Fig 2 - uploaded by Kenneth L Tyler
Content may be subject to copyright.
Coronal sections of neonatal mouse brain 7 days after intracranial inoculation of 10,000 plaque forming units (PFU) of reovirus strain T3D or mock inoculation. Hematoxylin and eosin-stained tissue reveals marked destruction of brain tissue in the T3D-infected brain ( B ) as compared to the uninfected brain ( A ). By immunohis- tochemistry, serial sections of T3D-infected brain tissue were stained for viral antigen ( C ), TUNEL/apoptosis marker ( D ), and active caspase 3 ( E ). Staining for viral antigen, TUNEL, and caspase 3 were undetectable in the mock-infected brains (data not shown). (With permission from Richardson-Burns et al. 2002) 

Coronal sections of neonatal mouse brain 7 days after intracranial inoculation of 10,000 plaque forming units (PFU) of reovirus strain T3D or mock inoculation. Hematoxylin and eosin-stained tissue reveals marked destruction of brain tissue in the T3D-infected brain ( B ) as compared to the uninfected brain ( A ). By immunohis- tochemistry, serial sections of T3D-infected brain tissue were stained for viral antigen ( C ), TUNEL/apoptosis marker ( D ), and active caspase 3 ( E ). Staining for viral antigen, TUNEL, and caspase 3 were undetectable in the mock-infected brains (data not shown). (With permission from Richardson-Burns et al. 2002) 

Source publication
Article
Full-text available
Reovirus infection has proven to be an excellent experimental system for studying mechanisms of virus-induced pathogenesis. Reoviruses induce apoptosis in a wide variety of cultured cells in vitro and in target tissues in vivo, including the heart and central nervous system. In vivo, viral infection, tissue injury, and apoptosis colocalize, suggest...

Contexts in source publication

Context 1
... (Diaz-Meco et al. 1999). Reovirus infection results in a viral strain-specific pattern activation of the c-Jun N-terminal kinase (JNK) and the JNK-associated transcription factor c-Jun (Clarke et al. 2001b). The capacity of reovirus strains to activate JNK correlates closely with their capacity to induce apoptosis (Clarke et al. 2001b). In addition, experiments using T1L T3D reassortants indicate that the same viral gene segments that determine apoptosis induction (S1 and M2) are also key determinants of JNK activation (Clarke et al. 2001b). Furthermore, our preliminary experiments indicate that reovirus-induced apoptosis is inhibited in cells deficient in MEK kinase 1, an upstream activator of JNK in reovirus-infected cells and in cells treated with inhibitors of JNK activity. These results indicate that JNK is required for reovirus-induced apoptosis. Our recent experiments also indicate that JNK is required for the efficient release of smac and cytochrome c from the mitochondria of reovirus-infected cells, suggesting that JNK promotes mitochondrial pathways of apoptosis in reovirus-infected cells (Clarke et al. 2004). Both JNK-induced phosphorylation of Bcl-2 family proteins and c-Jun-in- duced expression of the BH3-only protein Bim have previously been shown to promote mitochondrial apoptotic signaling. Experiments to determine the mechanism by which JNK and c-Jun influence reovirus- induced apoptosis are currently under way. Reovirus infection induces the activation of transcription factors NF- k B and c-Jun (Clarke et al. 2001b, 2003a; Connolly et al. 2000). This suggests that activation of specific cellular genes contributes to virus-induced cellular signaling, including apoptotic signaling, in infected cells. High- density oligonucleotide microarrays used to perform a global analysis of virus-induced cellular gene expression after reovirus infection of HEK293 cells (DeBiasi et al. 2003; Poggioli et al. 2002) showed that the expression of 24 genes related to apoptosis were altered in cells infected with the apoptosis-inducing reovirus strain T3A (Table 1). These genes encode proteins with potential roles in DR, endoplasmic reticulum stress, and mitochondrial apoptotic signaling and cysteine proteases (caspases and calpains) (DeBiasi et al. 2003). Only five of these genes were also differentially expressed after T1L (weakly apoptotic) infection, emphasizing their potential importance in reovirus-induced apoptosis. To date the best-characterized example of a potentially apoptosis-inducing gene identified by microarray analysis to be differentially expressed after reovirus infection is the survival motor neuron (SMN) gene. This gene was found by polymerase chain reaction (PCR) analysis to be upregulated at the transcriptional level in reovirus-infected HEK293 cells and at the translational level in the hearts of reovirus-infected baby mice (DeBiasi et al. 2003). The SMN protein has been shown to interact with Bcl-2, conferring a synergistic protective effect against Bax-induced or Fas-mediated apoptosis that has been shown to underlie the pathogenesis of spinal muscular atrophy (Sato et al. 2000; Iwahashi et al. 1997). Changes in the expression of genes encoding proteins known to be involved in DNA repair and cell cycle regulation were also identified in this study and may also affect virus-induced pathogenesis (DeBiasi et al. 2003). In addition to inducing TRAIL-mediated apoptosis, reovirus infection also sensitizes cells to TRAIL-induced apoptosis by a mechanism that results in an increase in the activation of caspases 8 and 3 and is blocked by the caspase 8 inhibitor IETD-FMK (Clarke et al. 2000, 2001a). Reovirus infection and TRAIL treatment have synergistic rather than merely additive effects on apoptosis, and infection can confer TRAIL sensitivity to previously TRAIL-resistant cells as well as increasing the TRAIL sensitivity of partially resistant lines (Clarke et al. 2000, 2001a). This finding may increase the potential utility of TRAIL as an agent for cancer thera- py, which is currently limited by the fact that cancer cells of all types differ in sensitivity to TRAIL-induced apoptosis. The ability of reovirus to sensitize cells to TRAIL does not appear to reflect an increase in the expression of TRAIL receptors as assayed in several human cancer cell lines (Clarke et al. 2001a) and may instead be the result of inhibition of TRAIL-induced activation of NF- k B in reovirus-infected cells (Clarke et al. 2003a). The ability of reovirus to sensitize cells to TRAIL also suggests that reovirus-infected cells in vivo are also susceptible to killing through the TRAIL pathway by immune cells such as natural killer and CD4 + cells that bear membrane-bound TRAIL. T3 reovirus strains infect neurons within specific regions of neonatal mouse brains, producing a lethal meningoencephalitis. Viral antigen and pathology colocalize in the brain and have a predilection for the cortex, hippocampus and thalamus (Fig. 2). T3 reovirus infection also induces apoptosis in the brains of newborn mice (Oberhaus et al. 1997; Richardson-Burns et al. 2002). Thus fragmentation of DNA into oligonu- cleosomal-length ladders can be detected in tissue samples prepared from T3D- but not mock-infected brains at 8–9 days PI, which coincides with maximal viral growth (Oberhaus et al. 1997). The presence of apoptotic cells also correlates with areas of tissue injury and viral infection in T3-infected brain sections (Fig. 2) (Oberhaus et al. 1997; Richardson- Burns et al. 2002). Most cells in infected brain regions are both TUNEL (TdT-mediated dUTP nick-end labeling)-positive (apoptotic) and reovirus antigen-positive (infected). However, there are cells in these regions that are apoptotic but antigen negative, suggesting that apoptosis occurs both as a result of direct viral infection and in uninfected “bystander” cells (Oberhaus et al. 1997). Reovirus infection in a mouse neuroblasto- ma-derived cell line (NB41a3) and in primary mouse cortical cultures (MCC) derived from embryonic (E20) mice is also associated with increased levels of caspase 3 activity and is blocked with the caspase 3 inhibitor DEVD-FMK (Richardson-Burns et al. 2002). Studies of reovirus infection in neuronal cultures also provides further evidence of bystander apoptosis. In both MCC and NB4 cells dual labeling with immunocy- tochemistry and TUNEL showed that although a great majority of infected cells were undergoing apoptosis there was also a subset of apoptotic cells that were uninfected but located in proximity to virus-infected cells (Richardson-Burns et al. 2002). Bystander apoptosis could result from the release of TRAIL, or other death ligands, from reovirus-infected cells. If this is the case, the amount of bystander apoptosis would reflect the sensitivity of the surrounding cells to the released ligand. Reovirus infection also induces increased caspase 8 activation in infected neurons, indicating that neuronal apoptosis, like that in its epi- thelial cell counterparts, involves DR activation (Richardson-Burns et al. 2002). However, the ligand-receptor trigger for this activation appears to be less specific. Thus, whereas reovirus-induced apoptosis in HEK293 cells is selectively inhibited by blocking TRAIL ligand-receptor interac- tion, reovirus-induced apoptosis in NB4 cells is inhibited by treating cells with both soluble TRAIL receptors (Fc:DR5) and soluble TNF receptors (TNFR) (FcTNFR-1), and reovirus-induced apoptosis in MCCs is inhibited by Fc:TNFR-1 and Fc:FasL. Mitochondrial apoptotic pathways are also activated after reovirus infection of neurons. Preliminary studies indicate that proapoptotic Bcl-2 family proteins, including Bid, Bax, and Bim, are activated in virus-infected neurons, resulting in the release of proapoptotic mitochondrial factors. However, there are again differences between mitochondrial signaling pathways activated after reovirus infection of neuronal and epi- thelial cells. In HEK293 cells, reovirus infection is associated with robust release of cytochrome c and smac and the subsequent activation of caspase 9 and inhibition of cellular IAPs. In neuronal cultures, however, our preliminary results indicate that reovirus infection results in the discor- dant release of smac and cytochrome c . Smac is released around 17 h PI and coincides with the cleavage of cellular IAPs. In contrast, cytochrome c release occurs only at low levels and at later times after infection, resulting in only low levels of activation of caspase 9 in these cells (Richardson-Burns et al. 2002). Consistent with these findings, the caspase 9 inhibitor Z-LEHD-FMK has little effect on reovirus-induced neuronal apoptosis, which is significantly inhibited by caspase 8 (Z-IETD-FMK), caspase 3 (Z-DEVD-FMK), or pan-caspase inhibitors. The T1L T3D reassortant virus 8B efficiently produces myocarditis in infected neonatal mice. Similar to results seen in mouse brain, DNA ex- tracted from the hearts of 8B-infected mice is fragmented into oligonu- cleosomal-length ladders, indicative of apoptosis (DeBiasi et al. 2001) and areas of TUNEL-positive cells in 8B-infected hearts correlates with areas of histological damage and reovirus antigen (DeBiasi et al. 2001). Injury to the heart following reovirus infection occurrs in the absence of an inflammatory response, also suggesting that it results from apoptotic cell death (DeBiasi et al. 2001). Treatment of mice with the calpain inhibitor CX295 [dipeptide a -ke- toamide calpain inhibitor z-Leu-aminobutyric acid-CONH(CH 2 )-3-mor- pholine] is protective against reovirus-induced myocarditis and results in a dramatic reduction in histopathologic evidence of myocardial injury (Fig. 3), a reduction in serum creatine phosphokinase (an intracellular enzyme whose release into the serum is a quantitative marker of skeletal and cardiac muscle damage), and improved weight gain (DeBiasi et al. 2001). Prevention of myocardial injury ...
Context 2
... determinants of JNK activation (Clarke et al. 2001b). Furthermore, our preliminary experiments indicate that reovirus-induced apoptosis is inhibited in cells deficient in MEK kinase 1, an upstream activator of JNK in reovirus-infected cells and in cells treated with inhibitors of JNK activity. These results indicate that JNK is required for reovirus-induced apoptosis. Our recent experiments also indicate that JNK is required for the efficient release of smac and cytochrome c from the mitochondria of reovirus-infected cells, suggesting that JNK promotes mitochondrial pathways of apoptosis in reovirus-infected cells (Clarke et al. 2004). Both JNK-induced phosphorylation of Bcl-2 family proteins and c-Jun-in- duced expression of the BH3-only protein Bim have previously been shown to promote mitochondrial apoptotic signaling. Experiments to determine the mechanism by which JNK and c-Jun influence reovirus- induced apoptosis are currently under way. Reovirus infection induces the activation of transcription factors NF- k B and c-Jun (Clarke et al. 2001b, 2003a; Connolly et al. 2000). This suggests that activation of specific cellular genes contributes to virus-induced cellular signaling, including apoptotic signaling, in infected cells. High- density oligonucleotide microarrays used to perform a global analysis of virus-induced cellular gene expression after reovirus infection of HEK293 cells (DeBiasi et al. 2003; Poggioli et al. 2002) showed that the expression of 24 genes related to apoptosis were altered in cells infected with the apoptosis-inducing reovirus strain T3A (Table 1). These genes encode proteins with potential roles in DR, endoplasmic reticulum stress, and mitochondrial apoptotic signaling and cysteine proteases (caspases and calpains) (DeBiasi et al. 2003). Only five of these genes were also differentially expressed after T1L (weakly apoptotic) infection, emphasizing their potential importance in reovirus-induced apoptosis. To date the best-characterized example of a potentially apoptosis-inducing gene identified by microarray analysis to be differentially expressed after reovirus infection is the survival motor neuron (SMN) gene. This gene was found by polymerase chain reaction (PCR) analysis to be upregulated at the transcriptional level in reovirus-infected HEK293 cells and at the translational level in the hearts of reovirus-infected baby mice (DeBiasi et al. 2003). The SMN protein has been shown to interact with Bcl-2, conferring a synergistic protective effect against Bax-induced or Fas-mediated apoptosis that has been shown to underlie the pathogenesis of spinal muscular atrophy (Sato et al. 2000; Iwahashi et al. 1997). Changes in the expression of genes encoding proteins known to be involved in DNA repair and cell cycle regulation were also identified in this study and may also affect virus-induced pathogenesis (DeBiasi et al. 2003). In addition to inducing TRAIL-mediated apoptosis, reovirus infection also sensitizes cells to TRAIL-induced apoptosis by a mechanism that results in an increase in the activation of caspases 8 and 3 and is blocked by the caspase 8 inhibitor IETD-FMK (Clarke et al. 2000, 2001a). Reovirus infection and TRAIL treatment have synergistic rather than merely additive effects on apoptosis, and infection can confer TRAIL sensitivity to previously TRAIL-resistant cells as well as increasing the TRAIL sensitivity of partially resistant lines (Clarke et al. 2000, 2001a). This finding may increase the potential utility of TRAIL as an agent for cancer thera- py, which is currently limited by the fact that cancer cells of all types differ in sensitivity to TRAIL-induced apoptosis. The ability of reovirus to sensitize cells to TRAIL does not appear to reflect an increase in the expression of TRAIL receptors as assayed in several human cancer cell lines (Clarke et al. 2001a) and may instead be the result of inhibition of TRAIL-induced activation of NF- k B in reovirus-infected cells (Clarke et al. 2003a). The ability of reovirus to sensitize cells to TRAIL also suggests that reovirus-infected cells in vivo are also susceptible to killing through the TRAIL pathway by immune cells such as natural killer and CD4 + cells that bear membrane-bound TRAIL. T3 reovirus strains infect neurons within specific regions of neonatal mouse brains, producing a lethal meningoencephalitis. Viral antigen and pathology colocalize in the brain and have a predilection for the cortex, hippocampus and thalamus (Fig. 2). T3 reovirus infection also induces apoptosis in the brains of newborn mice (Oberhaus et al. 1997; Richardson-Burns et al. 2002). Thus fragmentation of DNA into oligonu- cleosomal-length ladders can be detected in tissue samples prepared from T3D- but not mock-infected brains at 8–9 days PI, which coincides with maximal viral growth (Oberhaus et al. 1997). The presence of apoptotic cells also correlates with areas of tissue injury and viral infection in T3-infected brain sections (Fig. 2) (Oberhaus et al. 1997; Richardson- Burns et al. 2002). Most cells in infected brain regions are both TUNEL (TdT-mediated dUTP nick-end labeling)-positive (apoptotic) and reovirus antigen-positive (infected). However, there are cells in these regions that are apoptotic but antigen negative, suggesting that apoptosis occurs both as a result of direct viral infection and in uninfected “bystander” cells (Oberhaus et al. 1997). Reovirus infection in a mouse neuroblasto- ma-derived cell line (NB41a3) and in primary mouse cortical cultures (MCC) derived from embryonic (E20) mice is also associated with increased levels of caspase 3 activity and is blocked with the caspase 3 inhibitor DEVD-FMK (Richardson-Burns et al. 2002). Studies of reovirus infection in neuronal cultures also provides further evidence of bystander apoptosis. In both MCC and NB4 cells dual labeling with immunocy- tochemistry and TUNEL showed that although a great majority of infected cells were undergoing apoptosis there was also a subset of apoptotic cells that were uninfected but located in proximity to virus-infected cells (Richardson-Burns et al. 2002). Bystander apoptosis could result from the release of TRAIL, or other death ligands, from reovirus-infected cells. If this is the case, the amount of bystander apoptosis would reflect the sensitivity of the surrounding cells to the released ligand. Reovirus infection also induces increased caspase 8 activation in infected neurons, indicating that neuronal apoptosis, like that in its epi- thelial cell counterparts, involves DR activation (Richardson-Burns et al. 2002). However, the ligand-receptor trigger for this activation appears to be less specific. Thus, whereas reovirus-induced apoptosis in HEK293 cells is selectively inhibited by blocking TRAIL ligand-receptor interac- tion, reovirus-induced apoptosis in NB4 cells is inhibited by treating cells with both soluble TRAIL receptors (Fc:DR5) and soluble TNF receptors (TNFR) (FcTNFR-1), and reovirus-induced apoptosis in MCCs is inhibited by Fc:TNFR-1 and Fc:FasL. Mitochondrial apoptotic pathways are also activated after reovirus infection of neurons. Preliminary studies indicate that proapoptotic Bcl-2 family proteins, including Bid, Bax, and Bim, are activated in virus-infected neurons, resulting in the release of proapoptotic mitochondrial factors. However, there are again differences between mitochondrial signaling pathways activated after reovirus infection of neuronal and epi- thelial cells. In HEK293 cells, reovirus infection is associated with robust release of cytochrome c and smac and the subsequent activation of caspase 9 and inhibition of cellular IAPs. In neuronal cultures, however, our preliminary results indicate that reovirus infection results in the discor- dant release of smac and cytochrome c . Smac is released around 17 h PI and coincides with the cleavage of cellular IAPs. In contrast, cytochrome c release occurs only at low levels and at later times after infection, resulting in only low levels of activation of caspase 9 in these cells (Richardson-Burns et al. 2002). Consistent with these findings, the caspase 9 inhibitor Z-LEHD-FMK has little effect on reovirus-induced neuronal apoptosis, which is significantly inhibited by caspase 8 (Z-IETD-FMK), caspase 3 (Z-DEVD-FMK), or pan-caspase inhibitors. The T1L T3D reassortant virus 8B efficiently produces myocarditis in infected neonatal mice. Similar to results seen in mouse brain, DNA ex- tracted from the hearts of 8B-infected mice is fragmented into oligonu- cleosomal-length ladders, indicative of apoptosis (DeBiasi et al. 2001) and areas of TUNEL-positive cells in 8B-infected hearts correlates with areas of histological damage and reovirus antigen (DeBiasi et al. 2001). Injury to the heart following reovirus infection occurrs in the absence of an inflammatory response, also suggesting that it results from apoptotic cell death (DeBiasi et al. 2001). Treatment of mice with the calpain inhibitor CX295 [dipeptide a -ke- toamide calpain inhibitor z-Leu-aminobutyric acid-CONH(CH 2 )-3-mor- pholine] is protective against reovirus-induced myocarditis and results in a dramatic reduction in histopathologic evidence of myocardial injury (Fig. 3), a reduction in serum creatine phosphokinase (an intracellular enzyme whose release into the serum is a quantitative marker of skeletal and cardiac muscle damage), and improved weight gain (DeBiasi et al. 2001). Prevention of myocardial injury by apoptosis inhibitors is accompa- nied by a virtually complete inhibition of apoptotic myocardial cell death, strongly suggesting that virus-induced apoptosis is a key mechanism of cell death, tissue injury, and mortality in reovirus-infected mice and that inhibitors of apoptosis may prove useful in the treatment of virus-induced diseases (DeBiasi et al. 2001). Early studies showed that there is little correlation in continuous non- neuronal cell lines between the efficiency with which ...

Citations

... Mature particles are released from cells lytically or non-lytically, but it is currently unclear what mechanism controls these egress phenotypes [38]. Two strains of reovirus, type 1 Lang (T1L) and type 3 Dearing (T3D), differ in their pathogenesis, tropism, and capacity to induce apoptosis, with T3D inducing apoptosis more efficiently than T1L [39][40][41][42][43]. There appears to be little connection between apoptosis induction and progeny virus yield [40,44,45]. ...
Article
Full-text available
Several egress pathways have been defined for many viruses. Among these pathways, extracellular vesicles (EVs) have been shown to function as vehicles of non-lytic viral egress. EVs are heterogenous populations of membrane-bound structures released from cells as a form of intercellular communication. EV-mediated viral egress may enable immune evasion and collective viral transport. Strains of nonenveloped mammalian orthoreovirus (reovirus) differ in cell lysis phenotypes, with T3D disrupting cell membranes more efficiently than T1L. However, mechanisms of reovirus egress and the influence of transport strategy on infection are only partially understood. To elucidate reovirus egress mechanisms, we infected murine fibroblasts (L cells) and non-polarized human colon epithelial (Caco-2) cells with T1L or T3D reovirus and enriched cell culture supernatants for large EVs, medium EVs, small EVs, and free reovirus. We found that both reovirus strains exit cells in association with large and medium EVs and as free virus particles, and that EV-enriched fractions are infectious. While reovirus visually associates with large and medium EVs, only medium EVs offer protection from antibody-mediated neutralization. EV-mediated protection from neutralization is virus strain- and cell type-specific, as medium EVs enriched from L cell supernatants protect T1L and T3D, while medium EVs enriched from Caco-2 cell supernatants largely fail to protect T3D and only protect T1L efficiently. Using genetically barcoded reovirus, we provide evidence that large and medium EVs can convey multiple particles to recipient cells. Finally, T1L or T3D infection increases the release of all EV sizes from L cells. Together, these findings suggest that in addition to exiting cells as free particles, reovirus promotes egress from distinct cell types in association with large and medium EVs during lytic or non-lytic infection, a mode of exit that can mediate multiparticle infection and, in some cases, protection from antibody neutralization.
... inducing signalling complex (DISC)-mediated apoptosis. RT3D has previously been reported to induce apoptosis [19][20][21] . To probe the mechanisms of cell death, a caspase inhibitor screen was used to identify key caspases involved in RT3D plus talazoparib-mediated cell death. ...
... associated with activation of Caspase-8-mediated extrinsic apoptosis [20,21] . While we observed an increase in RT3D-induced TRAIL and TNF-α plus their receptors, DR4/DR5, or TNFR1/TNFR2 respectively, this was not increased further by combination with talazoparib. ...
Preprint
Full-text available
Oncolytic Reovirus type 3 Dearing (RT3D), is a naturally occurring double-stranded (ds) RNA virus that is under development as an oncolytic immunotherapy We used an unbiased high-throughput cytotoxicity screen of different targeted therapeutic agents with the aim of identifying potential drug-viral sensitizers to enhance RT3D tumour killing. Talazoparib, a clinical poly(ADP)-ribose polymerase 1 (PARP-1) inhibitor, was identified as a top hit and found to cause profound sensitisation to RT3D. This effect was not seen with other classes of oncolytic virus and was not mediated by enhanced viral replication or PARP inhibitor-related effects on the DNA damage response. RT3D interacts with retinoic acid-induced gene-1 (RIG-I) and activates PARP-1, with consequent PARylation of components of the extrinsic apoptosis pathway. Pharmacological and genetic inhibition of PARP-1 abrogates this PARylation and increases levels of extrinsic apoptosis, NF-kB signalling and pro-inflammatory cell death. Direct interaction between PARP-1 and RIG-I following RT3D/talazoparib treatment is a key factor in activating downstream signaling pathways that lead to IFN-β and TNF-α/TRAIL production which, in turn, amplify the therapeutic effect through positive feedback. Critically, it was possible to phenocopy the effect of RT3D through the use of non-viral ds-RNA therapy and RIG-I agonism. In in vivo studies, we demonstrated profound combinatorial efficacy of RT3D and talazoparib in human A375 melanoma in immunodeficient mice. More impressively, in immunocompetent mouse models of 4434 murine melanoma, we achieved 100% tumour control and protection from subsequent tumour rechallenge with the combination regimen. Correlative immunophenotyping confirmed significant innate and adaptive immune activation with the combination of RT3D and PARP inhibition. Taken together, these data provide a clear line of sight to clinical translation of combined regimens of PARP inhibition or ds-RNA agonism, with either viral or non-viral agents, in tumour types beyond the relatively narrow confines of current licensed indications for PARP inhibition.
... Consistent with a reduced glycolytic flux and reduced mitochondrial function, reovirus + FK866 caused a reduction in overall redox capacity compared with Figure 4E). Similar to findings from both monotherapies in other cancer models, 49,50 our data suggest that the reovirus + FK866 combination results in an early synergistic negative effect on central energy metabolism with a drop in mitochondrial function in KMS12 cells ( Figure 4F). ...
Article
Full-text available
Cancer cell energy metabolism plays an important role in dictating the efficacy of oncolysis by oncolytic viruses. To understand the role of multiple myeloma metabolism in reovirus oncolysis, we performed semi-targeted mass spectrometry-based metabolomics on twelve multiple myeloma cell lines and revealed a negative correlation between NAD+ levels and susceptibility to oncolysis. Likewise, a negative correlation was observed between the activity of the rate-limiting NAD+ synthesis enzyme, NAMPT, and oncolysis. Indeed, depletion of NAD+ levels by pharmacological inhibition of NAMPT using FK866 sensitized several myeloma cell lines to reovirus-induced killing. The myelomas that were most sensitive to this combination therapy expressed a functional p53 and had a metabolic and transcriptomic profile favoring mitochondrial metabolism over glycolysis, with the highest synergistic effect in KMS12 cells. Mechanistically, U-¹³C-labelled glucose flux, extracellular flux analysis, multiplex proteomics, and cell death assays revealed that reovirus+FK866 combination caused mitochondrial dysfunction and energy depletion, leading to enhanced autophagic cell death in KMS12 cells. Finally, the combination of reovirus and NAD+ depletion achieved greater antitumor effects in KMS12 tumors in vivo and patient-derived CD138+ multiple myeloma cells. These findings identify NAD+ depletion as a potential combinatorial strategy to enhance efficacy of oncolytic virus-based therapies in multiple myeloma.
... Consistent with previous studies showing that reovirus infection results in cell cycle arrest at G2/M in several cancer cell lines, [22][23][24][25][26] we observed that treatment with reovirus alone induced cell cycle arrest at G2/M in CMeC1. Moreover, treatment of CMeC1 with the combination of KU60019 and reovirus resulted in strong cell cycle arrest at G2/M and a reduction of the G1 interval ( Figure 4A). ...
Article
Full-text available
Oncolytic virotherapy using reovirus is a promising new anti-cancer treatment with potential for use in humans and dogs. Because reovirus monotherapy shows limited efficacy in human and canine cancer patients, the clinical development of a combination therapy is necessary. To identify candidate components of such a combination, we screened a 285-compound drug library for those that enhanced reovirus cytotoxicity in a canine melanoma cell line. Here, we show that exposure to an inhibitor of the ataxia telangiectasia mutated protein (ATM) enhances the oncolytic potential of reovirus in five of six tested canine melanoma cell lines. Specifically, the ATM inhibitor potentiated reovirus replication in cancer cells along with promoting the lysosomal activity, resulting in an increased proportion of caspase-dependent apoptosis and cell cycle arrest at G2/M compared to those observed with reovirus alone. Overall, our study suggests that the combination of reovirus and the ATM inhibitor may be an attractive option in cancer therapy.
... The most studied reovirus-induced cell death mechanism is the apoptotic pathway [19][20][21][22][23][24]. As is mentioned before, induction of apoptosis is enhanced in reovirus-infected Rastransformed cells and may stimulate virus release and spread of the virus to neighbouring cells. ...
Article
Full-text available
While the mammalian orthoreovirus type 3 dearing (reovirus T3D) infects many different tumour cells, various cell lines resist the induction of reovirus-mediated cell death. In an effort to increase the oncolytic potency, we introduced transgenes into the S1 segment of reovirus T3D. The adenovirus E4orf4 gene was selected as transgene since the encoded E4orf4 protein induces cell death in transformed cells. The induction of cell death by E4orf4 depends in part on its binding to phosphatase 2A (PP2A). In addition to the S1-E4orf4 reovirus, two other reoviruses were employed in our studies. The reovirus rS1-RFA encodes an E4orf4 double-mutant protein that cannot interact with PP2A and the rS1-iLOV virus encoding the fluorescent marker iLOV as a reporter. The replacement of the codons for the junction adhesion molecule-A (JAM-A) binding head domain of the truncated spike protein blocks the entry of these recombinant viruses via the reovirus receptor JAM-A. Instead these viruses rely on internalization via binding to sialic acids on the cell surface. This expands their tropism and allows infection of JAM-A-deficient tumour cells. Here we not only demonstrate the feasibility of this approach but also established that the cytolytic activity of these recombinant viruses is largely transgene independent.
... Bid -/mice showed no lung injury after Lipopolysaccharides (LPS) stimulation [17]. Survival of Bid-deficient mice was significantly increased when compared with wild type mice after reovirus infection [7,18]. However, there is limited knowledge of Bid in teleost fish, and the specific role of Bid during the virusinduced apoptosis in teleost fish is still unclear. ...
Article
Full-text available
Bid, BH3-interacting domain death agonist, is a pro-apoptotic BH3-only member of Bcl-2 family, playing an important role in apoptosis. In the study, Bid genes from grass carp (Ctenopharyngodon idellus) and rare minnow (Gobiocypris rarus), named CiBid and GrBid, were cloned and analyzed. Bid was constitutively expressed in all examined tissues of grass carp, but the expression level varied in different tissues. Following grass carp reovirus (GCRV) stimulation in vivo, Bid and apoptosis related genes Caspase-9 and Caspase-3 was up-regulated significantly at the late stage of infection. Moreover, we generated a Bid-deficient rare minnow (Bid-/-) to investigate the possible role of Bid in GCRV-triggered apoptosis. We found that the survival time of Bid-/- rare minnow after GCRV infection was extended when compared with wild-type fish, the relative copy number of GCRV in Bid-/- rare minnow was lower than that in wild-type fish, and the expression level of Caspase-9 and Caspase-3 in Bid-/- rare minnow were significantly lower than that in the wild-type fish. Collectively, the current data revealed the important role of Bid during virus-induced apoptosis in teleost fish. Our study would provide new insight into understanding the GCRV induced apoptosis and may provide a target gene for virus-resistant breeding in grass carp.
... Finally, reovirus-induced myocarditis is the result of a direct viral cytopathic effect to cardiac myocytes (Baty and Sherry, 1993), which is largely apoptotic (Miyamoto et al., 2009). Reovirus-induced apoptosis requires activation of NF-κB in most cell types (O'Donnell et al., 2005; Connolly et al., 2000; Holm et al., 2007; Stebbing et al., 2014; Clarke et al., 2005a Clarke et al., , 2005b Clarke et al., , 2003) but not in cardiac myocytes (O'Donnell et al., 2005; Stebbing et al., 2014). Thus reovirus-induced myocarditis does not require activation of NF-κB in cardiac myocytes, consistent with evidence here that cardiac myocytes are relatively recalcitrant to NF-κB activation (Figs. 1 and 6–8). ...
... CsCl-purified reovirus type 3 Dearing (T3D) was maintained as a low-passage laboratory stock and stored at –80 °C. T3D was chosen as a strong inducer of NF-κB activity (O'Donnell et al., 2005; Connolly et al., 2000; Holm et al., 2007; Stebbing et al., 2014; Clarke et al., 2005a Clarke et al., , 2005b Clarke et al., , 2003). Unless stated otherwise, cardiac cultures plated in 8-well chamber slides were inoculated with reovirus T3D at 50 plaque forming units (PFU) per cell and were fixed at the indicated times post-infection. ...
... However, there is limited knowledge of expression pattern of Bax in response to dsRNA viruses, and the specific role of Bax during the virus-induced apoptosis in teleost fish is still unclear. In addition, given that apoptosis can affect pathogenesis and accelerate the replication of virus, identification of a Bax-dependent apoptotic pathway in grass carp after GCRV infection might be a viable way to develop a virus resistance strategy for breeding of grass carp (Clarke et al. 2005). ...
Article
Full-text available
Multidomain proapoptotic Bcl-2-associated X (Bax) protein is an essential effector responsible for mitochondrial outer membrane permeabilization, resulting in cell death via apoptosis. In this study, two Bax genes of grass carp (Ctenopharyngodon idellus), designated as CiBax1 and CiBax2, were isolated and analyzed. The obtained CiBax1 cDNA is 2058 bp long, with a 579 bp open reading frame (ORF) coding a protein of 192 amino acid residues. The full-length cDNA of CiBax2 is 1161 bp, with a 618 bp ORF coding 205 amino acids. Both CiBax1 and CiBax2 are typical members of Bcl-2 family containing conserved Bcl and C-terminal domains, and they share conserved synteny with zebrafish Bax genes despite the grass carp Bax mapping to different linkage groups. Phylogenetic analysis showed that CiBax1 was clustered with Bax from most teleost fish, and CiBax2 was close to Bax2 from teleost fish but far separated from that of Salmo salar. Quantitative real-time PCR analysis revealed broad expression of CiBax1 and CiBax2 in tissues from healthy grass carp, but the relative expression level differed. The mRNA expression of CiBax1 and CiBax2 was both upregulated significantly and peaked in all examined tissues at days 5 or 6 post-infection with grass carp reovirus. Subcellular localization indicated that CiBax1 protein was localized in both nucleus and cytosol, while CiBax2 protein only in cytosol. Moreover, CiBax2, but not CiBax1 was colocalized with mitochondrion under normal condition. Taken together, the findings would be helpful for further understanding of the function of Bax in teleost fish.
... Respiratory Enteric Orphan Virus or Reovirus is a non-pathogenic virus initially isolated from human respiratory and gastrointestinal tracts in the 1950s [5,6]. Reovirus is not associated with any known disease and has been shown to induce apoptosis in cells carrying a mutation in a certain pathway, the Ras pathway, which is found in a wide variety of tumour types [7,8]. In vitro, over 80 % of cell lines of tumour origin tested have been found to be susceptible to Reovirus-related oncolysis [9]. ...
Article
Full-text available
Background: The ever increasing knowledge in the areas of cell biology, the immune system and the mechanisms of cancer are allowing a new phase of immunotherapy to develop. The aim of cancer vaccination is to activate the host immune system and some success has been observed particularly in the use of the BCG vaccine for bladder cancer as an immunostimulant. Reovirus, an orphan virus, has proven itself as an oncolytic virus in vitro and in vivo. Over 80 % of tumour cell lines have been found to be susceptible to Reovirus infection and it is currently in phase III clinical trials. It has been shown to induce immune responses to tumours with very low toxicities. Methods: In this study, Reovirus was examined in two main approaches in vivo, in mice, using the melanoma B16F10 and Lewis Lung Carcinoma (LLC) models. Initially, mice were treated intratumourally (IT) with Reovirus and the immune responses determined by cytokine analysis. Mice were also vaccinated using a cell-based Reovirus vaccine and subsequently exposed to a tumourigenic dose of cells (B16F10 or LLC). Using the same cell-based Reovirus vaccine, established tumours were treated and subsequent immune responses and virus retrieval investigated. Results: Upregulation of several cytokines was observed following treatment and replication-competent virus was also retrieved from treated tumours. Varying levels of cytokine upregulation were observed and no replication-competent virus was retrieved in vaccine-treated mice. Prolongation of survival and delayed tumour growth were observed in all models and an immune response to Reovirus, either using Reovirus alone or a cell-based vaccine was also observed in all mice. Conclusion: This study provides evidence of immune response to tumours using a cell-based Reovirus vaccine in both tumour models investigated, B16F10 and LLC, cytokine induction was observed with prolongation of survival in almost all cases which may suggest a new method for using Reovirus in the clinic.
... Like VV, MCMV, the murine counterpart of the human cytomegalovirus (HCMV), member of the herpes virus family, also encodes a viral inhibitor of caspase-8-dependent apoptosis [105,164] and can induce necroptosis as an antiviral mechanism [165]. Reovirus, a small nonenveloped double-stranded RNA virus, induces apoptotic regulated cell-death in several different cell types and involves death receptor signaling, especially the TNF-related apoptosis-inducing ligand (TRAIL) [166,167]. Recent data indicate that Reovirus can also induce RIP-1 associated necroptosis in the murine fibroblasts cell line L-929 [168]. Using its non-structural protein 1 (nsp1), Influenza A virus (IAV), a single-stranded multi-segment RNA virus causing the flu is able to inhibit apoptosis early during infection and to induce it at later stages [169]. ...
Chapter
Full-text available
Both autophagy and programmed cell death (PCD) are fundamental processes of cellular maintenance that are closely interrelated in plant and animal cells under physiological and stressful conditions. Differentiation, immune response, lack of nutrients, and wide range of abiotic factors induce their development and realization of survival or cell death scenario. Microtubular cytoskeleton is known as one of the principle players in the mediation of PCD/autophagic signals. Microtubules (MTs) were shown involved in numeral PCD-like processes as well as in autophagosome maturation and traffic. As very dynamic structures MTs are highly sensitive to external stimuli and their functional state is basic for their role in realization of pro-death or pro-recycling cellular program. Generally, it provided with cytoskeletal network reorganization via balancing between stable/dynamic MTs behavior that is closely depends on variety of post-translational modifications of tubulin and its situational protein microenvironments. This chapter represents an attempt to overview the current knowledge about the role of MTs in PCD- and autophagy-related processes in plants.