http://www.ncku.edu.tw/~biotech/teacher/Jiann-Ruey Hong/18.pdf

publicité
Available online at www.sciencedirect.com
Molecular Immunology 45 (2008) 1146–1152
The interferon response is involved in nervous necrosis virus acute and
persistent infection in zebrafish infection model
Ming-Wei Lu a , Yung-Mei Chao a , Tz-Chun Guo b , Nina Santi b , Øystein Evensen b ,
Siti Khadijah Kasani a , Jiann-Ruey Hong c , Jen-Leih Wu a,∗
a
Laboratory of Marine Molecular Biology and Biotechnology, Institute of Cellular & Organismic Biology, Academia Sinica, Nankang, Taipei 115, Taiwan
b Department of Basic Sciences and Aquatic Medicine, Norwegian School of Veterinary Science, Oslo, Norway
c Laboratory of Molecular Virology and Biotechnology, Institute of Biotechnology, National Cheng Kung University, Tainan, Taiwan
Received 29 June 2007; received in revised form 12 July 2007; accepted 18 July 2007
Available online 28 August 2007
Abstract
Betanodavirus, a small positive-sense bipartite RNA virus notoriously affecting marine aquaculture worldwide has been extensively studied in
vitro. However, impending studies in elucidating virus–host interactions have been limiting due to the lack of appropriate animal disease models.
Therefore, in this study, we have attempted to successfully establish NNV infection in zebrafish (Danio rerio) showing typical NNV symptoms and
which could potentially serve as an in vivo model for studying virus pathogenesis. Zebrafish being already a powerful research tool in developmental
biology and having its genome completely sequenced by the end of 2007 would expedite NNV research. We have observed viral titers peaked
at 3 days post-infection and histological study showing lesions in brain tissues similar to natural host infection. Further, we used this infection
model to study the acute and persistence infection during NNV infection. Interestingly, RT-PCR and immunoblotting assays revealed that the acute
infection in larvae and juveniles is largely due to inactive interferon response as opposed to activated innate immune response during persistent
infection in adult stage. This study is the first to demonstrate NNV infection of zebrafish, which could serve as a potential animal model to study
virus pathogenesis and neuron degeneration research.
© 2007 Elsevier Ltd. All rights reserved.
Keywords: Betanodavirus; Zebrafish; Animal model; Interferon; Persistent infection
1. Introduction
Nervous necrosis virus (NNV) is an important fish pathogen
belonging to the virus family Nodaviridae that targets nervous
tissues primarily brain and retina. The manifestation of NNV for
example, Piscine nodaviruses (betanodaviruses), the causative
agents of viral nervous necrosis (VNN) or viral encephalopathy
and retinopathy (VER) (Munday et al., 2002) in a wide range
of host fish species, has resulted in major economic losses for
the marine aquaculture industry. Nodaviruses are small, nonenveloped and isometric particles containing a bipartite genome
of two positive-sense RNA molecules; RNA1 encoding the RNA
replicase and RNA2 encoding the capsid protein precursor,
which are capped but not polyadenylated (Delsert et al., 1997;
Lin et al., 2001).
∗
Corresponding author. Tel.: +886 2 27899500; fax: +886 2 27858059.
E-mail address: [email protected] (J.-L. Wu).
0161-5890/$ – see front matter © 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.molimm.2007.07.018
Extensive studies have revealed abnormal swimming behavior and sight defect of NNV infected fish. In addition,
histological examination of tissues from the central nervous
system and the retina of infected fish often reveals areas of conspicuous tissue vacuolation and necrosis (Barke et al., 2002;
Dannevig et al., 2000; Johansen et al., 2004). However, the
pathogenesis of NNV infection in which acute infection in
larvae and juvenile stages caused mass mortality while persistently infecting adult remains poorly understood. To this end,
we sought to use zebrafish as a potential NNV disease model
to provide insights into NNV pathogenesis and its host immune
response.
At present, zebrafish is rapidly becoming a valuable molecular genetics model in understanding vertebrate organogenesis
and disease development (Glass and Dahm, 2004; Yee and Pack,
2005). To date, several viruses are known to infect zebrafish,
such as spring viraemia of carp virus (SVCV), a member of
the Rhabdoviridae, that causes significant mortality in common
carp (Cyprinus carpio) (Sanders et al., 2003) and snakehead
M.-W. Lu et al. / Molecular Immunology 45 (2008) 1146–1152
rhabdovirus (SHRV) which was shown to cause mortalities
exceeding 40% in zebrafish (Phelan et al., 2005).
In view of the differential effect of viral infection, we looked
into the zebrafish interferon response, in which the zebrafish
interferon gene (zIFN) has been recently identified as to having
anti-virus function (Altmann et al., 2003) and may contribute to
both induction and regulation of the innate and adaptive immunity. Downstream interferon activated Mx gene has also been
identified in zebrafish, grouper, salmon, trout, and halibut upon
infection with aquatic viruses (Chen et al., 2006; Kibenge et
al., 2005; Lin et al., 2005; McBeath et al., 2006) suggesting
the importance of the interferon regulatory pathway including
RNA-activated protein kinase (PKR) and the 2-5A proteins during viral infection. In our present study, we have observed an
elevated interferon expression in infected adult zebrafish relative to infected larvae resulting in higher rate of mortality in the
latter. This may indicate that the susceptibility to NNV infection
is dependent on the enhancement of IFN system. However, the
mechanism in which interferon response is activated upon NNV
viral infection needs to be further elucidated.
2. Materials and methods
2.1. Fish
AB (−) inbred strain zebrafish were obtained at 2 months
stage from Institute of Cellular & Organismic Biology,
Academia Sinica, Taiwan. The zebrafish were handled according
to Institutional Animal Care and Use Committee guidelines.
2.2. Cell lines and virus
The SSN-1 derivate cell line E-11 (Iwamoto et al., 2000)
was used to isolate and titrate NNV. Cells were propagated and
maintained at 28 ◦ C in L-15 medium (GIBCO) supplemented
with 10% fetal bovine serum (Sigma), 100 I.U./ml penicillin,
and 0.1 mg/ml streptomycin.
The nodavirus used in this study was isolated from malabaricus grouper (Epinephelus malabaricus) juveniles during
an outbreak of VER at a fish farm in southern Taiwan. Sequence
analysis of partial viral genome (data not shown) revealed that
it shares more than 95% homology with other betanodaviruses
isolated from fish in Taiwan (Chi et al., 2003). E-11 cells were
inoculated with a fourth passage virus supernatant and incubated
at 28 ◦ C. When a cytopathic effect (CPE) was observed on 3 days
post-infection (p.i.), the culture supernatant was harvested, clarified by centrifugation (3000 × g for 5 min) and stored at 4 ◦ C
until challenge. The virus titer in the supernatant was determined
by using an infectivity assay and was calculated to be 1 × 108
TCID50 ml−1 .
2.3. Fish challenge
Two groups of 60 adult zebrafish, one infected and one mockinfected, were used in the challenge experiment. The fish were
held in 5 liter tanks at a water temperature of 28 ◦ C. Challenge was performed by intraperitoneal (i.p.) injection of 1 × 105
1147
TCID50 ml−1 of NNV in 20 ␮l. The mock-infected group was
injected with PBS. After challenge, the fish were monitored daily
over a 14-day period for signs of disease and mortality. Six fish
in each group were sampled each day and tissue samples from
different organs were isolated for virus detection by RT-PCR and
infectivity assay (TCID50 ). In a separate experiment, fish were
maintained and challenged similarly, but three fish from each
group were sampled daily for histology and immunohistochemistry. Some fish in parallel were also collected for confirmation
of infection by RT-PCR and virus re-isolation.
2.4. Histology and immunohistochemistry (IHC)
The heads of fish used for histology were fixed for at
least 24 h in neutral phosphate-buffered 10% formalin and
embedded in paraffin wax. Five micrometers brain sections
were stained with haematoxylin and eosin (H&E) whereas the
other parallel sections were processed for immunohistochemical
detection of nodavirus protein using anti-NNV antiserum and
a streptavidin-biotin-alkaline-phosphatase complex antibody
detection technique (Chemicon IHC Select Kit). Mock-infected
fish served as negative controls for the experiment.
2.5. RNA isolation and RT-PCR
RNA was extracted using TRIZOL reagent (Life Technologies) according to the manufacturer’s instructions. The RT-PCR
was performed using a Superscript III one-step RT-PCR system (Invitrogen) and each reaction includes 2 ␮l total RNA
(20 ng) and a primer set. The primers were designed to amplify
the variable T4 region of the coat protein gene (Nishizawa
et al., 1994), Mx (sense 5 -AGTACCGGGGAAGAGAGCTA-3 antisense 5 -AAGGTGGCATGATTGT CTGT-3 ), IFN
(sense 5 -ATGAGAACTCAAATGTGGAC-3 antisense 5 -TTACA CTCGAGGATTGAC-3 ), and ␤-actin (sense 5 -ATGGATGAGGAAATCGCTG-3 antisense 5 -ATGCCAACCATCACTCCCTG-3 ) gene.
2.6. Microinjection of virus into zebrafish larvae
The NNV titer for microinjection was adjusted to 1 × 103
TCID50 ml−1 . Injections were conducted by PLI-100 airinjection apparatus (Medical System Co.). The amount
(approximately 0.1 ␮l) injected into larvae was estimated by
visualizing the injection volume. After injection, the larvae were
incubated in the Petri-dish at 28 ◦ C.
2.7. Immunoblot
Fish samples were collected, homogenized, and solubilized
in disruption buffer containing 50 mM Tris–HCl (pH 7.0), 5%
2-mercaptoethanol, 2% sodium dodecyl sulfate (SDS), and
2.75% sucrose. Samples were then sonicated, boiled, subjected to electrophoresis on denaturing 12% polyacrylamide
gels, transferred to nitrocellulose membranes, blocked with
5% nonfat milk, and reacted with antibodies against phospho-
1148
M.-W. Lu et al. / Molecular Immunology 45 (2008) 1146–1152
rylated eIF2␣ protein (Cell Signaling). The membranes were
rinsed in phosphate-buffered saline (PBS) and reacted with antirabbit immunoglobulin conjugated to horseradish peroxidase
and developed with an enhanced chemiluminescence Western
blot detection system kit (Amersham Pharmacia).
2.8. Quantitative RT-PCR for NNV RNA detection
A 100 ng aliquot of total RNA was used to quantify NNV-specific RNA levels using an ABI Prism 7000
sequence detector (Applied Biosystems). Real-time reverse
transcription-polymerase chain reaction (RT-PCR) amplifications were performed by the High Capacity cDNA Archive kit
(ABI, Applied Biosystems) and primer specific for NNV 5 CGAGTCAACACGGGTGAAGA-3 . RT reactions were incubated for 10 min at 25 ◦ C, 2 h at 37 ◦ C and cooling to 4 ◦ C
for 5 min. The quantitative RCR protocol provided by the ABI
real-time instruments by using Platinum® SYBR® GreenqPCR
SuperMix-UDG, and the primers specific for NNV were 5
NTR: 5 -GCCCCTGATGGAGCAGTCT-3 (sense 10 ␮M); 5 AGCACGGTCAACATCTCCAGTT-3 (antisense, 10 ␮M); 45
cycles of PCR were performed with cycling conditions of 3 s at
95 ◦ C, 30 s at 60 ◦ C. Standard curve was generated using vector
(pDA8; Lu et al., 2003) bearing NNV s RNA2 gene of variable
known concentration. The real-time PCR signals were analyzed
in a multiplex format using SDS software (Version 1.7; Applied
Biosystems).
3. Results
3.1. NNV replication in different organs
Six zebrafish were sacrificed each day over 14 days and
pooled tissues from different organs were tested by RT-PCR and
TCID50 titration for NNV. Brain, eye, heart, liver and gut were
tested positive for NNV, while muscle was negative for NNV
by RT-PCR from 3 days p.i. (Fig. 1a). The yields of RT-PCR
product from brain, gut and eye were greater compared to heart
and liver. It was observed to being most abundant in the brain
which is the major target organ for NNV propagation.
Virus titers were detected in brain, eye, and gut at 3.6 × 105 ,
3.6 × 102 , and 3.6 × 102 TCID50 g−1 , respectively, but not in
the heart, liver or muscle (Fig. 1b). Detection of NNV in gut,
heart and liver from RT-PCR was probably due to excessive virus
introduced during inoculation but had decreased significantly on
Fig. 2. NNV replication in infected zebrafish brains at different times postinfection. (a) Agarose gel showing RT-PCR product specific for the NNV
coat protein. (b) Virus titers in brains from NNV infected and mock-infected
zebrafish.
2 days p.i. (data not shown). In addition, TCID50 assay indicated
no considerable amount of virus was in the heart and liver while
the low copy number of NNV could still be within the detection
of RT-PCR.
NNV titers increased in brain tissue (Fig. 2a) from 1.4 × 103
TCID50 g−1 on 1 day p.i. to peak levels of 3.6 × 105 TCID50 g−1
after 3 days p.i. (Fig. 2b). The finding of NNV in the target
organs brain and eye implies that the virus can overcome the
blood–brain barrier after i.p. injection.
3.2. Nodavirus detection by histopathology
Histopathological lesions involving vacuolated cells were
observed in brain tissue from 4 days p.i. onwards. Using
immunohistochemistry, positive NNV-specific signal was
observed in brain tissue surrounding these lesions (Fig. 3).
These results demonstrated that the nodavirus not only replicates
but also cause pathological lesions in brain tissue of infected
zebrafish.
3.3. Re-isolation of NNV from zebrafish in E-11 cells
Brain tissue from zebrafish that tested positive by RT-PCR
was used to re-isolate NNV in E-11 cells. Development of CPE
manifest the cell death and detachment was observed in the E11 cells 3–4 days after inoculation of the tissue homogenate
(Fig. 4). The presence of NNV in cell cultures displaying CPE
was confirmed by using RT-PCR and sequence analysis of PCR
product.
3.4. Nodavirus infection in zebrafish larvae
Fig. 1. Tissue distribution of virus replication in infected zebrafish. (a) Agarose
gel showing RT-PCR products specific for the NNV coat protein. (b) Virus titers
in different organs from fish exposed to NNV via i.p. injection. B, brain; G, gut;
H, heart; E, eye; L, liver; M, muscle; ND, not-detected.
The NNV infection can spread both by horizontal and vertical
transmission (Peducasse et al., 1999). Hence, we simulated vertical transmission of NNV in zebrafish by injecting the virus to
M.-W. Lu et al. / Molecular Immunology 45 (2008) 1146–1152
1149
Fig. 3. Detection of NNV antigen in the brain of infected zebrafish by immunohistochemistry. (a) Negative control obtained from the brain of a mock-infected
zebrafish (scale bar = 100 ␮m); (b) brain of an infected zebrafish on day 4 p.i., showing immuno-positive cells associated with brain tissue vacuolization (arrow)
(scale bar = 100 ␮m).
zebrafish larvae. NNV was propagated in E-11 cells and injected
to zebrafish larvae at 1 × 103 TCID50 ml−1 . The result showed
98% mortality at 24 h p.i. compared to 24% mortality in the
mock injected group (Fig. 5a). Quantitative RT-PCR confirmed
the presence of NNV in dead larvae from the NNV infected
group. Furthermore, we also compared the amounts of viral
RNA between larval and adult stage by quantitative RT-PCR.
The amount of viral RNA2 in larval stage was 0.06 ng, which was
much higher than adult brain sample (1.33 × 10−6 ng) (Fig. 5b).
lyzed for innate immune response. The IFN-␣ gene was not
detectable in the larvae stage but expressed normally in the adult
stage (Fig. 6a). The differential expression pattern of activated
Mx gene and phosphorylated eIF2␣ gene, which are downstream
genes of interferon pathway, between larval and adult stage at
16 h p.i. (Fig. 6a and b) indicated the importance of interferon
response during NNV pathogenesis. We further confirmed the
observations by introducing interferon in the early stages of
zebrafish larvae followed by NNV infection. The dosage of uni-
3.5. The interferon response in larval and adult stage after
NNV infection
The mortality of larvae is very distinct from adult fish after
NNV infection. In larval stage, the mortality is higher than 95%
amongst all of marine species which could be infected by NNV
(Barke et al., 2002). In this study, we collected the larval and
adult stage samples which have been infected by NNV and ana-
Fig. 4. Re-isolation of NNV from infected zebrafish. (a) Normal E-11 cell monolayer was grown in L-15 with 10% fetal bovine serum for 3 days. Cells were
maintained at 28 ◦ C. (b) Monolayer of E-11 cells showing CPE upon infection
with NNV isolated from zebrafish for 3 days.
Fig. 5. Cumulative mortality for the three groups of zebrafish larvae and the
differential amount of viral RNA between zebrafish larval and adult stage.
(a) Control group was untreated. Mock-injection group was injected with L15 medium. NNV injected group was injected with supernatant from a NNV
infected cell culture at a dosage of 1 × 103 TCID50 ml−1 . (b) Quantitative RTPCR of the NNV RNA2 was performed in larval and adult stage after NNV
infection.
1150
M.-W. Lu et al. / Molecular Immunology 45 (2008) 1146–1152
Fig. 6. Interferon related gene expression in larval and adult stage. (a) Semiquantitative RT-PCR analysis to detect the temporal expression of interferon
and Mx gene in larval and adult stage after NNV infection. (b) Immuno-blot of
the eIF2␣ phosphorylation protein was performed in larval and adult stage after
NNV infection.
versal IFN-␣2a (Roche) that was introduced by microinjection
was 1000U. We designed four experimental groups, the first
group was the positive control for NNV infection, the second
group was co-injected with NNV and interferon, the third group
was pretreated with interferon for 48 h before inoculated with
NNV, and the last group was mock inoculated with PBS. The
group that was co-injected with interferon and NNV displayed
delayed mortality rate, however, after 96 h the mortality still
reached 100%, similar to positive control at 72 h. In addition, the
pretreatment with interferon for 48 h could rescue the survival
rate of NNV infected group from 0 to 76% which is similar to
mock-infected group (Fig. 7a). The viral RNA was also beyond
detectable amounts as shown by RT-PCR (Fig. 7b).
4. Discussion
Betanodavirus, one of the most important causal agent of viral
nervous necrosis resulting in economical losses to the aquaculture industry worldwide. In recent years, betanodaviruses have
been detected in a variety of fish species both in temperate and
tropical areas, as well as in fresh water species (Azad et al.,
2005; Hegde et al., 2003; Johansen et al., 2003, 2004; Sommer
et al., 2004; Starkey et al., 2001; Ucko et al., 2004). Experimental infection models have been established in different host
species, such as halibut, sea bass and wolfish (Grove et al.,
2003; Peducasse et al., 1999; Sommer et al., 2004) but unsuccessful in the mouse system (Banu and Nakai, 2004). In this
study, we demonstrated that zebrafish are also susceptible to
NNV infection. This is also the first NNV infection model in
Fig. 7. The effects of interferon in zebrafish larval stage after NNV infection. (a)
Survival rate of groups with differential interferon treatment. The arrow indicates
the injection time. (b) Semi-quantitative of the viral RNA amount of differential
interferon treatment.
zebrafish in which, its genome project is near completion, thus
will accelerate the mechanistic study of NNV pathogenesis. RTPCR showed that viral RNA was most abundant in the brain
followed by eye and gut but was less in liver and heart. Infectivity assay also confirmed the presence of virus in the brain,
eye and gut. The high virus titer in the gut compared to liver
and heart, which do not characteristically support replication of
NNV, may imply a role of the gut in the disease pathogenesis and
as a possible virus entry route via feed, such as contaminated
plankton. In persistently infected juvenile Atlantic halibut, viral
RNA can also be detected in abdominal organs by RT-PCR, but
it is not clear which organ the virus resides in (Johans et al.,
2002).
Herpes simplex virus, and rabies virus are spread by axonal
transport from the skin or muscle to the corresponding dorsal
root ganglion or anterior horn cells and then to populations of
neurons throughout the CNS. In contrast, arboviruses (mainly
togaviruses, flaviviruses, and bunyaviruses) spread to the brain
via the blood. In NNV-infected zebrafish, high titers of virus
occurred in the gut and eye as well as brain. These results indicated a possible spread of virus from the central lacteal via blood
into the brain and eye. However, this hypothesis needs more
detailed investigation.
Beginning from 14 days post-NNV-infection, brain tissue was
collected daily from zebrafish that was in persistent infection.
From 3 days to 14 days p.i., no apparent symptom was observed
in the adult zebrafish. Hand in zebrafish larvae, however, in the
zebrafish larvae, mortality following injection of a brain tissue homogenate from experimentally infected zebrafish reached
M.-W. Lu et al. / Molecular Immunology 45 (2008) 1146–1152
1151
Fig. 8. The molecular mechanism of NNV infection in acute and persistence infection. The interferon was activated after NNV infection in adult stage, however, no
activation observed in larval stage. The downstream antiviral responses of interferon pathway in the larval stage were also not activated leading to acute infection and
mass mortality. In contrast with larval stage, the adult stage interferon related antiviral genes were expressed normally and repressed virus replication thus leading
to persistent infection. The 2-5A system has not been reported in fish.
98% within 24 h p.i. These results mirror infection in natural
hosts of NNV in which mortality primarily occurs in larvae
and juveniles. There are some pathogens infecting human also
shown to have the same phenomenon, such as alphavirus, JEV,
and enterovirus. Therefore, it is vital to elucidate the mechanism
that discriminate acute infection in the juvenile from persistence
or non-infective stage in the adult.
The acute and persistent infection in the in vivo host is highly
related to immunity, signal transduction and viral replication.
The viral persistence involves several mechanisms especially
host innate/adaptive immune response. In HCV, control of acute
primary viral replication is associated with expansion of antiviral CD4+ (helper) and CD8+ (cytotoxic) T cells (Bowen and
Walker, 2005). The innate immune response also plays a very
important role in acute persistent infection. In a recent paper,
the in vitro results indicated that Mx gene expression may play
an important role in NNV persistence in BB cell (Wu and Chi,
2006). However, the involvement of interferon response in the
age-dependent viral disease remains unclear. We tried to use
zebrafish as an infection model to understand the relationship
between interferon response and NNV infection. According to
our results, the virus titer and mortality rate was significantly
different between larval and adult stage (Fig. 5). This phenomenon was similar to the acute and persistence infection in
NNV pathogenesis in nature. We first examined the interferon
related expression profile of NNV infected zebrafish in larval
and adult stage. We found that interferon related gene could
not be activated by NNV in larval stage. The similar observation was also reported in NNV infected adult sea bass and
sea bream recently (Poisa-Beiro et al., 2008). That might be
the reason for causing acute infection in larval stage but persistent infection in adult stage. The same phenomena was also
observed in the mice system, that is mice lacking functional
IFN-I receptors are extremely susceptible to a wide variety of
viral infections (Alsharifi et al., 2006). In the gain of function
experiment, we observed the inhibition of virus replication after
interferon treatment. The reaction time of interferon treatment
was also recorded to be 48 h. In conclusion, these results indicated that we could generate the possible mechanism that leads
to the difference between acute and persistence infection in NNV
pathogenesis (Fig. 8).
The zebrafish infection model will provide a convenient system to study NNV pathogenesis and its potential drug-mediated
intervention. Other important advantages of zebrafish are the
availability of developmental and genomic databases. Additionally, their immune systems are well characterized compared to
other fish species. By detecting the IFN-related gene expression
in our zebrafish infection model, we first suggest a hypothesis
that the antagonistic effect seen in acute and persistent infection
in NNV is age dependent. In addition, since the innate immunity
is conserved amongst fly, plant and mammals, the finding and
discovery in this study could significantly increase our understanding of innate immunity in human thereby lead to medical
application. This model can be further used for virus pathogenesis studies, such as apoptosis regulation, signal transduction and
immune response.
Acknowledgements
This work was funded by the Taiwanese Council of Agriculture (95AS-5.1.6-FA-F1), National Science Council of Taiwan
(NSC 95-2313-B-001-004) and Institute of Cellular and Organismic Biology, Academia Sinica, Taiwan.
References
Alsharifi, M., Regner, M., Blanden, R., Lobigs, M., Lee, E., Koskinen, A., Mullbacher, A., 2006. Exhaustion of type I interferon response following an acute
viral infection. J. Immunol. 177, 3235–3241.
1152
M.-W. Lu et al. / Molecular Immunology 45 (2008) 1146–1152
Altmann, S.M., Mellon, M.T., Distel, D.L., Kim, C.H., 2003. Molecular and
functional analysis of an interferon gene from the zebrafish, Danio rerio. J.
Virol. 77, 1992–2002.
Azad, I.S., Shekhar, M.S., Thirunavukkarasu, A.R., Poornima, M., Kailasam,
M., Rajan, J.J., Ali, S.A., Abraham, M., Ravichandran, P., 2005. Nodavirus
infection causes mortalities in hatchery produced larvae of Lates calcarifer:
first report from India. Dis. Aquat. Organ. 63, 113–118.
Banu, G.R., Nakai, T., 2004. Inoculation of BALB/c mice with fish-pathogenic
nodaviruses. J. Comp. Pathol. 130, 202–204.
Barke, D.E., MacKinnon, A.M., Boston, L., Burt, M.D., Cone, D.K., Speare,
D.J., Griffiths, S., Cook, M., Ritchie, R., Olivier, G., 2002. First report of
piscine nodavirus infecting wild winter flounder Pleuronectes americanus
in Passamaquoddy Bay, New Brunswick, Canada. Dis. Aquat. Organ. 49,
99–105.
Bowen, D.G., Walker, C.M., 2005. Adaptive immune responses in acute and
chronic hepatitis C virus infection. Nature 436, 946–952.
Chen, Y.M., Su, Y.L., Lin, J.H., Yang, H.L., Chen, T.Y., 2006. Cloning of an
orange-spotted grouper (Epinephelus coioides) Mx cDNA and characterisation of its expression in response to nodavirus. Fish Shellfish Immunol. 20,
58–71.
Chi, S.C., Shieh, J.R., Lin, S.J., 2003. Genetic and antigenic analysis of betanodaviruses isolated from aquatic organisms in Taiwan. Dis. Aquat. Organ.
55, 221–228.
Dannevig, B.H., Nilsen, R., Modahl, I., Jankowska, M., Taksdal, T., Press, C.M.,
2000. Isolation in cell culture of nodavirus from farmed Atlantic halibut
Hippoglossus hippoglossus in Norway. Dis. Aquat. Organ. 43, 183–189.
Delsert, C., Morin, N., Comps, M., 1997. A fish encephalitis virus that differs
from other nodaviruses by its capsid protein processing. Arch. Virol. 142,
2359–2371.
Glass, A.S., Dahm, R., 2004. The zebrafish as a model organism for eye development. Ophthalmic Res. 36, 4–24.
Grove, S., Johansen, R., Dannevig, B.H., Reitan, L.J., Ranheim, T., 2003.
Experimental infection of Atlantic halibut Hippoglossus hippoglossus with
nodavirus: tissue distribution and immune response. Dis. Aquat. Organ. 53,
211–221.
Hegde, A., The, H.C., Lam, T.J., Sin, Y.M., 2003. Nodavirus infection in
freshwater ornamental fish, guppy, Poicelia reticulata—comparative characterization and pathogenicity studies. Arch. Virol. 148, 575–586.
Iwamoto, T., Nakai, T., Mori, K., Arimoto, M., Furusawa, I., 2000. Cloning
of the fish cell line SSN-1 for piscine nodaviruses. Dis. Aquat. Organ. 43,
81–89.
Johans, R., Ranheim, T., Hansen, M.K., Taksdal, T., Totland, G.K., 2002.
Pathological changes in juvenile Atlantic halibut Hippoglossus hippoglossus
persistently infected with nodavirus. Dis. Aquat. Organ. 50, 161–169.
Johansen, R., Amundsen, M., Dannevig, B.H., Sommer, A.I., 2003. Acute and
persistent experimental nodavirus infection in spotted wolffish Anarhichas
minor. Dis. Aquat. Organ. 57, 35–41.
Johansen, R., Sommerset, I., Torud, B., Korsnes, K., Hjortaas, M.J., Nilsen,
F., Nerland, A.H., Dannevig, B.H., 2004. Characterization of nodavirus
and viral encephalopathy and retinopathy in farmed turbot, Scophthalmus
maximus (L.). J. Fish Dis. 27, 591–601.
Kibenge, M.J., Munir, K., Kibenge, F.S., 2005. Constitutive expression of
Atlantic salmon Mx1 protein in CHSE-214 cells confers resistance to infectious salmon anaemia virus. Virol. J. 2, 75.
Lin, C.S., Lu, M.W., Tang, L., Liu, W., Chao, C.B., Lin, C.J., Krishna, N.K.,
Johnson, J.E., Schneemann, A., 2001. Characterization of virus-like particles
assembled in a recombinant baculovirus system expressing the capsid protein
of a fish nodavirus. Virology 290, 50–58.
Lin, O.E., Ohira, T., Hirono, I., Saito-Taki, T., Aoki, T., 2005. Immunoanalysis of antiviral Mx protein expression in Japanese flounder (Paralichthys
olivaceus) cells. Dev. Comp. Immunol. 29, 443–455.
Lu, M.W., Liu, W., Lin, C.S., 2003. Infection competition against grouper nervous necrosis virus by virus-like particles produced in Escherichia coli. J.
Gen. Virol. 84, 1577–1582.
McBeath, A.J., Snow, M., Secombes, C.J., Ellis, A.E., Collet, B., 2006.
Expression kinetics of interferon and interferon-induced genes in Atlantic
salmon (Salmo salar) following infection with infectious pancreatic necrosis virus and infectious salmon anaemia virus. Fish Shellfish Immunol. 22,
230–241.
Munday, B.L., Kwang, J., Moody, N., 2002. Betanodavirus infections of teleost
fish: a review. J. Fish Dis. 25, 127–142.
Nishizawa, T., Mori, K., Nakai, T., Furusawa, I., Muroga, K., 1994. Polymerase
chain reaction (PCR) amplification of RNA of striped jack nervous necrosis
virus (SJNNV). Dis. Aquat. Organ. 18, 103–107.
Peducasse, S., Castric, J., Thiery, R., Jeffroy, J., Le Ven, A., Baudin Laurencin, F.,
1999. Comparative study of viral encephalopathy and retinopathy in juvenile
sea bass Dicentrarchus labrax infected in different ways. Dis. Aquat. Organ.
36, 11–20.
Phelan, P.E., Pressley, M.E., Witten, P.E., Mellon, M.T., Blake, S., Kim, C.H.,
2005. Characterization of snakehead rhabdovirus infection in zebrafish
(Danio rerio). J. Virol. 79, 1842–1852.
Poisa-Beiro, L., Dios, S., Montes, A., Aranguren, R., Figueras, A., Novoa, B.,
2008. Nodavirus increases the expression of Mx and inflammatory cytokines
in fish brain. Mol. Immunol. 45, 218–225.
Sanders, G.E., Batts, W.N., Winton, J.R., 2003. Susceptibility of zebrafish
(Danio rerio) to a model pathogen, spring viremia of carp virus. Comp.
Med. 53, 514–521.
Sommer, A.I., Amundsen Strand, M., Rasmussen, E., Mennen, S., 2004. Susceptibility of spotted wolffish Anarhichas minor to experimental infection
with nodavirus and infectious pancreatic necrosis virus. Dis. Aquat. Organ.
59, 101–108.
Starkey, W.G., Ireland, J.H., Muir, K.F., Jenkins, M.E., Roy, W.J., Richards,
R.H., Ferguson, H.W., 2001. Nodavirus infection in Atlantic cod and Dover
sole in the UK. Vet. Rec. 149, 179–181.
Ucko, M., Colorni, A., Diamant, A., 2004. Nodavirus infections in Israeli mariculture. J. Fish Dis. 27, 459–469.
Wu, Y.C., Chi, S.C., 2006. Persistence of betanodavirus in Barramundi brain
(BB) cell line involves the induction of Interferon response. Fish Shellfish
Immunol. 21, 540–547.
Yee, N.S., Pack, M., 2005. Zebrafish as a model for pancreatic cancer research.
Methods Mol. Med. 103, 273–298.
Téléchargement