Journal articles on the topic 'Flock house virus'

To see the other types of publications on this topic, follow the link: Flock house virus.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Flock house virus.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Fisher, A. J., B. R. McKinney, J. P. Wery, and J. E. Johnson. "Crystallization and preliminary data analysis of flock house virus." Acta Crystallographica Section B Structural Science 48, no. 4 (August 1, 1992): 515–20. http://dx.doi.org/10.1107/s0108768192000053.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

A. McCormick, Alison, Payal D. Maharaj, Jyothi K. Mallajosyula, Philip Thi, Gloria Lee, Yiyang Zhou, and Christopher Kearney. "Trans-Encapsidation of Flock House virus with Tobacco Mosaic virus Structural Proteins." Journal of Advanced Biotechnology and Bioengineering 2, no. 2 (November 2014): 49–59. http://dx.doi.org/10.12970/2311-1755.2014.02.02.2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Johnson, Kyle L., and L. Andrew Ball. "Induction and Maintenance of Autonomous Flock House Virus RNA1 Replication." Journal of Virology 73, no. 10 (1999): 7933–42. http://dx.doi.org/10.1128/jvi.73.10.7933-7942.1999.

Full text
Abstract:
The nodavirus flock house virus (FHV) has a bipartite, positive-sense, RNA genome that encodes the catalytic subunit of the RNA replicase and the viral capsid protein precursor on separate genomic segments (RNA1 and RNA2, respectively). RNA1 can replicate autonomously when transfected into permissive cells, allowing study of the kinetics of RNA1 replication in the absence of either RNA2 or capsid proteins. However, RNA1 replication ceases ca. 3 days after transfection despite the presence of replication-competent RNA. We examined this inhibition by inducing the expression of RNA1 in cells from a cDNA copy that was under the control of a hormone-regulated RNA polymerase II promoter. This system reproduced the shutoff of RNA replication when DNA-templated primary transcription was turned off. Continued primary transcription partially alleviated the shutoff and maintained the rate of RNA replication for several days at a steady-state level approximately one-third that of the peak rate. After shutoff, RNA replication could be restored by transferring the resulting intracellular RNA to fresh cells or by reinducing primary transcription, indicating that cessation of replication occurred despite the competence of both the viral RNA and the cytoplasmic environment. These data suggest that there is a mechanism by which replication is shut off at late times after transfection, which may reflect the natural endpoint of the replicative cycle.
APA, Harvard, Vancouver, ISO, and other styles
4

Short, James R., Jeffrey A. Speir, Radhika Gopal, Logan M. Pankratz, Jason Lanman, and Anette Schneemann. "Role of Mitochondrial Membrane Spherules in Flock House Virus Replication." Journal of Virology 90, no. 7 (January 20, 2016): 3676–83. http://dx.doi.org/10.1128/jvi.03080-15.

Full text
Abstract:
ABSTRACTViruses that generate double-stranded RNA (dsRNA) during replication must overcome host defense systems designed to detect this infection intermediate. All positive-sense RNA viruses studied to date modify host membranes to help facilitate the sequestration of dsRNA from host defenses and concentrate replication factors to enhance RNA production. Flock House virus (FHV) is an attractive model for the study of these processes since it is well characterized and infectsDrosophilacells, which are known to have a highly effective RNA silencing system. During infection, FHV modifies the outer membrane of host mitochondria to form numerous membrane invaginations, called spherules, that are ∼50 nm in diameter and known to be the site of viral RNA replication. While previous studies have outlined basic structural features of these invaginations, very little is known about the mechanism underlying their formation. Here we describe the optimization of an experimental system for the analysis of FHV host membrane modifications using crude mitochondrial preparations from infectedDrosophilacells. These preparations can be programmed to synthesize both single- and double-stranded FHV RNA. The system was used to demonstrate that dsRNA is protected from nuclease digestion by virus-induced membrane invaginations and that spherules play an important role in stimulating RNA replication. Finally, we show that spherules generated during FHV infection appear to be dynamic as evidenced by their ability to form or disperse based on the presence or absence of RNA synthesis.IMPORTANCEIt is well established that positive-sense RNA viruses induce significant membrane rearrangements in infected cells. However, the molecular mechanisms underlying these rearrangements, particularly membrane invagination and spherule formation, remain essentially unknown. How the formation of spherules enhances viral RNA synthesis is also not understood, although it is assumed to be partly a result of evading host defense pathways. To help interrogate some of these issues, we optimized a cell-free replication system consisting of mitochondria isolated from Flock House virus-infectedDrosophilacells for use in biochemical and structural studies. Our data suggest that spherules generated during Flock House virus replication are dynamic, protect double-stranded RNA, and enhance RNA replication in general. Cryo-electron microscopy suggests that the samples are amenable to detailed structural analyses of spherules engaged in RNA synthesis. This system thus provides a foundation for understanding the molecular mechanisms underlying spherule formation, maintenance, and function during positive-sense viral RNA replication.
APA, Harvard, Vancouver, ISO, and other styles
5

Thomson, Travis C., and Joshua Johnson. "Infection with Flock House Virus Induces Oocyte Destruction in Drosophila." Biology of Reproduction 83, Suppl_1 (November 1, 2010): 662. http://dx.doi.org/10.1093/biolreprod/83.s1.662.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Bong, Dennis T., Andreas Janshoff, Claudia Steinem, and M. Reza Ghadiri. "Membrane Partitioning of the Cleavage Peptide in Flock House Virus." Biophysical Journal 78, no. 2 (February 2000): 839–45. http://dx.doi.org/10.1016/s0006-3495(00)76641-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Hiscox, J. A., and L. A. Ball. "Cotranslational disassembly of flock house virus in a cell-free system." Journal of virology 71, no. 10 (1997): 7974–77. http://dx.doi.org/10.1128/jvi.71.10.7974-7977.1997.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Jovel, Juan, and Anette Schneemann. "Molecular characterization of Drosophila cells persistently infected with Flock House virus." Virology 419, no. 1 (October 2011): 43–53. http://dx.doi.org/10.1016/j.virol.2011.08.002.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Dasgupta, Ranjit, Li-Lin Cheng, Lyric C. Bartholomay, and Bruce M. Christensen. "Flock house virus replicates and expresses green fluorescent protein in mosquitoes." Journal of General Virology 84, no. 7 (July 1, 2003): 1789–97. http://dx.doi.org/10.1099/vir.0.18938-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

McMenamin, Alexander J., Fenali Parekh, Verena Lawrence, and Michelle L. Flenniken. "Investigating Virus–Host Interactions in Cultured Primary Honey Bee Cells." Insects 12, no. 7 (July 17, 2021): 653. http://dx.doi.org/10.3390/insects12070653.

Full text
Abstract:
Honey bee (Apis mellifera) health is impacted by viral infections at the colony, individual bee, and cellular levels. To investigate honey bee antiviral defense mechanisms at the cellular level we further developed the use of cultured primary cells, derived from either larvae or pupae, and demonstrated that these cells could be infected with a panel of viruses, including common honey bee infecting viruses (i.e., sacbrood virus (SBV) and deformed wing virus (DWV)) and an insect model virus, Flock House virus (FHV). Virus abundances were quantified over the course of infection. The production of infectious virions in cultured honey bee pupal cells was demonstrated by determining that naïve cells became infected after the transfer of deformed wing virus or Flock House virus from infected cell cultures. Initial characterization of the honey bee antiviral immune responses at the cellular level indicated that there were virus-specific responses, which included increased expression of bee antiviral protein-1 (GenBank: MF116383) in SBV-infected pupal cells and increased expression of argonaute-2 and dicer-like in FHV-infected hemocytes and pupal cells. Additional studies are required to further elucidate virus-specific honey bee antiviral defense mechanisms. The continued use of cultured primary honey bee cells for studies that involve multiple viruses will address this knowledge gap.
APA, Harvard, Vancouver, ISO, and other styles
11

Miller, David J., Michael D. Schwartz, and Paul Ahlquist. "Flock House Virus RNA Replicates on Outer Mitochondrial Membranes in Drosophila Cells." Journal of Virology 75, no. 23 (December 1, 2001): 11664–76. http://dx.doi.org/10.1128/jvi.75.23.11664-11676.2001.

Full text
Abstract:
ABSTRACT The identification and characterization of host cell membranes essential for positive-strand RNA virus replication should provide insight into the mechanisms of viral replication and potentially identify novel targets for broadly effective antiviral agents. The alphanodavirus flock house virus (FHV) is a positive-strand RNA virus with one of the smallest known genomes among animal RNA viruses, and it can replicate in insect, plant, mammalian, and yeast cells. To investigate the localization of FHV RNA replication, we generated polyclonal antisera against protein A, the FHV RNA-dependent RNA polymerase, which is the sole viral protein required for FHV RNA replication. We detected protein A within 4 h after infection ofDrosophila DL-1 cells and, by differential and isopycnic gradient centrifugation, found that protein A was tightly membrane associated, similar to integral membrane replicase proteins from other positive-strand RNA viruses. Confocal immunofluorescence microscopy and virus-specific, actinomycin D-resistant bromo-UTP incorporation identified mitochondria as the intracellular site of protein A localization and viral RNA synthesis. Selective membrane permeabilization and immunoelectron microscopy further localized protein A to outer mitochondrial membranes. Electron microscopy revealed 40- to 60-nm membrane-bound spherical structures in the mitochondrial intermembrane space of FHV-infected cells, similar in ultrastructural appearance to tombusvirus- and togavirus-induced membrane structures. We concluded that FHV RNA replication occurs on outer mitochondrial membranes and shares fundamental biochemical and ultrastructural features with RNA replication of positive-strand RNA viruses from other families.
APA, Harvard, Vancouver, ISO, and other styles
12

Chao, Jeffrey A., June Hyung Lee, Brian R. Chapados, Erik W. Debler, Anette Schneemann, and James R. Williamson. "Dual modes of RNA-silencing suppression by Flock House virus protein B2." Nature Structural & Molecular Biology 12, no. 11 (October 9, 2005): 952–57. http://dx.doi.org/10.1038/nsmb1005.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Dasgupta, Ranjit, Heather M. Free, Suzanne L. Zietlow, Susan M. Paskewitz, Serap Aksoy, Lei Shi, Jeremy Fuchs, Changyun Hu, and Bruce M. Christensen. "Replication of Flock House Virus in Three Genera of Medically Important Insects." Journal of Medical Entomology 44, no. 1 (January 1, 2007): 102–10. http://dx.doi.org/10.1093/jmedent/41.5.102.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Ball, L. A., and Y. Li. "cis-acting requirements for the replication of flock house virus RNA 2." Journal of Virology 67, no. 6 (1993): 3544–51. http://dx.doi.org/10.1128/jvi.67.6.3544-3551.1993.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Li, Y., and L. A. Ball. "Nonhomologous RNA recombination during negative-strand synthesis of flock house virus RNA." Journal of Virology 67, no. 7 (1993): 3854–60. http://dx.doi.org/10.1128/jvi.67.7.3854-3860.1993.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Ball, L. A. "Requirements for the self-directed replication of flock house virus RNA 1." Journal of virology 69, no. 2 (1995): 720–27. http://dx.doi.org/10.1128/jvi.69.2.720-727.1995.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Ball, L. A. "Requirements for the self-directed replication of flock house virus RNA 1." Journal of virology 69, no. 4 (1995): 2722. http://dx.doi.org/10.1128/jvi.69.4.2722-2722.1995.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Venter, P. A., and A. Schneemann. "Recent insights into the biology and biomedical applications of Flock House virus." Cellular and Molecular Life Sciences 65, no. 17 (June 2, 2008): 2675–87. http://dx.doi.org/10.1007/s00018-008-8037-y.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Stapleford, Kenneth A., Doron Rapaport, and David J. Miller. "Mitochondrion-Enriched Anionic Phospholipids Facilitate Flock House Virus RNA Polymerase Membrane Association." Journal of Virology 83, no. 9 (February 25, 2009): 4498–507. http://dx.doi.org/10.1128/jvi.00040-09.

Full text
Abstract:
ABSTRACT One characteristic of all positive-strand RNA viruses is the necessity to assemble viral RNA replication complexes on host intracellular membranes, a process whose molecular details are poorly understood. To study viral replication complex assembly we use the established model system of Flock House virus (FHV), which assembles its replication complexes on the mitochondrial outer membrane. The FHV RNA-dependent RNA polymerase, protein A, is the only viral protein necessary for genome replication in the budding yeast Saccharomyces cerevisiae. To examine the host components involved in protein A-membrane interactions, an initial step of FHV RNA replication complex assembly, we established an in vitro protein A membrane association assay. Protein A translated in vitro rapidly and specifically associated with mitochondria isolated from yeast, insect, and mammalian cells. This process was temperature dependent but independent of protease-sensitive mitochondrial outer membrane components or the host mitochondrial import machinery. Furthermore, lipid-binding studies revealed that protein A preferentially bound to specific anionic phospholipids, in particular the mitochondrion-specific phospholipid cardiolipin. These studies implicate membrane phospholipids as important host determinants for FHV RNA polymerase membrane association and provide evidence for the involvement of host phospholipids in positive-strand RNA virus membrane-specific targeting.
APA, Harvard, Vancouver, ISO, and other styles
20

Kim, Hui-Bae, Do-Yeong Kim, and Tae-Ju Cho. "Replication and packaging of Turnip yellow mosaic virus RNA containing Flock house virus RNA1 sequence." BMB Reports 47, no. 6 (June 30, 2014): 330–35. http://dx.doi.org/10.5483/bmbrep.2014.47.6.187.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Maharaj, Payal, Jyothi Mallajosyula, Gloria Lee, Phillip Thi, Yiyang Zhou, Christopher Kearney, and Alison McCormick. "Nanoparticle Encapsidation of Flock House Virus by Auto Assembly of Tobacco Mosaic Virus Coat Protein." International Journal of Molecular Sciences 15, no. 10 (October 14, 2014): 18540–56. http://dx.doi.org/10.3390/ijms151018540.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Walukiewicz, Hanna E., John E. Johnson, and Anette Schneemann. "Morphological Changes in the T=3 Capsid of Flock House Virus during Cell Entry." Journal of Virology 80, no. 2 (January 15, 2006): 615–22. http://dx.doi.org/10.1128/jvi.80.2.615-622.2006.

Full text
Abstract:
ABSTRACT We report the identification and characterization of a viral intermediate formed during infection of Drosophila cells with the nodavirus Flock House virus (FHV). We observed that even at a very low multiplicity of infection, only 70% of the input virus stayed attached to or entered the cells, while the remaining 30% of the virus eluted from cells after initial binding. The eluted FHV particles did not rebind to Drosophila cells and, thus, could no longer initiate infection by the receptor-mediated entry pathway. FHV virus-like particles with the same capsid composition as native FHV but containing cellular RNA also exhibited formation of eluted particles when incubated with the cells. A maturation cleavage-defective mutant of FHV, however, did not. Compared to naïve FHV particles, i.e., particles that had never been incubated with cells, eluted particles showed an acid-sensitive phenotype and morphological alterations. Furthermore, eluted particles had lost a fraction of the internally located capsid protein gamma. Based on these results, we hypothesize that FHV eluted particles represent an infection intermediate analogous to eluted particles observed for members of the family Picornaviridae.
APA, Harvard, Vancouver, ISO, and other styles
23

Walukiewicz, Hanna E., Manidipa Banerjee, Anette Schneemann, and John E. Johnson. "Rescue of Maturation-Defective Flock House Virus Infectivity with Noninfectious, Mature, Viruslike Particles." Journal of Virology 82, no. 4 (December 12, 2007): 2025–27. http://dx.doi.org/10.1128/jvi.02278-07.

Full text
Abstract:
ABSTRACT The infectivity of flock house virus (FHV) requires autocatalytic maturation cleavage of the capsid protein at residue 363, liberating the C-terminal 44-residue γ peptides, which remain associated with the particle. In vitro studies previously demonstrated that the amphipathic, helical portion (amino acids 364 to 385) of γ is membrane active, suggesting a role for γ in RNA membrane translocation during infection. Here we show that the infectivity of a maturation-defective mutant of FHV can be restored by viruslike particles that lack the genome but undergo maturation cleavage. We propose that the colocalization of the two defective particle types in an entry compartment allows the restoration of infectivity by γ.
APA, Harvard, Vancouver, ISO, and other styles
24

Banerjee, Manidipa, Reza Khayat, Hanna E. Walukiewicz, Amy L. Odegard, Anette Schneemann, and John E. Johnson. "Dissecting the Functional Domains of a Nonenveloped Virus Membrane Penetration Peptide." Journal of Virology 83, no. 13 (April 15, 2009): 6929–33. http://dx.doi.org/10.1128/jvi.02299-08.

Full text
Abstract:
ABSTRACT Recent studies have established that several nonenveloped viruses utilize virus-encoded lytic peptides for host membrane disruption. We investigated this mechanism with the “gamma” peptide of the insect virus Flock House virus (FHV). We demonstrate that the C terminus of gamma is essential for membrane disruption in vitro and the rescue of immature virus infectivity in vivo, and the amphipathic N terminus of gamma alone is not sufficient. We also show that deletion of the C-terminal domain disrupts icosahedral ordering of the amphipathic helices of gamma in the virus. Our results have broad implications for understanding membrane lysis during nonenveloped virus entry.
APA, Harvard, Vancouver, ISO, and other styles
25

Settles, Erik W., and Paul D. Friesen. "Flock House Virus Induces Apoptosis by Depletion of Drosophila Inhibitor-of-Apoptosis Protein DIAP1." Journal of Virology 82, no. 3 (November 7, 2007): 1378–88. http://dx.doi.org/10.1128/jvi.01941-07.

Full text
Abstract:
ABSTRACT The molecular mechanisms by which RNA viruses induce apoptosis and apoptosis-associated pathology are not fully understood. Here we show that flock house virus (FHV), one of the simplest RNA viruses (family, Nodaviridae), induces robust apoptosis of permissive Drosophila Line-1 (DL-1) cells. To define the pathway by which FHV triggers apoptosis in this model invertebrate system, we investigated the potential role of Drosophila apoptotic effectors during infection. Suggesting the involvement of host caspases, the pancaspase inhibitor benzyloxycarbonyl-Val-Ala-Asp-fluromethylketone (z-VAD-fmk) prevented FHV-induced cytopathology and prolonged cell survival. RNA interference-mediated ablation of the principal Drosophila effector caspase DrICE or its upstream initiator caspase DRONC prevented FHV-induced apoptosis and demonstrated direct participation of this intrinsic caspase pathway. Prior to the FHV-induced activation of DrICE, the intracellular level of inhibitor-of-apoptosis (IAP) protein DIAP1, the principal caspase regulator in Drosophila melanogaster, was dramatically reduced. DIAP1 was depleted despite z-VAD-fmk-mediated caspase inhibition during infection, suggesting that the loss of DIAP1 was caused by an upstream FHV-induced signal. The RNA interference-mediated knockdown of DIAP1 caused rapid and uniform apoptosis of DL-1 cells and thus indicated that DIAP1 depletion is sufficient to trigger apoptosis. Confirming this conclusion, the elevation of intracellular DIAP1 levels in stable diap1-transfected cells blocked caspase activation and prevented FHV-induced apoptosis. Collectively, our findings suggest that DIAP1 is a critical sensor of virus infection, which upon virus-signaled depletion relieves caspase inhibition, which subsequently executes apoptotic death. Thus, our study supports the hypothesis that altering the level or the activity of cellular IAP proteins is a general mechanism by which RNA viruses trigger apoptosis.
APA, Harvard, Vancouver, ISO, and other styles
26

Kim, Doyeong, and Tae-Ju Cho. "Replication of Recombinant Flock House Virus RNA Encapsidated by Turnip Yellow Mosaic Virus Coat Proteins inNicotiana benthamiana." Journal of Bacteriology and Virology 47, no. 2 (2017): 87. http://dx.doi.org/10.4167/jbv.2017.47.2.87.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Lanman, Jason, John Crum, Thomas J. Deerinck, Guido M. Gaietta, Anette Schneemann, Gina E. Sosinsky, Mark H. Ellisman, and John E. Johnson. "Visualizing flock house virus infection in Drosophila cells with correlated fluorescence and electron microscopy." Journal of Structural Biology 161, no. 3 (March 2008): 439–46. http://dx.doi.org/10.1016/j.jsb.2007.09.009.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Albariño, César G., B. Duane Price, Lance D. Eckerle, and L. Andrew Ball. "Characterization and Template Properties of RNA Dimers Generated during Flock House Virus RNA Replication." Virology 289, no. 2 (October 2001): 269–82. http://dx.doi.org/10.1006/viro.2001.1125.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Krishna, Neel K., and Anette Schneemann. "Formation of an RNA Heterodimer upon Heating of Nodavirus Particles." Journal of Virology 73, no. 2 (February 1, 1999): 1699–703. http://dx.doi.org/10.1128/jvi.73.2.1699-1703.1999.

Full text
Abstract:
ABSTRACT Flock House virus is a small icosahedral insect virus of the familyNodaviridae. Its genome consists of two positive-sense RNA molecules, which are believed to be encapsidated into a single viral particle. However, evidence to support this claim is circumstantial. Here we demonstrate that exposure of nodavirus particles to heat causes the two strands of viral RNA to form a stable complex, directly establishing that both RNAs are copackaged into one virion. The physical properties of the RNA complex, the effect of heat on the particles per se, and the possible relevance of these findings to the nodavirus life cycle are presented.
APA, Harvard, Vancouver, ISO, and other styles
30

Quirin, Tania, Yu Chen, Maija Pietilä, Deyin Guo, and Tero Ahola. "The RNA Capping Enzyme Domain in Protein A is Essential for Flock House Virus Replication." Viruses 10, no. 9 (September 9, 2018): 483. http://dx.doi.org/10.3390/v10090483.

Full text
Abstract:
The nodavirus flock house virus (FHV) and the alphavirus Semliki Forest virus (SFV) show evolutionarily intriguing similarities in their replication complexes and RNA capping enzymes. In this study, we first established an efficient FHV trans-replication system in mammalian cells, which disjoins protein expression from viral RNA synthesis. Following transfection, FHV replicase protein A was associated with mitochondria, whose outer surface displayed pouch-like invaginations with a ‘neck’ structure opening towards the cytoplasm. In mitochondrial pellets from transfected cells, high-level synthesis of both genomic and subgenomic RNA was detected in vitro and the newly synthesized RNA was of positive polarity. Secondly, we initiated the study of the putative RNA capping enzyme domain in protein A by mutating the conserved amino acids H93, R100, D141, and W215. RNA replication was abolished for all mutants inside cells and in vitro except for W215A, which showed reduced replication. Transfection of capped RNA template did not rescue the replication activity of the mutants. Comparing the efficiency of SFV and FHV trans-replication systems, the FHV system appeared to produce more RNA. Using fluorescent marker proteins, we demonstrated that both systems could replicate in the same cell. This work may facilitate the comparative analysis of FHV and SFV replication.
APA, Harvard, Vancouver, ISO, and other styles
31

Odegard, Amy L., Maggie H. Kwan, Hanna E. Walukiewicz, Manidipa Banerjee, Anette Schneemann, and John E. Johnson. "Low Endocytic pH and Capsid Protein Autocleavage Are Critical Components of Flock House Virus Cell Entry." Journal of Virology 83, no. 17 (June 24, 2009): 8628–37. http://dx.doi.org/10.1128/jvi.00873-09.

Full text
Abstract:
ABSTRACT The process by which nonenveloped viruses cross cell membranes during host cell entry remains poorly defined; however, common themes are emerging. Here, we use correlated in vivo and in vitro studies to understand the mechanism of Flock House virus (FHV) entry and membrane penetration. We demonstrate that low endocytic pH is required for FHV infection, that exposure to acidic pH promotes FHV-mediated disruption of model membranes (liposomes), and particles exposed to low pH in vitro exhibit increased hydrophobicity. In addition, FHV particles perturbed by heating displayed a marked increase in liposome disruption, indicating that membrane-active regions of the capsid are exposed or released under these conditions. We also provide evidence that autoproteolytic cleavage, to generate the lipophilic γ peptide (4.4 kDa), is required for membrane penetration. Mutant, cleavage-defective particles failed to mediate liposome lysis, regardless of pH or heat treatment, suggesting that these particles are not able to expose or release the requisite membrane-active regions of the capsid, namely, the γ peptides. Based on these results, we propose an updated model for FHV entry in which (i) the virus enters the host cell by endocytosis, (ii) low pH within the endocytic pathway triggers the irreversible exposure or release of γ peptides from the virus particle, and (iii) the exposed/released γ peptides disrupt the endosomal membrane, facilitating translocation of viral RNA into the cytoplasm.
APA, Harvard, Vancouver, ISO, and other styles
32

Nakase, Ikuhiko, Hisaaki Hirose, Gen Tanaka, Akiko Tadokoro, Sachiko Kobayashi, Toshihide Takeuchi, and Shiroh Futaki. "Cell-surface Accumulation of Flock House Virus-derived Peptide Leads to Efficient Internalization via Macropinocytosis." Molecular Therapy 17, no. 11 (November 2009): 1868–76. http://dx.doi.org/10.1038/mt.2009.192.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Eckerle, Lance D., César G. Albariño, and L. Andrew Ball. "Flock House virus subgenomic RNA3 is replicated and its replication correlates with transactivation of RNA2." Virology 317, no. 1 (December 2003): 95–108. http://dx.doi.org/10.1016/j.virol.2003.08.029.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Stamilla, Alessandro, Antonino Messina, Lucia Condorelli, Francesca Licitra, Francesco Antoci, Massimiliano Lanza, Guido Ruggero Loria, Giuseppe Cascone, and Roberto Puleio. "Morphological and Immunohistochemical Examination of Lymphoproliferative Lesions Caused by Marek’s Disease Virus in Breeder Chickens." Animals 10, no. 8 (July 27, 2020): 1280. http://dx.doi.org/10.3390/ani10081280.

Full text
Abstract:
Marek’s disease is widely controlled by vaccination programs; however, chickens are not totally protected, especially immediately after the vaccination when a strong challenge could interfere with the effectiveness of vaccination in the absence of proper biosecurity practice. This case report describes the occurrence of Marek’s disease (MD) observed in a breeder chicken flock reared southeast of Sicily. MD outbreak occurred from 32 to 47 weeks with an increase in weekly mortality rate (+0.4–0.6%). Overall, mortality rate related to Marek’s disease was about 6% at the end of the cycle. Carcasses of chickens found during the occurrence of disease underwent necropsy, and tissues were collected to confirm the infection. Gizzard, cecal tonsil, intestine, spleen and tumor mass were collected and analyzed from a carcass of one hen, 32 weeks old and apparently asymptomatic. Multiplex real-time PCR performed on spleen tissues detected the presence of MD virus pathogenic strain. Macroscopic and microscopic evaluation of the rest of the samples confirmed the neoplastic disease. Moreover, the immunophenotype of the tumor cells was identified as CD3 positive by immunohistochemical (IHC) staining. The vaccinated flock had become rapidly infected with the MD virus, which proves that the challenge of the MD virus was too strong in the rearing house at the beginning of the cycle, causing the outbreak.
APA, Harvard, Vancouver, ISO, and other styles
35

Yang, Chenghuai, Leyi Wang, Kent Schwartz, Eric Burrough, Jennifer Groeltz-Thrush, Qi Chen, Ying Zheng, Huigang Shen, and Ganwu Li. "Case Report and Genomic Characterization of a Novel Porcine Nodavirus in the United States." Viruses 13, no. 1 (January 7, 2021): 73. http://dx.doi.org/10.3390/v13010073.

Full text
Abstract:
Nodaviruses are small bisegmented RNA viruses belonging to the family Nodaviridae. Nodaviruses have been identified in different hosts, including insects, fishes, shrimps, prawns, dogs, and bats. A novel porcine nodavirus was first identified in the United States by applying next-generation sequencing on brain tissues of pigs with neurological signs, including uncontrollable shaking. RNA1 of the porcine nodavirus had the highest nucleotide identity (51.1%) to the Flock House virus, whereas its RNA2 shared the highest nucleotide identity (48%) with the RNA2 segment of caninovirus (Canine nodavirus). Genetic characterization classified porcine nodavirus as a new species under the genus Alphanodavirus. Further studies are needed to understand the pathogenicity and clinical impacts of this virus.
APA, Harvard, Vancouver, ISO, and other styles
36

Banerjee, Manidipa, Jeffrey A. Speir, Maggie H. Kwan, Rick Huang, Peyman P. Aryanpur, Brian Bothner, and John E. Johnson. "Structure and Function of a Genetically Engineered Mimic of a Nonenveloped Virus Entry Intermediate." Journal of Virology 84, no. 9 (February 17, 2010): 4737–46. http://dx.doi.org/10.1128/jvi.02670-09.

Full text
Abstract:
ABSTRACT Divalent metal ions are components of numerous icosahedral virus capsids. Flock House virus (FHV), a small RNA virus of the family Nodaviridae, was utilized as an accessible model system with which to address the effects of metal ions on capsid structure and on the biology of virus-host interactions. Mutations at the calcium-binding sites affected FHV capsid stability and drastically reduced virus infectivity, without altering the overall architecture of the capsid. The mutations also altered the conformation of gamma, a membrane-disrupting, virus-encoded peptide usually sequestered inside the capsid, by increasing its exposure under neutral pH conditions. Our data demonstrate that calcium binding is essential for maintaining a pH-based control on gamma exposure and host membrane disruption, and they reveal a novel rationale for the metal ion requirement during virus entry and infectivity. In the light of the phenotypes displayed by a calcium site mutant of FHV, we suggest that this mutant corresponds to an early entry intermediate formed in the endosomal pathway.
APA, Harvard, Vancouver, ISO, and other styles
37

Johnson, K. L., and L. A. Ball. "Replication of flock house virus RNAs from primary transcripts made in cells by RNA polymerase II." Journal of virology 71, no. 4 (1997): 3323–27. http://dx.doi.org/10.1128/jvi.71.4.3323-3327.1997.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Fisher, A. J., B. R. McKinney, A. Schneemann, R. R. Rueckert, and J. E. Johnson. "Crystallization of viruslike particles assembled from flock house virus coat protein expressed in a baculovirus system." Journal of Virology 67, no. 5 (1993): 2950–53. http://dx.doi.org/10.1128/jvi.67.5.2950-2953.1993.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Bong, Dennis T., Claudia Steinern, Andreas Janshoff, John E. Johnson, and M. Reza Ghadiri. "A highly membrane-active peptide in Flock House virus: implications for the mechanism of nodavirus infection." Chemistry & Biology 6, no. 7 (July 1999): 473–81. http://dx.doi.org/10.1016/s1074-5521(99)80065-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Kampmueller, Kathryn M., and David J. Miller. "The Cellular Chaperone Heat Shock Protein 90 Facilitates Flock House Virus RNA Replication in Drosophila Cells." Journal of Virology 79, no. 11 (June 1, 2005): 6827–37. http://dx.doi.org/10.1128/jvi.79.11.6827-6837.2005.

Full text
Abstract:
ABSTRACT The assembly of viral RNA replication complexes on intracellular membranes represents a critical step in the life cycle of positive-strand RNA viruses. We investigated the role of the cellular chaperone heat shock protein 90 (Hsp90) in viral RNA replication complex assembly and function using Flock House virus (FHV), an alphanodavirus whose RNA-dependent RNA polymerase, protein A, is essential for viral RNA replication complex assembly on mitochondrial outer membranes. The Hsp90 chaperone complex transports cellular mitochondrial proteins to the outer mitochondrial membrane import receptors, and thus we hypothesized that Hsp90 may also facilitate FHV RNA replication complex assembly or function. Treatment of FHV-infected Drosophila S2 cells with the Hsp90-specific inhibitor geldanamycin or radicicol potently suppressed the production of infectious virions and the accumulation of protein A and genomic, subgenomic, and template viral RNA. In contrast, geldanamycin did not inhibit the activity of preformed FHV RNA replication complexes. Hsp90 inhibitors also suppressed viral RNA and protein A accumulation in S2 cells expressing an FHV RNA replicon. Furthermore, Hsp90 inhibition with either geldanamycin or RNAi-mediated chaperone downregulation suppressed protein A accumulation in the absence of viral RNA replication. These results identify Hsp90 as a host factor involved in FHV RNA replication and suggest that FHV uses established cellular chaperone pathways to assemble its RNA replication complexes on intracellular membranes.
APA, Harvard, Vancouver, ISO, and other styles
41

Lindenbach, Brett D., Jean-Yves Sgro, and Paul Ahlquist. "Long-Distance Base Pairing in Flock House Virus RNA1 Regulates Subgenomic RNA3 Synthesis and RNA2 Replication." Journal of Virology 76, no. 8 (April 15, 2002): 3905–19. http://dx.doi.org/10.1128/jvi.76.8.3905-3919.2002.

Full text
Abstract:
ABSTRACT Replication of flock house virus (FHV) RNA1 and production of subgenomic RNA3 in the yeast Saccharomyces cerevisiae provide a useful tool for the dissection of FHV molecular biology and host-encoded functions involved in RNA replication. The replication template activity of RNA1 can be separated from its coding potential by supplying the RNA1-encoded replication factor protein A in trans. We constructed a trans-replication system in yeast to examine cis-acting elements in RNA1 that control RNA3 production, as well as RNA1 and RNA2 replication. Two cis elements controlling RNA3 production were found. A proximal subgenomic control element was located just upstream of the RNA3 start site (nucleotides [nt] 2282 to 2777). A short distal element also controlling RNA3 production (distal subgenomic control element) was identified 1.5 kb upstream, at nt 1229 to 1239. Base pairing between these distal and proximal elements was shown to be essential for RNA3 production by covariation analysis and in vivo selection of RNA3-expressing replicons from plasmid libraries containing random sequences in the distal element. Two distinct RNA1 replication elements (RE) were mapped within the 3′ quarter of RNA1: the intRE (nt 2322 to 2501) and the 3′RE (nt 2735 to 3011). The 3′RE significantly overlaps the RNA3 region in RNA1, and this information was applied to produce improved RNA3-based vectors for foreign-gene expression. In addition, replication of an RNA2 derivative was dependent on RNA1 templates capable of forming the long-distance interaction that controls RNA3 production.
APA, Harvard, Vancouver, ISO, and other styles
42

Zhou, Yiyang, Payal D. Maharaj, Jyothi K. Mallajosyula, Alison A. McCormick, and Christopher M. Kearney. "In planta Production of Flock House Virus Transencapsidated RNA and Its Potential Use as a Vaccine." Molecular Biotechnology 57, no. 4 (November 29, 2014): 325–36. http://dx.doi.org/10.1007/s12033-014-9826-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Jeřábková, Jana, Růžena Juranová, Kateřina Rosenbergová, Libuše Kulíková, Alfred Hera, Petr Lány, Oldřich Kubíček, and Josef Koláček. "Detection of the Newcastle disease virus and its effect on development of post-vaccination immunity in a commercial flock of laying hens." Acta Veterinaria Brno 81, no. 1 (2012): 3–8. http://dx.doi.org/10.2754/avb201281010003.

Full text
Abstract:
The aim of this study was to monitor the concentration of antibodies against Newcastle disease after vaccination of laying hens at the beginning and in the end of the laying period. The study was carried out in one commercial flock of laying hens in Opatovice in the Czech Republic in the years 2008-2010. A total of 280 samples of blood sera were taken from laying hens coming from four poultry houses. The sera were tested by the haemagglutination inhibition test according to the OIE Manual. Virological testing was conducted as a consequence of atypical results of serological testing. Newcastle disease virus RNA was proved by the RT-nested PCR method in the pooled tissue samples of 5 hens, in the samples of intestines with ileocaecal tonsila, in trachea and also in one swab sample from the environment of one house. Based on sequencing analysis and subsequent phylogenetic analysis, the virus was identified as a low pathogenic strain of paramyxovirus (PMV-1). This low pathogenic strain did not have any impact on the health of laying hens.
APA, Harvard, Vancouver, ISO, and other styles
44

Bedasa, Chala, Ararsa Duguma, Asamenew Tesfaye, and Tadele Tolosa. "Seroprevalence of infectious bursal disease and its potential risk factors in backyard chicken production of Waliso district, South Western Shoa Zone, Ethiopia." Veterinary Integrative Sciences 20, no. 1 (September 6, 2021): 95–105. http://dx.doi.org/10.12982/vis.2022.009.

Full text
Abstract:
A cross sectional study on infectious bursal disease was conducted in apparently healthy backyard chicken at Waliso district of Southwestern Shoa, central oromia, Ethiopia from from November, 2018 to October, 2019. A total of 282 chickens were randomly selected to estimate seroprevalence of IBD infection and to identify the likely potential risk factors for the disease. Serum samples collected and serological test conducted in laboratory at National Animal Health Diagnosis and Investigation Center Sebeta, Ethopia. Out of 282 serum samples tested 224 were positive for indirect ELISA technique and the overall seroprevalence of IBDV in the study area was found to be 79.43% at individual level. Educational level of owners, kebeles and flock size significantly affect seroprevalence of IBD in the study area. The effect of difference in managements like source of replacement, frequency of house cleaning, use of disinfectant and isolation practice has a significant effect on IBDV sero-prevalence. A lower seroprevalence of IBDV was reported in good hygienic level of house (26.7%) than poor level of chicken house hygiene (96.4%) with statistically significant difference (P < 0.05). The seroprevalence of IBDV in the present study associated with chicken management, flock size, owner education level and other animal related risk factors for occurrence of the disease. Therefore, awareness on chicken health management, and importance of immunization would help to minimize the prevalence of the disease and play crucial role in the control of the disease. Furthermore, characterizing virus strains circulating in the area in future study is recommended.
APA, Harvard, Vancouver, ISO, and other styles
45

Jaworski, Elizabeth, and Andrew Routh. "Parallel ClickSeq and Nanopore sequencing elucidates the rapid evolution of defective-interfering RNAs in Flock House virus." PLOS Pathogens 13, no. 5 (May 5, 2017): e1006365. http://dx.doi.org/10.1371/journal.ppat.1006365.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Wu, Wenzhe, Zhaowei Wang, Hongjie Xia, Yongxiang Liu, Yang Qiu, Yujie Liu, Yuanyang Hu, and Xi Zhou. "Flock House Virus RNA Polymerase Initiates RNA Synthesis De Novo and Possesses a Terminal Nucleotidyl Transferase Activity." PLoS ONE 9, no. 1 (January 23, 2014): e86876. http://dx.doi.org/10.1371/journal.pone.0086876.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Schneemann, A., R. Dasgupta, J. E. Johnson, and R. R. Rueckert. "Use of recombinant baculoviruses in synthesis of morphologically distinct viruslike particles of flock house virus, a nodavirus." Journal of Virology 67, no. 5 (1993): 2756–63. http://dx.doi.org/10.1128/jvi.67.5.2756-2763.1993.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Albariño, César G., Lance D. Eckerle, and L. Andrew Ball. "The cis-acting replication signal at the 3′ end of Flock House virus RNA2 is RNA3-dependent." Virology 311, no. 1 (June 2003): 181–91. http://dx.doi.org/10.1016/s0042-6822(03)00190-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Buratti, E., S. G. Tisminetzky, E. S. Scodeller, and F. E. Baralle. "Conformational display of two neutralizing epitopes of HIV-1 gp41 on the Flock House virus capsid protein." Journal of Immunological Methods 197, no. 1-2 (January 1996): 7–18. http://dx.doi.org/10.1016/0022-1759(96)00097-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Venter, P. Arno, and Anette Schneemann. "Assembly of Two Independent Populations of Flock House Virus Particles with Distinct RNA Packaging Characteristics in the Same Cell." Journal of Virology 81, no. 2 (November 1, 2006): 613–19. http://dx.doi.org/10.1128/jvi.01668-06.

Full text
Abstract:
ABSTRACT Flock House virus (FHV; Nodaviridae) is a positive-strand RNA virus that encapsidates a bipartite genome consisting of RNA1 and RNA2. We recently showed that specific recognition of these RNAs for packaging into progeny particles requires coat protein translated from replicating viral RNA. In the present study, we investigated whether the entire assembly pathway, i.e., the formation of the initial nucleating complex and the subsequent completion of the capsid, is restricted to the same pool of coat protein subunits. To test this, coat proteins carrying either FLAG or hemagglutinin epitopes were synthesized from replicating or nonreplicating RNA in the same cell, and the resulting particle population and its RNA packaging phenotype were analyzed. Results from immunoprecipitation analysis and ion-exchange chromatography showed that the differentially tagged proteins segregated into two distinct populations of virus particles with distinct RNA packaging phenotypes. Particles assembled from coat protein that was translated from replicating RNA contained the FHV genome, whereas particles assembled from coat protein that was translated from nonreplicating mRNA contained random cellular RNA. These data demonstrate that only coat proteins synthesized from replicating RNA partake in the assembly of virions that package the viral genome and that RNA replication, coat protein translation, and virion assembly are processes that are tightly coupled during the life cycle of FHV.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography