www.fgks.org   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

Family Iridoviridae: Molecular and Ecological Studies of a Family Infecting Invertebrates and Ectothermic Vertebrates

A special issue of Viruses (ISSN 1999-4915). This special issue belongs to the section "Animal Viruses".

Deadline for manuscript submissions: closed (28 February 2019) | Viewed by 73526

Printed Edition Available!
A printed edition of this Special Issue is available here.

Special Issue Editors


E-Mail Website
Guest Editor
Department of Microbiology and Immunology, University of Mississippi Medical Center, Jackson, MS, USA
Interests: viruses infecting lower vertebrates; specifically Iridoviruses; anti-viral immune responses in cold-blooded vertebrates; viral taxonomy
Special Issues, Collections and Topics in MDPI journals

E-Mail Website
Guest Editor
Department of Mathematics and Natural Science, Gordon State College, Barnesville, GA 30204, USA
Interests: ranavirus distributions; ranavirus ecology; ranavirus community dynamics; mathematical modelling of ranavirus dynamics; ranavirus phylogenetics

Special Issue Information

Dear Colleagues,

Iridovirids, a generic term describing viruses within the family Iridoviridae, comprise a diverse array of large, icosahedral, double-stranded DNA viruses that infect insects, other invertebrates, and three classes of ectothermic vertebrates (bony fish, reptiles, and amphibians). Iridovirid infections trigger considerable morbidity and have been linked to die-offs among ecologically- and commercially-important fish, reptiles, and amphibians. In this Special Issue of Viruses we will provide a representative sample of current research focused on cellular, molecular and ecological aspects of iridovirid biology. Although clinical disease linked to iridovirid infection, i.e., lymphocystis disease, has been known since the turn of the 20th century, concerted study of iridovirid biology did not begin until Allan Granoff’s identification of frog virus 3 (FV3) in the mid-1960s. Through his efforts and those of others, FV3 became the best characterized member of the family. These studies focused mainly on replicative events in FV3-infected cells and defined the essential elements of virus replication. Although early work was centered primarily on molecular aspects of FV3, research efforts during the last 30 years have expanded to include numerous studies on the ecology of iridovirid infections, replicative events among other species within the genus Ranavirus, as well as other genera within the family, and immune responses to iridovirid infections. Recently, genomic sequence analysis of over 40 members of the family generated a phylogenetically robust basis for our understanding of viral taxonomy and provided a facile way to identify and classify newly identified viruses. In addition, sequence information has provided the basis for numerous studies of viral gene function using siRNA- and antisense morpholino oligonucleotide-mediated knock down, knock out and conditionally-lethal mutants, and ectopic expression studies. Using these approaches, the functions of several viral replicative and immune-modulating proteins have been determined. Moreover, study of virus-encoded immune evasion proteins combined with ongoing research into host anti-viral immunity has provided insights into the evolutionary origins of the vertebrate immune system and may further the development of protective vaccines. The world of iridovirid research has expanded greatly in the last 30 years and we look forward to providing a venue highlighting different facets of research focused on the family Iridoviridae.

Prof. Dr. Gregory Chinchar
Assoc. Prof. Dr. Amanda Duffus
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Viruses is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2600 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • iridovirus
  • Iridoviridae
  • ranavirus
  • viral pathogenesis
  • modeling of infectious disease
  • vaccine development
  • elucidation of viral gene function
  • anti-viral immune responses
  • host-virus interaction
  • viral immune evasion
  • viral ecology

Published Papers (17 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Editorial

Jump to: Research, Other

3 pages, 171 KiB  
Editorial
Molecular and Ecological Studies of a Virus Family (Iridoviridae) Infecting Invertebrates and Ectothermic Vertebrates
by V. Gregory Chinchar and Amanda L. J. Duffus
Viruses 2019, 11(6), 538; https://doi.org/10.3390/v11060538 - 9 Jun 2019
Cited by 10 | Viewed by 2407
Abstract
Research involving viruses within the family Iridoviridae (generically designated iridovirids to distinguish members of the family Iridoviridae from members of the genus Iridovirus) has markedly increased in recent years [...] Full article

Research

Jump to: Editorial, Other

25 pages, 3349 KiB  
Article
Detection and Characterization of Invertebrate Iridoviruses Found in Reptiles and Prey Insects in Europe over the Past Two Decades
by Tibor Papp and Rachel E. Marschang
Viruses 2019, 11(7), 600; https://doi.org/10.3390/v11070600 - 2 Jul 2019
Cited by 23 | Viewed by 4030
Abstract
Invertebrate iridoviruses (IIVs), while mostly described in a wide range of invertebrate hosts, have also been repeatedly detected in diagnostic samples from poikilothermic vertebrates including reptiles and amphibians. Since iridoviruses from invertebrate and vertebrate hosts differ strongly from one another based not only [...] Read more.
Invertebrate iridoviruses (IIVs), while mostly described in a wide range of invertebrate hosts, have also been repeatedly detected in diagnostic samples from poikilothermic vertebrates including reptiles and amphibians. Since iridoviruses from invertebrate and vertebrate hosts differ strongly from one another based not only on host range but also on molecular characteristics, a series of molecular studies and bioassays were performed to characterize and compare IIVs from various hosts and evaluate their ability to infect a vertebrate host. Eight IIV isolates from reptilian and orthopteran hosts collected over a period of six years were partially sequenced. Comparison of eight genome portions (total over 14 kbp) showed that these were all very similar to one another and to an earlier described cricket IIV isolate, thus they were given the collective name lizard–cricket IV (Liz–CrIV). One isolate from a chameleon was also subjected to Illumina sequencing and almost the entire genomic sequence was obtained. Comparison of this longer genome sequence showed several differences to the most closely related IIV, Invertebrate iridovirus 6 (IIV6), the type species of the genus Iridovirus, including several deletions and possible recombination sites, as well as insertions of genes of non-iridoviral origin. Three isolates from vertebrate and invertebrate hosts were also used for comparative studies on pathogenicity in crickets (Gryllus bimaculatus) at 20 and 30 °C. Finally, the chameleon isolate used for the genome sequencing studies was also used in a transmission study with bearded dragons. The transmission studies showed large variability in virus replication and pathogenicity of the three tested viruses in crickets at the two temperatures. In the infection study with bearded dragons, lizards inoculated with a Liz–CrIV did not become ill, but the virus was detected in numerous tissues by qPCR and was also isolated in cell culture from several tissues. Highest viral loads were measured in the gastro-intestinal organs and in the skin. These studies demonstrate that Liz–CrIV circulates in the pet trade in Europe. This virus is capable of infecting both invertebrates and poikilothermic vertebrates, although its involvement in disease in the latter has not been proven. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Infection of a cricket by dipping into virus suspension. (<b>B</b>) Individual beaker with a cricket used in the infection studies. (<b>C</b>) Beakers were kept in a thermoregulated incubator with lights.</p>
Full article ">Figure 2
<p>Transmission study of IIV in bearded dragons (<span class="html-italic">Pogona vitticeps</span>). (<b>A</b>) Administering virus preparate with an intra-gastric tube. (<b>B</b>) Force feeding with cricket-preparation. Note the bluish iridescence of the cricket preparate at the tip of the forceps.</p>
Full article ">Figure 3
<p>Putative genomic map of the chameleon IIV isolate (Ir. iso. 1) constructed based on nine contigs of an Illumina de novo assembly, and applying the ORF numbering of IIV6. Red dotted rectangles indicate the genome regions, which were earlier covered by a Sanger method based comparison of the isolates (please note that the gaps joining the contigs are not proportional to their actual probable size, but were arbitrarily set to 1 kbp). <span class="html-fig-inline" id="viruses-11-00600-i001"> <img alt="Viruses 11 00600 i001" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i001.png"/></span> ORF highly similiar (85%–100%) to IIV6 homolog. <span class="html-fig-inline" id="viruses-11-00600-i002"> <img alt="Viruses 11 00600 i002" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i002.png"/></span> ORF broken or much shorter (&gt;20%) than IIV6 homolog. <span class="html-fig-inline" id="viruses-11-00600-i003"> <img alt="Viruses 11 00600 i003" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i003.png"/></span> ORF is joined into one ORF from two IIV6 homologs. <span class="html-fig-inline" id="viruses-11-00600-i004"> <img alt="Viruses 11 00600 i004" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i004.png"/></span> ORF most similar to an IIV31 (genus <span class="html-italic">Iridovirus)</span> gene/ORF. <span class="html-fig-inline" id="viruses-11-00600-i005"> <img alt="Viruses 11 00600 i005" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i005.png"/></span> ORF most similar to a polyiridovirus (genus <span class="html-italic">Chlorididovirus</span>) gene. <span class="html-fig-inline" id="viruses-11-00600-i006"> <img alt="Viruses 11 00600 i006" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i006.png"/></span> ORF most similar to a gene from a non-IV large DNA virus. <span class="html-fig-inline" id="viruses-11-00600-i007"> <img alt="Viruses 11 00600 i007" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i007.png"/></span> ORF most similar to a gene from a eukaryotic organism. <span class="html-fig-inline" id="viruses-11-00600-i008"> <img alt="Viruses 11 00600 i008" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i008.png"/></span> ORF most similar to a gene from a prokaryote. <span class="html-fig-inline" id="viruses-11-00600-i009"> <img alt="Viruses 11 00600 i009" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i009.png"/></span> ORF with no homology to current GenBank entries. <span class="html-fig-inline" id="viruses-11-00600-i010"> <img alt="Viruses 11 00600 i010" src="/viruses/viruses-11-00600/article_deploy/html/images/viruses-11-00600-i010.png"/></span> GAP (not yet sequenced part, arbitraty 1000, N’s joining of the mapped contigs).</p>
Full article ">Figure 4
<p>SimPlot analysis of the <span class="html-italic">MCP</span> gene with its flanking regions from our isolates. Three further members of the genus <span class="html-italic">Iridovirus</span> were available for the analysis in GenBank (IIV22 is highly similar to IIV1 and thus was omitted from the figure, for a better overview). Gene homologs with the highest BLAST values are drawn under the diagram. The region downstream of the <span class="html-italic">MCP</span> gene is a non-functional pseudogene homolog of IIV9 <span class="html-italic">ORF011</span>. Exclamation marks indicate stop codons in this pseudogene.</p>
Full article ">Figure 5
<p>Graphic comparison of cricket bioassay results with different isolates at different temperatures. (<b>A</b>) Patent infection rates projected on mortality rates. (<b>B</b>) IIV infection rates detected by different methods.</p>
Full article ">Figure 6
<p>Malformations associated with IIV infection in crickets. (<b>A</b>) Negative control cricket. (<b>B</b>) Inability to complete ecdysis. (<b>C</b>,<b>D</b>) Distorted development of the wings in patently infected crickets. (<b>E,F</b>) Bluish iridescence in the fat body of patently infected crickets (on picture F right side of tube is infected, left is the negative control).</p>
Full article ">
11 pages, 1956 KiB  
Article
Ranaviruses Bind Cells from Different Species through Interaction with Heparan Sulfate
by Fei Ke, Zi-Hao Wang, Cheng-Yue Ming and Qi-Ya Zhang
Viruses 2019, 11(7), 593; https://doi.org/10.3390/v11070593 - 29 Jun 2019
Cited by 14 | Viewed by 2721
Abstract
Ranavirus cross-species infections have been documented, but the viral proteins involved in the interaction with cell receptors have not yet been identified. Here, viral cell-binding proteins and their cognate cellular receptors were investigated using two ranaviruses, Andrias davidianus ranavirus (ADRV) and Rana grylio [...] Read more.
Ranavirus cross-species infections have been documented, but the viral proteins involved in the interaction with cell receptors have not yet been identified. Here, viral cell-binding proteins and their cognate cellular receptors were investigated using two ranaviruses, Andrias davidianus ranavirus (ADRV) and Rana grylio virus (RGV), and two different cell lines, Chinese giant salamander thymus cells (GSTC) and Epithelioma papulosum cyprinid (EPC) cells. The heparan sulfate (HS) analog heparin inhibited plaque formation of ADRV and RGV in the two cell lines by more than 80% at a concentration of 5 μg/mL. In addition, enzymatic removal of cell surface HS by heparinase I markedly reduced plaque formation by both viruses and competition with heparin reduced virus-cell binding. These results indicate that cell surface HS is involved in ADRV and RGV cell binding and infection. Furthermore, recombinant viral envelope proteins ADRV-58L and RGV-53R bound heparin-Sepharose beads implying the potential that cell surface HS is involved in the initial interaction between ranaviruses and susceptible host cells. To our knowledge, this is the first report identifying cell surface HS as ranavirus binding factor and furthers understanding of interactions between ranaviruses and host cells. Full article
Show Figures

Figure 1

Figure 1
<p>Soluble heparin inhibits <span class="html-italic">Andrias davidianus</span> ranavirus (ADRV) and <span class="html-italic">Rana grylio</span> virus (RGV) infection of giant salamander thymus cells (GSTC) and <span class="html-italic">Epithelioma papulosum</span> cyprinid (EPC) cells. Cells were infected with ADRV or RGV that had been pre-incubated in the presence of different concentrations of heparin. The number of plaques obtained in the absence of heparin was set as 1. The data represent triplicate results and was analyzed with Student’s <span class="html-italic">t</span>-test. Significant differences (versus virus without heparin) are marked with * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Heparan sulfate (<b>a</b>) and chondroitin sulfate (<b>b</b>) inhibit ADRV and RGV infection of GSTC and EPC cells. GSTC and EPC cells were infected with ADRV or RGV in the presence of different concentrations of heparan sulfate and chondroitin sulfate. The number of plaques obtained without glycosaminoglycans (GAGs) was set as 1. Experiments were conducted in triplicate and analyzed using Student’s <span class="html-italic">t</span>-test. Significant differences (versus virus without exposure to GAGs) are marked with * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Heparinase I treatment reduced ADRV and RGV plaque formation in GSTC cells. Cells were infected with ADRV or RGV after treatment with different concentrations of heparinase I and plaque formation monitored. Plaque numbers obtained in the absence of heparinase treatment were set as 1. Triplicate results were analyzed by Student’s <span class="html-italic">t</span>-test, and significant differences are marked with * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Soluble heparin inhibits virus binding to GSTC cells. Viral suspensions or purified virions were added to GSTC cells in the presence of different concentrations of heparin. After incubation, virion binding was assessed by determining the number of bound viral genomes by qPCR. DNA levels observed in the absence of heparin pre-treatment were set as 1. The data were obtained from three experiments and analyzed with Student’s <span class="html-italic">t</span>-test. Significant differences are marked with * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Recombinant proteins bind heparin-Sepharose beads. (<b>a</b>) Schematic diagram of the recombinant proteins: The N-terminal domain of ADRV-58L (r58L-N), the N-terminal domain of RGV-53R (r53R-N), and the C-terminal domain of ADRV-58L (r58L-C) were expressed using pET32a or pET28a. The predicted transmembrane region is shown in the grey box. (<b>b</b>) Expression and purification of the three proteins (r58L-N, r53R-N, and r58L-C) with pET32a vector. M: protein marker; 1, 4, 8: Bacteria without induction; 2, 5, 9: Bacteria with induction; 3, 6, 10: Purified proteins. The recombinant proteins are indicated with asterisks, and their predicted molecular weights are shown on the right. (<b>c</b>) Expression and purification of the three proteins using pET28a. M: Protein marker; 1, 5, 10: Cacteria without induction; 2, 6, 11: Cacteria with induction; 3, 4, 7, 8, 12: Purified proteins. The recombinant proteins are indicated with asterisks, and their predicted molecular weights are shown on the right. (<b>d</b>) Binding of recombinant proteins and heparin-Sepharose beads. Recombinant proteins were incubated with heparin-Sepharose or Sepharose beads. The fractions of input (Input), supernatant after incubation (S), the fifth wash solution (W5), and the eluate (Eluate) were detected by Western blot with the anti-His antibody. Recombinant proteins expressed with pET32a or pET28a vectors were used. Recombinant proteins were observed in the Input and Elute fractions from heparin-Sepharose beads and S fraction from control beads.</p>
Full article ">
13 pages, 2481 KiB  
Article
Modelling Ranavirus Transmission in Populations of Common Frogs (Rana temporaria) in the United Kingdom
by Amanda L.J. Duffus, Trenton W.J. Garner, Richard A. Nichols, Joshua P. Standridge and Julia E. Earl
Viruses 2019, 11(6), 556; https://doi.org/10.3390/v11060556 - 15 Jun 2019
Cited by 6 | Viewed by 3870
Abstract
Ranaviruses began emerging in common frogs (Rana temporaria) in the United Kingdom in the late 1980s and early 1990s, causing severe disease and declines in the populations of these animals. Herein, we explored the transmission dynamics of the ranavirus(es) present in [...] Read more.
Ranaviruses began emerging in common frogs (Rana temporaria) in the United Kingdom in the late 1980s and early 1990s, causing severe disease and declines in the populations of these animals. Herein, we explored the transmission dynamics of the ranavirus(es) present in common frog populations, in the context of a simple susceptible-infected (SI) model, using parameters derived from the literature. We explored the effects of disease-induced population decline on the dynamics of the ranavirus. We then extended the model to consider the infection dynamics in populations exposed to both ulcerative and hemorrhagic forms of the ranaviral disease. The preliminary investigation indicated the important interactions between the forms. When the ulcerative form was present in a population and the hemorrhagic form was later introduced, the hemorrhagic form of the disease needed to be highly contagious, to persist. We highlighted the areas where further research and experimental evidence is needed and hope that these models would act as a guide for further research into the amphibian disease dynamics. Full article
Show Figures

Figure 1

Figure 1
<p>Annual cycle of important life history events for common frogs (<span class="html-italic">Rana temporaria</span>) and important events for ranavirus infections and diseases, for these animals. Boxes shaded in blue are those that occur in the aquatic environment, green boxes are those that straddle the land and water, grey boxes occur at an unknown location, and mauve boxes are life history events that are known to happen on both the land and in the water.</p>
Full article ">Figure 2
<p>The interaction between σ and Ψ, under which R<sub>o</sub> ≥ 1, when A<sub>S</sub> = 99 and an M<sub>N</sub> = 20% (upper dashed-curve) and when the initial conditions of A<sub>S</sub> = 49 and an M<sub>N</sub> = 10% (lower curve), where Ψ is the contact rate and σ is the likelihood of transmission for the model.</p>
Full article ">Figure 3
<p>Predicted values of A<sub>S</sub> with varying values of Ψ while other values remained constant at: σ = 0.3; M<sub>N</sub> = 0.2; and M<sub>D</sub> = 0.75. The starting population composition is A<sub>I</sub> = 1 and A<sub>S</sub> = 99 and time is in years. A<sub>S</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate, and M<sub>D</sub> is the mortality rate associated with ranavirosis.</p>
Full article ">Figure 4
<p>The average expectation of the ranavirus dynamics in a population of adult common frogs (<span class="html-italic">Rana temporaria</span>) through time (years). (Ψ = 0.45; σ = 0.3; M<sub>N</sub> = 0.2; M<sub>D</sub> = 0.775; starting population comprised of A<sub>I</sub> = 1 and A<sub>S</sub> = 29.) A<sub>I</sub> is the number of infected individuals, A<sub>S</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate, and M<sub>D</sub> is the mortality rate associated with ranavirosis.</p>
Full article ">Figure 5
<p>Illustration of the predicted values for A<sub>S</sub> with different disease-induced mortality rates, while other values remained constant at: Ψ = 0.45; σ = 0.3; M<sub>N</sub> = 0.2; the starting population comprised of A<sub>I</sub> = 1 and As = 29. A<sub>s</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate and M<sub>D</sub> is the mortality rate associated with ranavirosis.</p>
Full article ">Figure 6
<p>Illustration of the predicted dynamics of a common frog population, with the ranavirus factoring in an annual population decline of 8.1% for adult common frogs. (Ψ = 0.45; σ = 0.3; M<sub>N</sub> = 0.2; M<sub>D</sub> = 0.775; starting population comprised of A<sub>I</sub> = 1 and A<sub>S</sub> = 29; time is in years.) A<sub>s</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate, and M<sub>D</sub> is the mortality rate associated with the ranavirosis.</p>
Full article ">Figure 7
<p>Diagrammatic representations of the transmission dynamics of the ranavirus when only the A<sub>S</sub> or A<sub>H</sub> causing isolate of the ranavirus is present. (<b>A</b>) When only the ulcerative form of the ranavirus is present within the population. (<b>B</b>) When only the hemorrhagic form of the disease is present in the population. All of the variables present are the same as described above and all have a time component associated with them. Where A<sub>s</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate, and M<sub>D</sub> is the mortality rate associated with ranavirosis.</p>
Full article ">Figure 8
<p>Illustration of the complex transmission dynamics of the ranavirus, when both of the observed disease syndromes are present in the population. Dashed lines are used to make the disease syndrome-specific vectors of the transmission easier to follow. The box sizes are not representative of the number of individuals in each category. The order of the boxes does not indicate when the given disease syndrome was introduced. All parameters have time components associated with them. A<sub>s</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate, and M<sub>D</sub> is the mortality rate associated with ranavirosis.</p>
Full article ">Figure 9
<p>R<sub>o</sub> values for the introduction of one A<sub>H</sub> individual into a population of A<sub>S</sub> = 28 and A<sub>U</sub> = 1 (M<sub>D(U)</sub> = M<sub>D(H)</sub> = 0.775, Ψ = 0.45, M<sub>N</sub> = 0.2). A<sub>S</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate, and M<sub>D</sub> is the mortality rate associated with ranavirosis.</p>
Full article ">Figure 10
<p>Ro values for the introduction of one A<sub>H</sub> individual to populations with differing numbers of A<sub>U</sub>, while the total population size remains constant at 30. The number associated with each line indicates the number of A<sub>U</sub> individuals present in the population. (M<sub>D(U)</sub> = 0.775, Ψ = 0.45, M<sub>N</sub> = 0.2) A<sub>s</sub> is the number of susceptible individuals, Ψ is the contact rate, σ is the likelihood of transmission, M<sub>N</sub> is the natural mortality rate, and M<sub>D</sub> is the mortality rate associated with ranavirosis.</p>
Full article ">
10 pages, 783 KiB  
Article
Artemia spp., a Susceptible Host and Vector for Lymphocystis Disease Virus
by Estefania J. Valverde, Alejandro M. Labella, Juan J. Borrego and Dolores Castro
Viruses 2019, 11(6), 506; https://doi.org/10.3390/v11060506 - 1 Jun 2019
Cited by 11 | Viewed by 3035
Abstract
Different developmental stages of Artemia spp. (metanauplii, juveniles and adults) were bath-challenged with two isolates of the Lymphocystis disease virus (LCDV), namely, LCDV SA25 (belonging to the species Lymphocystis disease virus 3) and ATCC VR-342 (an unclassified member of the genus Lymphocystivirus [...] Read more.
Different developmental stages of Artemia spp. (metanauplii, juveniles and adults) were bath-challenged with two isolates of the Lymphocystis disease virus (LCDV), namely, LCDV SA25 (belonging to the species Lymphocystis disease virus 3) and ATCC VR-342 (an unclassified member of the genus Lymphocystivirus). Viral quantification and gene expression were analyzed by qPCR at different times post-inoculation (pi). In addition, infectious titres were determined at 8 dpi by integrated cell culture (ICC)-RT-PCR, an assay that detects viral mRNA in inoculated cell cultures. In LCDV-challenged Artemia, the viral load increased by 2–3 orders of magnitude (depending on developmental stage and viral isolate) during the first 8–12 dpi, with viral titres up to 2.3 × 102 Most Probable Number of Infectious Units (MPNIU)/mg. Viral transcripts were detected in the infected Artemia, relative expression values showed a similar temporal evolution in the different experimental groups. Moreover, gilthead seabream (Sparus aurata) fingerlings were challenged by feeding on LCDV-infected metanauplii. Although no Lymphocystis symptoms were observed in the fish, the number of viral DNA copies was significantly higher at the end of the experimental trial and major capsid protein (mcp) gene expression was consistently detected. The results obtained support that LCDV infects Artemia spp., establishing an asymptomatic productive infection at least under the experimental conditions tested, and that the infected metanauplii are a vector for LCDV transmission to gilthead seabream. Full article
Show Figures

Figure 1

Figure 1
<p>Temporal evolution of viral loads (<b>A</b>) and relative major capsid protein (<span class="html-italic">mcp</span>) gene expression values (<b>B</b>) in different developmental stages of <span class="html-italic">Artemia</span> inoculated with Lymphocystis disease virus (LCDV) SA25.</p>
Full article ">Figure 2
<p>Temporal evolution of viral loads (<b>A</b>) and relative <span class="html-italic">mcp</span> gene expression values (<b>B</b>) in <span class="html-italic">Artemia</span> metanauplii inoculated with LCDV ATCC VR-342.</p>
Full article ">Figure 3
<p>Viral loads (<b>A</b>) and relative <span class="html-italic">mcp</span> gene expression values (<b>B</b>) in gilthead seabream fingerlings orally challenged with LCDV-positive <span class="html-italic">Artemia</span> metanauplii (mean ± standard deviation; <span class="html-italic">n</span> = 7). Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.01) (Mann–Whitney U-test, Holm–Bonferroni correction).</p>
Full article ">
10 pages, 2382 KiB  
Article
Characterization of a Novel Megalocytivirus Isolated from European Chub (Squalius cephalus)
by Maya A. Halaly, Kuttichantran Subramaniam, Samantha A. Koda, Vsevolod L. Popov, David Stone, Keith Way and Thomas B. Waltzek
Viruses 2019, 11(5), 440; https://doi.org/10.3390/v11050440 - 15 May 2019
Cited by 14 | Viewed by 4366
Abstract
A novel virus from moribund European chub (Squalius cephalus) was isolated on epithelioma papulosum cyprini (EPC) cells. Transmission electron microscopic examination revealed abundant non-enveloped, hexagonal virus particles in the cytoplasm of infected EPC cells consistent with an iridovirus. Illumina MiSeq sequence [...] Read more.
A novel virus from moribund European chub (Squalius cephalus) was isolated on epithelioma papulosum cyprini (EPC) cells. Transmission electron microscopic examination revealed abundant non-enveloped, hexagonal virus particles in the cytoplasm of infected EPC cells consistent with an iridovirus. Illumina MiSeq sequence data enabled the assembly and annotation of the full genome (128,216 bp encoding 108 open reading frames) of the suspected iridovirus. Maximum Likelihood phylogenetic analyses based on 25 iridovirus core genes supported the European chub iridovirus (ECIV) as being the sister species to the recently-discovered scale drop disease virus (SDDV), which together form the most basal megalocytivirus clade. Genetic analyses of the ECIV major capsid protein and ATPase genes revealed the greatest nucleotide identity to members of the genus Megalocytivirus including SDDV. These data support ECIV as a novel member within the genus Megalocytivirus. Experimental challenge studies are needed to fulfill River’s postulates and determine whether ECIV induces the pathognomonic microscopic lesions (i.e., megalocytes with basophilic cytoplasmic inclusions) observed in megalocytivirus infections. Full article
Show Figures

Figure 1

Figure 1
<p>Microscopic examination of <span class="html-italic">epithelioma papulosum cyprini</span> cells infected with European chub iridovirus. (<b>A</b>) Control flask at 48 h post-inoculation (hpi); (<b>B</b>) control flask 96 hpi; (<b>C</b>) infected flask showing enlarged and refractile cells at 48 hpi; (<b>D</b>) infected flask showing enlarged and refractile cells at 96 hpi. Scale bars are 50 µm.</p>
Full article ">Figure 2
<p>(<b>A</b>) Transmission electron photomicrograph of an <span class="html-italic">epithelioma papulosum cyprini</span> cell infected with European chub iridovirus, displaying numerous non-enveloped, hexagonal viral particles within the viral assembly site (labeled as V) in the cytoplasm. Scale bar is 1 µm. (<b>B</b>) Higher magnification of the virus particles. Scale bar is 250 nm.</p>
Full article ">Figure 3
<p>Cladogram depicting the relationship of European chub iridovirus to 47 other members of the family <span class="html-italic">Iridoviridae</span> based on 25 core genes. The Maximum Likelihood tree was generated using 1000 bootstraps and the branch lengths are based on the number of inferred substitutions, as indicated by the scale. All nodes were supported by bootstrap values &gt;80% except those labeled with black circles. See <a href="#viruses-11-00440-t001" class="html-table">Table 1</a> for virus abbreviations.</p>
Full article ">
13 pages, 4459 KiB  
Article
Interaction between Two Iridovirus Core Proteins and Their Effects on Ranavirus (RGV) Replication in Cells from Different Species
by Xiao-Tao Zeng and Qi-Ya Zhang
Viruses 2019, 11(5), 416; https://doi.org/10.3390/v11050416 - 4 May 2019
Cited by 12 | Viewed by 3313
Abstract
The two putative proteins RGV-63R and RGV-91R encoded by Rana grylio virus (RGV) are DNA polymerase and proliferating cell nuclear antigen (PCNA) respectively, and are core proteins of iridoviruses. Here, the interaction between RGV-63R and RGV-91R was detected by a yeast two-hybrid (Y2H) [...] Read more.
The two putative proteins RGV-63R and RGV-91R encoded by Rana grylio virus (RGV) are DNA polymerase and proliferating cell nuclear antigen (PCNA) respectively, and are core proteins of iridoviruses. Here, the interaction between RGV-63R and RGV-91R was detected by a yeast two-hybrid (Y2H) assay and further confirmed by co-immunoprecipitation (co-IP) assays. Subsequently, RGV-63R or RGV-91R were expressed alone or co-expressed in two kinds of aquatic animal cells including amphibian Chinese giant salamander thymus cells (GSTCs) and fish Epithelioma papulosum cyprinid cells (EPCs) to investigate their localizations and effects on RGV genome replication. The results showed that their localizations in the two kinds of cells are consistent. RGV-63R localized in the cytoplasm, while RGV-91R localized in the nucleus. However, when co-expressed, RGV-63R localized in both the cytoplasm and the nucleus, and colocalized with RGV-91R in the nucleus. 91R△NLS represents the RGV-91R deleting nuclear localization signal, which is localized in the cytoplasm and colocalized with RGV-63R in the cytoplasm. qPCR analysis revealed that sole expression and co-expression of the two proteins in the cells of two species significantly promoted RGV genome replication, while varying degrees of viral genome replication levels may be linked to the cell types. This study provides novel molecular evidence for ranavirus cross-species infection and replication. Full article
Show Figures

Figure 1

Figure 1
<p>The formed yeast colonies during testing interactions between RGV-63R and four proteins by Y2H. Three parallel experiments were performed. 2Q: SD/-Trp/-Leu. 3Q/X: SD/-Trp/-Leu/-His/X-α-Gal. 91R, 97R, 98R, and 102R: RGV proteins, which are encoded by iridovirus core genes, respectively. T+P53: the positive control; T+Lam: the negative control.</p>
Full article ">Figure 2
<p>Western blot analysis for sample of testing interaction between RGV-63R and RGV-91R by co-IP. Cell lysates from HEK293T cells cotransfected with indicated plasmids (<span class="html-italic">91R-HA+ pcDNA3.1</span>, <span class="html-italic">91R-HA+ 63R</span>-<span class="html-italic">3Fla</span>g) and IP (immunoprecipitated) protein complexes with indicated plasmids (<span class="html-italic">91R-HA+ pcDNA3.1</span>, <span class="html-italic">91R-HA+ 63R</span>-<span class="html-italic">3Fla</span>g) are subjected to Western blot analysis using anti-HA and anti-Flag. Cells lysates and IP showed the bands of 91R-HA and 63R-3Flag. M: protein molecular mass marker.</p>
Full article ">Figure 3
<p>Fluorescence micrographs of cells expressing a single protein or co-expressing two proteins. (<b>A</b>) Expressing proteins 63R-EGFP, 91R-RFP, and 91R△NLS1-RFP alone in giant salamander thymus cells (GSTCs) or <span class="html-italic">Epithelioma papulosum cyprinid</span> cells (EPCs), respectively. (<b>B</b>) Co-expressing two proteins, 63R-EGFP + 91R-RFP and 63R-EGFP + 91R△NLS1-RFP, in GSTCs or EPCs, respectively. 63R-EGFP (green), 91R-RFP (red), 91R△NLS1-RFP (red), nucleus (blue), and colocalization (yellow). Scale bar: 10 μm.</p>
Full article ">Figure 3 Cont.
<p>Fluorescence micrographs of cells expressing a single protein or co-expressing two proteins. (<b>A</b>) Expressing proteins 63R-EGFP, 91R-RFP, and 91R△NLS1-RFP alone in giant salamander thymus cells (GSTCs) or <span class="html-italic">Epithelioma papulosum cyprinid</span> cells (EPCs), respectively. (<b>B</b>) Co-expressing two proteins, 63R-EGFP + 91R-RFP and 63R-EGFP + 91R△NLS1-RFP, in GSTCs or EPCs, respectively. 63R-EGFP (green), 91R-RFP (red), 91R△NLS1-RFP (red), nucleus (blue), and colocalization (yellow). Scale bar: 10 μm.</p>
Full article ">Figure 4
<p>Western blot analysis of protein samples from cells transfected with plasmids for 24 h using anti-Flag and anti-HA. GSTC: in GSTCs expressing protein RGV-63R (63R + pcDNA3.1) or RGV-91R (91R + pcDNA3.1) alone or co-expressing two proteins RGV-63R and RGV-91R (63R + 91R), respectively. EPC: in EPCs expressing protein RGV-63R (63R + pcDNA3.1) or RGV-91R (91R + pcDNA3.1) alone or co-expressing two proteins RGV-63R and RGV-91R (63R + 91R), respectively. The empty vector pcDNA3.1 was used as a control. M: protein molecular mass marker.</p>
Full article ">Figure 5
<p>qPCR analysis of RGV genomic copies in GSTCs or EPCs. GSTC: in GSTCs expressing protein RGV-63R (63R + pcDNA3.1) or RGV-91R (91R + pcDNA3.1) alone or co-expressing two proteins RGV-63R and RGV-91R (63R + 91R), respectively. EPC: in EPCs expressing protein RGV-63R (63R + pcDNA3.1) or RGV-91R (91R + pcDNA3.1) alone or co-expressing two proteins RGV-63R and RGV-91R (63R + 91R), respectively. The empty vector pcDNA3.1 was used as a control. The transfected cells were infected with RGV, and RGV genome DNA was extracted from the cells at 48 hpi and quantified by qPCR. Each data point represents the average value of three independent infections. Error bars indicate standard deviations. ** represents <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
15 pages, 2213 KiB  
Article
IIV-6 Inhibits NF-κB Responses in Drosophila
by Cara West, Florentina Rus, Ying Chen, Anni Kleino, Monique Gangloff, Don B. Gammon and Neal Silverman
Viruses 2019, 11(5), 409; https://doi.org/10.3390/v11050409 - 1 May 2019
Cited by 10 | Viewed by 5101
Abstract
The host immune response and virus-encoded immune evasion proteins pose constant, mutual selective pressure on each other. Virally encoded immune evasion proteins also indicate which host pathways must be inhibited to allow for viral replication. Here, we show that IIV-6 is capable of [...] Read more.
The host immune response and virus-encoded immune evasion proteins pose constant, mutual selective pressure on each other. Virally encoded immune evasion proteins also indicate which host pathways must be inhibited to allow for viral replication. Here, we show that IIV-6 is capable of inhibiting the two Drosophila NF-κB signaling pathways, Imd and Toll. Antimicrobial peptide (AMP) gene induction downstream of either pathway is suppressed when cells infected with IIV-6 are also stimulated with Toll or Imd ligands. We find that cleavage of both Imd and Relish, as well as Relish nuclear translocation, three key points in Imd signal transduction, occur in IIV-6 infected cells, indicating that the mechanism of viral inhibition is farther downstream, at the level of Relish promoter binding or transcriptional activation. Additionally, flies co-infected with both IIV-6 and the Gram-negative bacterium, Erwinia carotovora carotovora, succumb to infection more rapidly than flies singly infected with either the virus or the bacterium. These findings demonstrate how pre-existing infections can have a dramatic and negative effect on secondary infections, and establish a Drosophila model to study confection susceptibility. Full article
Show Figures

Figure 1

Figure 1
<p>IIV-6 inhibits Imd and Toll Signaling. (<b>A</b>) S2* cells were treated with 20-hydroxyecdysone (EcD) as indicated for 18 hours and then infected with IIV-6 (blue circles) or uninfected (black circles) for six hours. Cells were then stimulated with DAP-type PGN for six hours, where indicated. <span class="html-italic">Diptericin</span> (<span class="html-italic">Dpt</span>) levels were monitored by qRT-PCR. (<b>B</b>) S2* cells were treated with 20-hydroxyecdysone (EcD) as indicated for 18 h and then infected with IIV-6 (blue circles) or uninfected (black circles) for six hours. Cells were then stimulated with cleaved Spätzle (spz) for 18 h, where indicated. <span class="html-italic">Drosomycin</span> (<span class="html-italic">Drs</span>) levels were monitored by qRT-PCR. (<b>A</b>,<b>B</b>) Black bars indicate mean and error bars indicate standard deviation. Statistics were determined using two-way ANOVA and Sidak’s multiple comparisons test; ns, not significant; ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Both Imd and Toll regulated AMPs are suppressed by IIV-6 infection. S2* cells were treated with 20-hydroxyecdysone (EcD) for 18 h, and then infected with IIV-6 for 6 h. (<b>A</b>) Cells were stimulated with peptidoglycan (PGN) for 6 h prior to RNA isolation. (<b>B</b>) S2* cells were then stimulated with cleaved Spätzle (spz) for 18 h prior to RNA isolation, and then analyzed by Nanostring nCounter. Heatmaps display Z scores of mRNA levels of immune genes in the presence or absence of virus and pathway stimulation clustered by expression pattern. Biologically independent duplicates are shown. Untr., untreated cells. EcD, cells treated with 20-hydroxyecdysone. EcD + PGN, cells treated with 20-hydroxyecdysone and peptidoglycan. EcD + spz, cells treated with 20-hydroxyecdysone and Spätzle.</p>
Full article ">Figure 3
<p>Some AMPs are elevated in vivo upon IIV-6 infection, before returning to baseline. (<b>A</b>) Heatmap of Z score transformed mRNA levels for AMP genes following IIV-6 infection of adult male <span class="html-italic">w</span><sup>1118</sup> flies for the indicated timepoints, assayed by Nanostring nCounter. RNA was isolated from PBS-injected flies at the same time points as a control. Biologically independent samples were analyzed in duplicate. (<b>A’</b>) Detailed comparison of mRNA levels for selected Imd-regulated AMP genes following IIV-6 infection of adult <span class="html-italic">w</span><sup>1118</sup> flies for 12, 24, and 48 h assayed by Nanostring nCounter.</p>
Full article ">Figure 4
<p>Imd signaling remains intact in the presence of IIV-6. (<b>A</b>) S2* cells were treated with 20-hydroxyecdysone (EcD) for 18 h. Cells were then infected with IIV-6, where indicated, for six hours. Samples were then stimulated with PGN for 15 min, where indicated, and lysed in standard lysis buffer. Endogenous Imd was monitored by immunoblotting; arrows indicate cleaved (c-Imd) or full-length (FL-Imd) Imd; ns, non-specific. (<b>B</b>) Cells were treated with 20-hydroxyecdysone (EcD), where indicated, for 18 h and infected with IIV-6, as indicated, for six hours. Samples stimulated with EcD were then stimulated with PGN for 15 min, and lysed in standard lysis buffer. Endogenous Relish was probed by immunoblotting using a C-terminal Relish antibody. MOI used was as follows. Lane 2: MOI =2; Lane 4: MOI = 0.2; Lane 5: MOI = 2; Lane 6: MOI = 5.</p>
Full article ">Figure 5
<p>Nuclear localization of Relish remains intact in the presence of IIV-6. (<b>A</b>) S2* cells stably expressing YFP-Relish were treated with 20-hydroxyecdysone (EcD) for 2 h and then infected with mCherry-IIV-6 (right) or left uninfected (left) for 18 h. Cells were then stimulated with DAP-type PGN for 15 min (lower panels), prior to fixation and staining with anti-Lamin (shown in green) and Hoechst 33342 (shown in blue). Representative images from four biologically independent experiments are shown. Arrowheads mark cells with nuclear localized Relish. Left panels display the overlay of YFP-Relish, Hoeschst, Lamin, and mCherry-IIV-6 where applicable, while the right panels exhibit the YFP-Relish alone. Scale bars are 10 μm. (<b>B</b>) Quantification of Relish nuclear translocation. Between 400–1800 cells were scored for each condition as displaying Relish localization in the nucleus, in the cytoplasm, or as indeterminate when staining was diffuse throughout both compartments or the signal was too weak to discern. Statistics were determined using two-way ANOVA and Tukey’s multiple comparisons test; *, <span class="html-italic">p</span> &lt; 0.05; ns, not significant. Error bars represent standard deviation.</p>
Full article ">Figure 6
<p>Viral replication is not needed for NF-κB inhibition. (<b>A</b>,<b>B</b>) S2* cells were treated with 20-hydroxyecdysone (EcD) as indicated for 18 h and then infected with IIV-6 (blue circles) or uninfected (black circles) for six hours. Cells were then stimulated with DAP-type PGN for six hours, where indicated. <span class="html-italic">Diptericin</span> levels were monitored by qRT-PCR. (<b>A</b>) Cells were treated with cidofovir, a viral polymerase inhibitor, where indicated. Mock cells untreated with EcD, PGN, or cidofovir are shown as a control. Each data point is a biologically independent replicate; <span class="html-italic">n</span> = 3. Error bars represent standard deviation. Statistics were determined using two-way ANOVA and Sidak’s multiple comparisons test; ns, not significant; ****, <span class="html-italic">p</span> &lt; 0.0001. (<b>B</b>) S2* cells were infected with IIV-6 (blue circles), or treated with UV- (purple circles) or heat- (red circles) inactivated IIV-6. Uninfected controls are shown in black. Each data point is a biologically independent replicate; <span class="html-italic">n</span> = 3. Error bars represent standard deviation. Statistics were determined using one-way ANOVA and Tukey’s multiple comparisons test; ns, not significant; ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Flies infected with IIV-6 have lower AMP levels. (<b>A</b>) Adult <span class="html-italic">w</span><sup>1118</sup> flies were infected with IIV-6, injected with PBS, or left unmanipulated for 7 days, and then snap frozen for RNA isolation. <span class="html-italic">Dpt</span> levels were assayed by qRT-PCR. Black bars indicate mean and error bars indicate standard deviation. Statistics were determined using one-way ANOVA and Tukey’s multiple comparisons test; ns, not significant; *, <span class="html-italic">p</span> = 0.0463; **, <span class="html-italic">p</span> = 0.0033. (<b>B</b>) Kaplan-Meier plots of adult <span class="html-italic">w</span><sup>1118</sup> flies infected with IIV-6 (blue lines) or PBS-injected (black lines) for 8 days prior to bacterial infection. On day 0, flies were pricked with <span class="html-italic">Ecc15</span> (solid lines) or were sterile pricked (dashed lines) with a microsurgery needle. Flies were counted daily for survivors. Statistics were determined using log-rank test; ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
12 pages, 355 KiB  
Article
Evaluating the Within-Host Dynamics of Ranavirus Infection with Mechanistic Disease Models and Experimental Data
by Joseph R. Mihaljevic, Amy L. Greer and Jesse L. Brunner
Viruses 2019, 11(5), 396; https://doi.org/10.3390/v11050396 - 27 Apr 2019
Cited by 3 | Viewed by 3294
Abstract
Mechanistic models are critical for our understanding of both within-host dynamics (i.e., pathogen replication and immune system processes) and among-host dynamics (i.e., transmission). Within-host models, however, are not often fit to experimental data, which can serve as a robust method of hypothesis testing [...] Read more.
Mechanistic models are critical for our understanding of both within-host dynamics (i.e., pathogen replication and immune system processes) and among-host dynamics (i.e., transmission). Within-host models, however, are not often fit to experimental data, which can serve as a robust method of hypothesis testing and hypothesis generation. In this study, we use mechanistic models and empirical, time-series data of viral titer to better understand the replication of ranaviruses within their amphibian hosts and the immune dynamics that limit viral replication. Specifically, we fit a suite of potential models to our data, where each model represents a hypothesis about the interactions between viral replication and immune defense. Through formal model comparison, we find a parsimonious model that captures key features of our time-series data: The viral titer rises and falls through time, likely due to an immune system response, and that the initial viral dosage affects both the peak viral titer and the timing of the peak. Importantly, our model makes several predictions, including the existence of long-term viral infections, which can be validated in future studies. Full article
Show Figures

Figure 1

Figure 1
<p>The fit of model B2 to the experimental data. Circles are data points representing the viral DNA copies from individual bullfrog tadpoles that were sampled on a given day. The median model fit (solid red line) and 95% Bayesian credible interval (CI) of the fit (dashed red lines) are shown. Additionally, the median (dashed vertical line) and 95% CI (light red polygon) are shown for the time of the maximum viral titer predicted by the model.</p>
Full article ">
15 pages, 3646 KiB  
Article
Susceptibility of Exopalaemon carinicauda to the Infection with Shrimp Hemocyte Iridescent Virus (SHIV 20141215), a Strain of Decapod Iridescent Virus 1 (DIV1)
by Xing Chen, Liang Qiu, Hailiang Wang, Peizhuo Zou, Xuan Dong, Fuhua Li and Jie Huang
Viruses 2019, 11(4), 387; https://doi.org/10.3390/v11040387 - 25 Apr 2019
Cited by 53 | Viewed by 6819
Abstract
In this study, ridgetail white prawns—Exopalaemon carinicauda—were infected per os (PO) with debris of Penaeus vannamei infected with shrimp hemocyte iridescent virus (SHIV 20141215), a strain of decapod iridescent virus 1 (DIV1), and via intramuscular injection (IM with raw extracts of [...] Read more.
In this study, ridgetail white prawns—Exopalaemon carinicauda—were infected per os (PO) with debris of Penaeus vannamei infected with shrimp hemocyte iridescent virus (SHIV 20141215), a strain of decapod iridescent virus 1 (DIV1), and via intramuscular injection (IM with raw extracts of SHIV 20141215. The infected E. carinicauda showed obvious clinical symptoms, including weakness, empty gut and stomach, pale hepatopancreas, and partial death with mean cumulative mortalities of 42.5% and 70.8% by nonlinear regression, respectively. Results of TaqMan probe-based real-time quantitative PCR showed that the moribund and surviving individuals with clinical signs of infected E. carinicauda were DIV1-positive. Histological examination showed that there were darkly eosinophilic and cytoplasmic inclusions, of which some were surrounded with or contained tiny basophilic staining, and pyknosis in hemocytes in hepatopancreatic sinus, hematopoietic cells, cuticular epithelium, etc. On the slides of in situ DIG-labeling-loop-mediated DNA amplification (ISDL), positive signals were observed in hematopoietic tissue, stomach, cuticular epithelium, and hepatopancreatic sinus of infected prawns from both PO and IM groups. Transmission electron microscopy (TEM) of ultrathin sections showed that icosahedral DIV1 particles existed in hepatopancreatic sinus and gills of the infected E. carinicauda from the PO group. The viral particles were also observed in hepatopancreatic sinus, gills, pereiopods, muscles, and uropods of the infected E. carinicauda from the IM group. The assembled virions, which mostly distributed along the edge of the cytoplasmic virogenic stromata near cellular membrane of infected cells, were enveloped and approximately 150 nm in diameter. The results of molecular tests, histopathological examination, ISDL, and TEM confirmed that E. carinicauda is a susceptible host of DIV1. This study also indicated that E. carinicauda showed some degree of tolerance to the infection with DIV1 per os challenge mimicking natural pathway. Full article
Show Figures

Figure 1

Figure 1
<p>Gross signs of prawns <span class="html-italic">Exopalaemon carinicauda</span> in different groups of the challenge test. CPO: Prawn in per os control group; PO: Prawn in per os group; IM: Prawn in intermuscular injection group. Solid arrows indicate stomach (ST), hepatopancreas (HP), midgut (MG), and hematopoietic tissue (HM). Red bar = 10 mm.</p>
Full article ">Figure 2
<p>Cumulative mortalities of <span class="html-italic">Exopalaemon carinicauda</span> in the challenge test. IM group, prawns were challenged with filtrated viral suspension via intermuscular injection; PO group, prawns were fed with tissues of DIV1-infected <span class="html-italic">P. vannamei</span>; CIM group, prawns were injected with sterile PPB-His buffer; CPO group, prawns were fed with commercial feed. Cumulative mortalities are shown as means of data from three replicates for each experimental group (each replicate contained 10 individuals). The mean points with same color indicate no significant difference (<span class="html-italic">P</span> &gt; 0.05), and the mean points with different colors indicate a significant difference (<span class="html-italic">P</span> &lt; 0.05). Overall analysis indicated there are very significant differences among IM, PO, and the two controls (<span class="html-italic">P</span> &lt; 0.01). The curves were drawn based on the nonlinear regression following the three-parameter sigmoid equation.</p>
Full article ">Figure 3
<p>Decapod iridescent virus 1 (DIV1) loads in <span class="html-italic">Exopalaemon carinicauda</span> from the challenge test. Different letters above the bars indicate significant difference (<span class="html-italic">P</span> &lt; 0.01). IM: prawn group challenged var intramuscular injection, PO: prawn group challenged per os; CIM: control group of injected with sterile buffer; CPO: control group fed with commercial feed. CIM and CPO were negative.</p>
Full article ">Figure 4
<p>Histopathological examination of <span class="html-italic">Exopalaemon carinicauda</span> tissues infected with DIV1 and controls. Black arrows show karyopyknosis and white arrows show eosinophilic inclusions. Pictures (<b>a</b>,<b>b</b>) are hepatopancreas; (<b>c</b>,<b>d</b>) are hematopoietic tissues; and (<b>e</b>,<b>f</b>) are cuticular epithelium on which the cuticles were removed before dehydration. The pictures in the left column (a, c, and e) are the tissues of the infected prawn; the pictures in the right column (b, d, and f) are the tissues of the control prawn. HpT: hepatopancreatic tubule; HpS: hepatopancreatic sinus; Hm: hematopoietic tissue; Ep: epithelium; Cn: connective tissue; m: mitotic phase. Bar = 50 µm.</p>
Full article ">Figure 5
<p>Information on primer design and primers used for Loop-mediated isothermal amplification (LAMP), based on the reference sequence of DIV1. Sequences of Shrimp hemocyte iridescent virus (SHIV 20141215 MF599468) and <span class="html-italic">Cherax quadricarinatus</span> iridovirus (CQIV CN01 MF197913) were obtained from GenBank.</p>
Full article ">Figure 6
<p>In situ DIG-labeling-loop-mediated DNA amplification (ISDL) micrographs of DIV1-infected and control <span class="html-italic">Exopalaemon carinicauda</span>. Hepatopancreas, stomach, cuticular epithelium, and hematopoietic tissue from a challenged prawn of the PO group, respectively (<b>a</b>, <b>c</b>, <b>e</b>, and <b>g</b>); hepatopancreas, stomach, cuticular epithelium, and hematopoietic tissue from a healthy prawn of the CPO group, respectively (<b>b</b>, <b>d</b>, <b>f</b>, and <b>h</b>). In pictures a, c, e, and g, blue-violet signals were observed in hepatopancreatic sinus, stomach epithelium, cuticular epithelium, and hematopoietic tissues of prawns, respectively. In pictures b, d, f, and h, no hybridization signal was seen in the same tissues of DIV1-negative <span class="html-italic">E. carinicauda</span>, except some non-specific signals on gastric sieve of stomach.</p>
Full article ">Figure 7
<p>Transmission electron microscopy (TEM) of DIV1-infected <span class="html-italic">Exopalaemon carinicauda</span>. (<b>A</b>–<b>E</b>): hepatopancreatic sinus, pereiopods, uropods, muscle and gills, respectively, of a prawn in the IM group; (<b>F</b> and <b>G</b>): hepatopancreatic sinus and gills, respectively, of a prawn in the PO group; (<b>H</b>): cytoplasmic inclusions with a cluster of viral particles.</p>
Full article ">
16 pages, 3506 KiB  
Article
Distribution and Phylogeny of Erythrocytic Necrosis Virus (ENV) in Salmon Suggests Marine Origin
by Veronica A. Pagowski, Gideon J. Mordecai, Kristina M. Miller, Angela D. Schulze, Karia H. Kaukinen, Tobi J. Ming, Shaorong Li, Amy K. Teffer, Amy Tabata and Curtis A. Suttle
Viruses 2019, 11(4), 358; https://doi.org/10.3390/v11040358 - 18 Apr 2019
Cited by 8 | Viewed by 4139
Abstract
Viral erythrocytic necrosis (VEN) affects over 20 species of marine and anadromous fishes in the North Atlantic and North Pacific Oceans. However, the distribution and strain variation of its viral causative agent, erythrocytic necrosis virus (ENV), has not been well characterized within Pacific [...] Read more.
Viral erythrocytic necrosis (VEN) affects over 20 species of marine and anadromous fishes in the North Atlantic and North Pacific Oceans. However, the distribution and strain variation of its viral causative agent, erythrocytic necrosis virus (ENV), has not been well characterized within Pacific salmon. Here, metatranscriptomic sequencing of Chinook salmon revealed that ENV infecting salmon was closely related to ENV from Pacific herring, with inferred amino-acid sequences from Chinook salmon being 99% identical to those reported for herring. Sequence analysis also revealed 89 protein-encoding sequences attributed to ENV, greatly expanding the amount of genetic information available for this virus. High-throughput PCR of over 19,000 fish showed that ENV is widely distributed in the NE Pacific Ocean and was detected in 12 of 16 tested species, including in 27% of herring, 38% of anchovy, 17% of pollock, and 13% of sand lance. Despite frequent detection in marine fish, ENV prevalence was significantly lower in fish from freshwater (0.03%), as assessed with a generalized linear mixed effects model (p = 5.5 × 10−8). Thus, marine fish are likely a reservoir for the virus. High genetic similarity between ENV obtained from salmon and herring also suggests that transmission between these hosts is likely. Full article
Show Figures

Figure 1

Figure 1
<p>Substitution-rate optimized maximum likelihood phylogenetic trees of putative ENV DNA-dependent DNA polymerase (<b>A</b>), DNA-dependent RNA polymerase (<b>B</b>), major capsid protein (<b>C</b>), and ATPase (<b>D</b>) sequences (SEQ#) mapped with related iridoviruses. Alignments were based on nucleotide sequences for the major capsid protein and amino acid sequences for all other trees. Branch labels indicate bootstrap support values for 100 re-samplings and the scale bar indicates substitution rate. GenBank reference sequences and contig IDs are listed below (<a href="#viruses-11-00358-t002" class="html-table">Table 2</a>), with colors indicating genera groups. Putative ENV sequence lengths are listed in <a href="#viruses-11-00358-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>Apparent ENV prevalence in mixed tissue smolt samples by species. Error bars indicate 95% confidence intervals. Blue bars indicate proportions with limit of detection (LOD) criteria applied and coral bars indicate proportions without LOD criteria applied. Values indicate sample sizes for each species.</p>
Full article ">Figure 3
<p>(<b>A</b>) Full sampling area heat map of the mean calculated ENV copy number, based on interpolated values. Copy number values are binned into 6 numerical categories (see color legend), (<b>B</b>) shows Vancouver Island inset map. Adults and smolts of all species are shown and LOD criteria are not applied.</p>
Full article ">Figure 4
<p>Saline (SW) and freshwater (FW) ENV prevalence (<b>A</b>) and load (<b>B</b>) among salmon species and herring. Samples with LOD criteria applied are in blue and samples without LOD criteria applied are in coral. Printed values indicate sample sizes and error bars indicate 95% confidence intervals. Only smolts are shown.</p>
Full article ">
14 pages, 3509 KiB  
Article
Description of a Natural Infection with Decapod Iridescent Virus 1 in Farmed Giant Freshwater Prawn, Macrobrachium rosenbergii
by Liang Qiu, Xing Chen, Ruo-Heng Zhao, Chen Li, Wen Gao, Qing-Li Zhang and Jie Huang
Viruses 2019, 11(4), 354; https://doi.org/10.3390/v11040354 - 17 Apr 2019
Cited by 72 | Viewed by 8946
Abstract
Macrobrachium rosenbergii is a valuable freshwater prawn in Asian aquaculture. In recent years, a new symptom that was generally called “white head” has caused high mortality in M. rosenbergii farms in China. Samples of M. rosenbergii, M. nipponense, Procambarus clarkii, [...] Read more.
Macrobrachium rosenbergii is a valuable freshwater prawn in Asian aquaculture. In recent years, a new symptom that was generally called “white head” has caused high mortality in M. rosenbergii farms in China. Samples of M. rosenbergii, M. nipponense, Procambarus clarkii, M. superbum, Penaeus vannamei, and Cladocera from a farm suffering from white head in Jiangsu Province were collected and analyzed in this study. Pathogen detection showed that all samples were positive for Decapod iridescent virus 1 (DIV1). Histopathological examination revealed dark eosinophilic inclusions and pyknosis in hematopoietic tissue, hepatopancreas, and gills of M. rosenbergii and M. nipponense. Blue signals of in situ digoxigenin-labeled loop-mediated isothermal amplification appeared in hematopoietic tissue, hemocytes, hepatopancreatic sinus, and antennal gland. Transmission electron microscopy of ultrathin sections showed a large number of DIV1 particles with a mean diameter about 157.9 nm. The virogenic stromata and budding virions were observed in hematopoietic cells. Quantitative detection with TaqMan probe based real-time PCR of different tissues in naturally infected M. rosenbergii showed that hematopoietic tissue contained the highest DIV1 load with a relative abundance of 25.4 ± 16.9%. Hepatopancreas and muscle contained the lowest DIV1 loads with relative abundances of 2.44 ± 1.24% and 2.44 ± 2.16%, respectively. The above results verified that DIV1 is the pathogen causing white head in M. rosenbergii. M. nipponense and Pr. clarkii are also species susceptible to DIV1. Full article
Show Figures

Figure 1

Figure 1
<p>Clinical symptoms of <span class="html-italic">M. rosenbergii</span> (20180620) naturally infected with DIV1. (<b>A</b>) Overall appearance of a diseased prawn in water. (<b>B</b>) Close-up of cephalothoraxes. Blue arrows show white area under the cuticle at the base of rostrum. White arrows indicate hepatopancreas atrophy, color fading, and yellowing.</p>
Full article ">Figure 2
<p>Histopathological features of Davidson’s alcohol-formalin-acetic acid fixative (DAFA) fixed <span class="html-italic">M. rosenbergii</span> (<b>A</b>–<b>C</b>) and <span class="html-italic">M. nipponense</span> (<b>D</b>) samples 20180620. White arrows show the eosinophilic inclusions and black arrows show the karyopyknotic nuclei. (<b>A</b>) Hematoxylin and eosin (H&amp;E) staining of the hematopoietic tissue. (<b>B</b>,<b>D</b>) H&amp;E staining of hepatopancreas. (<b>C</b>) H&amp;E staining of gills. Bar, 20 μm (<b>A</b>,<b>B</b>), 50 μm (<b>C</b>), and 10 μm (<b>D</b>), respectively.</p>
Full article ">Figure 3
<p>In situ digoxigenin-labeled loop-mediated isothermal amplification (ISDL) targeting the gene of the second largest subunit of DNA-directed RNA polymerase II of DIV1 on histological sections of <span class="html-italic">M. rosenbergii</span> (<b>A</b>–<b>F</b>), <span class="html-italic">M. nipponense</span> (<b>G</b>), and <span class="html-italic">Pr. clarkii</span> (<b>H</b>,<b>I</b>) samples 20180620. (<b>A</b>,<b>H</b>) Hematopoietic tissue; (<b>B</b>,<b>D</b>,<b>G</b>,<b>I</b>) hepatopancreas; (<b>C)</b> gills; (<b>E</b>) antennal gland; (<b>F</b>) ovaries. In (<b>A</b>–<b>C</b>,<b>H</b>), blue signals were observed in hematopoietic tissue, hemocytes in the sinus of the hepatopancreas, and in gills. In (<b>D</b>,<b>G</b>,<b>I</b>), blue signals exist in some hepatopancreatic R-cells and myoepithelial fibers. In (<b>E</b>), blue signals exist in the coelomosac epithelium. In (<b>F</b>), blue signals exist in the epithelium. Bar, 20 µm (<b>A</b>,<b>D</b>–<b>F</b>,<b>I</b>), 50 µm (<b>B</b>,<b>C</b>), and 10 µm (<b>G</b>,<b>H</b>), respectively.</p>
Full article ">Figure 4
<p>TEM of hematopoietic tissue of naturally infected <span class="html-italic">M. rosenbergii</span> samples 20180620. (<b>A</b>) A large numbers of virions in hematopoietic tissue. (<b>B</b>) DIV1 budded and acquired an envelope from the plasma membrane. (<b>C</b>) DIV1 replication and assembly in hematopoietic cells. (<b>D</b>) The stages of nucleocapsid assembly, which are indicated with numbers 1–3, and a complete nucleocapsid is indicated with number 4. The capsids at stage 2 and 3 should have a small opening at one vertex but may not be visible in the picture due to the ultrathin section. (<b>E</b>) Crescent-shaped structures. (<b>F</b>–<b>I</b>) As the assembling process continues, the crescent-shaped structure curves to form icosahedral capsids. (<b>J</b>) A mature virion with a dense core was eventually formed. N: nucleus; *: a large electron-lucent virogenic stroma; white arrows: paracrystalline array of viral particles; black arrows: budding virions; and white triangles: budded virions that acquired an envelope.</p>
Full article ">Figure 5
<p>Relative abundance of DIV1 for different tissues of fifteen DIV1-infected <span class="html-italic">M. rosenbergii</span> samples. Columns without sharing of a same letter indicate significant difference of <span class="html-italic">p</span> &lt; 0.05; columns without a same color indicate a highly significant difference of <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
14 pages, 6251 KiB  
Article
Critical Role of an MHC Class I-Like/Innate-Like T Cell Immune Surveillance System in Host Defense against Ranavirus (Frog Virus 3) Infection
by Eva-Stina Isabella Edholm, Francisco De Jesús Andino, Jinyeong Yim, Katherine Woo and Jacques Robert
Viruses 2019, 11(4), 330; https://doi.org/10.3390/v11040330 - 6 Apr 2019
Cited by 11 | Viewed by 3077
Abstract
Besides the central role of classical Major Histocompatibility Complex (MHC) class Ia-restricted conventional Cluster of Differentiation 8 (CD8) T cells in antiviral host immune response, the amphibian Xenopus laevis critically rely on MHC class I-like (mhc1b10.1.L or XNC10)-restricted innate-like (i)T cells (iVα6 T [...] Read more.
Besides the central role of classical Major Histocompatibility Complex (MHC) class Ia-restricted conventional Cluster of Differentiation 8 (CD8) T cells in antiviral host immune response, the amphibian Xenopus laevis critically rely on MHC class I-like (mhc1b10.1.L or XNC10)-restricted innate-like (i)T cells (iVα6 T cells) to control infection by the ranavirus Frog virus 3 (FV3). To complement and extend our previous reverse genetic studies showing that iVα6 T cells are required for tadpole survival, as well as for timely and effective adult viral clearance, we examined the conditions and kinetics of iVα6 T cell response against FV3. Using a FV3 knock-out (KO) growth-defective mutant, we found that upregulation of the XNC10 restricting class I-like gene and the rapid recruitment of iVα6 T cells depend on detectable viral replication and productive FV3 infection. In addition, by in vivo depletion with XNC10 tetramers, we demonstrated the direct antiviral effector function of iVα6 T cells. Notably, the transitory iVα6 T cell defect delayed innate interferon and cytokine gene response, resulting in long-lasting negative inability to control FV3 infection. These findings suggest that in Xenopus and likely other amphibians, an immune surveillance system based on the early activation of iT cells by non-polymorphic MHC class-I like molecules is important for efficient antiviral immune response. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of infection with attenuated knockout (KO) FV3 recombinant or bacterial stimulation on iVα6 T cell response. Peritoneal leukocytes (PLs) were collected at 1 dpi from adult frogs infected with 1 × 10<sup>6</sup> PFUs of WT-FV3 or ∆64-FV3, or 100 µl heat-killed (HB) <span class="html-italic">E. coli</span>. (<b>A</b>) Genome copy number using absolute qPCR with primers against FV3 DNA polymerase II. (<b>B</b>) XNC10 relative gene expression and (<b>C</b>) iVα6-Jα1.43 relative gene expression. Gene expression was determined relative to an endogenous control (GAPDH) and fold changes were calculated using the unstimulated sample (injected with equivalent volume of APBS) collected at the same time point. Data are pooled from three independent experiments with <span class="html-italic">n</span> = 4–5 animals in each experiment and each dot represents an individual animal. The line intersecting the <span class="html-italic">y</span>-axis at 0 represents the unstimulated control that the fold changes of the treatments are in relation to; (#) <span class="html-italic">p</span> &lt; 0.05 significant differences compared to unchallenged (APBS) injected controls; (*) <span class="html-italic">p</span> &lt; 0.05 and (***) <span class="html-italic">p</span> &lt; 0.001 statistically significant differences between the indicated groups (one way ANOVA and Dunns’s multiple comparison test).</p>
Full article ">Figure 2
<p>Magnitude of iVα6 T cell response in adult frogs is associated to the level of viral replication and production of infectious particles. PLs and kidneys were collected at 1, 3, and 6 dpi from adult frogs infected with 1 × 10<sup>6</sup> PFUs of WT-or ∆64-FV3. FV3 genome copy number by absolute qPCR in PLs (<b>A</b>) and kidneys (<b>B</b>) were determined. The total number of infectious particles (nd: not detected) in the kidney was determined by plaque assay (<b>C</b>).Gene expression of iVα6-Jα1.43 and XNC10 in PLs (<b>D</b>,<b>F</b>) and kidneys (<b>E</b>,<b>G</b>) were determined relative to an endogenous control (GAPDH) and fold changes were calculated using mock-infected frogs as a control. Each dot represents an individual animal (<span class="html-italic">n</span> = 4–5). The line intersecting the <span class="html-italic">y</span>-axis at 0 represents the APBS control that the fold changes of the treatments are in relation to. Note: * <span class="html-italic">p</span> &lt; 0,05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 above the line denotes statistically significant differences between the different treatment groups; significant differences between time points within each treatment group are indicated within parentheses; NS indicates no significant differences (one way ANOVA and Dunns’s multiple comparison test).</p>
Full article ">Figure 3
<p>The iVα6 T cell response in tadpoles depends on active viral replication and productive FV3 infection. Three week-old (stage 55) tadpoles were infected with 10,000 PFUs of WT-or ∆64-FV3. At 1, 3, and 6 dpi, kidneys were collected and the total number of infectious particles was determined by plaque assay, respectively (<b>A</b>). Gene expression of iVα6-Jα1.43 (<b>B</b>) and XNC10 (<b>C</b>) was determined relative to an endogenous control (GAPDH) and relative expression was calculated against the lowest observed expression according to the ∆∆Ct method (<span class="html-italic">n</span> = 5). No iVα6-Jα1.43 transcripts were detected in the kidney uninfected tadpoles. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 above the line denotes statistically significant differences between treatment groups; NS indicates no significant differences (one way ANOVA and Dunn’s multiple comparison test).</p>
Full article ">Figure 4
<p>XNC10 tetramer-mediated iVα6T cell depletion in tadpoles affects iVα6-Jα1.43 transcript levels and viral replication. PLs and kidneys were collected from three week-old (stage 55) tadpoles that had been injected with 1 μg XNC10 tetramers (XNC10-T) or vehicle control 1 day pre, and 1 day post i.p. injected with 10,000 PFUs of FV3, at the indicated time points (<span class="html-italic">n</span> = 9). A schematic of the injection regime is shown in (<b>A</b>). Gene expression of iVα6-Jα1.43 in PLs (<b>B</b>) and kidneys (<b>C</b>) is shown. Results are normalized to an endogenous control and presented as relative expression compared with the lowest observed value according to the ∆∆Ct method. FV3 loads in PLs (<b>D</b>) and kidneys (<b>E</b>) were measured using absolute qPCR with primers against FV3 polymerase II. For PLs, each dot represents a pool of 3 tadpoles, while for kidneys, each dot represents a single tadpole; * <span class="html-italic">p</span> &lt; 0.05 denotes statistically significant differences between the indicated groups; NS indicates no significant differences (One way ANOVA followed by Tukey’s multiple comparison test).</p>
Full article ">Figure 5
<p>Specificity and long term impact of transitory iVα6T cell depletion. (<b>A</b>) Three week-old (stage 55) tadpoles were injected with either 1 μg XNC10-tetramers (FV3/XNC10-T), 1 μg XNC10-monomers (FV3/XNC10-M), or vehicle control (FV3/APBS) 1 day pre and 1 day post i.p. infection with 10,000 PFUs of FV3 (<span class="html-italic">n</span> = 8), and kidneys were collected at 6 dpi. The last group (FV3/XNC10-T priming) was first injected with 1 μg XNC10-tetramers, then 3 days later infected with FV3, and kidneys were collected at 6 dpi. Viral loads were assessed by absolute qPCR with primers against FV3 polymerase II. The results are combined from two separate experiments, and each dot represents an individual tadpole; ** <span class="html-italic">p</span> &lt; 0.01 above the line denotes statistically significant differences between the indicated groups (One way ANOVA followed by Tukey’s multiple comparison test). (<b>B</b>) Three week-old (stage 55) tadpoles were injected with either 1 μg XNC10 tetramers (FV3/XNC10-T) or vehicle control (FV3/APBS) 1 day pre, and 1 day post were i.p injected with 10,000 PFUs of FV3 (<span class="html-italic">n</span> = 15) and survival was monitored daily over a 30-day period. Survival was determined using Kaplan-Meier, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, and *** <span class="html-italic">p</span> &lt; 0.0005. Uninfected controls (white circle), FV3 infected tadpoles (black circle), and XNC10 tetramer treated FV3 infected tadpoles (grey circle).</p>
Full article ">Figure 6
<p>Effect of XNC10 tetramer treatment on the expression of the macrophage stimulating factor genes CSF-1 and IL-34 in the peritoneal cavity and kidneys of FV3 infected tadpoles. PLs and kidneys were collected from three week-old (stage 55) tadpoles that had been injected with 1 μg XNC10 tetramers or vehicle control 1 day pre- and 1 day post-i.p. injection with 10,000 PFUs of FV3, at the indicated time points (<span class="html-italic">n</span> = 8–9). Quantitative gene expression analysis of IL-34 (<b>A</b>,<b>B</b>) and CSF-1(<b>C</b>,<b>D</b>) were determined relative to a endogenous control (GAPDH), and relative expression was calculated against the lowest observed expression according to the ∆∆Ct method (<span class="html-italic">n</span> = 9); ** <span class="html-italic">p</span> &lt; 0.005 above the line denotes statistically significant differences between the different treatment groups; NS indicates no significant differences (One way ANOVA followed by Tukey’s multiple comparison test).</p>
Full article ">Figure 7
<p>XNC10 tetramer treatment results in a delayed antiviral response in the peritoneal cavity and kidneys of FV3 infected tadpoles. PLs and kidneys were collected from three week-old (stage 55) tadpoles that had been injected with 1 μg XNC10 tetramers (XNC10-T) or vehicle control 1 day pre- and 1 day post-i.p. injection with 10,000 PFUs of FV3, at the indicated time points (<span class="html-italic">n</span> = 8–9). Quantitative gene expression analysis of type I (<b>A</b>,<b>D</b>), type II (<b>B</b>,<b>E</b>), and type III IFN (<b>C</b>,<b>F</b>) were determined relative to a endogenous control (GAPDH), and relative expression was calculated against the lowest observed expression according to the ∆∆Ct method (<span class="html-italic">n</span> = 9); * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, and *** <span class="html-italic">p</span> &lt; 0.001 above the line denote statistically significant differences between the different treatment groups; significant differences between time points within each treatment group is indicated within parentheses; NS indicates no significant differences (One way ANOVA followed by Tukey’s multiple comparison test).</p>
Full article ">Figure 8
<p>Effects of iVα6T cell depletion on IL-18 and IL-12 responses. PLs and kidneys were collected from three week-old (stage 55) tadpoles that had been injected with 1 μg XNC10 tetramers (XNC10-T) or vehicle control 1 day pre- and 1 day post-i.p. injection with 10,000 PFUs of FV3, at the indicated time points (<span class="html-italic">n</span> = 8–9). Quantitative gene expression of IL-18 (<b>A</b>,<b>B</b>) and type IL-12 (<b>C</b>,<b>D</b>) was determined relative to an endogenous control (GAPDH) and relative expression was calculated against the lowest observed expression according to the ∆∆Ct method (<span class="html-italic">n</span> = 9); * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005, above the line denotes statistically significant differences between the different treatment groups; significant differences between time points within each treatment group is indicated within parentheses; NS indicates no significant differences (One way ANOVA followed by Tukey’s multiple comparisons test).</p>
Full article ">
15 pages, 2455 KiB  
Article
Geographic Distribution of Epizootic haematopoietic necrosis virus (EHNV) in Freshwater Fish in South Eastern Australia: Lost Opportunity for a Notifiable Pathogen to Expand Its Geographic Range
by Joy A. Becker, Dean Gilligan, Martin Asmus, Alison Tweedie and Richard J. Whittington
Viruses 2019, 11(4), 315; https://doi.org/10.3390/v11040315 - 1 Apr 2019
Cited by 3 | Viewed by 3718
Abstract
Epizootic haematopoietic necrosis virus (EHNV) was originally detected in Victoria, Australia in 1984. It spread rapidly over two decades with epidemic mortality events in wild redfin perch (Perca fluviatilis) and mild disease in farmed rainbow trout (Oncorhynchus mykiss) being [...] Read more.
Epizootic haematopoietic necrosis virus (EHNV) was originally detected in Victoria, Australia in 1984. It spread rapidly over two decades with epidemic mortality events in wild redfin perch (Perca fluviatilis) and mild disease in farmed rainbow trout (Oncorhynchus mykiss) being documented across southeastern Australia in New South Wales (NSW), the Australian Capital Territory (ACT), Victoria, and South Australia. We conducted a survey for EHNV between July 2007 and June 2011. The disease occurred in juvenile redfin perch in ACT in December 2008, and in NSW in December 2009 and December 2010. Based on testing 3622 tissue and 492 blood samples collected from fish across southeastern Australia, it was concluded that EHNV was most likely absent from redfin perch outside the endemic area in the upper Murrumbidgee River catchment in the Murray–Darling Basin (MDB), and it was not detected in other fish species. The frequency of outbreaks in redfin perch has diminished over time, and there have been no reports since 2012. As the disease is notifiable and a range of fish species are known to be susceptible to EHNV, existing policies to reduce the likelihood of spreading out of the endemic area are justified. Full article
Show Figures

Figure 1

Figure 1
<p>Maximum extent of study area and catchments of the Murray–Darling Basin.</p>
Full article ">Figure 2
<p>Distribution and number of tissue and serum samples collected from all fish species between July 2007 and June 2011.</p>
Full article ">Figure 3
<p>Distribution and number of tissue and serum samples collected from redfin perch between July 2007 and June 2011.</p>
Full article ">Figure 4
<p>Moribund and dead redfin perch collected from Lake Ginnindera, ACT in December 2008 and 2010, and Blowering Dam, NSW in December 2009 were infected with EHNV.</p>
Full article ">
15 pages, 7414 KiB  
Article
Pathogen Risk Analysis for Wild Amphibian Populations Following the First Report of a Ranavirus Outbreak in Farmed American Bullfrogs (Lithobates catesbeianus) from Northern Mexico
by Bernardo Saucedo, José M. Serrano, Mónica Jacinto-Maldonado, Rob S. E. W. Leuven, Abraham A. Rocha García, Adriana Méndez Bernal, Andrea Gröne, Steven J. Van Beurden and César M. Escobedo-Bonilla
Viruses 2019, 11(1), 26; https://doi.org/10.3390/v11010026 - 3 Jan 2019
Cited by 15 | Viewed by 5268
Abstract
Ranaviruses are the second deadliest pathogens for amphibian populations throughout the world. Despite their wide distribution in America, these viruses have never been reported in Mexico, the country with the fifth highest amphibian diversity in the world. This paper is the first to [...] Read more.
Ranaviruses are the second deadliest pathogens for amphibian populations throughout the world. Despite their wide distribution in America, these viruses have never been reported in Mexico, the country with the fifth highest amphibian diversity in the world. This paper is the first to address an outbreak of ranavirus in captive American bullfrogs (Lithobates catesbeianus) from Sinaloa, Mexico. The farm experienced high mortality in an undetermined number of juveniles and sub-adult bullfrogs. Affected animals displayed clinical signs and gross lesions such as lethargy, edema, skin ulcers, and hemorrhages consistent with ranavirus infection. The main microscopic lesions included mild renal tubular necrosis and moderate congestion in several organs. Immunohistochemical analyses revealed scant infected hepatocytes and renal tubular epithelial cells. Phylogenetic analysis of five partial ranavirus genes showed that the causative agent clustered within the Frog virus 3 clade. Risk assessment with the Pandora+ protocol demonstrated a high risk for the pathogen to affect amphibians from neighboring regions (overall Pandora risk score: 0.619). Given the risk of American bullfrogs escaping and spreading the disease to wild amphibians, efforts should focus on implementing effective containment strategies and surveillance programs for ranavirus at facilities undertaking intensive farming of amphibians. Full article
Show Figures

Figure 1

Figure 1
<p>Location of city of the outbreak in Mexico. The city of Guasave is indicated by a black circle within the province of Sinaloa (grey). Mexico City, the place from which the bullfrogs were imported from, is shown in black.</p>
Full article ">Figure 2
<p>Macroscopic lesions of ranavirus-infected bullfrogs. (<b>A</b>) Extensive ulceration of the forelimb. (<b>B</b>) Extensive hepatic necrosis and epicardial pallor and hemorrhages (black arrow).</p>
Full article ">Figure 3
<p>Histopathology of ranavirus-infected American bullfrogs. (<b>A</b>) Liver section from an affected bullfrog with mild congestion of sinusoids. (<b>B</b>) Kidney section shows an area of mild tubular necrosis characterized by cytoplasmic hyper-eosinophilia, loss of nuclei, and mild nuclear pyknosis (black square). Intraluminal protein casts are also seen (black arrows). Hematoxylin/Eosin staining. Magnification for all photos 20×/100 μm.</p>
Full article ">Figure 4
<p>Immunohistochemistry of ranavirus-infected American bullfrogs (<span class="html-italic">Lithobates catesbeianus</span>). (<b>A</b>) Immunohistochemical staining of kidney, in which positive immunolabelling (red staining) is present in the renal cortical interstitium along with areas of necrosis and scant inflammation. (<b>B</b>) Immunohistochemical staining of a liver from an affected bullfrog with a focal area of positive immunolabelling (red staining) in the cytoplasm of a Kupffer cell. (<b>C</b>) Serial section of the kidney from the same animal without secondary antibody (control), (<b>D</b>) Serial section of liver from the same animal without secondary antibody (control). (<b>E</b>) Histological kidney section from an anuran which was PCR negative for ranavirus infection (control). (<b>F</b>) Histological liver section from an anuran which was PCR negative for ranavirus infection (control). Magnification for all photos 40×/50 μm.</p>
Full article ">Figure 5
<p>PCR products of five ranavirus genes. All products of the Mexican virus samples (S) are of similar size to those of the FV3 virus positive control (P). Abbreviations: W (water), S (Mexican virus sample), P (Positive control frog virus 3 isolate <span class="html-italic">Oophaga pumilio</span>/2015/Netherlands/UU3150324001 (GenBank no. MF360246)). 100 bp ladder.</p>
Full article ">Figure 6
<p>Phylogeny of five partial ranavirus gene sequences (1000 bootstrap values). The Mexican isolate FV3 <span class="html-italic">L. catesbeianus</span> (underlined) clusters closely within the FV3 ranavirus clade. Isolates and Genbank numbers used: Frog virus 3/<span class="html-italic">Lithobates catesbeianus</span>/2018/Mexico (GenBank accession numbers for partial sequences available in <a href="#app1-viruses-11-00026" class="html-app">Supplementary Table S1</a>), frog virus 3 (AY548484), frog virus 3 isolate SMME (KJ175144), tortoise ranavirus isolate 1 (882/96) (KP266743), Bosca’s newt virus isolate GA11001 (KJ703118), frog virus 3 isolate <span class="html-italic">Oophaga pumilio</span>/2015/Netherlands/UU3150324001 (MF360246), common midwife toad ranavirus isolate <span class="html-italic">Mesotriton alpestris</span>/2008/E (JQ231222), common midwife toad virus isolate P11114 (KJ703146), <span class="html-italic">Testudo hermanni</span> ranavirus isolate CH8/96 (KP266741), tiger frog virus (AF389451), soft-shelled turtle iridovirus (EU627010), European sheatfish virus (JQ724856), epizootic hematopoietic necrosis virus (FJ433873), <span class="html-italic">Ambystoma tigrinum stebbensi</span> virus (AY150217), <span class="html-italic">Andrias davidianus</span> ranavirus isolate (KF033124), <span class="html-italic">Andrias davidianus</span> ranavirus isolate 1201 (KC865735), Chinese giant salamander iridovirus, isolate CGSIV-HN1104, (KF512820), and <span class="html-italic">Rana grylio</span> virus (JQ654586). Only bootstrap values higher than 50 are shown.</p>
Full article ">

Other

Jump to: Editorial, Research

15 pages, 1053 KiB  
Technical Note
eDNA Increases the Detectability of Ranavirus Infection in an Alpine Amphibian Population
by Claude Miaud, Véronique Arnal, Marie Poulain, Alice Valentini and Tony Dejean
Viruses 2019, 11(6), 526; https://doi.org/10.3390/v11060526 - 6 Jun 2019
Cited by 33 | Viewed by 5431
Abstract
The early detection and identification of pathogenic microorganisms is essential in order to deploy appropriate mitigation measures. Viruses in the Iridoviridae family, such as those in the Ranavirus genus, can infect amphibian species without resulting in mortality or clinical signs, and they can [...] Read more.
The early detection and identification of pathogenic microorganisms is essential in order to deploy appropriate mitigation measures. Viruses in the Iridoviridae family, such as those in the Ranavirus genus, can infect amphibian species without resulting in mortality or clinical signs, and they can also infect other hosts than amphibian species. Diagnostic techniques allowing the detection of the pathogen outside the period of host die-off would thus be of particular use. In this study, we tested a method using environmental DNA (eDNA) on a population of common frogs (Rana temporaria) known to be affected by a Ranavirus in the southern Alps in France. In six sampling sessions between June and September (the species’ activity period), we collected tissue samples from dead and live frogs (adults and tadpoles), as well as insects (aquatic and terrestrial), sediment, and water. At the beginning of the breeding season in June, one adult was found dead; at the end of July, a mass mortality of tadpoles was observed. The viral DNA was detected in both adults and tadpoles (dead or alive) and in water samples, but it was not detected in insects or sediment. In live frog specimens, the virus was detected from June to September and in water samples from August to September. Dead tadpoles that tested positive for Ranavirus were observed only on one date (at the end of July). Our results indicate that eDNA can be an effective alternative to tissue/specimen sampling and can detect Ranavirus presence outside die-offs. Another advantage is that the collection of water samples can be performed by most field technicians. This study confirms that the use of eDNA can increase the performance and accuracy of wildlife health status monitoring and thus contribute to more effective surveillance programs. Full article
Show Figures

Figure 1

Figure 1
<p>Common frog <span class="html-italic">Rana temporaria</span> tadpoles feeding on their dead congeners (photo L. Miaud).</p>
Full article ">Figure 2
<p>Change in water temperature and <span class="html-italic">Ranavirus</span> prevalence and load in a pond where the mass mortality of common frog tadpoles was observed. Water temperature was recorded at the surface and 30 cm below the surface. Top row of histograms: <span class="html-italic">Ranavirus</span> DNA quantities in common frogs (adults and tadpoles) and water samples (log scales, mean value, and SD). Bottom row of histograms: Prevalence (number of <span class="html-italic">Ranavirus</span> positive specimens/total number of sampled specimens, with sample size indicated above the bars). Developmental stages are based on Gosner stages; mortality was observed only on July 27.</p>
Full article ">
8 pages, 1821 KiB  
Brief Report
Localization of Frog Virus 3 Conserved Viral Proteins 88R, 91R, and 94L
by Emily Penny and Craig R. Brunetti
Viruses 2019, 11(3), 276; https://doi.org/10.3390/v11030276 - 19 Mar 2019
Cited by 2 | Viewed by 2861
Abstract
The characterization of the function of conserved viral genes is central to developing a greater understanding of important aspects of viral replication or pathogenesis. A comparative genomic analysis of the iridoviral genomes identified 26 core genes conserved across the family Iridoviridae. Three [...] Read more.
The characterization of the function of conserved viral genes is central to developing a greater understanding of important aspects of viral replication or pathogenesis. A comparative genomic analysis of the iridoviral genomes identified 26 core genes conserved across the family Iridoviridae. Three of those conserved genes have no defined function; these include the homologs of frog virus 3 (FV3) open reading frames (ORFs) 88R, 91R, and 94L. Conserved viral genes that have been previously identified are known to participate in a number of viral activities including: transcriptional regulation, DNA replication/repair/modification/processing, protein modification, and viral structural proteins. To begin to characterize the conserved FV3 ORFs 88R, 91R, and 94L, we cloned the genes and determined their intracellular localization. We demonstrated that 88R localizes to the cytoplasm of the cell while 91R localizes to the nucleus and 94L localizes to the endoplasmic reticulum (ER). Full article
Show Figures

Figure 1

Figure 1
<p>FV3 88R localizes to the cytoplasm. Baby Green Monkey Kidney (BGMK) cells were transfected with pcDNA3-88R. 48 hours post transfection, cells were fixed, and indirect immunofluorescence was performed using rabbit anti-myc antibodies (red) and TO-PRO-3 (blue). Images were captured at 100× magnification using a confocal microscope.</p>
Full article ">Figure 2
<p>91R localizes to the nucleus. BGMK cells were transfected with pcDNA3-91R-myc. 48 hours post-transfection, the cells were fixed and indirect immunofluorescence was performed using rabbit anti-myc antibodies (red) and TO-PRO-3 (blue). Images were captured at 100× magnification using a confocal microscope. White arrows highlight nuclear areas that lack 91R and TO-PRO-3.</p>
Full article ">Figure 3
<p>94L localizes to the Endoplasmic Reticulum. BGMK cells were transfected with pcDNA3-94L-myc. Forty-eight hours post-transfection, the cells were fixed and indirect immunofluorescence was performed using mouse anti-myc antibodies (red) and rabbit anti-PDI antibodies (green). Images were captured at 100× magnification using a confocal microscope.</p>
Full article ">
Back to TopTop