Skip to main content
Journal of Virology logoLink to Journal of Virology
. 2006 Oct;80(19):9349–9360. doi: 10.1128/JVI.01122-06

Pathogenesis of West Nile Virus Infection: a Balance between Virulence, Innate and Adaptive Immunity, and Viral Evasion

Melanie A Samuel 1, Michael S Diamond 1,2,3,*
PMCID: PMC1617273  PMID: 16973541

West Nile virus (WNV) is a neurotropic flavivirus that has emerged globally as a significant cause of viral encephalitis. WNV is maintained in an enzootic cycle between mosquitoes and birds (reviewed in reference 75) but can also infect and cause disease in horses and other vertebrate animals. Infection of humans is associated with a febrile illness that can progress to a lethal encephalitis with symptoms including cognitive dysfunction and flaccid paralysis (30, 148, 167). Since the mid-1990s, outbreaks of WNV fever and encephalitis have occurred in regions throughout the world where WNV is endemic, including the Middle East, Europe, and Africa (43). Following its introduction into the United States in 1999, WNV rapidly disseminated across North America and more recently has been reported in Mexico, South America, and the Caribbean (45, 95, 102). Annual outbreaks of WNV fever and neuroinvasive disease occur in the United States (67), with ∼19,000 diagnosed human cases between 1999 and 2005 (http://www.cdc.gov/ncidod/dvbid/westnile/surv&control.htm#maps) and an estimated 750,000 undiagnosed infections in 2003 (25). Although vaccines are available for animal use, no vaccines or specific therapies for WNV are currently approved for humans.

WNV is a member of the Flaviviridae family of RNA viruses and is related to other important human pathogens, including dengue viruses, yellow fever viruses, and Japanese encephalitis viruses. Similar to other flaviviruses, WNV is an enveloped virus with a single-stranded, positive sense, ∼11-kb RNA genome. The genome is transcribed as a single polyprotein that is cleaved by host and viral proteases into three structural and seven nonstructural proteins (32). The structural proteins include a capsid protein (C) that binds viral RNA, a premembrane (prM) protein that blocks premature viral fusion and may chaperone envelope (E)-protein folding, and an E protein that mediates viral attachment, membrane fusion, and viral assembly (131). The viral nonstructural proteins (NS1, NS2A, NS2B, NS3, NS4A NS4B, and NS5) regulate viral transcription and replication and attenuate host antiviral responses (17, 71, 87, 110, 111, 115, 132). NS1 has cofactor activity for the viral replicase and is secreted from infected cells, NS2A inhibits interferon (IFN) responses and may participate in virus assembly, and NS3 has protease, NTPase, and helicase activities (87, 113-115, 118). NS2B is a cofactor required for NS3 proteolytic activity, NS4A and NS4B modulate IFN signaling, and NS5 encodes the RNA-dependent RNA polymerase and a methyltransferase (49, 86, 115, 202).

Several molecules have been implicated in the attachment of WNV to cells in vitro, including DC-SIGN, DC-SIGN-R, and the integrin αvβ3 (37, 44). However, attachment receptors in vivo for physiologically important cell types, such as neurons, remain uncharacterized. After binding, WNV enters cells via receptor-mediated endocytosis that may involve clathrin-coated pits (36, 63). Following a pH-dependent conformational change in the E protein, the viral and endosomal membranes fuse, releasing the viral nucleocapsid into the cytoplasm (2, 23, 65). WNV is believed to replicate on endoplasmic reticulum-associated membranes to generate a negative-strand RNA intermediate that serves as a template for nascent positive-strand RNA synthesis (119). Positive-strand RNA is either packaged within progeny virions or used to translate additional viral proteins. WNV assembles and buds into the endoplasmic reticulum to form immature particles that contain the prM protein. Following transport through the trans-Golgi network, furin-mediated cleavage of prM to M generates mature, infectious virions that are released by exocytosis (51, 70).

The emergence of WNV in the western hemisphere and the corresponding increase in disease burden have been accompanied by intensive study. Epidemiological analyses indicate that the elderly and immunocompromised are at greatest risk of developing severe neurological disease (20, 35, 67, 135, 179). Experimental studies have made significant progress in dissecting the viral and host factors that determine the pathogenesis and outcome of WNV infection. This review will focus on the interactions between WNV pathogenesis and the innate and adaptive host immune responses.

PATHOGENESIS OF WNV INFECTION

Clinical manifestations.

Seroprevalence studies suggest that while the majority of WNV infections are asymptomatic, approximately 20 to 30% of infected individuals develop flu-like clinical manifestations characterized as WNV fever (148, 189). In a subpopulation of individuals (approximately 1 in 150), a neuroinvasive disease develops (149, 190). The clinical features of severe WNV infection vary and include severe headache, ocular manifestations, muscle weakness, cognitive impairment, tremors, and a poliomyelitis-like flaccid paralysis (8, 30, 148, 167, 168). The mortality rate following neuroinvasive infection is approximately 10% (138, 148), and long-term neurological sequelae are common (>50%) (92, 166). Neuronal damage is most prevalent in the brain stem and anterior-horn neurons of the spinal cord, although in immunosuppressed individuals infection can disseminate throughout the central nervous system (CNS) (69, 94).

Viral dissemination and pathogenesis in vivo.

Rodent models have provided insight into the mechanisms of WNV dissemination and pathogenesis (Table 1). Following peripheral inoculation, initial WNV replication is thought to occur in skin Langerhans dendritic cells (26). These cells migrate to and seed draining lymph nodes, resulting in a primary viremia and subsequent infection of peripheral tissues such as the spleen and kidney. By the end of the first week, WNV is largely cleared from the serum and peripheral organs and infection of the CNS is observed in a subset of immunocompetent animals. Rodents that succumb to infection develop a CNS pathology similar to that observed in human WNV cases, including infection and injury of brain stem, hippocampal, and spinal cord neurons (46, 50, 55, 140, 172, 195). WNV infection is not significantly detected in nonneuronal CNS cell populations in humans or animals. In surviving wild-type mice, WNV is cleared from all tissue compartments within 2 to 3 weeks after infection. However, persistent viral infection in the brains of CD8+ T-cell (171)- or perforin-deficient mice (173) and in the brains and kidneys of infected hamsters has been reported (178, 195). Persistent infection has also been documented in a WNV-infected immunosuppressed patient in whom viremia was detected for more than 60 days (22). Although less in known about WNV pathogenesis in avian hosts, the virus has been detected by histology, reverse transcription-PCR, and virologic assays in the brains, livers, lungs, spleens, hearts, and kidneys of naturally infected crows and blue jays (59, 144, 194).

TABLE 1.

WNV infection in immunodeficient mice after peripheral infection

Immune function and mouse straina Viral strain (lineage) % Lethality Virologic and immune characterizationb Reference(s)
WT
    C57BL/6 NY2000 (I) 30-35 V—LN, S, SPL, CNS; I—innate, T, B 46, 47, 93, 125, 126, 161, 171, 173, 174
    129 Sv/Ev NY2000 (I) 60 V—LN, S, SPL, CNS; I—innate, T, B 126, 160
    C57BL/6 × 129 Sv/Ev NY2000 (I) 50 V—LN, S, SPL, CNS; I—B 126
    BALB/C (5-6 wk) NY99 NY2000 (I) 80 V—S, SPL, CNS, KD, HT 97, 170
    ICR WN25 (I) 0 V—none in S or CNS 72
IFN
    IFN-α/βR−/− (129 Sv/Ev) NY2000 (I) 100 V—↑in all tissues 160
    IFN-γ−/− or IFN-γR−/− NY2000 (I) 95 V—↑in S, LN, SPL, early CNS entry; I—IFN-γ+ γδ T cells,↓viral loads 174
    IFN-γ−/− 2741 (I) 90 No data reported 185
    RNase L−/− NY2000 (I) 51 V—↑in LN and SPL 161
    PKR−/− × RNase L−/− NY2000 (I) 90 V—↑in LN and SPL, early CNS entry 161
Complement
    C3−/− (C57BL/6 × 129 Sv/Ev) NY2000 (I) 100 V—↑in CNS; I—↓IgM and IgG 126
    CR1/CR2−/− NY2000 (I) 90 V—↑in CNS; I—↓IgM and IgG 126
    C4−/− NY2000 (I) 100 V—↓in SPL,↑in CNS; I—↓IgM and IgG;↓ T-cell responses 125
    C1q−/− NY2000 (I) 83 V—no SPL infection,↑in CNS; I—delayed IgG production 125
    fB−/− NY2000 (I) 96 V—↑in SPL and CNS, early CNS entry; I—↓ T-cell response 125
    fD−/− NY2000 (I) 74 V—↑in SPL and CNS, early CNS entry 125
    C1q−/− × fD−/− NY2000 (I) 90 No data reported 125
    C5aR−/− NY2000 (I) 40 No Δ in lethality 125
Other innate immunity effectors
    Flv resistance gene carriers IS-98-ST1(I) 0 V—↓in CNS, decreased viral spread 122
B-cell function
    μMT (B cell deficient) NY2000 (I) 100 V—↑in S, SPL, and CNS; I—protected by immune serum 46
    Secretory IgM−/− NY2000 (I) 100 V—↑in S and CNS; I—↓IgG responses, protected by transfer of immune IgM 47
T-cell function
    CD8 α chain−/− NY2000 (I) 84 V—↑in CNS,↓clearance in SPL, early CNS infection, persistent virus in brain 171
    TCR β chain−/− 2741 (I) 90 V—↑in SPL and brain 185
    β2-Microglobulin−/− Sarafend (II) 80 V—early CNS infection↑in brain 187
    MHC-Ia−/− NY2000 (I) 88 V—persistent virus in brain 171
    Perforin−/− NY2000 (I) 78 V—↑in SPL and CNS, persistent virus in brain; I—transfer of WT CD8+ T cells↓viral load 173
    Perforin−/− Sarafend (II) 18 No Δ in lethality 188
    Granzyme A and B−/− Sarafend (II) 70 V—↑in CNS 188
    Perforin−/− × granzyme A and B−/− Sarafend (II) 50 V—↑in CNS 188
    Fas−/− Sarafend (II) 32 No Δ in lethality 188
    Gld (Fas ligand−/−) Sarafend (II) 35 No Δ in lethality 188
    Perforin−/− × Gld Sarafend (II) 62 V—↑in CNS 188
    TCR δ−/− (γδ T cell deficient) 2741 (I) 100 V—↑in S, SPL, and CNS, early spread to CNS; I—protected by transfer of IFN-γ-sufficient γδ T cells, role in priming adaptive immunity 184, 185
Chemokines and chemokine receptors
    CXCL10−/− NY2000 (I) 90 V—↑in CNS; I—↓CD8+ and CD4+ T-cell trafficking to CNS 93
    CCR5−/− NY1999 (I) 100 V—↑in CNS; I—↓leukocyte trafficking to CNS, transfer of WT CD8+ T cells improved survival 61
    CCR1−/− NY1999 (I) ∼40 No Δ in lethality 61
    CX3CR1−/− NY1999 (I) ∼40 No Δ in lethality 61
Multiple deficiencies
    RAG (B and T cell deficient) NY2000 (I) 100 V—↑in S, SPL, and CNS 46
    SCID (ICR) WN25 (I) ND V—persistent viremia and↑in brain 72
Cell depletion studies
    Myeloid cells WN25 (I) 75 V—↑in S and entry into CNS 15
    NK cells NY2000 (I) ∼35 No Δ in lethality 173
    CD8+ T cells Sarafend (II) 70 V—↑in brain 187
a

C57BL/6 unless otherwise noted.

b

V, virologic; I, immune; LN, lymph node; S, serum; SPL, spleen; KD, kidney; HT, heart; ↑, increase; ↓, decrease; Δ, change; T, T cell; B, B cell.

The mechanisms by which WNV and other neurotropic flaviviruses cross the blood-brain barrier (BBB) remain largely unknown, although tumor necrosis factor alpha (TNF-α)-mediated changes in endothelial cell permeability may facilitate CNS entry (186). It is likely that WNV infects the CNS at least in part via hematogenous spread, as an increased viral burden in serum correlates with earlier viral entry into the brain (47, 160). Additional mechanisms may contribute to WNV CNS infection, including (i) infection or passive transport through the endothelium or choroid plexus epithelial cells (98), (ii) infection of olfactory neurons and spread to the olfactory bulb (129), (iii) a “Trojan horse” mechanism in which the virus is transported by infected immune cells that traffic to the CNS (58), and (iv) direct axonal retrograde transport from infected peripheral neurons (78, 81). Although the precise mechanisms of WNV CNS entry in humans require additional study, changes in cytokine levels that may modulate BBB permeability and infection of blood monocytes and choroid plexus cells have been documented in animal models (58, 160, 186).

INNATE IMMUNE RESPONSES TO WNV INFECTION

Interferons.

Type I (IFN-α and IFN-β), type II (IFN-γ), and type III (IFN-λ) IFNs act as important innate immune system controls of viral infections (reviewed in references 147 and 151). IFN-α/β is produced by most cell types following virus infection and induces an antiviral state by upregulating genes with direct and indirect antiviral functions. IFN-α/β also links innate and adaptive immune responses through stimulation of dendritic-cell maturation (7), by direct activation of B and T cells (105), and by preventing the death of recently activated T cells (121). Pretreatment with IFN-α/β inhibits WNV replication in vitro, but treatment after infection is much less effective (3, 41, 160). Although WNV can directly antagonize IFN-induced responses after infection (71, 114, 115), type I IFN is still required to restrict WNV replication and spread (160). Mice lacking the IFN-α/β receptor (IFN-α/βR) had uncontrolled viral replication, rapid dissemination to the CNS, and enhanced lethality. Altered viral tropism in IFN-α/βR−/− mice was also observed with enhanced infection in normally resistant cell populations and peripheral tissues.

IFN-γ is produced primarily by γδ T cells, CD8+ T cells, and natural killer (NK) cells and limits virus infection through several mechanisms. IFN-γ restricts viral replication directly by inducing an antiviral state or indirectly by modulating the adaptive immune response through activation of myeloid cells, inducing CD4+ T-cell activation and TH1-TH2 polarization, and increasing cell surface expression of major histocompatibility complex (MHC) class I molecules (34, 165). Although WNV is also resistant to the antiviral effects of IFN-γ after infection in vitro (160), in vivo IFN-γ limits early viral dissemination to the CNS: mice deficient in either IFN-γ or the IFN-γ receptor showed a greater peripheral viral burden, earlier entry into the CNS, and increased lethality (174, 185). Interestingly, no major deficits in adaptive immune responses were observed in these studies, suggesting that the dominant function of IFN-γ in controlling WNV infection is innate and antiviral. Additional experiments demonstrated a cell-specific requirement for IFN-γ, as γδ T cells utilized IFN-γ to limit WNV dissemination whereas CD8+ T cells did not (174, 185).

Viral sensors and IFN-induced effector molecules.

Cells recognize and respond to RNA virus infection through several nucleic acid sensors, including Toll-like receptor 3 (TLR3) and the cytoplasmic double-stranded RNA (dsRNA) sensors retinoic acid-inducible gene I (RIG-I) and melanoma differentiation-associated gene 5 (MDA5) (10, 83, 200, 201). Binding of RNA to these pathogen recognition receptors results in downstream activation of transcription factors such as IFN regulatory factor 3 (IRF3) and IRF7 and expression of IFN-stimulated genes. An emerging literature suggests that RIG-I and IRF3 have essential functions in responding to WNV infection. Murine embryonic fibroblasts deficient in RIG-I demonstrated delayed induction of host responses, decreased IRF3 activation, and augmented viral replication (56, 57). In contrast, MDA5 may be less essential for cellular recognition of WNV, as IFN production by MDA5-deficient dendritic cells remained largely intact after WNV infection (60). Unlike RIG-I and MDA5, TLR3 is expressed primarily in endosomes and activates IRF3 downstream of the kinases TBK1 and IKKɛ (54, 124, 169). Although WNV appears to interfere with poly(I-C)-induced IFN responses (164), initial studies suggest that TLR3 is somewhat dispensable for recognition of WNV in some cell types in vitro (56). TLR3−/− mice paradoxically showed decreased lethality after intraperitoneal infection, presumably because of blunted cytokine responses (e.g., TNF-α), which apparently facilitate WNV entry into the CNS (186). Nonetheless, it remains possible that TLR3 contributes to WNV recognition and IFN induction in specific immune cell populations, such as macrophages and dendritic cells.

Activation of pathogen recognition receptors stimulates IFN production and feedback amplification of the IFN-stimulated gene response. Studies have begun to elucidate the specific antiviral effector molecules that control WNV infection. Recent reports indicate that the dsRNA-dependent protein kinase (PKR) and 2′-5′-oligoadenylate synthase (OAS) proteins mediate intracellular resistance to WNV. PKR is activated by binding dsRNA and phosphorylates eukaryotic translation initiation factor 2α, resulting in attenuation of protein synthesis (127). RNase L is activated by 2′-5′-linked oligoadenylates synthesized by OAS enzymes and functions as an endoribonuclease that cleaves viral and host RNAs (175). RNase L−/− murine embryonic fibroblasts and PKR−/− × RNase L−/− bone marrow-derived macrophages supported increased WNV replication in vitro (161, 162). Moreover, mice deficient in both PKR and RNase L showed increased lethality following WNV infection, with greater viral loads in peripheral tissues at early time points after infection (161). Flavivirus susceptibility in mice has been mapped to a mutation in the Oas1b gene that results in the expression of a truncated OAS isoform (122, 146). However, the mechanisms by which Oas gene alleles affect flavivirus pathogenesis remain uncertain; recent reports suggest that Oas1b gene effects on WNV replication are independent of RNase L (161, 162).

Complement.

The complement system is a family of serum proteins and cell surface molecules that participate in pathogen recognition and clearance (158). Complement activation occurs through the classical, lectin, and alternative pathways, which are initiated by binding of C1q or mannan-binding lectins or through spontaneous hydrolysis of C3, respectively. Complement contributes to host protection through direct opsonization and/or cytolysis, chemotaxis, immune clearance, and modulation of B- and T-cell functions (29). Complement is required for protection from lethal WNV infection in mice. WNV activates complement in vivo, and mice lacking in the central complement component C3 or complement receptors 1 and 2 showed enhanced lethality after WNV infection (125, 126). All three pathways of complement activation are important for controlling WNV, as mice deficient in alternative, classical, or lectin pathway molecules exhibited increased mortality. Interestingly, the activation pathways modulated WNV infection through distinct mechanisms. Alternative-pathway-deficient mice demonstrated normal B-cell function but impaired CD8+ T-cell responses, whereas classical-pathway- and lectin pathway-deficient mice had defects in both WNV-specific antibody production and T-cell responsiveness (125).

Cellular innate immunity.

While few studies have directly addressed the function of cellular innate immunity in WNV infection, limited data suggest that macrophages and dendritic cells may directly inhibit WNV. Macrophage uptake of WNV could control infection through direct viral clearance, enhanced antigen presentation, and cytokine and chemokine secretion. Consistent with this, depletion of myeloid cells in mice enhanced lethality after WNV infection (15). Macrophages may also control flaviviruses through the production of nitric oxide intermediates (99, 112), although the role of nitric oxide in WNV infection has not been established. Less is known about the specific protective roles of dendritic cells in WNV infection, although it is likely that they produce IFN-α/β soon after infection and function as antigen-presenting cells to prime the adaptive immune response.

γδ T cells also function in early immune responses and directly limit WNV infection. As they lack classical MHC restriction, γδ T cells can react with viral antigens in the absence of conventional antigen processing (28, 176). γδ T cells expand following WNV infection in wild-type mice, and increases in viral burden and mortality were observed in mice deficient in γδ T cells (185). Bone marrow chimera reconstitution experiments demonstrated that γδ T cells require IFN-γ to limit WNV infection (174). NK cells also have the potential to control WNV infection through recognition and elimination of virus-infected cells. NK cell activity was transiently activated and then suppressed following flavivirus infection in mice, suggesting that these viruses may have developed a mechanism to inhibit NK cell responses (180). Some viruses evade natural killing by increasing surface expression of MHC or MHC-like molecules (77, 89, 90). Although it is not clear whether WNV utilizes this strategy in vivo, WNV infection in vitro increases surface MHC-I expression in some cell types by enhancing the activity of the transporter associated with antigen processing and by NF-κB-dependent transcriptional activation of MHC class I genes (33, 48, 91, 117). Notably, antibody depletion of NK cells in mice did not alter morbidity or mortality after WNV infection (173), and similar results were obtained by using Ly49A transgenic mice that lack functional circulating NK cells (88) (M. Engle and M. Diamond, unpublished results).

ADAPTIVE IMMUNE RESPONSES TO WNV INFECTION

Humoral responses.

Humoral immunity is an essential aspect of immune system-mediated protection from WNV (16, 27, 46, 47, 139, 177, 183). B-cell-deficient mice uniformly died after WNV infection but were protected by passive transfer of immune sera (46). Soluble IgM is required, as IgM−/− mice developed high viral loads in all tested tissues and demonstrated complete lethality after WNV infection (47). In prospective studies, the level of WNV-specific IgM at day 4 after infection predicted disease outcome. While it is apparent from passive-transfer studies that immune IgG can protect against flavivirus infection, the function of IgG during primary infection is less clear. In mice, WNV-specific IgG is not produced until somewhat late in infection (days 6 to 8), after both viral seeding of the CNS and clearance from peripheral tissues have occurred (46, 171). Thus, while it is possible that WNV-specific IgG alters WNV infection of the CNS, current data suggest that by the time IgG is produced, the survival of the animal may already have been largely determined. Additional studies are required to more clearly establish how IgG modulates WNV pathogenesis during primary infection.

Mechanisms of antibody protection.

The majority of neutralizing antibodies are directed against regions of the WNV E protein, although a subset likely recognizes the prM protein (40, 52, 150, 181). The E protein has three structural domains that mediate viral attachment, entry, and viral assembly and elicit antibodies with distinct neutralization potentials. Domain III (DIII) contains the putative receptor binding domain, DII encodes the fusion loop involved in pH-dependent fusion of virus and host cell membranes, and DI participates in E-protein structural rearrangements required for fusion (1, 18, 38, 128). Crystallography, nuclear magnetic resonance, and epitope-mapping studies have shown that E-specific neutralizing antibodies map to all three domains (11, 12, 31, 42, 109, 139, 156, 182, 193). Human single-chain variable-region antibody fragments against DII were protective in mice when administered prior to and after WNV infection (66). However, the most potent inhibitory antibodies recognize a dominant neutralizing epitope on the lateral face of DIII (12, 139, 182). Neutralizing antibodies to this DIII epitope were detected in serum from WNV-convalescent patients and protected mice even when administered after WNV had spread to the CNS. Recent data suggest that DIII-specific antibodies may be particularly potent inhibitors because they block at a postattachment stage, possibly by inhibiting viral fusion (64, 137). The structural basis of DIII-specific antibody neutralization has been investigated by crystallographic analysis of DIII in complex with the Fab fragments of a neutralizing monoclonal antibody (137). These studies demonstrated that the neutralizing antibody E16 engaged 16 residues in four discontinuous DIII regions that localize to the amino terminus (residues 302 to 309) and three strand-connecting loops (residues 330 to 333, 365 to 368, and 389 to 391). Additional crystallographic modeling and cryoelectron microscopy studies have shown that only 120 of the available 180 E-protein epitopes are occupied by E16, suggesting that epitope saturation is not required for neutralization (83a, 137) (Fig. 1).

FIG. 1.

FIG. 1.

E16 contact residues and binding of WNV virions. (A) E16 contact residues (red) on DIII of the WNV envelope protein are located in the amino terminus (residues 302 to 309) and three strand-connecting loops, BC (residues 330 to 333), DE (residues 365 to 368), and FG (residues 389 to 391). (B) Pseudoatomic model of the cryoelectron microscopic reconstruction of the WNV virion. The E16 structural epitope is mapped in magenta. (C) Saturation binding of E16 on the WNV particle. E16 is predicted to bind 120 out of 180 potential epitopes with exclusion from the inner fivefold axis. (D and E) Magnified regions of the boxed areas in panel C. This figure is reprinted with permission from Macmillan Publishers Ltd. (Nature 437:764-768, copyright 2005).

Although neutralizing antibodies generated during WNV infection predominantly bind structural proteins, antibodies to the nonstructural protein NS1 also protect mice against WNV infection (39). NS1 is a highly conserved glycoprotein that is not packaged within the virus but is secreted at high levels from infected cells and associates with cell surface membranes through undefined mechanisms (118, 123, 191, 192). NS1 is a cofactor in replication and is detected in the serum of infected animals during the acute phase of WNV disease (32, 118). Although the function of soluble NS1 in WNV pathogenesis is not well understood, a recent study suggests that NS1 may inhibit complement activation by binding regulatory protein factor H (K. M. Chung, K. Liszewski, G. Nybakken, R. R. Townsend, D. Fremont, J. P. Atkinson, and M. S. Diamond, submitted for publication). Competitive-binding assays and yeast surface display mapping studies demonstrated that antibodies against WNV NS1 localize to at least three discrete regions of the protein (39). Individual NS1 antibodies protected mice from lethal WNV infection through distinct mechanisms, as some required Fc-γ receptors whereas others did not. Thus, while significant progress has been made in understanding the mechanisms of protection by passive antibody transfer, further studies are needed to define the epitope specificity and function of antibodies that develop during the course of WNV infection.

T-cell responses during primary infection.

Experiments with small-animal models have demonstrated that T lymphocytes are an essential component of protection against WNV (27, 84, 116, 136, 171, 173, 187). Consistent with this, individuals with hematologic malignancies and impaired T-cell function have an increased risk of neuroinvasive WNV infection (134, 152). Upon recognition of a WNV-infected cell that expresses class I MHC molecules, antigen-restricted cytotoxic T lymphocytes proliferate, release proinflammatory cytokines (48, 84, 100), and lyse cells directly through the delivery of perforin and granzymes A and B or via Fas-Fas ligand interactions (74, 159). Mice deficient in CD8+ T cells or class I MHC molecules had normal humoral responses but greater and sustained WNV burdens in the spleen and CNS and increased mortality (171, 187). Granzymes appear to be important for control of the lineage II isolate Sarafend, with perforin, Fas, and Fas ligand having a more limited role in modulating infection (188). In contrast, CD8+ T cells require perforin to control lineage I WNV as mice deficient in perforin molecules had increased CNS viral burdens and lethality (173). Moreover, adoptive transfer of wild-type but not perforin-deficient CD8+ T cells decreased CNS viral burdens and enhanced survival. CD4+ T cells likely contribute to the control of infection through multiple mechanisms, including CD8+ T-cell priming, cytokine production, B-cell activation and priming, and direct cytotoxicity. Preliminary data suggest that CD4+ T cells restrict WNV pathogenesis in vivo. CD4+ T-cell depletion or a genetic deficiency in MHC class II molecules results in attenuated WNV-specific antibody responses and increased lethality (E. Sitati and M. Diamond, unpublished data).

Memory responses.

Although much remains to be learned about the specific mechanisms that define immunity against WNV, both humoral and cellular responses likely are essential for protection. Specific neutralizing antibodies are generated at late times after primary WNV infection (46, 139, 141, 177), and the development of high-titer neutralizing antibodies after vaccination correlates with protection against challenge (5, 106, 177). Nevertheless, it remains unknown how the epitope specificity of the antibody repertoire contributes to protective memory responses. T-cell-mediated immunity is also likely necessary to resist a second WNV challenge. Recent studies suggest that γδ T cells contribute directly to priming protective adaptive immune responses to WNV infection in mice (184). γδ T-cell-deficient mice infected with WNV had decreased memory CD8+ T-cell responses and were more susceptible to secondary infection, and adoptive transfer of CD8+ T cells from WNV-primed wild-type but not T-cell receptor (TCR) δ−/− mice increased the survival of naive mice. Immunization with different WNV vaccine preparations can induce memory T cells (4, 130, 197), although the relative contribution of humoral versus cellular memory to vaccine efficacy remains largely undefined.

CNS IMMUNE RESPONSES TO WNV

Since WNV is naturally acquired through peripheral inoculation and then spreads to the CNS, the immune response must control viral replication in both compartments. The CNS represents a challenge to the immune system, as effective resolution of CNS infection requires clearance with limited damage to critical nonrenewing cells such as neurons (68). Several features of the innate and adaptive immune response mitigate WNV pathogenesis in the CNS. IFN-α/β controls WNV replication in neurons and has an independent role in modulating survival. Mice that lack the IFN-α/βR had increased viral loads and died earlier following intracranial WNV inoculation (160). Recent data suggest that IFN-α/β induces cell-specific antiviral programs in different neuronal populations: PKR and RNase L mediated the antiviral effects of IFN in cortical neurons after WNV infection, but these molecules were dispensable for protection in peripheral neurons (161). For some encephalitic viruses, IFN-γ can mediate noncytolytic viral clearance from infected neurons in vitro and in vivo (19, 24, 155); however, control of WNV infection in the CNS appears to be largely independent of IFN-γ. Mice deficient in IFN-γ did not demonstrate delayed viral clearance from the CNS or enhanced lethality after intracranial infection (174).

T-cell-mediated immunity is essential for controlling WNV infection in the CNS. CD8+ T cells traffic to the brain after WNV infection in mice, and their presence correlates temporally with viral clearance (126, 171, 173, 187). CD8+ T cell−/−, MHC class I−/−, and perforin−/− mice all showed WNV persistence in the brain, with detectable infectious virus up to 35 days postinfection (171, 173). Similarly, mice deficient in MHC class II molecules showed persistent WNV infection in the brain and spinal cord up to 40 days postinfection (Sitati and Diamond, unpublished). Thus, the absence of functional CD8+ or CD4+ T cells results in failure to clear WNV from infected neurons in the CNS. Since the CNS experiences limited immune surveillance in the absence of inflammation, chemokine-dependent T-cell recruitment to infected CNS tissues modulates viral pathogenesis. Following viral infection of the CNS, inflammatory chemokines (e.g., CCL5 and RANTES) are expressed by trafficking leukocytes and resident astrocytes and microglia (6, 103, 143, 154). Surprisingly, WNV infection induced expression of the chemokine CXCL10 in neurons, which recruited effector CD8+ T cells through its cognate ligand, CXCR3 (93). A genetic deficiency in CXCL10 resulted in reduced T-cell trafficking to the CNS, greater viral loads in the brain, and enhanced mortality. The chemokine receptor CCR5 also regulates T-cell trafficking to the brain during WNV infection; its absence resulted in depressed CNS leukocyte migration and increased lethality in mice (61). Collectively, these experiments suggest that multiple aspects of the innate and adaptive immune responses function together to control WNV infection in the periphery and in the CNS (Fig. 2).

FIG. 2.

FIG. 2.

WNV dissemination and immune system control. (A) WNV is maintained in nature in an enzootic mosquito-bird-mosquito transmission cycle. (B) Following Culex mosquito inoculation, WNV replicates in skin Langerhans dendritic cells, which traffic the virus to the lymph node, where further replication ensues. Following induction of a primary viremia, WNV spreads to other peripheral organs. Several aspects of the innate and adaptive immune response limit WNV replication in the periphery. IFN-α/β acts as an antiviral agent that restricts viral translation and replication soon after infection. B cells and antibody (primarily IgM) modulate viral levels in serum and prevent early CNS seeding, while complement is required for efficient priming of humoral and cellular immune responses. IFN-γ-secreting γδ T cells control viral replication through direct antiviral mechanisms and contribute to the generation of adaptive immune responses. CD4+ and CD8+ T cells participate in viral clearance from peripheral tissues. (C) Following replication in the periphery, WNV spreads to the CNS possibly through TNF-α-mediated changes in BBB permeability. Neurons are the primary target of WNV in the brain and spinal cord. IFN-α/β is required to control WNV infection in the CNS and may prolong neuronal survival. The chemokines CXCL10 and CCL5 and their cognate ligands CXCR3 and CCR5 aid in recruiting CD4+ and CD8+ T cells and monocytes to the CNS, where they function to clear virus from infected tissues.

VIRAL AND HOST FACTORS THAT MODULATE DISEASE OUTCOME

Molecular basis of virulence.

The ability of WNV to survive and cause disease within the host depends on its capacity to infect target cells and evade immune system recognition (Table 2). Certain aspects of the biology of WNV facilitate its ability to cause severe disease. WNV productively infects diverse cell populations from many animal species, suggesting usage of multiple and/or well-conserved receptors (9, 53, 58, 79, 157, 178, 195). The relatively diverse tropism of WNV allows viral replication in several tissues in animal and human hosts and may contribute to the wide spectrum of clinical manifestations (76, 142, 167, 199). WNV is cytolytic and induces apoptosis in a variety of cell types, including neurons (145, 172). Although few studies have investigated the mechanisms of WNV-induced cell death in vivo, individual WNV proteins may contribute to virus-mediated cytotoxicity. In vitro, expression of either NS3 or capsid protein induced rapid, caspase-dependent apoptosis, and capsid protein expression in vivo resulted in cell death (153, 198).

TABLE 2.

WNV virulence factors

Trait Source Effect on pathogenesis Reference(s)
Diverse cellular and species tropism Use of multiple and/or well-conserved receptors Replication in multiple tissues and diverse clinical manifestations 46, 53, 58, 79, 96, 157, 195
Induction of rapid cell death NS3, capsid May contribute to neuropathology 145, 153, 172, 198
N-linked glycosylation of E protein Genetic variation Alters virion stability, replication, particle assembly, pathogenesis 14, 73, 107, 170
N-linked glycosylation of NS1 protein Genetic variation Lack of glycosylation reduces pathogenesis in vivo Whiteman et al., submitted
IFN resistance NS2A, NS2B, NS3, NS4A, NS4B, genetic variation May suppress IFN production in vivo and allow increased replication or pathogenesis 71, 114, 115, 132
Quasispecies generation Low-fidelity RNA-dependent RNA polymerase Potential for escape from antibody neutralization and T-cell lysis 12, 80, 108
Upregulation of MHC class I expression Dependent on NF-κB activation May inhibit NK cell responses 33, 48, 85, 90, 117

Genetic variation also affects WNV virulence. Sequence-based phylogenic analyses have defined two genetic lineages of WNV, lineages I and II (101, 163). Lineage I strains are detected worldwide and have been responsible for recent lethal human outbreaks, including those in Romania (1996), Russia (1999), Israel (1998 to 2000), and the Americas (1999 to 2005) (43, 120). In contrast, lineage II strains appear to be localized to central and southern Africa and have caused occasional human infections (82, 101). In birds and mammals, lineage I strains generally induce significant encephalitis and mortality, although isolates from both lineages can be neuroinvasive (13, 21, 46, 104, 195). Thus, while lineage I WNV isolates appear to be linked to the recent increase in severe infection of humans, pathogenic lineage II isolates have been identified and have the potential to induce significant human disease. The specific sequence determinants of virulence are an area of intensive study. N-linked glycosylation of the E protein appears important for neuroinvasion as mutations of E-protein glycosylation sites attenuated viral replication and pathogenesis (13, 14, 170). Moreover, E-protein glycosylation modulates WNV virulence by altering virion stability, viral replication, and particle assembly (14, 73, 107). Glycosylation of the NS1 protein has also been linked to WNV pathogenesis. WNV NS1 contains three N-linked glycosylation sites (residues 130, 175, and 203); mutants lacking glycosylation at either two or three sites induced lower viremia and decreased lethality in vivo (M. C. Whiteman, R. M. Kinney, D. W. C. Beasley, K. M. Chung, M. S. Diamond, T. Solomon, and A. D. T. Barrett, submitted for publication).

WNV has also evolved specific strategies to avoid and/or attenuate innate and adaptive immune responses. Flaviviruses, including WNV, are largely resistant to the antiviral effects of IFN once cellular infection is established. Through studies with WNV, Japanese encephalitis virus, dengue virus, yellow fever virus, and Langat virus, this phenotype has been largely ascribed to the actions of nonstructural proteins NS2A, NS2B, NS3, NS4A, NS4B, and NS5. These nonstructural proteins disenable IFN-induced responses at multiple stages within the cell by delaying IRF-3 activation and IFN-β gene transcription and by impairing JAK1 and Tyk2 phosphorylation (17, 71, 110, 111, 115, 132, 133). The replication fitness and virulence of lineage I and lineage II strains has been linked to control of host IFN responses (83b). A pathogenic lineage I Texas isolate actively antagonized IFN signaling, whereas an attenuated lineage II strain from Madagascar lacked this activity. The replication and virulence of the lineage II isolate were restored in cells and mice that lacked the IFN-α/βR. These data suggest that inhibition of type I IFN responses may be a key feature in the evolution of pathogenic WNV strains. Consistent with this, a mutation in the NS2A protein of Kunjin virus resulted in increased IFN production in weanling mice and attenuated neurovirulence (114). Escape from the humoral immune response may also contribute to WNV pathogenesis. Flaviviruses have a low-fidelity RNA-dependent RNA polymerase that generates quasispecies in vivo (80). This antigenic variation may allow viral quasispecies to escape antibody-mediated neutralization (12), as strains with mutations at the dominant neutralizing epitope in DIII of the E protein can emerge (108).

Host genetic determinates of WNV susceptibility.

A limited number of genetic risk factors have been correlated with increased susceptibility to WNV. In inbred mouse strains, susceptibility to infection by flaviviruses, including WNV, has been mapped to a truncated isoform of the gene for OAS1b (122, 146). Although an analogous deletion has not been linked to severe WNV infection in humans, a recent study with 33 WNV-infected patients showed an increased frequency of an OAS splice-enhancer site that could result in a dominant-negative protein (196). Another study examined CCR5Δ32, a relatively common mutant allele of the chemokine receptor CCR5, in WNV-infected cohorts (62). A greater incidence of CCR5Δ32 homozygosity was observed in symptomatic and lethal WNV cases, suggesting that CCR5 may mediate resistance to WNV infection in humans.

FUTURE PERSPECTIVES

The use of animal models has fostered a significant understanding of the balance between WNV pathogenesis and immune system-mediated control. These studies establish that innate immunity, humoral immunity, and T-cell-mediated immunity are all required to orchestrate effective control of WNV. The absence of one or more of these factors often results in fulminant and fatal infection. To modulate these host defenses, WNV has developed countermeasures that attenuate innate and adaptive immune responses and facilitate its infectivity in vivo. Several important questions regarding WNV infection and host interaction remain unanswered. Although WNV induces lethal encephalitis, the exact mechanisms by which WNV crosses the BBB, spreads within the CNS, and causes neuronal dysfunction and death require further study. Additionally, although antibodies can protect against WNV infection, much remains to be learned about the specific epitopes and mechanisms that define neutralization. Finally, few studies have characterized which aspects of the memory response to WNV mediate protection. Enhanced investigation of the immunologic basis of protection may facilitate the design of improved vaccines and more-targeted immunotherapies. Such efforts also will likely enhance our understanding of the pathogenesis of related encephalitic flaviviruses that cause human disease.

Acknowledgments

We thank A. Barrett, D. Beasley, E. Fikrig, M. Gale, and T. Pierson for critical reviews of the manuscript, G. Nybakken for help with the figures, and D. Fremont and R. Klein for discussions.

This work was supported by a predoctoral fellowship from the Howard Hughes Medical Institute (M.A.S.), by the Pediatric Dengue Vaccine Initiative, and by NIH grants AI53870, AI061373 (M.S.D.), and U54 AI057160 (to the Midwest Regional Center of Excellence for Biodefense and Emerging Infectious Diseases Research).

REFERENCES

  • 1.Allison, S. L., J. Schalich, K. Stiasny, C. W. Mandl, and F. X. Heinz. 2001. Mutational evidence for an internal fusion peptide in flavivirus envelope protein E. J. Virol. 75:4268-4275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Allison, S. L., J. Schalich, K. Stiasny, C. W. Mandl, C. Kunz, and F. X. Heinz. 1995. Oligomeric rearrangement of tick-borne encephalitis virus envelope proteins induced by an acidic pH. J. Virol. 69:695-700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Anderson, J. F., and J. J. Rahal. 2002. Efficacy of interferon alpha-2b and ribavirin against West Nile virus in vitro. Emerg. Infect. Dis. 8:107-108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Anraku, I., T. J. Harvey, R. Linedale, J. Gardner, D. Harrich, A. Suhrbier, and A. A. Khromykh. 2002. Kunjin virus replicon vaccine vectors induce protective CD+ T-cell immunity. J. Virol. 76:3791-3799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Arroyo, J., C. Miller, J. Catalan, G. A. Myers, M. S. Ratterree, D. W. Trent, and T. P. Monath. 2004. ChimeriVax-West Nile virus live-attenuated vaccine: preclinical evaluation of safety, immunogenicity, and efficacy. J. Virol. 78:12497-12507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Asensio, V. C., J. Maier, R. Milner, K. Boztug, C. Kincaid, M. Moulard, C. Phillipson, K. Lindsley, T. Krucker, H. S. Fox, and I. L. Campbell. 2001. Interferon-independent, human immunodeficiency virus type 1 gp120-mediated induction of CXCL10/IP-10 gene expression by astrocytes in vivo and in vitro. J. Virol. 75:7067-7077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Asselin-Paturel, C., G. Brizard, K. Chemin, A. Boonstra, A. O'Garra, A. Vicari, and G. Trinchieri. 2005. Type I interferon dependence of plasmacytoid dendritic cell activation and migration. J. Exp. Med. 201:1157-1167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Bakri, S. J., and P. K. Kaiser. 2004. Ocular manifestations of West Nile virus. Curr. Opin. Ophthalmol. 15:537-540. [DOI] [PubMed] [Google Scholar]
  • 9.Banet-Noach, C., L. Simanov, and M. Malkinson. 2003. Direct (non-vector) transmission of West Nile virus in geese. Avian Pathol. 32:489-494. [DOI] [PubMed] [Google Scholar]
  • 10.Barton, G. M., and R. Medzhitov. 2003. Linking Toll-like receptors to IFN-α/β expression. Nat. Immunol. 4:432-433. [DOI] [PubMed] [Google Scholar]
  • 11.Beasley, D. W., and J. G. Aaskov. 2001. Epitopes on the dengue 1 virus envelope protein recognized by neutralizing IgM monoclonal antibodies. Virology 279:447-458. [DOI] [PubMed] [Google Scholar]
  • 12.Beasley, D. W., and A. D. Barrett. 2002. Identification of neutralizing epitopes within structural domain III of the West Nile virus envelope protein. J. Virol. 76:13097-13100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Beasley, D. W., L. Li, M. T. Suderman, and A. D. Barrett. 2002. Mouse neuroinvasive phenotype of West Nile virus strains varies depending upon virus genotype. Virology 296:17-23. [DOI] [PubMed] [Google Scholar]
  • 14.Beasley, D. W., M. C. Whiteman, S. Zhang, C. Y. Huang, B. S. Schneider, D. R. Smith, G. D. Gromowski, S. Higgs, R. M. Kinney, and A. D. Barrett. 2005. Envelope protein glycosylation status influences mouse neuroinvasion phenotype of genetic lineage 1 West Nile virus strains. J. Virol. 79:8339-8347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Ben-Nathan, D., I. Huitinga, S. Lustig, N. van Rooijen, and D. Kobiler. 1996. West Nile virus neuroinvasion and encephalitis induced by macrophage depletion in mice. Arch. Virol. 141:459-469. [DOI] [PubMed] [Google Scholar]
  • 16.Ben-Nathan, D., S. Lustig, G. Tam, S. Robinzon, S. Segal, and B. Rager-Zisman. 2003. Prophylactic and therapeutic efficacy of human intravenous immunoglobulin in treating West Nile virus infection in mice. J. Infect. Dis. 188:5-12. [DOI] [PubMed] [Google Scholar]
  • 17.Best, S. M., K. L. Morris, J. G. Shannon, S. J. Robertson, D. N. Mitzel, G. S. Park, E. Boer, J. B. Wolfinbarger, and M. E. Bloom. 2005. Inhibition of interferon-stimulated JAK-STAT signaling by a tick-borne flavivirus and identification of NS5 as an interferon antagonist. J. Virol. 79:12828-12839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Bhardwaj, S., M. Holbrook, R. E. Shope, A. D. Barrett, and S. J. Watowich. 2001. Biophysical characterization and vector-specific antagonist activity of domain III of the tick-borne flavivirus envelope protein. J. Virol. 75:4002-4007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Binder, G. K., and D. E. Griffin. 2001. Interferon-gamma-mediated site-specific clearance of alphavirus from CNS neurons. Science 293:303-306. [DOI] [PubMed] [Google Scholar]
  • 20.Bode, A. V., J. J. Sejvar, W. J. Pape, G. L. Campbell, and A. A. Marfin. 2006. West Nile virus disease: a descriptive study of 228 patients hospitalized in a 4-county region of Colorado in 2003. Clin. Infect. Dis. 42:1234-1240. [DOI] [PubMed] [Google Scholar]
  • 21.Brault, A. C., S. A. Langevin, R. A. Bowen, N. A. Panella, B. J. Biggerstaff, B. R. Miller, and K. Nicholas. 2004. Differential virulence of West Nile strains for American crows. Emerg. Infect. Dis. 10:2161-2168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Brenner, W., G. Storch, R. Buller, R. Vij, S. Devine, and J. DiPersio. 2005. West Nile virus encephalopathy in an allogeneic stem cell transplant recipient: use of quantitative PCR for diagnosis and assessment of viral clearance. Bone Marrow Transplant. 36:369-370. [DOI] [PubMed] [Google Scholar]
  • 23.Brinton, M. A. 2002. The molecular biology of West Nile virus: a new invader of the western hemisphere. Annu. Rev. Microbiol. 56:371-402. [DOI] [PubMed] [Google Scholar]
  • 24.Burdeinick-Kerr, R., and D. E. Griffin. 2005. Gamma interferon-dependent, noncytolytic clearance of Sindbis virus infection from neurons in vitro. J. Virol. 79:5374-5385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Busch, M. P. 2006. West Nile virus infections projected from blood donor screening data, United States, 2003. Emerg. Infect. Dis. 12:395-402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Byrne, S. N., G. M. Halliday, L. J. Johnston, and N. J. King. 2001. Interleukin-1β but not tumor necrosis factor is involved in West Nile virus-induced Langerhans cell migration from the skin in C57BL/6 mice. J. Investig. Dermatol. 117:702-709. [DOI] [PubMed] [Google Scholar]
  • 27.Camenga, D. L., N. Nathanson, and G. A. Cole. 1974. Cyclophosphamide-potentiated West Nile viral encephalitis: relative influence of cellular and humoral factors. J. Infect. Dis. 130:634-641. [DOI] [PubMed] [Google Scholar]
  • 28.Carding, S. R., and P. J. Egan. 2002. γδ T cells: functional plasticity and heterogeneity. Nat. Rev. Immunol. 2:336-345. [DOI] [PubMed] [Google Scholar]
  • 29.Carroll, M. C. 2004. The complement system in regulation of adaptive immunity. Nat. Immunol. 5:981-986. [DOI] [PubMed] [Google Scholar]
  • 30.Ceausu, E., S. Erscoiu, P. Calistru, D. Ispas, O. Dorobat, M. Homos, C. Barbulescu, I. Cojocaru, C. V. Simion, C. Cristea, C. Oprea, C. Dumitrescu, D. Duiculescu, I. Marcu, C. Mociornita, T. Stoicev, I. Zolotusca, C. Calomfirescu, R. Rusu, R. Hodrea, S. Geamai, and L. Paun. 1997. Clinical manifestations in the West Nile virus outbreak. Rom. J. Virol. 48:3-11. [PubMed] [Google Scholar]
  • 31.Cecilia, D., and E. A. Gould. 1991. Nucleotide changes responsible for loss of neuroinvasiveness in Japanese encephalitis virus neutralization-resistant mutants. Virology 181:70-77. [DOI] [PubMed] [Google Scholar]
  • 32.Chambers, T. J., C. S. Hahn, R. Galler, and C. M. Rice. 1990. Flavivirus genome organization, expression, and replication. Annu. Rev. Microbiol. 44:649-688. [DOI] [PubMed] [Google Scholar]
  • 33.Cheng, Y., N. J. King, and A. M. Kesson. 2004. Major histocompatibility complex class I (MHC-I) induction by West Nile virus: involvement of 2 signaling pathways in MHC-I up-regulation. J. Infect. Dis. 189:658-668. [DOI] [PubMed] [Google Scholar]
  • 34.Chesler, D. A., and C. S. Reiss. 2002. The role of IFN-γ in immune responses to viral infections of the central nervous system. Cytokine Growth Factor Rev. 13:441-454. [DOI] [PubMed] [Google Scholar]
  • 35.Chowers, M. Y., R. Lang, F. Nassar, D. Ben-David, M. Giladi, E. Rubinshtein, A. Itzhaki, J. Mishal, Y. Siegman-Igra, R. Kitzes, N. Pick, Z. Landau, D. Wolf, H. Bin, E. Mendelson, S. D. Pitlik, and M. Weinberger. 2001. Clinical characteristics of the West Nile fever outbreak, Israel, 2000. Emerg. Infect. Dis. 7:675-678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Chu, J. J., and M. L. Ng. 2004. Infectious entry of West Nile virus occurs through a clathrin-mediated endocytic pathway. J. Virol. 78:10543-10555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Chu, J. J., and M. L. Ng. 2004. Interaction of West Nile virus with αvβ3 integrin mediates virus entry into cells. J. Biol. Chem. 279:54533-54541. [DOI] [PubMed] [Google Scholar]
  • 38.Chu, J. J., R. Rajamanonmani, J. Li, R. Bhuvanakantham, J. Lescar, and M. L. Ng. 2005. Inhibition of West Nile virus entry by using a recombinant domain III from the envelope glycoprotein. J. Gen. Virol. 86:405-412. [DOI] [PubMed] [Google Scholar]
  • 39.Chung, K. M., G. E. Nybakken, B. S. Thompson, M. J. Engle, A. Marri, D. H. Fremont, and M. S. Diamond. 2006. Antibodies against West Nile virus nonstructural protein NS1 prevent lethal infection through Fcγ receptor-dependent and -independent mechanisms. J. Virol. 80:1340-1351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Colombage, G., R. Hall, M. Pavy, and M. Lobigs. 1998. DNA-based and alphavirus-vectored immunisation with prM and E proteins elicits long-lived and protective immunity against the flavivirus, Murray Valley encephalitis virus. Virology 250:151-163. [DOI] [PubMed] [Google Scholar]
  • 41.Crance, J. M., N. Scaramozzino, A. Jouan, and D. Garin. 2003. Interferon, ribavirin, 6-azauridine and glycyrrhizin: antiviral compounds active against pathogenic flaviviruses. Antiviral Res. 58:73-79. [DOI] [PubMed] [Google Scholar]
  • 42.Crill, W. D., and J. T. Roehrig. 2001. Monoclonal antibodies that bind to domain III of dengue virus E glycoprotein are the most efficient blockers of virus adsorption to Vero cells. J. Virol. 75:7769-7773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Dauphin, G., S. Zientara, H. Zeller, and B. Murgue. 2004. West Nile: worldwide current situation in animals and humans. Comp. Immunol. Microbiol. Infect. Dis. 27:343-355. [DOI] [PubMed] [Google Scholar]
  • 44.Davis, C. W., H. Y. Nguyen, S. L. Hanna, M. D. Sanchez, R. W. Doms, and T. C. Pierson. 2006. West Nile virus discriminates between DC-SIGN and DC-SIGNR for cellular attachment and infection. J. Virol. 80:1290-1301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Deardorff, E., J. Estrada-Franco, A. C. Brault, R. Navarro-Lopez, A. Campomanes-Cortes, P. Paz-Ramirez, M. Solis-Hernandez, W. N. Ramey, C. T. Davis, D. W. Beasley, R. B. Tesh, A. D. Barrett, and S. C. Weaver. 2006. Introductions of West Nile virus strains to Mexico. Emerg. Infect. Dis. 12:314-318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Diamond, M. S., B. Shrestha, A. Marri, D. Mahan, and M. Engle. 2003. B cells and antibody play critical roles in the immediate defense of disseminated infection by West Nile encephalitis virus. J. Virol. 77:2578-2586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Diamond, M. S., E. M. Sitati, L. D. Friend, S. Higgs, B. Shrestha, and M. Engle. 2003. A critical role for induced IgM in the protection against West Nile virus infection. J. Exp. Med. 198:1853-1862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Douglas, M. W., A. M. Kesson, and N. J. King. 1994. CTL recognition of West Nile virus-infected fibroblasts is cell cycle dependent and is associated with virus-induced increases in class I MHC antigen expression. Immunology 82:561-570. [PMC free article] [PubMed] [Google Scholar]
  • 49.Egloff, M. P., D. Benarroch, B. Selisko, J. L. Romette, and B. Canard. 2002. An RNA cap (nucleoside-2′-O-)-methyltransferase in the flavivirus RNA polymerase NS5: crystal structure and functional characterization. EMBO J. 21:2757-2768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Eldadah, A. H., and N. Nathanson. 1967. Pathogenesis of West Nile virus encephalitis in mice and rats. II. Virus multiplication, evolution of immunofluorescence, and development of histological lesions in the brain. Am. J. Epidemiol. 86:776-790. [DOI] [PubMed] [Google Scholar]
  • 51.Elshuber, S., S. L. Allison, F. X. Heinz, and C. W. Mandl. 2003. Cleavage of protein prM is necessary for infection of BHK-21 cells by tick-borne encephalitis virus. J. Gen. Virol. 84:183-191. [DOI] [PubMed] [Google Scholar]
  • 52.Falconar, A. K. 1999. Identification of an epitope on the dengue virus membrane (M) protein defined by cross-protective monoclonal antibodies: design of an improved epitope sequence based on common determinants present in both envelope (E and M) proteins. Arch. Virol. 144:2313-2330. [DOI] [PubMed] [Google Scholar]
  • 53.Farajollahi, A., N. A. Panella, P. Carr, W. Crans, K. Burguess, and N. Komar. 2003. Serologic evidence of West Nile virus infection in black bears (Ursus americanus) from New Jersey. J. Wildl. Dis. 39:894-896. [DOI] [PubMed] [Google Scholar]
  • 54.Fitzgerald, K. A., S. M. McWhirter, K. L. Faia, D. C. Rowe, E. Latz, D. T. Golenbock, A. J. Coyle, S. M. Liao, and T. Maniatis. 2003. IKKɛ and TBK1 are essential components of the IRF3 signaling pathway. Nat. Immunol. 4:491-496. [DOI] [PubMed] [Google Scholar]
  • 55.Fratkin, J. D., A. A. Leis, D. S. Stokic, S. A. Slavinski, and R. W. Geiss. 2004. Spinal cord neuropathology in human West Nile virus infection. Arch. Pathol. Lab. Med. 128:533-537. [DOI] [PubMed] [Google Scholar]
  • 56.Fredericksen, B. L., and M. Gale, Jr. 2006. West Nile virus evades activation of interferon regulatory factor 3 through RIG-I-dependent and -independent pathways without antagonizing host defense signaling. J. Virol. 80:2913-2923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Fredericksen, B. L., M. Smith, M. G. Katze, P. Y. Shi, and M. Gale, Jr. 2004. The host response to West Nile virus infection limits viral spread through the activation of the interferon regulatory factor 3 pathway. J. Virol. 78:7737-7747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Garcia-Tapia, D., C. M. Loiacono, and S. B. Kleiboeker. 2006. Replication of West Nile virus in equine peripheral blood mononuclear cells. Vet. Immunol. Immunopathol. 110:229-244. [DOI] [PubMed] [Google Scholar]
  • 59.Gibbs, S. E., A. E. Ellis, D. G. Mead, A. B. Allison, J. K. Moulton, E. W. Howerth, and D. E. Stallknecht. 2005. West Nile virus detection in the organs of naturally infected blue jays (Cyanocitta cristata). J. Wildl. Dis. 41:354-362. [DOI] [PubMed] [Google Scholar]
  • 60.Gitlin, L., W. Barchet, S. Gilfillan, M. Cella, B. Beutler, R. A. Flavell, M. S. Diamond, and M. Colonna. 2006. Essential role of mda-5 in type I IFN responses to polyriboinosinic:polyribocytidylic acid and encephalomyocarditis picornavirus. Proc. Natl. Acad. Sci. USA 103:8459-8464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Glass, W. G., J. K. Lim, R. Cholera, A. G. Pletnev, J. L. Gao, and P. M. Murphy. 2005. Chemokine receptor CCR5 promotes leukocyte trafficking to the brain and survival in West Nile virus infection. J. Exp. Med. 202:1087-1098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Glass, W. G., D. H. McDermott, J. K. Lim, S. Lekhong, S. F. Yu, W. A. Frank, J. Pape, R. C. Cheshier, and P. M. Murphy. 2006. CCR5 deficiency increases risk of symptomatic West Nile virus infection. J. Exp. Med. 203:35-40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Gollins, S. W., and J. S. Porterfield. 1985. Flavivirus infection enhancement in macrophages: an electron microscopic study of viral cellular entry. J. Gen. Virol. 66(Pt. 9):1969-1982. [DOI] [PubMed] [Google Scholar]
  • 64.Gollins, S. W., and J. S. Porterfield. 1986. A new mechanism for the neutralization of enveloped viruses by antiviral antibody. Nature 321:244-246. [DOI] [PubMed] [Google Scholar]
  • 65.Gollins, S. W., and J. S. Porterfield. 1986. The uncoating and infectivity of the flavivirus West Nile on interaction with cells: effects of pH and ammonium chloride. J. Gen. Virol. 67(Pt. 9):1941-1950. [DOI] [PubMed] [Google Scholar]
  • 66.Gould, L. H., J. Sui, H. Foellmer, T. Oliphant, T. Wang, M. Ledizet, A. Murakami, K. Noonan, C. Lambeth, K. Kar, J. F. Anderson, A. M. de Silva, M. S. Diamond, R. A. Koski, W. A. Marasco, and E. Fikrig. 2005. Protective and therapeutic capacity of human single-chain Fv-Fc fusion proteins against West Nile virus. J. Virol. 79:14606-14613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Granwehr, B. P., K. M. Lillibridge, S. Higgs, P. W. Mason, J. F. Aronson, G. A. Campbell, and A. D. Barrett. 2004. West Nile virus: where are we now? Lancet Infect. Dis. 4:547-556. [DOI] [PubMed] [Google Scholar]
  • 68.Griffin, D. E. 2003. Immune responses to RNA-virus infections of the CNS. Nat. Rev. Immunol. 3:493-502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Guarner, J., W. J. Shieh, S. Hunter, C. D. Paddock, T. Morken, G. L. Campbell, A. A. Marfin, and S. R. Zaki. 2004. Clinicopathologic study and laboratory diagnosis of 23 cases with West Nile virus encephalomyelitis. Hum. Pathol. 35:983-990. [DOI] [PubMed] [Google Scholar]
  • 70.Guirakhoo, F., F. X. Heinz, C. W. Mandl, H. Holzmann, and C. Kunz. 1991. Fusion activity of flaviviruses: comparison of mature and immature (prM-containing) tick-borne encephalitis virions. J. Gen. Virol. 72(Pt. 6):1323-1329. [DOI] [PubMed] [Google Scholar]
  • 71.Guo, J. T., J. Hayashi, and C. Seeger. 2005. West Nile virus inhibits the signal transduction pathway of alpha interferon. J. Virol. 79:1343-1350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Halevy, M., Y. Akov, D. Ben-Nathan, D. Kobiler, B. Lachmi, and S. Lustig. 1994. Loss of active neuroinvasiveness in attenuated strains of West Nile virus: pathogenicity in immunocompetent and SCID mice. Arch. Virol. 137:355-370. [DOI] [PubMed] [Google Scholar]
  • 73.Hanna, S. L., T. C. Pierson, M. D. Sanchez, A. A. Ahmed, M. M. Murtadha, and R. W. Doms. 2005. N-linked glycosylation of West Nile virus envelope proteins influences particle assembly and infectivity. J. Virol. 79:13262-13274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Harty, J. T., and V. P. Badovinac. 2002. Influence of effector molecules on the CD8+ T cell response to infection. Curr. Opin. Immunol. 14:360-365. [DOI] [PubMed] [Google Scholar]
  • 75.Hayes, E. B., N. Komar, R. S. Nasci, S. P. Montgomery, D. R. O'Leary, and G. L. Campbell. 2005. Epidemiology and transmission dynamics of West Nile virus disease. Emerg. Infect. Dis. 11:1167-1173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Hayes, E. B., J. J. Sejvar, S. R. Zaki, R. S. Lanciotti, A. V. Bode, and G. L. Campbell. 2005. Virology, pathology, and clinical manifestations of West Nile virus disease. Emerg. Infect. Dis. 11:1174-1179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Herzer, K., C. S. Falk, J. Encke, S. T. Eichhorst, A. Ulsenheimer, B. Seliger, and P. H. Krammer. 2003. Upregulation of major histocompatibility complex class I on liver cells by hepatitis C virus core protein via p53 and TAP1 impairs natural killer cell cytotoxicity. J. Virol. 77:8299-8309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Hunsperger, E. A., and J. T. Roehrig. 2006. Temporal analyses of the neuropathogenesis of a West Nile virus infection in mice. J. Neurovirol. 12:129-139. [DOI] [PubMed] [Google Scholar]
  • 79.Jacobson, E. R., P. E. Ginn, J. M. Troutman, L. Farina, L. Stark, K. Klenk, K. L. Burkhalter, and N. Komar. 2005. West Nile virus infection in farmed American alligators (Alligator mississippiensis) in Florida. J. Wildl. Dis. 41:96-106. [DOI] [PubMed] [Google Scholar]
  • 80.Jerzak, G., K. A. Bernard, L. D. Kramer, and G. D. Ebel. 2005. Genetic variation in West Nile virus from naturally infected mosquitoes and birds suggests quasispecies structure and strong purifying selection. J. Gen. Virol. 86:2175-2183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Johnson, R. T. 1998. Viral infections of the nervous system, 2nd ed. Lippincott Williams & Wilkins, Philadelphia, Pa.
  • 82.Jupp, P. G. 2001. The ecology of West Nile virus in South Africa and the occurrence of outbreaks in humans. Ann. N. Y. Acad. Sci. 951:143-152. [DOI] [PubMed] [Google Scholar]
  • 83.Kang, D. C., R. V. Gopalkrishnan, Q. Wu, E. Jankowsky, A. M. Pyle, and P. B. Fisher. 2002. mda-5: an interferon-inducible putative RNA helicase with double-stranded RNA-dependent ATPase activity and melanoma growth-suppressive properties. Proc. Natl. Acad. Sci. USA 99:637-642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83a.Kaufmann, B., G. E. Nybakken, P. R. Chipman, W. Zhang, M. S. Diamond, D. H. Fremont, R. J. Kuhn, and M. G. Rossmann. West Nile virus in complex with a Fab fragment of a neutralizing monoclonal antibody. Proc. Natl. Acad. Sci. USA, in press. [DOI] [PMC free article] [PubMed]
  • 83b.Keller, B. C., B. L. Fredericksen, M. A. Samuel, R. E. Mock, P. W. Mason, M. S. Diamond, and M. Gale, Jr. 2006. Resistance to alpha/beta interferon is a determinant of West Nile virus replication fitness and virulence. J. Virol. 80:9424-9434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Kesson, A. M., R. V. Blanden, and A. Mullbacher. 1987. The primary in vivo murine cytotoxic T cell response to the flavivirus, West Nile. J. Gen. Virol. 68(Pt. 7):2001-2006. [DOI] [PubMed] [Google Scholar]
  • 85.Kesson, A. M., and N. J. King. 2001. Transcriptional regulation of major histocompatibility complex class I by flavivirus West Nile is dependent on NF-κB activation. J. Infect. Dis. 184:947-954. [DOI] [PubMed] [Google Scholar]
  • 86.Khromykh, A. A., P. L. Sedlak, K. J. Guyatt, R. A. Hall, and E. G. Westaway. 1999. Efficient trans-complementation of the flavivirus Kunjin NS5 protein but not of the NS1 protein requires its coexpression with other components of the viral replicase. J. Virol. 73:10272-10280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Khromykh, A. A., P. L. Sedlak, and E. G. Westaway. 2000. cis- and trans-acting elements in flavivirus RNA replication. J. Virol. 74:3253-3263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Kim, S., K. Iizuka, H. L. Aguila, I. L. Weissman, and W. M. Yokoyama. 2000. In vivo natural killer cell activities revealed by natural killer cell-deficient mice. Proc. Natl. Acad. Sci. USA 97:2731-2736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.King, N. J., and A. M. Kesson. 2003. Interaction of flaviviruses with cells of the vertebrate host and decoy of the immune response. Immunol. Cell Biol. 81:207-216. [DOI] [PubMed] [Google Scholar]
  • 90.King, N. J., and A. M. Kesson. 1988. Interferon-independent increases in class I major histocompatibility complex antigen expression follow flavivirus infection. J. Gen. Virol. 69(Pt. 10):2535-2543. [DOI] [PubMed] [Google Scholar]
  • 91.King, N. J., L. E. Maxwell, and A. M. Kesson. 1989. Induction of class I major histocompatibility complex antigen expression by West Nile virus on gamma interferon-refractory early murine trophoblast cells. Proc. Natl. Acad. Sci. USA 86:911-915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Klee, A. L., B. Maidin, B. Edwin, I. Poshni, F. Mostashari, A. Fine, M. Layton, and D. Nash. 2004. Long-term prognosis for clinical West Nile virus infection. Emerg. Infect. Dis. 10:1405-1411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Klein, R. S., E. Lin, B. Zhang, A. D. Luster, J. Tollett, M. A. Samuel, M. Engle, and M. S. Diamond. 2005. Neuronal CXCL10 directs CD8+ T-cell recruitment and control of West Nile virus encephalitis. J. Virol. 79:11457-11466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Kleinschmidt-DeMasters, B. K., B. A. Marder, M. E. Levi, S. P. Laird, J. T. McNutt, E. J. Escott, G. T. Everson, and K. L. Tyler. 2004. Naturally acquired West Nile virus encephalomyelitis in transplant recipients: clinical, laboratory, diagnostic, and neuropathological features. Arch. Neurol. 61:1210-1220. [DOI] [PubMed] [Google Scholar]
  • 95.Komar, N., and G. G. Clark. 2006. West Nile virus activity in Latin America and the Caribbean. Rev. Panam. Salud Publica 19:112-117. [DOI] [PubMed] [Google Scholar]
  • 96.Komar, N., S. Langevin, S. Hinten, N. Nemeth, E. Edwards, D. Hettler, B. Davis, R. Bowen, and M. Bunning. 2003. Experimental infection of North American birds with the New York 1999 strain of West Nile virus. Emerg. Infect. Dis. 9:311-322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Kramer, L. D., and K. A. Bernard. 2001. West Nile virus infection in birds and mammals. Ann. N. Y. Acad. Sci. 951:84-93. [DOI] [PubMed] [Google Scholar]
  • 98.Kramer-Hammerle, S., I. Rothenaigner, H. Wolff, J. E. Bell, and R. Brack-Werner. 2005. Cells of the central nervous system as targets and reservoirs of the human immunodeficiency virus. Virus Res. 111:194-213. [DOI] [PubMed] [Google Scholar]
  • 99.Kreil, T. R., and M. M. Eibl. 1996. Nitric oxide and viral infection: NO antiviral activity against a flavivirus in vitro, and evidence for contribution to pathogenesis in experimental infection in vivo. Virology 219:304-306. [DOI] [PubMed] [Google Scholar]
  • 100.Kulkarni, A. B., A. Mullbacher, and R. V. Blanden. 1991. In vitro T-cell proliferative response to the flavivirus, West Nile. Viral Immunol. 4:73-82. [DOI] [PubMed] [Google Scholar]
  • 101.Lanciotti, R. S., G. D. Ebel, V. Deubel, A. J. Kerst, S. Murri, R. Meyer, M. Bowen, N. McKinney, W. E. Morrill, M. B. Crabtree, L. D. Kramer, and J. T. Roehrig. 2002. Complete genome sequences and phylogenetic analysis of West Nile virus strains isolated from the United States, Europe, and the Middle East. Virology 298:96-105. [DOI] [PubMed] [Google Scholar]
  • 102.Lanciotti, R. S., J. T. Roehrig, V. Deubel, J. Smith, M. Parker, K. Steele, B. Crise, K. E. Volpe, M. B. Crabtree, J. H. Scherret, R. A. Hall, J. S. MacKenzie, C. B. Cropp, B. Panigrahy, E. Ostlund, B. Schmitt, M. Malkinson, C. Banet, J. Weissman, N. Komar, H. M. Savage, W. Stone, T. McNamara, and D. J. Gubler. 1999. Origin of the West Nile virus responsible for an outbreak of encephalitis in the northeastern United States. Science 286:2333-2337. [DOI] [PubMed] [Google Scholar]
  • 103.Lane, T. E., V. C. Asensio, N. Yu, A. D. Paoletti, I. L. Campbell, and M. J. Buchmeier. 1998. Dynamic regulation of alpha- and beta-chemokine expression in the central nervous system during mouse hepatitis virus-induced demyelinating disease. J. Immunol. 160:970-978. [PubMed] [Google Scholar]
  • 104.Langevin, S. A., A. C. Brault, N. A. Panella, R. A. Bowen, and N. Komar. 2005. Variation in virulence of West Nile virus strains for house sparrows (Passer domesticus). Am. J. Trop. Med. Hyg. 72:99-102. [PubMed] [Google Scholar]
  • 105.Le Bon, A., C. Thompson, E. Kamphuis, V. Durand, C. Rossmann, U. Kalinke, and D. F. Tough. 2006. Cutting edge: enhancement of antibody responses through direct stimulation of B and T cells by type I IFN. J. Immunol. 176:2074-2078. [DOI] [PubMed] [Google Scholar]
  • 106.Ledizet, M., K. Kar, H. G. Foellmer, T. Wang, S. L. Bushmich, J. F. Anderson, E. Fikrig, and R. A. Koski. 2005. A recombinant envelope protein vaccine against West Nile virus. Vaccine 23:3915-3924. [DOI] [PubMed] [Google Scholar]
  • 107.Li, J., R. Bhuvanakantham, J. Howe, and M. L. Ng. 2006. The glycosylation site in the envelope protein of West Nile virus (Sarafend) plays an important role in replication and maturation processes. J. Gen. Virol. 87:613-622. [DOI] [PubMed] [Google Scholar]
  • 108.Li, L., A. D. Barrett, and D. W. Beasley. 2005. Differential expression of domain III neutralizing epitopes on the envelope proteins of West Nile virus strains. Virology 335:99-105. [DOI] [PubMed] [Google Scholar]
  • 109.Lin, B., C. R. Parrish, J. M. Murray, and P. J. Wright. 1994. Localization of a neutralizing epitope on the envelope protein of dengue virus type 2. Virology 202:885-890. [DOI] [PubMed] [Google Scholar]
  • 110.Lin, R.-J., B.-L. Chang, H.-P. Yu, C.-L. Liao, and Y.-L. Lin. 2006. Blocking of interferon-induced Jak-Stat signaling by Japanese encephalitis virus NS5 through a protein tyrosine phosphatase-mediated mechanism. J. Virol. 80:5908-5918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Lin, R.-J., C.-L. Liao, E. Lin, and Y.-L. Lin. 2004. Blocking of the alpha interferon-induced Jak-Stat signaling pathway by Japanese encephalitis virus infection. J. Virol. 78:9285-9294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Lin, Y.-L., Y.-L. Huang, S.-H. Ma, C.-T. Yeh, S.-Y. Chiou, L.-K. Chen, and C.-L. Liao. 1997. Inhibition of Japanese encephalitis virus infection by nitric oxide: antiviral effect of nitric oxide on RNA virus replication. J. Virol. 71:5227-5235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Liu, W. J., H. B. Chen, and A. A. Khromykh. 2003. Molecular and functional analyses of Kunjin virus infectious cDNA clones demonstrate the essential roles for NS2A in virus assembly and for a nonconservative residue in NS3 in RNA replication. J. Virol. 77:7804-7813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Liu, W. J., X. J. Wang, D. C. Clark, M. Lobigs, R. A. Hall, and A. A. Khromykh. 2006. A single amino acid substitution in the West Nile virus nonstructural protein NS2A disables its ability to inhibit alpha/beta interferon induction and attenuates virus virulence in mice. J. Virol. 80:2396-2404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Liu, W. J., X. J. Wang, V. V. Mokhonov, P.-Y. Shi, R. Randall, and A. A. Khromykh. 2005. Inhibition of interferon signaling by the New York 99 strain and Kunjin subtype of West Nile virus involves blockage of STAT1 and STAT2 activation by nonstructural proteins. J. Virol. 79:1934-1942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Liu, Y., R. V. Blanden, and A. Mullbacher. 1989. Identification of cytolytic lymphocytes in West Nile virus-infected murine central nervous system. J. Gen. Virol. 70(Pt. 3):565-573. [DOI] [PubMed] [Google Scholar]
  • 117.Liu, Y., N. King, A. Kesson, R. V. Blanden, and A. Mullbacher. 1988. West Nile virus infection modulates the expression of class I and class II MHC antigens on astrocytes in vitro. Ann. N. Y. Acad. Sci. 540:483-485. [DOI] [PubMed] [Google Scholar]
  • 118.Macdonald, J., J. Tonry, R. A. Hall, B. Williams, G. Palacios, M. S. Ashok, O. Jabado, D. Clark, R. B. Tesh, T. Briese, and W. I. Lipkin. 2005. NS1 protein secretion during the acute phase of West Nile virus infection. J. Virol. 79:13924-13933. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Mackenzie, J. M., and E. G. Westaway. 2001. Assembly and maturation of the flavivirus Kunjin virus appear to occur in the rough endoplasmic reticulum and along the secretory pathway, respectively. J. Virol. 75:10787-10799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Mackenzie, J. S., D. J. Gubler, and L. R. Petersen. 2004. Emerging flaviviruses: the spread and resurgence of Japanese encephalitis, West Nile and dengue viruses. Nat. Med. 10:S98-S109. [DOI] [PubMed] [Google Scholar]
  • 121.Marrack, P., J. Kappler, and T. Mitchell. 1999. Type I interferons keep activated T cells alive. J. Exp. Med. 189:521-530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Mashimo, T., M. Lucas, D. Simon-Chazottes, M. P. Frenkiel, X. Montagutelli, P. E. Ceccaldi, V. Deubel, J. L. Guenet, and P. Despres. 2002. A nonsense mutation in the gene encoding 2′-5′-oligoadenylate synthetase/L1 isoform is associated with West Nile virus susceptibility in laboratory mice. Proc. Natl. Acad. Sci. USA 99:11311-11316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Mason, P. W. 1989. Maturation of Japanese encephalitis virus glycoproteins produced by infected mammalian and mosquito cells. Virology 169:354-364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Matsumoto, M., K. Funami, M. Tanabe, H. Oshiumi, M. Shingai, Y. Seto, A. Yamamoto, and T. Seya. 2003. Subcellular localization of Toll-like receptor 3 in human dendritic cells. J. Immunol. 171:3154-3162. [DOI] [PubMed] [Google Scholar]
  • 125.Mehlhop, E., and M. S. Diamond. 2006. Protective immune responses against West Nile virus are primed by distinct complement activation pathways. J. Exp. Med. 203:1371-1381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Mehlhop, E., K. Whitby, T. Oliphant, A. Marri, M. Engle, and M. S. Diamond. 2005. Complement activation is required for induction of a protective antibody response against West Nile virus infection. J. Virol. 79:7466-7477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Meurs, E. F., Y. Watanabe, S. Kadereit, G. N. Barber, M. G. Katze, K. Chong, B. R. Williams, and A. G. Hovanessian. 1992. Constitutive expression of human double-stranded RNA-activated p68 kinase in murine cells mediates phosphorylation of eukaryotic initiation factor 2 and partial resistance to encephalomyocarditis virus growth. J. Virol. 66:5805-5814. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Modis, Y., S. Ogata, D. Clements, and S. C. Harrison. 2004. Structure of the dengue virus envelope protein after membrane fusion. Nature 427:313-319. [DOI] [PubMed] [Google Scholar]
  • 129.Monath, T. P., C. B. Cropp, and A. K. Harrison. 1983. Mode of entry of a neurotropic arbovirus into the central nervous system. Reinvestigation of an old controversy. Lab. Investig. 48:399-410. [PubMed] [Google Scholar]
  • 130.Monath, T. P., J. Liu, N. Kanesa-Thasan, G. A. Myers, R. Nichols, A. Deary, K. McCarthy, C. Johnson, T. Ermak, S. Shin, J. Arroyo, F. Guirakhoo, J. S. Kennedy, F. A. Ennis, S. Green, and P. Bedford. 2006. A live, attenuated recombinant West Nile virus vaccine. Proc. Natl. Acad. Sci. USA 103:6694-6699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Mukhopadhyay, S., R. J. Kuhn, and M. G. Rossmann. 2005. A structural perspective of the flavivirus life cycle. Nat. Rev. Microbiol. 3:13-22. [DOI] [PubMed] [Google Scholar]
  • 132.Muñoz-Jordán, J. L., M. Laurent-Rolle, J. Ashour, L. Martínez-Sobrido, M. Ashok, W. I. Lipkin, and A. García-Sastre. 2005. Inhibition of alpha/beta interferon signaling by the NS4B protein of flaviviruses. J. Virol. 79:8004-8013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Muñoz-Jordán, J. L., G. G. Sanchez-Burgos, M. Laurent-Rolle, and A. García-Sastre. 2003. Inhibition of interferon signaling by dengue virus. Proc. Natl. Acad. Sci. USA 100:14333-14338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Murray, K., S. Baraniuk, M. Resnick, R. Arafat, C. Kilborn, K. Cain, R. Shallenberger, T. L. York, D. Martinez, J. S. Hellums, D. Hellums, M. Malkoff, N. Elgawley, W. McNeely, S. A. Khuwaja, and R. B. Tesh. 4. May 2006. Risk factors for encephalitis and death from West Nile virus infection. Epidemiol. Infect. 134:1-8. [Epub ahead of print.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Nash, D., F. Mostashari, A. Fine, J. Miller, D. O'Leary, K. Murray, A. Huang, A. Rosenberg, A. Greenberg, M. Sherman, S. Wong, and M. Layton. 2001. The outbreak of West Nile virus infection in the New York City area in 1999. N. Engl. J. Med. 344:1807-1814. [DOI] [PubMed] [Google Scholar]
  • 136.Nathanson, N., and G. A. Cole. 1970. Immunosuppression and experimental virus infection of the nervous system. Adv. Virus Res. 16:397-448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Nybakken, G. E., T. Oliphant, S. Johnson, S. Burke, M. S. Diamond, and D. H. Fremont. 2005. Structural basis of West Nile virus neutralization by a therapeutic antibody. Nature 437:764-769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.O'Leary, D. R., A. A. Marfin, S. P. Montgomery, A. M. Kipp, J. A. Lehman, B. J. Biggerstaff, V. L. Elko, P. D. Collins, J. E. Jones, and G. L. Campbell. 2004. The epidemic of West Nile virus in the United States, 2002. Vector Borne Zoonotic Dis. 4:61-70. [DOI] [PubMed] [Google Scholar]
  • 139.Oliphant, T., M. Engle, G. E. Nybakken, C. Doane, S. Johnson, L. Huang, S. Gorlatov, E. Mehlhop, A. Marri, K. M. Chung, G. D. Ebel, L. D. Kramer, D. H. Fremont, and M. S. Diamond. 2005. Development of a humanized monoclonal antibody with therapeutic potential against West Nile virus. Nat. Med. 11(5):522-530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Omalu, B. I., A. A. Shakir, G. Wang, W. I. Lipkin, and C. A. Wiley. 2003. Fatal fulminant pan-meningo-polioencephalitis due to West Nile virus. Brain Pathol. 13:465-472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Ozkul, A., Y. Yildirim, D. Pinar, A. Akcali, V. Yilmaz, and D. Colak. 2005. Serological evidence of West Nile virus (WNV) in mammalian species in Turkey. Epidemiol. Infect. 134(4):826-829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Paddock, C. D., W. L. Nicholson, J. Bhatnagar, C. S. Goldsmith, P. W. Greer, E. B. Hayes, J. A. Risko, C. Henderson, C. G. Blackmore, R. S. Lanciotti, G. L. Campbell, and S. R. Zaki. 2006. Fatal hemorrhagic fever caused by West Nile virus in the United States. Clin. Infect. Dis. 42:1527-1535. [DOI] [PubMed] [Google Scholar]
  • 143.Palma, J. P., and B. S. Kim. 2004. The scope and activation mechanisms of chemokine gene expression in primary astrocytes following infection with Theiler's virus. J. Neuroimmunol. 149:121-129. [DOI] [PubMed] [Google Scholar]
  • 144.Panella, N. A., A. J. Kerst, R. S. Lanciotti, P. Bryant, B. Wolf, and N. Komar. 2001. Comparative West Nile virus detection in organs of naturally infected American crows (Corvus brachyrhynchos). Emerg. Infect. Dis. 7:754-755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Parquet, M. C., A. Kumatori, F. Hasebe, K. Morita, and A. Igarashi. 2001. West Nile virus-induced bax-dependent apoptosis. FEBS Lett. 500:17-24. [DOI] [PubMed] [Google Scholar]
  • 146.Perelygin, A. A., S. V. Scherbik, I. B. Zhulin, B. M. Stockman, Y. Li, and M. A. Brinton. 2002. Positional cloning of the murine flavivirus resistance gene. Proc. Natl. Acad. Sci. USA 99:9322-9327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Pestka, S., C. D. Krause, and M. R. Walter. 2004. Interferons, interferon-like cytokines, and their receptors. Immunol. Rev. 202:8-32. [DOI] [PubMed] [Google Scholar]
  • 148.Petersen, L. R., and A. A. Marfin. 2002. West Nile virus: a primer for the clinician. Ann. Intern. Med. 137:173-179. [DOI] [PubMed] [Google Scholar]
  • 149.Petersen, L. R., and J. T. Roehrig. 2001. West Nile virus: a reemerging global pathogen. Emerg. Infect. Dis. 7:611-614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Pincus, S., P. W. Mason, E. Konishi, B. A. Fonseca, R. E. Shope, C. M. Rice, and E. Paoletti. 1992. Recombinant vaccinia virus producing the prM and E proteins of yellow fever virus protects mice from lethal yellow fever encephalitis. Virology 187:290-297. [DOI] [PubMed] [Google Scholar]
  • 151.Platanias, L. C. 2005. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat. Rev. Immunol. 5:375-386. [DOI] [PubMed] [Google Scholar]
  • 152.Pruitt, A. A. 2004. Central nervous system infections in cancer patients. Semin. Neurol. 24:435-452. [DOI] [PubMed] [Google Scholar]
  • 153.Ramanathan, M. P., J. A. Chambers, P. Pankhong, M. Chattergoon, W. Attatippaholkun, K. Dang, N. Shah, and D. B. Weiner. 2006. Host cell killing by the West Nile virus NS2B-NS3 proteolytic complex: NS3 alone is sufficient to recruit caspase-8-based apoptotic pathway. Virology 345:56-72. [DOI] [PubMed] [Google Scholar]
  • 154.Ransohoff, R. M., T. Wei, K. D. Pavelko, J. C. Lee, P. D. Murray, and M. Rodriguez. 2002. Chemokine expression in the central nervous system of mice with a viral disease resembling multiple sclerosis: roles of CD4+ and CD8+ T cells and viral persistence. J. Virol. 76:2217-2224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Rodriguez, M., L. J. Zoecklein, C. L. Howe, K. D. Pavelko, J. D. Gamez, S. Nakane, and L. M. Papke. 2003. Gamma interferon is critical for neuronal viral clearance and protection in a susceptible mouse strain following early intracranial Theiler's murine encephalomyelitis virus infection. J. Virol. 77:12252-12265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Roehrig, J. T., R. A. Bolin, and R. G. Kelly. 1998. Monoclonal antibody mapping of the envelope glycoprotein of the dengue 2 virus, Jamaica. Virology 246:317-328. [DOI] [PubMed] [Google Scholar]
  • 157.Root, J. J., J. S. Hall, R. G. McLean, N. L. Marlenee, B. J. Beaty, J. Gansowski, and L. Clark. 2005. Serologic evidence of exposure of wild mammals to flaviviruses in the central and eastern United States. Am. J. Trop. Med. Hyg. 72:622-630. [PubMed] [Google Scholar]
  • 158.Roozendaal, R., and M. C. Carroll. 2006. Emerging patterns in complement-mediated pathogen recognition. Cell 125:29-32. [DOI] [PubMed] [Google Scholar]
  • 159.Russell, J. H., and T. J. Ley. 2002. Lymphocyte-mediated cytotoxicity. Annu. Rev. Immunol. 20:323-370. [DOI] [PubMed] [Google Scholar]
  • 160.Samuel, M. A., and M. S. Diamond. 2005. Alpha/beta interferon protects against lethal West Nile virus infection by restricting cellular tropism and enhancing neuronal survival. J. Virol. 79:13350-13361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Samuel, M. A., K. Whitby, B. C. Keller, A. Marri, W. Barchet, B. R. G. Williams, R. H. Silverman, M. Gale, Jr., and M. S. Diamond. 2006. PKR and RNase L contribute to protection against lethal West Nile virus infection by controlling early viral spread in the periphery and replication in neurons. J. Virol. 80:7009-7019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Scherbik, S. V., J. M. Paranjape, B. M. Stockman, R. H. Silverman, and M. A. Brinton. 2006. RNase L plays a role in the antiviral response to West Nile virus. J. Virol. 80:2987-2999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Scherret, J. H., J. S. Mackenzie, R. A. Hall, V. Deubel, and E. A. Gould. 2002. Phylogeny and molecular epidemiology of West Nile and Kunjin viruses. Curr. Top. Microbiol. Immunol. 267:373-390. [DOI] [PubMed] [Google Scholar]
  • 164.Scholle, F., and P. W. Mason. 2005. West Nile virus replication interferes with both poly(I:C)-induced interferon gene transcription and response to interferon treatment. Virology 342:77-87. [DOI] [PubMed] [Google Scholar]
  • 165.Schroder, K., P. J. Hertzog, T. Ravasi, and D. A. Hume. 2004. Interferon-gamma: an overview of signals, mechanisms and functions. J. Leukoc. Biol. 75:163-189. [DOI] [PubMed] [Google Scholar]
  • 166.Sejvar, J. J. 2006. West Nile virus-associated flaccid paralysis outcome. Emerg. Infect. Dis. 12:514-516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Sejvar, J. J., M. B. Haddad, B. C. Tierney, G. L. Campbell, A. A. Marfin, J. A. Van Gerpen, A. Fleischauer, A. A. Leis, D. S. Stokic, and L. R. Petersen. 2003. Neurologic manifestations and outcome of West Nile virus infection. JAMA 290:511-515. [DOI] [PubMed] [Google Scholar]
  • 168.Sejvar, J. J., A. A. Leis, D. S. Stokic, J. A. Van Gerpen, A. A. Marfin, R. Webb, M. B. Haddad, B. C. Tierney, S. A. Slavinski, J. L. Polk, V. Dostrow, M. Winkelmann, and L. R. Petersen. 2003. Acute flaccid paralysis and West Nile virus infection. Emerg. Infect. Dis. 9:788-793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Sharma, S., B. R. tenOever, N. Grandvaux, G. P. Zhou, R. Lin, and J. Hiscott. 2003. Triggering the interferon antiviral response through an IKK-related pathway. Science 300:1148-1151. [DOI] [PubMed] [Google Scholar]
  • 170.Shirato, K., H. Miyoshi, A. Goto, Y. Ako, T. Ueki, H. Kariwa, and I. Takashima. 2004. Viral envelope protein glycosylation is a molecular determinant of the neuroinvasiveness of the New York strain of West Nile virus. J. Gen. Virol. 85:3637-3645. [DOI] [PubMed] [Google Scholar]
  • 171.Shrestha, B., and M. S. Diamond. 2004. Role of CD8+ T cells in control of West Nile virus infection. J. Virol. 78:8312-8321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Shrestha, B., D. Gottlieb, and M. S. Diamond. 2003. Infection and injury of neurons by West Nile encephalitis virus. J. Virol. 77:13203-13213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Shrestha, B., M. A. Samuel, and M. S. Diamond. 2006. CD8+ T cells require perforin to clear West Nile virus from infected neurons. J. Virol. 80:119-129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Shrestha, B., T. Wang, M. A. Samuel, K. Whitby, J. Craft, E. Fikrig, and M. S. Diamond. 2006. Gamma interferon plays a crucial early antiviral role in protection against West Nile virus infection. J. Virol. 80:5338-5348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Silverman, R. H. 1997. 2-5A-dependent RNase L: a regulated endoribonuclease in the interferon system, p. 115. In E. G. D'Alessio and J. F Riordan (ed.), Ribonuclease: structure and function. Academic Press, Inc., New York, N.Y.
  • 176.Steele, C. R., D. E. Oppenheim, and A. C. Hayday. 2000. γδ T cells: non-classical ligands for non-classical cells. Curr. Biol. 10:R282-R285. [DOI] [PubMed] [Google Scholar]
  • 177.Tesh, R. B., J. Arroyo, A. P. Travassos Da Rosa, H. Guzman, S. Y. Xiao, and T. P. Monath. 2002. Efficacy of killed virus vaccine, live attenuated chimeric virus vaccine, and passive immunization for prevention of West Nile virus encephalitis in hamster model. Emerg. Infect. Dis. 8:1392-1397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Tesh, R. B., M. Siirin, H. Guzman, A. P. Travassos da Rosa, X. Wu, T. Duan, H. Lei, M. R. Nunes, and S. Y. Xiao. 2005. Persistent West Nile virus infection in the golden hamster: studies on its mechanism and possible implications for other flavivirus infections. J. Infect. Dis. 192:287-295. [DOI] [PubMed] [Google Scholar]
  • 179.Tsai, T. F., F. Popovici, C. Cernescu, G. L. Campbell, and N. I. Nedelcu. 1998. West Nile encephalitis epidemic in southeastern Romania. Lancet 352:767-771. [DOI] [PubMed] [Google Scholar]
  • 180.Vargin, V. V., and B. F. Semenov. 1986. Changes of natural killer cell activity in different mouse lines by acute and asymptomatic flavivirus infections. Acta Virol. 30:303-308. [PubMed] [Google Scholar]
  • 181.Vazquez, S., M. G. Guzman, G. Guillen, G. Chinea, A. B. Perez, M. Pupo, R. Rodriguez, O. Reyes, H. E. Garay, I. Delgado, G. Garcia, and M. Alvarez. 2002. Immune response to synthetic peptides of dengue prM protein. Vaccine 20:1823-1830. [DOI] [PubMed] [Google Scholar]
  • 182.Volk, D. E., D. W. Beasley, D. A. Kallick, M. R. Holbrook, A. D. Barrett, and D. G. Gorenstein. 2004. Solution structure and antibody binding studies of the envelope protein domain III from the New York strain of West Nile virus. J. Biol. Chem. 279:38755-38761. [DOI] [PubMed] [Google Scholar]
  • 183.Wang, T., J. F. Anderson, L. A. Magnarelli, S. J. Wong, R. A. Koski, and E. Fikrig. 2001. Immunization of mice against West Nile virus with recombinant envelope protein. J. Immunol. 167:5273-5277. [DOI] [PubMed] [Google Scholar]
  • 184.Wang, T., Y. Gao, E. Scully, C. T. Davis, J. F. Anderson, T. Welte, M. Ledizet, R. Koski, J. A. Madri, A. D. Barrett, Z. Yin, J. Craft, and E. Fikrig. 2006. γ-δ T cells facilitate adaptive immunity against West Nile virus. J. Immunol. 177:1825-1832. [DOI] [PubMed] [Google Scholar]
  • 185.Wang, T., E. Scully, Z. Yin, J. H. Kim, S. Wang, J. Yan, M. Mamula, J. F. Anderson, J. Craft, and E. Fikrig. 2003. IFN-γ-producing γδ T cells help control murine West Nile virus infection. J. Immunol. 171:2524-2531. [DOI] [PubMed] [Google Scholar]
  • 186.Wang, T., T. Town, L. Alexopoulou, J. F. Anderson, E. Fikrig, and R. A. Flavell. 2004. Toll-like receptor 3 mediates West Nile virus entry into the brain causing lethal encephalitis. Nat. Med. 10:1366-1373. [DOI] [PubMed] [Google Scholar]
  • 187.Wang, Y., M. Lobigs, E. Lee, and A. Mullbacher. 2003. CD8+ T cells mediate recovery and immunopathology in West Nile virus encephalitis. J. Virol. 77:13323-13334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Wang, Y., M. Lobigs, E. Lee, and A. Mullbacher. 2004. Exocytosis and Fas mediated cytolytic mechanisms exert protection from West Nile virus induced encephalitis in mice. Immunol. Cell Biol. 82:170-173. [DOI] [PubMed] [Google Scholar]
  • 189.Watson, J. T., P. E. Pertel, R. C. Jones, A. M. Siston, W. S. Paul, C. C. Austin, and S. I. Gerber. 2004. Clinical characteristics and functional outcomes of West Nile fever. Ann. Intern. Med. 141:360-365. [DOI] [PubMed] [Google Scholar]
  • 190.Weaver, S. C., and A. D. Barrett. 2004. Transmission cycles, host range, evolution and emergence of arboviral disease. Nat. Rev. Microbiol. 2:789-801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Winkler, G., S. E. Maxwell, C. Ruemmler, and V. Stollar. 1989. Newly synthesized dengue-2 virus nonstructural protein NS1 is a soluble protein but becomes partially hydrophobic and membrane-associated after dimerization. Virology 171:302-305. [DOI] [PubMed] [Google Scholar]
  • 192.Winkler, G., V. B. Randolph, G. R. Cleaves, T. E. Ryan, and V. Stollar. 1988. Evidence that the mature form of the flavivirus nonstructural protein NS1 is a dimer. Virology 162:187-196. [DOI] [PubMed] [Google Scholar]
  • 193.Wu, S. C., W. C. Lian, L. C. Hsu, and M. Y. Liau. 1997. Japanese encephalitis virus antigenic variants with characteristic differences in neutralization resistance and mouse virulence. Virus Res. 51:173-181. [DOI] [PubMed] [Google Scholar]
  • 194.Wunschmann, A., J. Shivers, L. Carroll, and J. Bender. 2004. Pathological and immunohistochemical findings in American crows (Corvus brachyrhynchos) naturally infected with West Nile virus. J. Vet. Diagn. Investig. 16:329-333. [DOI] [PubMed] [Google Scholar]
  • 195.Xiao, S. Y., H. Guzman, H. Zhang, A. P. Travassos da Rosa, and R. B. Tesh. 2001. West Nile virus infection in the golden hamster (Mesocricetus auratus): a model for West Nile encephalitis. Emerg. Infect. Dis. 7:714-721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Yakub, I., K. M. Lillibridge, A. Moran, O. Y. Gonzalez, J. Belmont, R. A. Gibbs, and D. J. Tweardy. 2005. Single nucleotide polymorphisms in genes for 2′-5′-oligoadenylate synthetase and RNase L in patients hospitalized with West Nile virus infection. J. Infect. Dis. 192:1741-1748. [DOI] [PubMed] [Google Scholar]
  • 197.Yang, J. S., J. J. Kim, D. Hwang, A. Y. Choo, K. Dang, H. Maguire, S. Kudchodkar, M. P. Ramanathan, and D. B. Weiner. 2001. Induction of potent Th1-type immune responses from a novel DNA vaccine for West Nile virus New York isolate (WNV-NY1999). J. Infect. Dis. 184:809-816. [DOI] [PubMed] [Google Scholar]
  • 198.Yang, J. S., M. P. Ramanathan, K. Muthumani, A. Y. Choo, S. H. Jin, Q. C. Yu, D. S. Hwang, D. K. Choo, M. D. Lee, K. Dang, W. Tang, J. J. Kim, and D. B. Weiner. 2002. Induction of inflammation by West Nile virus capsid through the caspase-9 apoptotic pathway. Emerg. Infect. Dis. 8:1379-1384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Yim, R., K. M. Posfay-Barbe, D. Nolt, G. Fatula, and E. R. Wald. 2004. Spectrum of clinical manifestations of West Nile virus infection in children. Pediatrics 114:1673-1675. [DOI] [PubMed] [Google Scholar]
  • 200.Yoneyama, M., M. Kikuchi, K. Matsumoto, T. Imaizumi, M. Miyagishi, K. Taira, E. Foy, Y. M. Loo, M. Gale, Jr., S. Akira, S. Yonehara, A. Kato, and T. Fujita. 2005. Shared and unique functions of the DExD/H-box helicases RIG-I, MDA5, and LGP2 in antiviral innate immunity. J. Immunol. 175:2851-2858. [DOI] [PubMed] [Google Scholar]
  • 201.Yoneyama, M., M. Kikuchi, T. Natsukawa, N. Shinobu, T. Imaizumi, M. Miyagishi, K. Taira, S. Akira, and T. Fujita. 2004. The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses. Nat. Immunol. 5:730-737. [DOI] [PubMed] [Google Scholar]
  • 202.Yusof, R., S. Clum, M. Wetzel, H. M. Murthy, and R. Padmanabhan. 2000. Purified NS2B/NS3 serine protease of dengue virus type 2 exhibits cofactor NS2B dependence for cleavage of substrates with dibasic amino acids in vitro. J. Biol. Chem. 275:9963-9969. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Virology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES