Next Article in Journal
Immune Profile Analysis in Peripheral Blood and Tumor in Patients with Malignant Melanoma
Next Article in Special Issue
Bcl-xL: A Focus on Melanoma Pathobiology
Previous Article in Journal
Instability of the NS1 Glycoprotein from La Reunion 2018 Dengue 2 Virus (Cosmopolitan-1 Genotype) in Huh7 Cells Is Due to Lysine Residues on Positions 272 and 324
Previous Article in Special Issue
Regulation of Platelet Production and Life Span: Role of Bcl-xL and Potential Implications for Human Platelet Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Bcl-xL Protein in Viral Infections

1
Institute of Biological Sciences, Cardinal Stefan Wyszyński University in Warsaw, 01-815 Warsaw, Poland
2
Department of Biochemistry and Microbiology, Institute of Biology, Warsaw University of Life Sciences-SGGW, 02-787 Warsaw, Poland
3
Institute of Veterinary Medicine, Warsaw University of Life Sciences, 02-787 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(4), 1956; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22041956
Submission received: 28 January 2021 / Revised: 13 February 2021 / Accepted: 14 February 2021 / Published: 16 February 2021
(This article belongs to the Special Issue Advances in Bcl-xL Research 2.0)

Abstract

:
Bcl-xL represents a family of proteins responsible for the regulation of the intrinsic apoptosis pathway. Due to its anti-apoptotic activity, Bcl-xL co-determines the viability of various virally infected cells. Their survival may determine the effectiveness of viral replication and spread, dynamics of systemic infection, and viral pathogenesis. In this paper, we have reviewed the role of Bcl-xL in the context of host infection by eight different RNA and DNA viruses: hepatitis B virus (HBV), hepatitis C virus (HCV), human immunodeficiency virus (HIV), influenza A virus (IAV), Epstein-Barr virus (EBV), human T-lymphotropic virus type-1 (HTLV-1), Maraba virus (MRBV), Schmallenberg virus (SBV) and coronavirus (CoV). We have described an influence of viral infection on the intracellular level of Bcl-xL and discussed the impact of Bcl-xL-dependent cell survival control on infection-accompanying pathogenic events such as tissue damage or oncogenesis. We have also presented anti-viral treatment strategies based on the pharmacological regulation of Bcl-xL expression or activity.
Keywords:
Bcl-xL; HBV; HCV; HIV; IAV; EBV; HTLV-1; MRBV; SBV; CoV

1. Introduction

Bcl-extra-large (Bcl-xL; UniProtKB: Q07817) is a member of the Bcl-2 protein family, a group of regulatory molecules involved in intrinsic apoptosis pathway regulation. The Bcl-2 family consists of both pro-apoptotic and anti-apoptotic factors. The first group is represented by three BH domains (i.e., Bax and Bak) as well as BH3-only proteins (i.e., Bim, Bad, Bid, Bik, Bmf, and Hrk). The anti-apoptotic Bcl-2 family members comprise a protein with four BH domains, including Bcl-2, Bcl-w, A1, Mcl-1, and Bcl-xL [1,2,3,4].
The effect of Bcl-xL activity on the apoptotic cell potential has been widely studied in the areas of tumor development [5,6]. Indeed, cell viability can be a potent factor to determine various diseases. Upregulation of the intracellular level of Bcl-xL is associated with the development of tumors, such as prostate cancer [7], lymphomas [8], advanced malignant melanoma [9], neuroblastoma [10], breast cancer [11], and many others. Bcl-xL also seems to be an important player in viral infections. In this paper, we have reviewed reports presenting the role of Bcl-xL in infections caused by different viral pathogens: hepatitis B virus (HBV), hepatitis C virus (HCV), human immunodeficiency virus (HIV), influenza A virus (IAV), Epstein-Barr virus (EBV), human T-lymphotropic virus type-1 (HTLV-1), Maraba virus (MRBV), Schmallenberg virus (SBV) and coronavirus (CoV).

2. The Role of Bcl-xL and Other Bcl-2 Family Members in the Regulation of Apoptosis

Apoptosis may be induced via two different cascades of signal transduction, the extrinsic and intrinsic pathway. The first one requires extracellular ligands to activate the cell surface receptors of the tumor necrosis factor (TNF) superfamily (TNFSF) [12]. For example, the first apoptosis signal (Fas) ligand (FasL) and TNF-α can stimulate Fas protein and TNFR1, respectively [13,14]. Activation of TNFSF members initiates intracellular signal transduction, and the cell undergoes apoptosis via a caspase-8/-3-mediated mechanism [15]. The intrinsic pathway is strictly dependent on the integrity and functionality of mitochondria as its essential property is to release pro-apoptotic markers (i.e., cytochrome c) from the mitochondrial intermembrane space (IMS) to the cytosol [16,17]. As a result, it increases the outer mitochondrial membrane (OMM) permeability. The unsealing of mitochondria and the consecutive caspase activation cascade leads to programmed cell death [18,19,20].
Permeabilization of the OMM consistently results in the mitochondrial permeability transition (MPT). The MPT is a process that opens non-specific permeability transition pores (PTPs) located in the inner mitochondrial membrane (IMM) to communicate with the OMM. The PTP is a protein complex composed of a voltage-dependent anion channel (VDAC), benzodiazepine peripheral receptor (BPR), and an adenine nucleotide translocator (ANT) [21,22,23,24]. Under normal circumstances, the MPT appears in a small number of mitochondria without any substantial influence on the mitochondrial membrane potential (ΔΨm). Pro-apoptotic stimulation increases the MPT frequency, resulting in the loss of ΔΨm, permeabilization of the OMM, and the release of mitochondrial proteins to the cytosol [25]. The Bcl-2 protein family members can be recruited to the PTP and contribute to MPT regulation. Bax and Bak indirectly interact with the VDAC, resulting in the increase in PTP diameter until the release of cytochrome c, whereas Bcl-xL is a negative regulator of the mechanism [26].
Another inherent operation of the intrinsic apoptosis induction is the PTP-independent oligomerization of the pro-apoptotic Bcl-2 family members. The group consists of proteins that interact with each other to promote a mitochondrial pathway of type I programmed cell death, or to counteract it, affecting mitochondrial integrity [25,26,27]. Active forms of the three domain members of the Bcl-2 protein family, Bax and Bak, are able to integrate into the OMM, oligomerize within a lipid bilayer, and form pores that enable cytochrome c and other pro-apoptotic factors to escape the intermembrane mitochondrial space and move to the cytosol. In the cytosol, cytochrome c binds apoptotic protease activating factor-1 (Apaf-1) to form an apoptosome, a complex responsible for the initiation of the caspase cascade via caspase-9 activation. The anti-apoptotic Bcl-2 family proteins, such as Bcl-xL, can bind Bax and Bak to inactivate them and prevent PTP-independent mitochondrial permeabilization [28,29,30]. Many viruses are able to upregulate or downregulate Bcl-xL to affect the viability of the cell (Figure 1).

3. The Role of Bcl-xL in HBV Infection

Hepatitis B virus (HBV) is a member of the Hepadnaviridae family [31] and the Orthohepadnavirus genus [32]. Its genome has the form of partially double-stranded DNA that undergoes reverse transcription during infection. Genetic material is enclosed in an icosahedral capsid, encircled by a lipoprotein envelope [33,34]. The diameter of infectious viral particles is 44 nm [35]. The DNA genome comprises four open reading frames that overlap each other and encode seven proteins: polymerase (pol), a structural capsid protein named HBV core antigen (HBcAg), HBV x protein (HBx), secreted HBV e antigen (HBeAg), and three forms of surface proteins (HBsAg): large (L), medium (M), and small (S) ones [36,37,38]. Studies on HBV are very important, because a combination of pathogenicity, genome variability, and high prevalence of the virus may pose serious problems for modern medicine. HBV is an etiological agent of both acute and chronic hepatitis [36]. Viral infection may lead to liver cirrhosis and hepatocellular carcinoma (HCC) [39]. Deficiency in the proofreading activity of HBV reverse transcriptase causes high mutation rates of the viral genome, resulting in the emergence of ten HBV genotypes with different geographic localization [39,40,41]. It is estimated that approximately 257 million people worldwide are infected with HBV [36].
Research has shown that HBV-host interactions included contact between HBx and anti-apoptotic members of the Bcl-2 family. HBx targets Bcl-xL and Bcl-2 via its BH3-like motif. Geng et al. [42] have determined that interaction between HBx and Bcl-2 family proteins has a significant impact on viral pathogenesis. During the experiment, the mutations in the BH3-like motif deprived HBx of the ability to bind Bcl-xL and Bcl-2, abolishing the HBx-dependent increase in the cytosolic calcium level in infected hepatocytes. Mutational inactivation of HBx prevented cell death and markedly reduced viral replication, a process requiring high levels of cytosolic Ca2+. Silencing Bcl-xL expression with the use of short hairpin RNA (shRNA) has shown the importance of anti-apoptotic protein in viral replication. Bcl-xL knockdown reduced the cytosolic calcium level in the infected HepG2 cells, decreasing HBV replication. Summarizing, the obtained results have suggested that Bcl-xL cooperated with HBx to promote viral reproduction and infected cell death [42]. Zhang et al. have analyzed the crystal structure of the HBx BH-3-like motif bound to Bcl-xL protein. Exploring the motif further, the team has identified a short α-helix and its ability to enter the hydrophobic pocket of Bcl-xL [43]. Because of its role in the HBV-host cell interaction, Bcl-xL should be considered an antiviral treatment [42,43].
Another reason behind examining Bcl-xL further is its association with hepatocarcinogenesis. Upregulation of Bcl-xL expression causes a decrease in the apoptotic potential of HCC cells that contributes to their survival, intensified growth, and colony formation [44]. The study of Kong et al. [45] on HBV-positive HCC has shown that HBV causes an increase in Bcl-xL in an IL-34-dependent manner [45].

4. The Role of Bcl-xL in HCV Infection

Hepatitis C virus (HCV) belongs to the Flaviviridae family [46] and the Hepacivirus genus [47]. Its genome has the form of positive-sense, single-stranded RNA (ssRNA), and is enclosed in an icosahedral capsid. The enveloped virion is 50–60 nm in diameter. The RNA genome contains one open reading frame [48] that is flanked by two non-coding regions (5’-NTR and 3’-NTR) and encodes one viral polyprotein. After expression, the product is cleaved into the structural proteins (i.e., protein capsid and two envelope proteins) and the non-structural (NS) ones, including proteases and viral RNA-dependent RNA polymerase (vRdRp) [49,50]. HCV is the etiological agent of chronic liver disease that can develop into cirrhosis. HCV infection may also lead to hepatocellular carcinoma [51]. It is estimated that 170 million people worldwide are chronically infected with HCV [52].
Scientific studies have suggested that the downregulation of Bcl-xL may play an important role in intracellular HCV infection and pathogenesis [53]. Scientific findings suggest that the pathogenic processes promote the death of infected cells. Apart from necrosis and autophagy, hepatocyte apoptosis plays a significant role in the development of liver disease [54]. Javed and Manzoor [53] have examined levels of two Bcl-2 family members, anti-apoptotic Bcl-xL and pro-apoptotic Bax, in Huh-7 cells (hepatocytes derived from the cellular carcinoma cell line) transfected with NS4A and NS3-NS4A of HCV genotype 3a. The presence of viral proteins resulted in a decrease in the intracellular level of Bcl-xL and a parallel promotion of Bax expression. A substantial change in the proportion Bcl-xL/Bax led to a series of pro-apoptotic molecular events, including Bax translocation to the mitochondria and the initiation of a caspase cascade preceding cell death [53]. In another study, Joyce et al. [55] observed damage to and the apoptosis of human hepatocytes in chimeric mice (animals transplanted with human liver cells) inoculated with HCV. The viral infection led to an increase in total Bcl-xL levels in the liver, however, within HCV-positive organs, Bcl-xL expression in the HCV-infected cells was lower than in the non-infected ones. The results suggested that hepatocytes responded to the liver infection-associated stress by upregulation of Bcl-xL and promoting their survival, whereas HCV impaired this response by counteracting the increase in Bcl-xL and favoring apoptosis [55].
Pharmacological regulation of hepatocyte apoptosis is considered a promising therapeutic strategy against HCV. Yang et al. [56] have used an animal model of HCV infection, Tupaia belangeri, to examine the efficacy of xanthohumol, a prenylated chalcone isolated from hops, in liver disease treatment. The therapeutic agent had a beneficial influence on the HCV-infected animals, which was manifested by a reduction of steatosis, fibrosis, and hepatic inflammation. The findings of the team enabled them to determine the possible association between liver condition improvement and the decrease in the apoptotic potential of hepatocytes. Xanthohumol exerted its therapeutic effect by increasing the expression of Bcl-xL and lowering the level of Bax in the liver tissues. Therefore, the Bcl-xL was shown to be a factor that might affect the clinical picture of liver pathogenesis by participating in the regulation of hepatocyte apoptosis and the enhancement of Bcl-xL’s contribution to the treatment effect by promoting cell survival [56].
On the other hand, Bcl-xL may also be co-responsible for HCV-induced oncogenesis. Research by Guo et al. [57] on human fetal liver stem cells (hFLSCs) infected with HCV has revealed nuclear factor-ĸB (NF-ĸB)-dependent mechanisms that counteract HCV-induced apoptosis at the late stage of infection. One of the cell survival-promoting events was the upregulation of intracellular Bcl-xL [57]. Meanwhile, previous studies had found the overexpression of Bcl-xL in HCCs and suggested the contribution of its anti-apoptotic activity to HCC survival under various stress conditions, and to hepatocellular carcinoma development [58,59].

5. The Role of Bcl-xL in HIV Infection

Human immunodeficiency virus (HIV) belong to the Retroviridae family and the Lentivirus genus [60]. Its genome has the form of two copies of identical, positive-sense, single-stranded RNA, each of them containing a complete set of genetic information necessary for HIV to replicate in the host cell. An essential step in viral reproduction is reverse transcription that is indispensable in the production of the double-stranded DNA (dsDNA) molecule that integrates into the host genome [60,61]. Genetic material is encased in an enveloped, cone-shaped capsid [62]. An average mature virion diameter is 145 nm [63]. The HIV DNA genome comprises several genes located between long terminal repeats (5’-LTR and 3’-LTR) at either end. Among the HIV genes, gag, pol, and env encode precursors of inner structural proteins, viral enzymes (including reverse transcriptase), and envelope proteins, respectively. The env open reading frame is followed by the nef gene encoding multifunctional negative regulating factor (Nef). Another important HIV gene is vpr and its product is pro-apoptotic viral protein R (Vpr) [60]. HIV is the etiological agent of acquired immunodeficiency syndrome (AIDS) and remains one of the major challenges for modern medicine. It poses a serious public health problem due to its ability to hide inside the immune cells of an HIV-positive individual [64]. Estimates show that the number of HIV-1-positive people is approximately 38 million individuals worldwide [65]. The HIV pandemic has spread across the whole world and, at present, there is no region free from the virus and the disease it causes [64].
An early study of Petrovas et al. [66] examined the influence of HIV infection on the Bcl-xL level in HIV-specific CD8+ T lymphocytes [66], host cells responsible for mediating the immune response against the pathogen [67]. The team found an HIV-dependent decrease in the Bcl-xL expression, a phenomenon that, along with Bcl-2 deficiency, seemed to elevate cell sensitivity to apoptosis. To confirm the potent role of the anti-apoptotic Bcl-2 family members in maintaining CD8+ T cell viability, IL-15, a cytokine promoting their activation and proliferation, was implemented. HIV-infected CD8+ T cell cultures treated with IL-15 were resistant to apoptosis, and their survival coincided with the stimulation of Bcl-xL and Bcl-2 expression. The study indicated the contribution of the anti-apoptotic Bcl-2 protein family to the antiviral effect of the IL-15 treatment [66].
HIV infection is extremely difficult to treat because of the virus’s ability to hide inside immune cells. The cellular reservoirs consist of memory CD4+ T lymphocytes and myeloid cells including macrophages. Those cells constitute a “refuge” responsible for the persistence of the viral infection. Therefore, the ability of infected macrophages to counteract apoptosis promotes the spread of the virus [68]. Meanwhile, previous studies of Sanz et al. [69] have shown that the differentiation of hematopoietic progenitor cells in macrophages is associated with a substantial increase in the intracellular level of Bcl-xL [69]. Research by Choi and Smithgall [70] has revealed that the viral Nef protein enhances the expression of Bcl-xL in a human macrophage precursor cell line, TF-1. Nef-dependent upregulation of Bcl-xL via the Erk signaling pathway promoted cell resistance to apoptotic stimulation by cytokine deprivation. Further analysis by Busca et al. [71] has elucidated that Bcl-xL (together with another anti-apoptotic representative of the Bcl-2 family, Mcl-1) determines the overall cell viability, as the silenced expression of Bcl-xL and Mcl-1 results in macrophage apoptosis. However, the team revealed that resistance to Vpr-induced cell death was achieved via a Bcl-xL-independent mechanism [71].
Many reports are devoted to the mechanisms of AIDS-associated disorders. HIV infection may lead to neurocognitive impairment. It can replicate in the immune system and supporting cells of the central nervous system (CNS) to provoke an immune–inflammatory response [72,73]. HIV-positive macrophages release neurotoxins that cause the death of host neuronal cells [74]. Chen et al. [74] have studied the potential role of Bcl-xL in the cellular response to HIV infection. The team used NT2.N/Bcl-xL neuronal cell lines characterized by the overexpression of Bcl-xL to determine whether the anti-apoptotic Bcl-2 family members are able to protect neurons from HIV-1-infected macrophages. The experiment showed that high levels of Bcl-xL result in resistance to the intrinsic apoptosis pathway induced by macrophage neurotoxins. The results suggest considering Bcl-xL overexpression in neuronal cells as part of a strategy for the treatment of AIDS-associated disorders [74].

6. The Role of Bcl-xL in Influenza A Virus Infection

Influenza A virus (IAV) is a viral genus representing the Orthomyxoviridae family [75]. The IAV genome has the form of negative-sense ssRNA divided into eight segments and enclosed in an enveloped virion [76] of 100 nm in diameter [77]. Each segment encodes at least one viral protein. Segments 1, 2, and 3 consist of genes for PB2, PB1, and PA proteins, respectively, and are subunits of vRdRP. Segments 4 and 6 encode hemagglutinin (HA) and neuraminidase (NA), respectively, and are both located on the surface of the virion envelope. The IAV serotypes are dependent on the variants of HA and NA [76]. IAV infects animals such as birds [78], seals [79], ferrets [80], mink [81], pigs [82], horses, and dogs, but also poses a threat to humans as a zoonotic agent. The main reservoir for IAV is aquatic birds [83]. IAV is often responsible for mild infections in humans. However, a few serotypes, such as H1N1, H5N1, H5N6, H10N8, and H7N9, have proven to be the cause of severe illnesses or even death in infected patients [84].
Research by Kakkola et al. [85] has indicated that the anti-apoptotic Bcl-2 family members may be involved in effective viral replication at the early stage of infection. The team has revealed that, in humans, in IAV-infected retinal pigment epithelium (RPE) cells, the interaction of Bcl-xL, Bcl-2, and Bcl-w with Bax, Bad, and the uveal autoantigen with coiled-coil domains and ankyrin repeats (UACA) counteracted the OMM permeabilization, thwarted the consequent release of pro-apoptotic factors from the mitochondria to the cytosol, and prevented the downstream signaling cascade. The role of Bcl-xL, Bcl-2, and Bcl-w seemed to be to delay the intrinsic apoptosis induction at the same time as maintaining cell viability at the early stage of infection to enable the synthesis of progeny viral ribonucleoprotein (vRNP) in the nucleus. The report of Kakkola et al. [85] suggests improving the anti-IAV therapies by diminishing the activity or reducing the intracellular level of the anti-apoptotic Bcl-2 family members, including Bcl-xL. The team has examined the pro-apoptotic activity of the anti-cancer compound ABT-263 in IAV-infected cells. In light of the proposed model, at the early stage of the infection, ABT-263 binds Bcl-xL and enhances its dissociation from Bax, Bad, and UACA. After release from the inactivating complex with Bcl-xL, Bax and Bad cause mitochondria permeabilization and start the intrinsic apoptosis pathway (with free UACA as one of the downstream pro-apoptotic molecular players). Therefore, the ABT-263–Bcl-xL interaction impairs the metabolism and functions of an infected cell and thus interrupts viral replication. The obtained results have suggested the possible effectiveness of treatment strategies that use Bcl-xL as a potential drug target [85].
On the other hand, the report of Lee et al. [86] has suggested that Bcl-xL may counteract virus spread by directing the cellular signaling pathways to pyroptosis. The team studied the possible role of Bcl-xL in the response of respiratory epithelial cells to IAV infection. The obtained results suggested that viral infection led to the substantial upregulation of Bcl-xL via Janus kinase/signal transducers and the activators of transcription (JAK-STAT) signaling pathway, a molecular cascade induced by type I interferon (IFN). In light of the report, it seems that the anti-apoptotic activity of Bcl-xL contributed to the switch of cellular signaling from the apoptotic program to the pyroptotic one at the late stage of IAV infection. The pyroptosis of infected respiratory epithelial cells induced an inflammatory response, leading to the mobilization of macrophages and neutrophils to eliminate the infected cells and stop the viral spread [86].

7. The Role of Bcl-xL in EBV Virus Infection

Epstein–Barr virus (EBV) is a member of the Herpesviridae family [87] and the Lymphocryptovirus genus [88]. Its genome has the form of linear double-stranded DNA [89,90] and is encased in an enveloped [91] icosahedral capsid [92]. The EBV virion is approximately 115 nm in diameter [93]. The viral DNA encodes over 85 genes, including DNA polymerase [94]. EBV is characterized by its very high prevalence—it is estimated that over 90% of the human population is persistently infected with the virus. The primary infection of EBV in infants does not occur with any specific symptoms—it can even be completely asymptomatic [88]. The virus is able to persist in some infected B lymphocytes and cause reactive lymphoid hyperplasia in immunodeficient hosts [95]. EBV-positive adults may also suffer from B cell lymphoproliferative disorders and lymphomas (i.e., Burkitt’s lymphoma) [96,97], undifferentiated nasopharyngeal carcinoma (NPC) [98], gastric carcinoma [99], chronic active EBV (CAEBV), or extranodal natural killer (NK)/T cell lymphoma (ENKTL) [100].
In light of scientific reports, it seems that Bcl-xL plays an important role in EBV-associated diseases. Portis et. al. [101] have revealed that one of the viral late gene expression products, latent membrane protein-2A (LMP-2A), is responsible for the upregulation of Bcl-xL expression in EBV-infected B lymphocytes, favoring cell viability and latent EBV infection, possibly contributing to EBV-associated tumorigenesis [101,102]. The study of Zhang et al. [103] has suggested that the EBV protein, BARF1, promoted gastric carcinoma development by inducing Bcl-xL and Bcl-2 expression in a process involving the mitogen-activated protein kinase (MAPK)/c-Jun signaling pathway [103]. Sejic et al. [100] have performed studies on cancer cell lines derived from individuals with CAEBV or the ENKTL. All of the tested tumor cells were resistant to DNA damage-inducing chemotherapeutic agents. Meanwhile, A-1331852, a chemical compound mimicking the BH-3 domain of Bcl-xL, was able to induce apoptosis. The results have suggested that Bcl-xL activity was responsible for CAEBV- and ENKTL-derived tumor cell viability and their sustained growth [100].

8. The Role of Bcl-xL in HTLV-1 Virus Infection

Human T-lymphotropic virus type-1 (HTLV-1) is a member of the Retroviridae family [104] and the Deltaretrovirus genus [105]. It contains two strands of positive-sense, single-stranded RNA. An enveloped virion is approximately 100 nm in diameter. During the infection, viral RNA undergoes reverse transcription, leading to the synthesis of double-stranded DNA that integrates into the host genome [106]. HTLV-1 DNA contains genes for structural and functional proteins (i.e., capsid protein, protease, polymerase, and envelope protein), and also comprises four overlapping open reading frames encoding regulatory factors [106,107]. HTLV-1 infects immune system cells, such as CD4+ and CD8+ T lymphocytes and monocytes. The infection tends to be asymptomatic, however, in 2–5% of infected people, it is associated with adult T cell leukemia/lymphoma (ATLL) or HTLV-1-associated myelopathy (HAM)/tropical spastic paraparesis (TSP) [107].
ATLL development is driven by the clonal expansion of CD4+CD25+ T cells, therefore, their viability seems to be a key factor determining the progress of the disease [108]. Studies have revealed that the transactivator from the X-gene region (Tax), a protein encoded by HTLC-1, is responsible for the upregulation of Bcl-xL levels in ATLL T lymphocytes [109,110]. Zhang et al. [108] have proposed Bcl-xL as the target of antiviral therapy. The team has examined navitoclax, an inhibitor of Bcl-xL activity. The tested agent, used in combination with the JAK inhibitor ruxolitinib, exerted a strong, anti-tumor effect in ATLL treatment while, acting alone, it showed modest therapeutic effectiveness. In ATLL cells, navitoclax combined with ruxolitinib caused pro-apoptotic events, including caspase-3/7 activation. An experiment on tumor-bearing mice showed that a reduction in cancer cell viability led to the inhibition of tumor growth and promoted survival of the tested animals [108]. Mori et al. [111] have determined the ability of rottlerin, a protein kinase C (PKC)-δ inhibitor, to kill HTLV-1 infected cells. The research showed that rottlerin prevented PKC-δ phosphorylation and caused several events that significantly decreased cell viability. One of them was the downregulation of Bcl-xL [111]. Another study on ATLL treatment strategies was devoted to the evaluation of cerdulatinib, a chemical agent inhibiting JAK and spleen tyrosine kinase (SYK). Ishikawa et al. [112] have determined its anti-proliferative and pro-apoptotic impact on HTLV-1 infected T lymphocytes. The activity of cerdulatinib led to a number of pro-apoptotic events, including a decrease in the intracellular level of Bcl-xL [112].

9. The Role of Bcl-xL in Maraba Virus Infection

Maraba virus (MRBV) is a member of the Rhabdoviridae family and the Vesiculovirus genus that contains a negative-sense ssRNA genome [113]. Its enveloped bullet-shaped virion is 170 nm in length and 70 nm in diameter. The viral genome comprises five open reading frames that encode a nucleocapsid protein, phosphoprotein, matrix protein, glycoprotein, and polymerase, respectively [114]. MRBV is closely related to vesicular stomatitis virus (VSV), an archetypical vesiculovirus that is well known for its oncolytic properties. The antigenic relationship with VSV makes MRBV a potential candidate in anti-tumor treatment [113,115]. Many scientific reports have shown that MRBV is non-pathogenic to humans, although its antigens have caused one documented case of seroconversion. Summarizing, the advantages presented above prove MRBV to be an effective and safe instrument in oncomedicine [114]. Brun et al. [116] found that some mutations of the viral genome can elevate the selectivity of infection, directing it to the tumor cells. In the study, MRBV was examined for its lytic properties towards human tumor cells, such as breast, colon, lung, ovarian, prostate, renal, CNS, and melanoma cancer cells. The virus was highly effective in destroying cancerous cells: all tested tumor cell lines were susceptible to the cytolytic activity. In addition, MRBV was characterized by rapid virion production and high viral burst size. The results have encouraged the researchers to produce mutants with increased selectivity towards cancer cells. Double amino acid substitution (Q242R within glycoprotein and L123W in matrix protein) in the MG1 mutant diminished the potency of the modified virus to destroy normal cells while maintaining hypervirulence in the cancerous ones [116]. Subsequent studies have proven the therapeutic properties of Maraba MG1 virus in sarcoma treatment [117], human papillomavirus-positive (HPV+) malignancies [118], and ovarian cancer [119,120]. Apart from performing its direct oncolytic activity, the MG1 mutant exerts its therapeutic effect by modifying the tumor microenvironment (TME) and sensitizing it to the mechanism of the host immune response [121].
Because of MRBV’s anti-tumor potency, the molecular framework of viral infection requires closer attention and deeper investigation. An interaction between virus and host cells is reflected in the systemic scale of oncogenesis, therefore, it should be well characterized. To date, Hassanzadeh et al. [113] have conceptualized the influence of Maraba MG1 virus on the expression of Blc-xL protein. In infected cells, two factors, 4E-BP1 and eIF2α, were involved in the downregulation of global protein synthesis. However, the whole-cell translational arrest was contrasted with a substantial elevation of Bcl-xL protein levels. A study on the U2OS cells has indicated the intensification of the Bcl-xL mRNA translation process, probably in an eIF2α-dependent manner. Upregulation of Bcl-xL increased cell viability and consequently made it very potent to survival during infection. By counteracting the induction of the mitochondrion-dependent apoptosis pathway, it promoted the propagation of MRBV. The study on the role of Bcl-xL in viral infection control has enriched the view of intracellular mechanisms of host-virus interactions. The results have shown that Bcl-xL may be a highly important protein in viral anti-tumor therapies [113].

10. The Role of Bcl-xL in Schmallenberg Virus Infection

Schmallenberg virus (SBV) is a member of the Peribunyaviridae family and the Orthobunyavirus genus and contains a negative-sense ssRNA genome divided into three segments: large (L), medium (M), and small (S) ones, enclosed in an enveloped virion of 100 nm in diameter. The vRdRp is derived from the L-segment of SBV RNA. The M segment encodes a polyprotein that further undergoes proteolytic cleavage, splitting into two glycoproteins (Gn and Gc) and a nonstructural protein, NSm. The products of S-segment expression are a nucleocapsid protein (N) and nonstructural protein (NS) that are derived from the two overlapping open reading frames. SBV infects ruminants and causes hyperthermia and diarrhea and decreases milk supply. Symptoms of the infection include, but are not limited to, abortions and congenital anomalies such as hydranencephaly, scoliosis, brachygnathia, hydrocephalus, and arthrogryposis [122,123,124].
The molecular aspects of SBV infection are still not well known. The most important question is its influence on infected cell viability. Aksoy and Azkur [123] have studied the impact of SBV on the apoptotic potential of Vero cells. The research revealed that SBV infection triggered the intrinsic and extrinsic apoptosis pathways. The host cell–virus interaction resulted in the activation of both caspase-9 and caspase-8. The analyses showed that the anti-apoptotic members of the Bcl-2 family, Bcl-xL and Bcl-2, were downregulated. The outcome suggested that, to some extent, the pro-apoptotic effect of SBV infection may be due to the change in the levels of Bcl-xL. Nevertheless, further studies are necessary to investigate the importance of Bcl-xL in SBV infection and pathogenesis [123].

11. The Role of Bcl-xL in CoV Infection

Coronaviruses (CoVs) are a group of viruses belonging to the Coronaviridae family [125] and the Orthocoronavirinae subfamily [126]. The CoV genome has the form of positive-sense ssRNA [125] and is encased in an enveloped helical nucleocapsid. The virion is approximately 125 nm in diameter [126]. Severe acute respiratory syndrome (SARS)coronavirus (SARS-CoV) and SARS-CoV-2 are two highly pathogenic groups of CoV. SARS-CoV is characterized by the ability to cause SARS in infected humans. An outbreak of the disease emerged in 2002 and ceased in 2003. SARS-CoV-2 is the etiological agent of coronavirus disease 2019 (COVID-19) that has not yet been eradicated [127]. On 11 March 2020, the World Health Organization labeled its outbreak a pandemic, and the status of the disease has not changed since then [128].
An early report of Yang et al. [129] revealed the possible role of Bcl-xL in the pathogenesis of SARS. A feature of the disease is the substantial loss of host lymphocytes that leads to lymphopenia, therefore, the team used Jurkat T cells to examine the pro-apoptotic ability of SARS-CoV. The cells were transfected with cDNA encoding viral envelope (E) protein. SARS-CoV E protein increased the level of apoptosis, indicating the possible reason for the SARS-associated lymphopenia. Other tests revealed that Bcl-xL overexpression prevented the SARS-CoV E-induced death of Jurkat T cells (co-transfected with both SARS-CoV E and Bcl-xL cDNA). Further experiments showed that SARS-CoV E was able to interact with Bcl-xL through the BH3-like motif of SARS-CoV E and the BH3 domain of Bcl-xL [129]
Recent bioinformatic analysis of Navratil et al. [130] revealed structural homology between the C-terminal BH3-like motifs of SARS-CoV E and SARS-CoV-2 E. It suggested that SARS-CoV-2 E protein may also target Bcl-xL to affect the viability of infected cells [130].

12. Conclusions

Bcl-xL plays an important role in scenarios of various viral infections. SBV has been reported to cause a decrease in the intracellular level of Bcl-xL, whereas HBV, IAV, MRBV, EBV, and HTLV-1 are responsible for its overexpression. HCV and HIV infection can exert either a stimulatory or suppressive influence on Bcl-xL expression, according to the circumstances. Overexpression of Bcl-xL may counteract apoptosis induced by the envelope protein of SARS-CoV, while downregulation of Bcl-xL may be a molecular event leading to the development of viral pathogenesis, such as HIV-driven immune impairment or HCV-associated liver disease. On the other hand, the anti-apoptotic activity of Bcl-xL can also promote viral infections. By prolonging the survival of infected cells, Bcl-xL may favor the replication of IAV and MRBV. Bcl-xL overexpression is co-responsible for the persistence of HIV infection in immune cells acting as a viral “refuge”. Bcl-xL-dependent cell resistance to pro-apoptotic stimuli can also contribute to HBV-, HCV-, EBV-, and HTLV-1-associated tumorigenesis. Moreover, Bcl-xL may promote HCV replication by affecting the cytosolic level of Ca2+ in cooperation with HBx. Therefore, Bcl-xL is proposed by many reports to be a promising target molecule in anti-viral treatment.

Author Contributions

Z.W.; writing—original draft preparation, W.Ś.; investigation, M.B.M.; investigation and artwork, K.P.G.-Z.; investigation. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AIDSacquired immunodeficiency syndrome
ANT
Apaf-1
adenine nucleotide translocator
apoptotic protease activating factor-1
ATLLadult T cell leukemia/lymphoma
Bcl-xLBcl-extra-large
BPRbenzodiazepine peripheral receptor
CAEBVchronic active EBV
CNS
CoV
central nervous system
coronavirus
ds-DNA
Nef
double-stranded DNA
negative regulating factor
EBVEpstein–Barr virus
ENKTL
Fas
FasL
extranodal natural killer (NK)/T cell lymphoma
first apoptosis signal
first apoptosis signal (Fas) ligand
HAhemagglutinin
HAMHTLV-1-associated myelopathy
HBcAgHBV core antigen
HBeAgHBV e antigen
HBsAgHBV surface antigen
HBxHBV x protein
HCV
ssRNA
hepatitis C virus
single-stranded RNA
hFLSCshuman fetal liver stem cells
HIVhuman immunodeficiency virus
HPV+
HBV
human papillomavirus-positive
hepatitis B virus
HTLV-1human T-lymphotropic virus type-1
IAVinfluenza A virus
IFNinterferon
IMMinner mitochondrial membrane
IMSintermembrane mitochondrial space
JAKJanus kinase
LMP-2Alatent membrane protein-2A
MAPKmitogen-activated protein kinase
MPTmitochondrial permeability transition
MRBVMaraba virus
NAneuraminidase
NF-ĸBnuclear factor-ĸB
NPCnasopharyngeal carcinoma
NSnon-structural protein
OMMouter mitochondrial membrane
PKCprotein kinase C
polpolymerase
PTPpermeability transition pores
RPE
SARS
SARS-CoV
retinal pigment epithelium
severe acute respiratory syndrome
severe acute respiratory syndrome (SARS) coronavirus
SBVSchmallenberg virus
shRNAshort hairpin RNA
STATsignal transducer and activator of transcription
SYKspleen tyrosine kinase
Taxtransactivator from the X-gene region
TME
TNF
TNFSF
tumor microenvironment
tumor necrosis factor
tumor necrosis factor (TNF) superfamily
TSPtropical spastic paraparesis
UACAuveal autoantigen with coiled-coil domains and ankyrin repeats
VDACvoltage-dependent anion channel
Vprviral protein R
vRdRpviral RNA-dependent RNA polymerase
vRNPviral ribonucleoprotein
VSVvesicular stomatitis virus
ΔΨmmitochondrial membrane potential

References

  1. Choudhury, S. A comparative analysis of BCL-2 family. Bioinformation 2019, 15, 299–306. [Google Scholar] [CrossRef] [Green Version]
  2. Raemy, E.; Martinou, J.C. Involvement of cardiolipin in tBID-induced activation of BAX during apoptosis. Chem. Phys. Lipids 2014, 179, 70–74. [Google Scholar] [CrossRef]
  3. Roy, M.J.; Vom, A.; Czabotar, P.E.; Lessene, G. Cell death and the mitochondria: Therapeutic targeting of the BCL-2 family-driven pathway. Br. J. Pharmacol. 2014, 171, 1973–1987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Voss, A.K.; Strasser, A. The essentials of developmental apoptosis. F1000Research 2020, 9. [Google Scholar] [CrossRef]
  5. De Carné Trécesson, S.; Souazé, F.; Basseville, A.; Bernard, A.C.; Pécot, J.; Lopez, J.; Bessou, M.; Sarosiek, K.A.; Letai, A.; Barillé-Nion, S.; et al. BCL-XL directly modulates RAS signalling to favour cancer cell stemness. Nat. Commun. 2017, 8, 1–11. [Google Scholar] [CrossRef]
  6. Trisciuoglio, D.; Tupone, M.G.; Desideri, M.; Di Martile, M.; Gabellini, C.; Buglioni, S.; Pallocca, M.; Alessandrini, G.; D’aguanno, S.; Del Bufalo, D. BCL-XL overexpression promotes tumor progression-associated properties article. Cell Death Dis. 2017, 8, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Castilla, C.; Congregado, B.; Chinchón, D.; Torrubia, F.J.; Japón, M.A.; Sáez, C. Bcl-xL Is Overexpressed in Hormone-Resistant Prostate Cancer and Promotes Survival of LNCaP Cells via Interaction with Proapoptotic Bak. Endocrinology 2006, 147, 4960–4967. [Google Scholar] [CrossRef] [PubMed]
  8. Adams, C.M.; Mitra, R.; Vogel, A.N.; Liu, J.; Gong, J.Z.; Eischen, C.M. Targeting BCL-W and BCL-XL as a therapeutic strategy for Hodgkin lymphoma. Leukemia 2020, 34, 947–952. [Google Scholar] [CrossRef] [PubMed]
  9. Leiter, U.; Schmid, R.M.; Kaskel, P.; Peter, R.U.; Krähn, G. Antiapoptotic bcl-2 and bcl-xL in advanced malignant melanoma. Arch. Dermatol. Res. 2000, 292, 225–232. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, J.R.; Fletcher, B.; Page, C.; Hu, C.; Nunez, G.; Baker, V. Bcl-x(L) is expressed in ovarian carcinoma and modulates chemotherapy- induced apoptosis. Gynecol. Oncol. 1998, 70, 398–403. [Google Scholar] [CrossRef] [PubMed]
  11. España, L.; Fernández, Y.; Rubio, N.; Torregrosa, A.; Blanco, J.; Sierra, A. Overexpression of Bcl-x L in human breast cancer cells enhances organ-selective lymph node metastasis. Breast Cancer Res. Treat. 2004, 87, 33–44. [Google Scholar] [CrossRef]
  12. Lee, W.H.; Seo, D.; Lim, S.G.; Suk, K. Reverse Signaling of Tumor Necrosis Factor Superfamily Proteins in Factor Superfamily Proteins in Macrophages and microgia: Superfamily portrait in the neuroimmune interface. Front. Immunol. 2019, 10, 262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Du, P.; Li, S.-J.; Ojcius, D.M.; Li, K.-X.; Hu, W.-L.; Lin, X.; Sun, A.-H.; Yan, J. A novel Fas-binding outer membrane protein and lipopolysaccharide of Leptospira interrogans induce macrophage apoptosis through the Fas/FasL-caspase-8/-3 pathway. Emerg. Microbes Infect. 2018, 7, 1–17. [Google Scholar] [CrossRef] [Green Version]
  14. Jiang, Y.; Yu, M.; Hu, X.; Han, L.; Yang, K.; Ba, H.; Zhang, Z.; Yin, B.; Yang, X.P.; Li, Z.; et al. STAT1 mediates transmembrane TNF-alpha-induced formation of death-inducing signaling complex and apoptotic signaling via TNFR1. Cell Death Differ. 2017, 24, 660–671. [Google Scholar] [CrossRef] [Green Version]
  15. Kalkavan, H.; Green, D.R. MOMP, cell suicide as a BCL-2 family business. Cell Death Differ. 2018, 25, 46–55. [Google Scholar] [CrossRef]
  16. Gyulkhandanyan, A.V.; Mutlu, A.; Freedman, J.; Leytin, V. Mitochondrial permeability transition pore (MPTP)-dependent and -independent pathways of mitochondrial membrane depolarization, cell shrinkage and microparticle formation during platelet apoptosis. Br. J. Haematol. 2015, 169, 142–145. [Google Scholar] [CrossRef]
  17. Shamas-Din, A.; Bindner, S.; Chi, X.; Leber, B.; Andrews, D.W.; Fradin, C.É. Distinct lipid effects on tBid and Bim activation of membrane permeabilization by pro-apoptotic Bax. Biochem. J. 2015, 467, 495–505. [Google Scholar] [CrossRef]
  18. Fleury, C.; Mignotte, B.; Vayssière, J.L. Mitochondrial reactive oxygen species in cell death signaling. Biochimie 2002, 84, 131–141. [Google Scholar] [CrossRef]
  19. Sinha, K.; Das, J.; Pal, P.B.; Sil, P.C. Oxidative stress: The mitochondria-dependent and mitochondria-independent pathways of apoptosis. Arch. Toxicol. 2013, 87, 1157–1180. [Google Scholar] [CrossRef] [PubMed]
  20. Via, L.; Garcia-Argaeza, A.; Martínez-Vázquez, M.; Grancara, S.; Martinis, P.; Toninello, A. Mitochondrial Permeability Transition as Target of Anticancer Drugs. Curr. Pharm. Des. 2014, 20, 223–244. [Google Scholar] [CrossRef] [PubMed]
  21. Fang, D.; Maldonado, E.N. VDAC Regulation: A Mitochondrial Target to Stop Cell Proliferation. Adv. Cancer Res. 2018, 138, 41–69. [Google Scholar] [PubMed]
  22. Neginskaya, M.A.; Pavlov, E.V.; Sheu, S.S. Electrophysiological properties of the mitochondrial permeability transition pores: Channel diversity and disease implication. Biochim. Biophys. Acta Bioenerg. 2021, 1862, 148357. [Google Scholar] [CrossRef]
  23. Pastorino, J.; Hoek, J. Hexokinase II: The Integration of Energy Metabolism and Control of Apoptosis. Curr. Med. Chem. 2005, 10, 1535–1551. [Google Scholar] [CrossRef]
  24. Ueda, N. Ceramide-induced apoptosis in renal tubular cells: A role of mitochondria and sphingosine-1-phoshate. Int. J. Mol. Sci. 2015, 16, 5076–5124. [Google Scholar] [CrossRef] [Green Version]
  25. Gogvadze, V.; Orrenius, S.; Zhivotovsky, B. Multiple pathways of cytochrome c release from mitochondria in apoptosis. Biochim. Biophys. Acta Bioenerg. 2006, 1757, 639–647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Peng, T.I.; Jou, M.J. Oxidative stress caused by mitochondrial calcium overload. Ann. N. Y. Acad. Sci. 2010, 1201, 183–188. [Google Scholar] [CrossRef]
  27. Estaquier, J.; Vallette, F.; Vayssiere, J.L.; Mignotte, B. The mitochondrial pathways of apoptosis. Adv. Exp. Med. Biol. 2012, 942, 157–183. [Google Scholar] [PubMed]
  28. Chen, H.C.; Kanai, M.; Inoue-Yamauchi, A.; Tu, H.C.; Huang, Y.; Ren, D.; Kim, H.; Takeda, S.; Reyna, D.E.; Chan, P.M.; et al. An interconnected hierarchical model of cell death regulation by the BCL-2 family. Nat. Cell Biol. 2015, 17, 1270–1281. [Google Scholar] [CrossRef] [Green Version]
  29. Dai, H.; Pang, Y.P.; Ramirez-Alvarado, M.; Kaufmann, S.H. Evaluation of the BH3-only protein Puma as a direct Bak activator. J. Biol. Chem. 2014, 289, 89–99. [Google Scholar] [CrossRef] [Green Version]
  30. Xu, X.; Iwasa, H.; Hossain, S.; Sarkar, A.; Maruyama, J.; Arimoto-Matsuzaki, K.; Hata, Y. BCL-XL binds and antagonizes RASSF6 tumor suppressor to suppress p53 expression. Genes Cells 2017, 22, 993–1003. [Google Scholar] [CrossRef]
  31. Littlejohn, M.; Locarnini, S.; Yuen, L. Origins and evolution of hepatitis B virus and hepatitis D virus. Cold Spring Harb. Perspect. Med. 2016, 6, a021360. [Google Scholar] [CrossRef]
  32. Jacquet, S.; Pons, J.-B.; De Bernardo, A.; Ngoubangoye, B.; Cosset, F.-L.; Régis, C.; Etienne, L.; Pontier, D. Evolution of Hepatitis B Virus Receptor NTCP Reveals Differential Pathogenicities and Species Specificities of Hepadnaviruses in Primates, Rodents, and Bats. J. Virol. 2018, 93, 1738–1756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. McNaughton, A.L.; D’Arienzo, V.; Ansari, M.A.; Lumley, S.F.; Littlejohn, M.; Revill, P.; McKeating, J.A.; Matthews, P.C. Insights From Deep Sequencing of the HBV Genome—Unique, Tiny, and Misunderstood. Gastroenterology 2019, 156, 384–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Venkatakrishnan, B.; Zlotnick, A. The Structural Biology of Hepatitis B Virus: Form and Function. Annu. Rev. Virol. 2016, 3, 429–451. [Google Scholar] [CrossRef] [Green Version]
  35. Cao, J.; Zhang, J.; Lu, Y.; Luo, S.; Zhang, J.; Zhu, P. Cryo-EM structure of native spherical subviral particles isolated from HBV carriers. Virus Res. 2019, 259, 90–96. [Google Scholar] [CrossRef]
  36. Al-Sadeq, D.W.; Taleb, S.A.; Zaied, R.E.; Fahad, S.M.; Smatti, M.K.; Rizeq, B.R.; Al Thani, A.A.; Yassine, H.M.; Nasrallah, G.K. Hepatitis B virus molecular epidemiology, host-virus interaction, coinfection, and laboratory diagnosis in the MENA region: An update. Pathogens 2019, 8, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Churin, Y.; Roderfeld, M.; Roeb, E. Hepatitis B virus large surface protein: Function and fame. Hepatobiliary Surg. Nutr. 2015, 4, 1–10. [Google Scholar]
  38. Tu, T.; Budzinska, M.; Shackel, N.; Urban, S. HBV DNA Integration: Molecular Mechanisms and Clinical Implications. Viruses 2017, 9, 75. [Google Scholar] [CrossRef]
  39. Tong, S.; Revill, P. Overview of hepatitis B viral replication and genetic variability. J. Hepatol. 2016, 64, S4–S16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Yin, F.; Xie, Y.; Fan, H.; Zhang, J.; Guo, Z. Mutations in hepatitis B virus polymerase are associated with the postoperative survival of hepatocellular carcinoma patients. PLoS ONE 2017, 12, e0189730. [Google Scholar] [CrossRef] [Green Version]
  41. Yano, Y.; Azuma, T.; Hayashi, Y. Variations and mutations in the hepatitis B virus genome and their associations with clinical characteristics. World J. Hepatol. 2015, 7, 583–592. [Google Scholar] [CrossRef]
  42. Geng, X.; Huang, C.; Qin, Y.; McCombs, J.E.; Yuan, Q.; Harry, B.L.; Palmer, A.E.; Xia, N.S.; Xue, D. Hepatitis B virus X protein targets Bcl-2 proteins to increase intracellular calcium, required for virus replication and cell death induction. Proc. Natl. Acad. Sci. USA 2012, 109, 18471–18476. [Google Scholar] [CrossRef] [Green Version]
  43. Zhang, T.Y.; Chen, H.Y.; Cao, J.L.; Xiong, H.L.; Mo, X.B.; Li, T.L.; Kang, X.Z.; Zhao, J.H.; Yin, B.; Zhao, X.; et al. Structural and functional analyses of hepatitis B virus X protein BH3-like domain and Bcl-xL interaction. Nat. Commun. 2019, 10, 1–14. [Google Scholar] [CrossRef] [PubMed]
  44. Ge, H.; Zou, D.; Wang, Y.; Jiang, H.; Wang, L. MicroRNA-377 downregulates Bcl-xL and increases apoptosis in hepatocellular carcinoma cells. Oncol. Res. 2017, 25, 29–34. [Google Scholar] [CrossRef] [PubMed]
  45. Kong, F.; Zhou, K.; Zhu, T.; Lian, Q.; Tao, Y.; Li, N.; Tu, T.; Bi, Y.; Yang, X.; Pan, X.; et al. Interleukin-34 mediated by hepatitis B virus X protein via CCAAT/enhancer-binding protein α contributes to the proliferation and migration of hepatoma cells. Cell Prolif. 2019, 52, e12703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Zhang, Z.; Rong, L.; Li, Y.P. Flaviviridae Viruses and Oxidative Stress: Implications for Viral Pathogenesis. Oxid. Med. Cell. Longev. 2019, 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Billerbeck, E.; Wolfisberg, R.; Fahnøe, U.; Xiao, J.W.; Quirk, C.; Luna, J.M.; Cullen, J.M.; Hartlage, A.S.; Chiriboga, L.; Ghoshal, K.; et al. Mouse models of acute and chronic hepacivirus infection. Science 2017, 357, 204–208. [Google Scholar] [CrossRef] [Green Version]
  48. Yu, M.L.; Chuang, W.L. Treatment of chronic hepatitis C in Asia: When East meets West. J. Gastroenterol. Hepatol. 2009, 24, 336–345. [Google Scholar] [CrossRef]
  49. Moradpour, D.; Penin, F. Hepatitis C Virus Proteins: From Structure to Function. Curr. Top. Microbiol. Immunol. 2013, 369, 113–142. [Google Scholar]
  50. Vieyres, G.; Dubuisson, J.; Pietschmann, T. Incorporation of hepatitis C virus E1 and E2 glycoproteins: The keystones on a peculiar virion. Viruses 2014, 6, 1149–1187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Sanna, A.; Le Strat, Y.; Roudot-Thoraval, F.; Deuffic Burban, S.; Carrieri, P.; Delarocque-Astagneau, E.; Larsen, C. Severe liver disease related to chronic hepatitis c virus infection in treatment-naive patients: Epidemiological characteristics and associated factors at first expert centre visit, France, 2000 to 2007 and 2010 to 2014. Eurosurveillance 2017, 22, 1. [Google Scholar] [CrossRef] [Green Version]
  52. Mohamed, A.A.; Elbedewy, T.A.; El-Serafy, M.; El-Toukhy, N.; Ahmed, W.; El Din, Z.A. Hepatitis C virus: A global view. World J. Hepatol. 2015, 7, 2676–2680. [Google Scholar] [CrossRef]
  53. Javed, F.; Manzoor, S. HCV non-structural NS4A protein of genotype 3a induces mitochondria mediated death by activating Bax and the caspase cascade. Microb. Pathog. 2018, 124, 346–355. [Google Scholar] [CrossRef] [PubMed]
  54. Masalova, O.V.; Lesnova, E.I.; Solyev, P.N.; Zakirova, N.F.; Prassolov, V.S.; Kochetkov, S.N.; Ivanov, A.V.; Kushch, A.A. Modulation of cell death pathways by hepatitis C virus proteins in Huh7.5 hepatoma cells. Int. J. Mol. Sci. 2017, 18, 2346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Joyce, M.A.; Walters, K.A.; Lamb, S.E.; Yeh, M.M.; Zhu, L.F.; Kneteman, N.; Doyle, J.S.; Katze, M.G.; Tyrrell, D.L. HCV induces oxidative and ER stress, and sensitizes infected cells to apoptosis in SCID/Alb-uPA mice. PLoS Pathog. 2009, 5, e1000291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Yang, M.; Li, N.; Li, F.; Zhu, Q.; Liu, X.; Han, Q.; Wang, Y.; Chen, Y.; Zeng, X.; Lv, Y.; et al. Xanthohumol, a main prenylated chalcone from hops, reduces liver damage and modulates oxidative reaction and apoptosis in hepatitis C virus infected Tupaia belangeri. Int. Immunopharmacol. 2013, 16, 466–474. [Google Scholar] [CrossRef]
  57. Guo, X.; Liu, W.L.; Yang, D.; Shen, Z.Q.; Qiu, Z.G.; Jin, M.; Li, J.W. Hepatitis C virus infection induces endoplasmic reticulum stress and apoptosis in human fetal liver stem cells. J. Pathol. 2019, 248, 155–163. [Google Scholar] [CrossRef] [Green Version]
  58. Takehara, T.; Liu, X.; Fujimoto, J.; Friedman, S.L.; Takahashi, H. Expression and role of Bcl-xL in human hepatocellular carcinomas. Hepatology 2001, 34, 55–61. [Google Scholar] [CrossRef] [Green Version]
  59. Watanabe, J.; Kushihata, F.; Honda, K.; Sugita, A.; Tateishi, N.; Mominoki, K.; Matsuda, S.; Kobayashi, N. Prognostic significance of Bcl-xL in human hepatocellular carcinoma. Surgery 2004, 135, 604–612. [Google Scholar] [CrossRef]
  60. Seitz, R. Human Immunodeficiency Virus (HIV). Transfus. Med. Hemotherapy 2016, 43, 203–222. [Google Scholar]
  61. Moore, M.D.; Hu, W.S. HIV-1 RNA dimerization: It takes two to tango. AIDS Rev. 2009, 11, 91–102. [Google Scholar] [PubMed]
  62. Campbell, E.M.; Hope, T.J. HIV-1 capsid: The multifaceted key player in HIV-1 infection. Nat. Rev. Microbiol. 2015, 13, 471–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Briggs, J.A.G.; Wilk, T.; Welker, R.; Kräusslich, H.G.; Fuller, S.D. Structural organization of authentic, mature HIV-1 virions and cores. EMBO J. 2003, 22, 1707–1715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Simon, V.; Ho, D.D.; Abdool Karim, Q. HIV/AIDS epidemiology, pathogenesis, prevention, and treatment. Lancet 2006, 368, 489–504. [Google Scholar] [CrossRef] [Green Version]
  65. Horton, R. Offline: A dangerous virus, but not the one you think. Lancet 2020, 395, 854. [Google Scholar] [CrossRef]
  66. Petrovas, C.; Mueller, Y.M.; Dimitriou, I.D.; Bojczuk, P.M.; Mounzer, K.C.; Witek, J.; Altman, J.D.; Katsikis, P.D. HIV-Specific CD8 + T Cells Exhibit Markedly Reduced Levels of Bcl-2 and Bcl-x L. J. Immunol. 2004, 172, 4444–4453. [Google Scholar] [CrossRef] [Green Version]
  67. Collins, D.R.; Gaiha, G.D.; Walker, B.D. CD8+ T cells in HIV control, cure and prevention. Nat. Rev. Immunol. 2020, 20, 471–482. [Google Scholar] [CrossRef]
  68. Kruize, Z.; Kootstra, N.A. The Role of Macrophages in HIV-1 Persistence and Pathogenesis. Front. Microbiol. 2019, 10, 2828. [Google Scholar] [CrossRef] [Green Version]
  69. Sanz, C.; Benito, A.; Silva, M.; Albella, B.; Richard, C.; Segovia, J.C.; Insunza, A.; Bueren, J.A.; Fernández-Luna, J.L. The expression of Bcl-x is downregulated during differentiation of human hematopoietic progenitor cells along the granulocyte but not the monocyte/macrophage lineage. Blood 1997, 89, 3199–3204. [Google Scholar] [CrossRef] [Green Version]
  70. Choi, H.J.; Smithgall, T.E. HIV-1 Nef promotes survival of TF-1 macrophages by inducing Bcl-X L expression in an extracellular signal-regulated kinase-dependent manner. J. Biol. Chem. 2004, 279, 51688–51696. [Google Scholar] [CrossRef] [Green Version]
  71. Busca, A.; Saxena, M.; Kumar, A. Critical role for antiapoptotic Bcl-xL and Mcl-1 in human macrophage survival and cellular IAP1/2 (cIAP1/2) in resistance to HIV-Vpr-induced apoptosis. J. Biol. Chem. 2012, 287, 15118–15133. [Google Scholar] [CrossRef] [Green Version]
  72. Olivier, I.S.; Cacabelos, R.; Naidoo, V. Risk factors and pathogenesis of HIV-associated neurocognitive disorder: The role of host genetics. Int. J. Mol. Sci. 2018, 19, 3594. [Google Scholar] [CrossRef] [Green Version]
  73. Saylor, D.; Dickens, A.M.; Sacktor, N.; Haughey, N.; Slusher, B.; Pletnikov, M.; Mankowski, J.L.; Brown, A.; Volsky, D.J.; McArthur, J.C. HIV-associated neurocognitive disorder—Pathogenesis and prospects for treatment. Nat. Rev. Neurol. 2016, 12, 234–248. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, W.; Sulcove, J.; Frank, I.; Jaffer, S.; Ozdener, H.; Kolson, D.L. Development of a Human Neuronal Cell Model for Human Immunodeficiency Virus (HIV)-Infected Macrophage-Induced Neurotoxicity: Apoptosis Induced by HIV Type 1 Primary Isolates and Evidence for Involvement of the Bcl-2/Bcl-xL-Sensitive Intrinsic Apoptosis Pathway. J. Virol. 2002, 76, 9407–9419. [Google Scholar] [PubMed] [Green Version]
  75. Hause, B.M.; Collin, E.A.; Liu, R.; Huang, B.; Sheng, Z.; Lu, W.; Wang, D.; Nelson, E.A.; Li, F. Characterization of a novel influenza virus in cattle and swine: Proposal for a new genus in the Orthomyxoviridae family. MBio 2014, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Dawson, W.K.; Lazniewski, M.; Plewczynski, D. RNA structure interactions and ribonucleoprotein processes of the influenza A virus. Brief. Funct. Genom. 2018, 17, 402–414. [Google Scholar] [CrossRef] [Green Version]
  77. Vajda, J.; Weber, D.; Brekel, D.; Hundt, B.; Müller, E. Size distribution analysis of influenza virus particles using size exclusion chromatography. J. Chromatogr. A 2016, 1465, 117–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Ramey, A.M.; Reeves, A.B. Ecology of Influenza A Viruses in Wild Birds and Wetlands of Alaska. Avian Dis. 2020, 64, 109–122. [Google Scholar] [CrossRef]
  79. Puryear, W.B.; Keogh, M.; Hill, N.; Moxley, J.; Josephson, E.; Davis, K.R.; Bandoro, C.; Lidgard, D.; Bogomolni, A.; Levin, M.; et al. Prevalence of influenza A virus in live-captured North Atlantic gray seals: A possible wild reservoir. Emerg. Microbes Infect. 2016, 5, 1–9. [Google Scholar] [CrossRef]
  80. Guan, M.; Hall, J.S.; Zhang, X.; Dusek, R.J.; Olivier, A.K.; Liu, L.; Li, L.; Krauss, S.; Danner, A.; Li, T.; et al. Aerosol Transmission of Gull-Origin Iceland Subtype H10N7 Influenza A Virus in Ferrets. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Zhang, B.-S.; Li, L.-J.; Zhu, Q.; Wang, Z.; Yuan, P.; Zhou, G.-D.; Shi, W.-J.; Chu, X.-F.; Jiang, S.; Xie, Z.-J. Co-infection of H9N2 influenza virus and Pseudomonas aeruginosa contributes to the development of hemorrhagic pneumonia in mink. Vet. Microbiol. 2020, 240, 108542. [Google Scholar]
  82. Schwaiger, T.; Sehl, J.; Karte, C.; Schäfer, A.; Hühr, J.; Mettenleiter, T.C.; Schröder, C.; Köllner, B.; Ulrich, R.; Blohm, U. Experimental H1N1pdm09 infection in pigs mimics human seasonal influenza infections. PLoS ONE 2019, 14, e0222943. [Google Scholar] [CrossRef]
  83. Parrish, C.R.; Murcia, P.R.; Holmes, E.C. Influenza Virus Reservoirs and Intermediate Hosts: Dogs, Horses, and New Possibilities for Influenza Virus Exposure of Humans. J. Virol. 2015, 89, 2990–2994. [Google Scholar] [CrossRef] [Green Version]
  84. Bui, C.M.; Chughtai, A.A.; Adam, D.C.; MacIntyre, C.R. An overview of the epidemiology and emergence of influenza A infection in humans over time. Arch. Public Health 2017, 75, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Kakkola, L.; Denisova, O.V.; Tynell, J.; Viiliäinen, J.; Ysenbaert, T.; Matos, R.C.; Nagaraj, A.; Öhman, T.; Kuivanen, S.; Paavilainen, H.; et al. Anticancer compound ABT-263 accelerates apoptosis in virus-infected cells and imbalances cytokine production and lowers survival rates of infected mice. Cell Death Dis. 2013, 4, e742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Lee, S.; Hirohama, M.; Noguchi, M.; Nagata, K.; Kawaguchi, A. Influenza A Virus Infection Triggers Pyroptosis and Apoptosis of Respiratory Epithelial Cells through the Type I Interferon Signaling Pathway in a Mutually Exclusive Manner. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Keorochana, N.; Intaraprasong, W.; Choontanom, R. Herpesviridae prevalence in aqueous humor using PCR. Clin. Ophthalmol. 2018, 12, 1707–1711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Evans, T.S.; Lowenstine, L.J.; Gilardi, K.V.; Barry, P.A.; Ssebide, B.J.; Kinani, J.F.; Nizeyimana, F.; Noheri, J.B.; Cranfield, M.R.; Mudakikwa, A.; et al. Mountain gorilla lymphocryptovirus has Epstein-Barr virus-like epidemiology and pathology in infants. Sci. Rep. 2017, 7, 1–11. [Google Scholar]
  89. Hau, P.M.; Tsao, S.W. Epstein–Barr virus hijacks DNA damage response transducers to orchestrate its life cycle. Viruses 2017, 9, 341. [Google Scholar] [CrossRef] [Green Version]
  90. Tao, Q.; Young, L.S.; Woodman, C.B.J.; Murray, P.G. Epstein-Barr virus (EBV) and its associated human cancers—Genetics, epigenetics, pathobiology and novel therapeutics. Front. Biosci. 2006, 11, 2672–2713. [Google Scholar] [CrossRef] [Green Version]
  91. Li, Z.; Zhang, X.; Dong, L.; Pang, J.; Xu, M.; Zhong, Q.; Zeng, M.S.; Yu, X. CryoEM structure of the tegumented capsid of Epstein-Barr virus. Cell Res. 2020, 30, 873–884. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, W.-H.; Kuo, C.-W.; Chang, L.-K.; Hung, C.-C.; Chang, T.-H.; Liu, S.-T. Assembly of Epstein-Barr Virus Capsid in Promyelocytic Leukemia Nuclear Bodies. J. Virol. 2015, 89, 8922–8931. [Google Scholar] [CrossRef] [Green Version]
  93. Nemerow, G.R.; Cooper, N.R. Isolation of Epstein Barr-virus and studies of its neutralization by human IgG and complement. J. Immunol. 1981, 127, 272–278. [Google Scholar] [PubMed]
  94. Smatti, M.K.; Al-Sadeq, D.W.; Ali, N.H.; Pintus, G.; Abou-Saleh, H.; Nasrallah, G.K. Epstein-barr virus epidemiology, serology, and genetic variability of LMP-1 oncogene among healthy population: An update. Front. Oncol. 2018, 8, 211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Dojcinov, S.D.; Venkataraman, G.; Pittaluga, S.; Wlodarska, I.; Schrager, J.A.; Raffeld, M.; Hills, R.K.; Jaffe, E.S. Age-related EBV-associated lymphoproliferative disorders in theWestern population: A spectrum of reactive lymphoid hyperplasia and lymphoma. Blood 2011, 117, 4726–4735. [Google Scholar] [CrossRef] [Green Version]
  96. Pannone, G.; Zamparese, R.; Pace, M.; Pedicillo, M.; Cagiano, S.; Somma, P.; Errico, M.; Donofrio, V.; Franco, R.; De Chiara, A.; et al. The role of EBV in the pathogenesis of Burkitt’s Lymphoma: An Italian hospital based survey. Infect. Agent. Cancer 2014, 9, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Marques-Piubelli, M.L.; Salas, Y.I.; Pachas, C.; Becker-Hecker, R.; Vega, F.; Miranda, R.N. Epstein–Barr virus-associated B-cell lymphoproliferative disorders and lymphomas: A review. Pathology 2020, 52, 40–52. [Google Scholar] [CrossRef] [Green Version]
  98. Tsao, S.W.; Tsang, C.M.; Lo, K.W. Epstein-barr virus infection and nasopharyngeal carcinoma. Philos. Trans. R. Soc. B Biol. Sci. 2017, 372, 20160270. [Google Scholar] [CrossRef]
  99. Nishikawa, J.; Iizasa, H.; Yoshiyama, H.; Shimokuri, K.; Kobayashi, Y.; Sasaki, S.; Nakamura, M.; Yanai, H.; Sakai, K.; Suehiro, Y.; et al. Clinical importance of Epstein–Barr virus-associated gastric cancer. Cancers 2018, 10, 167. [Google Scholar] [CrossRef] [Green Version]
  100. Sejic, N.; George, L.C.; Tierney, R.J.; Chang, C.; Kondrashova, O.; MacKinnon, R.N.; Lan, P.; Bell, A.I.; Lessene, G.; Long, H.M.; et al. BCL-XL inhibition by BH3-mimetic drugs induces apoptosis in models of Epstein-Barr virus-associated T/NK-cell lymphoma. Blood Adv. 2020, 4, 4775–4787. [Google Scholar]
  101. Portis, T.; Longnecker, R. Epstein-Barr virus (EBV) LMP2A mediates B-lymphocyte survival through constitutive activation of the Ras/PI3K/Akt pathway. Oncogene 2004, 23, 8619–8628. [Google Scholar] [CrossRef] [Green Version]
  102. Ok, C.Y.; Li, L.; Young, K.H. EBV-driven B-cell lymphoproliferative disorders: From biology, classification and differential diagnosis to clinical management. Exp. Mol. Med. 2015, 47, e132. [Google Scholar] [CrossRef] [Green Version]
  103. Zhang, Y.; Xu, M.; Zhang, X.; Chu, F.; Zhou, T. MAPK/c-jun signaling pathway contributes to the upregulation of the anti-apoptotic proteins Bcl-2 and Bcl-xL induced by epstein-barr virus-encoded BARF1 in gastric carcinoma cells. Oncol. Lett. 2018, 15, 7537–7544. [Google Scholar] [CrossRef] [Green Version]
  104. Futsch, N.; Mahieux, R.; Dutartre, H. HTLV-1, the other pathogenic yet neglected human retrovirus: From transmission to therapeutic treatment. Viruses 2018, 10, 1. [Google Scholar] [CrossRef] [Green Version]
  105. Quaresma, J.A.S.; Yoshikawa, G.T.; Koyama, R.V.L.; Dias, G.A.S.; Fujihara, S.; Fuzii, H.T. HTLV-1, immune response and autoimmunity. Viruses 2015, 8, 5. [Google Scholar] [CrossRef] [Green Version]
  106. Kannian, P.; Green, P.L. Human T lymphotropic virus type 1 (HTLV-1): Molecular biology and oncogenesis. Viruses 2010, 2, 2037–2077. [Google Scholar] [CrossRef]
  107. Galli, V.; Nixon, C.C.; Strbo, N.; Artesi, M.; de Castro-Amarante, M.F.; McKinnon, K.; Fujikawa, D.; Omsland, M.; Washington-Parks, R.; Romero, L.; et al. Essential Role of Human T Cell Leukemia Virus Type 1 orf-I in Lethal Proliferation of CD4 + Cells in Humanized Mice. J. Virol. 2019, 93, 565–584. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, M.; Griner, L.A.M.; Ju, W.; Duveau, D.Y.; Guha, R.; Petrus, M.N.; Wen, B.; Maeda, M.; Shinn, P.; Ferrer, M.; et al. Selective targeting of JAK/STAT signaling is potentiated by Bcl-xL blockade in IL-2-dependent adult T-cell leukemia. Proc. Natl. Acad. Sci. USA 2015, 112, 12480–12485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Nicot, C.; Mahieux, R.; Takemoto, S.; Franchini, G. Bcl-X(L) is up-regulated by HTLV-I and HTLV-II in vitro and in ex vivo ATLL samples. Blood 2000, 96, 275–281. [Google Scholar] [CrossRef] [PubMed]
  110. Witzens-Harig, M.; Giaisi, M.; Köhler, R.; Krammer, P.H.; Li-Weber, M. HTLV-1-associated adult T cell leukemia is highly susceptible to Navitoclax due to enhanced Bax expression. Int. J. Cancer 2016, 138, 507–514. [Google Scholar] [CrossRef] [PubMed]
  111. Mori, N.; Ishikawa, C.; Senba, M. Activation of PKC-d in HTLV-1-infected T cells. Int. J. Oncol. 2015, 46, 1609–1618. [Google Scholar] [CrossRef] [Green Version]
  112. Ishikawa, C.; Senba, M.; Mori, N. Anti-adult T-cell leukemia/lymphoma activity of cerdulatinib, a dual SYK/JAK kinase inhibitor. Int. J. Oncol. 2018, 53, 1681–1690. [Google Scholar] [CrossRef] [Green Version]
  113. Hassanzadeh, G.; Naing, T.; Graber, T.; Jafarnejad, S.M.; Stojdl, D.F.; Alain, T.; Holcik, M. Characterizing cellular responses during oncolytic maraba virus infection. Int. J. Mol. Sci. 2019, 20, 580. [Google Scholar] [CrossRef] [Green Version]
  114. Pol, J.; Atherton, M.; Bridle, B.; Stephenson, K.; Le Boeuf, F.; Hummel, J.; Martin, C.; Pomoransky, J.; Breitbach, C.; Diallo, J.-S.; et al. Development and applications of oncolytic Maraba virus vaccines. Oncolytic Virotherapy 2018, 7, 117–128. [Google Scholar] [CrossRef] [Green Version]
  115. Breitbach, C.J. Considerations for Clinical Translation of MG1 Maraba Virus. Methods Mol. Biol. 2020, 2058, 285–293. [Google Scholar]
  116. Brun, J.; McManus, D.; Lefebvre, C.; Hu, K.; Falls, T.; Atkins, H.; Bell, J.C.; McCart, J.A.; Mahoney, D.; Stojdl, D.F. Identification of genetically modified maraba virus as an oncolytic rhabdovirus. Mol. Ther. 2010, 18, 1440–1449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Le Boeuf, F.; Selman, M.; Son, H.H.; Bergeron, A.; Chen, A.; Tsang, J.; Butterwick, D.; Arulanandam, R.; Forbes, N.E.; Tzelepis, F.; et al. Oncolytic Maraba Virus MG1 as a Treatment for Sarcoma. Int. J. Cancer 2017, 141, 1257–1264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Atherton, M.J.; Stephenson, K.B.; Pol, J.; Wang, F.; Lefebvre, C.; Stojdl, D.F.; Nikota, J.K.; Dvorkin-Gheva, A.; Nguyen, A.; Chen, L.; et al. Customized viral immunotherapy for HPV-associated cancer. Cancer Immunol. Res. 2017, 5, 847–859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Tong, J.G.; Valdes, Y.R.; Sivapragasam, M.; Barrett, J.W.; Bell, J.C.; Stojdl, D.; DiMattia, G.E.; Shepherd, T.G. Spatial and temporal epithelial ovarian cancer cell heterogeneity impacts Maraba virus oncolytic potential. BMC Cancer 2017, 17, 594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Tong, J.G.; Valdes, Y.R.; Barrett, J.W.; Bell, J.C.; Stojdl, D.; McFadden, G.; McCart, J.A.; DiMattia, G.E.; Shepherd, T.G. Evidence for differential viral oncolytic efficacy in an in vitro model of epithelial ovarian cancer metastasis. Mol. Ther. Oncolytics 2015, 2, 15013. [Google Scholar] [CrossRef]
  121. Pol, J.G.; Acuna, S.A.; Yadollahi, B.; Tang, N.; Stephenson, K.B.; Atherton, M.J.; Hanwell, D.; El-Warrak, A.; Goldstein, A.; Moloo, B.; et al. Preclinical evaluation of a MAGE-A3 vaccination utilizing the oncolytic Maraba virus currently in first-in-human trials. Oncoimmunology 2019, 8, e1512329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Wernike, K.; Beer, M. Schmallenberg Virus: A Novel Virus of Veterinary Importance. Adv. Virus Res. 2017, 99, 39–60. [Google Scholar]
  123. Aksoy, E.; Azkur, A.K. Schmallenberg virus induces apoptosis in vero cell line via extrinsic and intrinsic pathways in a time and dose dependent manner. J. Vet. Med. Sci. 2018, 81, 204–212. [Google Scholar] [CrossRef] [Green Version]
  124. Endalew, A.D.; Faburay, B.; Wilson, W.C.; Richt, J.A. Schmallenberg disease-a newly emerged culicoides-borne viral disease of ruminants. Viruses 2019, 11, 1065. [Google Scholar] [CrossRef] [Green Version]
  125. Leao, J.C.; Gusmao, T.P.D.L.; Zarzar, A.M.; Leao Filho, J.C.; Barkokebas Santos de Faria, A.; Morais Silva, I.H.; Gueiros, L.A.M.; Robinson, N.A.; Porter, S.; Carvalho, A.D.A.T. Coronaviridae—Old friends, new enemy! Oral Dis. 2020, 00, 1–9. [Google Scholar] [CrossRef]
  126. Malik, Y.A. Properties of coronavirus and SARS-CoV-2. Malays. J. Pathol. 2020, 42, 3–11. [Google Scholar] [PubMed]
  127. Petersen, E.; Koopmans, M.; Go, U.; Hamer, D.H.; Petrosillo, N.; Castelli, F.; Storgaard, M.; Al Khalili, S.; Simonsen, L. Comparing SARS-CoV-2 with SARS-CoV and influenza pandemics. Lancet Infect. Dis. 2020, 20, e238–e244. [Google Scholar] [CrossRef]
  128. Cucinotta, D.; Vanelli, M. WHO declares COVID-19 a pandemic. Acta Biomed. 2020, 91, 157–160. [Google Scholar] [PubMed]
  129. Yang, Y.; Xiong, Z.; Zhang, S.; Yan, Y.; Nguyen, J.; Ng, B.; Lu, H.; Brendese, J.; Yang, F.; Wang, H.; et al. Bcl-xL inhibits T-cell apoptosis induced by expression of SARS coronavirus E protein in the absence of growth factors. Biochem. J. 2005, 392, 135–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Navratil, V.; Lionnard, L.; Longhi, S.; Combet, C.; Aouacheria, A. The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) envelope (E) protein harbors a conserved BH3-like sequence. bioRxiv 2020. [Google Scholar]
Figure 1. The influence of viral infections on the cell viability via regulation of Bcl-extra-large (Bcl-xL) expression. The color of arrows represents the influence of viral infections on Bcl-xL expression (green: upregulation; red: downregulation; blue: upregulation or downregulation, according to the circumstances).
Figure 1. The influence of viral infections on the cell viability via regulation of Bcl-extra-large (Bcl-xL) expression. The color of arrows represents the influence of viral infections on Bcl-xL expression (green: upregulation; red: downregulation; blue: upregulation or downregulation, according to the circumstances).
Ijms 22 01956 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wyżewski, Z.; Świtlik, W.; Mielcarska, M.B.; Gregorczyk-Zboroch, K.P. The Role of Bcl-xL Protein in Viral Infections. Int. J. Mol. Sci. 2021, 22, 1956. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22041956

AMA Style

Wyżewski Z, Świtlik W, Mielcarska MB, Gregorczyk-Zboroch KP. The Role of Bcl-xL Protein in Viral Infections. International Journal of Molecular Sciences. 2021; 22(4):1956. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22041956

Chicago/Turabian Style

Wyżewski, Zbigniew, Weronika Świtlik, Matylda Barbara Mielcarska, and Karolina Paulina Gregorczyk-Zboroch. 2021. "The Role of Bcl-xL Protein in Viral Infections" International Journal of Molecular Sciences 22, no. 4: 1956. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22041956

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop