Sei sulla pagina 1di 15

Annual Review of Fish Diseases, Vol 5, pp.

209-223, 1995
Pergamon Copyright 0 1996 Elsevier Science Ltd
Printed in Great Britain. All rights reserved
0959-8030195 $29.00 + .OO

0959-8030(95)00001-1

IMMUNOLOGICAL AND PATHOLOGICAL RESPONSES


OF SALMONIDS TO INFECTIOUS PANCREATIC
NECROSIS VIRUS (IPNV)
Eileen C. Saahsiv
Department of Fisheries, Animal and Veterinary Science, The University of Rhode Island,
Kingston, Rhode Island, 02881-0804, USA

Abstract The aquatic bimavims IPNV is commonly found in association with apparently healthy,
mature salmonids. Some bimaviruses cause lethal diseases in fry; however, many of those discovered
in fish may not be pathogenic for the species from which they were isolated. More pathogenic
virus may be produced from the acinar cells of the pancreas; less pathogenic virus by skin and gut
cells. Skin infection could explain the lack of virus clearance after development of circulating
antibody; the less pathogenic form of the virus may not induce protective antibodies. True vertical
transmission to progeny fish would seem not to occur, but virus may adhere to egg cases and
spread to fry which ingest them at the time of first feeding. When viruses are found with moribund
mature fish, alternative causes of death must be considered. Atlantic salmon Salmo salar smelts,
which have high levels of virus during pre-smoking, can develop pancreatic lesions (possibly
mediated by the immune system) on transfer to salt water. IPNV infection may be linked to
immunosuppression; possible controlling genes have been found.

Keywords: Bimavirus, Persistent infection, Aquaculture, Salmo salar

INTRODUCTION

One of the important diseases of farmed salmonids is infectious pancreatic necrosis (IPN)
which is caused by a birnavirus, infectious pancreatic necrosis virus (IPNV). Multiple strains
of the virus exist that differ in virulence and in evoking variable serologic responses (1).
Although the virus can be a highly destructive pathogen of hatchery-reared salmonids, it has
been reported to be carried and possibly replicated in hosts for long periods without causing
clinical disease. Broodstock carriage has been considered a likely source of virus for the
lethal infection of progeny fish.
IPN disease happens only with the coming together of susceptible fish and virus able to
attach to, enter and replicate in cells of those fish. Salmonid defensive systems are either not
present or are overcome. Each cell infected with virus can produce up to 1000 new infectious
particles within a few hours, depending upon the temperature and cell type (2).
Numerous studies of the disease processes caused by IPNV have been conducted. This
paper will be limited to a discussion of IPNV infection of three species of salmonids: brook
trout (Salvelinus fintinalis), rainbow trout (Oncorhynchus mykiss) and Atlantic salmon (Salmo
salar). Brook trout have been found to be the most sensitive to IPN and almost all survivors
become carriers (3, 4). They are followed in sensitivity by rainbow trout, which are better
antibody producers and have a lower and more rapidly declining carrier rate, and by Atlantic
salmon which seem to be the least sensitive of the three species to the lethal effects of the
virus.
With the improvements in virus detection methods (5-8), it is apparent that IPNV, or
IPNV-like viruses, are quite ubiquitous (9-14). Most hatchery efforts to date have been
directed towards eliminating virus during juvenile development. An integral part of avoiding

209
210 E. C. Sadasiv

juvenile disease has been to select broodstock which have never encountered the virus.
Control of IPN by vaccination has not been achieved and serological identification of the
virus able to cause lethal disease has not been reliable (14). The purpose of this review is to
explore how fish can have an antiviral immune response which does not result in elimination
of virus from the fish.

THE VIRUS
IPNV is an aquatic birnavirus having at least four different proteins which are coded for by
two segments of double-stranded RNA (15, 16). It is a simple and remarkably stable virus. It
can be added to silage and pass through the digestive system of cows (17) or it can be
retained within scallops for almost a year without being degraded (18).
Pathogenicity for the host and tropism, or the likelihood that the virus will attach to and
enter particular types of cells where it can be replicated, is determined by the interaction of
viral proteins with host proteins (19, 20). The largest of the viral proteins, VP-I, is the
polymerase which is coded for by the smaller RNA segment (21). It is not associated with
virulence (19-22). The larger RNA segment encodes three proteins which are produced as a
polyprotein and cleaved to yield the major capsid protein VP-2, a non-structural protein NS or
VP-4, and an interior structural protein VP-3 (23). An exact cleavage site between VP-2 and
NS has not been determined (24). VP-2 is the outermost protein and not surprisingly is
involved in attachment to cells. Estay et al. (25) have reported that VP-2 is glycoslyated.
VP-3 is associated with VP-2, but does not contain neutralizing epitopes and thus would be a
less important target of the host’s defensive mechanisms. According to the work of Manning,
Mason, and Leong in 1990 (26) and Heppell et d. in 1993 (24), NS is likely to function as a
co-enzyme, pointing towards the importance of cellular enzymes for viral formation.
The invaded host can respond to infection with a number of defensive strategies including
neutralization by the production of specific antibodies against the virus surface. The viral
protein epitope important in its neutralization would appear to be either a conformational
arrangement of VP-2 or a sequential-conformational epitope (27-29). IPNV strains have outer
proteins differing in molecular weight between 63K and 54K. This variability in length of the
surface covering would be one way in which their ability to be neutralized would vary. The
surface pattern formed by different length VP-2s would certainly present an antigenically
altered antibody attachment surface.
In 1986, Wolski, Roberson, and Hetrick (28) reported the production of monoclonal
antibodies (MAB) to the Sp strain of IPNV Ten of the MAB were found to be broadly
reactive against representatives of the 3 major serotypes of IPNV. Two others were specific
for the Sp strain which was the immunogen. One of the specific MAB’s was shown to be
directed against the major capsid protein while the other specific one, and the broadly reacting
ones, reacted with the low molecular weight viral polypeptides. &swell-Reno and her
associates (29-34) have used a panel of MAB to 11 aquatic birnavirus epitopes to group
IPNV viral stains. Lecomte, Arella, and Bertiaume (31, 35) have described a MAB to VP-3
that can interact with the virus surface without neutralizing it and were able to show some
differences between pathogenic and non-pathogenic virus.
DNA probes have been developed to detect the virus (32-37). Probes, coupled with
polymerase chain reaction (PCR) amplification techniques, can greatly increase the sensitivity
of viral detection, but they have not performed well to date in field tests with infected fish.
Some probes can detect areas of the viral genome not concerned with producing structures
which form the virus neutralization epitope (38).
IPNV in salmonids 211

IPN virus has been found to propagate in several lines of cultured cells, generally of
epithelial or fibroblastic origin, in which cytopathic effects are produced or a persistent
infection is achieved (3). Kidney cells have been noted to be infected at similar rates (O-OS%)
as are persistently infected cultured cells (O.l-1.0%) (39). In the case of some smaller RNA
viruses such as poliomyelitis, expression of viral proteins is altered during a persistent infection
(4m2), but this has not yet been reported for birnaviruses. Kennedy and MacDonald (43)
postulated that persistent infection with IPNV could be mediated by virus particles having
altered properties, which might allow those cells to survive.
Pathogenic properties of IPNV can be changed by culture conditions; cell-culture passage
may produce a less pathogenic virus which in some cases may have an altered reaction with
antiserum (44, 45). In susceptible fish, more highly pathogenic strains (Buhl, VR299), have
been observed by Sano and associates to be produced in all tissues; whereas the avirulent
EVE (Ab strain) was produced from epidermis only (19). They also agreed that the virulent
strain becomes avirulent by passage in cell culture, but they did not suggest how these two
observations might be related. Viral attenuation can occur in a number of ways. One is by
accumulation of genetic mutations; these would not revert to the original form in a single
step. Another way is by change due to non-hereditary causes (14) such as trimming the size
of VP-2 or varying the sugar residues or their placement on the viral surface (gylcosylation).
Each of these processes could yield an altered virus which might evade the immune system
and attach to different cells. The attenuation of virus by a single passage in cell culture could
be an example of a phenotypic alteration.

SALMONID DISEASE AND VIRUS CARRIAGE

Brook trout
Yamamoto (46) reported that IPNV could be isolated from the kidneys of carrier brook
trout but their antibody level could not be correlated with virus yield. Swanson (47) reported
changes in the pathogenicity of IPNV due to its passage in cell culture. He used virus after it
had been repeatedly passed in culture and contrasted its pathologic effects with those of a
field isolate which had undergone only one culture pass. The high passage virus replicated in
the intestinal cells while the field isolate replicated in the pancreas. Brook trout infected
several months after first feeding were not found to have either virus-caused lesions or
significant growth retardation. He first reported the presence of mononuclear white blood cells
around damaged pancreatic acini. Swanson and Gillespie (48) reported that kidneys did not
show evidence of spreading foci of viral infections, although they contained macrophages
carrying cell debris and necrotic cells. Virus replicated in the intestine and was detected in the
kidney, with the viral titer peaking at 7 days post-inoculation (49). It is also replicated in the
digestive tract of yearlings without being found in the internal organs. They concluded that
virus was picked up by phagocytic cells and transported to the kidney following intraperitoneal
injection. Virus titers in blood cells were lower and variable compared to samples of kidneys,
spleens or sexual excretions (50). It is notable that replication in older fish did not involve
disease or viremia, which would release the virus to other organs. Around this same time Yu,
MacDonald, and Moore (51) reported that they could recover IPNV consistently from the
leukocytes of asymptomatic year-old brook trout.
Although birnaviruses are good antigens, an immune response does not result in the
elimination of the virus. Bootland, Stevenson, and Dobos (52) documented the puzzling
212 E. C. Sadasiv

difficulty of vaccination, showing that immunization by immersion could increase the percentage
of fry succumbing to IPN. Their work confirmed that IPNV could not be found in fertilized
eggs from brook trout carrier broodstock (53). Transmission from broodstock to progeny was
demonstrated, with progeny having low mortality, low prevalence and low titers. They
concluded that the carrier state and vertical transmission of IPNV was unpredictable. The
mechanism of transmission was not determined. IPNV in yearling brook trout induced an
asymptomatic chronic virus presence that persisted for at least 76 weeks in 95% of the fish.
They confirmed that the fish have an immune response, but that it did not result in clearance
of virus from the fish (54). The prevalence of “shedders” and their viral titers varied over
time. The infection did not affect growth rate, which argued against an ongoing infection, and
a wide variability was noted in the humoral immune responses of individual fish.
In his continuing efforts to explain long-term virus carriage, McAllister et al. (55) confirmed
that those brook trout which survive can remain lifelong asymptomatic reservoirs of infection,
shedding virus in urine, feces and sexual fluids. They reported that although virus could be
detected in the leukocyte cell fraction from a small portion of asymptomatic 2 year old brook
trout it was present in the majority of 4 year old fish, with up tolO infectious doses being
detected in an infected cell. The virus was universally detectable with high titers in the
kidney-spleen, pancreas-pyloric cecum and in the cell fraction of ovarian fluid from heavily
infected brook trout. Chemical immunosuppression accompanied by elevated water temperatures
increased both the percentage of fingerlings which were IPNV positive, as well as their virus
titers. The offspring of infected broodstock were not infected, but could become infected by
exposure (56).

Rainbow trout
Although rainbow and brook trout are genetically similar, their susceptibility to IPNV is
not identical. McKnight and Roberts (57) noted that the extent of pancreatic damage was
variable in moribund fish, but there was always severe enteritis. Sano, Tanaka and Fukuzaki
(58) felt that they could at least show virus clearance in fish with high antibody titers. They,
like Swanson, also noted that older fish which had survived infection later developed pancreatic
damage. Yu, MacDonald and Moore (51) found that rainbow trout leukocytes rarely harbored
IPNV or supported virus replication. They recovered up to 400 infectious units of virus per
cell from a very few (0.01%) adherent leukocytes (which they suggested could have been
from a subpopulation of cells) of asymptomatically infected three and four year old fish. They
found no IPNV-specific fluorescence, suggesting little or no viral replication. Swanson and
Gillespie (49) recovered IPNV from white blood cells for up to forty days after experimental
infection, but found that the virus localized in the kidneys thereafter. They found a persistent
viremia in yearling rainbow trout, and destruction of the pancreatic acinar cells with necrosis.
Macrophages appeared in the affected tissue. Swanson and Gillespie (49) then followed the
effects produced by the virus and found IPNV primarily in the gasterointestinal tract, using
IPNV strain VR-229 and a field isolate. Although virus titers rose, VR-229 did not cause
mortalities in fry or yearlings. The field isolate replicated in the pancreatic acinar cells,
Kidney and spleen samples have been consistently noted to have higher virus titers than
pyloric cecum (59, 60). Again, it was noted that there appeared to be no correlation between
antibody levels and virus isolation.
Okamoto and Sano (61) were able to delineate more of the pathogenic effects of IPNV on
fry. Using high and low virulent strains of IPNV and resistant and sensitive rainbow trout,
IPNV in salmonids 213

they showed that virus propagated in the pancreas of healthy and moribund fry. However,
only in the moribund fry was the virus found to multiply in the intestines. Water temperature
at the time of viral exposure did not affect the likelihood of illness, but the temperature of the
rearing water (5-20°C) was related to disease outbreaks. Mortalities at 5°C were minimal,
whereas at 10, 15, and 2O”C, 80-100% of the fish died. Later, Okamoto and Kamon (62)
confirmed the protective effects of low temperatures in IPN. Fry infected and held at 5°C
suffered no mortalities; however, similar fry succumbed when moved to 15°C even after 93
days. Their work confirmed that the cause of death in fry is intestinal virus, as had been
shown by McKnight and Roberts in 1976 (57) and that viral replication in the pancreas was
not, at least immediately, lethal. More recently, Okamoto et al. (63) have shown that stocks of
rainbow trout can develop a genetically stable resistance to IPN.
Ahne and Negele (64) reported that eyed eggs of rainbow trout and Arctic char could be
infected via water and that the egg cases remained infectious even after hatching. Sperm and
eggs rapidly lost virus and hatched fry were free of virus. The source of infection, at least in
the case of the Arctic char, appeared to be ingestion of discarded virus-coated egg cases.
Dorson (45) noted that virus could be found in the skin mucus of rainbow trout for long
periods of time, along with specific antibody in the mucus. This antibody was described as
being elicited independently of the circulating serum antibody.
Virus has been repeatedly found in leukocytes, although its ability to replicate in them
would appear limited. Mangunwiryo and Agius (65) described the isolation of IPNV from
leukocytes of carrier fish. They found no IPNV (strain Sp) in juvenile hatchery-held rainbow
trout but virus and antibody were detected in healthy larger fish in open ponds, implying that
fish beyond the stage of disease sensitivity could mount an antibody response should they
become exposed. Virus titers peaked as waters warmed in the spring while antibody titers
peaked as waters cooled in the fall. Rodriguez et al. (6) also described the presence and
possible multiplication of IPNV in leukocytes. Ledo et al. (3) ascribed adult rainbow trout
mortalities to the concurrent presence of bacterial pathogens with IPNV. Dorson, DeKinkelin
and Torchy (66) showed interferon synthesis in fry infected with pathogenic IPNV-Sp. Adult
rainbow trout injected with the virus did not produce interferon, suggesting that the virus did
not replicate in adult fish as it did in fry.

Atlantic salmon
Virus has been isolated from disease-free Atlantic salmon (67). Dorson (68) reported that
salmon infected at 3 weeks of age could still be carriers of IPNV after 6 months, along with
detectable antibody. Although he tried several different isolates, Dorson (4) was unable to
reproduce IPN disease in salmon fry. He pointed out that although virus had frequently been
isolated, no clinical cases were seen.
An atypical finding of IPNV-related late pathology in Atlantic salmon was reported by
Smail and Munro (69) who found IPNV-Sp associated with subclinical pancreatic necrosis in
Atlantic salmon post-smolts, and correlated mortality with rising virus titers in the pre-smolting
fish. They suggested that this late response might be due to immune recognition of viral
antigens on the acinar cells of the pancreas. This deduction would support the earlier findings
with older surviving rainbow trout (49, 58) showing pancreatic damage. All populations of
survivors had a significant number of individuals with life-long infections. In the great majority
of the fish (141/143), virus could be detected yet only 65% of the fish developed antibody,
whether infected with the more pathogenic Sp or the less pathogenic Ab strains. The kidneys
214 E. C. Sadasiv

of smolts with pancreatic lesions were normal; the presence of virus did not seem to
compromise the digestive function of the pancreas. In 1986, Knott and Munro (70) isolated
virus from what had been considered virus-negative fish. They pointed out that only one case
of clinical IPN had been found in Scottish Atlantic salmon, despite numerous isolations of
IPNV from carriers. Smail et al. (71) confirmed that a high level of IPNV-Sp was associated
with the mortality of post-smolts. Moribund fish were thin, with virus in the intestinal cells
and necrosis in the pancreatic acinar cells.
Disease was noted as Atlantic salmon culture increased. In 1989, epizootics of farmed
salmon smolts were reported by Krogsrod, Hastein, and Ronningten (72). More than 50% of
the farmed fish in Norway were IPNV carriers (73). Ledo et al. (3) found that, in Spain,
IPNV-Sp but not the disease could be detected in juvenile Atlantic salmon and rainbow trout.
They concluded that virus had been acquired after the fish had been placed on fish farms,
rather than being vertically passed.
It has been difficult to account for the quantity of virus found in apparently healthy fish.
Evensen and Rimstad (74) pointed out that relatively high virus titers (up tolO infectious
doses) can be found in normal fish, but titers of lo6 to 10’ infectious doses per gram of tissue
would indicate IPNV as a cause of death. Using virus-negative smolts (110 g, 4-month,
post-transfer to sea water), Rimstad et al. (75) found that when these fish were inoculated
with a cell-culture-grown Sp strain of IPNV and kept at 11°C none died, no clinical signs
developed, and no pathological or immunohistological evidence for virus replication was
noted. Any pancreatic damage noted was due to the more pathogenic strain Sp rather than the
less pathogenic Ab strain. Small amounts of virus were isolated only sporadically from the
kidneys (but not the pyloric ceca or gills) of presumptively virus-negative fish. However,
virus was consistently isolated from fish inoculated with cell-culture-propagated virus. This
virus also readily spread to contact control fish from which it could be isolated. Virus could
be reisolated from inoculated groups of fish for 80 days even though they were
immuno-histochemically negative for virus replication and without disease symptoms. Rimstad’s
work (75) emphasized the fact that cell-culture grown virus differs in some way from virus
passed either in pancreatic cells or sensitive fry, certainly in its ability to spread to other fish
without appearing to replicate in them. In order to demonstrate such a viral difference in fry,
sensitive fry were exposed to a highly passed strain of IPNV (VR-229). There was 80%
mortality, peaking at 6 days post-exposure. One month later, the survivors were exposed to a
field isolate of IPNV (passed once in cell culture), and again 80% mortality was produced,
with deaths peaking at 6 days (R.E. Wolke and E.C. Sadasiv, unpublished data). Virus was
reisolated from both epizootics, but in agreement with Swanson and Gillespie’s (67) work no
histological lesions were found in the fish. The full interpretation of this finding has not yet
been developed.
McAllister et al. (76) have reported that older Atlantic salmon exposed to IPNV rarely
suffer mortalities due to IPN and their antibody response is variable. Disease seemed to occur
only under conditions of virus exposure accompanied by additional stress, such as might be
found when the fish were undergoing a concurrent bacterial infection.
In 1993, Smail and Munro (77) reported on vertical transmission. Virus could bind to
sperm and enter the eggs of Atlantic salmon, yet the progeny were not infected. Even when
ovarian fluid was carrying virus, the progeny fish were not infected.
IPNV in salmonids 215

ANTIBODY AND VACCINATION


Antigenic relationships among IPNV isolates have been studied by a number of groups, but
IPNV does not fit well into classification schemes. It has been categorized by Hill and Way
(78) in 1988; Christie, Ness, and Djupvik (79) in 1990; Lecomte, Arella, and Bertiaume (31)
in 1992; and Tarrab et al. (80) in 1993. Genotype does not equal serotype (81) and
non-pathogenic viral forms can be serologically identical to pathogenic forms (82, 83).
Antibody response in salmonids is inherently variable (84), but the majority unquestionably
develop antibody. Whether the antibody is protective is not easily demonstrated, for it must be
tested by exposing susceptible fry to virus. In trying to interpret virus-antibody reactions, one
must bear in mind that both the virus and the antibody, as well as the strain of fish or the cells
being used as the test system, may vary. It is not even certain that antibody is produced only
in response to infection, as opposed to viral exposure. In Norway, Havarstein et al. (73) found
that adult Atlantic salmon produced variable levels of IPNV-specific immunoglobins in
response to intraperitoneal injection with cell-culture-passed IPNV. In our laboratory, we
confirmed that Atlantic salmon exposed to IPNV by immersion can produce neutralizing
antibodies following exposure to virus with a history of many passages in cell culture. We
tested a limited number of mature returning Atlantic salmon and found that specific IPNV
antibody levels were higher in spawning fish (presumptively virus-free, hatchery-raised and
released) than in similar, hatchery-held fish. Our conclusion is that during their maturation in
the oceans, they could have been exposed to virus and developed antibodies. These antibodies
were detected by ELISA, so they were not necessarily neutralizing antibodies.
The failure of fish to develop protective immunity after exposure or vaccination has been
ascribed to tolerance or to immune suppression. Tolerance was expected to result from an
encounter with virus before immunological competence had developed. Immune suppression
would have been due to low temperatures. In 1990, Bootland, Dobos, and Stevenson (53), did
a careful study to address the first point. They reported that immunization with inactivated
virus reduced brook trout mortalities in some cases but did not prevent infection of any age
group of fry. Immunization could lessen mortalities of 2 and 3 week eleutheroembryos and 6
week alevins but not 1, 4, 5, 7, or 8 week old fish. They proposed that lymphoid organs are
not only important in the immune response but that they could also be involved in IPNV
pathogenesis. This supported the earlier suggestion made by Hedrick and Fryer (39) that,
upon IPNV introduction, the virus initiated widespread infection of the visceral organs and
then became sequestered from antibody exposure before an adequate immune response could
be mounted.
Temperature affects virus production as well as immunity. Telost immune function in
relation to temperature has been reviewed by Bly and Clem (85). The production of antibody
has been found to be faster and of a higher magnitude at higher temperatures, with lower
temperatures tending to inhibit immune responses. Bly and Clem believe that the adaptive
immune responses of virgin T-cells, rather than memory T-cells, B-cells or accessory cells (at
least of channel catfish), are particularly susceptible to low temperatures (85). Immunization
at low temperatures does not induce antigen-specific tolerance, but rather specifically inhibits
T-helper cells, with the immunologically non-permissive temperature for salmonids being
4°C. Thus, based upon Bly and Clem’s summary, fish exposed to IPNV supposedly would not
develop an immunological tolerance to the virus and argues against any role for it in virus
persistence. Low temperature would also not explain the mechanism of IPNV persistence. In
our laboratory, we have shown that Atlantic salmon kept at 6°C can develop a good antibody
response to IPNV, although somewhat more slowly than those held at higher temperatures
(86).
216 E. C. Sadasiv

There has been question of whether IPNV can replicate in white blood cells or is simply
carried by them. Both Yu, MacDonald, and Moore (51) and McAllister et al. (55) have found
virus-positive brook trout leukocytes, with each cell containing up to 400 virus infectious
doses, which suggests virus multiplication in the cells. Johansen and Sommer (87) found that
Atlantic salmon adherent leukocytes (primarily macrophages) from head kidney could carry
IPNV. Adherent leukocytes were described by Tate, Kodoma, and Izawa (88) as possessing
surface immunoglobulins and corresponding to avian and mammalian B-cells. The number of
virus-specific fluorescent cells increased from lo2 to 106, during one week in culture which
suggests replication (87). These studies do not, however, conclusively prove that virus or
virus-infected cells could not have been ingested by scavenging macrophages, from which the
virus could be released undegraded at a later time. Dannevig et al. (89) demonstrated that the
sinusoidal endothelium of the head kidney can take up particles the size of IPNV.
Koumans-vanDiepen et al. (90) deduced that rainbow trout immunoglobulin-bearing cells first
appear in the kidney 4-5 days after hatching, while those of salmon were noted by Ellis (91)
at first feeding, times that coincide with IPN sensitivity. All of these points taken together
suggest that IPNV infects leukocytes.
There have been several reports that IPNV could cause immunosuppression, as avian
birnaviruses do. Tate, Kodama, and Izawa (88) proposed that IPNV could have an
immunosuppressive effect on those rainbow trout mononuclear leukocytes which correspond
in function to avian and mammalian B-cells, and that this could be involved with the
establishment of a carrier state. These putative B-cells would appear to be the cells found by
Johansen and Sommer (87) to allow replication of IPNV. Older salmonids exposed to IPNV
were reported to have variable antibody responses to virus. The establishment of a carrier
state was attributed to the ability of the virus to suppress mononuclear leukocytes.
In birds, birnaviruses cause immunodeficiency (92). Avian birnaviruses have a predilection
for actively dividing and differentiating lymphocytes of B lineage (93), particularly those
which carry IgM, the immunoglobulin most similar to piscine immunoglobulins (94). Infected
young birds experience a decline in the percentage of immunoglobulin-expressing cells, due
to lysis of antibody-producing B-cells (95). All fish lymphocytes are similar to mammalian
B-cells, in that they possess membrane immunoglobulins (91).
Great advances in understanding the salmonid immune system are currently being made in
laboratories around the world (96), but a discussion of them is beyond the scope of this
review. However, a few points should be noted. Salmonids have been described as also
possessing a T-cell-like immunity, possibly based in the non-adherent leukocytes (88). Viruses,
such as IPNV, are described as T-dependent antigens (54). In mammals, carbohydrate antigens
can directly induce antibody synthesis without T-lymphocyte activation (97). If IPNV is
indeed glycosylated, would it then be capable of directly inducing antibody? At this time, not
enough is known about the development of the salmonid immune system to accurately
determine the processes which occur with IPNV, but these observations suggest plausible
lines of investigation. However, if IPNV compromised the immunity of fish, it should be
reflected in a decreased resistance to other pathogens. Bruno and Munro (98) have shown that
is not so, at least for bacterial antigens.
Vaccine development has not been easy despite great efforts on the part of many researchers.
In 1982, Dorson (4) reported that protection could be produced with inactivated pathogenic
virus, but when he (45) summarized the work on immunization against IPN in 1988, the
production of an effective vaccine seemed more elusive. Avirulent cell-culture virus, capable
of producing an active infection, has not proven to be of great value as a live vaccine. Fry
infected with it have not been protected against subsequent challenge with virulent virus of
IPNV in salmonids 217

the same serotype. Fish surviving IPN became poor neutralizing antibody producers, despite
the fact that they reacted to injection of virus by producing high antibody titers. Viral antigen
could be found in epithelial macrophages (99). Kelly and Nielsen (100) studied the neutralizing
antibodies induced by vaccination of rainbow trout with different IPNV strains, using
cell-culture adapted virus. By using methods designed to avoid the effects of non-specific
viral inhibitors, they also found that the immune response was highly variable.
Manning and Leong (38) have been the only researchers able to induce protective immunity,
and they did it in rainbow trout fry using cloned viral proteins of the entire larger RNA
segment. This indicates that the genome products produced by bacterial cells, and possibly
insect cells (16), are different from the virus produced in fish-derived cell-culture. Lawrence
et al. (lOl), Havarstein et al. (102), Hah, Park, and Jeong (103), Bootland (104) and Singer
(105) have also had some recent success with subunit vaccines. Leong and Fryer (97) have
summarized the problems and pointed out that there are still no commercially available IPN
vaccines.

CONCLUSIONS

No single answer can yet be given as to the way in which IPNV is able to be carried in
fish. A few points do seem to be clear, however. Late-occurring fish deaths, some noted to be
related to increasing water temperatures, could be explained by antibody-mediated processes
(4, 61, 69). Low temperature is found to have a sparing effect on fry. Cold temperatures
would decrease the rate of virus production by infected cells, as well as slow the production
of antibodies. The mechanism for such sparing has not been demonstrated, but would likely
relate to the maturation of vulnerable cells to a stage where they either would not be sensitive
to the virus or they would produce only non-pathogenic virus. The evidence for autoimmune
cellular damage is circumstantial, but rather convincing.
Although the vertical passage of virus inside sperm or eggs from brood stock to progeny
does not seem to occur, virus can be carried on sperm and egg cases (6). Infection has been
reported to coincide with time of first feeding (l), which would be logical if its occurrence
were due to fry eating their virus-coated egg cases. But, if the chain of infection can be
broken by protecting fry at their vulnerable stage, much of the concern about carriage and
passage of IPNV becomes less important. This does not resolve the problem of whether any
aquatic birnavirus carried over to fry might be lethal to them. Our single test with Atlantic
salmon suggests that virus with low pathogenicity can become more pathogenic again by
passage in susceptible fish.
Not all virus detected fish indicates systemic infections. Virus found in surface mucus could
result from an active infection of epidermal cells or it could be attributable to absorption from
water. Similarly, virus found in the digestive system or feces could be ingested material.
Aquatic bimaviruses have been isolated from numerous invertebrates and IPNV can be
harbored by many non-salmonid species. Of these species, only a few have proved susceptible
to IPN disease. Roberts (106) reported on the IPNV prevalence of steelhead trout broodstock,
which do not develop lethal IPN disease although the virus is often found in association with
them (P McAllister, personal communication). Roberts (106) noted that no pattern of
prevalence was found (O-10% positive pooled samples over a period of 10 years). Virus did
not easily spread to other broodfish, as would be expected in the course of a true infection.
Although the virus is most often found in kidneys, the kidney has not been considered to
be the prime site for virus replication (50). Histopathologic changes can be visualized there
218 E. C. Sadasiv

(1) and multiple tiny viral foci have been seen within its hematopoietic cells (48). Salmonid
kidneys perform a blood filtration function, as well as being a hematopoietic organ. The
kidney, with its reservoir of progenitor blood cells, could be the key to viral carriage. One
might speculate that, if IPNV were to be sequestered in the hematopoietic kidney cells (in
some yet-to-be-discovered manner), those cells might produce infected leukocytes in which
the virus could be replicated.
Viral pathogenicity is controlled by viral structural proteins, which can be altered phenotypically
as well as genotypically (44). The mechanisms of alterations have not been unequivocally
established. Ahne (107) pointed out the influence of host cells on the formation of VP-2 from
preVP-2. The size of avian birnavirus VP-2 can correlate with pathogenicity. When the outer
protein is produced in fibroblastic cells, it can be larger, less pathogenic, elicit antibody which
is less protective and to have a different net charge from that produced from a hemopoietic
organ. Cellular environments may not be capable of cooperating with the viral NS protein to
suitably truncate VP-2. Additional changes in shape and weight of VP-2 would come from
altered numbers or locations of carbohydrate moieties; cellular-mediated glycosylation might
produce virus with altered glycosylation patterns (108). Glycosylation has been reported in
IPNV by Estay et al. (25) but it has not been found in related avian birnavirus. Either or both
of these mechanisms could account for the lack of coincidence between serotype and genotype.
Structurally altered particles could play a role in modulating virus production, allowing for
the different responses detected. If either the trimming of VP-2 or its glycosylation are
critical, the differences should be detected after only a single passage in cultured cells. Since
Sano et al. (19) reported pathogenic as well as non-pathogenic virus being produced in cell
culture, this may be too simplistic an explanation. To my knowledge no work has been done
using pathogenic virus directly taken from fish.
No alteration of expression of IPNV proteins during persistent infection has been found;
however, several groups have recently reported on additional proteins. Work from Dobos’
laboratory (101, 109) described one produced by a frame-shift reading of the end section of
the larger RNA which encodes a 17K polypeptide found in infected cells. Manning and Leong
(38) describe a protein expressed from the putative negative strand of one of the RNAs.
Havarstein et al. (102) reported one specified by the N terminus of the larger RNA near the
area coding for VP-2. These findings point towards additional gene products which could be
involved in modulating persistence (or any other yet undiscovered function). Knott and Munro’s
(70) finding of IPNV in leukocytes only after they were stimulated suggests that an altered
form of the virus could have initially been present in those cells. This hypothesis is also
supported by the work of Tate et al. (88), as well as the work of Rimstad et al. (75). However,
it could also be explained by a low sensitivity of the detecting systems.
Virus with a low potential for disease-causation, as well as a limited ability to induce
protective antibody, could be produced from persistently infected epidermis or other epithelial-derived
cells. This occurs with the avian birnavirus (110). For this birnavirus, vaccine produced from
fibroblasts or whole embryos is not as protective as a vaccine in which the antigen is
produced in cells of the hematopoietic organ. As Dorson (3) noted with kidney cells,
epithelial-derived cells (skin and gut, including kidney cells) might sustain long-term infections,
with low rates of production of non-pathogenic virus. This could be one possible explanation
for virus production which can extend for years. Virus production from the skin could also
account for the lack of clearance of virus from fish having circulating serum antibodies. Virus
produced from the skin might be less affected by circulating antibodies; the locally-induced
antibody from skin might not be directed at pathogenic virus. It is worth noting that those rare
persistently infected cells of the kidney might, indeed, be hematopoietic cells.
Aquareoviruses 219

For virus to persist for years in a fish without being detected, it would have to either be
sequestered and present in such low numbers that current systems do not detect it; or be
present only as genetic material which would not interact with the immune system. One
would not expect that such a simple virus as IPNV would be able to carry that quantity of
genetic information which would allow persistence of the type seen, for example, with
herpesviruses. With the DNA-containing herpesvirus, only the nucleic acid, rather than the
complete virus with its proteins, is latent in cell nuclei of nerve and brain tissue. This would
not likely be the case with a virus, like IPNV, which is not known to enter nuclei or produce
a DNA intermediate which could integrate the cellular genome.
Virus would not be expected to remain in circulating leukocytes for years; however, B-cell
immortalization by a herpesvirus suggests that viruses can alter cell functioning (111). Could
virus be sequestered in the renal hematopoietic stem or in blast cells, as suggested and
discounted by Yu, MacDonald, and Moore (51), or could the virus influence the production/activation
of leukocytes? There is currently no information to prove either hypothesis. If IPNV were
able to suppress the immune system (70, SS), it would explain the variable immune responses,
as well as being an alternative explanation for the difficulty in producing protective vaccines.
Virus can be found in leukocytes and replication may occur in them. Is it possible that
there could be some interaction of stable viral antigens with macrophages of the head kidney
(112)? Investigation of this might also elucidate the mechanism of the long-term presence of
putative pathogens.
It is interesting to note that several other fish pathogens can persist for long periods of time
including the rhabdoviruses infectious hematopoietic necrosis (IHNV) and viral hemorrhagic
septicemia (VHS). These viruses are quite labile, in contrast to the birnaviruses. Since
birnaviruses and rhabdoviruses are so different structurally, it seems unlikely that the two
virus families would have evolved similar mechanisms for latency. More plausibly, similarities
should be looked for in the response of the fish to these very different infections.

Note added in proof: Some additional recently published work by Johansen and Sommer (Johansen L.-H. and A.-I
Sommer, Aquaculture 132: 91-95. 1995; Johansen L.-H. and A.-I. Sommer, J. Fish Diseases 18: 147-156. 1995) has
shown that IPNV replication was found in adherent head kidney leukocytes (macrophages). Viral specific fluorescence
was detected in both carrier and in vitro infected macrophages, with the number of infected cells increasing during a
week in culture. IPNV infected macrophages were found to have reduced respiratory burst activity. They theorize that
non-cytolytic infected macrophages may protect virus from immune defenses of host, leading to a “carrier” condition.

Acknowledgments - The author wishes to thank Drs. Pei W. Chang, Richard E. Wolke, and Robert E. Carleson for
critical readings of the manuscript; and for helpful discussion with Drs. Philip E. McAllister, David A. Smail, Sharon
A. MacLean, Eva Nagy and Ahmed A. Azad. The author is also indebted to Ms. Carole Klawansky for gracious
assistance in manuscript preparation. Contribution No. 2950 of the College of Resource Development, University of
Rhode Island, with support from the Rhode Island Agricultural Experiment Station and the US Department of
Agriculture. This work was also supported by a gift from Abbott Laboratories, in recognition of the achievements of
Dr. George Dawson.

REFERENCES
I. Wolf, K. (1988). Infectious pancreatic necrosis. In: Wolf, K. (ed.) Fish viruses and fish viral diseases, Come11
Univ. Press., Ithaca, NY, pp. 115-157.
2. Dobos, P., Roberts, T.E. (1983). The molecular biology of infectious pancreatic necrosis virus: a review. Can. J.
Viral. 29: 337-384.
3. Ledo, A., Lupiani, B., Dopaza, C.P., Toranzo, A.E., Barja, J.L. (1990). Fish viral infections in northwestern
Spain. Microbiologia Sem. 6: 21-29.
4. Dorson, M. (1982). Infectious pancreatic necrosis of salmonids: overview of current problems. In: Anderson, D.,
Dorson, P., Dubourget, M. (eds.) Les antigens des micro-organisms pathogenic des poissons. Symposium
International de Talloires. Assoc. Corp. Etudients Med., Lyon, pp. 7-32.
220 B. Lupiani, K. Subramanian and S. K. Samal

5. Maheshkumar, S., Goyal, S.M., Peterson, R.B., Economon, PP. (1991). Method of the concentration of infectious
pancreatic necrosis virus from hatchery water. J. Virol Methods 31: 211-218.
6. Rodriguez Saint-Jean, S., Pilar Vilas Minando, M., Angel Palacios, M., Perez Prieto, S. (1991). Detection od
infectious pancreatic necrosis in a carrier population of rainbow trout Oncorhynchus mykiss, by flow cytometry.
J. Fish Dis. 14: 545-553.
7. Dominguez, J., Babin, M., Sanchez, C., Hedrick, R.D. (1991). Rapid serotyping of infectious pancreatic necrosis
virus by one-step enzyme-linked immunosorbent assay using monoclonal antibodies. J. Viral. Meth. 31: 93-103.
8. Davis, P.J., Laidler, L.A., Perry, P.W., Rossington, D., Alcolk, R. (1994). The detection of infectious pancreatic
necrosis virus in asymptomatic carrier fish by an integrated cell-culture and ELISA technique. J. Fish Dis. 17:
99-110.
9. Diamant, A., Smail, D.A., McFarlane, L., Thomson, A.M. (1988). An infectious pancreatic necrosis virus isolated
from common dab Limanda limanda previously affected with X-cell disease, a disease apparently unrelated to
the presence of the virus. Dis. Aquat. Org. 4: 223-227.
10. Lipipun V., Caswell-Reno, F?, Hsu, Y.L., Wu, L., Tung, M.C., Reno, P.W., Wattauavijarn, W., Nicholson, B.L.
(1989). Antigenic analysis of Asian aquatic birnavirus isolates using monoclonal antibodies. Fish Pathol. 24:
155-160.
11. Novoa, B., Figueras, A., Puentes, C.F., Ledo, A., Toranzo, A.E. (1993). Characterization of a bimavirus isolated
from diseased turbot cultured in Spain. Dis. Aquat. Org. 5: 163-169.
12. Rivas, C., Cepeda, C., Dopazo, C.P, Novoa, B., Noya, M. (1993). Marine environment as reservoir of bimaviruses
from poikilothermic animals. Aquaculture :15: 183-194.
13. Hsu, Y.L., Chen, B.S., Wu, J.L. (1993). Demonstration of infectious pancreatic necrosis virus strain VR-229 in
Japanese eel, Anguilfu japonica Temminck & Schlegel. J. Fish Dis. 16: 123-129.
14. Melby, HP, Caswell-Reno, P, Falk, K. (1994). Antigenic analysis of Norwegian aquatic bimavirus isolates using
monoclonal antibodies with special reference to fish species, age and health status. J. Fish Dis. 17: 85-91.
15. Dobos, P (1991). Bimavitidae. In: Fran&i, R.I.B., Fouquet, C.M., Knudson, D.L., Brown, F. (eds.) Classification
and nomenclature of viruses: Fifth Report Int. Comm. Taxonomy Viruses. Springer-Verlag, New York, NY, pp.
200-202.
16. Magyar, G., Dobos, P. (1994). Expression of infectious pancreatic necrosis virus polyprotein and VP1 in insect
cells and the detection of the polyprotein in purified virus. Virology 198: 4374t5.
17. Smail, D.A., Irwin, N., Harrison, D., Munro, A.L.S. (1993). Passage and survival of infectious pancreatic necrosis
(IPN) virus in the cow’s gut after feeding a silage mixture containing IPN virus. Aquaculture 113: 183-187.
18. Mortensen, S.H., Bachere, E., LeGall, G., Mialhe, E. (1992). Persistence of infectious pancreatic necrosis virus
(IPNV) in scallops Pecren maximus. Dis. Aquat. Org. 12: 211-227.
19. Sano, M., Okamoto, N., Fukuda, H., Saneyoshi, M., Sane, T. (1992). Virulence of IPNV is associated with the
larger RNA segment (RNA segment A). J. Fish Dis. IS: 283- 293.
20. Barrie, R.J., Mason, CL., Leong, J.C. (1992). Identification of a conserved antigenic domain in the major capsid
protein of infectious pancreatic necrosis virus. National Oceanic and Atmospheric Administration Technical
Reports of National Marine Fisheries Service, 111.
21. Duncan, R., Mason, CL., Nagy, E., Leong J.A., Dobos, P. (1991). Sequence analysis of infectious pancreatic
necrosis virus genome segment B and its encoded VP-I protein: a putative RNA-dependent RNA polymerase
lacking the Gly-Asp-Asp motif. Virology 181: 541-552.
22. Calve& J.G., Nagy, E., Soler, M., Dobos, P (1991). Characterization of the VPg-dsRNA linkage of infectious
pancreatic necrosis virus. J. Gen. Virol. 72: 2563-2567.
23. Duncan, R., Nagy, E., Krell, P.J., Dobos P. (1987). Synthesis of the infectious pancreatic necrosis virus polyprotein,
detection of virus-encoded protease, and fine structure mapping of genome segment A coding regions. J. Virol. 61:
3655-3664.
24. Heppell, J., Berthiaume, L., Corbin, F., Tarrab, E., Lecomte, J., Arella, M. (1993). Comparison of amino acid
sequences deduced from a cDNA fragment obtained from infectious pancreatic necrosis virus (IPNV) strains of
different serotypes. Virology 195: 840-844.
25. Estay, A., Farias,G., Soler, M., Kuznar, J. (1990). Further analysis of the structural proteins of infectious pancreatic
necrosis virus. Virus Res. 15: 85-95.
26. Manning, D.S., Mason, CL., Leong, J.C. (1990). Cell-free translational analysis of the processing of infectious
pancreatic necrosis virus. Virology 179: 9-15.
27. Caswell-Reno, P., Reno, P., Nicholson, B.L. (1986). Monoclonal antibodies to infectious pancreatic necrosis virus:
analysis of viral epitopes and comparison of different isolates. J. Gen. Viral. 67: 2193-2205.
28. Wolski, S.C., Roberson, G.S., Hetrick, EM. (1986). Monoclonal antibodies to the Sp strain of infectious pancreatic
necrosis virus. Vet. Immunol. Immunopath. 12: 373-381.
29. Caswell-Reno, P., Lipipun, V., Reno, PW., Nicholson, B.L. (1989). Utilization of a group reactive and other
monoclonal antibodies in an enzyme immunoblot assay for identification and presumptive serotyping of aquatic
bimavirus. J. Clin. Microbial. 27: 1924-1929.
30. Lipipun, V., Caswell-Reno, P., Hsu, Y.L., Wu, J-L., Tung, M-C., Reno, P.W., Wattanavijam, W., Nicholson, B.L.
(1989). Antigenic analysis of Asian aquatic bimavirus isolates using monoclonal antibodies. Fish Pathol. 24:
155-160.
Aquareoviruses 221

3 1. Lecomte, J., Arella, M., Bertiaume, L. (1992). Comparison of polyclonal and monoclonal antibodies for serotyping
infectious pancreatic necrosis virus (IPNV) strains isolated in eastern Canada. J. Fish Dis. 15: 431436.
32. Christie, K.E., Havarstein, L.S., Djupvik, H.O., Ness, S., Endresen, C. (1988). Characterization of a new serotype
of infectious pancreatic necrosis virus isolated from Atlantic salmon. Arch. Virol. 103: 167-177.
33. Rimstad, E., Homes, E., Olsvik, 0.. Hyllseth, B. (1990). Identification of double-stranded RNA virus by using
polymerase chain reaction and magnetic separation of the synthesized DNA segments. J. Clin. Microbial. 28:
22752278.
34. Rimstad, E., Krona, R., Homes, E., Olsvik, 0.. Hyllseth, B. (1990). Detection of infectious pancreatic necrosis virus
(IPNV) RNA by hybridization with an oligonucleotide DNA probe. Vet. Microbial. 23: 211-219.
35. Heppell, J., Berthiaume, L., Tarrab, E., Lecomte, J., Arella, M. (1992). Evidence of genomic variations between
infectious pancreatic necrosis virus strains determined by restriction fragment profiles. J. Gen. Virol. 73:
2863-2870.
36. Pryde, A., Melvin, W.T., Munro, A.L. (1993). Nucleotide sequence analysis of the serotype-specific epitope of
infectious pancreatic necrosis virus. Arch. Virol. 129; 287-293.
37. Dopazo, C.P., Hetrick, EM., Samal, S.K. (1994). Use of cloned cDNA probes for diagnosis of infectious pancreatic
necrosis virus infections. J. Fish Dis. 17: I-16.
38. Manning, D.S., Leong, J.C. (1990). Expression in Escherichia coli of the large genomic segment of infectious
pancreatic necrosis virus. Virology 179: 16-25.
39. Hedrick, R.P., Fryer, J. (1981). Persistent infection of three salmonid cell lines with infectious pancreatic necrosis
virus (IPNV). Fish Pathol. 15: 163-172.
40. Borzakian, S., Pellitier, I., Calvez, V., Colbere-Garapin, F. (1993). Precise missense and silent point mutations are
fixed in the genomes of poliovirus mutants from persistently infected cells. J. Virol. 67: 2914-2917.
41. Lloyd, R.E., Bovee, M. (1993). Persistent infection of human erythroblastoid cells by poliovirus, Virology 194:
200-204.
42. Lopez, O.J., Osorio, EA., Kelling, C.L., Donis, R.O. (1993). Presence of bovine viral diarrhoea virus in lymphoid
cell populations of persistently infected cattle. J. Gen. Virol. 74: 1201-1206.
43. Kennedy, J.C., MacDonald, R.D. (1982). Persistent infection with infectious pancreatic necrosis virus mediated by
defective-interfering (DI) virus particles in a cell line showing strong interference but little DI replication. Virology
58: 361-371.
44. Dorson, M., Gastric, J., Torchy, C. (1978). Infectious pancreatic necrosis virus of salmonids: biological and
antigenic features of a pathogenic strain and of a non-pathogenic variant selected in RTG-2 cells. J. Fish Dis. 1:
309-320.
45. Dorson, M. (1988). Vaccination against IPN. In: Ellis, A.E.(ed.) Fish vaccination. Academic Press, London, pp.
162-171.
46. Yamamoto, T. (1975). Frequency of detection and survival of infectious pancreatic necrosis virus in a canier
population of brook trout (Salvelinus fonrinalis) in a lake. J. Res. Bd. Can. 32: 568-570.
47. Swanson, R.N. (1981). Use of the indirect fluorescent antibody test to study the pathogenesis of infectious
pancreatic necrosis virus infection in trout. In: International Symposium on Fish Biologics: Serodiagnostics and
Vaccines, Develop. Biol. Standard. 49, S. Karger, New York, NY, pp. 71-77.
48. Swanson, R.N., Gillespie, J.H. (1981). An indirect FA test for rapid detection of IPNV in tissues. J. Fish Dis. 4:
309-316.
49. Swanson, R.N., Gillespie, J.H. (1982). Isolation of infectious pancreatic necrosis virus from blood and blood
components of experimentally infected trout. Can. J. Fish. Aquat. Sci. 39: 225-228.
50. Swanson, R.N., Carlisle, J.C., Gillespie, J.H. (1982). Pathogenesis of IPNV infection in brook trout Salvelinus
fontinalis (Mitchell), following intraperitoneal injection. J. Fish Dis. 5: 449-460.
51. Yu, K.K.Y., MacDonald, R.D. Moore, A.R. (1982). Replication of IPNV in trout leucocytes and detection of the
carrier state. J. Fish Dis. 5: 401-410.
52. Bootland, L.M., Stevenson, R.M.W., Dobos, P. (1986). Experimental induction of the carrier state in yearling brook
trout: a model challenge protocol for IPNV immunization. Vet. Immunol. Immunopath. 12: 365-372.
53. Bootland, L.M., Dobos, P., Stevenson, R.M.W. (1990). Fry age and size effects on immersion immunization of
brook trout, Salvelinus fonfinalis Mitchell, against IPN. J. Fish. Dis. 13: 113-125.
54. Bootland, L.M., Dobos, P., Stevenson, R.M.W. (1991). The IPNV carrier state and demonstration of vertical
transmission in experimentally infected brook trout. Dis. Aquat. Org. IO: 13-21.
55. McAllister, P.E., Schill, W.B., Owens, W.J., Hodge, D.L. (1993). Determining the prevalence of infectious
pancreatic necrosis virus in asymptomatic brook trout Salvelinus fontinalis: a study of clinical samples and
processing methods. Dis. Aquat. Org. 15: 157-162.
56. McAllister, P.E. (1994). Comparison of methods for detection of infectious pancreatic necrosis virus in fluids
and tissues of virus-carrier brook trout. In: Chang, P.W. (ed.) Detection of fish pathogens for fish health
inspection by non-lethal methods. Cooperative Regional Project Termination Report NRAC #90-38500-5211,
Northeastern Regional Aquaculture Center, University of Massachusettes, Dartmouth, North Dartmouth, MA,
52~~.
57. McKnight, I.J., Roberts, R.J., (1976). The pathology of infectious pancreatic necrosis virus. 1. The sequential
histopathology of the naturally occurring condition. Br. Vet. J. 132: 76-84.
222 B. Lupiani, K. Subramanian and S. K. Samal

58. Sano, T., Tanaka, K., Fukuzaki, S. (1981). lmmune response in adult trout against formalin killed concentrated
IPNV. In: International Symposium on Fish Biologics: Serodiagnostics and Vaccines, Develop. Biol. Standard.
49, S. Karger, New York, NY, pp. 63-70.
59. Agius, C., Mangunwiryo, H., Johnson, R.H., Smail, D.A. (1982). A more sensitive technique for isolating
IPNV from asymptomatic carrier trout Sulmo gairdneri Richardson. J. Fish Dis. 5: 285-292.
60. Agius, C., Richardson, A., Walker W. (1983). Further observations on the co-cultivation method for isolating
IPNV from asymptomatic carrier trout Salmo gairdneri Richardson, J. Fish Dis. 6: 477-480.
61. Okamoto, N., Sano, T. (1984). Study of the manifestation mechanism of IPN (Abstract 37-5, p. 377) Sixth
International Congress Virol. Sendai, Japan.
62. Okamoto, N., Kanon, T. (1991). Effects of water temperature on mortality of rainbow trout infected with
IPNV. (Abstract) 14th Annual American Fisheries Soceity/Fish Health Section Meeting, 32nd Western Fish
Disease Conference, (31 July-3 August, 1991).
63. Okamoto, N., Tayama, T., Kawanobe, M., Fujiki, N., Yasuda, Y., Sano, T. (1993). Resistance of a rainbow
trout strain to infectious pancreatic necrosis virus. Aquaculture 117: l-76.
64. Ahne, W., Negele, R.D. (1985). Studies on the transmission of infectious pancreatic necrosis virus via eyed
eggs and sexual products of salmonid fish. In: Ellis, A.E. (ed.) Fish shellfish pathology. Academic Press,
Orlando, FL, Ch. 25, pp. 251-260.
65. Mangunwiryo, H., Agius, C. (1988). Studies on the carrier state of infectious pancreatic necrosis virus
infections in rainbow trout, Salmo gairdneri Richardson. J. Fish Dis. 11: 125-132.
66. Dorson, M., DeKinkeIin, P., Torchy, C. (1992). Interferon synthesis in rainbow trout fry following infection
with infectious pancreatic necrosis virus. Fish Shellfish Immunol 2: 311-313.
67. Swanson, R.N., Gillespie, J.H. (1979). Pathogenesis of infectious pancreatic necrosis in Atlantic salmon (Safmo
sulur). J. Fish. Res. Bd. Can. 36: 587-591.
68. Dorson, M. (1981). Role and characterization of fish antibody. In: International Symposium on Fish Biologics:
Serodiagnostics and Vaccines, Develop. Biol. Standard. 49, S. Karger, New York, NY, pp. 307-319.
69. Smail, D.A., Munro. A.L.S. (1985). Infectious pancreatic necrosis virus persistence in farmed Atlantic salmon
(Salmo safar). In: Ellis, A.E. (ed.) Fish shellfish pathology, Acad. Press, Orlando, FL, Ch. 28, pp. 277-288.
70. Knott, R.M., Munro, A.L.S. (1986). The persistence of infectious pancreatic necrosis virus in Atlantic salmon.
Vet. Immunol. Immunopathol. 12: 359-364.
71. Smail, D.A., Bruno, D.W., Dear, G., McFarlane, L.A., Ross, K. (1992). Infectious pancreatic necrosis (IPN)
virus Sp serotype in farmed Atlantic salmon, Salmo s&r L., post-smolts associated with mortality and clinical
disease. J. Fish Dis. 15: 77-83.
72. Krogsrood, J., Hastein, T., Ronningten, K. (1989). Infectious pancreatic necrosis virus in Norwegian fish farms.
In: Ahne, W., Kurstak, E. (eds.) Viruses of lower vertebrates. Springer-Verlag, New York, NY, pp. 284-291.
73. Havarstein, L.S., Endresen, C., Hjeltnes, B., Christie, K.E., Glette, 3. (1990). Specific immunoglobuins in
serum from Atlantic salmon, Salmo salar L., immunized with fibro salmonicida and infectious pancreatic
necrosis virus. J. Fish Dis. 13: 101-111.
74. Evensen, O., Rimstad, E. (1990). Immunohistochemical identification of infectious pancreatic necrosis virus in
pa&in-embedded tissues of Atlantic salmon (Safmo s&r). J. Vet. Diagn. Invest. 2: 288-293.
75. Rimstad, E., Poppe, T. Evensen, O., Hyllseth, B. (1991). Inoculation of infectious pancreatic necrosis virus
serotype Sp did not cause pancreas disease in Atlantic salmon (Salmo salar L.). Acta. Vet. Stand. 32:
503-510.
76. McAllister, P.E., Schill, W.B., Owens, W.J., Hodge, D.L. (1993). Determining the prevalence of infectious
pancreatic necrosis in asymptomatic brook trout Salvelinus fonrinalis: a study of clinical samples and processing
methods. Dis. Aquat. Org. 15: 157-162.
77. Smail, D.A., Munro, A.L.S. (1993). Vertical transmission studies on IPNV in Atlantic salmon (Salmo salur L.).
International Council for the Exploration of the Seas/MaricuIture Committee I F:36.
78. Hill, B.J., Way, K. (1988). Epidemological aspects of the proposed serological classification of aquatic
bimaviruses. In: Abstracts of International Fish Health Conference. Vancouver, B.C. (19-21 July, 1988).
79. Christie, K.E., Ness, S., Djupvik, H.O. (1990). Infectious pancreatic necrosis virus in Norway: partial serotyping
by monoclonal antibodies. J. Fish Dis. 13: 323-327.
80. Tarrab, E., Berthiaume, L., Heppell, J., Arella, M., Lecomte, J. (1993). Antigenic characterization of serogroup
‘A’ of infectious pancreatic necrosis with three panels of monoclonal antibodies. J. Gen. Viral 74: 2025-2030.
81. Pryde, A., Melvin, W.T., Munro, A.L. (1993). Nucleotide sequence analysis of the serotype-specific epitope of
infectious pancreatic necrosis virus. Arch. Virol. 129: 287-293.
82. Berthiaume, L., Tarrab, E., Heppell, J., Arella, M., Dobos, P., Duncan, R., Lecomte, J. (1992). Antigenic and
genomic differences of two Jasper strains of infectious pancreatic necrosis virus. Intervirology 34: 197-201,
83. Heppell, J., Berthiaume, L., Corbin, F., Tarrab, E., Lecomte, J., Arella, M. (1993). Comparison of amino acid
sequences deduced from a cDNA fragment obtained from infectious pancreatic necrosis virus (IPNV) strains of
different serotypes. Virology 195: 840-844.
84. Olesen, N.J., Vestergaard-Jorgensen, PE. (1986). Quantification of serum immunoglobulins in rainbow trout
Safmo gairdneri under various environmental conditions. Dis. Aquat. Org. 1: 183-189.
85. Bly, J.E., Clem, L.W. (1993). Temperature and immune function. Fish Shellfish Immunol. 2: 159-171.
Aquareoviruses 223

86. Sadasiv, EC., Chang, P.W., Lin, W. (1993). Infectious pancreatic necrosis IPNV antibody as a means of
detection of possible virus carriage in Atlantic salmon surviving virus challenge. J. Shellfish Res. 12:
113-114.
87. Johansen, L.H., Sommer, AI. (1993). Detection of infectious pancreatic necrosis virus (IPNV) in leucocytesfrom
Atlantic salmon (Salmo salar L.) (Abstract P 50-16) IX International Congress of Virology, Glasgow,
Scotland (8-13 August, 1993).
88. Tate, H., Kodama, H., Izawa, H. (1990). Immunosuppressive effect of infectious pancreatic necrosis virus on
rainbow trout (Oncorhynchus mykiss) Jap. .I. Vet. Sci. 52: 931-937.
89 Dannevig, B.H., Lame, A., Press, C.M., Landsverk, T. (1994). Receptor-mediated endocytosis and phagocytosis
by rainbow trout head kidney sinusoidal cells. Fish Shellfish Immunol. 4: 3-18.
90. Kownans-vanDiepen, J.C.F. Taveme-Thiele, J.J., Van Rens, B.T.T.M., Rombout, J.H.W.M. (1994). Immunocytochemical
and flow cytometric analysis of B cells and plasma cells in carp (Cyprinus carpio L.); An ontogenic study.
Fish Shellfish Immunol. 4: 19-28.
91. Ellis, A.E. (1977). Ontogeny of the immune response in Salmo salar. Histogenesis of the lymphoid organs
and the appearance of membrane immunoglobulin and mixed leucocyte reactivity. In: Solomon, J.B., Horton,
J.D. (eds.) Developmental Immunobiology. ElsevierMorth Holland Biomedical Press, Amsterdam, pp. 225-23 1.
92 Azad, A.A., Jagadish, M.N., Brown, M.A., Hudson, P.J. (1987). Deletion mapping and expression in Escherichia
coli of the large genomic segment of a bimavirus. Virology 161: 145-152.
93 Burkhardt, E., Muller, H. (1987). Susceptibility of chicken blood lymphoblasts and monocytes to infectious
bursal disease virus (IBDV). Arch Virol 94: 297-303.
94 Hirai, K., Calnek, B. (1979). In vitro replication of infectious bursal disease virus in established cell lines
and chicken B lymphocytes. Infect. Immun. 25: 964-970.
95 Rosenberg, J., Sharma, J.M., Belzer, S.W., Nordgren, R.M., Naqi, S. (1994). Flow cytometric analysis of B
cell and T cell subpopulations in specific-pathogen-free chickens infected with infectious bursal disease virus.
Avian Dis. 38: 16-21.
96 Stolen, J.S., Fletcher, T.C. (1994). Modulators of fish immune responses. SOS Publications, Fair Haven, NJ,
254~~.
97 Leong, J.C., Fryer, J.L. (1993). Viral vaccines for aquaculture. Ann. Rev. Fish Dis. 3: 225-240.
98. Bnmo, D.W., Munro, A.L.S. (1989). Immunity in Atlantic salmon, Salmo salar L. fry, following vaccination
against Yersinia ruckeri, and the influence of body weight and infectious pancreatic necrosis virus (IPNV) on
the detection of carriers. Aquaculture 81: 205-211.
99. Dorson, M., Perrier, H., Perrier, C., Torchy, C. (1990). Antibody in mucus and other fish secretions: research
on antibody of rainbow trout immunized against infectious pancreatic necrosis, Ichtyophysiol. Acta, 13:
31-12.
100. Kelly, R.K., Nielsen, 0. (1990). Serological propetites of neutralizing antibodies induced by vaccination of
rainbow trout with distinct strains of infectious pancreatic necrosis virus. J. Aquat. Anim. Health 22: 56-60.
101. Lawrence, W.R., Nagy, E., Duncan, R., Krell, P., Dobos, P. (1989). Expression in E. coli of the major outer
capsid protein of infectious pancreatic necrosis virus. Gene 79: 369-374.
102. Havarstein, L.S., Kalland, K.H., Christie, K.E., Endresen, C. (1990). Sequence of large dsRNA segment of
N-l strain of infectious pancreatic necrosis virus: a comparison with other bimaviridae. J. Gen. Virol. 71:
299-308.
103. Hah, Y.C., Park, J.W., Jeong, G. (1992). Neutralization epitope of DRT serotype of infectious pancreatic
necrosis virus (IPNV) isolated in Korea. Proceedings of International Symposium on Infectious Virus in Fish,
9-11 October, Seoul, Korea, pp. 13-22.
104. Bootland, L.M. (1993). Development of a recombinant IPNV vaccine for protection of salmonid fish. Current
Research Information System, United States Department of Agriculture Report, (February, 1993).
105. Singer, J. (1993). Virulence, transmission and bivalent recombinant vaccines against vibriosis and IPN virus.
Current Research Information System, United States Department of Agriculture Report (Feb., 1993).
106. Roberts, S.D. (1993). Infectious pancreatic necrosis virus prevalence in the Wells Summer broodstock. Fish
Health Section/American Fisheries Society Newsletter, 21: 5-7.
107. Ahne, W. (1985). Virusinfektionen bei Fischen: Atiologie, Diagnose und Bekampfung. Zentralbl. Vetrinarmed.
[B] 32: 327-264.
108. Klenk, H.D. (1990). Influence of glycosylation on antigenicity of viral proteins. In: Van Regenmorel, M.H.V.,
Neurath, A.R. (eds.) Immunochemistry of viruses II. The basis of serodiagnosis and vaccines. Elsevier
Science, Amsterdam, The Netherlands.
109. Magyar, G., Dobos, P. (1994). Expression of infectious pancreatic necrosis virus polyprotein and VP! in
insect cells and the detection of the polyprotein in purified virus. Virology 198: 437445.
110. Lange, H., Muller, H., Kaufer, I., Becht, H. (1987). Pathogenic and structural properties of wild type
infectious bursal disease virus (IBDV) and virus grown in vitro. Arch. Viral. 92: 187-196.
111. Ring, C.J.A. (1994) The B cell-immortalizing functions of Epstein-Barr virus. J. Gen. Viral. 75: I-13.
112. Wolke, R.E. (1992). Piscine macrophage aggregates: a review. Ann. Rev. Fish Dis. 2: 91-108.

Potrebbero piacerti anche