Sie sind auf Seite 1von 14

REVIEWS

Molecular mechanisms of varicella


zoster virus pathogenesis
Leigh Zerboni, Nandini Sen, Stefan L. Oliver and Ann M. Arvin
Abstract | Varicella zoster virus (VZV) is the causative agent of varicella (chickenpox) and
zoster (shingles). Investigating VZV pathogenesis is challenging as VZV is a human-specific
virus and infection does not occur, or is highly restricted, in other species. However, the use
of human tissue xenografts in mice with severe combined immunodeficiency (SCID)
enables the analysis of VZV infection in differentiated human cells in their typical tissue
microenvironment. Xenografts of human skin, dorsal root ganglia or foetal thymus that
contains T cells can be infected with mutant viruses or in the presence of inhibitors of viral or
cellular functions to assess the molecular mechanisms of VZV–host interactions. In this
Review, we discuss how these models have improved our understanding of VZV pathogenesis.

Dermatome
Varicella zoster virus (VZV), which is a human alphaher- predominantly composed of viral regulatory proteins
An area of skin that is primarily pesvirus of the genus Varicellovirus, causes varicella (also surrounds an icosahedral nucleocapsid core that contains
innervated by a single sensory known as chickenpox) and zoster (also known as shin- the linear double-stranded DNA genome1. The viral life
ganglion of the spinal cord. gles)1. Epidemiological evidence suggests that primary cycle begins with VZV entry, which is a poorly under-
VZV infection begins with replication in epithelial cells stood process that is presumed to involve either direct
of the upper respiratory mucosa, which is followed by fusion of viral particles with the plasma membrane or
the widely distributed vesicular rash that is typical of endocytosis (FIG. 1b). Viral envelope proteins are pre-
varicella after an incubation period of 10–21 days. This dicted to interact with cell surface molecules, such as
pattern probably reflects viral spread to the tonsils and mannose-6-phosphate receptor8 or myelin-associated
other local lymphoid tissues, from where infected T cells glycoprotein9. VZV glycoprotein B (gB), gH and gL
can transport the virus via the bloodstream to the skin2 function as the core fusion complex9, but other envelope
(FIG. 1a). During primary infection, virions presumably glycoproteins probably contribute as accessory proteins.
gain access to the sensory nerve cell bodies in ganglia by After entry, the virions undergo uncoating, and tegument
retrograde axonal transport from skin sites of replication proteins, including the immediate-early protein 62 (IE62)
or by T cell viraemia, and latent infection is established3. — which is the major viral protein that functions as a
When viral replication is reactivated, VZV reaches the transcription factor (that is, as a viral transactivator)1,10
skin via anterograde axonal transport to cause the symp- — are released and might be transported to the nucleus
toms of zoster, which is characterized by a vesicular rash before de novo protein synthesis occurs. Nucleo­capsids
in the dermatome that is innervated by the affected gan- anchor at nuclear pores, where viral genomes are injected
glion. Both varicella and zoster skin lesions contain high into the nucleus. VZV genome replication and viral gene
concentrations of infectious virus and are thus responsi- expression depend on virus-encoded and host cell tran-
ble for transmission to susceptible individuals. Varicella scription factors and cellular translation systems11. Fur-
epidemics occurred annually in the United States until thermore, the tegument proteins ORF47 and ORF66 are
Departments of Pediatrics a varicella vaccine (which is a live attenuated form of the important serine/threonine kinases that autophospho-
and of Microbiology &
VZV Oka strain) was introduced in 1995, but epidemics rylate and phosphorylate viral transcription factors and
Immunology, Stanford
University School of continue among children in countries that do not have other VZV proteins12–18. IE62 forms regulatory complexes
Medicine, Stanford, immunization programs4–6. with cellular factors, such as transcription factor speci-
California 94305, USA. The VZV genome has at least 71 known or predicted ficity protein 1 (Sp1), which has binding sites in many
Correspondence to A.M.A. ORFs1,7 (BOX 1). Similar to all herpesviruses, VZV has viral promoters11, to transactivate VZV genes. Similarly
e-mail: aarvin@stanford.edu
doi:10.1038/nrmicro3215
a lipid-rich envelope, which is acquired from cellu- to other herpesviruses, nucleocapsids undergo primary
Published online lar membranes and into which viral glycoproteins are envelopment, fusion with nuclear membranes and
10 February 2014 inserted. Within the envelope, a tegument layer that is de‑envelopment during transfer to the cytoplasm (FIG. 1b).

NATURE REVIEWS | MICROBIOLOGY VOLUME 12 | MARCH 2014 | 197

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Inoculation
Waldeyer’s Ring
VZV
T cell viremia
Blood vessel Dermatome
Primary skin
infection T cell Spinal cord Reactivated
skin infection

Retrograde Anterograde
transport Neuronal axon transport

Dorsal root ganglion

Varicella Latency Zoster

b Expression of viral proteins


Immediate early Early (E) Late (L)
(IE) proteins proteins proteins
Nucleocapsid proteins

Golgi
Nucleus

Nuclear pore
Uncoated capsid
Vesicles Virion release
Attachment
DNA
Fusion and
Varicella zoster virion uncoating Assembly and
DNA duplication secondary
envelopment
Budding across
nuclear membrane

Figure 1 | VZV life cycle and replication.  a | Model of the varicella zoster virus (VZV) life cycle. VZV infects the human
Nature Reviews | Microbiology
host when virus particles reach mucosal epithelial sites of entry. Local replication is followed by spread to tonsils and other
regional lymphoid tissues, where VZV gains access to T cells. Infected T cells then deliver the virus to cutaneous sites of
replication. VZV establishes latency in sensory ganglia after transport to neuronal nuclei along neuronal axons or by
viraemia. Reactivation from latency enables a second phase of replication to occur in skin, which typically causes lesions in
the dermatome that is innervated by the affected sensory ganglion. b | Model of VZV replication. Enveloped VZV particles
attach to cell membranes, fuse and release tegument proteins. Uncoated capsids dock at nuclear pores, where genomic
Trans-Golgi network DNA is injected into the nucleus and circularizes. On the basis of events that have been documented in herpes simplex
(TGN). An intracellular virus 1 (HSV‑1) replication, immediate-early genes are expressed, followed by early and late genes. Nucleocapsids are
collection of tubules and assembled and package newly synthesized genomic DNA, move to the inner nuclear membrane and bud across the
vesicles that are located at the
nuclear membrane. Capsids enter the cytoplasm, and virion glycoproteins mature in the trans-Golgi region and tegument
trans face of the Golgi stack; it
is involved in processing
proteins assemble in vesicles; capsids undergo secondary envelopment and are transported to cell surfaces, where newly
glycoproteins. assembled virus particles are released.   

Syncytia
Multinucleated cells that are
created by fusion of Secondary envelopment occurs in the cisternae of the of infection19. VZV differs from other herpesviruses
membranes between cells with trans-Golgi network (TGN), where the capsids acquire in that assembled virions typically remain highly cell-
single nuclei; multinucleated tegument proteins and glycoprotein-containing mem- associated. The same viral glycoproteins that are predicted
cells in tissue are referred to as branes. Nascent virus particles then move to the cell to mediate entry are expressed on cell membranes and
polykaryocytes.
surface in post-Golgi compartment vesicles; the first induce fusion of infected and uninfected cells, producing
Polykaryocytes enveloped progeny virions are detected 9 hours after infec- syncytia and multinucleated polykaryocytes, which further
Cells that contain many nuclei. tion and many are present on cell surfaces within 12 hours contributes to virus spread20–25.

198 | MARCH 2014 | VOLUME 12 www.nature.com/reviews/micro

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Box 1 | VZV genome and virion structure


Genome structure a S/L;ORF0
ORF1
The VZV genome is a linear double-stranded DNA molecule of ORF2
~125,000 bp that encodes at least 71 unique ORFs and related ORF3 gK;ORF5 ORF4
promoter sequences1,7. Five phylogenetic VZV clades have ORF7 ORF6
dUTPase;ORF8
been identified, but the most disparate still have 99.8% gN;ORF9a
sequence conservation69. The genome consists of a unique ORF9
ORF10
long region (UL) of ~105,000 bp, a unique short region (US) of ORF11
ORF12
~5,232 bp, and internal repeat (IR) and terminal repeat (TR) TS;ORF13
regions (see the figure, part a). The genes that encode ORF62 gC;ORF14
ORF15
and ORF70, ORF63 and ORF71, and ORF64 and ORF69 are DNApol;ORF16 ORF20
duplicated. The origin of replication (ori) is located in the repeat ORF17 ORF21
Ribonucleotide ORF18
region. The VZV genome can circularize via unpaired bases at reductase ORF19
ORF22
each end. About two-thirds of VZV ORFs are necessary for
replication in vitro77, most of which are among the ~40 genes ORF24 ORF23
ORF25
that are conserved in all herpesviruses, including eight ORF26 ORF27
glycoproteins (gB, gC, gE, gH, gI, gK, gL, gN; green), proteins DNApol;ORF28
that are involved in DNA replication (purple) and other ORF29
UL
functions, such as DNA cleavage and packaging, nucleic acid ORF30
ORF32 gB;ORF31
metabolism and capsid assembly. Replication proteins include ORF33
the small and large subunits of the viral ribonucleotide ORF34
ORF35
reductase (known as ORF18 and ORF19), the two subunits TK;ORF36 gH;ORF37
ORF38
of the viral DNA polymerase (known as ORF16 and ORF28), ORF39
the single-stranded DNA-binding protein (known as ORF29), the ORF40
ORF41
origin of DNA replication binding protein (known as ORF51), ORF43
ORF45/42 ORF44
two viral protein kinases (known as ORF47 and ORF66) and
ORF46
other enzymes that are involved in DNA replication, including PK;ORF47 ORF49 ORF48
dUTPase (known as ORF 8), thymidylate synthetase (known as gM;ORF50 DNAss;ORF51
ORF13), DNase (known as ORF48) and uracil DNA glycosylase ORF52
ORF53
(known as ORF59). Some VZV gene products have functional ORF56 ORF54
ORF57 ORF55
subdomains that are dispensable in cultured cells; others are ORF58
dispensable for replication in vitro but are necessary for ORF59 ORF61
pathogenesis. The ORF9–ORF12 cluster of tegument proteins gL;ORF60
IE63;ORF63 IE62;ORF62
(blue) is conserved in the alphaherpesviruses. The products of ori IR
ORF64
the dispensable genes are of interest for their potential ORF65
PK;ORF66 US
differential functions in tropism. Cloning the VZV genome
gI;ORF67
into bacterial artificial chromosome vectors or as four or five gE;ORF68 ori
ORF71 TR
overlapping fragments in cosmids enables the deletion of ORFs ORF69
or targeted mutations of coding and non-coding sequences ORF70
to define functions in vitro and in vivo (ORTs evaluated for
pathogenesis indicated in bold, part a)40,53,107,108. b Capsid DNA Envelope
Virion formation and structure Tegument proteins
VZV particles are ~80–120 nm in diameter (see the figure,
part b). Linear VZV genomes are packaged into an icosahedral
nucleocapsid core that is formed from proteins encoded by
orf20, orf21, orf23, orf33, orf40 and orf41 (REF. 1) Capsids are
surrounded by a tegument layer, which is a less well-defined
structure that is made up of proteins with known or predicted
regulatory functions, including the immediate-early (IE) viral
Xenografts
Tissues from one species that
transactivating factors that are encoded by orf4, orf62 and
have been transplanted into a orf63, those that are encoded by the orf9–orf12 gene cluster,
different species. the two viral kinases ORF47 and ORF66, and others. The outer
virion component is a lipid membrane envelope that is derived
Severe combined from cellular membranes with incorporated viral glycoproteins,
immunodeficiency including gB/gH–gL, which form the minimal fusion complex.
(SCID). A genetic defect that
blocks the function of T cells
and B cells, interfering with the Nature Reviews | Microbiology
capacity of the host to mount
an effective immune response Investigating VZV pathogenesis is challenging as major tissue tropisms of VZV: T cell-, skin- and neuro­
to foreign proteins. VZV is a highly human-specific virus that has little or tropism13,26,27. In these models, innate responses that
no capacity to infect other species. This obstacle can modulate infectious processes can be assessed indepen-
Dorsal root ganglia be overcome by using human tissue xenografts in mice dently of adaptive immunity, which is absent in SCID
(DRG). Nodules on the dorsal
root of the spinal cord that
with severe combined immunodeficiency (SCID) (BOX 2). mice. VZV-specific T cells are necessary to clear primary
contain sensory nerve cell Infecting foetal thymus-liver T cell, skin and dorsal root infection and prevent symptomatic reactivation from
bodies. ganglia (DRG) xenografts enables studies of the three latency, but the xenograft models show the importance

NATURE REVIEWS | MICROBIOLOGY VOLUME 12 | MARCH 2014 | 199

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Box 2 | Modelling the pathogenesis of varicella zoster virus infection


conditions that are present in intact tissues and fully dif-
ferentiated human cells in vivo. To investigate which host
The pathogenesis of VZV infection can be modelled by infecting xenografts of human cell factors are required during infection, small-molecule
foetal tissues in mice with severe combined immunodeficiency (SCID)26,27,33. The inhibitors or antibodies that block cell functions can be
xenografts are maintained for prolonged periods, as SCID mice lack T cell responses administered. In this Review, we summarize the char-
that mediate rejection of these foreign tissues. Skin xenografts are implanted beneath
acteristics of the infectious process in T cells, skin and
the mouse skin, and thymus-liver or dorsal root ganglia (DRG) xenografts are placed
under the kidney capsule (see the figure). Then the human tissues become vascularized
dorsal root ganglia, the contributions of VZV proteins
by anastomosis of capillaries in the xenografts with those in the surrounding murine and functional motifs within these proteins to the capac-
tissues; endothelial cells of the microvasculature express human platelet endothelial ity of VZV to infect differentiated human cells (TABLE 1)
cell adhesion molecule 1 (PECAM‑1). Skin engraftment is established within 3–4 weeks; and the modulation of VZV infection by host cell fac-
T cell and DRG xenografts require >8–12 weeks. Skin xenografts differentiate to have tors within the tissue microenvironment. Investigating
typical layers of dermal and epidermal cells, and epidermal keratinocytes differentiate pathogenesis using these tools offers insights into how
into the anuclear corneocytes that comprise the stratum corneum, as in postnatal skin VZV causes the clinical manifestations of varicella and
in humans. Thymus-liver xenografts contain immature dual-positive CD4+ CD8+ T cells zoster and how this ubiquitous virus has so successfully
but predominantly contain mature CD4+ T cell and CD8+ T cell populations. Neurons in survived in the human population.
DRG xenografts are surrounded by satellite cells, express neuronal subtype-specific
proteins (such as TRKA and RT97), neural cell adhesion molecule (NCAM), synaptophysin
and other markers such as herpes viral entry mediator (HVEM). After the period of
T cell tropism
engraftment, functional signalling pathways and effector proteins, including those that Discovering VZV tropism for T cells. VZV was initially
are involved in blocking apoptosis and innate antiviral responses (for example, signal classified as a neurotropic herpesvirus, but experiments
transducer and activator of transcription (STAT) proteins, interferons and promyelocytic using T cell xenografts in SCID mice in vivo and tonsil
leukaemia protein (PML)) are present, as expected, in postnatal tissues. Effects of T cells in vitro have revealed that VZV also shows T cell
targeted mutations in the VZV genome are assessed by inoculating surgically exposed tropism13,26,28 (FIG. 2). CD3+ T cells, including CD4+, CD8+
xenografts with fibroblasts that are infected with the parent wild-type or mutant virus, and dual CD4+CD8+ T cell subpopulations, are fully
or with uninfected fibroblasts. The infectious process is similar when infected permissive for the replication and release of infectious
fibroblasts or when infected T cells — which are introduced into the mouse circulation virions. VZV infects tonsil T cells with high efficiency,
— are used to deliver and release virions. To study the consequences of interfering with
which suggests that the virus is transferred from respira-
the functions of viral or cellular proteins, the mice are treated with antibodies or
small-molecule inhibitors. Xenografts are recovered at different intervals after
tory epithelial cells to T cells, presumably in the tonsils
inoculation to assess viral titres, genome copies and viral transcripts; quantitative and other lymphoid tissues that comprise the Waldeyer’s
microscopy of tissue sections shows viral and cellular protein expression, genome ring (FIG. 1), similarly to the transfer of Epstein–Barr virus
localization, virion assembly and ultrastructural changes in cells. Furthermore, to tonsil B cells29. VZV can also infect dendritic cells,
progression of infection can be monitored in vivo using recombinant VZV that which might facilitate spread to lymph nodes30,31. VZV-
expresses firefly luciferase. The SCID mouse model also provides a system for infected CD4+ T cells predominantly show a memory
translational research to assess live attenuated VZV vaccines and antiviral drugs26,96,109. T cell phenotype and express activation markers and
Uninfected VZV-infected
skin-homing proteins, such as cutaneous leukocyte anti-
gen (CLA) and CC‑chemokine receptor 4 (CCR4), and
are thus more likely to circulate through skin and other
Skin tissues28. In addition, VZV induces activation and skin
xenograft
homing proteins on naive T cells.
SCID mouse
In SCID mice with skin xenografts, infected human
tonsil T cells can transport VZV and initiate skin infec-
tion2. At 24 hours after injection of T cells into the circu-
Thymus-liver lation, infected T cells appear in the skin that surrounds
xenograft hair follicles, where the capillary microvasculature is
extensive, which indicates that these cells retain the abil-
ity to transit across capillary endothelial walls by diapede-
Kidney sis. Infected T cells also deliver VZV to DRG xenografts
DRG in vivo, which suggests that viraemia facilitates the estab-
xenograft
lishment of latency27. VZV promotes survival of infected
T cells by inducing signal transducer and activator of
transcription 3 (STAT3), which gives these cells time to
Nature Reviews | Microbiology reach tissues32. Notably, VZV does not trigger fusion of
of intrinsic responses of differentiated cells in the infected T cells, which indicates that virions must enter
Anastomosis absence of an adaptive immune response. Such studies each T cell separately13,26,33. This characteristic differs
The end‑to‑end connection of
can be done in knockout mouse models that have defects from the multinucleated syncytia that are formed in
tubular structures, such as
capillaries. in adaptive immunity, but VZV does not infect mice. VZV-infected skin34 and suggests that VZV has a cell
Furthermore, the xenograft models have the advantage type-specific capacity to suppress fusion, probably to
Waldeyer’s ring of investigating infection in the various human tissue retain the capacity of T cells to enter and exit tissues.
An annular arrangement of microenvironments that are targeted by VZV. Inocu- The ability of VZV to infect memory T cells without dis-
lymphoid tissue in the
oropharynx; it consists of the
lating human tissue xenografts with mutant VZV can rupting their trafficking explains the development of the
pharyngeal, tubal, palatine and show functions of viral genes that are dispensable in scattered varicella lesions and is consistent with the cell-
lingual tonsils. tissue culture but necessary under the more stringent associated viraemia that is observed in clinical cases35–37.

200 | MARCH 2014 | VOLUME 12 www.nature.com/reviews/micro

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Viral determinants of T cell infection. The only glyco- binding sites for cellular transcription factors shows the
proteins that are encoded by the short unique region of importance of the synergistic regulation of viral genes
the VZV genome (BOX 1), gE and gI, have been inves- by IE62 and cellular cofactors. For example, the tran-
tigated for their functions in T cells33,38–40 (TABLE 1). In scription factor specificity protein 1 (Sp1) increases the
contrast to its homologues in other alphaherpesviruses, recruitment of IE62 to viral promoters11,54, and mutation
the VZV membrane glycoprotein gE has a large unique of the two Sp1-binding motifs in the gE promoter pre-
amino-terminal region (consisting of amino acids vents VZV replication55. The gE promoter also has a site
1–188), which is essential for replication39. Within this for upstream stimulatory factor (USF) binding, but it is
region, gE amino acid residues 51–187 are important not needed for the infection of tonsil T cells. Inhibiting
for secondary virion envelopment and T cell infection. the binding of Sp1 and USF to the gI promoter impairs
Although gE typically forms a heterodimer with gI via VZV replication in T cell xenografts43. Thus, cellular
a cysteine-rich ectodomain region at residues 208–236, cofactors that interact with IE62 have differential effects
this interaction is not needed to infect T cells38,39. In addi- on particular viral gene promoters, which influence
tion, amino acid residues 27–90 are required for binding VZV replication in T cells in vivo.
to insulin-degrading enzyme (IDE), which is a cellular
protein that contributes to VZV replication in melanoma Skin tropism. Innate cellular responses regulate skin
cells in vitro and is a proposed entry receptor41. However, pathogenesis. Lesion formation in infected skin xeno-
disrupting these interactions by mutating gE does not grafts is a highly regulated process that is determined by
affect T cell-xenograft infection, which indicates that robust innate responses of epidermal and dermal cells
IDE is not needed for T cell entry. The gE cytoplasmic (FIG. 3). Delivery of VZV to skin xenografts by infected
domain contains an endocytosis motif that is required T cells leads to the gradual formation of skin lesions over
for replication, whereas a TGN-targeting motif in this 10–21 days, which is consistent with the varicella incu-
domain is dispensable for replication in vitro but is bation period2. Infected cells initially appear around hair
important in T cells in vivo33. VZV gI is required for T follicles and viral proteins are then detected in clusters
cell infection42, and reducing gI synthesis by disrupting of adjacent cells, some of which fuse to create multinu-
gI promoter sites for interactions with cellular transcrip- cleated polykaryocytes37. The uninfected cells that sur-
tion factors or ORF29 DNA-binding protein impairs round infectious foci show upregulation of interferon-α
replication in T cell xenografts43. (IFNα) and IFNβ and the cellular transcription factors
Five of the tegument proteins that surround the viral STAT1 and nuclear factor-κB (NF-κB), which orches-
capsid have been evaluated for their contributions to trate innate immune responses2,56.
T cell infection, including ORF10 protein, IE63, ORF65 The importance of the IFNα and IFNβ response is
protein and the two viral kinases, ORF47 protein and evident from the enlarged lesions in skin xenografts
ORF66 protein13,16,17,44,45. ORF10 protein increases IE62 that form if it is blocked with an antibody that targets
expression but is not required for T cell infection44. IE63 the IFNα and IFNβ receptor2. During varicella in the
also regulates IE62 expression and transactivates cel- human host, T cell-mediated immunity only occurs late
lular elongation factor 1α (EF‑1α)49–52: IE63 functions in infection and is rarely detected until skin lesions have
that are needed to infect T cell xenografts are preserved developed1. Adaptive responses are important to stop
even when the extensive serine/threonine phosphoryla- the infection57, but the initial control of viral replication
tion domain is disrupted45. The ORF65 protein is also by innate responses probably contributes to the persis-
dispensable53. In contrast, the VZV kinases have a major tence of VZV in the population, as a severe infection
role in T cells: the kinase activity of ORF47 directs the that overwhelms the host would limit opportunities for
intracellular localization of IE62, gE and ORF47 and is VZV transmission to other susceptible individuals.
needed for virion assembly. Although these functions are Notably, replication of the vaccine virus Oka is
dispensable in vitro12,14,34, blocking ORF47 expression or reduced in skin xenografts, which is consistent with
disabling its kinase activity prevents VZV replication in the situation in humans, who rarely develop lesions
T cell xenografts, which shows their importance in vivo13. after subcutaneous inoculation of the vaccine4,5,26. Thus,
Whereas ORF47 is conserved in all herpesviruses, only the mutations that have accumulated in this attenu-
alphaherpesviruses have ORF66 kinase homologues1,7. ated strain (via passage in fibroblasts) have reduced its
ORF66 is important for virion assembly in differentiated capacity to overcome intrinsic cutaneous barriers, even
T cells in vivo17,18 and in human corneal fibroblasts but not though the mutations have no effect on T cell tropism26
in other cells cultured in vitro15,18,46 (FIG. 1b). Furthermore, or neurotropism5. The Oka vaccine consists of a mix-
the ORF66 kinase inhibits apoptosis and counteracts the ture of polymorphic viral clones, and some polymor-
Diapedesis induction of interferon (IFN) signalling in infected phisms seem to be common, but none were identified
The passage of lymphocytes
T cells17,18. ORF66 also contributes to the downregulation as the molecular basis for attenuation. The evaluation
and cells of the immune system
through intact vessel walls. of the major histocompatibility complex I and thereby inter- of chimeric viruses constructed from segments of the
feres with CD8+ T cell-mediated elimination of infected pathogenic parental Oka strain and the attenuated vac-
Major histocompatibility cells47,48. Thus, the ORF66 kinase, which is dispensable cine in skin xenografts indicated that several genome
complex I in vitro, has multiple functions in T cells in vivo. regions contribute to viral attenuation in the skin 58,
A protein that is present on the
surfaces of most mammalian
which is consistent with the presence of a mixture
cells that functions to present Cellular transcription factors as determinants of T cell of VZV clones with varying sequence changes in Oka
peptide epitopes to T cells. tropism. Investigating VZV mutants that have disrupted vaccine preparations59.

NATURE REVIEWS | MICROBIOLOGY VOLUME 12 | MARCH 2014 | 201

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Table 1 | VZV protein functions in the pathogenesis of T cell, skin and DRG infection
Protein Characteristics Cell culture* Xenografts*
Glycoproteins
gB (ORF31) Mature form generated by furin cleavage; Essential Mutation of furin-cleavage site: impaired replication in skin
ITIM motif in cytoplasmic domain; required
for cell fusion together with gH–gL ITIM mutation; severely impaired replication and enhanced fusion in skin

gH (ORF37) Ectodomain comprising domains I–III; forms Essential Mutation of domain I amino terminus: impaired replication in skin
heterodimer with gL; required for cell fusion
together with gB and gL Mutation of domain III fusion loop; impaired replication in skin
gL (ORF60) Forms heterodimer with gH Presumed Not tested
essential
gE (ORF68) Large unique N terminus (amino acids Essential; Deletion of amino acid residues 51–187: no replication in T cells and skin
1–187); forms heterodimer with gI (amino deletion of
acids 208–236); binds to IDE; TGN-targeting amino acids Mutation of gI-binding domain: no effect in T cells, severely impaired
and endocytosis motifs in cytoplasmic 1–187 blocks replication in skin, prolonged replication and severe tissue damage in DRG
domain; MSP‑gE is a natural variant endocytosis Mutation of IDE-binding domain: no effect in T cells, impaired
replication in skin, no effect in DRG
Mutation of TGN-targeting motif: impaired replication in T cells and
skin, no effect in DRG
MSP-gE: enhanced replication in skin
gI (ORF67) Forms heterodimer with gE (via amino acids Dispensable, Deletion of gI: no replication in T cells and skin, prolonged replication
105–125 and conserved cysteine residues); decreased in DRG
required for incorporation of gE into virions titres and
plaque size Mutation of gE-binding domain: prolonged replication in DRG
Deletion of amino acids 105–125: no replication in skin
Mutation of Sp1–USF-binding motif: no replication in T cells, impaired
replication in skin, no effect in DRG
Mutation of ORF29-binding motif: impaired replication in T cells
Regulatory proteins
ORF9 Tegument protein; binds to IE62 and Essential Not tested
ORF11; functions unknown
ORF10 Tegument protein; ORF62 and ORF71 Dispensable Deletion of ORF10: no effect in T cells, impaired replication in skin
transactivator; required for efficent virion
assembly
ORF11 Tegument protein; required for normal Dispensable Deletion of ORF11: impaired replication in skin
levels of IE4, IE62, IE63 and gE; RNA-binding
domain (amino acids 1–22); binding to Mutation of RNA-binding motif: no effect in skin
ORF9 required for efficient virion assembly Mutation of ORF9-binding motif: no replication in skin

VZV manipulation of cellular responses. VZV sup- Skin infection was severely impaired by treating SCID
presses innate responses to produce a virus-filled lesion mice with a small-molecule inhibitor of STAT3, S31‑201
at the skin surface, where infected keratinocytes release (FIG. 3). STAT3-mediated upregulation of survivin, which
newly-assembled virions60 (FIG. 3). Several viral proteins is a cell protein that inhibits apoptosis, was necessary to
interfere with IFN-mediated responses in infected cells: support VZV infection. Although oncogenic herpesvi-
IE62 inhibits the phosphorylation of IFN regulatory fac- ruses manipulate this pathway to cause tumours, these
tor 3 (IRF3) by TANK-binding kinase 1 (TBK‑1), thus experiments showed that VZV, which is a lytic herpes-
blocking IFNβ production61, ORF47 kinase reduces IRF3 virus, must also induce survivin. Interestingly, survivin
phosphorylation62 and IE63 inhibits the phosphorylation expression is constitutive in the epithelial cells that line
of the eukaryotic initiation factor 2 and its downstream the hair follicles, which is the initial site of VZV replica-
IFNα effects63. The ORF66 kinase inhibits the activation tion in skin. VZV triggers autophagosome formation in
of STAT1, which is induced in response to IFNα and skin cells, which might also reinforce the delay of apop-
IFNβ signaling and upregulates IFN-stimulated genes tosis, thereby favouring infection64.
(ISGs)17. VZV also sequesters the p50–p65 heterodimer,
which is the most abundant form of NFκB, in the cyto- Promyelocytic leukemia protein (PML) in the regulation
plasm of epidermal cells, by preventing degradation of of skin infection. PML is a multifunctional protein that
the NF-κB inhibitor-α (IκBα)56. The latter effect is cell has antiviral effects against many viruses and is upreg-
type-specific, as VZV induces NF-κB in monocytes48. ulated by IFNs. In VZV-infected cells, PML can form
In addition to its inhibitory effects, VZV repro- intranuclear cages that trap nascent virions and restrict
grammes cell signalling to activate STAT3 in epidermal their egress from the nucleus to the cytoplasm65,66. This
cells as well as T cells in vivo and fibroblasts in vitro32. antiviral effect is mediated by PML isoform IV, which

202 | MARCH 2014 | VOLUME 12 www.nature.com/reviews/micro

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Table 1 (cont.) | VZV protein functions in the pathogenesis of T cell, skin and DRG infection
Protein Characteristics Cell culture* Xenografts*
ORF12 Tegument protein; activates cell signalling Dispensable Deletion of ORF12: no effect in skin
pathways (such as ERK, p38, JNK and PI3K–
AKT) that inhibit apoptosis and support viral
replication
ORF47 Serine/threonine kinase (conserved); Dispensable Deletion of ORF47: no replication in T cells and skin
phosphorylates regulatory proteins and
glycoproteins; binds to IE62; required for Mutation of kinase motif: no replication in T cells, impaired replication
virion assembly in skin

ORF61 Transactivator or repressor; dimerization Essential‡ Deletion of SIMs: impaired replication in skin
required for regulatory functions; E3 ligase;
PML dispersal by SIMs Deletion of dimerization domain (amino acids 250–320): impaired
replication in skin
ORF62 and IE62 is major viral transactivator; IFN Essential Deletion of ORF62 and ORF71 with ectopic ORF62: no replication in
ORF71 inhibition skin
ORF63 and IE63 phosphoprotein; represses IE62; Essential‡ Deletion of ORF63 and ORF70 with ectopic ORF63: no effect in T cells
ORF70 transactivates EF‑1α and skin
Mutation of phosphorylation sites: no effect in T cells, impaired
replication in skin
ORF66 Serine/threonine kinase Dispensable Deletion of ORF66: impaired replication in T cells, slightly impaired
(alphaherpesviruses); phosphorylates IE62; replication in skin
inhibits apoptosis
Mutation of kinase motif: impaired replication in T cells, slightly
impaired replication in skin
Other VZV proteins
ORF23 Small capsid surface protein; required for Dispensable Deletion of ORF23: no replication in skin
nuclear transport of other capsid proteins
and capsid assembly
ORF35 Cell fusion Dispensable, Deletion of ORF35: slightly impaired replication in T cells, impaired
decreased cell replication in skin
fusion
ORF64 and Cell fusion Dispensable, Deletion of ORF64 and ORF69: no effect in T cells and skin
ORF69 increased cell
fusion
ORF65 Virion protein Dispensable Deletion of ORF65: no effect in T cells and skin
DRG, dorsal root ganglia; EF-1α, elongation factor 1α; ERK, extracellular signal-regulated kinase; IDE, insulin-degrading enzyme; IE, immediate-early; IFN,
interferon; ITIM, immunoreceptor tyrosine-based inhibition; JNK, c-JUN N-terminal kinase; PML, promyelocytic leukaemia protein; SIMs, SUMO-binding motifs;
TGN, trans-Golgi network; VZV, varicella zoster virus. *Function in cells and tissues was assessed by mutagenesis of the VZV genome using cosmids or bacterial
artificial chromosomes (BACs) to delete or insert stop codons or introduce targeted changes in the coding sequence ‡Indicates differences among published
observations about whether the gene is essential or dispensable, including variations in the mutations tested and/or experimental conditions used to assess
growth requirement.

binds to the ORF23 outer capsid protein via a unique Cell–cell fusion and skin tropism. Cell–cell fusion is
carboxy‑terminal domain and sequesters virions. not strictly required for VZV spread, as virions that are
PML is a predominant component of nuclear struc- released from infected cells can enter adjacent cells19.
tures known as PML nuclear bodies (PML-NBs; also However, polykaryocyte formation is the classic patho-
known as nuclear domain 10). PML-NBs are more abun- logical change that is induced by VZV in the skin. This
dant in skin than in cultured cells, and VZV infection and pattern suggests that VZV reprogrammes infected cells to
the concomitant IFN secretion cause further accumula- overcome the normal preservation of plasma membrane
tion in uninfected, adjacent epidermal and dermal cells, boundaries between differentiated cells and to mediate
thus potentially limiting viral replication in vivo67 (FIG. 3). fusion of human skin cells in vivo, leading to facilitated
The ORF61 protein, which is expressed very early after virus spread, which overcomes innate barriers.
infection19, functions to counteract this antiviral effect by Glycoprotein gB and the heterodimer that is formed
binding to PML via SUMO-interacting motifs (SIMs) and by gH and gL constitute the minimal VZV fusion com-
disrupting the architecture of PML-NBs (FIG. 3). Mutation plex and are candidates for mediating virus-induced
of the ORF61 SIMs has no consequences in vitro, but, in cell fusion as well as virion entry9,20,23,24 (FIG. 1b; TABLE 1).
skin, it enables PML-NBs to remain abundant, severely The gB ectodomain has a highly conserved primary
impairs VZV replication and prevents viral penetration fusion loop that is essential for replication, and a furin
of the cutaneous basement membrane67. Again, IE62 protease recognition motif, which is not found in most
synergism with Sp1 is necessary, as interfering with the alphaherpesvirus homologues24. Disrupting furin cleav-
binding of Sp1 to the ORF61 promoter limits ORF61 age attenuates viral infection of skin xenografts, which
expression and reduces the extent of skin lesions68. indicates that this post-translational modification of gB

NATURE REVIEWS | MICROBIOLOGY VOLUME 12 | MARCH 2014 | 203

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Lymphocyte Tonsillar crypt

Mutant VZV Memory marker


(∆gl, ∆gE, ∆OR47,
∆ORF66) Skin homing Hair folicle
markers
Wild-type VZV

pSTAT3,
pZap70,
pAkt
pSTAT1

Efferent lymphatics connected Skin microvasculature


to systemic circulation
Activated CD4+ T cell

Figure 2 | VZV T cell tropism.  According to the model of varicella zoster virus (VZV) cell-associated viraemia, tonsil
T cells are infected following VZV inoculation and replication in respiratory mucosal epithelial Nature
cells.Reviews | Microbiology
T cells traffic into and
out of tonsils across the squamous epithelial cells that line the tonsilar crypts (left panel). VZV has increased tropism for
activated memory T cells that have skin-homing markers, which are common in tonsils (centre panel). These T cells are
programmed for immune surveillance and can transport the virus across capillary endothelial cells into skin. VZV
glycoprotein E (gE) (through its unique amino terminus), gI and the viral kinases ORF47 and ORF66 are important for T cell
infection. Proteins that regulate cellular gene expression are activated (in the case of signal transducer and activator of
transcription 3 (STAT3)) or inhibited (in the case of STAT1) in infected T cells. The microvasculature is extensive at the base
of hair follicles, where T cells transit into the surrounding skin and initial VZV replication is observed (right panel).

is important for pathogenesis. According to the current is genetically stable, and unrelated isolates exhibit little
model, herpesvirus gH (together with gL) activates gB variability in virulence69. However, VZV-MSP, which is
to trigger its fusogenic function. For VZV, this function a naturally occurring variant with a single amino acid
seems to depend, in part, on gH ectodomain residues and change (D150N) in the gE ectodomain, has increased
is required for skin pathogenesis, as administering a gH- cell–cell spread in vitro and accelerated growth in skin
specific monoclonal antibody blocks replication in skin xenografts compared with other isolates70. Highlighting
xenografts22. Residues in the extreme amino‑terminus the potential effect of changing a single gE residue, a ser-
of gH (which are dispensable in vitro) also contribute to ine to alanine substitution at gE position 31 markedly
skin infection in vivo23. impairs skin replication39.
Although cell–cell fusion is a prominent character- Residues 51–187 in the non-conserved ectodomain
istic of VZV pathogenesis, it must be tightly controlled of gE are essential for both skin tropism and T cell tro-
during skin infection. Appropriately regulated cell–cell pism39, which reinforces the importance of this unique
fusion depends on the cytoplasmic domain of gB and is gE region for the VZV life cycle (TABLE 1). Blocking gE–gI
mediated by a previously unrecognized immunorecep- heterodimerization interferes with gE maturation and
tor tyrosine-based inhibition motif (ITIM) in the gB surface expression and inhibits the incorporation of gI
cytoplasmic domain25 (FIG. 3). Phosphorylation of tyro­ into virions. These functions that involve gE–gI-binding
sine 881 (Y881) in the ITIM modulates VZV-induced are important for skin tropism but are dispensable in
fusion, and blocking ITIM phosphorylation leads to T cells38. Binding of gE to IDE also contributes to skin
a hyperfusogenic phenoptye in vitro and reduces the tropism, but blocking the interaction of gE with this cel-
production of infectious virions. Inhibition of ITIM lular protein has much less effect than eliminating gE–gI
phosphorylation in vivo causes the aberrant fusion of binding. Overall, gE domains, including the TGN-tar-
skin cells in the outermost layer and markedly impairs geting motif in the C-terminus of gE33, are more impor-
lesion formation and infectious virus yields, which sug- tant for the infection of skin than of T cells, which sug-
gests that the gB cytoplasmic domain is essential for the gests that gE functions determine cell–cell spread rather
control of polykaryocyte formation and optimal skin than initial virion entry. Binding sites for cellular factors
infection. in the gE promoter, other than Sp1, are not necessary for
replication or skin tropism55.
Glycoproteins gE and gI as determinants of skin infec- Deleting gI interferes with gE trafficking and sec-
tion. Although VZV isolates can be classified into several ondary envelopment of virions in vitro and completely
distinct clades that reflect their geographical origin, VZV blocks infection of skin and T cells in vivo, which can

204 | MARCH 2014 | VOLUME 12 www.nature.com/reviews/micro

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Viral factors Cellular factors

IFN-α pre IL-1α(+)


Epidermis
pSTAT1 IFN-α pSTAT1
Multinucleated cell
Innate cytokine barrier
pSTAT3 Survivin
Dermis

Cytokine PML Cell with


expressing cell pSTAT3
upregulation
Nerve axons T cell
VZV lesion VZV lesion Uninfected skin cells

ORF61 dispersal gB regulation pSTAT3 and survivin pSTAT1 upregulation


of PML bodies of cell fusion upregulation

Figure 3 | VZV skin tropism.  The schematic illustrates viral factors that ensure spread to the skin surface after varicella
Nature Reviews | Microbiology
zoster virus (VZV) is delivered to cutaneous sites of replication by infected T cells or by retrograde axonal transport from
neurons (left-hand side). Two examples of VZV proteins that are important for pathogenesis are shown: ORF61 protein has
SUMO-interacting motifs that are important for dispersal of promyelocytic leukaemia nuclear bodies (PML-NBs)67 and the
cytoplasmic domain of glycoprotein B (gB) has an immunoreceptor tyrosine-based inhibition motif that regulates cell–cell
fusion and polykaryocyte formation25. VZV replication in skin triggers cellular responses, including changes that are
induced in infected cells and changes in the uninfected cells adjacent to infected cells. Examples of VZV effects within
infected cells are illustrated (right-hand side). VZV induces signal transducer and activator of transcription 3 (STAT3)
activation, which triggers the expression of the anti-apoptotic protein survivin and inhibits the expression of interferon-α
(IFNα) and STAT1 (REF. 32). In contrast to infected cells, surrounding uninfected cells exhibit upregulation of IFNs, STAT1,
which activates IFN-stimulated factors such as PML, and other cell transacivators and innate cytokines2.

be attributed to defective virion assembly41,71. Ensur- A single copy of orf63 at a non-native site in the genome
ing that gI expression levels are sufficient for normal provides sufficient IE63 for replication in skin; however,
skin pathogenesis depends on the enhancing effect of lesion formation is substantially impaired if IE63 phos-
the ORF29 DNA-binding protein for IE62 transactiva- phorylation is disrupted, which indicates that there are
tion43. Co‑regulation of the gI promoter by Sp1 and USF differential requirements for the IE63 phosphoprotein
with IE62 is essential in skin, whereas it enables a low in skin and in T cells45.
level of replication in T cell xenografts43, which indi- The VZV tegument contains proteins that are
cates tissue-specific differences in the requirement for encoded by a cluster of genes (known as orf9–orf12) that
cell transactivators that influence levels of gI expression. is conserved in alphaherpesvirus genomes7. Of these
genes, only orf9 is essential in vitro. The ORF10 protein
Tegument and capsid proteins in skin infection. Like increases expression of IE62 and is important for skin
other herpesviruses, the VZV genome has duplicate infection, which shows that interactions of IE62 with
copies of genes in the repeat regions, including those viral cofactors — which are dispensable in vitro — are
that encode the tegument proteins, IE62 (encoded by necessary for pathogenesis in vivo44,73. The ORF11 pro-
orf62 and orf71), IE63 (encoded by orf63 and orf70) tein contributes essential functions in skin xenografts via
and ORF64 protein (encoded by orf64 and orf69)1,7 its binding to the ORF9 protein during virion tegument
(BOX 1). The importance of this genomic organization assembly; however, the capacity of the ORF11 protein to
for pathogenesis is not known. Deleting the genes for bind to mRNA is not necessary, which shows that distinct
IE62 or IE63 is lethal for VZV replication but the orf64– functions of the same VZV protein differ in their effects
orf69 gene pair is dispensable, which indicates that gene on skin tropism74. The ORF12 protein — which activates
duplication is not conserved to protect against the loss ERK1/2, inhibits apoptosis75 and manipulates cell cycle
of an essential gene50. However, when skin xenografts progression by activating the PI3K/AKT pathway76 — is
are infected with VZV mutants that have single copies dispensable in skin73. Thus, although functions of some
of the genes for IE62 and IE63 genes, the second copy gene products are essential, conservation of the orf9–12
is restored by recombination during replication, which cluster does not signify a uniform requirement of these
suggests that gene duplication is important in vivo45,72. genes for pathogenesis in differentiated human tissues.
IE62 production in skin is necessary, as is evident from Disrupting the ORF47 protein kinase function, which
impaired infection when the binding site for the ORF29 causes the nuclear retention of ORF47 and IE62 proteins
viral cofactor is mutated in the orf62–orf70 promoter. in vitro, severely impairs skin pathogenesis14,34. Deleting

NATURE REVIEWS | MICROBIOLOGY VOLUME 12 | MARCH 2014 | 205

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

Active infection Latency


Axon

Satellite cell

Neuronal
nucleus

Genome
Virion

PML-NB
Neuron– IE62 mRNA
satellite IE63 mRNA
fusion gB mRNA

Fibrosis

Aberrant chronic infection:


• gE binding to gI blocked
• gI deleted

Figure 4 | VZV neurotropism in DRG xenografts.  This schematic illustrates active infection of dorsal root ganglia
(DRG) which is characterized by the transcription of genes (for example, genes encoding glycoprotein B (gB),| Microbiology
Nature Reviews immediate
early protein 62 (IE62) and IE63) that produce proteins that are required for lytic infection, varicella zoster virus (VZV)
genome synthesis, virus assembly in neurons and satellite cells, release of VZV into intracellular spaces and fusion of some
neurons and satellite cells27 (left panel). Virions are captured in cages that are formed by promyelocytic leukaemia nuclear
bodies (PML-NBs) in some neurons and satellite cells66. By contrast, latency (right panel) is associated with the persistence
of VZV genomes and immediate-early (IE) transcripts, whereas late gene transcription, such as transcription of gB, ceases
and virion formation ceases. When DRG are infected with VZV mutants in which binding of gE to gI is blocked or in which
gI is deleted, the transition to latency is disrupted (right panel; outlined box), infectious virions continue to be produced at
low levels and in the case of disrupted binding of gE to gI, tissue destruction is extensive, which is associated with
disruption of the cell matrix, elimination of many neurons and the proliferation of satellite cells.

orf47 also reduces replication in skin organ cultures77. These studies showed that VZV genomes (~2–9 cop-
However, despite reduced virion formation, the ORF47 ies per cell) are present in about 4% of neurons dur-
kinase mutant elicits some polykaryocyte formation in ing latency79,80. Transcripts of ten VZV genes have been
skin xenografts as long as binding of the ORF47 pro- reported in autopsy ganglia, and orf63 transcripts are
tein to IE62 is preserved34. By contrast, replication of the the most abundant81,82, but the extent to which detec-
ORF47 kinase mutant in T cell xenografts is completely tion represents post-mortem release of gene silencing
blocked, which suggests that even minimal cell fusion is not known. In some studies, VZV proteins were rare
can support VZV spread in the skin, whereas T cell or absent, whereas others reported frequent expression
infection depends on the complete assembly and release in the neuronal cytoplasm83,84; however, eliminating an
of infectious virions 13,14. Notably, disabling ORF66 artefact, which results from antibody reactivity against
kinase activity has very little effect on virulence in the blood group A determinants, confirms that VZV pro-
skin, despite its importance for T cell tropism13. Thus, tein expression is rare in latency85. By contrast, ganglia
the ORF47 and ORF66 viral kinases make different and from individuals who had zoster a few months before
tissue-specific contributions to VZV pathogenesis. death harbour VZV proteins and inflammatory proteins
The small capsid protein that is encoded by orf23 is in up to 25% of neurons86,87. Prolonged replication in
dispensable for replication in vitro, owing to redundant ganglia might be a factor in some cases of postherpetic
ORF33.5‑mediated transport of the major capsid pro- neuralgia.
tein into the nucleus78. However, this mechanism is not
sufficient in skin, which indicates that there are strict VZV infection and latency in DRG xenografts. In DRG
requirements for VZV capsid assembly in vivo. xenografts, viral proteins, genomes and infectious virus
are detectable for 3–4 weeks after VZV inoculation27
Neurotropism (FIG. 4). Productive infection is followed by a transition to
Before the DRG-xenograft model was developed, VZV VZV latency after 4–8 weeks. During this phase, infec-
neurotropism in human tissues could only be exam- tious virus is no longer detectable, viral genome copies
ined in sensory ganglia that were obtained at autopsy. decline, levels of orf62 and orf63 transcripts are reduced

206 | MARCH 2014 | VOLUME 12 www.nature.com/reviews/micro

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

and gB transcription ceases. This pattern of DRG infec- no effect on neurotropism95. By contrast, blocking gE–gI
tion is identical after direct inoculation or VZV transfer heterodimer formation impairs cell–cell spread in DRG
from infected T cells. The transition to persistence in and reduces replication during acute infection, with no
neurons occurs despite the absence of adaptive immu- infectious virus detected until four weeks after inocula-
nity and markedly contrasts with the progressive lytic tion. However, instead of transitioning to persistence,
infection that is observed in skin and T cell xenografts. the gE mutant causes widespread cytopathic changes in
Thus, VZV gene silencing in neurons is a cell type-spe- neurons, satellite cells and the surrounding tissue, for at
cific characteristic of the VZV–host interaction. Infec- least two months after inoculation96,97. Thus, the gE–gI
tion of explanted foetal DRG neurons in vitro showed interaction is crucial for preventing a chronic, highly
that IE63 has anti-apoptotic functions that might facili- destructive process in sensory ganglia.
tate this transition88,89. Deleting gI, similarly to blocking heterodimer forma-
As observed in skin xenografts, VZV infection of tion, results in the prolonged production of infectious
DRG is regulated by the capacity of PML-NBs to seques- virus particles despite reduced virion assembly and gE
ter nucleocapsids in neurons and satellite cells65. Cages mislocalization97. Impaired virion assembly is associated
that consist of PML fibres surround both nascent and with the accumulation of aberrant membrane stacks in
fully formed virions in infected cell nuclei. Notably, the TGN region and altered Golgi structures both in
neurodegenerative disorders, such as Huntington’s dis- DRG neurons and in vitro98. These effects, like those that
ease, are also associated with large PML-NBs that retain are observed when gE cannot bind to gI, might be due
aberrant polyglutamine proteins. PML isoform IV cages to the absence of cell fusion, such that the virus can only
have the capacity to sequester both viral capsids and the spread slowly from cell to cell and host innate responses
aberrant Huntington’s disease protein, HttQ72 (REF. 90), that faciliate the transition to persistence are less effec-
which suggests that neurons have a conserved mecha- tively triggered. Of interest, gI promoter regulation by
nism to sense and entrap protein aggregates. Sp1 and USF is not required for neuropathogenesis97, in
How VZV reactivation is triggered is not known, but contrast to T cell and skin infection, which, again, indi-
intrinsic cellular mechanisms might also restore gene cates that cellular factors contribute in a cell type-specific
silencing, resulting in abortive replication. If replication manner to viral gene expression and therefore to tissue
continues, VZV is transported to skin, causing zoster tropism. Surprisingly, the ability of VZV to replicate in
(FIG. 1). VZV-specific T cells then function together with DRG in the absence of gI suggests that the requirements
innate cutaneous responses to control the infection. for VZV infection of neural cells are less stringent than
VZV cell-mediated immunity is deficient in elderly those that are required for infection of T cells or the
and immunocompromised patients, which accounts for skin. Finally, the unexpected finding that disrupting the
more frequent and severe zoster in these populations92. formation of gE–gI heterodimers, or deleting gI, causes
Whether VZV-specific T cells also limit reactivation in chronic infection highlights the need to evaluate effects
ganglia is not known. on neurovirulence in vivo and shows that the conse-
quences of VZV mutations cannot be predicted from
Neuron–satellite-cell fusion. When VZV replicates in their effects in other tissue microenvironments.
DRG xenografts, VZV genomic DNA, viral proteins and
virion production are detectable in both neurons and sat- Perspective
ellite cells27,92. Importantly, VZV can induce fusion and These studies in the SCID mouse model show that virus–
polykaryocyte formation between differentiated neurons host cell interactions result in a well-regulated infectious
and surrounding satellite cells (FIG. 4), whereas herpes process in each of the tissue microenvironments that is
simplex virus 1 (HSV‑1) does not induce cell fusion in important for VZV pathogenesis. In infected cells, the
DRG92,93. Neuron–satellite cell fusion is reported in gan- virus reprogrammes cell signalling pathways by inducing
glia from patients with zoster at the time of death94, which or downregulating cellular factors, such as the STATs, to
indicates that DRG xenograft infections are a model of support replication. VZV also has tissue-specific effects
this consequence of VZV reactivation. This formation that are important for pathogenesis, such as suppressing
of neuron–satellite-cell polykaryons amplifies the the fusion of infected T cells. Although many viral pro-
spread of VZV to neuronal cell bodies in the ganglion. teins are incidental for VZV replication in vitro, special-
As VZV is transported to the skin by neuronal axons that ized functions, not only of complete VZV proteins but
extend from neuronal cell bodies, spread within ganglia also small motifs, as well as single amino acids, are often
would increase the extent of infection of the derma­ crucial for pathogenesis in vivo. To protect these crucial
tome during zoster. These pathological changes are not functions, VZV has a high degree of genome stability69.
Satellite cells readily reversible and help to explain why zoster can be At the same time, innate responses of uninfected cells
Glial cells that surround the
associated with prolonged neuropathic pain1. modulate infection so that the host is not overwhelmed,
surfaces of neurons in the
peripheral nervous system. which benefits both the host and the virus by ensuring
Glycoproteins gE and gI as determinants of neu- that there are opportunities for transmission and persis-
Neuropathic pain rotropism. Both gE and gI have functions that influ- tence in the host population.
Pain that is caused by damage ence VZV pathogenesis in DRG xenografts (TABLE 1). Despite these advances in understanding VZV–host
to the somatosensory system;
it is associated with increased
Infection is not altered when binding of gE to IDE is interactions, many questions about VZV pathogenesis
sensitivity to touch, disrupted, although neural cells express IDE, and inter- remain unresolved; for example, as is the case for other
temperature and other stimuli. ference with the TGN-trafficking motif of gE also has herpesviruses, the triggers of VZV reactivation from

NATURE REVIEWS | MICROBIOLOGY VOLUME 12 | MARCH 2014 | 207

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

latency are unknown. DRG xenografts can be used advances in vaccines and antiviral drugs to prevent or
to investigate VZV–neuron interactions, but axonal mitigate VZV infection. Live attenuated VZV vaccines
transport to skin cannot be examined in this model, are effective in healthy individuals but are not safe for
and stimuli of viral reactivation from latency have immunocompromised patients, in whom they cause
not been explored. By contrast, simian varicella virus viraemia, and they can establish latency and reactivate
(SVV) can be studied in the natural non-human primate in healthy and in immunodeficient individuals4–6,104,105,
host99,100 and has been used to investigate reactivation which is consistent with the lack of attenuation that is
that is associated with immunosuppression84. The SVV observed in T cell and DRG xenografts. The genetic
model avoids the need to acquire human tissue to make basis of their attenuation is not defined58,59. The SCID
xenografts, which is a limitation of the SCID model, mouse model can be exploited for rational VZV vac-
but genetic dissimilarities between SVV and VZV are cine design by incorporating mutations that dampen
likely to lead to differences in pathogenesis in their replication in skin into the viral genome, similarly to the
native hosts. Mechanisms of neuropathic pain cannot current vaccine, and, in contrast to the current vaccine,
be assessed in DRG xenografts, but they can be investi- that also interfere with the capacity to infect T cells or
gated in the rat footpad model101, and infection of enteric to persist in neurons. Antiviral drugs, such as acyclo-
neurons has recently been reported in guinea pigs102. vir and related agents, substantially reduce the risk of
Although cultured neurons lack the surrounding satel- severe or fatal VZV infection in immunocompromised
lite cells that mediate neuronal homeostasis in vivo, viral patients but have little or no effect on postherpetic
protein functions of interest for further study in vivo can neuralgia following zoster in the elderly106. Knowledge
be identified in neuron cultures, as was shown for the about functional motifs of VZV proteins and how the
ORF7 protein103. Although the xenograft models are val- virus reprogrammes differentiated human cells in vivo
uable for probing intrinsic antiviral defences, advances might help in designing small-molecule inhibitors
in establishing a human immune system in SCID mice with antiviral activity that would also decrease post-
might make it possible to assess adaptive immune clear- herpetic neuralgia. Understanding the principles of
ance of VZV in conjunction with innate control. VZV pathogenesis at the molecular level has the poten-
From a clinical perspective, VZV remains a medi- tial to yield new approaches to prevent and treat VZV
cally important human herpesvirus despite major infections.

1. Arvin, A. M. & Gilden, D. in Fields Virology 6th Edn for phosphorylation of ORF63 protein, the VZV 22. Vleck, S. E. et al. Anti-glycoprotein H antibody impairs
(Eds Knipe, D., Howley, P.) 2015–2057 (Lippincott homolog of herpes simplex virus ICP22. J. Virol. 69, the pathogenicity of varicella-zoster virus in skin
Williams & Wilkins, 2013). 7367–7370 (1995). xenografts in the SCID mouse model. J. Virol. 84,
2. Ku, C. C. et al. Varicella-zoster virus transfer to skin by 13. Moffat, J. F. et al. The ORF47 and ORF66 putative 141–152 (2010).
T cells and modulation of viral replication by epidermal protein kinases of varicella-zoster virus determine 23. Vleck, S. E. et al. Structure–function analysis of
cell interferon‑α. J. Exp. Med. 200, 917–925 (2004). tropism for human T cells and skin in the SCID‑hu varicella-zoster virus glycoprotein H identifies domain-
This paper shows the role of T cells in the transport mouse. Proc. Natl Acad. Sci. USA 95, 11969–11974 specific roles for fusion and skin tropism. Proc. Natl
of VZV and the control of replication by the potent (1998). Acad. Sci. USA 108, 18412–18417 (2011).
innate response of skin cells. 14. Besser, J. et al. Differentiation of varicella-zoster virus This study maps the functional domains in the gH
3. Gilden, D. H. et al. Varicella-zoster virus DNA in human ORF47 protein kinase and IE62 protein binding ectodomain that are required for replication and
sensory ganglia. Nature. 306, 478–480 (1983). domains and their contributions to replication in skin infection.
This study provides the first evidence of VZV human skin xenografts in the SCID‑hu mouse. J. Virol. 24. Oliver, S. L. et al. Mutagenesis of varicella-zoster virus
latency in neurons, using in situ hybridization to 77, 5964–5974 (2003). glycoprotein B: putative fusion loop residues are
detect VZV genomes. 15. Erazo, A. & Kinchington, P. R. Varicella-zoster virus essential for viral replication, and the furin cleavage
4. Takahashi, M. Clinical overview of varicella vaccine: open reading frame 66 protein kinase and its motif contributes to pathogenesis in skin tissue
development and early studies. Pediatrics. 78, relationship to alphaherpesvirus US3 kinases. Curr. in vivo. J. Virol. 83, 7495–7506 (2009).
736–741 (1986). Top. Microbiol. Immunol. 342, 79–98 (2010). 25. Oliver, S. L. et al. An immunoreceptor tyrosine-based
5. Gershon, A. A. & Gershon, M. D. Perspectives on 16. Schaap, A. et al. T‑cell tropism and the role of ORF66 inhibition motif in varicella-zoster virus glycoprotein B
vaccines against varicella-zoster virus infections. Curr. protein in pathogenesis of varicella-zoster virus regulates cell fusion and skin pathogenesis. Proc. Natl
Top. Microbiol. Immunol. 342, 359–372 (2010). infection. J. Virol. 79, 12921–12933 (2005). Acad. Sci. USA 110, 1911–1916 (2013).
6. Weibel, R. E. et al. Live attenuated varicella virus 17. Schaap-Nutt, A. et al. ORF66 protein kinase function This paper shows that the gB cytoplasmic domain
vaccine: efficacy trial in healthy children. N. Engl. is required for T‑cell tropism of varicella-zoster virus has a previously unrecognized motif that mimicks
J. Med. 310, 1409–1415 (1984). in vivo. J. Virol. 80, 11806–11816 (2006). cellular ITIMs and that must be phosphorylated to
7. Cohen, J. I. The varicella-zoster virus genome. Curr. 18. Eisfeld, A. J. et al. Phosphorylation of the varicella- control the fusogenic properties of gB and the
Top. Microbiol. Immunol. 342, 1–14 (2010). zoster virus (VZV) major transcriptional regulatory detrimental effects of exaggerated cell fusion on
8. Chen, J. J. et al. Mannose 6‑phosphate receptor protein IE62 by the VZV open reading frame 66 VZV infection of skin.
dependence of varicella zoster virus infection in vitro protein kinase. J. Virol. 80, 1710–1723 (2006). 26. Moffat, J. F. et al. Attenuation of the vaccine Oka
and in the epidermis during varicella and zoster. Cell. This paper shows the role of the ORF66 viral strain of varicella-zoster virus and role of
119, 915–926 (2004). kinase in activating the transcriptional function of glycoprotein C in alphaherpesvirus virulence
9. Suenaga, T. et al. Myelin-associated glycoprotein IE62. demonstrated in the SCID‑hu mouse. J. Virol. 72,
mediates membrane fusion and entry of neurotropic 19. Reichelt, M., Brady, J. & Arvin, A. M. The replication 965–974 (1998).
herpesviruses. Proc. Natl Acad. Sci. USA 107, cycle of varicella-zoster virus: analysis of the kinetics of This study is the first report to model VZV
866–871 (2010). viral protein expression, genome synthesis, and virion pathogenesis using T cell and skin xenografts in
10. Yang, M., Hay, J. & Ruyechan, W. T. Varicella-zoster assembly at the single-cell level. J. Virol. 83, SCID mice.
virus IE62 protein utilizes the human mediator 3904–3918 (1999). 27. Zerboni, L. et al. Varicella-zoster virus infection of
complex in promoter activation. J. Virol. 82, This study documents the dynamics of VZV protein human dorsal root ganglia in vivo. Proc. Natl Acad.
12154–12163 (2008). expression, genome replication and progeny virus Sci. USA 102, 6490–6495 (2005).
This paper documents the VZV takeover of the host assembly in newly infected cells from entry to virion This paper is the first report to model the
cell gene transcription complex for viral gene release. neurotropism of VZV using DRG xenografts in SCID
transcription in combination with the IE62 major 20. Grose, C., Carpenter, J. E., Jackson, W. & Duus, K. M. mice.
viral transactivator. Overview of varicella-zoster virus glycoproteins gC, gH 28. Ku, C. C. et al. Tropism of varicella-zoster virus for
11. Ruyechan, W. T. Roles of cellular transcription factors and gL. Curr. Top. Microbiol. Immunol. 342, 113–128 human tonsillar CD4+ T lymphocytes that express
in VZV replication. Curr. Top. Microbiol. Immunol. 42, (2010). activation, memory and skin homing markers. J. Virol.
43–65 (2010). 21. Maresova, L. et al. Characterization of interaction of 76, 11425–11433 (2002).
12. Heineman, T. C. & Cohen, J. I. The varicella-zoster virus gH and gL glycoproteins of varicella-zoster virus: their 29. Longnecker, R., Kieff, E. & Cohen, J. in Fields Virology
(VZV) open reading frame 47 (ORF47) protein kinase processing and trafficking. J. Gen. Virol. 81, 6th Edn (Eds Knipe, D., Howley, P) 1898–1957
is dispensable for viral replication and is not required 1545–1552 (2000). (Lippincott Williams & Wilkins, 2013).

208 | MARCH 2014 | VOLUME 12 www.nature.com/reviews/micro

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

30. Abendroth, A. et al. Varicella-zoster virus infection of 49. Bontems, S. et al. Phosphorylation of varicella-zoster 68. Wang, L., Sommer, M., Rajamani, J. & Arvin, A. M.
human dendritic cells and transmission to T cells: virus IE63 protein by casein kinase influences its Regulation of the ORF61 promoter and ORF61
implications for virus dissemination in the host. cellular localization and gene regulation activity. functions in varicella-zoster virus replication and
J. Virol. 75, 6183–6192 (2001). J. Biol. Chem. 277, 21050–21060 (2002). pathogenesis. J. Virol. 83, 7560–7572 (2009).
This paper shows the capacity of VZV to infect 50. Sommer, M. H. et al. Mutational analysis of varicella- 69. Breuer, J. VZV molecular epidemiology. Curr. Top.
dendritic cells in skin and their capacity to transfer zoster virus ORF63/70 and ORF64/69. J. Virol. 75, Microbiol. Immunol. 342, 15–42 (2010).
the virus into T cells. 8224–8239 (2001). 70. Santos, R. A. et al. Varicella-zoster virus gE escape
31. Huch, J. H. et al. Impact of varicella-zoster virus on 51. Hoover, S. E. et al. Downregulation of varicella-zoster mutant VZV-MSP exhibits an accelerated cell‑to‑cell
dendritic cell subsets in human skin during natural virus (VZV) immediate-early ORF62 transcription by spread phenotype in both infected cell cultures and
infection. J. Virol. 84, 4060–4072 (2010). VZV ORF63 correlates with virus replication in vitro SCID‑hu mice. Virology 275, 306–317 (2000).
32. Sen, N. et al. STAT3 activation and survivin induction and with latency. J. Virol. 80, 3459–3468 (2006). This paper provides the first evidence of a naturally
by varicella zoster virus promotes viral replication and 52. Zuranski, T. et al. Cell-type-dependent activation of the occurring variant of VZV that has an increased
skin pathogenesis. Proc. Natl Acad. Sci. USA 109, cellular EF‑1α promoter by the varicella-zoster virus capacity to replicate in skin.
600–605 (2012). IE63 protein. Virology 338, 35–42 (2005). 71. Oliver, S. L. et al. Mutagenesis of varicella-zoster virus
This paper shows that VZV, which is a lytic 53. Niizuma, T. et al. Construction of varicella-zoster virus glycoprotein I (gI) identifies a cysteine residue critical
herpesvirus, induces STAT3 activation — leading to recombinants from parent Oka cosmids and for gE/gI heterodimer formation, gI structure, and
survivin expression, which was previously demonstration that ORF65 protein is dispensable for virulence in skin cells. J. Virol. 85, 4095–4110 (2011).
associated with oncogenic herpesviruses — and infection of human skin and T cells in the SCID‑hu 72. Sato, B. et al. Mutational analysis of open reading
that takeover of this cellular pathway is necessary mouse model. J. Virol. 77, 6062–6065 (2003). frames 62 and 71, encoding the varicella-zoster virus
for VZV skin infection. 54. Peng, H. et al. Interaction between the varicella-zoster immediate-early transactivating protein, IE62, and
33. Moffat, J. F. et al. Functions of C‑terminal domain of virus IE62 major transactivator and cellular effects on replication in vitro and in skin xenografts in
varicella-zoster virus glycoprotein E in viral replication transcription factor SP1. J. Biol. Chem. 278, the SCID‑hu mouse in vivo. J. Virol. 77, 5607–5620
in vitro and skin and T‑cell tropism in vivo. J. Virol. 78, 38068–38075 (2003). (2003).
12406–12415 (2004). 55. Berarducci, B., Sommer, M., Zerboni, L., Rajamani, J. 73. Che, X. et al. Functions of the ORF9‑to‑ORF12 gene
34. Besser, J., Ikoma, M. & Fabel, K. Differential & Arvin, A. M. Cellular and viral factors regulate the cluster in varicella-zoster virus replication and in the
requirement for cell fusion and virion formation in the varicella-zoster virus gE promoter during viral pathogenesis of skin infection. J. Virol. 82,
pathogenesis of varicella-zoster virus infection of skin replication. J. Virol. 81, 10258–10267 (2007). 5825–5834 (2008).
and T cells. J. Virol. 78, 13293–13305 (2004). 56. Jones, J. O. & Arvin, A. M. Inhibition of the NF‑κB 74. Che, X. et al. ORF11 protein interacts with the ORF9
35. Koropchak, C. M. et al. Investigation of varicella-zoster pathway by varicella-zoster virus in vitro and in human essential tegument protein in varicella-zoster virus
virus infection of lymphocytes by in situ hybridization. epidermal cells in vivo. J. Virol. 80, 5113–5124 (2006). infection. J. Virol. 87, 5106–5117 (2013).
J. Virol. 63, 2392–2395 (1989). 57. Arvin, A. M. et al. The early immune response in 75. Liu, X., Li, Q., Dowdell, K., Fischer, E. R. & Cohen, J. I.
36. Ozaki, T. et al. Viremic phase in healthy and immunocompromised subjects with Varicella-zoster virus ORF12 protein triggers
nonimmunocompromised children with varicella. primary varicella-zoster virus infection. J. Infect. Dis. phosphorylation of ERK1/2 and inhibits apoptosis.
J. Pediatr. 104, 85–87 (1984). 154, 422–429 (1986). J. Virol. 86, 3143–3151 (2012).
37. Arvin, A. M. et al. Varicella-zoster virus T cell tropism 58. Zerboni, L. et al. Analysis of varicella zoster virus 76. Liu, X. & Cohen, J. I. Varicella-zoster virus ORF12
and the pathogenesis of skin infection. Curr. Top. attenuation by evaluation of chimeric parent Oka/ protein activates the phosphatidylinositol 3‑kinase/Akt
Microbiol. Immunol. 342, 189–209 (2010). vaccine Oka recombinant viruses in skin xenografts in pathway to regulate cell cycle progression. J. Virol. 87,
38. Berarducci, B. et al. Deletion of the first cysteine-rich the SCIDhu mouse model. Virology. 332, 337–346 1842–1848 (2013).
region of the varicella-zoster virus glycoprotein E (2005). 77. Zhang, Z. et al. Genome-wide mutagenesis reveals
ectodomain abolishes the gE and gI interaction and 59. Schmid, D. S. Varicella-zoster virus vaccine: molecular that ORF7 is a novel VZV skin-tropic factor. PLoS
differentially affects cell–cell spread and viral entry. genetics. Curr. Top. Microbiol. Immunol. 342, 323–340 Pathog. 6, e1000971 (2010).
J. Virol. 83, 228–240 (2009). (2010). This study uses comprehensive screening to
39. Berarducci, B. et al. Functions of the unique 60. Gershon, M. D., Gershon A. A. VZV infection of establish which VZV genes are required and which
N‑terminal region of glycoprotein E in the keratinocytes: production of cell-free infectious virions genes are dispensable for replication.
pathogenesis of varicella-zoster virus infection. Proc. in vivo. Curr. Top. Microbiol. Immunol. 342, 173–188 78. Chaudhuri, V., Sommer, M., Rajamani, J., Zerboni, L.
Natl Acad. Sci. USA 107, 282–287 (2010). (2010). & Arvin, A. M. Functions of varicella-zoster virus
This study identifies a large non-conserved region This paper reviews the process of VZV replication ORF23 capsid protein in viral replication and the
of gE and maps its functions in skin infection. and release of cell-free virus into cutaneous skin pathogenesis of skin infection. J. Virol. 82,
40. Mallory, S., Sommer, M. & Arvin, A. M. Mutational lesions that are needed for transmission to other 10231–10246 (2008).
analysis of varicella-zoster virus (VZV) glycoproteins, gI susceptible individuals. 79. Pevenstein, S. R. et al. Quantitation of latent varicella-
and gE, in recombinant VZV strains. J. Virol. 71, 61. Sen, N. et al. Varicella-zoster virus immediate-early zoster virus and herpes simplex virus genomes in
8279–8288 (1997). protein 62 blocks interferon regulatory factor 3 (IRF3) human trigeminal ganglia. J. Virol. 73, 10514–10518
41. Li, Q. et al. Insulin degrading enzyme induces a phosphorylation at key serine residues: a novel (1999).
conformational change in varicella-zoster virus gE, and mechanism of IRF3 inhibition among herpesviruses. 80. Wang, K. et al. Laser-capture microdissection: refining
enhances virus infectivity and stability. PLoS ONE. 5, J. Virol. 84, 9240–9253 (2010). estimates of the quantity and distribution of latent
e11327 (2010). 62. Vandevenne, P. et al. The varicella-zoster virus ORF47 herpes simplex virus 1 and varicella-zoster virus DNA
42. Moffat, J. F. et al. Glycoprotein I of varicella-zoster kinase interferes with host innate immune response by in human trigeminal ganglia at the single cell level.
virus is required for viral replication in skin and T cells. inhibiting the activation of IRF3. PLoS ONE 6, e16870 J. Virol. 79, 14079–14087 (2005).
J. Virol. 76, 8468–8471 (2002). (2011). 81. Azarkh, Y., Gilden, D. & Cohrs, R. J. Molecular
43. Ito, H. et al. Promoter sequences of varicella-zoster 63. Ambagala, A. P. & Cohen, J. I. Varicella-Zoster virus characterization of varicella zoster virus in latently
virus glycoprotein I targeted by cellular transactivating IE63, a major viral latency protein, is required to infected human ganglia: physical state and abundance
factors Sp1 and USF determine virulence in skin and inhibit the alpha interferon-induced antiviral response. of VZV DNA, quantitation of viral transcripts and
T cells in SCIDhu mice in vivo. J. Virol. 77, 489–498 J. Virol. 81, 7844–7851 (2007). detection of VZV-specific proteins. Curr. Top. Microbiol.
(2003). 64. Carpenter, J. E., Jackson, W., Benetti, L. & Grose, C. Immunol. 342, 229–241 (2010).
This paper reports that cell transcription factors Autophagosome formation during varicella-zoster 82. Nagel, M. A. et al. Varicella-zoster virus transcriptome
have an essential role as determinants of VZV virus infection following endoplasmic reticulum stress in latently infected human ganglia. J. Virol. 85,
pathogenesis. and the unfolded protein response. J. Virol. 85, 2276–2287 (2011).
44. Che, X., Zerboni, L., Sommer, M. H. & Arvin, A. M. 9414–9424 (2011). 83. Mahalingam, R. et al. Expression of protein encoded
Varicella-zoster virus open reading frame 10 is a This paper shows the role of autophagy in the by varicella-zoster virus open reading frame 63 in
virulence determinant in skin cells but not in T cells intrinsic response of skin cells to VZV infection. latently infected human ganglionic neurons. Proc. Natl
in vivo. J. Virol. 80, 3238–3248 (2006). 65. Reichelt, M. et al. Entrapment of viral capsids in nuclear Acad. Sci. USA 93, 2122–2124 (1996).
45. Baiker, A. et al. The immediate-early 63 protein of PML cages is an intrinsic antiviral host defense against 84. Lungu, O. et al. Aberrant intracellular localization of
varicella-zoster virus: analysis of functional domains varicella-zoster virus. PLoS Pathog. 7, e1001266 (2011). varicella-zoster virus regulatory proteins during
required for replication in vitro and T cell and skin This paper identifies a previously unrecognized latency. Proc. Natl Acad. Sci. USA 95, 7080–7085
tropism in the SCIDhu model in vivo. J. Virol.78, function of PML in sequestering virions in the (1998).
1181–1194 (2004). nuclei of neurons, satellite cells and epidermal cells 85. Zerboni, L. et al. Apparent expression of varicella-
46. Kinchington, P. R. et al. Virion association of IE62, the and the antiviral effects of PML isoform IV that are zoster virus proteins in latency resulting from
varicella-zoster virus (VZV) major transcriptional mediated by this process. reactivity of murine and rabbit antibodies with human
regulatory protein, requires expression of the VZV 66. Reichelt, M. et al. 3D reconstruction of herpesvirus blood group A determinants in sensory neurons.
open reading frame 66 protein kinase. J. Virol. 75, infected cell nuclei and PML nuclear cages by serial J. Virol. 86, 578–583 (2012).
9106–9113 (2001). section array scanning electron microscopy and electron 86. Gowrishankar, K. et al. Characterization of the host
47. Eisfeld, A. J., Yee, M. B., Erazo, A., Abendroth, A. & tomography. PLoS Pathog. 8, e1002740 (2012). immune response in human ganglia after herpes
Kinchington, P. R. Downregulation of class I major 67. Wang, L. et al. Disruption of PML nuclear bodies is zoster. J. Virol. 84, 8861–8870 (2010).
histocompatibility complex surface expression by mediated by ORF61 sumo-interacting motifs and This paper shows the prominent characteristics of
varicella-zoster virus involves open reading frame 66 required for varicella-zoster virus pathogenesis in skin. the host immune cell response to VZV reactivation
protein kinase-dependent and -independent PloS Pathog. 7, e1002157 (2011). in ganglia.
mechanisms. J. Virol. 81, 9034–9049 (2007). This study shows that VZV must disrupt the 87. Steain, M., Gowrishankar, K., Rodriguez, M.,
48. Abendroth, A., Kinchington, P. R. & Slobedman, B. architecture of PML-NBs in order to replicate in Slobedman, B. & Abendroth, A. Upregulation of
Varicella zoster virus immune evasion strategies. skin and documents the importance of the CXCL10 in human dorsal root ganglia during
Curr. Top. Microbiol. Immunol. 342, 155–171 SUMO-interacting motifs of the ORF61 protein in experimental and natural varicella-zoster virus
(2010). disarming this antiviral mechanism. infection. J. Virol. 85, 626–631 (2011).

NATURE REVIEWS | MICROBIOLOGY VOLUME 12 | MARCH 2014 | 209

© 2014 Macmillan Publishers Limited. All rights reserved


REVIEWS

88. Steain, M., Slobedman, B. & Abendroth, A. 97. Zerboni, L. et al. Aberrant infection and persistence of laboratory features 10 years after the introduction of the
Experimental models to study varicella-zoster virus varicella-zoster virus in human dorsal root ganglia in varicella vaccine. J. Infect. Dis. 203, 316–323 (2011).
infection of neurons. Curr. Top. Microbiol. Immunol. vivo in the absence of glycoprotein I. Proc. Natl Acad. 106. Tyring, S. et al. Famciclovir for the treatment of acute
342, 211–228 (2010). Sci. USA 104, 14086–14091 (2007). herpes zoster: effects on acute disease and
89. Hood, C. et al. Varicella-zoster virus ORF63 inhibits This study documents the unexpected finding that postherpetic neuralgia. A randomized, double-blind,
apoptosis of primary human neurons. J. Virol. 80, the mutation of VZV proteins can result in chronic placebo-controlled trial. Collaborative Famciclovir
1025–1031 (2006). infectious processes in DRG. Herpes Zoster Study Group. Ann. Intern. Med. 123,
90. Janer, A. et al. PML clastosomes prevent nuclear 98. Wang, Z. H. et al. Essential role played by the 89–96 (1995).
accumulation of mutant ataxin‑7 and other C‑terminal domain of glycoprotein I in envelopment of 107. Cohen, J. I. & Seidel, K. E. Generation of varicella-
polyglutamine proteins. J. Cell Biol. 174, 65–76 varicella-zoster virus in the trans-Golgi network: zoster virus (VZV) and viral mutants from cosmid
(2006). interactions of glycoproteins with tegument. J. Virol. DNAs: VZV thymidylate synthetase is not essential for
91. Weinberg, A. et al. Varicella-zoster virus-specific 75, 323–340 (2001). replication in vitro. Proc. Natl Acad. Sci. USA 90,
immune responses to herpes zoster in elderly 99. Mahalingam, R., Messaoudi, I. & Gilden, D. Simian 7376–7380 (1993).
participants in a trial of a clinically effective zoster varicella virus pathogenesis. Curr. Top. Microbiol. This paper reports the first use of VZV cosmids for
vaccine. J. Infect. Dis. 200, 1068–1077 (2009). Immunol. 342, 309–321 (2010). generating mutant viruses.
92. Reichelt, M., Zerboni, L. & Arvin, A. M. Mechanisms of 100. Ouwendijk, W. J. et al. T‑cell infiltration correlates with 108. Tischer, B. K. et al. A self-excisable infectious bacterial
varicella-zoster virus neuropathogenesis in human CXCL10 expression in ganglia of cynomolgus artificial chromosome clone of varicella-zoster virus
dorsal root ganglia. J. Virol. 82, 3971–3983 macaques with reactivated simian varicella virus. allows analysis of the essential tegument protein
(2008). J. Virol. 87, 2979–2982 (2013). encoded by ORF9. J. Virol. 81, 13200–13208 (2007).
This paper reports that the capacity of VZV to 101. Kinchington, P. R. & Goins, W. F. Varicella zoster virus- This paper reports the first use of a VZV bacterial
induce cell fusion includes the fusion of satellite induced pain and post-herpetic neuralgia in the human artificial chromosome (BAC) for generating mutant
cells and neurons in infected DRG in addition to host and in rodent animal models. J. Neurovirol. 17, viruses.
polykaryocte formation in skin. 590–599 (2011). 109. Rowe, J., Greenblatt, R. J., Liu, D. & Moffat, J. F.
93. Zerboni, L. et al. Herpes simplex virus 1 tropism for This paper shows that VZV is transferred from local Compounds that target host cell proteins prevent
human sensory ganglion neurons in the severe sites of inoculation into sensory ganglia and causes varicella-zoster virus replication in culture, ex vivo, and
combined immunodeficiency mouse model of hypersensitivity to peripheral nerve stimulation. in SCID‑Hu mice. Antiviral Res. 86, 276–285 (2010).
neuropathogenesis. J. Virol. 87, 2791–2802 (2013). 102. Chen, J. J., Gershon, A. A., Li, Z., Cowles, R. A. &
94. Esiri, M. M. & Tomlinson, A. H. Herpes zoster: Gershon, M. D. Varicella zoster virus (VZV) infects and Acknowledgements
demonstration of virus in trigeminal nerve and establishes latency in enteric neurons. The authors thank M. Sommer, X. Che, L. Wang and
ganglion by immunofluorescence and electron J. Neurovirol.17, 578–589 (2011). M. Reichelt, past postdoctoral fellows and students and col-
microscopy. J. Neurol. Sci. 15, 35–48 (1972). 103. Selariu, A. et al. ORF7 of varicella-zoster virus is a laborators and many dedicated colleagues in the field of VZV
95. Zerboni, L. et al. Varicella-zoster virus glycoprotein E is neurotropic factor. J. Virol. 86, 8614–8624 (2012). research for their invaluable contributions. J. Moffat initiated
a critical determinant of virulence in the SCID mouse- 104. Oxman, M. N. et al. A vaccine to prevent herpes the development of the SCID mouse model as a postdoctoral
human model of neuropathogenesis. J. Virol. 85, zoster and postherpetic neuralgia in older adults. fellow in the Arvin laboratory. This work is supported by US
98–111 (2011). N. Engl. J. Med. 352, National Institutes of Health (NIH) grants, AI20459,
96. Zerboni, L., Reichelt, M. & Arvin, A. Varicella-zoster 2271–2284 (2005). AI05346 and AI102546.
virus neurotropism in SCID mouse-human dorsal root 105. Pahud, B. A., Glaser, C. A., Dekker, C. L., Arvin, A. M.
ganglia xenografts. Curr. Top. Microbiol. Immunol. & Schmid, D. S. Varicella zoster disease of the central Competing interests statement
342, 255–276 (2010). nervous system: epidemiological, clinical, and The authors declare no competing interests.

210 | MARCH 2014 | VOLUME 12 www.nature.com/reviews/micro

© 2014 Macmillan Publishers Limited. All rights reserved

Das könnte Ihnen auch gefallen