Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 162
Filter
1.
Virus Res ; 246: 28-34, 2018 02 15.
Article in English | MEDLINE | ID: mdl-29337162

ABSTRACT

Recently, a novel antiviral compound (K22) that inhibits replication of a broad range of animal and human coronaviruses was reported to interfere with viral RNA synthesis by impairing double-membrane vesicle (DMV) formation (Lundin et al., 2014). Here we assessed potential antiviral activities of K22 against a range of viruses representing two (sub)families of the order Nidovirales, the Arteriviridae (porcine reproductive and respiratory syndrome virus [PRRSV], equine arteritis virus [EAV] and simian hemorrhagic fever virus [SHFV]), and the Torovirinae (equine torovirus [EToV] and White Bream virus [WBV]). Possible effects of K22 on nidovirus replication were studied in suitable cell lines. K22 concentrations significantly decreasing infectious titres of the viruses included in this study ranged from 25 to 50 µM. Reduction of double-stranded RNA intermediates of viral replication in nidovirus-infected cells treated with K22 confirmed the anti-viral potential of K22. Collectively, the data show that K22 has antiviral activity against diverse lineages of nidoviruses, suggesting that the inhibitor targets a critical and conserved step during nidovirus replication.


Subject(s)
Antiviral Agents/pharmacology , Arterivirus/drug effects , Benzamides/pharmacology , Coronaviridae/drug effects , Equartevirus/drug effects , Piperidines/pharmacology , Porcine respiratory and reproductive syndrome virus/drug effects , Torovirus/drug effects , Animals , Arterivirus/genetics , Arterivirus/growth & development , Arterivirus/metabolism , Carps , Cell Line , Chlorocebus aethiops , Coronaviridae/genetics , Coronaviridae/growth & development , Coronaviridae/metabolism , Epithelial Cells/drug effects , Epithelial Cells/virology , Equartevirus/genetics , Equartevirus/growth & development , Equartevirus/metabolism , Mesocricetus , Porcine respiratory and reproductive syndrome virus/genetics , Porcine respiratory and reproductive syndrome virus/growth & development , Porcine respiratory and reproductive syndrome virus/metabolism , RNA, Double-Stranded/antagonists & inhibitors , RNA, Double-Stranded/biosynthesis , RNA, Double-Stranded/genetics , RNA, Viral/antagonists & inhibitors , RNA, Viral/biosynthesis , RNA, Viral/genetics , Torovirus/genetics , Torovirus/growth & development , Torovirus/metabolism , Virus Replication/drug effects
2.
mBio ; 4(1): e00611-12, 2013 Feb 19.
Article in English | MEDLINE | ID: mdl-23422412

ABSTRACT

The recent emergence of a novel human coronavirus (HCoV-EMC) in the Middle East raised considerable concerns, as it is associated with severe acute pneumonia, renal failure, and fatal outcome and thus resembles the clinical presentation of severe acute respiratory syndrome (SARS) observed in 2002 and 2003. Like SARS-CoV, HCoV-EMC is of zoonotic origin and closely related to bat coronaviruses. The human airway epithelium (HAE) represents the entry point and primary target tissue for respiratory viruses and is highly relevant for assessing the zoonotic potential of emerging respiratory viruses, such as HCoV-EMC. Here, we show that pseudostratified HAE cultures derived from different donors are highly permissive to HCoV-EMC infection, and by using reverse transcription (RT)-PCR and RNAseq data, we experimentally determined the identity of seven HCoV-EMC subgenomic mRNAs. Although the HAE cells were readily responsive to type I and type III interferon (IFN), we observed neither a pronounced inflammatory cytokine nor any detectable IFN responses following HCoV-EMC, SARS-CoV, or HCoV-229E infection, suggesting that innate immune evasion mechanisms and putative IFN antagonists of HCoV-EMC are operational in the new host. Importantly, however, we demonstrate that both type I and type III IFN can efficiently reduce HCoV-EMC replication in HAE cultures, providing a possible treatment option in cases of suspected HCoV-EMC infection. IMPORTANCE A novel human coronavirus, HCoV-EMC, has recently been described to be associated with severe respiratory tract infection and fatalities, similar to severe acute respiratory syndrome (SARS) observed during the 2002-2003 epidemic. Closely related coronaviruses replicate in bats, suggesting that, like SARS-CoV, HCoV-EMC is of zoonotic origin. Since the animal reservoir and circumstances of zoonotic transmission are yet elusive, it is critically important to assess potential species barriers of HCoV-EMC infection. An important first barrier against invading respiratory pathogens is the epithelium, representing the entry point and primary target tissue of respiratory viruses. We show that human bronchial epithelia are highly susceptible to HCoV-EMC infection. Furthermore, HCoV-EMC, like other coronaviruses, evades innate immune recognition, reflected by the lack of interferon and minimal inflammatory cytokine expression following infection. Importantly, type I and type III interferon treatment can efficiently reduce HCoV-EMC replication in the human airway epithelium, providing a possible avenue for treatment of emerging virus infections.


Subject(s)
Coronaviridae/physiology , Coronaviridae/pathogenicity , Epithelial Cells/virology , Virus Replication , Animals , Coronaviridae/growth & development , Coronaviridae/isolation & purification , Coronavirus Infections/virology , Cytokines/biosynthesis , Humans , Immune Evasion , Middle East , RNA, Messenger/biosynthesis , RNA, Messenger/genetics , RNA, Viral/biosynthesis , RNA, Viral/genetics , Respiratory Tract Infections/virology , Reverse Transcriptase Polymerase Chain Reaction , Sequence Analysis, RNA , Zoonoses/virology
3.
New Microbiol ; 25(3): 285-90, 2002 Jul.
Article in English | MEDLINE | ID: mdl-12173769

ABSTRACT

A strain of porcine epidemic diarrhea virus (PEDV), P-5V, utilized as a live virus vaccine in Japan was infected to a swine cell lines, KSEK6 and IB-RS-2 cells. Clear CPE, characterized by cellular destruction, started to appear in the infected cells on 2-3 days post infection (DPI) and affected cells was completely degenerated on 4 DPI. The virus was serially passaged in the cells even without addition of trypsin. Small but clear plaques were formed under an agar overlay medium on the cells. The infective titer in the order of 10(7.00-7.50) TCID50 per ml was obtained at usual incubation temperature.


Subject(s)
Coronaviridae Infections/veterinary , Coronaviridae/growth & development , Swine Diseases/virology , Animals , Cells, Cultured , Coronaviridae/immunology , Coronaviridae Infections/immunology , Coronaviridae Infections/prevention & control , Coronaviridae Infections/virology , Cytopathogenic Effect, Viral/immunology , Swine , Swine Diseases/immunology , Swine Diseases/prevention & control , Viral Plaque Assay/veterinary , Viral Vaccines/immunology
4.
J Virol ; 68(10): 6523-34, 1994 Oct.
Article in English | MEDLINE | ID: mdl-8083990

ABSTRACT

The prevailing hypothesis is that the intracellular site of budding of coronaviruses is determined by the localization of its membrane protein M (previously called E1). We tested this by analyzing the site of budding of four different coronaviruses in relation to the intracellular localization of their M proteins. Mouse hepatitis virus (MHV) and infectious bronchitis virus (IBV) grown in Sac(-) cells, and feline infectious peritonitis virus (FIPV) and transmissible gastroenteritis virus (TGEV) grown in CrFK cells, all budded exclusively into smooth-walled, tubulovesicular membranes located intermediately between the rough endoplasmic reticulum and Golgi complex, identical to the so-called budding compartment previously identified for MHV. Indirect immunofluorescence staining of the infected cells showed that all four M proteins accumulated in a perinuclear region. Immunogold microscopy localized MHV M and IBV M in the budding compartment; in addition, a dense labeling in the Golgi complex occurred, MHV M predominantly in trans-Golgi cisternae and trans-Golgi reticulum and IBV M mainly in the cis and medial Golgi cisternae. The corresponding M proteins of the four viruses, when independently expressed in a recombinant vaccinia virus system, also accumulated in the perinuclear area. Quantitative pulse-chase analysis of metabolically labeled cells showed that in each case the majority of the M glycoproteins carried oligosaccharide side chains with Golgi-specific modifications within 4 h after synthesis. Immunoelectron microscopy localized recombinant MHV M and IBV M to the same membranes as the respective proteins in coronavirus-infected cells, with the same cis-trans distribution over the Golgi complex. Our results demonstrate that some of the M proteins of the four viruses are transported beyond the budding compartment and are differentially retained by intrinsic retention signals; in addition to M, other viral and/or cellular factors are probably required to determine the site of budding.


Subject(s)
Coronaviridae/growth & development , Golgi Apparatus/microbiology , Viral Matrix Proteins/biosynthesis , Animals , Cats , Cell Line , Coronaviridae/metabolism , Coronaviridae/ultrastructure , Coronavirus, Feline/growth & development , Coronavirus, Feline/metabolism , Coronavirus, Feline/ultrastructure , Fetus , Golgi Apparatus/metabolism , Golgi Apparatus/ultrastructure , Humans , Infectious bronchitis virus/growth & development , Infectious bronchitis virus/metabolism , Infectious bronchitis virus/ultrastructure , Kidney , Microscopy, Electron , Microscopy, Immunoelectron , Murine hepatitis virus/growth & development , Murine hepatitis virus/metabolism , Murine hepatitis virus/ultrastructure , Recombination, Genetic , Swine , Transmissible gastroenteritis virus/growth & development , Transmissible gastroenteritis virus/metabolism , Transmissible gastroenteritis virus/ultrastructure , Tumor Cells, Cultured , Viral Matrix Proteins/analysis
5.
J Virol Methods ; 43(1): 53-64, 1993 Jun.
Article in English | MEDLINE | ID: mdl-7689580

ABSTRACT

Plaque assays under Sephadex or agarose overlays are described for rat coronaviruses (RCVs) grown in L2 mouse fibroblasts. A plaque assay using Sephadex was simple; however, viable plaques could not be collected for propagation, and fixation was necessary before evaluation. Plaque formation under agarose was optimized using diethylaminoethyl-dextran (DEAE-D) in the pre-treatment and absorption media and trypsin added to the absorption media and agarose overlay. The use of DEAE-D alone, trypsin alone or trypsin combined with DEAE-D significantly increased plaque numbers and visibility. Plaque numbers were highest when pre-treatment media contained DEAE-D, absorption media contained DEAE-D and trypsin, and the agarose overlay contained trypsin. The assay was useful for plaque isolation and quantification of sialodacryoadenitis virus (SDA), Parker's rat coronavirus (PRCV) and other coronavirus isolates from rats and its specificity was demonstrated by plaque-reduction neutralization testing. These methods will facilitate production of cloned virus stocks for study of RCV biology and virus quantification for in vitro and in vivo studies of RCVs.


Subject(s)
Coronaviridae/growth & development , Rats/microbiology , Viral Plaque Assay , Animals , Antibodies, Viral/immunology , Cell Line , Coronaviridae/immunology , Coronaviridae/isolation & purification , Coronaviridae Infections/immunology , DEAE-Dextran , Dextrans , Fibroblasts/microbiology , Mice , Neutralization Tests , Rats/immunology , Sepharose , Specific Pathogen-Free Organisms , Trypsin , Virus Cultivation
6.
J Gen Virol ; 73 ( Pt 12): 3285-8, 1992 Dec.
Article in English | MEDLINE | ID: mdl-1335030

ABSTRACT

The growth of feline enteric coronavirus strain 79-1683 in whole feline embryo cells was inhibited by the presence of 1 microgram/ml of actinomycin D in the culture fluid. No virus-specific mRNAs could be detected in such cultures and yields of infectious virus were depressed by > 99%. By contrast, the antigenically related feline infectious peritonitis virus strain 79-1146 was unaffected by the presence of actinomycin D, indicating a fundamental difference between the two feline coronavirus strains in their requirements for host-encoded function(s).


Subject(s)
Coronaviridae/growth & development , Coronavirus, Feline/growth & development , Dactinomycin/pharmacology , Virus Replication/drug effects , Animals , Cats/microbiology , Drug Administration Schedule , RNA, Viral/metabolism
7.
J Gen Virol ; 73 ( Pt 10): 2731-6, 1992 Oct.
Article in English | MEDLINE | ID: mdl-1402806

ABSTRACT

The gene encoding the membrane (M) protein of the OC43 strain of human coronavirus (HCV-OC43) was amplified by a reverse transcription-polymerase chain reaction of viral RNA with HCV-OC43- and bovine coronavirus (BCV)-specific primers. The nucleotide sequence of the cloned 1.5 kb fragment revealed an open reading frame (ORF) of 690 nucleotides which was identified as the M protein gene from its homology to BCV. This ORF encodes a protein of 230 amino acids with an M(r) of 26416. The gene is preceded by the motif UCCAAAC, analogous to the consensus coronavirus transcription initiation sequence. The M protein of HCV-OC43 shows features typical of all coronavirus M proteins studied: a hydrophilic, presumably external N terminus including about 10% of the protein, and a potential N-glycosylation site followed by three major hydrophobic transmembrane domains. The amino acid sequence of the M protein of HCV-OC43 has 94% identity with that of the Mebus strain of BCV, and also contains six potential O-glycosylation sites in the exposed N-terminal domain. Indeed, the glycosylation of the M protein was not inhibited in the presence of tunicamycin, which is indicative of O-glycosylation, as previously reported for BCV and murine hepatitis virus. Virions released from tunicamycin-treated cells contained the M glycoprotein but were devoid of both peplomer (S) and haemagglutinin-esterase (HE) proteins. Thus, inhibition of the N-glycosylation of the S and HE structural proteins prevented their incorporation into progeny virions, an indication that they are dispensable for virion morphogenesis, unlike the M protein.


Subject(s)
Coronaviridae/genetics , Genes, Viral/genetics , RNA, Viral/genetics , Viral Matrix Proteins/genetics , Viral Structural Proteins/genetics , Amino Acid Sequence , Base Sequence , Cells, Cultured , Cloning, Molecular , Coronaviridae/growth & development , Coronaviridae/metabolism , Coronavirus M Proteins , Glycosylation , Humans , Molecular Sequence Data , Open Reading Frames/genetics , Polymerase Chain Reaction , Protein Processing, Post-Translational/drug effects , Sequence Analysis , Sequence Homology, Amino Acid , Tunicamycin/pharmacology , Viral Matrix Proteins/metabolism , Virion/growth & development
8.
Microbiologica ; 15(3): 313-7, 1992 Jul.
Article in English | MEDLINE | ID: mdl-1323743

ABSTRACT

The viral susceptibility of a cell line, named KSEK6, newly established from the kidney cortex of swine embryo was tested for the indication of CPE occurrences and also plaque formations. The multiplication of porcine adenoviruses was considerably high in the cells among the virus strains tested though plaques of these viruses were hardly visible under agar overlay medium. Two strains of swine enteroviruses, Aujeszky's disease virus and hemagglutinating encephalomyelitis virus also multiplied well in a similar order to those received in the other cells employed.


Subject(s)
Cell Line , Kidney Cortex/microbiology , Viruses/growth & development , Adenoviridae/growth & development , Animals , Coronaviridae/growth & development , Cytopathogenic Effect, Viral , Herpesvirus 1, Suid/growth & development , Swine , Viral Plaque Assay
9.
J Vet Med Sci ; 54(2): 313-8, 1992 Apr.
Article in English | MEDLINE | ID: mdl-1318752

ABSTRACT

Porcine epidemic diarrhea virus (PEDV) was isolated in Vero cell cultures from the small intestine of a piglet experimentally infected with porcine coronavirus 83P-5, that had been isolated during outbreaks of porcine acute diarrhea and passaged in piglets. The isolation of the PEDV was successful only in Vero cells maintained in the maintenance medium (MM) containing trypsin. Infected Vero cell cultures exhibited CPE characterized by cell-fusion and syncytial formation, as well as cytoplasmic fluorescence when examined by the indirect immunofluorescent test using rabbit anti-83P-5 virus serum. The isolate was adapted to serial propagation in Vero cell cultures by adding trypsin to MM. Vero cell-adapted PEDV was successfully propagated in the MA104, CPK and ESK cell lines in the presence of trypsin in MM. Vero cell-adapted PEDV had morphologic and physicochemical characteristics similar to those of other members of the coronaviridae. The isolate differed serologically from porcine transmissible gastroenteritis (TGE) and porcine hemagglutinating encephalomyelitis viruses, and no antigenic relationship between the isolate and TGE virus could be detected by the indirect immunofluorescent test. Attempts to isolate PEDV in 6 types of primary fetal pig cell cultures and 6 of 10 established cell lines resulted in the failure, probably because these cells were damaged by the action of trypsin.


Subject(s)
Coronaviridae Infections/veterinary , Coronaviridae/growth & development , Diarrhea/veterinary , Swine Diseases/microbiology , Animals , Cell Fusion , Cell Line , Cells, Cultured , Coronaviridae/isolation & purification , Coronaviridae/ultrastructure , Coronaviridae Infections/microbiology , Culture Media , Cytopathogenic Effect, Viral , Diarrhea/microbiology , Fluorescent Antibody Technique , Giant Cells , Microscopy, Electron , Serial Passage , Swine , Trypsin , Vero Cells
10.
J Med Virol ; 35(2): 142-9, 1991 Oct.
Article in English | MEDLINE | ID: mdl-1662706

ABSTRACT

Cellular susceptibility to fusion mediated by murine coronavirus (mouse hepatitis virus, MHV strain A59) was separated into lipid-dependent and lipid-independent mechanisms with the use of subclones and selected mutants of mouse L-2 fibroblasts. Fusion-resistant L-2 cell mutants had similar cholesterol and fatty acid composition as did their fusion-susceptible parent subclone, and were presumably deficient in a genetically mutable non-lipid, host cell factor (e.g., fusion protein receptor). On the other hand, cellular sensitivity to virus fusion, which is known to be influenced by cell cholesterol content [Daya et al., 1988], was shown further to be modulated by homeostatic alterations in fatty acid metabolism. Cholesterol supplementation of mouse L-2 fibroblasts or of peritoneal macrophages from MHV-susceptible mice elevated susceptibility to viral fusion. Increased fusion susceptibility occurred in cholesterol-supplemented L-2 cells in the absence of any detectable alterations in host cell fatty acid composition, thus demonstrating fusion enhancement by cholesterol alone. L-2 cells cloned by limiting dilution in normal (not cholesterol-supplemented) medium were found to be heterogeneous in cholesterol content. Interestingly, high cholesterol-containing subclones had increased levels of C-18:0, C-18:2, C-20:4, and C-22:6 and markedly reduced levels of C-18:1 fatty acids when compared to low cholesterol-containing subclones. High cholesterol-containing subclones did not show enhanced susceptibility to viral fusion, suggesting that homeostatic alteration of fatty acid metabolism compensated for the increased cholesterol levels and countered the normally fusion-enhancing effect of cholesterol alone.(ABSTRACT TRUNCATED AT 250 WORDS)


Subject(s)
Cholesterol/metabolism , Coronaviridae/growth & development , Fatty Acids/metabolism , Animals , Cell Fusion , Cells, Cultured , Cholesterol, Dietary/pharmacology , Chromatography, Thin Layer , Coronaviridae Infections/metabolism , Coronaviridae Infections/microbiology , Fibroblasts/microbiology , Homeostasis , Macrophages/microbiology , Mice , Mice, Inbred Strains
11.
J Virol ; 65(6): 3219-26, 1991 Jun.
Article in English | MEDLINE | ID: mdl-2033672

ABSTRACT

Two murine hepatitis virus strain A59 defective interfering (DI) RNAs were generated by undiluted virus passages. The DI RNAs were encapsidated efficiently. The smallest DI particle, DI-a, contained a 5.5-kb RNA consisting of the following three noncontiguous regions from the MHV-A59 genome, which were joined in frame: the 5'-terminal 3.9 kb, a 798-nucleotide fragment from the 3' end of the polymerase gene, and the 3'-terminal 805 nucleotides. A full-length cDNA clone of the DI-a genome was constructed and cloned downstream of the bacteriophage T7 promoter. Transcripts derived from this clone, pMIDI, were used for transfection of MHV-A59-infected cells and found to be amplified and packaged. Deletion analysis of pMIDI allowed us to identify a 650-nucleotide region derived from the 3' end of the second open reading frame of the polymerase gene that was required for efficient encapsidation.


Subject(s)
Coronaviridae/genetics , DNA-Directed RNA Polymerases/genetics , Defective Viruses/genetics , RNA, Viral/chemistry , Virus Replication/genetics , Amino Acid Sequence , Animals , Base Sequence , Cells, Cultured , Cloning, Molecular , Coronaviridae/growth & development , Cytopathogenic Effect, Viral , Mice , Molecular Sequence Data , Mutagenesis , Open Reading Frames , Promoter Regions, Genetic , Restriction Mapping , T-Phages/genetics
13.
Virology ; 180(1): 221-8, 1991 Jan.
Article in English | MEDLINE | ID: mdl-1984649

ABSTRACT

Bovine coronavirus (BCV) and hemagglutinating encephalomyelitis virus (HEV) from swine were found to grow to high titers in MDCK I cells, a subline of Madin Darby canine kidney cells. Virus grown in these cells was used to isolate and purify the HE-protein. This protein has been shown recently to have acetylesterase activity and to function as the receptor-destroying enzyme of BCV. Here we show that HEV contains this enzyme, too. The glycoproteins were solubilized by treatment of virions with octylglucoside. Following centrifugation through a sucrose gradient the surface proteins S and HE (hemagglutinin-esterase) were obtained in purified form. After removal of the detergent by dialysis, HE formed rosettes as shown by electron microscopy. The purified HE protein retained acetylesterase activity and was able to function as a receptor-destroying enzyme rendering red blood cells resistant against agglutination by both coronaviruses. HE protein released from the viral membrane failed to agglutinate red blood cells. However, it was found to recognize glycoconjugates containing N-acetyl-9-O-acetylneuraminic acid as indicated by a binding assay with rat serum proteins blotted to nitrocellulose and by its ability to inhibit the hemagglutinating activity of BCV, HEV, and influenza C virus. The purified enzyme provides a useful tool for analyzing the cellular receptors for coronaviruses.


Subject(s)
Coronaviridae/enzymology , Hemagglutinins, Viral/metabolism , Receptors, Virus/metabolism , Viral Fusion Proteins , Viral Proteins/metabolism , Acetylesterase/isolation & purification , Acetylesterase/metabolism , Animals , Cattle , Cell Line , Coronaviridae/growth & development , Coronaviridae/ultrastructure , Electrophoresis, Polyacrylamide Gel , Hemagglutination Inhibition Tests , Hemagglutination Tests , Hemagglutination, Viral , Hemagglutinins, Viral/isolation & purification , Swine , Viral Proteins/isolation & purification , Virus Cultivation
14.
Arch Virol ; 118(1-2): 57-66, 1991.
Article in English | MEDLINE | ID: mdl-2048975

ABSTRACT

Cell lines of rodent origin were tested for susceptibility to infection with rat coronavirus (RCV), including sialodacryoadenitis virus (SDAV) and Parker's rat coronavirus (PRCV). LBC rat mammary adenocarcinoma cells were susceptible only if the cells were treated with diethylaminoethyl-dextran (DEAE-D). A recent report that RCVs grow well in L2 mouse fibroblast cells was confirmed and expanded. RCV infection of L2 cells was substantially enhanced by treatment of cells with trypsin but not by treatment with DEAE-D. Primary isolation of SDAV from experimentally infected rats was accomplished using trypsin-treated L2 cells. One of 13 additional cell lines tested (rat urinary bladder epithelium, RBL-02) supported growth of RCVs, and growth was slightly enhanced by DEAE-D, but not by trypsin. These refinements of in vitro growth conditions for RCVs should facilitate further studies of their basic biology and improve options for primary isolation.


Subject(s)
Cell Line/microbiology , Coronaviridae/growth & development , Animals , Coronaviridae/drug effects , Coronaviridae/isolation & purification , DEAE-Dextran/pharmacology , Mice , Microscopy, Fluorescence , Rats , Trypsin/pharmacology
15.
Neurology ; 40(6): 904-10, 1990 Jun.
Article in English | MEDLINE | ID: mdl-2161091

ABSTRACT

Vacuolar degeneration was constantly induced in the CNS of 4-week-old ICR mice by intracerebral or intranasal inoculation of JHM-CC virus, a small plaque mutant of mouse hepatitis virus (JHM). Most animals showed no symptoms or only mild hindlimb paresis. Irrespective of clinical manifestations, the virus was isolated from the CNS up to days 14 to 16. Viral antigen expression in the CNS tissue was most extensive around days 5 to 7 and became undetectable on day 14. Viral antigens were localized almost exclusively to neurons, and the temporal sequence of viral antigen distribution after intranasal inoculation clearly indicated the virus spread through the olfactory and limbic systems into the brainstem and spinal cord, and possible cell-to cell transmission of the virus within the CNS. Vacuolar changes, most conspicuous in the brainstem and spinal cord, were steadily progressive up to 4 weeks after infection, but became indistinct by 4 months. Although the distribution of vacuolar lesions largely agreed with that of viral antigen-positive cells, the severity of vacuolation did not correlate with that of inflammation. Intramyelinic splitting, periaxonal edema, and swollen neurites were major ultrastructural substrates for vacuolar changes. This model could provide a better understanding of new types of neurologic disorders associated with viral infections, including vacuolar myelopathy in AIDS.


Subject(s)
Central Nervous System Diseases/microbiology , Coronaviridae Infections/microbiology , Vacuoles/microbiology , Animals , Antigens, Viral/analysis , Central Nervous System Diseases/pathology , Coronaviridae/growth & development , Coronaviridae/immunology , Coronaviridae/ultrastructure , Coronaviridae Infections/pathology , Immunohistochemistry , Male , Mice , Mice, Inbred ICR , Microscopy, Electron , Vacuoles/pathology , Vacuoles/ultrastructure
16.
J Clin Microbiol ; 28(6): 1454-7, 1990 Jun.
Article in English | MEDLINE | ID: mdl-2166085

ABSTRACT

Fecal filtrates from cows with winter dysentery were inoculated into gnotobiotic and conventional calves, and a coronavirus was isolated from calf feces. Cytopathic effects were observed on human rectal tumor cells but not bovine cell cultures. The winter dysentery isolates morphologically and antigenically resembled the Mebus strain of bovine coronavirus.


Subject(s)
Cattle Diseases/microbiology , Coronaviridae Infections/veterinary , Coronaviridae/growth & development , Dysentery/veterinary , Animals , Cattle , Coronaviridae/isolation & purification , Coronaviridae Infections/complications , Cytopathogenic Effect, Viral , Dysentery/complications , Germ-Free Life , Humans , Tumor Cells, Cultured , Virus Cultivation
17.
J Virol ; 64(6): 3112-8, 1990 Jun.
Article in English | MEDLINE | ID: mdl-2159566

ABSTRACT

Antigenic and genomic relationships among tissue culture-adapted turkey enteric coronavirus (TCV) isolates, three strains of avian infectious bronchitis virus (IBV), and mammalian coronaviruses were investigated. Immunoblotting and immunoprecipitation experiments using polyclonal antisera showed that the four major structural proteins of TCV cross-reacted with the four homologous proteins of bovine enteric coronavirus (BCV), the N and M proteins of mouse hepatitis virus serotype 3, and the N protein of IBV. Close antigenic relationships between TCV and BCV were also established by seroneutralization and hemagglutination-inhibition. Of 49 monoclonal antibodies produced against either TCV or BCV, 11 differentiated the two viruses. Five of these monoclonal antibodies had neutralizing activities and were directed to either the peplomeric S (gp200-gp100) or hemagglutinin HE (gp140-gp65) glycoproteins. BCV cDNA probes tested on purified viral preparations and coronavirus-positive (by electron microscopy) fecal samples from diarrheic turkey poults confirmed the relatedness of TCV and BCV. The two viruses produced distinct cytopathic changes in HRT-18 cells in the presence of trypsin, whereas only TCV isolates were able to reproduce the clinical symptoms in turkey poults. Their matrix (M) proteins undergo different glycosylation processes.


Subject(s)
Antigens, Viral/analysis , Coronaviridae/genetics , Coronavirus, Turkey/genetics , Genes, Viral , Animals , Antibodies, Monoclonal , Cattle , Cell Line , Coronaviridae/growth & development , Coronaviridae/immunology , Coronavirus, Turkey/growth & development , Coronavirus, Turkey/immunology , Enteritis, Transmissible, of Turkeys/pathology , Humans , Molecular Weight , Tumor Cells, Cultured/cytology , Tunicamycin/pharmacology , Turkeys , Viral Structural Proteins/isolation & purification
18.
Nihon Juigaku Zasshi ; 52(2): 217-24, 1990 Apr.
Article in English | MEDLINE | ID: mdl-2161476

ABSTRACT

Plaque formation, replication and related cytopathic function of 9 strains of transmissible gastroenteritis (TGE) virus were examined in primary cells and cell lines such as CPK, IB-RS-2, ESK, and PK-15 originated from porcine kidney and the effects of trypsin on the replication of TGE virus were examined in CPK cells. All strains produced a cytopathic effect and grew well in CPK cells as well as in primary porcine kidney cells. The effect of trypsin on the plaque formation was different from strains. The number of plaques produced by strains TO-163, Ukiha and Niigata increased from 2.6 to 3.52 times when trypsin was present in the medium during incubation at 37 degrees C for 1 hr after adsorption of the virus at 4 degrees C for 2 hr. The plaque sizes of TO-163, h-5, Ukiha and Niigata became larger from 1.4 to 1.7 times, when trypsin was present in the agar MEM overlay.


Subject(s)
Coronaviridae/growth & development , Transmissible gastroenteritis virus/growth & development , Trypsin/pharmacology , Viral Plaque Assay/methods , Cell Line , Species Specificity , Transmissible gastroenteritis virus/drug effects , Transmissible gastroenteritis virus/physiology , Virus Replication/drug effects
19.
Zentralbl Veterinarmed B ; 37(2): 91-6, 1990 Mar.
Article in English | MEDLINE | ID: mdl-2363328

ABSTRACT

Various factors influencing the plaque formation of rat coronavirus (RCV) and sialodacryoadenitis virus (SDAV) in LBC cell monolayers were studied to develop the practical method for plaque assay. By this method, 4 Japanese isolates of SDAV also produced clear plaques. In one-step growth experiments of these viruses, newly formed virus was first recognized within 7.5 h postinfection and showed subsequently a rapid exponential increase. The virus was released rapidly from the infected cells. By indirect immunofluorescence virus specific antigen was detected as perinuclear granules in the cytoplasm of the cells at 5-6 h postinfection, and all the cells revealed fluorescence at 12 h postinfection.


Subject(s)
Coronaviridae/growth & development , Animals , Cell Line , Male , Rats , Rats, Inbred Strains , Viral Plaque Assay
20.
Lab Anim Sci ; 40(2): 138-43, 1990 Mar.
Article in English | MEDLINE | ID: mdl-2157091

ABSTRACT

To determine whether SDAV infection persists in athymic rats, weanling athymic rats and euthymic rats were inoculated intranasally with 10(4) TCID50 of SDAV and examined periodically for up to 90 days. Viral antigen and lesions characteristic of acute SDAV infection, including rhinotracheitis, bronchitis and sialodacryoadenitis, were detected in both groups of rats during the first week. In euthymic rats, tissues were under repair and viral antigen was undetectable by day 17, and tissues were histologically normal by day 31 except for mild focal dacryoadenitis. In athymic rats, viral antigen and chronic active inflammation of respiratory tract, salivary and lacrimal glands persisted through day 90. Inflammation and viral antigen also were observed in the transitional epithelium of the renal pelvis and urinary bladder as late as day 90. Virus was isolated from nasopharynx, lung, salivary gland and Harderian gland of athymic rats through day 90. All euthymic rats seroconverted to SDAV by day 6, whereas all athymic rats remained seronegative through day 31, and two of six were seropositive by day 90. As judged by seroconversion of contact sentinels, six of six athymic rats shed virus through 6 weeks, and five of six through 10 weeks. These results indicate that SDAV persists in athymic rats, and that normal T cell function is required for host defenses against SDAV.


Subject(s)
Antigens, Viral/analysis , Coronaviridae Infections/veterinary , Coronaviridae/growth & development , Rats, Mutant Strains/microbiology , Rats, Nude/microbiology , Animals , Coronaviridae/isolation & purification , Coronaviridae Infections/etiology , Coronaviridae Infections/pathology , Female , Male , Necrosis , Rats , Specific Pathogen-Free Organisms
SELECTION OF CITATIONS
SEARCH DETAIL
...