首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Despite detailed knowledge of the components of the spindle assembly checkpoint, a molecular explanation of how cells die after prolonged spindle checkpoint activation, and thus how microtubule inhibitors and other antimitotic drugs ultimately elicit their lethal effects, has yet to emerge. Mitotically arrested cells typically display extensive phosphorylation of two key antiapoptotic proteins, Bcl-xL and Bcl-2, and evidence suggests that phosphorylation disables their antiapoptotic activity. However, the responsible kinase has remained elusive. In this report, evidence is presented that cyclin-dependent kinase 1 (CDK1)/cyclin B catalyzes mitotic-arrest-induced Bcl-xL/Bcl-2 phosphorylation. Furthermore, we show that CDK1 transiently and incompletely phosphorylates these proteins during normal mitosis. When mitosis is prolonged in the absence of microtubule inhibition, Bcl-xL and Bcl-2 become highly phosphorylated. Transient overexpression of nondegradable cyclin B1 caused apoptotic death, which was blocked by a phosphodefective Bcl-xL mutant but not by a phosphomimetic Bcl-xL mutant, confirming Bcl-xL as a key target of proapoptotic CDK1 signaling. These findings suggest a model whereby a switch in the duration of CDK1 activation, from transient during mitosis to sustained during mitotic arrest, dramatically increases the extent of Bcl-xL/Bcl-2 phosphorylation, resulting in inactivation of their antiapoptotic function. Thus, phosphorylation of antiapoptotic Bcl-2 proteins acts as a sensor for CDK1 signal duration and as a functional link coupling mitotic arrest to apoptosis.The cell division cycle is controlled by checkpoints, which ensure the fidelity of chromosome replication and segregation, as well as orderly progression through the cell cycle. If these critical events cannot be completed as scheduled, damaged cells, which might otherwise pose a threat to the organism as precancerous cells, are eliminated (16). The mitotic checkpoint, for example, produces a “prevent anaphase” signal until all the chromosomes are properly attached to kinetochores (22). Microtubule inhibitors (MTIs) and other antimitotic agents prolong the activation of this checkpoint, causing mitotic arrest, which culminates in cell death generally via intrinsic apoptosis, providing a rationale for the use of these agents as antitumor agents (20, 31). Intrinsic or mitochondrial apoptosis is regulated by the Bcl-2 family of proteins, which exhibit either pro- or antiapoptotic properties (17, 37). The BH3-only proapoptotic members act as essential initiators of intrinsic apoptosis, whereas the multidomain proapoptotic members, Bax and Bak, act as essential mediators of mitochondrial membrane permeability. Antiapoptotic Bcl-2 family members, including Bcl-xL, Bcl-2, and Mcl-1, oppose apoptosis by binding to the proapoptotic members and neutralizing their activity.The molecular mechanisms leading to cell death in response to spindle checkpoint activation have yet to be established. Indeed, how the spindle checkpoint couples to pathways regulating cell survival and death still represents an unresolved issue in cell biology (26, 35). Nonetheless, it seems reasonable to hypothesize that signals generated in response to prolonged mitotic arrest are eventually transduced to the apoptotic machinery. In this regard, it is striking that MTIs consistently induce the phosphorylation of two key antiapoptotic proteins, Bcl-2 and Bcl-xL, whereas other apoptotic stimuli fail to do so (9, 13, 25). The results of studies with phosphodefective mutants of Bcl-2 and Bcl-xL indicate that phosphorylation antagonizes their antiapoptotic function (2, 33, 36), but the precise mechanism(s) has yet to be fully clarified.The identity of the kinase responsible for the extensive phosphorylation of Bcl-xL and Bcl-2 that occurs in response to sustained spindle checkpoint activation is unresolved. Identification of this kinase is considered to be of critical importance, since it will provide insight into the molecular links between mitotic arrest and cell death, as well as the molecular mechanism of action of antimitotic drugs. Several candidates have been proposed, including Raf-1 (3), Jun N-terminal protein kinase (JNK) (2, 11, 36), protein kinase A (PKA) (32), cyclin-dependent kinase 1 (CDK1) (24), and mammalian target of rapamycin (mTOR) (4). In general, however, conclusions have been correlative or have been based on the use of kinase inhibitors tested under conditions that precluded mitotic arrest and thus indirectly blocked the effects of MTIs. Thus, strong experimental evidence supporting identification is lacking.Here we present evidence that the CDK1/cyclin B kinase complex is responsible for mitotic arrest-induced Bcl-xL/Bcl-2 phosphorylation. Furthermore, we show that CDK1 transiently and incompletely phosphorylates these proteins during normal mitosis. The findings suggest a model whereby a switch in the duration of CDK1 activation, from transient during mitosis to sustained during mitotic arrest, dramatically increases the extent of Bcl-xL/Bcl-2 phosphorylation, resulting in inactivation of the antiapoptotic function of Bcl-xL/Bcl-2. Thus, CDK1-mediated phosphorylation of antiapoptotic Bcl-2 proteins acts as a key link coupling mitotic arrest to apoptosis.  相似文献   

2.
Antagonists of anti-apoptotic Bcl-2 family members hold promise as cancer therapeutics. Apoptosis is triggered when a peptide containing a BH3 motif or a small molecule BH3 peptidomimetic, such as ABT 737, binds to the relevant Bcl-2 family members. ABT-737 is an antagonist of Bcl-2, Bcl-xL, and Bcl-w but not of Mcl-1. Here we describe new structures of mutant BH3 peptides bound to Bcl-xL and Mcl-1. These structures suggested a rationale for the failure of ABT-737 to bind Mcl-1, but a designed variant of ABT-737 failed to acquire binding affinity for Mcl-1. Rather, it was selective for Bcl-xL, a result attributable in part to significant backbone refolding and movements of helical segments in its ligand binding site. To date there are few reported crystal structures of organic ligands in complex with their pro-survival protein targets. Our structure of this new organic ligand provided insights into the structural transitions that occur within the BH3 binding groove, highlighting significant differences in the structural properties of members of the Bcl-2 pro-survival protein family. Such differences are likely to influence and be important in the quest for compounds capable of selectively antagonizing the different family members.Apoptosis, or programmed cell death, is a fundamental cellular process required by all multicellular organisms for the elimination of redundant, damaged, or potentially dangerous cells (1). A consequence of dysregulated cell death is the survival of abnormal cells, which in some cases can proliferate uncontrollably and form tumors. Hence, strategies to activate cell death pathways may represent one avenue by which cancer cells can be killed. A major pathway to cell death is regulated by the Bcl-2 family of proteins that consists of both pro-apoptotic and pro-survival members (2). Those family members that promote cell death are divided into two subgroups; that is, the Bax/Bak molecules, which are the essential mediators of apoptosis, and the BH3-only proteins (such as Bim, Bad, Puma, Noxa, and several others), which are the initiators of the apoptotic cascade. Within the pro-survival faction there are five members including Bcl-2 itself, Bcl-xL, Bcl-w, Mcl-1, and A1. Overexpression of pro-survival proteins can confer a survival advantage to cancer cells. Critically, conventional anti-cancer therapies are often rendered ineffective by this overexpression and other up-stream defects, most prominently mutations in the tumor suppressor p53.One strategy to kill cancer cells is to develop molecules that can mimic the BH3-only proteins (3). These proteins function by engaging the pro-survival proteins, although the downstream consequences of this interaction remain controversial (4). These interactions are mediated by a short sequence motif called the BH3 domain on the BH3-only protein. The structures of a number of BH3 domains in complex with pro-survival proteins have been solved, and all reveal that the BH3 sequence forms an amphipathic helix that inserts into a hydrophobic groove on the surface of the pro-survival proteins (510). These structures suggest that it might be possible to develop drugs based on small organic molecules that mimic natural BH3 ligands and activate the cell death pathways.Although a number of “BH3-mimetic” molecules have now been described (1121), many of these appear to kill cells in a non-mechanism-based manner (22). Only ABT-737, developed by Abbott Laboratories (15, 23), has been shown to be a bona fide BH3-mimetic (22, 24), binding in the same hydrophobic groove of Bcl-xL as do BH3 ligands (25). ABT-263, a molecule in the same chemical class as ABT-737, has shown efficacy as a single agent in cancer cell lines and animal models of cancer (2628) and is currently in a phase I/II clinical trial in leukemia and lymphoma patients.One of the important aspects of BH3 peptide interactions with pro-survival proteins is selectivity; some BH3 ligands only bind certain subsets of pro-survival proteins (e.g. Bad only binds Bcl-xL, Bcl-2, and Bcl-w, whereas Noxa only binds Mcl-1 and A1), but others such as Bim and Puma are more promiscuous and bind all pro-survival proteins tightly (2931). This selectivity has important implications as it appears that a range of pro-survival proteins needs to be neutralized for cell death to proceed in some cell types (30, 3234). Selectivity is also an important issue for drug potency. ABT-737 and ABT-263 share the same binding profile as Bad, and consequently, their efficacy seems to be restricted to cells/tumors in which Mcl-1 (to which they do not bind with high affinity) is inactivated or limiting (22, 28, 3538). Hence, only a small subset of all tumors, in particular some hematological malignancies and small-cell lung cancers, appear to respond to ABT-737/ABT-263 when used as a single agent (15, 28). Therefore, to expand the repertoire of cancers that could be treated with BH3 mimetics, it is likely that the range of binding specificities of such molecules also needs to be expanded, with Mcl-1 being an obvious target candidate. Obatoclax (GX15-070), a small molecule developed by Gemin X Biotechnologies, has been reported to function in this way (14). However, this compound can kill cells deficient in both Bax and Bak, the essential mediators of apoptotic cell death (28); hence, its true mechanism of action is uncertain.Peptide ligands can provide a template for small molecule drug design, particularly when a structure is available for it in complex with its target. For example, highly effective small molecule antagonists of the inhibitors of apoptosis proteins were designed using peptides as a starting point (39, 40). Antagonists of Bcl-xL based on derivatized terphenyl and terephthalamide scaffolds that apparently mimic the BH3 α-helix have also been reported (17, 18), although in this case no complexes with the protein target have yet been described.Here we use mutagenesis data and structures of BH3 peptides in complex with pro-survival proteins to guide attempts to develop a BH3-mimetic that binds to Mcl-1. Although that goal was not achieved, our results demonstrate extreme conformational flexibility in the pro-survival protein Bcl-xL and yielded a novel Bcl-xL-selective antagonist. In contrast, Mcl-1 displays no backbone conformational flexibility when presented with the various ligands we describe.  相似文献   

3.
4.
5.
The release of cytochrome c from mitochondria, which leads to activation of the intrinsic apoptotic pathway, is regulated by interactions of Bax and Bak with antiapoptotic Bcl-2 family members. The factors that regulate these interactions are, at the present time, incompletely understood. Recent studies showing preferences in binding between synthetic Bcl-2 homology domain 3 and antiapoptotic Bcl-2 family members in vitro have suggested that the antiapoptotic proteins Mcl-1 and Bcl-xL, but not Bcl-2, restrain proapoptotic Bak from inducing mitochondrial membrane permeabilization and apoptosis. Here we show that Bak protein has a much higher affinity than the 26-amino acid Bak Bcl-2 homology domain 3 for Bcl-2, that some naturally occurring Bcl-2 allelic variants have an affinity for full-length Bak that is only 3-fold lower than that of Mcl-1, and that endogenous levels of these Bcl-2 variants (which are as much as 40-fold more abundant than Mcl-1) restrain part of the Bak in intact lymphoid cells. In addition, we demonstrate that Bcl-2 variants can, depending on their affinity for Bak, substitute for Mcl-1 in protecting cells. Thus, the ability of Bcl-2 to protect cells from activated Bak depends on two important contextual variables, the identity of the Bcl-2 present and the amount expressed.The release of cytochrome c from mitochondria, which leads to activation of the intrinsic apoptotic pathway, is regulated by Bcl-2 family members (15). This group of proteins consists of three subgroups: Bax and Bak, which oligomerize upon death stimulation to form a putative pore in the outer mitochondrial membrane, thereby allowing efflux of cytochrome c and other mitochondrial intermembrane space components; Bcl-2, Bcl-xL, Mcl-1, and other antiapoptotic homologs, which antagonize the effects of Bax and Bak; and BH3-only proteins2 such as Bim, Bid, and Puma, which are proapoptotic Bcl-2 family members that share only limited homology with the other two groups in a single 15-amino acid domain (the BH3 domain, see Ref. 6). Although it is clear that BH3-only proteins serve as molecular sensors of various stresses and, when activated, trigger apoptosis (3, 611), the mechanism by which they do so remains incompletely understood. One current model suggests that BH3-only proteins trigger apoptosis solely by binding and neutralizing antiapoptotic Bcl-2 family members, thereby causing them to release the activated Bax and Bak that are bound (reviewed in Refs. 9 and 10; see also Refs. 12 and 13), whereas another current model suggests that certain BH3-only proteins also directly bind to and activate Bax (reviewed in Ref. 3; see also Refs. 1417). Whichever model turns out to be correct, both models agree that certain antiapoptotic Bcl-2 family members can inhibit apoptosis, at least in part, by binding and neutralizing activated Bax and Bak before they permeabilize the outer mitochondrial membrane (13, 18, 19).Much of the information about the interactions between pro- and antiapoptotic Bcl-2 family members has been derived from the study of synthetic peptides corresponding to BH3 domains. In particular, these synthetic peptides have been utilized as surrogates for the full-length proapoptotic proteins during structure determinations (2022) as well as in functional studies exploring the effect of purified BH3 domains on isolated mitochondria (14, 23) and on Bax-mediated permeabilization of lipid vesicles (15).Recent studies using these same peptides have suggested that interactions of the BH3 domains of Bax, Bak, and the BH3-only proteins with the “BH3 receptors” of the antiapoptotic Bcl-2 family members are not all equivalent. Surface plasmon resonance, a technique that is widely used to examine the interactions of biomolecules under cell-free conditions (2426), has demonstrated that synthetic BH3 peptides of some BH3-only family members show striking preferences, with the Bad BH3 peptide binding to Bcl-2 and Bcl-xL but not Mcl-1, and the Noxa BH3 peptide binding to Mcl-1 but not Bcl-2 or Bcl-xL (27). Likewise, the Bak BH3 peptide exhibits selectivity, with high affinity for Bcl-xL and Mcl-1 but not Bcl-2 (12). The latter results have led to a model in which Bcl-xL and Mcl-1 restrain Bak and inhibit Bak-dependent apoptosis, whereas Bcl-2 does not (10).Because the Bak protein contains multiple recognizable domains in addition to its BH3 motif (28, 29), we compared the binding of Bak BH3 peptide and Bak protein to Bcl-2. Surface plasmon resonance demonstrated that Bcl-2 binds Bak protein with much higher affinity than the Bak 26-mer BH3 peptide. Further experiments demonstrated that the KD for Bak differs among naturally occurring Bcl-2 sequence variants but is only 3-fold higher than that of Mcl-1 in some cases. In light of previous reports that Bcl-2 overexpression contributes to neoplastic transformation (3033) and drug resistance (3436) in lymphoid cells, we also examined Bcl-2 expression and Bak binding in a panel of neoplastic lymphoid cell lines. Results of these experiments demonstrated that Bcl-2 expression varies among different lymphoid cell lines but is up to 40-fold more abundant than Mcl-1. In lymphoid cell lines with abundant Bcl-2, Bak is detected in Bcl-2 as well as Mcl-1 immunoprecipitates; and Bak-dependent apoptosis induced by Mcl-1 down-regulation can be prevented by Bcl-2 overexpression. Collectively, these observations shed new light on the role of Bcl-2 in binding and neutralizing Bak.  相似文献   

6.
Pseudomonas exotoxin A (PE) is a bacterial toxin that arrests protein synthesis and induces apoptosis. Here, we utilized mouse embryo fibroblasts (MEFs) deficient in Bak and Bax to determine the roles of these proteins in cell death induced by PE. PE induced a rapid and dose-dependent induction of apoptosis in wild-type (WT) and Bax knockout (Bax−/−) MEFs but failed in Bak knockout (Bak−/−) and Bax/Bak double-knockout (DKO) MEFs. Also a loss of mitochondrial membrane potential was observed in WT and Bax−/− MEFs, but not in Bak−/− or in DKO MEFs, indicating an effect of PE on mitochondrial permeability. PE-mediated inhibition of protein synthesis was identical in all 4 cell lines, indicating that differences in killing were due to steps after the ADP-ribosylation of EF2. Mcl-1, but not Bcl-xL, was rapidly degraded after PE treatment, consistent with a role for Mcl-1 in the PE death pathway. Bak was associated with Mcl-1 and Bcl-xL in MEFs and uncoupled from suppressed complexes after PE treatment. Overexpression of Mcl-1 and Bcl-xL inhibited PE-induced MEF death. Our data suggest that Bak is the preferential mediator of PE-mediated apoptosis and that the rapid degradation of Mcl-1 unleashes Bak to activate apoptosis.Apoptosis is a mode of cell death utilized by multicellular organisms to remove unwanted cells. Also, many different cancer treatments, including chemotherapy and radiotherapy, induce apoptosis and result in the destruction of tumor cells. In some cases, apoptosis resistance can contribute to the failure of chemotherapy (14, 20, 24). Immunotoxins are a class of antitumor agents in which a powerful protein toxin is brought to the cancer cell by an antibody or an antibody fragment (for reviews, see references 28, 29, and 32). Several immunotoxins are currently in clinical trials, and one of these, BL22, targeting CD22, has shown excellent activity in drug-resistant hairy-cell leukemia (18, 19). Also, a fusion protein in which a fragment of diphtheria toxin is fused to the cytokine interleukin 2 (IL-2) (Ontak) is approved for the treatment of cutaneous T-cell lymphoma (26). Several studies carried out to determine how protein toxins and immunotoxins containing these toxins kill target cells have reported caspase activation (13, 16, 17, 30, 33). However, the steps leading up to caspase activation by these toxins that inhibit protein synthesis have not been elucidated.Bcl-2 family members are essential regulators of the mitochondrial (intrinsic) apoptosis pathway (1, 21). Proteins of this family have been divided into pro- and antiapoptotic proteins. Antiapoptotic proteins include the multi-Bcl-2 homology (BH) domain proteins Bcl-2, Bcl-xL, Bcl-w, Mcl-1, Bcl-b, and Bcl2a1. Proapoptotic members can be further classified into two subfamilies, the multi-BH domain Bax homologues, including Bax, Bak, and Bok, and the BH3-only proteins, including Nbk/Bik, Noxa, Hrk, Bad, Bim, Puma, and Bmf. Bax and Bak are the most extensively studied central mediators in the mitochondrial apoptosis pathway (4, 6). Various stimuli, including pathogens, toxic drugs, irradiation, and starvation, induce a conformational change and activation of Bak/Bax, usually via BH3-only proapoptosis proteins. This results in the disruption of mitochondrial membranes and the release of apoptotic factors, such as cytochrome c, SMAC, and apoptosis-inducing factor, which lead to the activation of effector caspases (5, 37, 40, 42, 43).The roles of Bax and Bak can be redundant or nonredundant, depending on the apoptotic stimuli. Bak and Bax can compensate for each other in apoptosis induced by staurosporine, etoposide, UV irradiation, serum deprivation, tBid, Bim, Bad, or Noxa (37, 43). Bak plays an essential role for apoptosis induced by Semliki Forest virus, gliotoxin, Bcl-xS, and vinblastine (22, 27, 34, 35), while Bax is favored for apoptosis induced by Nbk/Nik, a combination of tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) and ionizing irradiation, or TRAIL and 5-fluorouracil (5-FU) (9, 10, 36, 38). Silencing of either Bak or Bax resulted in resistance to apoptosis induced by Neisseria gonorrhoeae and cisplatin (15). Sometimes the same stimulus may result in different outcomes in different cell types. NBK/Bik mediated Bax-dependent cell death in one study (9), while in another study, NBK/Bik activated BAK-mediated apoptosis (31).In the current study, we utilized mutant mouse embryo fibroblasts (MEFs) deficient in Bak, Bax, or both proteins and provided evidence for an essential role of Bak in apoptosis induced by Pseudomonas exotoxin A (PE) and other protein synthesis inhibitors. We found that Bak−/− cells are resistant to killing by PE and that Mcl-1, which binds to Bak, controls apoptosis induced by PE.  相似文献   

7.
KS-Bcl-2, encoded by Kaposi''s sarcoma-associated herpesvirus (KSHV), is a structural and functional homologue of the Bcl-2 family of apoptosis regulators. Like several other Bcl-2 family members, KS-Bcl-2 protects cells from apoptosis and autophagy. Using a yeast two-hybrid screen and coimmunoprecipitation assays, we identified a novel KS-Bcl-2-interacting protein, referred to as protein interacting with carboxyl terminus 1 (PICT-1), encoded by a candidate tumor suppressor gene, GLTSCR2. Confocal laser scanning microscopy revealed nucleolar localization of PICT-1, whereas KS-Bcl-2 was located mostly at the mitochondrial membranes with a small fraction in the nucleoli. Ectopic expression of PICT-1 resulted in a large increase in the nucleolar fraction of KS-Bcl-2, and only a minor fraction remained in the cytoplasm. Furthermore, knockdown of endogenous PICT-1 abolished the nucleolar localization of KS-Bcl-2. However, ectopically expressed PICT-1 did not alter the cellular distribution of human Bcl-2. Subsequent analysis mapped the crucial amino acid sequences of both KS-Bcl-2 and PICT-1 required for their interaction and for KS-Bcl-2 targeting to the nucleolus. Functional studies suggest a correlation between nucleolar targeting of KS-Bcl-2 by PICT-1 and reduction of the antiapoptotic activity of KS-Bcl-2. Thus, these studies demonstrate a cellular mechanism to sequester KS-Bcl-2 from the mitochondria and to downregulate its virally encoded antiapoptotic activity. Additional characterization of the interaction of KS-Bcl-2 and PICT-1 is likely to shed light on the functions of both proteins.Kaposi''s sarcoma (KS)-associated herpesvirus (KSHV), also referred to as human herpesvirus 8 (HHV-8), is a gamma 2 herpesvirus implicated in several cancers, including KS, primary effusion lymphoma (PEL), and a subset of multicentric Castleman''s disease. Among human viruses, KSHV is most closely related to the Epstein-Barr virus (EBV), a tumorigenic gamma 1 herpesvirus known to be associated with lymphomas and nasopharyngeal carcinoma (10, 12).KSHV open reading frame 16 (orf16) encodes the KS-Bcl-2 protein, which shares sequence and functional homology with the Bcl-2 family (9, 31). Members of the Bcl-2 family are defined by the presence of up to four conserved domains known as the Bcl-2 homology (BH) domains. Several members also possess a carboxy-terminal transmembrane domain that mediates their association with intracellular membranes, such as the endoplasmic reticulum or mitochondria. Bcl-2 proteins are thought to serve primarily as cell death agonists or antagonists that integrate diverse survival and death signals, which are generated outside and within the cell (15, 37), yet Bcl-2 proteins also modulate cell cycle checkpoints, DNA repair/recombination pathways, calcium homeostasis, and cellular bioenergetics.All gammaherpesviruses encode Bcl-2 proteins that generally share 20 to 30% homology with one another and with their cellular counterparts (8, 11). The conservation of Bcl-2 homologues in these viruses indicates their importance for viral infection, with an evolutionarily conserved function of unknown nature. KS-Bcl-2, like most herpesvirus homologues of Bcl-2, contains a transmembrane domain and demonstrates conservation of sequences in both BH1 and BH2 but has only a low degree of homology with other regions of cellular Bcl-2 (18, 22). Still, KS-Bcl-2 shares 3-dimensional structural conservation with Bcl-2 family members and includes the conserved BH3 binding groove and a hydrophobic membrane anchor domain that also contains a mitochondrial outer membrane targeting signal (18). The BH3 binding cleft of KS-Bcl-2 binds with high affinity to peptides encoding BH3 domains present on the proapoptotic proteins Noxa, Bik, PUMA, Bak, Bax, Bid, Bim, and, to a much lesser extent, Bad (13, 18, 22). Based on these characteristics, KS-Bcl-2 has been suggested to have the closest resemblance to the cellular Bcl-2 family member Mcl-1 (13).Previous studies have demonstrated that KS-Bcl-2 protects various cell types from apoptosis mediated by the expression of BAX, tBid, or Bim through Sindbis virus infection or by ectopic expression of KSHV-cyclin-CDK6 (9, 13, 25, 31). However, unlike the cellular Bcl-2, KS-Bcl-2 is not a substrate for KSHV-cyclin-CDK6 phosphorylation (25) and cannot be converted into a proapoptotic protein via caspase cleavage (3). KS-Bcl-2 is able to form a stable complex with the cellular protein Aven, which binds Apaf-1 and is known as a regulator of caspase 9 and ataxia-telangiectasia (ATM) activation (7, 16). Like the cellular and other virus-encoded Bcl-2 proteins, KS-Bcl-2 binds Beclin and disrupts its lysosomal degradation pathway of autophagy (21, 29). However, since KS-Bcl-2 lacks the nonstructured loop located between the BH4 and BH3 domains, its binding to BH3-containing proapoptotic proteins and to the BH3-containing proautophagy protein Beclin is not modulated by phosphorylation (38).KS-Bcl-2 is transcribed during lytic virus infection (30, 31). Thus, inhibition of apoptosis and autophagy by KS-Bcl-2 may provide an attractive mechanism for prolonging the life span of KSHV-infected cells, which in turn enables increased virus production or establishment of latency. Whether the function of KS-Bcl-2 is necessary for KSHV-mediated oncogenesis is still unknown. Nevertheless, the KS-Bcl-2 protein is expressed in late-stage KS lesions but has not been detected in latent or in lytic KSHV-infected PEL cells (39).To explore the role of KS-Bcl-2 in cell signaling, we searched for its potential cellular-protein partners. In the present study, we describe a novel interaction between KS-Bcl-2 and the protein interacting with carboxyl terminus 1 (PICT-1) cellular protein, encoded by a candidate tumor suppressor gene, GLTSCR2. We show that this interaction specifically targets KS-Bcl-2 to the nucleolus and decreases its antiapoptotic activity.(Portions of this work were submitted to Bar Ilan Univeristy, Ramat Gan, Israel, by I. Kalt and T. Borodianskiy-Shteinberg in partial fulfillment of the requirements for the degree of Doctor of Philosophy.)  相似文献   

8.
Apoptosis is a potent immune barrier against viral infection, and many viruses, including poxviruses, encode proteins to overcome this defense. Interestingly, the avipoxviruses, which include fowlpox and canarypox virus, are the only poxviruses known to encode proteins with obvious Bcl-2 sequence homology. We previously characterized the fowlpox virus protein FPV039 as a Bcl-2-like antiapoptotic protein that inhibits apoptosis by interacting with and inactivating the proapoptotic cellular protein Bak. However, both Bak and Bax can independently trigger cell death. Thus, to effectively inhibit apoptosis, a number of viruses also inhibit Bax. Here we show that FPV039 inhibited apoptosis induced by Bax overexpression and prevented both the conformational activation of Bax and the subsequent formation of Bax oligomers at the mitochondria, two critical steps in the induction of apoptosis. Additionally, FPV039 interacted with activated Bax in the context of Bax overexpression and virus infection. Importantly, the ability of FPV039 to interact with active Bax and inhibit Bax activity was dependent on the structurally conserved BH3 domain of FPV039, even though this domain possesses little sequence homology to other BH3 domains. FPV039 also inhibited apoptosis induced by the BH3-only proteins, upstream activators of Bak and Bax, despite interacting detectably with only two: BimL and Bik. Collectively, our data suggest that FPV039 inhibits apoptosis by sequestering and inactivating multiple proapoptotic Bcl-2 proteins, including certain BH3-only proteins and both of the critical “gatekeepers” of apoptosis, Bak and Bax.Apoptosis is a highly conserved form of programmed cell death that plays an important role in the immune defense against pathogens. The controlled and deliberate destruction of virally infected cells comprises a potent innate immune barrier against rampant viral replication and infection. As such, many viruses, including poxviruses, encode numerous proteins that inhibit a variety of steps in the biochemical pathways that lead to cell death (29, 69).The mitochondria, and the Bcl-2 family of proteins that preside over them, serve as an important control point in the regulation of apoptosis (87). United by the presence of one to four highly conserved Bcl-2 homology (BH) domains, the Bcl-2 family regulates the integrity of the outer mitochondrial membrane (OMM) and controls the release of apoptogenic molecules from the mitochondrial intermembrane space. Bak and Bax, the two proapoptotic Bcl-2 proteins, possess BH domains 1 to 3 and, upon activation, commit the cell to death (53, 77). Whereas Bak resides constitutively at the OMM, Bax exists in an inactive form in the cytoplasm and, upon apoptotic insult, undergoes a conformational change that exposes its C-terminal transmembrane domain and results in its relocalization to the OMM (10, 34, 41, 56). The attendant exposure of the N termini of both Bak and Bax precedes Bak and Bax homooligomerization, which facilitates mitochondrial damage and, ultimately, the release of cytochrome c (3, 4, 36, 37, 76). Cytochrome c, in turn, triggers the activation of caspases, a group of cysteine proteases responsible for dismantling the apoptotic cell (59). Bak and Bax are therefore crucial for the induction of apoptosis and, because either Bak or Bax alone is sufficient to facilitate the release of cytochrome c, both must be inactivated to effectively inhibit apoptosis (53, 77, 90). The activation of Bak and Bax is counteracted by the antiapoptotic members of the Bcl-2 family, including Bcl-2, Bcl-XL, and Mcl-1. These three proteins, which possess all four BH domains, reside at the mitochondria and prevent apoptosis by directly interacting with and inhibiting Bak and Bax or the BH3-only proteins (87). The BH3-only proteins, which possess only the BH3 domain, act as sentinels responsive to a variety of cellular stresses, including virus infection (79). Upon receipt of an apoptotic stimulus, BH3-only proteins become activated and subsequently activate Bak and Bax or inhibit the antiapoptotic function of Bcl-2, Bcl-XL, and Mcl-1 (15). Of the eight BH3-only proteins that are directly involved in the induction of apoptosis—namely, Bim, Bid, Puma, Bik, Bmf, Bad, Noxa, and Hrk—each displays a specific and characteristic ability to bind and inhibit Bcl-2 proteins (79).Like cellular antiapoptotic Bcl-2 proteins, viral inhibitors of apoptosis have evolved especially to interfere with the activation of Bak and Bax (18, 40). For example, E1B 19K, encoded by adenovirus, and M11L, encoded by myxoma virus, bind and inactivate both Bak and Bax to inhibit apoptosis (26, 49, 65, 67, 72). Similarly, ORF125, the antiapoptotic protein encoded by the poxvirus Orf virus, also inactivates Bak and Bax, but exactly how ORF125 mediates this inactivation remains unknown (78). Although interacting with Bak and Bax is ostensibly the most direct way to prevent apoptosis, several viral antiapoptotic proteins appear to inhibit apoptosis by functioning upstream of Bak and Bax at the level of the BH3-only proteins. The vaccinia virus protein F1L, for example, interacts with Bak but not Bax, yet F1L is nonetheless capable of inactivating Bax, likely a result of F1L interacting with the BH3-only protein and Bax activator, Bim (61, 70, 74). Moreover, the Bcl-2 homolog encoded by Kaposi''s sarcoma-associated herpesvirus, and BHRF-1, encoded by Epstein-Barr virus, each interact with a specific and distinct array of BH3-only proteins, yet neither protein interacts detectably with Bak or Bax (14, 27, 44). Thus, to effectively inhibit apoptosis, it may not be necessary for viral proteins to directly target Bak and Bax but, instead, to prevent the activation of Bak and Bax by interfering with the upstream BH3-only proteins (15).Recently, our lab has shown that FPV039, encoded by fowlpox virus, localizes to the mitochondria, where it inhibits apoptosis induced by a variety of stimuli (6). Interestingly, FPV039 is the only characterized poxvirus protein that shares obvious, albeit limited, sequence homology with cellular Bcl-2 proteins (1, 6). FPV039 possesses a highly conserved BH1 and BH2 domain but lacks an obvious BH3 and BH4 domain. Importantly, however, we predicted structural homology between the Bcl-2 BH3 domain and a corresponding region in FPV039, and we validated the prediction by showing that this cryptic FPV039 BH3 domain is functionally important (6). Indeed, the ability of FPV039 to interact with the proapoptotic protein Bak is dependent on this cryptic BH3 domain (6). Thus, despite lacking sequence conservation of a highly conserved BH3 domain, FPV039 is able to interact with, and inactivate, the proapoptotic protein Bak. Nevertheless, to completely inhibit apoptosis, both Bak and Bax must be inactivated.Accordingly, we wanted to determine whether FPV039, in addition to inactivating Bak, could inactivate Bax. We report here that FPV039 inhibited Bax activity and prevented critical steps in Bax activation. FPV039 did not appear to interact with endogenous inactive Bax; however, FPV039 was able to interact with active Bax. Moreover, FPV039 inhibited apoptosis induced by the BH3-only proteins despite interacting with only BimL and Bik. Together, these data strongly suggest FPV039 inhibits apoptosis by inactivating multiple proapoptotic Bcl-2 proteins, including the critical Bak and Bax, as well as a discrete subset of BH3-only proteins.  相似文献   

9.
Simian varicella virus (SVV) causes varicella in primates, becomes latent in ganglionic neurons, and reactivates to produce zoster. SVV produces a cytopathic effect in monkey kidney cells in tissue culture. To study the mechanism by which SVV-infected cells die, we examined markers of apoptosis 24 to 64 h postinfection (hpi). Western blot analysis of virus-infected cell lysates revealed a significant increase in the levels of the cleaved active form of caspase-3, accompanied by a parallel increase in caspase-3 activity at 40 to 64 hpi. Caspase-9, a marker for the intrinsic pathway, was activated significantly in SVV-infected cells at all time points, whereas trace levels of the active form of caspase-8, an extrinsic pathway marker, was detected only at 64 hpi. Bcl-2 expression at the mRNA and protein levels was decreased by 50 to 70% throughout the course of virus infection. Release of cytochrome c, an activator of caspase-9, from mitochondria into the cytoplasm was increased by 200% at 64 hpi. Analysis of Vero cells infected with SVV expressing green fluorescent protein (SVV-GFP) at 64 hpi revealed colocalization of the active forms of caspase-3 and caspase-9 and terminal deoxynucleotidyltransferase-mediated dUTP-biotin nick end labeling (TUNEL) staining with GFP. A significant decrease in the bcl-2 mRNA levels along with an abundance of mRNA specific for SVV genes 63, 40, and 21 was seen in the fraction of Vero cells that were infected with SVV-GFP. Together, these findings indicate that SVV induces apoptosis in cultured Vero cells through the intrinsic pathway in which Bcl-2 is downregulated.Apoptosis, a regulated form of cell death, plays a critical role in the homeostasis of multicellular organisms. Key features include membrane blebbing, chromatin condensation, and cell shrinkage. UV irradiation, deprivation of growth factors, and viral infection all cause apoptosis in cultured cells. Apoptosis is triggered by sequential activation of a group of cysteine proteases known as caspases. Apoptosis proceeds primarily through two pathways. The extrinsic pathway involves activation of caspase-8 and is initiated by ligand interaction with Fas or death receptors, while the intrinsic pathway is activated by an imbalance between proapoptotic (e.g., Bad and Bax) and antiapoptotic (e.g., Bcl-2 and Bcl-xL) proteins in mitochondria (21), resulting in release of cytochrome c from mitochondria, which in turn activates caspase-9. Bcl-2 plays an important role in cell survival (22, 32). Both caspase-8 and caspase-9 activate caspase-3, which along with other effector caspases, cleave critical cellular proteins, resulting in apoptosis.Simian varicella virus (SVV), the primate counterpart of human varicella zoster virus (VZV), produces a naturally occurring exanthematous disease that mimics human varicella (9, 18). Clinical and pathological changes produced by SVV infection of primates are similar to those produced by human varicella, and both VZV and SVV reactivate from latently infected ganglionic neurons (4, 13, 23, 33). The SVV and VZV genomes share a high degree of nucleotide homology (3, 10), and SVV-specific antibodies cross-react with human VZV in serum neutralization and complement fixation tests (5, 6, 30). Both viruses produce a cytopathic effect in monkey kidney cells in tissue culture (2, 29, 31). VZV has been shown to cause apoptosis in cultured Vero cells, human foreskin fibroblasts, and peripheral blood mononuclear cells isolated from healthy donors but not in primary human dorsal root ganglionic neurons (12, 13, 16, 28). Apoptosis is also seen in peripheral blood mononuclear cells of children infected with VZV in vivo (25). Thus, VZV-induced apoptosis may be cell type specific. The main objectives of this study were to determine if SVV induces apoptosis in cultured Vero cells, a monkey kidney cell line, and to identify the specific pathways.  相似文献   

10.
Small cell lung cancer (SCLC) has an annual mortality approaching that of breast and prostate cancer. Although sensitive to initial chemotherapy, SCLC rapidly develops resistance, leading to less effective second-line therapies. SCLC cells often overexpress Bcl-2, which protects cells from apoptosis both by sequestering pro-apoptotic family members and by modulating inositol 1,4,5-trisphosphate receptor (IP3R)-mediated calcium signaling. BH3-mimetic agents such as ABT-263 disrupt the former activity but have limited activity in SCLC patients. Here we report for the first time that Bcl-2-IP3 receptor disruptor-2 (BIRD-2), a decoy peptide that binds to the BH4 domain of Bcl-2 and prevents Bcl-2 interaction with IP3Rs, induces cell death in a wide range of SCLC lines, including ABT-263-resistant lines. BIRD-2-induced death of SCLC cells appears to be a form of caspase-independent apoptosis mediated by calpain activation. By targeting different regions of the Bcl-2 protein and different mechanisms of action, BIRD-2 and ABT-263 induce cell death synergistically. Based on these findings, we propose that targeting the Bcl-2–IP3R interaction be pursued as a novel therapeutic strategy for SCLC, either by developing BIRD-2 itself as a therapeutic agent or by developing small-molecule inhibitors that mimic BIRD-2.Lung cancer accounts for 12% of all new cancers worldwide and is a leading cause of cancer-related mortality in the United States.1, 2, 3 Although small cell lung cancer (SCLC) comprises only 15% of lung cancer cases,2, 3 it has an annual mortality rate approaching that of breast and prostate cancer.4 Compared with the more common non-small cell lung cancer (NSCLC), SCLC is more aggressive and associated with rapid development of metastasis.2 Moreover, although SCLC is more responsive to chemotherapy and radiation therapy initially, it typically relapses quickly with treatment-resistant disease.2 In contrast to dramatic advances in chemotherapy and personalized medicine in other malignancies, the life expectancy of SCLC patients has remained <2 years for decades and is <1 year for patients with extensive disease.5, 6 The lethality of SCLC is attributed in part to the development of resistance to standard combination chemotherapies, underscoring the need to develop novel therapeutic approaches based on understanding the molecular and cellular biology of SCLC.5, 6Evasion from apoptosis is a major hallmark of cancer and a prominent factor underlying drug resistance in SCLC.3 Multiple mechanisms contribute to apoptosis resistance in SCLC, including elevated expression of the antiapoptotic Bcl-2 protein3 (Supplementary Figure S1). Tsujimoto and colleagues discovered elevated levels of Bcl-2 mRNA and protein in SCLC cells not long after their identification of Bcl-2 as the protein product of the bcl-2 gene in follicular lymphoma.7, 8 Subsequently, immunohistochemistry of 164 primary SCLC samples revealed 76% were positive for Bcl-2, a finding substantiated by microarray detection of increased BCL-2 mRNA levels in 84% of SCLC samples9, 10 and by genomic sequencing of circulating SCLC tumor cells.11 Moreover, proteomic profiling documented that Bcl-2 is more highly expressed in SCLC than in NSCLC, reflecting the vastly different biology of these lung cancer subtypes.12The major known function of Bcl-2 is to bind and sequester BH3-only proteins such as Bim, preventing these proteins from inducing apoptosis.13, 14, 15 Therefore, a major investment has been made in targeting this interaction for cancer treatment. The interaction takes place in a hydrophobic groove on Bcl-2 and the therapeutic strategy for targeting this interaction has been to develop small molecules, BH3-mimetic agents, which bind in the hydrophobic groove and induce apoptosis by displacing the BH3-only proteins. This approach has been reviewed in detail.14, 15, 16Among BH3-mimetic agents advancing through clinical trials for both hematological malignancies15, 17 and solid tumors18 are ABT-737 and its orally bioavailable derivative ABT-263 (Navitoclax). Reported studies of ABT-199, a selective inhibitor of Bcl-2, are at present limited to hematological malignancies.18 In screening a large number of cancer cell lines, the pioneering work of Oltersdorf et al.19 demonstrated potent single-agent activity of ABT-737 against cell lines representative of lymphoid malignancies and SCLC. Clinical trials of ABT-263, an orally bioavailable version of ABT-737, achieved overall response rates ranging from as high as 35% in relapsed/refractory chronic lymphocytic leukemia (CLL) and 22% in follicular lymphoma.17 Reported responses are generally less in solid tumors with the notable exception of SCLC.18 But even in SCLC, activity of ABT-263 is limited in comparison to hematological malignancies, with 1 of the 39 (3%) of patients achieving a partial response to ABT-263 and 9 of the 37 (23%) achieving stable disease in a phase I clinical trial.20 This experience suggests a need to develop additional ways of targeting Bcl-2 for cancer treatment.A potential alternative therapeutic target for Bcl-2-positive malignancies involves interaction of Bcl-2 with the inositol 1,4,5-trisphosphate receptor (IP3R), an IP3-gated Ca2+ channel located on the endoplasmic reticulum (ER). Bcl-2 is located not only on the outer mitochondrial membrane but also on the ER, and at both of these locations, it functions as a potent inhibitor of apoptosis.21, 22, 23 ER-localized Bcl-2 interacts with IP3Rs and inhibits apoptosis by preventing excessive IP3R-mediated Ca2+ transfer from the ER lumen into the cytoplasm and nearby mitochondria.24, 25, 26 Notably, regions of Bcl-2 involved in binding BH3-only proteins and IP3Rs are entirely different. Whereas BH3-only proteins and their BH3-mimetic counterparts bind in a hydrophobic groove composed of BH3 domains 1–3 of Bcl-2,13, 14 the BH4 domain of Bcl-2 is necessary for interaction with IP3Rs.27 To develop a peptide inhibitor of Bcl-2–IP3R interaction, we identified the Bcl-2-binding region on the IP3R and developed a small synthetic 20 amino-acid peptide corresponding to this region.28 This peptide, when fused to the cell-penetrating peptide of HIV TAT, binds to the BH4 domain of Bcl-2 and functions as a decoy peptide, inhibiting Bcl-2–IP3R interaction.29, 30 We currently refer to this peptide as BIRD-2 (Bcl-2-IP3 Receptor Disruptor-2), having formerly named it TAT-IDPDD/AA.31 By disrupting the Bcl-2–IP3R interaction, BIRD-2 abrogates Bcl-2 control over IP3R-mediated Ca2+ elevation and induces Ca2+-mediated apoptosis in primary human CLL cells29 and diffuse large B-cell lymphoma cells.32 Notably, BIRD-2 does not kill normal cells, including human lymphocytes isolated from peripheral blood29 and normal murine embryonic fibroblasts (F Zhong and C Distelhorst, unpublished data).The present investigation was undertaken to determine whether Bcl-2–IP3R interaction is a potentially useful therapeutic target in SCLC. In support of this concept, we find the majority of SCLC lines tested are sensitive to BIRD-2-induced apoptosis and that BIRD-2 induces apoptosis in several ABT-263-resistant SCLC lines. BIRD-2, we find, lacks generalized cytotoxicity as it does not induce cell death in NSCLC lines or a normal lung epithelial line. On the other hand, we find that BIRD-2 and ABT-263 synergize in killing SCLC cells. These findings for the first time provide preclinical evidence of the potential value of targeting both antiapoptotic mechanisms of Bcl-2 for the treatment of SCLC.  相似文献   

11.
12.
13.
Bcl2-modifying factor (Bmf) is a member of the BH3-only group of proapoptotic proteins. To test the role of Bmf in vivo, we constructed mice with a series of mutated Bmf alleles that disrupt Bmf expression, prevent Bmf phosphorylation by the c-Jun NH2-terminal kinase (JNK) on Ser74, or mimic Bmf phosphorylation on Ser74. We report that the loss of Bmf causes defects in uterovaginal development, including an imperforate vagina and hydrometrocolpos. We also show that the phosphorylation of Bmf on Ser74 can contribute to a moderate increase in levels of Bmf activity. Studies of compound mutants with the related gene Bim demonstrated that Bim and Bmf exhibit partially redundant functions in vivo. Thus, developmental ablation of interdigital webbing on mouse paws and normal lymphocyte homeostasis require the cooperative activity of Bim and Bmf.Bmf is a proapoptotic BH3-only member of the Bcl2-related protein family that is implicated in cell death caused by anoikis (23, 26, 27), arsenic trioxide (19), histone deacetylase inhibitors (33, 34), transforming growth factor β (24), and tumor necrosis factor alpha (8). Mice with a loss of Bmf expression exhibit B-cell hyperplasia and increased sensitivity to γ-radiation-induced B-cell lymphoma (14). These observations indicate that Bmf represents an important mediator of cell death signaling pathways.The structure of Bmf includes a BH3 domain that is essential for apoptosis induction. In addition, Bmf contains a sequence motif that is required for interactions with dynein light chain 2 (DLC2), a component of the myosin V motor complex (23). The interaction of Bmf with DLC2 is required for the recruitment of Bmf to the cytoskeleton. The release of Bmf from complexes sequestered on the cytoskeleton may contribute to anoikis (23). Interestingly, this regulatory mechanism is shared by the related proapoptotic BH3-only protein Bim, which interacts via a similar sequence motif with dynein light chain 1 (DLC1), a component of the dynein motor complex (22).The similarities between Bmf and Bim include the presence of a conserved phosphorylation site (Bmf Ser74 and Bim Thr112) that is a substrate for the c-Jun NH2-terminal kinase (JNK) (15). Data from biochemical studies indicate that the JNK-mediated phosphorylation of Bmf and Bim may increase apoptotic activity (15). Indeed, mice with a germ line point mutation in the Bim gene (Thr112 replaced with Ala) exhibit decreased apoptosis (10). These studies indicate that Bmf and Bim may mediate, in part, proapoptotic signaling by JNK (3, 30).The purpose of this study was to examine the role of Bmf using mouse models with germ line defects in the Bmf gene, including mice with Bmf alleles that disrupt Bmf expression, prevent Bmf phosphorylation, or mimic Bmf phosphorylation. We examined the effects of these mutations in mice with both wild-type and mutant alleles of the related gene Bim. The results of our analysis demonstrate that Bmf and Bim exhibit partially redundant functions, that phosphorylation on Ser74 is not essential for Bmf activity, and that phosphorylation on Ser74 can contribute to increased levels of Bmf activity in vivo.  相似文献   

14.
The ectopic overexpression of Bcl-2 restricts both influenza A virus-induced apoptosis and influenza A virus replication in MDCK cells, thus suggesting a role for Bcl-2 family members during infection. Here we report that influenza A virus cannot establish an apoptotic response without functional Bax, a downstream target of Bcl-2, and that both Bax and Bak are directly involved in influenza A virus replication and virus-induced cell death. Bak is substantially downregulated during influenza A virus infection in MDCK cells, and the knockout of Bak in mouse embryonic fibroblasts yields a dramatic rise in the rate of apoptotic death and a corresponding increase in levels of virus replication, suggesting that Bak suppresses both apoptosis and the replication of virus and that the virus suppresses Bak. Bax, however, is activated and translocates from the cytosol to the mitochondria; this activation is required for the efficient induction of apoptosis and virus replication. The knockout of Bax in mouse embryonic fibroblasts blocks the induction of apoptosis, restricts the infection-mediated activation of executioner caspases, and inhibits virus propagation. Bax knockout cells still die but by an alternative death pathway displaying characteristics of autophagy, similarly to our previous observation that influenza A virus infection in the presence of a pancaspase inhibitor leads to an increase in levels of autophagy. The knockout of Bax causes a retention of influenza A virus NP within the nucleus. We conclude that the cell and virus struggle to control apoptosis and autophagy, as appropriately timed apoptosis is important for the replication of influenza A virus.The pathology of influenza A virus infection usually arises from acute lymphopenia and inflammation of the lungs and airway columnar epithelial cells (23, 38). Influenza A virus induces apoptotic death in infected epithelial, lymphocyte, and phagocytic cells, and apoptosis is a source of tissue damage during infection (3, 22, 33) and increased susceptibility to bacterial pathogens postinfection (31). While the induction of apoptosis by influenza A virus has been well documented (4, 19-21, 28, 33, 37), the mechanisms of this interaction are not well understood. Two viral proteins, NS1 and PB1-F2, have been associated with viral killing of cells. NS1, originally characterized as being proapoptotic (34), was later identified as being an interferon antagonist, inhibiting the activation of several key antiviral responses and restricting the apoptotic response to infection (1, 10, 15, 18, 35, 39, 46). In contrast, PB1-F2 induces apoptosis primarily by localizing to the outer mitochondrial membrane, promoting cytochrome c release, and triggering the apoptotic cascade (43). This effect, however, is typically restricted to infected monocytes, leading to the hypothesis that PB1-F2 induces apoptosis specifically to clear the landscape of immune responders (5, 44). Although PB1-F2 activity does not directly manipulate virus replication or virus-induced apoptosis, PB1-F2 localization to the mitochondrial membrane during infection potentiates the apoptotic response in epithelial and fibroblastic cells through tBID signaling with proapoptotic Bcl-2 family protein members Bax and Bak (22, 43, 44).The Bcl-2 protein family consists of both pro- and antiapoptotic members that regulate cytochrome c release during mitochondrion-mediated apoptosis through the formation of pore-like channels in the outer mitochondrial membrane (12, 16). During the initiation of mitochondrion-mediated apoptosis, cytoplasmic Bid is cleaved to form tBID. This, in turn, activates proapoptotic Bax and Bak (40), which drive cytochrome c release and subsequent caspase activation. Bak is constitutively associated with the mitochondrial membrane, whereas inactive Bax is primarily cytosolic, translocating to the outer mitochondrial membrane only after activation (6). The activation of Bax and Bak results in homo- and heterodimer formation at the outer mitochondrial membrane, generating pores that facilitate mitochondrial membrane permeabilization and cytochrome c release (14, 17), leading to caspase activation and the apoptotic cascade (8). Antiapoptotic members of the Bcl-2 protein family, including Bcl-2, inhibit the activation of proapoptotic Bax and Bak primarily by sequestering inactive Bax and Bak monomers via interactions between their BH3 homology domains (7).Bcl-2 expression has been linked to decreased viral replication rates (26). Bcl-2 overexpression inhibits influenza A virus-induced cell death and reduces the titer and spread of newly formed virions (29). The activation of caspase-3 in the absence of sufficient Bcl-2 is critical to the influenza A virus life cycle. Both Bcl-2 expression and the lack of caspase activation during infection lead to the nuclear accumulation of influenza virus ribonucleoprotein (RNP) complexes, thereby leading to the improper assembly of progeny virions and a marked reduction in titers of infectious virus (26, 41, 42, 45).Here we show that influenza A virus induces mitochondrion-mediated (intrinsic-pathway) apoptosis signaled specifically through Bax and that this Bax signaling is essential for the maximum efficiency of virus propagation. In contrast, Bak expression is strongly downregulated during infection. Cells lacking Bak (while expressing Bax) display a much more severe apoptotic phenotype in response to infection and produce infectious virions at a higher rate than the wild type (WT), suggesting that Bak, which can suppress viral replication, is potentially downregulated by the virus. Our results indicate essential and opposing roles for Bax and Bak in both the response of cells to influenza A virus infection and the ability of the virus to maximize its own replicative potential.  相似文献   

15.
The filovirus VP40 protein is capable of budding from mammalian cells in the form of virus-like particles (VLPs) that are morphologically indistinguishable from infectious virions. Ebola virus VP40 (eVP40) contains well-characterized overlapping L domains, which play a key role in mediating efficient virus egress. L domains represent only one component required for efficient budding and, therefore, there is a need to identify and characterize additional domains important for VP40 function. We demonstrate here that the 96LPLGVA101 sequence of eVP40 and the corresponding 84LPLGIM89 sequence of Marburg virus VP40 (mVP40) are critical for efficient release of VP40 VLPs. Indeed, deletion of these motifs essentially abolished the ability of eVP40 and mVP40 to bud as VLPs. To address the mechanism by which the 96LPLGVA101 motif of eVP40 contributes to egress, a series of point mutations were introduced into this motif. These mutants were then compared to the eVP40 wild type in a VLP budding assay to assess budding competency. Confocal microscopy and gel filtration analyses were performed to assess their pattern of intracellular localization and ability to oligomerize, respectively. Our results show that mutations disrupting the 96LPLGVA101 motif resulted in both altered patterns of intracellular localization and self-assembly compared to wild-type controls. Interestingly, coexpression of either Ebola virus GP-WT or mVP40-WT with eVP40-ΔLPLGVA failed to rescue the budding defective eVP40-ΔLPLGVA mutant into VLPs; however, coexpression of eVP40-WT with mVP40-ΔLPLGIM successfully rescued budding of mVP40-ΔLPLGIM into VLPs at mVP40-WT levels. In sum, our findings implicate the LPLGVA and LPLGIM motifs of eVP40 and mVP40, respectively, as being important for VP40 structure/stability and budding.Ebola and Marburg viruses are members of the family Filoviridae. Filoviruses are filamentous, negative-sense, single-stranded RNA viruses that cause lethal hemorrhagic fevers in both humans and nonhuman primates (5). Filoviruses encode seven viral proteins including: NP (major nucleoprotein), VP35 (phosphoprotein), VP40 (matrix protein), GP (glycoprotein), VP30 (minor nucleoprotein), VP24 (secondary matrix protein), and L (RNA-dependent RNA polymerase) (2, 5, 10, 12, 45). Numerous studies have shown that expression of Ebola virus VP40 (eVP40) alone in mammalian cells leads to the production of virus-like particles (VLPs) with filamentous morphology which is indistinguishable from infectious Ebola virus particles (12, 17, 18, 25, 26, 27, 30, 31, 34, 49). Like many enveloped viruses such as rhabdovirus (11) and arenaviruses (44), Ebola virus encodes late-assembly or L domains, which are sequences required for the membrane fission event that separates viral and cellular membranes to release nascent virion particles (1, 5, 7, 10, 12, 18, 25, 27, 34). Thus far, four classes of L domains have been identified which were defined by their conserved amino acid core sequences: the Pro-Thr/Ser-Ala-Pro (PT/SAP) motif (25, 27), the Pro-Pro-x-Tyr (PPxY) motif (11, 12, 18, 19, 41, 53), the Tyr-x-x-Leu (YxxL) motif (3, 15, 27, 37), and the Phe-Pro-Ile-Val (FPIV) motif (39). Both PTAP and the PPxY motifs are essential for efficient particle release for eVP40 (25, 27, 48, 49), whereas mVP40 contains only a PPxY motif. L domains are believed to act as docking sites for the recruitment of cellular proteins involved in endocytic trafficking and multivesicular body biogenesis to facilitate virus-cell separation (8, 13, 14, 16, 28, 29, 33, 36, 43, 50, 51).In addition to L domains, oligomerization, and plasma-membrane localization of VP40 are two functions of the protein that are critical for efficient budding of VLPs and virions. Specific sequences involved in self-assembly and membrane localization have yet to be defined precisely. However, recent reports have attempted to identify regions of VP40 that are important for its overall function in assembly and budding. For example, the amino acid region 212KLR214 located at the C-terminal region was found to be important for efficient release of eVP40 VLPs, with Leu213 being the most critical (30). Mutation of the 212KLR214 region resulted in altered patterns of cellular localization and oligomerization of eVP40 compared to those of the wild-type genotype (30). In addition, the proline at position 53 was also implicated as being essential for eVP40 VLP release and plasma-membrane localization (54).In a more recent study, a YPLGVG motif within the M protein of Nipah virus (NiV) was shown to be important for stability, membrane binding, and budding of NiV VLPs (35). Whether this NiV M motif represents a new class of L domain remains to be determined. However, it is clear that this YPLGVG motif of NiV M is important for budding, perhaps involving a novel mechanism (35). Our rationale for investigating the corresponding, conserved motifs present within the Ebola and Marburg virus VP40 proteins was based primarily on these findings with NiV. In addition, Ebola virus VP40 motif maps close to the hinge region separating the N- and C-terminal domains of VP40 (4). Thus, the 96LPLGVA101 motif of eVP40 is predicted to be important for the overall stability and function of VP40 during egress. Findings presented here indicate that disruption of these filovirus VP40 motifs results in a severe defect in VLP budding, due in part to impairment in overall VP40 structure, stability and/or intracellular localization.  相似文献   

16.
To assess interchangeability of estimates of bacterial abundance by different epifluorescence microscopy methods, total bacterial numbers (TBNs) determined by most widely accepted protocols were statistically compared. Bacteria in a set of distinctive samples were stained with acridine orange (AO), 4′-6-diamidino-2-phenylindole (DAPI), and BacLight and enumerated by visual counting (VC) and supervised image analysis (IA). Model II regression and Bland-Altman analysis proved general agreements between IA and VC methods, although IA counts tended to be lower than VC counts by 7% on a logarithmic scale. Distributions of cells and latex beads on polycarbonate filters were best fitted to negative binomial models rather than to Poisson or log-normal models. The fitted models revealed higher precisions of TBNs by the IA method than those by the VC method. In pairwise comparisons of the staining methods, TBNs by AO and BacLight staining showed good agreement with each other, but DAPI staining had tendencies of underestimation. Although precisions of the three staining methods were comparable to one another (intraclass correlation coefficients, 0.97 to 0.98), accuracy of the DAPI staining method was rebutted by disproportionateness of TBNs between pairs of samples that carried 2-fold different volumes of identical cell suspensions. It was concluded that the TBN values estimated by AO and BacLight staining are relatively accurate and interchangeable for quantitative interpretation and that IA provides better precision than does VC. As a prudent measure, it is suggested to avoid use of DAPI staining for comparative studies investigating accuracy of novel cell-counting methods.Bacterial abundance is an instrumental parameter in assessing the roles of bacteria in the environments (18, 27, 30, 45). While a variety of techniques are available (1, 30, 53, 60), staining bacterial cells with acridine orange (AO) (29) or 4′,6-diamidino-2-phenylindole (DAPI) (48) and counting them on black polycarbonate (PC) filters by epifluorescence microscopy have become the standard procedure for direct counting (9, 18, 30). The Live/Dead BacLight staining kit, which is widely accepted as a rapid measure of viability of individual cells, also provides a total count of bacteria (10). Currently, most studies reporting total bacterial numbers (TBNs) use one of the three staining methods described above. However, the basic question of which fluorochrome to use for a given samples still presents challenges, as comparative studies using two or more of these fluorochromes have often yielded conflicting results (10, 17, 20, 34, 37, 40, 49, 52, 54, 57, 58).A more perplexing question is whether TBN values based on different fluorochromes are interchangeable for a quantitative interpretation incorporating TBN data from different methods. A large-scale intersystem study, an analysis of long-term collection of longitudinal data, or a collaborative study by multiple laboratories often requires an amalgamated use of TBN values from different fluorochromes. Apart from the interchangeability of fluorochromes, there is another complication at the step of cell enumeration. For example, TBN estimates by digital image analysis (IA) on microscope fields were often either slightly higher (3, 44) or significantly lower (25) than those found by visual counting (VC). With the introduction of various instrument-aided enumeration methods, including photomicrography IA (43, 55, 59), laser-scanning microscopy (8, 36), flow cytometry (2, 27, 34), and microfluidic devices (1, 53), TBN values are now reported based on various combinations of fluorochromes and enumeration methods. Considering the rapid advancement of novel enumeration technologies, establishing a robust “gold standard” method that can estimate bacterial abundance with high accuracy and precision is more in demand than ever.However, the robust gold standard that can validate novel methods and calibrate different methods apparently does not exist yet, largely due to insufficient attention to random errors and biases involved with fluorochromes or enumeration methods (9, 30). In the studies reporting general agreement among TBN methods (22, 34, 41, 44, 53, 59), using correlation or ordinary linear regression as the only or major evidence of agreement appears to be a major analytical drawback. Since measurements under comparison are from the same quantity, i.e., the true value, intrinsic correlation is naturally expected. Therefore, analytical approaches based on correlation are biased toward finding an agreement (7), and hence, the strength of agreement cannot be objectively quantified. In cases reporting discrepancies between different TBN methods (17, 25, 35, 43, 48, 54, 57, 58), sources of biases were not identified due to the limitation of knowledge on the true abundance values or lack of estimation of precisions of methods. Error propagations of TBN methods were analyzed by several studies (13, 23, 32, 39) but have been limited to identification of sources of error for a specific method (35, 36), instead of comparing precisions and accuracies of commonly used TBN methods. Therefore, a comprehensive statistical study to reveal the intrinsic nature of the errors and biases of conventional TBN methods is necessary to establish the robust gold standard method for determining TBNs. In essence, the statistical study should compare different combinations of staining and enumeration methods that are used as the standard method for calibration of novel TBN methods or those that are most widely used for TBN estimation, either to establish a robust gold standard method for TBN estimation or to understand differences in TBN values reported in the literature.In this study, we performed intensive analyses on accuracies and precisions of the conventional TBN methods and determined agreements among their measurements. For the fluorochromes, the three most-used fluorochromes (AO, DAPI, and BacLight) were compared. For the enumeration methods, we employed VC, which is the traditional gold standard method for enumeration of bacteria, and a simple supervised IA method as a representative, using photographic images by imaging instruments. In comparison to other novel instrument-aided enumeration methods, these two methods could validate the objects being enumerated by human decision. Therefore, they had best potential as a part of the gold standard for TBN estimation. In many studies, these methods were implicitly regarded as the gold standard method in estimation of bacterial abundance. We applied Bland-Altman analysis (5) to quantify difference of measurements, characterized intrinsic errors of count data by generalized linear models (64), and determined accuracies of methods based on the confidence interval (CI) of ratios of average cell counts by a generalized pivotal approach (15). Based on these statistical properties of the methods, we identified biases intrinsic to each method and addressed which methods are accurate and interchangeable.  相似文献   

17.
18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号