK cells.75 Two recent studies have addressed the issue of what effect the natural variation in UL18 observed in clinical isolates of HCMV has on the interaction between UL18 and the LIR-1 receptor.76,77 These studies showed that sequence variation leads to different affinities of the UL18 molecule for the LIR-1 receptor, as shown by enzyme-linked immunosorbent assay, flow cytometry and Biacore analyses. Variant forms of UL18, expressed as soluble Ig fusion molecules, showed differential capacities to inhibit lysis of BHK-CD64 target cells by LIR-1 transfected NK cells.77 Cerboni et al.76 also found that soluble forms (UL18-Fc) of these UL18 variants differed in their ability to inhibit LIR-1-dependent protection of HLA-G1+ target cells from lysis by LIR-1+ NK cells. Together, these data indicate that sequence variability in HCMV UL18 affects the ability of UL18 to modify NK cell function. What remains to be defined is whether the UL18 sequence variation has functional consequences for HCMV disease outcome in vivo. Viral genes that code for proteins that are recognized by or that modulate the functions of cytotoxic T cells MCMV and ie1. In BALB/c mice, the cytotoxic T-lymphocyte (CTL) response to MCMV is predominantly directed to the H2 Ld-restricted nonameric peptide 168-YPHFMPTNL-176 of the pp89 protein encoded by the ie1 (m123) gene.78–80 Adoptive transfer of pp89specific CTL into naı¨ve BALB/c mice is sufficient to confer protection from challenge with lethal doses of MCMV.81 Analysis of natural Immunology and Cell Biology

sequence variation in the ie1 gene of MCMV identified extensive variation in the nonapeptide sequence that defines the immunodominant epitope.82 Importantly, these sequence variants are not efficiently recognized by CTL specific for the wild-type YPHFMPTNL sequence present in the laboratory strains Smith or K181 and indeed preimmunization with the YPHFMPTNL peptide conferred protection against K181 infection, but did not reduce the pathogenesis of MCMV isolates carrying sequence variants of this nonapeptide.82 The extent of genetic variation in the pp89-encoded nonapeptide from wild-derived MCMV isolates might suggest that these variants have arisen from immune-driven pressure that has selected for ie1 escape mutants. However, we have found that repeated sequential passage of the K181 strain through H2Ld BALB/c mice does not select for mutations in the ie1 gene (Scalzo et al., unpublished observations). This finding suggests that while this protein elicits a host protective CTL response, unlike the rapid Ly49H+ NK cell driven evolution of m157 in B6 mice,57–59 the CTL pressure in BALB/c mice is insufficient to drive the evolution of the ie1 epitope. Alternative explanations are that this region of the pp89 phosphoprotein is functionally conserved or that the wild-type K181 strain sequence in maintained to skew the CTL response in mice with the H2d haplotype. The latter possibility warrants further investigation. HCMV CTL epitopes. The specificity of CTLs specific for HCMV is largely directed towards the HCMV tegument phosphoprotein pp65.83,84 A survey of multiple HLA-A*0201-restricted CTL epitopes in the pp65 protein indicated that they were conserved among different strains of HCMV.85 The UL123-encoded 72 kDa IE1 immediate early protein also represents an important HCMV CTL epitope86,87 and, while natural sequence variation exists in the IE1 epitope (IE1199–206) of clinical isolates, T-lymphocyte responses are able to crossreact with these sequence variants.88 Variation in CMV genes that code for immunoevasins that target MHC class I presentation. In a preliminary analysis of the four HCMV genes (US2, US3, US6 and US11) that regulate the cell surface expression of human MHC class I, and thus antigen presentation to CD8+ T cells, we found that the levels of sequence variation among six clinical isolates and two laboratory strains (AD169 and Towne) were relatively low with the minimum level of amino-acid identity being 96% for US6 (Scalzo and Allan, unpublished data). These data suggest that these genes may be conserved because they bind to relatively structurally conserved regions of MHC class I molecules and hence are not the subject of significant evolutionary pressure. MCMV encodes three immunoevasins that target the MHC class I presentation pathway and inhibit CTL activation. The m06-encoded gp48 and m152-encoded gp40 glycoproteins downregulate surface expression of MHC class I molecules by redirecting them to lysosomal compartments (gp48)89 or by retaining them in the ER/golgi (gp40).90,91 In contrast, the m04-encoded gp34 glycoprotein binds MHC class I molecules in the ER and on the cell surface,92,93 and inhibits CTL activity without blocking MHC class I surface expression.92 The ability of MCMV Smith m04, m06 and m152 to block MHC class I presentation is haplotype dependent94,95 and the effect on CTL recognition varies depending on the CTL epitope tested.96 As MHC haplotypes vary between mouse strains or individual wild mice, it is conceivable that variation in these immunoevasins would confer selective advantages in different host populations. Sequence variation in these genes between wild mouse-derived MCMV isolates has recently been addressed. Both m06 and m152 appear to be relatively

The interplay between host and viral factors AA Scalzo et al 51

conserved between strains tested, whereas remarkable sequence diversity exists for m04, with m04 sequences forming six distinct groups (Smith et al.,67 Corbett et al., unpublished data). The functional implications of this variation are yet to be investigated, but the data suggest selection by different host haplotypes. It has been suggested that the lack of sequence variation in m152 may be due to functional constraints since, in addition to interfering with class I molecules, m152/gp40 also acts to downregulate the RAE-1 family of NKG2D ligands,67 as discussed above. Alternatively, these glycoproteins may bind to relatively structurally conserved regions of MHC class I molecules, and hence may not be the subject of significant evolutionary pressure. The gp34 protein contains a CTL epitope recognized by protective CTLs,97 which is conserved between wild MCMV strains (Corbett et al., unpublished data). The m04, m06 and m152 gene products have both cooperative and antagonistic effects on inhibition of MHC class I surface expression.94,95 Recently, m04 has been suggested to be both a positive and negative regulator of CTL activation98 and the combined action of all three genes may in fact misdirect, rather than block the host–CTL response. Understanding how the sequence variation between different MCMV strains relates to the host MHC haplotype and the individual CTL epitopes presented may be the key to unraveling the apparent contradiction between the MHC class I downregulation induced by these MCMV genes and the active and effective CTL response mounted by the host.99 Genetic variation in viral genes that are targets of antiviral drugs Of the eleven loci necessary for transient complementation of OriLytdependent DNA replication, CMV replication is dependent upon seven key genes involved in reproducing the double-stranded DNA.100–102 These include the CMV DNA polymerase catalytic subunit (HCMV pUL54, MCMV pM54), the DNA polymerase accessory protein (pUL44, pM44), the single-stranded DNA-binding protein (pUL57, pM57), the primase (pUL70, pM70), the helicase (pUL105, pM105), the helicase–primase associated factor (pUL102, pM102) and the OriLyt initiator protein (pUL84).100–102 Genes involved in CMV DNA replication are typically highly invariant in wild-type CMV strains (Scott, unpublished observations). Most currently available anti-CMV agents inhibit CMV DNA replication via different mechanisms of interaction with the CMV DNA polymerase catalytic subunit (pUL54).103 The HCMV-encoded protein kinase (pUL97) is also important for activation and phosphorylation of the most commonly used of these antiviral agents – the nucleoside analog, ganciclovir (GCV).104,105 The DNA polymerase and protein kinase genes are highly conserved in antiviral sensitive (naı¨ve) strains,106–108 and any significant genetic variation in these genes is predominantly the result of antiviral selective pressure.109–112 These patterns of DNA polymerase and protein kinase variation are similar to those seen in the MCMV homologs of these genes (M54 and M97, respectively), both for antiviral-sensitive strains and in vitro generated resistant isolates.113 Therefore, antiviral pressure drives mutation within the genes of replication in an analogous way as host immune pressure drives mutations within other genes. Removal of antiviral pressure either in vivo or in vitro, can lead to the reemergence of wild-type DNA polymerase and protein kinase sequence,114 although emerging evidence suggests certain antiviral resistant genotypes can persist in some individuals in the absence of antiviral pressure.112 The emergence of resistant CMV is also related to the degree of host immune competence; CMV challenge results in increased rates of viral replication in CMV-seronegative versus CMV-seropositive

immunosuppressed transplant recipients115 and good evidence exists that rates of CMV-resistant genotypes are higher in more highly immunosuppressed patients such as bone marrow recipients.116 Kinetic studies also suggest antiviral-resistant strains emerge alongside increased rates of wild-type viral replication in patients receiving subtherapeutic doses of antiviral drugs.114 This suggests host immunosuppression leads to the loss of control of viral replication with the capacity for mutants to emerge increased in such patients.114,115 The antiviral-resistant mutations generally arise within functional domains of the CMV DNA polymerase and protein kinase, interfering with processes such as substrate binding and incorporation, DNA template interaction or phosphotransfer.104,105,117,118 However, these changes are such that normal biological functions of the produced proteins is maintained and viral replication can still occur, albeit at reduced levels for certain mutations.119–121 The significant difference between mutations of DNA polymerase and protein kinase that confer resistance to antiviral drugs, and those occurring in CMV genes under immune pressure, is that mutations will be focused in certain areas because of the functional constraints of these essential replication genes.101,122 Consistent with this is the finding that certain mutations can confer resistance to different drugs of completely different inhibitory mechanisms; for example, a DNA polymerase mutation arising under GCV selective pressure can confer cross-resistance to different classes of nucleoside analogs (cidofovir) as well as the pyrophosphate analog, foscarnet.121 Interestingly, similar mutations are seen in the MCMV genome under selective pressure from the same antivirals as seen in human CMV,113 again consistent with the pressure resulting from functional constraints, rather than hostderived constraints. Conclusions The outcome of the interaction between infectious agents and their hosts is dependent on a variety of intrinsic factors pertaining to the pathogen and host. In the case of CMV the outcome of infection of immunocompetent hosts is usually asymptomatic infection and the establishment of persistence, but disease is pronounced in immunodeficient individuals. Analyses of host resistance genes controlling MCMV infection in mice have revealed a number of strategies that are mouse-strain dependent and, based on forward genetics approaches, likely to be conserved among individual animals in the population. The outcomes of host infection are highly dependent on the balance of host and viral factors that interact to shape infection and thus conclusions based on the use of certain combinations of host and viral strains may provide a limited perspective of the spectrum of biological outcomes. ACKNOWLEDGEMENTS We acknowledge financial support from the NH&MRC, ARC, The Wellcome Trust and Roche Pharmaceuticals. AAS and MADE are supported by fellowships from the NH&MRC. AJC is the recipient of a WA & MG Saw Medical Research Fellowship from the University of Western Australia. We also acknowledge Michael Brown for comments on the manuscript.

1

2 3

Chalmer JE, Mackenzie JS, Stanley NF. Resistance to murine cytomegalovirus linked to the major histocompatibility complex of the mouse. J Gen Virol 1977; 37: 107–114. Selgrade MK, Osborn JE. Role of macrophages in resistance to murine cytomegalovirus. Infect Immun 1974; 10: 1383–1390. Grundy (Chalmer) JE, Mackenzie JS, Stanley NF. Influence of H-2 and non-H-2 genes on resistance to murine cytomegalovirus infection. Infect Immun 1981; 32: 277–286.

Immunology and Cell Biology

The interplay between host and viral factors AA Scalzo et al 52 4

5 6

7 8

9

10

11

12 13

14

15 16 17

18

19

20

21

22

23

24 25

26

27

28

29

30

31

32

Price P, Winter JG, Nikoletti S, Hudson JB, Shellam GR. Functional changes in murine macrophages infected with cytomegalovirus relate to H-2-determined sensitivity to infection. J Virol 1987; 61: 3602–3606. Price P, Gibbons AE, Shellam GR. H-2 class I loci determine sensitivity to MCMV in macrophages and fibroblasts. Immunogenetics 1990; 32: 20–26. Wykes MN, Shellam GR, McCluskey J, Kast WM, Dallas PB, Price P. Murine cytomegalovirus interacts with major histocompatibility complex class I molecules to establish cellular infection. J Virol 1993; 67: 4182–4189. Tay CH, Welsh RM, Brutkiewicz RR. NK cell response to viral infections in b2-microglobulin-deficient mice. J Immunol 1995; 154: 780–789. Polic B, Jonjic S, Pavic I, Crnkovic I, Zorica I, Hengel H et al. Lack of MHC class I complex expression has no effect on spread and control of cytomegalovirus infection in vivo. J Gen Virol 1996; 77: 217–225. Harnett GB, Shellam GR. Variation in murine cytomegalovirus replication in fibroblasts from different mouse strains in vitro: correlation with in vivo resistance. J Gen Virol 1982; 62: 39–47. Desrosiers MP, Kielczewska A, Loredo-Osti JC, Adam SG, Makrigiannis AP, Lemieux S et al. Epistasis between mouse Klra and major histocompatibility complex class I loci is associated with a new mechanism of natural killer cell-mediated innate resistance to cytomegalovirus infection. Nat Genet 2005; 37: 593–599. Dighe A, Rodriguez M, Sabastian P, Xie X, McVoy M, Brown MG. Requisite H2k role in NK cell-mediated resistance in acute murine cytomegalovirus-infected MA/My mice. J Immunol 2005; 175: 6820–6828. Biron CA. Initial and innate responses to viral infections – pattern setting in immunity or disease. Curr Opin Microbiol 1999; 2: 374–381. Grundy (Chalmer) JE, Trapman J, Allan JE, Shellam GR, Melief CJM. Evidence for a protective role of interferon in resistance to murine cytomegalovirus and its control by non-H-2-linked genes. Infect Immun 1982; 37: 143–150. Dalod M, Salazar-Mather TP, Malmgaard L, Lewis C, Asselin-Paturel C, Briere F et al. Interferon alpha/beta and interleukin 12 responses to viral infections: pathways regulating dendritic cell cytokine expression in vivo. J Exp Med 2002; 195: 517–528. Bukowski JF, Warner JF, Dennert G, Welsh RM. Adoptive transfer studies demonstrating the antiviral effect of natural killer cells in vivo. J Exp Med 1985; 161: 40–52. Bukowski JF, Woda BA, Welsh RM. Pathogenesis of murine cytomegalovirus infection in natural killer cell-depleted mice. J Virol 1984; 52: 119–128. Bancroft GJ, Shellam GR, Chalmer JE. Genetic influences on the augmentation of natural killer (NK) cells during murine cytomegalovirus infection: correlation with patterns of resistance. J Immunol 1981; 126: 988–994. Scalzo AA, Fitzgerald NA, Simmons A, La Vista AB, Shellam GR. Cmv-1, a genetic locus that controls murine cytomegalovirus replication in the spleen. J Exp Med 1990; 171: 1469–1483. Scalzo AA, Fitzgerald NA, Wallace CR, Gibbons AE, Smart YC, Burton RC et al. The effect of the Cmv-1 resistance gene, which is linked to the natural killer cell gene complex, is mediated by natural killer cells. J Immunol 1992; 149: 581–589. Brown MG, Dokun AO, Heusel JW, Smith HRC, Beckman DL, Blattenberger EA et al. Vital involvement of a natural killer cell activation receptor in resistance to viral infection. Science 2001; 292: 934–937. Daniels KA, Devora G, Lai WC, O’Donnell CL, Bennett M, Welsh RM. Murine cytomegalovirus is regulated by a discrete subset of natural killer cells reactive with monoclonal antibody to Ly49H. J Exp Med 2001; 194: 29–44. Lee SH, Girard S, Macina D, Busa M, Zafer A, Belouchi A et al. Susceptibility to mouse cytomegalovirus is associated with deletion of an activating natural killer cell receptor of the C-type lectin superfamily. Nat Genet 2001; 28: 42–45. Gosselin P, Mason LH, Willette-Brown J, Ortaldo JR, McVicar DW, Anderson SK. Induction of DAP12 phosphorylation, calcium mobilization and cytokine secretion by Ly49H. J Leuk Biol 1999; 66: 165–171. Smith KM, Wu J, Bakker AB, Phillips JH, Lanier LL. Ly-49D and Ly-49H associate with mouse DAP12 and form activating receptors. J Immunol 1998; 161: 7–10. Lee SH, Zafer A, de Repentigny Y, Kothary R, Tremblay ML, Gros P et al. Transgenic expression of the activating natural killer receptor Ly49H confers resistance to cytomegalovirus in genetically susceptible mice. J Exp Med 2003; 197: 515–526. Sjolin H, Tomasello E, Mousavi-Jazi M, Bartolazzi A, Karre K, Vivier E et al. Pivotal role of KARAP/DAP12 adaptor molecule in the natural killer cell-mediated resistance to murine cytomegalovirus infection. J Exp Med 2002; 195: 825–834. Arase H, Mocarski ES, Campbell AE, Hill AB, Lanier LL. Direct recognition of cytomegalovirus by activating and inhibitory NK cell receptors. Science 2002; 296: 1323–1326. Smith HR, Heusel JW, Mehta IK, Kim S, Dorner BG, Naidenko OV et al. Recognition of a virus-encoded ligand by a natural killer cell activation receptor. Proc Natl Acad Sci USA 2002; 99: 8826–8831. Bubic´ I, Wagner M, Krmpotic´ A, Saulig T, Kim S, Yokoyama WM et al. Gain of virulence caused by loss of a gene in murine cytomegalovirus. J Virol 2004; 78: 7536–7544. Tripathy SK, Smith HR, Holroyd EA, Pingel JT, Yokoyama WM. Expression of m157, a murine cytomegalovirus-encoded putative major histocompatibility class I (MHC-I)like protein, is independent of viral regulation of host MHC-I. J Virol 2006; 80: 545–550. Dokun AO, Kim S, Smith HRC, Kang HSP, Chu DT, Yokoyama WM. Specific and nonspecific NK cell activation during virus infection. Nature Immunol 2001; 2: 951–956. Scalzo AA, Lyons PA, Fitzgerald NA, Forbes CA, Yokoyama WM, Shellam GR. Genetic mapping of Cmv1 in the region of mouse chromosome 6 encoding the NK gene complex-associated loci Ly49 and musNKR-P1. Genomics 1995; 27: 435–441.

Immunology and Cell Biology

33 Anderson SK, Dewar K, Goulet ML, Leveque G, Makrigiannis AP. Complete elucidation of a minimal class I MHC natural killer cell receptor haplotype. Genes Immun 2005; 6: 481–492. 34 Brown MG, Scalzo AA, Stone LR, Clark PY, Du Y, Palanca B et al. Natural killer gene complex (Nkc) allelic variability in inbred mice: evidence for Nkc haplotypes. Immunogenetics 2001; 53: 584–591. 35 Lee SH, Gitas J, Zafer A, Lepage P, Hudson TJ, Belouchi A et al. Haplotype mapping indicates two independent origins for the Cmv1(s) susceptibility allele to cytomegalovirus infection and refines its localization within the Ly49 cluster. Immunogenetics 2001; 53: 501–505. 36 Scalzo AA, Manzur M, Forbes CA, Brown MG, Shellam GR. NK gene complex haplotype variability and host resistance alleles to murine cytomegalovirus in wild mouse populations. Immunol Cell Biol 2005; 83: 144–149. 37 Rodriguez M, Sabastian P, Clark P, Brown MG. Cmv1-independent antiviral role of NK cells revealed in murine cytomegalovirus-infected New Zealand white mice. J Immunol 2004; 173: 6312–6318. 38 Adam SG, Caraux A, Fodil-Cornu N, Loredo-Osti JC, Lesjean-Pottier S, Jaubert J et al. Cmv4, a new locus linked to the NK cell gene complex, controls innate resistance to cytomegalovirus in wild-derived mice. J Immunol 2006; 176: 5478–5485. 39 Beutler B, Crozat K, Koziol JA, Georgel P. Genetic dissection of innate immunity to infection: the mouse cytomegalovirus model. Curr Opin Immunol 2005; 17: 36–43. 40 Tabeta K, Georgel P, Janssen E, Du X, Hoebe K, Crozat K et al. Toll-like receptors 9 and 3 as essential components of innate immune defense against mouse cytomegalovirus infection. Proc Natl Acad Sci USA 2004; 101: 3516–3521. 41 Crozat K, Georgel P, Rutschmann S, Mann N, Du X, Hoebe K et al. Analysis of the MCMV resistome by ENU mutagenesis. Mamm Genome 2006; 17: 398–406. 42 Steinman RM. Some interfaces of dendritic cell biology. Apmis 2003; 111: 675–697. 43 Degli-Esposti MA, Smyth MJ. Close encounters of different kinds: dendritic cells and NK cells take centre stage. Nat Rev Immunol 2005; 5: 112–124. 44 Andrews DM, Andoniou CE, Granucci F, Ricciardi-Castagnoli P, Degli-Esposti MA. Infection of dendritic cells by murine cytomegalovirus induces functional paralysis. Nature Immunol 2001; 2: 1077–1084. 45 Andoniou CE, van Dommelen SL, Voigt V, Andrews DM, Brizard G, Asselin-Paturel C et al. Interaction between conventional dendritic cells and natural killer cells is integral to the activation of effective antiviral immunity. Nature Immunol 2005; 6: 1011–1019. 46 Mathys S, Schroeder T, Ellwart J, Koszinowski UH, Messerle M, Just U. Dendritic cells under influence of mouse cytomegalovirus have a physiologic dual role: to initiate and to restrict T cell activation. J Infect Dis 2003; 187: 988–999. 47 Andrews DM, Scalzo AA, Yokoyama WM, Smyth MJ, Degli-Esposti MA. Functional interactions between dendritic cells and NK cells during viral infection. Nature Immunol 2003; 4: 175–181. 48 Shortman K, Liu YJ. Mouse and human dendritic cell subtypes. Nat Rev Immunol 2002; 2: 151–161. 49 Belz GT, Smith CM, Eichner D, Shortman K, Karupiah G, Carbone FR et al. Conventional CD8-alpha+ dendritic cells are generally involved in priming CTL immunity to viruses. J Immunol 2004; 172: 1996–2000. 50 Pignatelli S, Dal Monte P, Rossini G, Landini MP. Genetic polymorphisms among human cytomegalovirus (HCMV) wild-type strains. Rev Med Virol 2004; 14: 383–410. 51 Redwood AJ, Messerle M, Harvey NL, Hardy CM, Koszinowski UH, Lawson MA et al. Use of a murine cytomegalovirus K181-derived bacterial artificial chromosome as a vaccine vector for immunocontraception. J Virol 2005; 79: 2998–3008. 52 Wagner M, Jonjic S, Koszinowski UH, Messerle M. Systematic excision of vector sequences from the BAC-cloned herpesvirus genome during virus reconstitution. J Virol 1999; 73: 7056–7060. 53 Jordan MC, Takagi JL. Virulence characteristics of murine cytomegalovirus in cell and organ cultures. Infect Immun 1983; 41: 841–843. 54 Selgrade MK, Nedrud JG, Collier AM, Gardner DE. Effects of cell source, mouse strain, and immunosuppressive treatment on production of virulent and attenuated murine cytomegalovirus. Infect Immun 1981; 33: 840–847. 55 Delale T, Paquin A, Asselin-Paturel C, Dalod M, Brizard G, Bates EE et al. MyD88dependent and -independent murine cytomegalovirus sensing for IFN-alpha release and initiation of immune responses in vivo. J Immunol 2005; 175: 6723–6732. 56 Krug A, French AR, Barchet W, Fischer JA, Dzionek A, Pingel JT et al. TLR9dependent recognition of MCMV by IPC and DC generates coordinated cytokine responses that activate antiviral NK cell function. Immunity 2004; 21: 107–119. 57 Voigt V, Forbes CA, Tonkin JN, Degli-Esposti MA, Smith HR, Yokoyama WM et al. Murine cytomegalovirus m157 mutation and variation leads to immune evasion of natural killer cells. Proc Natl Acad Sci USA 2003; 100: 13483–13488. 58 French AR, Pingel JT, Kim S, Yang L, Yokoyama WM. Rapid emergence of escape mutants following infection with murine cytomegalovirus in immunodeficient mice. Clin Immunol 2005; 115: 61–69. 59 French AR, Pingel JT, Wagner M, Bubic I, Yang L, Kim S et al. Escape of mutant double-stranded DNA virus from innate immune control. Immunity 2004; 20: 747–756. 60 Cretney E, Degli-Esposti MA, Densley EH, Farrell HE, Davis-Poynter NJ, Smyth MJ. m144, a murine cytomegalovirus (MCMV)-encoded major histocompatibility complex class I homologue, confers tumor resistance to natural killer cell-mediated rejection. J Exp Med 1999; 190: 435–444. 61 Farrell HE, Vally H, Lynch DM, Fleming P, Shellam GR, Scalzo AA et al. Inhibition of natural killer cells by a cytomegalovirus MHC class I homologue in vivo. Nature 1997; 386: 510–514.

The interplay between host and viral factors AA Scalzo et al 53 62 Hasan M, Krmpotic A, Ruzsics Z, Bubic I, Lenac T, Halenius A et al. Selective downregulation of the NKG2D ligand H60 by mouse cytomegalovirus m155 glycoprotein. J Virol 2005; 79: 2920–2930. 63 Krmpotic´ A, Busch DH, Bubic´ I, Gebhardt F, Hengel H, Hasan M et al. MCMV glycoprotein gp40 confers virus resistance to CD8+ T cells and NK cells in vivo. Nature Immunol 2002; 3: 529–535. 64 Krmpotic A, Hasan M, Loewendorf A, Saulig T, Halenius A, Lenac T et al. NK cell activation through the NKG2D ligand MULT-1 is selectively prevented by the glycoprotein encoded by mouse cytomegalovirus gene m145. J Exp Med 2005; 201: 211–220. 65 Lodoen M, Ogasawara K, Hamerman JA, Arase H, Houchins JP, Mocarski ES et al. NKG2D-mediated natural killer cell protection against cytomegalovirus is impaired by viral gp40 modulation of retinoic acid early inducible 1 gene molecules. J Exp Med 2003; 197: 1245–1253. 66 Lodoen MB, Abenes G, Umamoto S, Houchins JP, Liu F, Lanier LL. The cytomegalovirus m155 gene product subverts natural killer cell antiviral protection by disruption of H60-NKG2D interactions. J Exp Med 2004; 200: 1075–1081. 67 Smith LM, Shellam GR, Redwood AJ. Genes of murine cytomegalovirus exist as a number of distinct genotypes. Virology 2006; 352: 450–465. 68 Beck S, Barrell BG. Human cytomegalovirus encodes a glycoprotein homologous to MHC class-I antigens. Nature 1988; 331: 269–272. 69 Browne H, Smith G, Beck S, Minson T. A complex between the MHC class I homologue encoded by human cytomegalovirus and b2 microglobulin. Nature 1990; 347: 770–772. 70 Fahnestock ML, Johnson JL, Feldman RMR, Neveu JM, Lane WS, Bjorkman PJ. The MHC class I homolog encoded by human cytomegalovirus binds endogenous peptides. Immunity 1995; 3: 583–590. 71 Leong CC, Chapman TL, Bjorkman PJ, Formankova D, Mocarski ES, Phillips JH et al. Modulation of natural killer cell cytotoxicity in human cytomegalovirus infection – the role of endogenous class i major histocompatibility complex and a viral class I homolog. J Exp Med 1998; 187: 1681–1687. 72 Reyburn HT, Mandelboim O, Vales-Gomez M, Davis DM, Pazmany L, Strominger JL. The class I MHC homologue of human cytomegalovirus inhibits attack by natural killer cells. Nature 1997; 386: 514–517. 73 Saverino D, Ghiotto F, Merlo A, Bruno S, Battini L, Occhino M et al. Specific recognition of the viral protein UL18 by CD85j/LIR-1/ILT2 on CD8+ T cells mediates the non-MHC-restricted lysis of human cytomegalovirus-infected cells. J Immunol 2004; 172: 5629–5637. 74 Vitale M, Castriconi R, Parolini S, Pende D, Hsu ML, Moretta L et al. The leukocyte Iglike receptor (LIR)-1 for the cytomegalovirus UL18 protein displays a broad specificity for different HLA class I alleles: analysis of LIR-1(+) NK cell clones. Int Immunol 1999; 11: 29–35. 75 Cosman D, Fanger N, Borges L. Human cytomegalovirus, MHC class I and inhibitory signalling receptors: more questions than answers. Immunol Rev 1999; 168: 177–185. 76 Cerboni C, Achour A, Warnmark A, Mousavi-Jazi M, Sandalova T, Hsu ML et al. Spontaneous mutations in the human CMV HLA class I homologue UL18 affect its binding to the inhibitory receptor LIR-1/ILT2/CD85j. Eur J Immunol 2006; 36: 732–741. 77 Vales-Gomez M, Shiroishi M, Maenaka K, Reyburn HT. Genetic variability of the major histocompatibility complex class I homologue encoded by human cytomegalovirus leads to differential binding to the inhibitory receptor ILT2. J Virol 2005; 79: 2251–2260. 78 Del Val M, Schlicht H-J, Volkmer H, Messerle M, Reddehase MJ, Koszinowski UH. Protection against lethal cytomegalovirus infection by a recombinant vaccine containing a single nonameric T-cell epitope. J Virol 1991; 65: 3641–3646. 79 Del Val M, Volkmer H, Rothbard JB, Jonjic S, Messerle M, Schickedanz J et al. Molecular basis for cytolytic T-lymphocyte recognition of the murine cytomegalovirus immediate-early protein pp89. J Virol 1988; 62: 3965–3972. 80 Reddehase MJ, Rothbard JB, Koszinowski UH. A pentapeptide as minimal antigenic determinant for MHC class I-restricted T lymphocytes. Nature 1989; 337: 651–653. 81 Jonjic S, Del Val M, Keil GM, Reddehase MJ, Koszinowski UH. A nonstructural viral protein expressed by a recombinant vaccinia virus protects against lethal cytomegalovirus infection. J Virol 1988; 62: 1653–1658. 82 Lyons PA, Allan JE, Carrello C, Shellam GR, Scalzo AA. Effect of natural sequence variation at the H-2Ld-restricted CD8+ T cell epitope of the murine cytomegalovirus ie1-encoded pp89 on T cell recognition. J Gen Virol 1996; 77: 2615–2623. 83 Riddell SR, Rabin M, Geballe AP, Britt WJ, Greenberg PD. Class I MHC-restricted cytotoxic T lymphocyte recognition of cells infected with human cytomegalovirus does not require endogenous viral gene expression. J Immunol 1991; 146: 2795–2804. 84 Wills MR, Carmichael AJ, Mynard K, Jin X, Weekes MP, Plachter B et al. The human cytotoxic T-lymphocyte (CTL) response to cytomegalovirus is dominated by structural protein pp65: frequency, specificity and T-cell receptor usage of pp65-specific CTL. J Virol 1996; 70: 7569–7579. 85 Solache A, Morgan CL, Dodi AI, Morte C, Scott I, Baboonian C et al. Identification of three HLA-A*0201-restricted cytotoxic T cell epitopes in the cytomegalovirus protein pp65 that are conserved between eight strains of the virus. J Immunol 1999; 163: 5512–5518. 86 Gyulai Z, Endresz V, Burian K, Pincus S, Toldy J, Cox WI et al. Cytotoxic T lymphocyte (CTL) responses to human cytomegalovirus pp65, IE1-Exon4, gB, pp150, and pp28 in healthy individuals: reevaluation of prevalence of IE1-specific CTLs. J Infect Dis 2000; 181: 1537–1546.

87 Khan N, Cobbold M, Keenan R, Moss PA. Comparative analysis of CD8+ T cell responses against human cytomegalovirus proteins pp65 and immediate early 1 shows similarities in precursor frequency, oligoclonality, and phenotype. J Infect Dis 2002; 185: 1025–1034. 88 Prod’homme V, Retiere C, Valtcheva R, Bonneville M, Hallet MM. Cross-reactivity of HLA-B*1801-restricted T-lymphocyte clones with target cells expressing variants of the human cytomegalovirus 72kDa-IE1 protein. J Virol 2003; 77: 7139–7142. 89 Reusch U, Muranyi W, Lucin P, Burgert HG, Hengel H, Koszinowski UH. A cytomegalovirus glycoprotein re-routes MHC class I complexes to lysosomes for degradation. EMBO J 1999; 18: 1081–1091. 90 Ziegler H, Muranyi W, Burgert HG, Kremmer E, Koszinowski UH. The luminal part of the murine cytomegalovirus glycoprotein gp40 catalyzes the retention of MHC class I molecules. EMBO J 2000; 19: 870–881. 91 Ziegler H, Thale R, Lucin P, Muranyi W, Flohr T, Hengel H et al. A mouse cytomegalovirus glycoprotein retains MHC class I complexes in the ERGIC/cis-Golgi compartments. Immunity 1997; 6: 57–66. 92 Kavanagh DG, Koszinowski UH, Hill AB. The murine cytomegalovirus immune evasion protein m4/gp34 forms biochemically distinct complexes with class I MHC at the cell surface and in a pre-Golgi compartment. J Immunol 2001; 167: 3894–3902. 93 Kleijnen MF, Huppa JB, Lucin P, Mukherjee S, Farrell H, Campbell AE et al. A mouse cytomegalovirus glycoprotein, gp34, forms a complex with folded class I MHC molecules in the ER which is not retained but is transported to the cell surface. EMBO J 1997; 16: 685–694. 94 Kavanagh DG, Gold MC, Wagner M, Koszinowski UH, Hill AB. The multiple immuneevasion genes of murine cytomegalovirus are not redundant: m4 and m152 inhibit antigen presentation in a complementary and cooperative fashion. J Exp Med 2001; 194: 967–978. 95 Wagner M, Gutermann A, Podlech J, Reddehase MJ, Koszinowski UH. Major histocompatibility complex class I allele-specific cooperative and competitive interactions between immune evasion proteins of cytomegalovirus. J Exp Med 2002; 196: 805–816. 96 Holtappels R, Grzimek NK, Simon CO, Thomas D, Dreis D, Reddehase MJ. Processing and presentation of murine cytomegalovirus pORFm164-derived peptide in fibroblasts in the face of all viral immunosubversive early gene functions. J Virol 2002; 76: 6044–6053. 97 Holtappels R, Thomas D, Podlech J, Geginat G, Steffens HP, Reddehase MJ. The putative natural killer decoy early gene m04 (gp34) of murine cytomegalovirus encodes an antigenic peptide recognized by protective antiviral CD8T cells. J Virol 2000; 74: 1871–1884. 98 Holtappels R, Gillert-Marien D, Thomas D, Podlech J, Deegen P, Herter S et al. Cytomegalovirus encodes a positive regulator of antigen presentation. J Virol 2006; 80: 7613–7624. 99 Gold MC, Munks MW, Wagner M, McMahon CW, Kelly A, Kavanagh DG et al. Murine cytomegalovirus interference with antigen presentation has little effect on the size or the effector memory phenotype of the CD8 T cell response. J Immunol 2004; 172: 6944–6953. 100 Colletti KS, Xu Y, Cei SA, Tarrant M, Pari GS. Human cytomegalovirus UL84 oligomerization and heterodimerization domains act as transdominant inhibitors of oriLyt-dependent DNA replication: evidence that IE2-UL84 and UL84-UL84 interactions are required for lytic DNA replication. J Virol 2004; 78: 9203–9214. 101 Pari GS, Anders DG. Eleven loci encoding trans-acting factors are required for transient complementation of human cytomegalovirus oriLyt-dependent DNA replication. J Virol 1993; 67: 6979–6988. 102 Xu Y, Cei SA, Huete AR, Pari GS. Human cytomegalovirus UL84 insertion mutant defective for viral DNA synthesis and growth. J Virol 2004; 78: 10360–10369. 103 De Clercq E. Antiviral drugs: current state of the art. J Clin Virol 2001; 22: 73–89. 104 Littler E, Stuart AD, Chee MS. Human cytomegalovirus UL97 open reading frame encodes a protein that phosphorylates the antiviral nucleoside analogue ganciclovir. Nature 1992; 358: 160–162. 105 Sullivan V, Talarico CL, Stanat SC, Davis M, Coen DM, Biron KK. A protein kinase homologue controls phosphorylation of ganciclovir in human CytomegalovirusInfected cells. Nature 1992; 358: 162–164. 106 Chou S, Erice A, Jordan MC, Vercellotti GM, Michels KR, Talarico CL et al. Analysis of the UL97 phosphotransferase coding sequence in clinical cytomegalovirus isolates and identification of mutations conferring ganciclovir resistance. J Infect Dis 1995; 171: 576–583. 107 Chou S, Lurain NS, Weinberg A, Cai GY, Sharma PL, Crumpacker CS. Interstrain variation in the human cytomegalovirus DNA polymerase sequence and its effect on genotypic diagnosis of antiviral drug resistance. Adult AIDS clinical trials group CMV laboratories. Antimicrob Agents Chemother 1999; 43: 1500–1502. 108 Fillet AM, Auray L, Alain S, Gourlain K, Imbert BM, Najioullah F et al. Natural polymorphism of cytomegalovirus DNA polymerase lies in two nonconserved regions located between domains delta-C and II and between domains III and I. Antimicrob Agents Chemother 2004; 48: 1865–1868. 109 Alain S, Hantz S, Scieux C, Karras A, Mazeron MC, Szelag JC et al. Detection of ganciclovir resistance after valacyclovir-prophylaxis in renal transplant recipients with active cytomegalovirus infection. J Med Virol 2004; 73: 566–573. 110 Chou S. Antiviral drug resistance in human cytomegalovirus. Transpl Infect Dis 1999; 1: 105–114. 111 Gilbert C, Boivin G. Human cytomegalovirus resistance to antiviral drugs. Antimicrob Agents Chemother 2005; 49: 873–883.

Immunology and Cell Biology

The interplay between host and viral factors AA Scalzo et al 54 112 Scott GM, Isaacs MA, Zeng F, Kesson AM, Rawlinson WD. Cytomegalovirus antiviral resistance associated with treatment induced UL97 (protein kinase) and UL54 (DNA polymerase) mutations. J Med Virol 2004; 74: 85–93. 113 Scott GM, Ng HL, Morton CJ, Parker MW, Rawlinson WD. Murine cytomegalovirus resistant to antivirals has genetic correlates with human cytomegalovirus. J Gen Virol 2005; 86: 2141–2151. 114 Emery VC, Griffiths PD. Prediction of cytomegalovirus load and resistance patterns after antiviral chemotherapy. Proc Natl Acad Sci USA 2000; 97: 8039–8044. 115 Emery VC, Hassan-Walker AF, Burroughs AK, Griffiths PD. Human cytomegalovirus (HCMV) replication dynamics in HCMV-naive and -experienced immunocompromised hosts. J Infect Dis 2002; 185: 1723–1728. 116 Emery VC. Investigation of CMV disease in immunocompromised patients. J Clin Pathol 2001; 54: 84–88. 117 Tchesnokov EP, Gilbert C, Boivin G, Gotte M. Role of helix P of the human cytomegalovirus DNA polymerase in resistance and hypersusceptibility to the antiviral drug foscarnet. J Virol 2006; 80: 1440–1450.

Immunology and Cell Biology

118 Ye LB, Huang ES. In vitro expression of the human cytomegalovirus DNA polymerase gene: effects of sequence alterations on enzyme activity. J Virol 1993; 67: 6339–6347. 119 Baldanti F, Underwood MR, Stanat SC, Biron KK, Chou S, Sarasini A et al. Single amino acid changes in the DNA polymerase confer foscarnet resistance and slowgrowth phenotype, while mutations in the UL97-encoded phosphotransferase confer ganciclovir resistance in three double-resistant human cytomegalovirus strains recovered from patients with AIDS. J Virol 1996; 70: 1390–1395. 120 Cihlar T, Fuller MD, Mulato AS, Cherrington JM. A point mutation in the human cytomegalovirus DNA polymerase gene selected in vitro by cidofovir confers a slow replication phenotype in cell culture. Virology 1998; 248: 382–393. 121 Scott GM, Weinberg A, Rawlinson WD, Chou S. Multidrug resistance conferred by novel DNA polymerase mutations in human cytomegalovirus isolates. Antimicrob Agents Chemother 51(1): accepted 16 October 2006; doi:10.1128/AAC.00633-06. 122 Prichard MN, Gao N, Jairath S, Mulamba G, Krosky P, Coen DM et al. A recombinant human cytomegalovirus with a large deletion in UL97 has a severe replication deficiency. J Virol 1999; 73: 5663–5670.