Vous êtes sur la page 1sur 20

Rev. Med. Virol. 2015; 25: 300–319.

Published online 23 July 2015 in Wiley Online Library


(wileyonlinelibrary.com)
Reviews in Medical Virology DOI: 10.1002/rmv.1848

REVIEW
Innate immune response during herpes simplex
virus encephalitis and development of
immunomodulatory strategies
Jocelyne Piret and Guy Boivin*
Research Center in Infectious Diseases, CHU de Québec and Laval University, Quebec City, Quebec, Canada

S U M M A RY
Herpes simplex viruses are large double-stranded DNA viruses. These viruses have the ability to establish a lifelong
latency in sensory ganglia and to invade and replicate in the CNS. Apart from relatively benign mucosal infections,
HSV is responsible for severe illnesses including HSV encephalitis (HSE). HSE is the most common cause of sporadic,
potentially fatal viral encephalitis in Western countries. If left untreated, the mortality rate associated with HSE is ap-
proximately 70%. Despite antiviral therapy, the mortality is still higher than 30%, and almost 60% of surviving individ-
uals develop neurological sequelae. It is suggested that direct virus-related and indirect immune-mediated mechanisms
contribute to the damages occurring in the CNS during HSE. In this manuscript, we describe the innate immune re-
sponse to HSV, the development of HSE in mice knock-out for proteins of the innate immune system as well as
inherited deficiencies in key components of the signaling pathways involved in the production of type I interferon that
could predispose individuals to develop HSE. Finally, we review several immunomodulatory strategies aimed at mod-
ulating the innate immune response at a critical time after infection that were evaluated in mouse models and could be
combined with antiviral therapy to improve the prognosis of HSE. In conclusion, the cerebral innate immune response
that develops during HSE is a “double-edged sword” as it is critical to control viral replication in the brain early after
infection, but, if left uncontrolled, may also result in an exaggerated inflammatory response that could be detrimental to
the host. Copyright © 2015 John Wiley & Sons, Ltd.

INTRODUCTION capacity to invade and replicate in the CNS [1]. A


Herpes simplex viruses are large double-stranded US survey from 2005 to 2010 estimated that 53.9%
DNA viruses. Among the key biological properties of adults aged 14–49 years are infected with HSV-
of HSV, there are the establishment of a lifelong 1 and 15.7% with HSV-2 [2]. Transmission of these
latency in dorsal root sensory ganglia and the viruses is dependent on intimate contacts through
mucosal surfaces. The virus is then transported by
*Corresponding author: G. Boivin, FRCP, Centre de recherche en retrograde route via sensory nerves to dorsal root
infectiologie, CHU de Québec, Pavillon CHUL, 2705, Blvd Laurier, ganglia where it establishes latency followed by pe-
RC-709, Québec (Québec), Canada, G1V 4G2. riodic reactivations and anterograde transport back
E-mail: Guy.Boivin@crchudequebec.ulaval.ca
to the mucous surfaces [1]. Although HSV-1 is
Abbreviations used
AIM2, absent in melanoma 2; AP-1, activator protein 1; BBB, blood– protein 88; NAP1, NAK-associated protein 1; NEMO, NF-κB essen-
brain barrier; CARD, caspase recruitment domain; Casp., caspase; tial modulator; NF-κB, nuclear factor kappa B; NOD, nucleotide-bind-
CCL, C–C motif chemokine ligand; cGAS, cytosolic guanosine- ing oligomerization domain; NLRs, NOD-like receptors; ODNs,
monophosphate adenosine-monophosphate (cGAMP) synthase; CXCL, oligodeoxynucleotides; PI3K, phosphoinositide-3-kinase; PAMPs,
C–X–C motif chemokine ligand; dsDNA, double-stranded DNA; pathogen-associated molecular patterns; PRRs, pathogen recognition
dsRNA, double-stranded RNA; FADD, Fas-associated death domain; receptors; poly I:C, polyinosinic:polycytidylic acid; RIG-I, retinoic
GFP, green fluorescent protein; HSE, herpes simplex virus encephali- acid-inducible gene I; RIP, receptor-interacting protein; RLH, RIG-I-
tis; IFI16, IFN-inducible protein 16; IFNAR, IFN-α/β receptor; IκB, like RNA helicase; ssRNA, single-stranded RNA; STAT, signal trans-
inhibitor of NF-κB; IKK, IκB kinase; IKKα/β, inhibitor of IκB kinase ducer and activator of transcription; STING, stimulator of IFN genes;
α/β; IKKi, inducible IκB kinase; iNOS, inducible nitric oxide synthase; TAB, TAK1-binding protein; TAK1, transforming growth factor-β-
IPS-1, IFN-β promoter stimulator 1; iPSCs, induced pluripotent stem activated kinase 1; TANK, TRAF family member-associated NF-κB ac-
cells; IRF, IFN regulatory factor; IRAK, IL-1 receptor-associated ki- tivator; TBK1, TANK-binding kinase 1; TIR, Toll-IL-1 receptor;
nase; ISGs, IFN-stimulated genes; ISRE, IFN-stimulated response ele- TIRAP, TIR-associated protein; TLR, Toll-like receptor; TRAF, TNF
ment; JAK1, Janus kinase 1; LRR, leucine-rich repeat protein; MAPK, receptor-associated factor; TNFR1, TNF-α receptor 1; TRAM, TRIF-
mitogen-activated protein kinase; Mda5, melanoma differentiation- related adaptor molecule; TRIF, TIR domain-containing adaptor induc-
associated gene 5; MyD88, myeloid differentiation primary response ing IFN-β; TYK2, tyrosine kinase 2.

Copyright © 2015 John Wiley & Sons, Ltd.


Innate immune response to herpes simplex virus 301

mainly associated with herpes labialis and HSV-2 levels and brain injury [13] and the correlation
with genital herpes, there is a considerable overlap between infiltration of immune cells in the CNS
in viral tropism [2]. Apart from these relatively be- and demyelination. It has also been reported that
nign conditions, HSV is also associated with severe depletion of macrophages and neutrophils in sus-
illnesses including ocular infections (keratitis), neo- ceptible 129 mice enhanced survival whereas resis-
natal infections (affecting the skin or disseminated) tant immunodeficient B6-E mice transplanted with
and CNS involvement (HSV encephalitis; HSE) [3]. bone marrow of 129 mice had accelerated fatal out-
come [14].
HERPES SIMPLEX VIRUS ENCEPHALITIS The clinical manifestations of HSE include fever,
Herpes simplex virus encephalitis is the most com- altered consciousness, abnormal behavior and lo-
mon cause of sporadic fatal viral encephalitis in calized neurological findings such as seizure and
Western countries [4]. The incidence of HSE is esti- paralysis [3,15]. Laboratory clues for the diagnosis
mated to be approximately 1 in 250 000 individuals include a lymphocytic pleocytosis and a proteinosis
[5]. Importantly, the mortality rate is in excess of of the CSF, characteristic electroencephalographic
70% with only 2.5% of all patients returning to nor- changes and abnormal computed tomographic or
mal function in the absence of antiviral treatment magnetic resonance imaging findings localized to
[6]. The incidence of HSE peaks in childhood the temporal lobe(s) [15,16]. PCR detection of
between the age of 6 months to 3 years during HSV DNA in the CSF has become the diagnostic
HSV-1 primary infection and thereafter in adults method of choice for HSE and has replaced the
> 50 years [7]. Virtually all adult cases of HSE are more invasive brain biopsy [17,18]. Acyclovir, an
caused by HSV-1 with about 1/3 of them being analogue of guanosine, is the antiviral of choice
the consequence of primary infection and the re- for the treatment of HSE. At a dosage of
maining 2/3 occurring in the presence of pre- 10 mg/kg intravenously every 8 h for 14–21 days,
existing antibodies [4]. acyclovir decreases the mortality attributable to
The pathogenesis of HSE has remained elusive. HSE to 30% (instead of 70%) although only 40%
Herpes simplex virus encephalitis results in acute of patients return to normal activities or have mi-
inflammation, congestion and/or hemorrhage, nor impairments [4,15]. A recent international
most prominently in the temporal lobes and usu- study suggested that rapid diagnosis and early ad-
ally asymmetrically [4]. At the microscopic level, ministration of antiviral therapy at the onset of
there is diffuse perivascular subarachnoid mono- HSE are keys to a favorable outcome [19].
nuclear cell infiltrate, gliosis and satellitosis [8]. Al- The topic of this review deals with the cerebral
though both primary and recurrent HSV infections innate immune response in experimental mouse
can lead to HSE, the route of access of the virus to models of HSE and the inherited deficiencies in
the CNS is still controversial. Various animal the immune signaling pathways that may predis-
models have been developed in an attempt to pose individuals to develop natural HSE. Finally,
mimic the possible routes of infection. Some have we discuss the rational to combine antiviral agents
suggested the olfactory and trigeminal nerves as with immune-based therapy to improve the prog-
potential pathways to the CNS with possible estab- nosis of HSE.
lishment of latency within the brain [9,10]. Intrana-
sal inoculation of HSV-1 to mice produces focal INNATE IMMUNE RESPONSE TO HERPES
lesions localized to the temporal lobe(s), similar to SIMPLEX VIRUS
what is observed in humans [11]. It is though that Innate immune response constitutes the first line of
both direct virus-mediated and indirect immune- host defense that limits viral spread and plays an
mediated mechanisms play a role in CNS damages, important role in the activation of the adaptive
which may explain why HSE is not more common response. Pathogen recognition receptors (PRRs)
among immunocompromised hosts despite recur- recognize conserved microbial molecules known as
rent mucocutaneous infections. Other lines of evi- pathogen-associated molecular patterns (PAMPs),
dence that support immune-mediated damages in to initiate innate immune response and inflammation
animal models of HSE include the predominance (Figure 1) [20–22]. PRRs include Toll-like recep-
of specific cytotoxic T cells at sites of infection tors (TLRs), retinoic acid-inducible gene I (RIG-I)-
[12], the correlation between cytokine/nitric oxide like RNA helicases (RLHs), nucleotide-binding

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
302 J. Piret and G. Boivin

Figure 1. Schematic representation of HSV sensing by pathogen recognition receptors. HSV PAMPS are recognized by host PRRs which
initiate several signaling pathways that activate IRFs, NF-κB and AP-1 transcription factors. This leads to the production of type I IFN and
pro-inflammatory cytokines. Viral glycoproteins are sensed by TLR2 located at the cell surface. The abundant CpG motifs present in the
viral DNA are recognized by endosomal TLR9. In the cytosol, viral dsDNA also induces the synthesis of cGAMP by cGAS which activates
the STING/TBK1 pathway. Intermediate dsRNA produced during the viral replication are recognized by endosomal TLR3 and cytosolic
RIG-I and Mda5 sensors. During the second wave of type I IFN production, IFN-α/β bind to the IFNAR receptor, the signal is transduced
to JAK1 and TYK2 that respectively phosphorylate STAT1 and STAT2. The complex of phosphorylated STAT1 and STAT2
heterodimerizes with IRF9 and translocates to the nucleus to encode antiviral effectors that establish an antiviral state in the infected cells
as well as in the neighboring non-infected cells

oligomerization domain (NOD)-like receptors (NLRs) pathway(s) can be activated to induce appropriate
as well as cytosolic DNA sensors. immune response to pathogens. Recognition of
TLRs are mainly expressed at the cell surface except HSV is mediated by TLRs 2, 3 and 9 (Figure 1)
TLRs 3, 7, 8 and 9 that are localized in the endosomes. [24–26]. HSV particles are first sensed by TLR2
The chaperone protein Unc93B is essential for the localized at the cell surface through the detection
translocation of TLRs 3, 7 and 9 from the endoplasmic of viral glycoproteins [27,28]. After entry, viral
reticulum to the endosomes [23]. The PAMPs recogni- genomic DNA, which contains abundant unme-
tion by TLRs induces the recruitment of adaptor pro- thylated CpG motifs, is detected by endosomal
teins containing a Toll/IL-1 receptor (TIR) domain TLR9 [29–31]. In recent years, the implication of
including myeloid differentiation primary response cyclic guanosine monophosphate-adenosine mo-
protein 88 (MyD88), TIR-associated protein (TIRAP), nophosphate (cGAMP) synthase (cGAS) in the
TIR-domain-containing adaptor inducing IFN-β recognition of cytosolic HSV dsDNA has been
(TRIF) and TRIF-related adaptor molecule (TRAM) described [32]. cGAS is activated by dsDNA and
[20,22]. According to the adaptor and to the TLR, catalyzes the synthesis of a noncanonical cyclic
the MyD88-dependent (TLRs 1, 2, 4, 5, 6, 7, 8 and dinucleotide from ATP and GTP. cGAMP is an
9) and/or the TRIF-dependent (TLRs 3 and 4) endogenous second messenger that binds to and

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 303

activates the adaptor protein stimulator of IFN element (ISRE) promoter regions and stimulates
genes (STING) [33,34]. It is suggested that other type I IFN-depending gene transcription, including
cytosolic dsDNA sensors, such as absent in mela- IRF7 [49,50]. This cascade also results in the induc-
noma 2 (AIM2), interferon-inducible protein 16 tion of the expression of IFN-stimulated genes
(IFI16) and RNA polymerase III, may also contrib- (ISGs) encoding antiviral effectors important to
ute to the innate immune response to HSV [35]. limit viral spread and to establish an antiviral state
Replication of the viral genome leads to accumula- in the infected cells as well as in the neighboring
tion of intermediate dsRNAs which are then sensed non-infected cells [51].
by endosomal TLR3 [36]. RLHs consist of a group
of RNA helicases, namely RIG-I and melanoma CEREBRAL INNATE IMMUNE RESPONSE TO
differentiation-associated gene 5 (Mda5), that are HERPES SIMPLEX VIRUS ENCEPHALITIS
involved in the recognition of viral RNA in the cy- Microglial cells are the first line of immune defense
toplasm [20]. RIG-I recognizes short dsRNA struc- in the brain parenchyma [52]. They can be activated
tures and 5′-triphosphate ssRNA whereas Mda5 during systemic infection without loss of integrity
senses long dsRNA that has higher order structures of the blood–brain barrier (BBB). The circum-
[37]. As TLRs, RLHs interact with adaptor proteins ventricular organs and the choroid plexus are stra-
to recruit downstream effectors [20]. RIG-I and tegically positioned to detect PAMPs in the blood.
Mda5 recognize replication intermediate dsRNA This leads to the progressive activation of resident
structures of HSV and interact with IFN-β promoter microglial cells and to the production of TNF-α,
stimulator-1 (IPS-1) adaptor [38–40]. Viral com- which acts in autocrine and paracrine manners to
ponents sensed by these specific PRRs trigger activate the immune cells across the brain paren-
activation of the transcription of nuclear factor-κB chyma [53]. Similar to macrophages, parenchymal
(NF-κB), activator protein 1 (AP-1) and IFN microglial cells express TLRs, respond to PAMPs
regulatory factor (IRF) family members. These and produce pro-inflammatory mediators. In nor-
transcription factors regulate the expression of mal conditions, the trafficking of leukocytes from
chemokines, pro-inflammatory cytokines and type the blood to the CNS is exquisitely controlled by
I IFNs [24–26] which consist of a single IFN-β and the BBB [54]. The activation of innate immune re-
more than a dozen IFN-α subtypes [41]. Proteins sponse in resident microglial cells following an
of the IRF family, in particular IRF3 and IRF7, infection induces a marked recruitment, prolifera-
tightly control the transcription of IFN-α/β genes tion and activation of microglial cell precursors
[42] and are strong inducers of type I IFNs [22]. from the blood to affected regions of the brain
During initial induction of type I IFNs, IRF3 and [55]. Stimulated astrocytes can also induce the
NF-κB, which are constitutively expressed, are expression of pro-inflammatory cytokines and
respectively phosphorylated by TANK-binding chemokines that can initiate leukocytes migration
kinase 1 (TBK1) and IκB kinase (IKK) complex across the BBB [56]. By using chimeric mice with
[43,44]. Once activated, they translocate to the nu- bone marrow-derived cells expressing the green
cleus, where they bind to the IFN promoter to form fluorescent protein (GFP), our group demonstrated
the IFN enhanceosome [45]. In concert with IRF3, a recruitment of GFP positive cells into the CNS af-
IRF7 amplifies and facilitates the expression of the ter intranasal infection with HSV-1 [57]. These infil-
type I IFN cascade [46]. Phosphorylated IRF7 can trating cells originating from the periphery could
form a homodimer or a heterodimer with IRF3 to contribute to the cerebral innate immune response
induce differential effects on the expression of type to the virus.
I IFN gene family members [47]. The expression of TLR2, TLR3 and TLR9 was
The initial type I IFNs produced act in both auto- strongly up-regulated and widely distributed
crine and paracrine manners through the IFN-α/β throughout the brain of mice infected by the intra-
receptor (IFNAR) composed of the IFNAR1 and nasal route with HSV-1 whereas the other TLRs
IFNAR2 subunits, which activates components of were barely detected in some regions or almost un-
the Janus kinase-signal transducer and activator of detectable [58]. More specifically, TLR2 and TLR9
transcription (JAK/STAT) signaling pathway [48]. are expressed by microglia and astrocytes whereas
STAT1 interacts with STAT2 and IRF9 to form a TLR3 is present in microglia, astrocytes, oligoden-
complex that binds to the IFN-stimulated response drocytes and neurons [59]. In vitro studies have

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
304 J. Piret and G. Boivin

demonstrated that cytosolic RIG-I and Mda5 sen- brain [66]. TLR2 and TLR9 double knock-out mice
sors are found in microglia, astrocytes and neurons were also more susceptible to viral dissemination
[60,61]. The adaptor protein STING has been de- to the CNS than either TLR2- or TLR9-deficient
tected in microglial cells and astrocytes [62]. Thus, mice after intraperitoneal or vaginal inoculation of
both hematopoietic and non-hematopoietic resi- HSV-2 [67]. It was thus proposed that both TLR2
dent cells of the CNS have been reported to partic- and TLR9 recognize HSV components and act in
ipate to the innate immune response to HSV. an additive manner on the activation of the adaptor
protein MyD88 to initiate an immune response. In
HERPES SIMPLEX VIRUS ENCEPHALITIS IN this respect, intranasal infection with HSV-1 to mice
KNOCK-OUT MOUSE MODELS deficient for MyD88 resulted in an increased
Different mouse strains (in a C57BL/6 or mortality rate because of an uncontrolled viral
C57BL/6x129 genetic background) knock-out for replication in the brain compared to wild-type [63].
PRRs, adaptor proteins, mediators of inflammation The viral dissemination to the CNS was also increased
and their receptors have been used to evaluate the in mice lacking MyD88 inoculated with HSV-2 by
specific role of several components of the innate the intraperitoneal and vaginal routes [67,68].
immune response to HSV in the CNS (Table 1). Sensing of HSV dsRNA is mediated by
Intranasal infection of mice deficient for TLR2 endosomal TLR3 and cytoplasmic RIG-I and
with HSV-1 had no effect on survival rate and viral Mda5 which signal through TRIF and IPS-1 adap-
titers in the brain compared to wild-type controls tor proteins, respectively [36,38–40]. The relative in-
[63,64]. Other groups reported that intracranial or fluence of these sensors following HSV-1 infection
intraperitoneal inoculation of HSV-1 to mice lack- has been investigated by our group in TRIF- and
ing TLR2 increased the survival rates but did not IPS-1-deficient mice [69]. The brain viral load and
affect the viral titers in the brain [65,66]. The latter the mortality rate were markedly increased in
mice responded to HSV-1 infection with an im- TRIF- and IPS-1-deficient mice compared to wild-
paired immune response [65]. Similarly, intraperi- type controls following intranasal infection with
toneal inoculation of HSV-2 to TLR2-deficient HSV-1. Moreover, the production of IFN-β was sig-
mice did not affect the viral titers in the brain com- nificantly reduced at a critical time during the in-
pared to wild-type [67]. These data suggest that the fection and could explain the higher susceptibility
inflammatory response mediated by TLR2 in the of both deficient mice. This study suggests that
CNS could be detrimental to the host. Mice lacking both endosomal TLR3 and cytoplasmic helicase
TLR9 infected intranasally with HSV-1 had in- sensors are implicated in the signaling pathways
creased mortality rates because of uncontrolled vi- leading to IFN-β production during HSE with the
ral replication and increased levels of chemokines TLR3-TRIF pathway being predominant over
in the brain but the mortality rate was lower than RIG-I/Mda5-IPS-1 pathways. TLR3-deficient mice
in mice deficient for both TLR2 and TLR9 [64]. Mice were highly susceptible to HSV-2 dissemination to
deficient for TLR9 infected with HSV-1 by the intra- the CNS after vaginal infection [70]. The viral tro-
cranial route or with HSV-2 by the intraperitoneal pism was extended to astrocytes that became per-
route had no increased susceptibility to infection missive to the infection. It was thus suggested
[66,67]. These data suggest that TLR9 does not ex- that TLR3 expressed by astrocytes senses HSV-2 af-
ert a strong effect in the innate immune defense ter entry into the CNS and limits the spread of the
against HSV infection. The potential cooperative virus to neurons.
action of TLR2 and TLR9 in the recognition of The deficiency of Unc93B, involved in endo-
HSV in the CNS has been investigated by several somal TLR3, TLR7 and TLR9 sorting, decreased
groups. Mice lacking both TLR2 and TLR9 infected the survival rate but did not affect viral replication
intranasally with HSV-1 had increased mortality or type I IFN levels in the brain of mice infected
rates because of uncontrolled viral replication, with HSV-1 by the intracranial route [66].
higher levels of chemokines and lower production More recently, it was also reported that mice
of IL-1β and IFN-γ in the brain compared to wild- lacking cGAS or STING, involved in the recogni-
type and TLR2- or TLR9-deficient mice [64]. In con- tion of cytosolic HSV dsDNA, infected intrave-
trast, inoculation of HSV-1 by the intracranial route nously with HSV-1 exhibited enhanced HSV titers
had no effect on survival rate and viral titers in the in the brain, decreased levels of IFN-α/β in serum

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 305

Table 1. Outcome of herpes simplex virus encephalitis in knock-out mouse models


Route of
inoculation Observations in
Viral strain Mouse genetic deficient mice
Deficiency Inoculum background compared to wild-type References

Pathogen recognition receptors


TLR2 HSV-1 EK Intranasal No effect on survival rate [63,64]
104/106 PFU C57BL/6 No effect on viral titers in the brain
HSV-1 7134R Intracranial Increased survival rate [66]
3 × 104 PFU C57BL/6 No effect on viral titers in
the brain
No effect on IFN-β levels in
the brain
HSV-1 KOS Intraperitoneal Increased survival rate [65]
109 PFU C57BL/6x129 No effect on viral titers in
the brain
Decrease in CCL2 levels in
the brain
Less brain inflammation by
histopathology
HSV-2 MS Intraperitoneal No effect on viral titers in the brain [67]
106 PFU C57BL/6
TLR3 HSV-2 333 Vaginal More susceptible to clinical [70]
disease in the CNS
6.7 × 104 PFU C57BL/6 Increase in viral titers in the
cerebellum
No effect on type I IFN and
TNF-α levels in the brain
Broader tropism to astrocytes
TLR9 HSV-1 EK Intranasal Increased mortality rates [64]
106 PFU C57BL/6 Increase in viral titers in the brain
Increase in CCL2 and CXCL10
levels in the brain
HSV-1 7134R Intracranial No effect on survival rate [66]
3 × 104 PFU C57BL/6 No effect on viral titers in the brain
No effect on IFN-β levels in
the brain
HSV-2 MS Intraperitoneal No effect on viral titers in the brain [67]
106 PFU C57BL/6
TLR2/TLR9 HSV-1 EK Intranasal Increased mortality rate [64]
106 PFU C57BL/6 Increase in viral titers in
the brain
Decrease in IL-1β and IFN-γ
levels in the brain
Increase in CCL2 and CXCL10
levels in the brain

(Continues)

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
306 J. Piret and G. Boivin

Table 1.
0. (Continued)
Route of
inoculation Observations in
Viral strain Mouse genetic deficient mice
Deficiency Inoculum background compared to wild-type References

HSV-1 7134R Intracranial No effect on survival rate [66]


3 × 104 PFU C57BL/6 No effect on viral titers in the brain
No effect on IFN-β levels in
the brain
HSV-2 MS Intraperitoneal Increased susceptibility to infection [67]
106 PFU C57BL/6 Increase in viral titers in the brain
HSV-2 333 Vaginal Increased mortality rate [67]
6.7 × 104 PFU C57BL/6 Increase in viral titers in
the brain stem
Decrease in TNF-α and
CXCL9 levels in the brain
cGAS HSV-1 aka KOS Intravenous Increased mortality rate [32]
106 PFU C57BL/6 Increase in viral titers in
the brain
Decrease in IFN-α/β
production in serum
TLRs 3, 7, 9 routing
Unc93B HSV-1 7134R Intracranial Increased mortality rate [66]
3 × 104 PFU C57BL/6 No effect on viral
replication in the brain
No effect on IFN-β levels
in the brain
Adaptor proteins
MyD88 HSV-1 EK Intranasal Increased mortality rate [63]
104 PFU C57BL/6 Increase in viral titers in the brain
Increased inflammation in
the brain by histopathology
HSV-2 MS Intraperitoneal Increase in viral titers in the brain [67]
106 PFU C57BL/6
HSV-2 333 Vaginal Increased mortality rate [68]
9 × 104 and 9 × 103 C57BL/6
PFU
TRIF HSV-1 H25 Intranasal Increased mortality rate [69]
7.5 × 105 PFU C57BL/6 Increase in the brain viral load
Decrease in IFN-β production
in the brain
IPS-1 HSV-1 H25 Intranasal Increased mortality rate [69]
7.5 × 105 PFU C57BL/6 Increase in the brain viral load
Decrease in IFN-β production
in the brain

(Continues)

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 307

Table 1.
0. (Continued)
Route of
inoculation Observations in
Viral strain Mouse genetic deficient mice
Deficiency Inoculum background compared to wild-type References

STING HSV-1 KOS Intravenous Increased mortality rate [71]


107 PFU Increase in viral titers in
the brain
Decrease in IFN-α/β
production in serum
Receptors of IFN-α/β and TNF-α
IFNAR HSV-1 7134R Intracranial Increased mortality rate [66]
3 × 104 PFU C57BL/6 Increase in viral titers in
the brain
Increase in IFN-β production
in the brain
HSV-2 333 Vaginal More susceptible to clinical [70]
6.7 × 104 PFU C57BL/6 disease in the CNS
No effect on viral titers
in the cerebellum
No effect on type I IFN
and TNF-α levels in the brain
Broader tropism to astrocytes
TNFR1 HSV-1 EK Intracranial Increased mortality rate [72]
102 PFU C57BL/6 Increase in viral titers in the brain
Decrease in CCL3 and CCL5
levels in the brain
Pro-inflammatory cytokines
IL-1β HSV-1 H25 Intranasal Increased mortality rate [73]
3 × 106 PFU C57BL/6 Increase in the brain viral load
Decrease in the expression of
TLR2 in the brain
Decrease in CCL2, IL-12,
IFN-β and IFN-γ levels in the brain
TNF-α HSV-1 H25 Intranasal Increased mortality rate [73]
3 × 106 PFU C57BL/6 Increase in the brain viral load
Decrease in the expression of
TLR2 in the brain
Decrease in CCL2, IL-12,
IFN-β and IFN-γ levels in the brain

and reduced survival rates compared to wild-type infection and demonstrated increased titers of HSV
controls [32,71]. in the brain [66]. IFNAR-deficient mice infected vagi-
Type I IFN is crucial for the survival of mice in- nally with HSV-2 failed to induce the expression of
fected by direct intracranial inoculation as mice lack- most PRRs (eg., TLR2, TLR3, RIG-I and Mda5) in
ing the type I IFN receptor died rapidly from the the brain which could explain the more severe disease

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
308 J. Piret and G. Boivin

and the increased mortality seen in these animals [70]. gene of two unrelated children with HSE from
The deficiency of the TNF-α receptor 1 (TNFR1) also nonconsanguineous families [75]. The mutation
resulted in an increased mortality rate because of an (P554S), located in the ectodomain of TLR3, re-
uncontrolled viral replication in the brain [72]. The sulted in a loss-of-function of the protein. The case
survival rate was decreased in mice deficient for of another child suffering from HSE and who car-
IL-1β and/or TNF-α compared to the wild-type ried a heterozygous autosomal recessive complete
following intranasal infection with HSV-1 because of mutation in the TLR3 gene was then described
an uncontrolled viral replication, a reduced expres- [76]. One allele harbored the amino acid substitu-
sion of TLR2 and a decrease in the production of tion P554S and the other allele had a stop codon
cytokines/chemokines in the CNS [73]. resulting in the premature termination and a loss-
Overall, these data suggest that the inability to of-function of the protein. The production of IFN-
rapidly generate an adequate innate immune re- β and -λ was reduced but not abolished in the fibro-
sponse to the virus could ultimately lead to a loss blasts of the first two patients after stimulation
of viral containment. with poly I:C and HSV-1 whereas it was completely
abolished in fibroblasts of the other one. All pa-
INHERITED DEFICIENCIES OF THE INNATE tients’ fibroblasts were more permissive to infection
IMMUNE SYSTEM IN HUMAN CASES OF with HSV-1. Viral replication and cell viability
HERPES SIMPLEX VIRUS ENCEPHALITIS could be restored to control levels by pre-treatment
Most genetic predispositions to HSE in humans re- of cells with recombinant IFN-α2b before infection.
sult from deficiencies in the type I IFN production In contrast, PBMCs from all patients responded
pathways (Table 2). The first report was the normally to poly I:C and HSV-1 activation. Re-
discovery of homozygous autosomal recessive cently, two other patients carrying heterozygous
complete mutations in the Unc93B1 gene in two un- mutations (G743D+R811I and L360P) in the TLR3
related consanguineous children suffering from re- gene that resulted in a loss-of-function respectively
current episodes of HSE [74]. Mutations (i.e. because of a lower expression or a lack of cleavage
1034del4/G781A) led to a premature termination of the protein as well as a single patient with a ho-
codon and a complete loss-of-expression of the pro- mozygous hypomorphic mutation (R867Q) were
tein. Dermal fibroblasts, which selectively express described [77]. Dermal fibroblasts derived from
TLR3, derived from these patients were stimulated these three patients had severely impaired response
with polyinosinic:polycytidylic acid (poly I:C; a to stimulation with poly I:C and HSV-1 as well as
synthetic form of dsRNA) and HSV-1 and demon- an increased susceptibility to HSV-1 infection. All
strated an impaired IFN-β and -λ response com- patients with deficiency in TLR3 were normally re-
pared to those of controls. Patients’ fibroblasts sistant to other infectious diseases including viral
were more permissive to HSV-1 infection, which re- infections. Four of the six children carrying TLR3
sulted in an increased viral replication and cell deficiency (i.e. one child with P554S mutation and
death compared to controls. Pre-treatment of cells patients with G743D/R811I, L360P and R867Q mu-
with exogenous IFN-α2b before viral infection re- tations) as well as the two children with autosomal
stored viral titers and cell viability to control levels. recessive Unc93B deficiency presented HSE re-
These data suggested that a production of type I lapses (Table 2) suggesting that recurrences could
IFN dependent of Unc93B and TLR3 is involved be more frequent in individuals with inborn errors
in the protective immunity against HSE. The pro- of TLR3 immunity [77].
duction of IFN-α/β and -λ in PBMCs derived from The importance of the TLR3-dependent signaling
Unc93B-deficient patients was abolished in re- pathway in the protective immunity against HSE
sponse to agonists of TLRs 7, 8 and 9. Despite this was confirmed by the identification of a series of
deficiency, these patients were normally resistant deficiencies in proteins involved in this axis. A het-
to other infectious diseases including viral infec- erozygous autosomal negative dominant mutation
tions suggesting that the responses of TLRs 7, 8 in the TRAF3 gene was discovered in one child suf-
and 9 could be redundant for the production of fering from HSE and no other unusually severe in-
IFN-α/β and -λ outside the CNS. fectious diseases, especially of viral origin [78]. The
A heterozygous autosomal dominant negative mutation (R118W) was associated with a reduced
mutation was subsequently identified in the TLR3 stability of the protein. The production of IFN-β

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 309

Table 2. Genetic predispositions to herpes simplex virus encephalitis in humans


Mutated gene/protein No. of identified patients
Mutation characteristics Susceptibility to infections Immunological findings References

TLR3-dependent pathway deficiency


Unc93B1/Unc93B Two children with Loss-of-expression [74]
1034del4 and G781A recurrent HSE Impaired IFN-β and -λ in
Homozygous Normal resistance to other fibroblasts in response to
Autosomal recessive infectious diseases poly I:C and HSV-1
complete including viral infections Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
Impaired IFN-α/β and -λ
in PBMCs in response to
agonists of TLRs 7, 8 and 9
TLR3/TLR3 Two children with HSE Loss-of-function [75]
P554S One of the two children Partial defect in IFN-β and
Heterozygous had recurrences -λ in fibroblasts in
Autosomal dominant Normal resistance to other response to poly I:C
negative infectious diseases and HSV-1
including viral infections Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
Normal response to poly
I:C in PBMCs
TLR3/TLR3 One child with HSE Premature termination, [76]
P554S and E746stop Normal resistance to other loss-of-function
Heterozygous infectious diseases Abolition of IFN-β and -λ
Autosomal recessive including viral infections in fibroblasts in response
complete to poly I:C and HSV-1
Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
Normal response to poly
I:C or HSV-1 in PBMCs
TLR3/TLR3 One child with recurrent Loss-of-expression, loss- [77]
G743D and R811I HSE of-function
Heterozygous Normal resistance to other IFN-β and -λ responses
Autosomal dominant infectious diseases including almost abolished in
partial viral infections fibroblasts stimulated
Haploinsufficiency with poly I:C and HSV-1
Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
(Continues)

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
310 J. Piret and G. Boivin

Table 2.
. (Continued)
(Continued)
Mutated gene/protein No. of identified patients
Mutation characteristics Susceptibility to infections Immunological findings References

TLR3/TLR3 One child with recurrent Lack of cleavage, loss-of-


L360P HSE function [77]
Heterozygous Normal resistance to other Severely impaired IFN-β
Autosomal dominant infectious diseases and -λ responses in
negative, partial including viral infections fibroblasts stimulated
with poly I:C and HSV-1
Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
TLR3/TLR3 One child with HSE Loss-of-function [77]
R867Q followed by HSE relapse Severely impaired IFN-β
Homozygous Normal resistance to other and -λ responses in
Autosomal recessive infectious diseases fibroblasts stimulated
partial including viral infections with poly I:C and HSV-1
Hypomorphic Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
TRAF3/TRAF3 One child with HSE Loss-of-expression, loss- [78]
R118W Normal resistance to other of-function
Heterozygous infectious diseases Impaired IFN-β and -λ
Autosomal dominant including viral responses in fibroblasts
negative, partial infections stimulated with poly I:C
and HSV-1
Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
TRIF/TRIF One child with HSE Loss-of-expression, loss- [80]
R141stop Normal resistance to other of-function
Homozygous infectious diseases Abolition of IFN-β and -λ
Autosomal recessive including viral infections in fibroblasts in response
complete to poly I:C and HSV-1
Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b
TRIF/TRIF One out of 3 children Not a complete loss-of- [80]
S186L with HSE function
Heterozygous Impaired IFN-β and -λ in
fibroblasts in response to
poly I:C and HSV-1

(Continues)

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 311

Table 2.
. (Continued)
(Continued)
Mutated gene/protein No. of identified patients
Mutation characteristics Susceptibility to infections Immunological findings References

Autosomal dominant Fibroblasts more


negative, partial permissive to HSV-1
Hypomorphic Rescue with exogenous
IFN-α2b

TBK1/TBK1 One child with HSE Loss-of-function [81]


G159A Normal resistance to other Severely impaired
infectious diseases IFN-β and -λ in
Heterozygous including viral infections fibroblasts in response
Autosomal dominant to poly I:C
negative, partial Impaired IFN-β and -λ in
fibroblasts in
response to HSV-1
Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b

TBK1/TBK1 One child with HSE Loss-of-expression, loss- [81]


D50A Normal resistance to other of-function
Heterozygous infectious diseases Normal IFN-β and -λ in
Autosomal dominant including viral infections fibroblasts in
partial response to poly I:C
Haploinsufficiency Impaired IFN-β and -λ in
fibroblasts in
response to HSV-1
Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b

Canonical IKK complex regulation deficiency


IKBKG/NEMO One child with lethal HSE Expression of a truncated [84]
110-111insC Susceptibility to infections protein
X-linked recessive by pyogenic bacteria, Impaired IFN-β and -λ in
Hypomorphic mycobacteria and to a fibroblasts in response to
lesser extent viruses poly I:C and HSV-1
and fungi Fibroblasts more
permissive to HSV-1
Rescue with exogenous
IFN-α2b

(Continues)

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
312 J. Piret and G. Boivin

Table 2.
. (Continued)
(Continued)
Mutated gene/protein No. of identified patients
Mutation characteristics Susceptibility to infections Immunological findings References

Type 1 IFN response deficiency


STAT1/STAT1 One child with recurrent Loss-of-expression, loss- [90]
1757-1758delAG HSE; lethal of-function
Homozygous Severe mycobacterial and Response of fibroblasts to
Autosomal recessive viral diseases IFN-α/β impaired
complete IFN-α2b had no effect on
HSV replication in
fibroblasts
Classical MyD88-dependent pathway deficiency
MyD88/MyD88 26 patients No IL-6 production by
Autosomal recessive No increase susceptibility to whole blood cells after [89,93]
HSE activation with IL-1 or
Susceptibility to infections by agonists of TLRs 7, 8 and 9
pyogenic bacteria and normal Normal responses to
resistance to fungi, parasites, TLR3 agonist
viruses and many bacteria
IRAK-4/IRAK-4 52 patients No IFN-α/β and -λ
Autosomal recessive No increase susceptibility to production by whole [89,94]
HSE blood cells and fibroblasts
Susceptibility to infections after activation with IL-1
by pyogenic bacteria and or agonists of TLRs 7, 8
normal resistance to and 9
fungi, parasites, viruses Normal responses to
and many bacteria TLR3 agonist

and -λ was impaired in the patient’s fibroblasts protein whereas the other one had a heterozygous
stimulated with poly I:C and HSV-1 and resulted autosomal dominant mutation (S186L) that re-
in an increased viral replication and cell death that sulted in an altered protein activity. The production
could be rescued by pre-treatment with exogenous of IFN-β and -λ in response to activation with poly
IFN-α2b. Apart from TLR3, the TRAF3 protein acts I:C and HSV-1 was respectively abolished or im-
downstream from other receptors inducing the pro- paired in fibroblasts of the first and the second pa-
duction of type I and III IFNs as well as the TNF re- tient. Fibroblasts of both patients were more
ceptor [79]. Therefore, the cellular response was permissive to HSV-1 infection and could be rescued
also impaired following specific stimulation of by pre-treatment with exogenous IFN-α2b. Finally,
these pathways. The TRAF3-deficient patient was mutations in the TBK1 gene were identified in
susceptible to HSE only suggesting that impair- two children [81]. TBK1 is a kinase at the crossroad
ment of TLR3-dependent induction of IFNs in the of signaling pathways leading to the production of
CNS could be involved. Two other children with type I IFN such as TLR3 and sensors of cytosolic
HSE who harbored mutations in the TRIF gene dsRNA and dsDNA [82]. The first patient had a
coding for the TRIF adaptor protein of TLR3 were heterozygous autosomal dominant negative muta-
described [80]. The first patient carried a homozy- tion (G159A) that resulted in a loss-of-function of
gous autosomal recessive complete mutation the kinase activity. The IFN-β and -λ responses
(R141stop) that led to a loss-of-expression of the were severely impaired in patient’s fibroblasts

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 313

stimulated with poly I:C and HSV-1. The second pyogenic bacteria, mycobacteria and to a lesser ex-
patient carried a heterozygous autosomal domi- tent viruses and fungi [86–89]. However, patients
nant mutation (D50A) that led to a decrease in the carrying NEMO mutations could potentially be
expression and stability of the protein. The re- more vulnerable to HSE.
sponse of patient’s fibroblasts to activation with Mutations in the STAT1 gene were identified in
poly I:C was normal whereas it was impaired fol- several patients and were mostly associated with
lowing stimulation with HSV-1. an increased susceptibility to mycobacterial and viral
Children with inborn errors in TLR3 immunity diseases [90–92]. However, one of these patients died
suffering from HSE are not susceptible to other vi- from disseminated HSV-1 infection with recurrent
ral infections suggesting that TLR3-independent HSE [90]. This child carried a homozygous autoso-
IFN responses to other viral components may con- mal recessive complete mutation (1757–1758delAG)
fer protection against a wide range of viruses in that resulted in a loss-of-expression and loss-of-
these patients. Moreover, these children are not function of the protein and the abolition of IFN-α/β
prone to HSV-1 infections outside the CNS. Thus, and -λ cellular response to poly I:C and HSV-1 stim-
in humans, TLR3 could play a non-redundant role ulations. Pre-treatment of patient’s fibroblasts with
in the CNS for protection against HSV-1 during pri- exogenous IFN-α2b had no effect on HSV-1 replica-
mary infection but could be otherwise largely re- tion. Because of the important role of STAT1 in the
dundant in host defense [75,76]. It was suggested type I IFN response, it is suggested that its defi-
that the restricted susceptibility to HSE seems to re- ciency could predispose individuals to multiple vi-
sult from impaired TLR3-dependent type I IFN ral infections, including HSE.
production by non-hematopoietic cells of the CNS Deficiencies in the MyD88 adaptor protein and
infected with HSV-1 and that TLR3-independent its associated cytosolic IL-1 receptor-associated ki-
or residual TLR3-dependent signaling outside the nase 4 (IRAK-4) were identified in several patients
CNS may contribute to antiviral immunity. To with increased susceptibility to infections caused
strengthen this hypothesis, induced pluripotent by pyogenic bacteria and normal resistance to
stem cells (iPSCs) derived from patients deficient fungi, parasites, many bacteria and viruses includ-
for Unc93B and TLR3 were differentiated into neu- ing HSV-1 [89]. The production of IL-6 was im-
rons, astrocytes and oligodendrocytes [83]. Neu- paired in whole blood cells from MyD88-deficient
rons and oligodendrocytes were more susceptible patients after activation with IL-1 or agonists of
to HSV-1 infection compared to differentiated con- TLRs 7, 8 and 9 but not following stimulation with
trol iPSCs and viral infection could be rescued by agonists of TLR3 [93]. Whole blood cells and fibro-
a pre-treatment with exogenous IFN-α or IFN-β indi- blasts of IRAK-4-deficient patients did not produce
cating that non-hematopoietic cells of the CNS could IFN-α/β and -λ in response to IL-1 and agonists of
be involved in the protective immunity against HSE. TLRs 7, 8 and 9 but the response to TLR3 agonist
It is thus suggested that non-hematopoietic intrinsic was normal [94]. These data suggest that the classi-
immunity could play a role in the protection against cal MyD88-dependent pathway is redundant in the
HSV-1 primary infection of the CNS beyond innate protective immunity against HSE and reinforce the
and adaptive hematopoietic immunity. notion that the TLR3-dependent pathway is pre-
A mutation in the IKBKG gene coding for the NF- dominant in that context.
κB essential modulator (NEMO) protein was
identified in a child who died from HSE [84]. IMMUNOMODULATORY STRATEGIES
NEMO is a regulatory subunit of the IKK complex AGAINST HERPES SIMPLEX VIRUS
activating the canonical NF-κB signaling pathway ENCEPHALITIS
which acts downstream of multiple receptors [85]. The cerebral innate immune response that develops
The X-linked recessive mutation (110-111insC) was during HSE is a «double-edged sword» as it is crit-
hypomorphic and resulted in the expression of a ical to control HSV replication in the brain early af-
truncated protein. Both fibroblasts and PBMCs iso- ter infection but, if uncontrolled, may also result in
lated from this patient had impaired IFN-β and -λ an exaggerated inflammatory response that could
production in response to poly I:C and HSV-1 acti- be detrimental to the host. Thus, the host innate im-
vation. This deficiency is mostly associated with an mune response must be finely balanced to suppress
increased susceptibility to infections caused by viral replication and cell death while limiting

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
314 J. Piret and G. Boivin

inflammation in a sensitive tissue such as the brain Our group demonstrated that the administration
(Figure 2). Immunomodulatory strategies aimed at of glucocorticoids initiated on day 3 (at the onset of
blocking the innate immune response at a critical symptoms) after intranasal infection of mice with
time after infection may thus have beneficial effects HSV-1 was associated with reduced viral replication
by controlling the late onset excessive inflamma- and cytokine levels in the CNS and an increased sur-
tion of the CNS that develops during HSE. Such vival rate [58]. In contrast, glucocorticoids treatment
immunomodulatory strategies could be eventually initiated the day of infection (day 0) prevented the
combined with specific antiviral therapy. control of viral replication and was detrimental to
The mechanism of action of glucocorticoids, which the host. Thus, contrary to other CNS infections
are broad-spectrum anti-inflammatory agents, im- (e.g. bacterial meningitis) which require early ad-
plies the promotion of anti-inflammatory monocytes ministration of steroids, this study demonstrated
which seem to be actively involved in the resolution that treatment with immunomodulatory drugs
of immune reactions [95,96]. Several human case re- should be delayed to improve the outcome of HSE.
ports and series have indicated a clinical improve- In this respect, the modulation of the immune re-
ment in patients suffering from HSE treated with sponse by TLR agonists or antagonists adminis-
corticosteroids combined with acyclovir [97–100]. A tered before or after infection of mice, respectively,
retrospective analysis of a larger number of patients could improve the outcome of HSE. For instance,
with HSE showed that administration of adjunctive stimulation of TLR3 by pre-treatment of mice with
glucocorticoids was an independent predictor of a poly I:C before HSV-1 challenge reduced viral rep-
better neurological outcome [101]. Also, a recent re- lication and the severity of infection [104]. In a
view of animal and clinical studies suggests a benefit similar manner, the pre-treatment of mice with
of adjunctive corticosteroids therapy in HSE [102]. TLR9 agonists [oligodeoxynucleotides (ODNs)
However, clear evidences are still lacking to translate containing unmethylated CpG motifs] before
this approach into clinical practice. A clinical trial HSV-1 infection reduced brain viral load and in-
evaluating a combination of acyclovir with glucocor- creased survival rate [105]. More clinically-relevant
ticoids is currently ongoing (German trial of acyclo- data showed that treatment of mice with an antag-
vir and corticosteroids in herpes simplex virus onist of TLR9 given on day 3 post-infection in-
encephalitis; GACHE) [103]. creased the survival rate likely by controlling the
inflammatory response that could be detrimental
to the host [105].
The added benefit of combining an antiviral
agent with an immunomodulatory drug for the
treatment of HSE has also been investigated in
mouse models. We reported that treatment of mice
with a combination of valacyclovir (a prodrug of
acyclovir) and an anti-TNF-α antibody (i.e.
etanercept) initiated on day 3 post-infection signifi-
cantly increased survival rate compared to antiviral
therapy alone [106].
Overall, these data suggest that antiviral agents
could be combined with immune-based therapy to
improve the outcome of patients suffering from HSE.
However, biomarkers of inflammation in the blood
or in the CSF [107,108] should be identified for timely
administration of immunomodulatory therapy.

Figure 2. Potential management of herpes simplex virus encepha- MANAGEMENT OF HSE IN CHILDREN WITH
litis. Antiviral agents that suppress the viral replication could be INHERITED DEFICIENCIES IN INNATE
combined with immunomodulatory drugs to reduce the exagger-
ated inflammatory response that develops in the CNS at later times
IMMUNITY
after infection. Biomarkers of inflammation should be identified As in the general population, antiviral therapy
for timely administration of adjunctive immune-based therapy should be initiated early in children with inherited

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 315

deficiencies of the innate immune system present- chemokines. Experimental mouse models of HSE
ing suspected or proven HSE [19]. It has been re- have demonstrated the essential role of the signal-
ported that the clinical penetrance of TLR3 ing pathways that are implicated in the recognition
pathway deficiencies is incomplete as some family of HSV and of the inflammatory mediators that
members of affected children did not develop participate in the protective immunity against this
HSE despite HSV-1 infection [109]. The age at infec- virus. In addition, recent studies have suggested
tion may be a key factor for the development of that primary natural infections in some children
HSE in genetically predisposed children probably may result from HSV-1 specific, CNS-intrinsic,
because of an immature immune system [7]. There- single-gene inborn errors of IFN-mediated immu-
fore, children in at risk families could be provided nity [112]. Several homozygous or heterozygous
with prenatal or neonatal genetic testing to allow mutations were identified in a gene of the TLR3 sig-
appropriate follow-up and clinical care until ado- naling pathway (TLR3, Unc93B1, TRIF, TRAF3 and
lescence or documented HSV-1 seroconversion. TBK1) [109]. Overall, these inherited deficiencies
Moreover, children with an inherited deficiency in have been reported in 13 (11%) of 120 children with
TLR3 immunity suffering from HSE should be care- HSE studied to date [77]. Nevertheless, these data
fully followed because of an increased risk of recur- allow a better understanding of the components
rences [77]. Such children may benefit from of the immune system implicated in the recogni-
antiviral prophylaxis. The use of adjunctive tion of HSV. However, at later times after the in-
immune-based therapy in such patients should re- fection, the immune response to the virus can
inforce the innate immune response rather than re- become exaggerated and contribute to the dam-
ducing an overzealous inflammation that would ages occurring in the CNS. Despite antiviral ther-
develop at later times after the infection. For in- apy with acyclovir, the mortality rate is still close
stance, it was suggested that children with first ep- to 30% with almost 60% of surviving individuals
isode of HSE or HSE relapses associated with a developing neurological sequelae. The host innate
genetic defect in IFN-α production or a partial de- immune response must be finely balanced to
fect in IFN-α response would benefit from a combi- suppress viral replication and cell death while
nation therapy of acyclovir and recombinant IFN-α limiting the inflammatory response in the CNS.
2b [110]. In this respect, a combination of acyclovir Thus, it would be beneficial to combine antiviral
and IFN-α administered on day 1 or 2 after infec- agents with immune-based therapy to improve
tion with HSV showed a 30% increase in the sur- the prognosis of patients suffering from HSE. Bio-
vival rate of mice compared to those treated with markers of inflammation should nevertheless be
acyclovir alone [111]. identified for timely administration of immuno-
modulatory compounds.
CONCLUSIONS AND PERSPECTIVES
The innate immune system is involved in the sens-
ing of HSV and induces the production of type I CONFLICT OF INTEREST
IFN as well as pro-inflammatory cytokines and The authors have no competing interest.

REFERENCES
1. Roizman B, Knipe DM, Whitley RJ. Her- edn, Richman DD, Whitley RJ, Hayden arabinoside therapy of biopsy-proved her-
pes simplex viruses. In Fields Virology, FG (eds). ASM Press: Washington, D.C., pes simplex encephalitis. National Institute
Vol. 2, 6th edn, Knipe DM, Howley PM 2009; 409–436. of Allergy and Infectious Diseases colla-
(eds). Lippincott Williams & Wilkins: Balti- 4. Whitley RJ. Herpes simplex encephalitis: borative antiviral study. The New England
more, 2013; 1823–1897. adolescents and adults. Antiviral Research Journal of Medicine 1977; 297: 289–294.
2. Bradley H, Markowitz LE, Gibson T, 2006; 71: 141–148. DOI: 10.1056/NEJM197708112970601
McQuillan GM. Seroprevalence of herpes 5. Dagsdottir HM, Sigurethardottir B, 7. Abel L, Plancoulaine S, Jouanguy E, et al.
simplex virus types 1 and 2—United States, Gottfreethsson M, et al. Herpes simplex Age-dependent Mendelian predisposition
1999–2010. The Journal of Infectious Diseases encephalitis in Iceland 1987–2011. Springerplus to herpes simplex virus type 1 encephalitis
2014; 209: 325–333. DOI: 10.1093/infdis/jit458 2014; 3: 524. DOI: 10.1186/2193-1801-3-524 in childhood. The Journal of Pediatrics 2010;
3. Whitley RJ, Roizman B. Herpes simplex 6. Whitley RJ, Soong SJ, Dolin R, Galasso GJ, 157: 623–629, 629 e621. DOI: 10.1016/j.
viruses. In Clinical Virology, Vol. 1, 3rd Ch’ien LT, Alford CA. Adenine jpeds.2010.04.020

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
316 J. Piret and G. Boivin

8. Garcia JH, Colon LE, Whitley RJ, Wilmes and correlation with disease. National In- 28. Aravalli RN, Hu S, Lokensgard JR. Toll-
FJ. Diagnosis of viral encephalitis by brain stitute of Allergy and Infectious Diseases like receptor 2 signaling is a mediator of
biopsy. Seminars in Diagnostic Pathology Collaborative Antiviral Study Group. The apoptosis in herpes simplex virus-
1984; 1: 71–81. Journal of Infectious Diseases 1995; 171: infected microglia. Journal of Neuroinflam-
9. Johnson RT, Olson LC, Buescher EL. Her- 857–863. mation 2007; 4: 11. DOI: 10.1186/1742-
pes simplex virus infections of the nervous 18. DeBiasi RL, Tyler KL. Polymerase chain re- 2094-4-11
system. Problems in laboratory diagnosis. action in the diagnosis and management 29. Lund J, Sato A, Akira S, Medzhitov R,
Archives of Neurology 1968; 18: 260–264. of central nervous system infections. Ar- Iwasaki A. Toll-like receptor 9-mediated
10. Schlitt M, Lakeman AD, Wilson ER, et al. A chives of Neurology 1999; 56: 1215–1219. recognition of Herpes simplex virus-2
rabbit model of focal herpes simplex en- 19. Erdem H, Cag Y, Ozturk-Engin D, et al. by plasmacytoid dendritic cells. The Jour-
cephalitis. The Journal of Infectious Diseases Results of a multinational study suggest nal of Experimental Medicine 2003; 198:
1986; 153: 732–735. the need for rapid diagnosis and early an- 513–520. DOI: 10.1084/jem.20030162
11. Hudson SJ, Dix RD, Streilein JW. Induction tiviral treatment at the onset of herpetic 30. Krug A, Luker GD, Barchet W, Leib DA,
of encephalitis in SJL mice by intranasal in- meningoencephalitis. Antimicrobial Agents Akira S, Colonna M. Herpes simplex virus
fection with herpes simplex virus type 1: a and Chemotherapy 2015; 59: 3084–3089. type 1 activates murine natural interferon-
possible model of herpes simplex encepha- DOI: 10.1128/AAC.05016-14 producing cells through toll-like receptor
litis in humans. The Journal of Infectious Dis- 20. Takeuchi O, Akira S. Pattern recognition re- 9. Blood 2004; 103: 1433–1437. DOI:
eases 1991; 163: 720–727. ceptors and inflammation. Cell 2010; 140: 10.1182/blood-2003-08-2674
12. Barnett EM, Jacobsen G, Evans G, Cassell 805–820. DOI: 10.1016/j.cell.2010.01.022 31. Hochrein H, Schlatter B, O’Keeffe M, et al.
M, Perlman S. Herpes simplex encephalitis 21. Akira S, Uematsu S, Takeuchi O. Pathogen Herpes simplex virus type-1 induces
in the temporal cortex and limbic system recognition and innate immunity. Cell IFN-alpha production via Toll-like recep-
after trigeminal nerve inoculation. The 2006; 124: 783–801. DOI: 10.1016/j.cell. tor 9-dependent and -independent path-
Journal of Infectious Diseases 1994; 169: 2006.02.015 ways. Proceedings of the National Academy
782–786. 22. Kawai T, Akira S. The role of pattern- of Sciences of the United States of America
13. Koprowski H, Zheng YM, Heber-Katz E, recognition receptors in innate immunity: 2004; 101: 11416–11421. DOI: 10.1073/
et al. In vivo expression of inducible nitric update on Toll-like receptors. Nature Im- pnas.0403555101
oxide synthase in experimentally induced munology 2010; 11: 373–384. DOI: 32. Li XD, Wu J, Gao D, Wang H, Sun L, Chen
neurologic diseases. Proceedings of the Na- 10.1038/ni.1863 ZJ. Pivotal roles of cGAS-cGAMP signal-
tional Academy of Sciences of the United 23. Kim YM, Brinkmann MM, Paquet ME, ing in antiviral defense and immune adju-
States of America 1993; 90: 3024–3027. Ploegh HL. UNC93B1 delivers nucleotide- vant effects. Science 2013; 341: 1390–1394.
14. Lundberg P, Ramakrishna C, Brown J, sensing toll-like receptors to endolysosomes. DOI: 10.1126/science.1244040
et al. The immune response to herpes sim- Nature 2008; 452: 234–238. DOI: 10.1038/ 33. Shu C, Li X, Li P. The mechanism of
plex virus type 1 infection in susceptible nature06726 double-stranded DNA sensing through
mice is a major cause of central nervous 24. Melchjorsen J. Sensing herpes: more than the cGAS–STING pathway. Cytokine &
system pathology resulting in fatal en- toll. Reviews in Medical Virology 2012; 22: Growth Factor Reviews 2014. DOI:
cephalitis. Journal of Virology 2008; 82: 106–121. DOI: 10.1002/rmv.716 10.1016/j.cytogfr.2014.06.006
7078–7088. DOI: 10.1128/JVI.00619-08 25. Vandevenne P, Sadzot-Delvaux C, Piette J. 34. Cai X, Chiu YH, Chen ZJ. The cGAS–
15. Sili U, Kaya A, Mert A, Group HSVES. Innate immune response and viral inter- cGAMP–STING pathway of cytosolic
Herpes simplex virus encephalitis: clinical ference strategies developed by human DNA sensing and signaling. Molecular Cell
manifestations, diagnosis and outcome in herpesviruses. Biochemical Pharmacology 2014; 54: 289–296. DOI: 10.1016/j.
106 adult patients. Journal of Clinical Virol- 2010; 80: 1955–1972. DOI: 10.1016/j.bcp. molcel.2014.03.040
ogy 2014; 60: 112–118. DOI: 10.1016/j. 2010.07.001 35. Luecke S, Paludan SR. Innate recognition
jcv.2014.03.010 26. Paludan SR, Bowie AG, Horan KA, of alphaherpesvirus DNA. Advances in Vi-
16. Steiner I, Budka H, Chaudhuri A, et al. Vi- Fitzgerald KA. Recognition of herpesvi- rus Research 2015; 92: 63–100. DOI:
ral meningoencephalitis: a review of diag- ruses by the innate immune system. 10.1016/bs.aivir.2014.11.003
nostic methods and guidelines for Nature Reviews Immunology 2011; 11: 36. Weber F, Wagner V, Rasmussen SB,
management. European Journal of Neurol- 143–154. DOI: 10.1038/nri2937 Hartmann R, Paludan SR. Double-
ogy 2010; 17: 999–e957. DOI: 10.1111/ 27. Aravalli RN, Hu S, Rowen TN, Palmquist stranded RNA is produced by positive-
j.1468-1331.2010.02970.x JM, Lokensgard JR. Cutting edge: TLR2- strand RNA viruses and DNA viruses
17. Lakeman FD, Whitley RJ. Diagnosis of mediated proinflammatory cytokine and but not in detectable amounts by
herpes simplex encephalitis: application chemokine production by microglial cells negative-strand RNA viruses. Journal of Vi-
of polymerase chain reaction to cerebro- in response to herpes simplex virus. Jour- rology 2006; 80: 5059–5064. DOI: 10.1128/
spinal fluid from brain-biopsied patients nal of Immunology 2005; 175: 4189–4193. JVI.80.10.5059-5064.2006

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 317

37. Kato H, Takeuchi O, Mikamo-Satoh E, et al. virus infection: activation of a transcrip- 57. Boivin N, Menasria R, Gosselin D, Rivest S,
Length-dependent recognition of double- tion factor complex containing IRF-3 and Boivin G. Impact of deficiency in CCR2 and
stranded ribonucleic acids by retinoic acid- CBP/p300. The EMBO Journal 1998; 17: CX3CR1 receptors on monocytes trafficking
inducible gene-I and melanoma 1087–1095. DOI: 10.1093/emboj/17.4.1087 in herpes simplex virus encephalitis. The
differentiation-associated gene 5. The Journal 46. Honda K, Taniguchi T. IRFs: master regu- Journal of General Virology 2012; 93:
of Experimental Medicine 2008; 205: lators of signalling by Toll-like receptors 1294–1304. DOI: 10.1099/vir.0.041046-0
1601–1610. DOI: 10.1084/jem.20080091 and cytosolic pattern-recognition recep- 58. Sergerie Y, Boivin G, Gosselin D, Rivest S.
38. Rasmussen SB, Sorensen LN, Malmgaard tors. Nature Reviews Immunology 2006; 6: Delayed but not early glucocorticoid treat-
L, et al. Type I interferon production dur- 644–658. DOI: 10.1038/nri1900 ment protects the host during experimen-
ing herpes simplex virus infection is 47. Honda K, Yanai H, Negishi H, et al. IRF-7 is tal herpes simplex virus encephalitis in
controlled by cell-type-specific viral recog- the master regulator of type-I interferon- mice. The Journal of Infectious Diseases
nition through Toll-like receptor 9, the mi- dependent immune responses. Nature 2005; 2007; 195: 817–825.
tochondrial antiviral signaling protein 434: 772–777. DOI: 10.1038/nature03464 59. Kielian T. Overview of toll-like receptors
pathway, and novel recognition systems. 48. Schindler C, Levy DE, Decker T. JAK–STAT in the CNS. Current Topics in Microbiology
Journal of Virology 2007; 81: 13315–13324. signaling: from interferons to cytokines. The and Immunology 2009; 336: 1–14. DOI:
DOI: 10.1128/JVI.01167-07 Journal of Biological Chemistry 2007; 282: 10.1007/978-3-642-00549-7_1
39. Rasmussen SB, Jensen SB, Nielsen C, et al. 20059–20063. DOI: 10.1074/jbc.R700016200 60. Furr SR, Chauhan VS, Sterka D Jr,
Herpes simplex virus infection is sensed 49. Prakash A, Smith E, Lee CK, Levy DE. Grdzelishvili V, Marriott I. Characterization
by both Toll-like receptors and retinoic Tissue-specific positive feedback require- of retinoic acid-inducible gene-I expression
acid-inducible gene- like receptors, which ments for production of type I interferon in primary murine glia following exposure
synergize to induce type I interferon pro- following virus infection. The Journal of Bi- to vesicular stomatitis virus. Journal of
duction. The Journal of General Virology ological Chemistry 2005; 280: 18651–18657. Neurovirology 2008; 14: 503–513. DOI:
2009; 90: 74–78. DOI: 10.1099/ DOI: 10.1074/jbc.M501289200 10.1080/13550280802337217
vir.0.005389-0 50. Sato M, Suemori H, Hata N, et al. Distinct 61. Peltier DC, Simms A, Farmer JR, Miller DJ.
40. Melchjorsen J, Rintahaka J, Soby S, et al. and essential roles of transcription factors Human neuronal cells possess functional
Early innate recognition of herpes simplex IRF-3 and IRF-7 in response to viruses for cytoplasmic and TLR-mediated innate
virus in human primary macrophages is IFN-alpha/beta gene induction. Immunity immune pathways influenced by
mediated via the MDA5/MAVS- 2000; 13: 539–548. phosphatidylinositol-3 kinase signaling.
dependent and MDA5/MAVS/RNA po- 51. Sadler AJ, Williams BR. Interferon-induc- Journal of Immunology 2010; 184:
lymerase III-independent pathways. Jour- ible antiviral effectors. Nature Reviews Im- 7010–7021. DOI: 10.4049/jimmunol.0904133
nal of Virology 2010; 84: 11350–11358. DOI: munology 2008; 8: 559–568. DOI: 10.1038/ 62. Furr SR, Chauhan VS, Moerdyk-
10.1128/JVI.01106-10 nri2314 Schauwecker MJ, Marriott I. A role for
41. van Pesch V, Lanaya H, Renauld JC, 52. Rivest S. Regulation of innate immune re- DNA-dependent activator of interferon
Michiels T. Characterization of the murine sponses in the brain. Nature Reviews Immu- regulatory factor in the recognition of her-
alpha interferon gene family. Journal of Vi- nology 2009; 9: 429–439. DOI: 10.1038/ pes simplex virus type 1 by glial cells. Jour-
rology 2004; 78: 8219–8228. DOI: 10.1128/ nri2565 nal of Neuroinflammation 2011; 8: 99. DOI:
JVI.78.15.8219-8228.2004 53. Nadeau S, Rivest S. Regulation of the gene 10.1186/1742-2094-8-99
42. Barnes B, Lubyova B, Pitha PM. On the encoding tumor necrosis factor alpha 63. Mansur DS, Kroon EG, Nogueira ML, et al.
role of IRF in host defense. Journal of Inter- (TNF-alpha) in the rat brain and pituitary Lethal encephalitis in myeloid differentia-
feron & Cytokine Research 2002; 22: 59–71. in response in different models of systemic tion factor 88-deficient mice infected with
DOI: 10.1089/107999002753452665 immune challenge. Journal of Neuropathol- herpes simplex virus 1. American Journal
43. Fitzgerald KA, McWhirter SM, Faia KL, ogy and Experimental Neurology 1999; 58: of Pathology 2005; 166: 1419–1426. DOI:
et al. IKKepsilon and TBK1 are essential 61–77. 10.1016/S0002-9440(10)62359-0
components of the IRF3 signaling path- 54. Galea I, Bechmann I, Perry VH. What is 64. Lima GK, Zolini GP, Mansur DS, et al.
way. Nature Immunology 2003; 4: 491–496. immune privilege (not)? Trends in Immu- Toll-like receptor (TLR) 2 and TLR9
DOI: 10.1038/ni921 nology 2007; 28: 12–18. DOI: 10.1016/j. expressed in trigeminal ganglia are criti-
44. Solt LA, May MJ. The IkappaB kinase it.2006.11.004 cal to viral control during herpes simplex
complex: master regulator of NF-kappaB 55. Soulet D, Rivest S. Bone-marrow-derived virus 1 infection. American Journal of Pathol-
signaling. Immunologic Research 2008; 42: microglia: myth or reality? Current Opinion ogy 2010; 177: 2433–2445. DOI: 10.2353/
3–18. DOI: 10.1007/s12026-008-8025-1 in Pharmacology 2008; 8: 508–518. DOI: ajpath.2010.100121
45. Yoneyama M, Suhara W, Fukuhara Y, 10.1016/j.coph.2008.04.002 65. Kurt-Jones EA, Chan M, Zhou S, et al. Her-
Fukuda M, Nishida E, Fujita T. Direct trig- 56. Dong Y, Benveniste EN. Immune function pes simplex virus 1 interaction with Toll-
gering of the type I interferon system by of astrocytes. Glia 2001; 36: 180–190. like receptor 2 contributes to lethal

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
318 J. Piret and G. Boivin

encephalitis. Proceedings of the National UNC-93B deficiency. Science 2006; 314: cells. Nature 2012; 491: 769–773. DOI:
Academy of Sciences of the United States of 308–312. DOI: 10.1126/science.1128346 10.1038/nature11583
America 2004; 101: 1315–1320. DOI: 75. Zhang SY, Jouanguy E, Ugolini S, et al. 84. Audry M, Ciancanelli M, Yang K, et al.
10.1073/pnas.0308057100 TLR3 deficiency in patients with herpes NEMO is a key component of NF-
66. Wang JP, Bowen GN, Zhou S, et al. Role simplex encephalitis. Science 2007; 317: kappaB- and IRF-3-dependent TLR3-
of specific innate immune responses in 1522–1527. DOI: 10.1126/science.1139522 mediated immunity to herpes simplex
herpes simplex virus infection of the cen- 76. Guo Y, Audry M, Ciancanelli M, et al. Her- virus. The Journal of Allergy and Clinical
tral nervous system. Journal of Virology pes simplex virus encephalitis in a patient Immunology 2011; 128: 610–617 e611-614.
2012; 86: 2273–2281. DOI: 10.1128/ with complete TLR3 deficiency: TLR3 is DOI: 10.1016/j.jaci.2011.04.059
JVI.06010-11 otherwise redundant in protective immu- 85. Clark K, Nanda S, Cohen P. Molecular
67. Sorensen LN, Reinert LS, Malmgaard L, nity. The Journal of Experimental Medicine control of the NEMO family of ubiquitin-
Bartholdy C, Thomsen AR, Paludan SR. 2011; 208: 2083–2098. DOI: 10.1084/ binding proteins. Nature Reviews Molecular
TLR2 and TLR9 synergistically control her- jem.20101568 Cell Biology 2013; 14: 673–685. DOI:
pes simplex virus infection in the brain. Jour- 77. Lim HK, Seppanen M, Hautala T, et al. 10.1038/nrm3644
nal of Immunology 2008; 181: 8604–8612. TLR3 deficiency in herpes simplex enceph- 86. Niehues T, Reichenbach J, Neubert J, et al.
68. Tengvall S, Harandi AM. Importance of alitis: high allelic heterogeneity and recur- Nuclear factor kappaB essential
myeloid differentiation factor 88 in innate rence risk. Neurology 2014; 83: 1888–1897. modulator-deficient child with immuno-
and acquired immune protection against DOI: 10.1212/WNL.0000000000000999 deficiency yet without anhidrotic ectoder-
genital herpes infection in mice. Journal of 78. Perez de Diego R, Sancho-Shimizu V, mal dysplasia. The Journal of Allergy and
Reproductive Immunology 2008; 78: 49–57. Lorenzo L, et al. Human TRAF3 adaptor Clinical Immunology 2004; 114: 1456–1462.
DOI: 10.1016/j.jri.2007.09.001 molecule deficiency leads to impaired DOI: 10.1016/j.jaci.2004.08.047
69. Menasria R, Boivin N, Lebel M, Piret J, Toll-like receptor 3 response and suscepti- 87. Puel A, Reichenbach J, Bustamante J, et al.
Gosselin J, Boivin G. Both TRIF and IPS-1 bility to herpes simplex encephalitis. Im- The NEMO mutation creating the most-
adaptor proteins contribute to the cerebral munity 2010; 33: 400–411. DOI: 10.1016/j. upstream premature stop codon is
innate immune response against herpes immuni.2010.08.014 hypomorphic because of a reinitiation of
simplex virus 1 infection. Journal of Virol- 79. Hacker H, Tseng PH, Karin M. Expanding translation. American Journal of Human
ogy 2013; 87: 7301–7308. DOI: 10.1128/ TRAF function: TRAF3 as a tri-faced Genetics 2006; 78: 691–701. DOI: 10.1086/
JVI.00591-13 immune regulator. Nature Reviews Immu- 501532
70. Reinert LS, Harder L, Holm CK, et al. TLR3 nology 2011; 11: 457–468. DOI: 10.1038/ 88. Puel A, Picard C, Ku CL, Smahi A,
deficiency renders astrocytes permissive to nri2998 Casanova JL. Inherited disorders of
herpes simplex virus infection and facili- 80. Sancho-Shimizu V, Perez de Diego R, NF-kappaB-mediated immunity in man.
tates establishment of CNS infection in Lorenzo L, et al. Herpes simplex encephali- Current Opinion in Immunology 2004; 16:
mice. The Journal of Clinical Investigation tis in children with autosomal recessive 34–41.
2012; 122: 1368–1376. DOI: 10.1172/ and dominant TRIF deficiency. The Journal 89. Picard C, Casanova JL, Puel A. Infectious
JCI60893 of Clinical Investigation 2011; 121: diseases in patients with IRAK-4, MyD88,
71. Ishikawa H, Ma Z, Barber GN. STING reg- 4889–4902. DOI: 10.1172/JCI59259 NEMO, or IkappaBalpha deficiency. Clini-
ulates intracellular DNA-mediated, type I 81. Herman M, Ciancanelli M, Ou YH, et al. cal Microbiology Reviews 2011; 24: 490–497.
interferon-dependent innate immunity. Heterozygous TBK1 mutations impair DOI: 10.1128/CMR.00001-11
Nature 2009; 461: 788–792. DOI: 10.1038/ TLR3 immunity and underlie herpes 90. Dupuis S, Jouanguy E, Al-Hajjar S, et al.
nature08476 simplex encephalitis of childhood. The Impaired response to interferon-alpha/
72. Vilela MC, Lima GK, Rodrigues DH, et al. Journal of Experimental Medicine 2012; 209: beta and lethal viral disease in human
TNFR1 plays a critical role in the control 1567–1582. DOI: 10.1084/jem.20111316 STAT1 deficiency. Nature Genetics 2003;
of severe HSV-1 encephalitis. Neuroscience 82. Helgason E, Phung QT, Dueber EC. 33: 388–391. DOI: 10.1038/ng1097
Letters 2010; 479: 58–62. DOI: 10.1016/j. Recent insights into the complexity of 91. Chapgier A, Wynn RF, Jouanguy E, et al.
neulet.2010.05.028 Tank-binding kinase 1 signaling net- Human complete Stat-1 deficiency is asso-
73. Sergerie Y, Rivest S, Boivin G. Tumor ne- works: the emerging role of cellular local- ciated with defective type I and II IFN re-
crosis factor-alpha and interleukin-1 beta ization in the activation and substrate sponses in vitro but immunity to some
play a critical role in the resistance against specificity of TBK1. FEBS Letters 2013; low virulence viruses in vivo. Journal of Im-
lethal herpes simplex virus encephalitis. 587: 1230–1237. DOI: 10.1016/j.febslet. munology 2006; 176: 5078–5083.
The Journal of Infectious Diseases 2007; 196: 2013.01.059 92. Boisson-Dupuis S, Kong XF, Okada S, et al.
853–860. 83. Lafaille FG, Pessach IM, Zhang SY, et al. Inborn errors of human STAT1: allelic
74. Casrouge A, Zhang SY, Eidenschenk C, et al. Impaired intrinsic immunity to HSV-1 in heterogeneity governs the diversity of im-
Herpes simplex virus encephalitis in human human iPSC-derived TLR3-deficient CNS munological and infectious phenotypes.

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv
Innate immune response to herpes simplex virus 319

Current Opinion in Immunology 2012; 24: 100. Lizarraga KJ, Alexandre LC, Ramos- 106. Boivin N, Menasria R, Piret J, Rivest S,
364–378. DOI: 10.1016/j.coi.2012.04.011 Estebanez C, Merenda A. Are steroids a Boivin G. The combination of valacyclovir
93. von Bernuth H, Picard C, Jin Z, et al. beneficial adjunctive therapy in the immu- with an anti-TNF alpha antibody increases
Pyogenic bacterial infections in humans nosuppressed patient with herpes simplex survival rate compared to antiviral ther-
with MyD88 deficiency. Science 2008; 321: virus encephalitis? Case Reports Neurology apy alone in a murine model of herpes
691–696. DOI: 10.1126/science.1158298 2013; 5: 52–55. DOI: 10.1159/000350572 simplex virus encephalitis. Antiviral Re-
94. Picard C, Puel A, Bonnet M, et al. Pyogenic 101. Kamei S, Sekizawa T, Shiota H, et al. Eval- search 2013; 100: 649–653.
bacterial infections in humans with IRAK- uation of combination therapy using 107. Rosler A, Pohl M, Braune HJ, Oertel WH,
4 deficiency. Science 2003; 299: 2076–2079. aciclovir and corticosteroid in adult pa- Gemsa D, Sprenger H. Time course of
DOI: 10.1126/science.1081902 tients with herpes simplex virus encephali- chemokines in the cerebrospinal fluid and
95. Ehrchen J, Steinmuller L, Barczyk K, et al. tis. Journal of Neurology, Neurosurgery, and serum during herpes simplex type 1 en-
Glucocorticoids induce differentiation of Psychiatry 2005; 76: 1544–1549. DOI: cephalitis. Journal of Neurological Sciences
a specifically activated, anti-inflammatory 10.1136/jnnp.2004.049676 1998; 157: 82–89.
subtype of human monocytes. Blood 2007; 102. Ramos-Estebanez C, Lizarraga KJ, 108. Kamei S, Taira N, Ishihara M, et al. Prog-
109: 1265–1274. DOI: 10.1182/blood-2006- Merenda A. A systematic review on the nostic value of cerebrospinal fluid cyto-
02-001115 role of adjunctive corticosteroids in herpes kine changes in herpes simplex virus
96. Varga G, Ehrchen J, Tsianakas A, et al. Glu- simplex virus encephalitis: is timing criti- encephalitis. Cytokine 2009; 46: 187–193.
cocorticoids induce an activated, anti- cal for safety and efficacy? Antiviral Ther- DOI: 10.1016/j.cyto.2009.01.004
inflammatory monocyte subset in mice apy 2014; 19: 133–139. DOI: 10.3851/ 109. Zhang SY, Abel L, Casanova JL. Mendelian
that resembles myeloid-derived suppres- IMP2683 predisposition to herpes simplex encephali-
sor cells. Journal of Leukocyte Biology 2008; 103. Martinez-Torres F, Menon S, Pritsch M, et al. tis. Handbook of Clinical Neurology 2013; 112:
84: 644–650. DOI: 10.1189/jlb.1107768 Protocol for German trial of Acyclovir and 1091–1097. DOI: 10.1016/B978-0-444-
97. Nakano A, Yamasaki R, Miyazaki S, corticosteroids in Herpes-simplex-virus- 52910-7.00027-1
Horiuchi N, Kunishige M, Mitsui T. Bene- encephalitis (GACHE): a multicenter, multi- 110. Jouanguy E, Zhang SY, Chapgier A, et al.
ficial effect of steroid pulse therapy on national, randomized, double-blind, Human primary immunodeficiencies of
acute viral encephalitis. European Neurol- placebo-controlled German, Austrian and type I interferons. Biochimie 2007; 89:
ogy 2003; 50: 225–229. DOI: 73864 Dutch trial [ISRCTN45122933]. BMC Neurol- 878–883. DOI: 10.1016/j.biochi.2007.04.016
98. Musallam B, Matoth I, Wolf DG, ogy 2008; 8: 40. DOI: 10.1186/1471-2377-8-40 111. Wintergerst U, Gangemi JD, Whitley RJ,
Engelhard D, Averbuch D. Steroids for de- 104. Boivin N, Sergerie Y, Rivest S, Boivin G. Chatterjee S, Kern ER. Effect of recom-
teriorating herpes simplex virus encephali- Effect of pretreatment with toll-like binant human interferon alpha B/D
tis. Pediatric Neurology 2007; 37: 229–232. receptor agonists in a mouse model of (rHu-IFN-alpha B/D) in combination with
DOI: 10.1016/j.pediatrneu-rol.2007.05.007 herpes simplex virus type 1 encephalitis. acyclovir in experimental HSV-1 encepha-
99. Mesker AJ, Bon GG, de Gans J, de Kruijk The Journal of Infectious Diseases 2008; litis. Antiviral Research 1999; 44: 75–78.
JR. Case report: a pregnant woman with 198: 664–672. 112. Sancho-Shimizu V, Perez de Diego R,
herpes simplex encephalitis successfully 105. Boivin N, Menasria R, Piret J, Boivin G. Jouanguy E, Zhang SY, Casanova JL. In-
treated with dexamethasone. European Modulation of TLR9 response in a mouse born errors of anti-viral interferon immu-
Journal of Obstetrics, Gynecology, and Repro- model of herpes simplex virus encephali- nity in humans. Current Opinion Virology
ductive Biology 2011; 154: 231–232. DOI: tis. Antiviral Research 2012; 96: 414–421. 2011; 1: 487–496. DOI: 10.1016/j.coviro.
10.1016/j.ejogrb.2010.10.014 DOI: 10.1016/j.antiviral.2012.09.022 2011.10.016

Copyright © 2015 John Wiley & Sons, Ltd. Rev. Med. Virol. 2015; 25: 300–319.
DOI: 10.1002/rmv

Vous aimerez peut-être aussi