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 62 毫秒
1.
Chemoresistance in cancer has previously been attributed to gene mutations or deficiencies. Bax or p53 deficiency can lead to resistance to cancer drugs. We aimed to find an agent to overcome chemoresistance induced by Bax or p53 deficiency. Here, we used immunoblot, flow-cytometry analysis, gene interference, etc. to show that genistein, a major component of isoflavone that is known to have anti-tumor activities in a variety of models, induces Bax/p53-independent cell death in HCT116 Bax knockout (KO), HCT116 p53 KO, DU145 Bax KO, or DU145 p53 KO cells that express wild-type (WT) Bak. Bak knockdown (KD) only partially attenuated genistein-induced apoptosis. Further results indicated that the release of AIF and endoG also contributes to genistein-induced cell death, which is independent of Bak activation. Conversely, AIF and endoG knockdown had little effect on Bak activation. Knockdown of either AIF or endoG alone could not efficiently inhibit apoptosis in cells treated with genistein, whereas an AIF, endoG, and Bak triple knockdown almost completely attenuated apoptosis. Next, we found that the Akt-Bid pathway mediates Bak-induced caspase-dependent and AIF- and endoG-induced caspase-independent cell death. Moreover, downstream caspase-3 could enhance the release of AIF and endoG as well as Bak activation via a positive feedback loop. Taken together, our data elaborate the detailed mechanisms of genistein in Bax/p53-independent apoptosis and indicate that caspase-3-enhanced Bid activation initiates the cell death pathway. Our results also suggest that genistein may be an effective agent for overcoming chemoresistance in cancers with dysfunctional Bax and p53.Mammalian cell death proceeds through a highly regulated program called apoptosis that is highly dependent on the mitochondria.1 Mitochondrial outer membrane (MOM) multiple apoptotic stresses permeabilize the MOM, resulting in the release of apoptogenic factors including cytochrome c, Smac, AIF, and endoG.2, 3, 4 Released cytochrome c activates Apaf-1, which assists in caspase activation. Then, activated caspases cleave cellular proteins and contribute to the morphological and biochemical changes associated with apoptosis. Bcl-2 family proteins control a crucial apoptosis checkpoint in the mitochondria.2, 5, 6, 7 Multidomain proapoptotic Bax and Bak are essential effectors responsible for the permeabilization of the MOM, whereas anti-apoptotic Bcl-2, Bcl-xL, and Mcl-1 preserve mitochondrial integrity and prevent cytochrome c efflux triggered by apoptotic stimuli. The third Bcl-2 subfamily of proteins, BH3-only molecules (BH3s), promotes apoptosis by either activating Bax/Bak or inactivating Bcl-2/Bcl-xL/Mcl-1.8, 9, 10, 11, 12 Upon apoptosis, the ‘activator'' BH3s, including truncated Bid (tBid), Bim, and Puma, activate Bax and Bak to mediate cytochrome c efflux, leading to caspase activation.8, 11, 12 Conversely, antiapoptotic Bcl-2, Bcl-xL, and Mcl-1 sequester activator BH3s into inert complexes, which prevents Bax/Bak activation.8, 9 Although it has been proposed that Bax and Bak activation occurs by default as long as all of the anti-apoptotic Bcl-2 proteins are neutralized by BH3s,13 liposome studies clearly recapitulate the direct activation model in which tBid or BH3 domain peptides derived from Bid or Bim induce Bax or Bak oligomerization and membrane permeabilization.12, 14, 15Numerous studies have demonstrated a critical role for Bax in determining tumor cell sensitivity to drug induction and in tumor development. Bax has been reported to be mutated in colon16, 17 and prostate cancers,18, 19 contributing to tumor cell survival and promoting clonal expansion. Bax has been shown to restrain tumorigenesis20 and is necessary for tBid-induced cancer cell apoptosis.21 Loss of Bax has been reported to promote tumor development in animal models.22 Bax knockout (KO) renders HCT116 cells resistant to a series of apoptosis inducers.23, 24, 25 p53 has been reported to be a tumor suppressor,26 and its mutant can cause chemoresistance in cancer cells.27, 28, 29 Moreover, p53 is often inactivated in solid tumors via deletions or point mutations.30, 31 Thus, it is necessary to find an efficient approach or agent to overcome chemoresistance caused by Bax and/or p53 mutants.Few studies have focused on the role of Bak in tumor cell apoptosis and cancer development. Bak mutations have only been shown in gastric and colon cancer cells.32 Some studies have revealed that Bak is a determinant of cancer cell apoptosis.33, 34 Some studies have even demonstrated that Bak renders Bax KO cells sensitive to drug induction.33, 35 In this study, we are the first group to show that tBid induces Bak activation and the release of AIF and endoG in colon cancer cells, which causes cellular apoptosis independent of Bax/p53. We also found that caspase-3 is activated in apoptosis. Interestingly, downstream caspase-3 can strengthen Bak activation and the release of AIF and endoG during apoptosis via a feedback loop. Furthermore, we reveal that Akt upregulates apoptosis progression. These results will help us to better understand the function of mitochondrial apoptotic protein members in apoptosis and cancer therapies. Furthermore, our experiments may provide a theoretical basis for overcoming chemoresistance in cancer cells.  相似文献   

2.
3.
Tumour progression and therapy resistance in squamous cell carcinoma of the skin (SCC) is strongly associated with resistance to intrinsic mitochondrial apoptosis. We thus investigated the role of various anti-apoptotic Bcl-2 proteins for apoptosis protection in SCC using the BH3 agonist ABT737 that can overcome multidomain Bcl-2 protein protection. Sensitive SCC cells underwent rapid loss of mitochondrial membrane potential (MMP), subsequent apoptosis concomitant with caspase-3 activation and an early release of mitochondria-derived cytochrome c and smac/DIABLO. In contrast, ABT737 resistance in subsets of SCC cells was not explained by XIAP, important for protection from DR-induced apoptosis in SCC. Of note, ABT737 did not prime SCC cells to DR-induced apoptosis. Interestingly, the ratio of Mcl-1 and Noxa determined sensitivity to ABT737: loss of Mcl-1 rendered resistant cells sensitive to ABT737, whereas loss of Noxa promoted resistance in sensitive cells. In line, suppression of Mcl-1 by the pan-Bcl-2 inhibitor Obatoclax or overexpression of Noxa rendered resistant SCC cells sensitive to BH3 mimetics. Our data indicate that targeting of the Mcl-1/Noxa axis is important to overcome resistance to mitochondrial apoptosis in SCC. Therefore, combination treatment of ABT737 or derivatives with Mcl-1 inhibitors, or inducers of Noxa, may represent a novel option of targeted therapy in metastatic SCC of the skin.Apoptosis is an indispensible process to maintain cellular homeostasis, in particular in highly dynamic tissues. Apoptosis can be induced by activation of death receptors (DRs; such as TRAIL-R1/R2 or cluster of differentiation 95 (CD95)) or by intrinsic disturbance of mitochondria.1 Death ligands (DLs; TNF-related apoptosis-inducing ligand (TRAIL) or CD95L), when bound to their respective DRs, induce apoptosis by activation of procaspase-8 within the death-inducing signalling complex (DISC).2 Caspase-8 activation is followed by proteolytic cleavage of caspase-3.3 Extrinsic and intrinsic cell death is negatively controlled by caspase inhibitors such as X-linked inhibitor of apoptosis protein (XIAP)4 or by B-cell lymphoma 2 (Bcl-2) proteins that suppress the mitochondria outer membrane permeability (MOMP) by limiting Bax (Bcl-2-associated X protein)/Bak (Bcl-2 homologous antagonist/killer) translocation into the mitochondrial outer membrane.5 The extrinsic signalling cascade communicates with the intrinsic death pathway by cleavage of Bid (BH3 interacting-domain death agonist), a pro-apoptotic member of the BH3 (Bcl-2 homology domain 3)-only subfamily of Bcl-2 proteins.1 Other stimuli such as genotoxic stress allow for translocation and pore formation of pro-apoptotic multidomain Bcl-2 proteins Bax and Bak in the outer mitochondrial membrane.6, 7, 8 This process promotes release of mitochondria-derived apoptogenic proteins, in particular cytochrome c,9 or smac/DIABLO (second mitochondria-derived activator of caspases/direct IAP binding protein with low pI).10 Within the apoptosome,11 active caspase-9 finally leads to activation of caspase-3,12 and subsequent cell death.Anti-apoptotic multidomain Bcl-2 proteins (Bcl-2, Bcl-2-like protein 2 (Bcl-w), B-cell lymphoma-extra large (Bcl-XL), induced myeloid leukaemia cell differentiation protein (Mcl-1) and Bcl-2-related protein A1 (A1)) with four Bcl-2 homology domains (BH1, BH2, BH3 and BH4) suppress the pro-apoptotic function of Bax-like proteins such as Bax, Bak and Bok (that contain BH1–BH3 domains) or the BH3-only proteins Bad (Bcl-2-associated death promoter), Bim (Bcl-2-like protein 11), Bid, Noxa (phorbol-12-myristate-13-acetate-induced protein 1) and Puma (p53 upregulated modulator of apoptosis).13 Regulation of mitochondria-mediated apoptosis is determined by the balance between pro- and anti-apoptotic Bcl-2 proteins.14In a variety of cancer types, a decrease of BH3-only protein or upregulation of pro-survival Bcl-2 proteins is associated with poor prognosis.15 In metastatic squamous cell carcinoma (SCC) of the skin or the so-called ‘head and neck SCC'' (HNSCC), high expression of pro-survival Bcl-2 proteins conferred radio- and chemotherapy resistance.16, 17 These findings mark Bcl-2 proteins as regulators of SCC apoptosis and indicate that BH3 mimetics may hold therapeutic potential for metastatic SCC. The BH3 mimetics navitoclax (ABT263) and ABT199 are currently under investigation in clinical studies.18, 19, 20 Mechanistically, their lead compound ABT737 suppresses Bcl-2 activity by binding to the hydrophobic groove of Bcl-2, Bcl-w and Bcl-XL.18 As ABT263 upregulates Mcl-1, resistance to a number of Bcl-2 inhibitors (ABT737 and ABT263) has been described.21 Another compound, Obatoclax, was developed to block all anti-apoptotic Bcl-2 proteins including Mcl-1.22 Obatoclax blocks the interaction of Bim or Bax with Mcl-1.23 In this report, we have studied the effect of ABT737 for cell death in SCC of the skin and investigated the molecular mechanisms of resistance to different BH3 mimetics.  相似文献   

4.
5.
6.
Caspases and the cytotoxic lymphocyte protease granzyme B (GB) induce reactive oxygen species (ROS) formation, loss of transmembrane potential and mitochondrial outer membrane permeabilization (MOMP). Whether ROS are required for GB-mediated apoptosis and how GB induces ROS is unclear. Here, we found that GB induces cell death in an ROS-dependent manner, independently of caspases and MOMP. GB triggers ROS increase in target cell by directly attacking the mitochondria to cleave NDUFV1, NDUFS1 and NDUFS2 subunits of the NADH: ubiquinone oxidoreductase complex I inside mitochondria. This leads to mitocentric ROS production, loss of complex I and III activity, disorganization of the respiratory chain, impaired mitochondrial respiration and loss of the mitochondrial cristae junctions. Furthermore, we have also found that GB-induced mitocentric ROS are necessary for optimal apoptogenic factor release, rapid DNA fragmentation and lysosomal rupture. Interestingly, scavenging the ROS delays and reduces many of the features of GB-induced death. Consequently, GB-induced ROS significantly promote apoptosis.To induce cell death, human granzyme B (GB) activates effector caspase-3 or acts directly on key caspase substrates, such as the proapoptotic BH3 only Bcl-2 family member Bid, inhibitor of caspase-activated DNase (ICAD), poly-(ADP-ribose) polymerase-1 (PARP-1), lamin B, nuclear mitotic apparatus protein 1 (NUMA1), catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) and tubulin.1, 2, 3 Consequently, caspase inhibitors have little effect on human GB-mediated cell death and DNA fragmentation.2 GB causes reactive oxygen species (ROS) production, dissipation of the mitochondrial transmembrane potential (ΔΨm) and MOMP, which leads to the release of apoptogenic factors such as cytochrome c (Cyt c), HtrA2/Omi, endonuclease G (Endo G), Smac/Diablo and apoptosis-inducing factor, from the mitochondrial intermembrane space to the cytosol.4, 5, 6, 7, 8, 9, 10, 11 Interestingly, cells deficient for Bid, Bax and Bak are still sensitive to human GB-induced cell death,5, 11, 12, 13 suggesting that human GB targets the mitochondria in another way that needs to be characterized. Altogether, much attention has been focused on the importance of MOMP in the execution of GB-mediated cell death, leaving unclear whether ROS production is a bystander effect or essential to the execution of GB-induced apoptosis. The mitochondrial NADH: ubiquinone oxidoreductase complex I is a key determinant in steady-state ROS production. This 1 MDa complex, composed of 44 subunits,14 couples the transfer of two electrons from NADH to ubiquinone with the translocation of four protons to generate the ΔΨm. The importance of ROS has been previously demonstrated for caspase-3 and granzyme A (GA) pathways through the cleavage of NDUFS1 and NDUFS3, respectively.15, 16, 17, 18 GA induces cell death in a Bcl-2-insensitive and caspase- and MOMP-independent manner that has all the morphological features of apoptosis.1, 16, 17, 18, 19, 20 As GA and GB cell death pathways are significantly different, whether ROS are also critical for GB still need to be tested. Here, we show that GB induces ROS-dependent apoptosis by directly attacking the mitochondria in a caspase-independent manner to cleave NDUFS1, NDUFS2 and NDUFV1 in complex I. Consequently, GB inhibits electron transport chain (ETC) complex I and III activities, mitochondrial ROS production is triggered and mitochondrial respiration is compromised. Interestingly, MOMP is not required for GB to cleave the mitochondrial complex I subunits and ROS production. Moreover, GB action on complex I disrupts the organization of the respiratory chain and triggers the loss of the mitochondrial cristae junctions. We also show that GB-mediated mitocentric ROS are necessary for proper apoptogenic factor release from the mitochondria to the cytosol and for the rapid DNA fragmentation, both hallmarks of apoptosis. Moreover, GB-induced ROS are necessary for lysosomal membrane rupture. Thus, our work brings a new light to the GB pathway, showing that GB-mediated mitochondrial ROS are not adventitious waste of cell death, but essential mediators of apoptosis.  相似文献   

7.
Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising agent for anticancer therapy; however, non-small-cell lung carcinoma (NSCLC) cells are relatively TRAIL resistant. Identification of small molecules that can restore NSCLC susceptibility to TRAIL-induced apoptosis is meaningful. We found here that rotenone, as a mitochondrial respiration inhibitor, preferentially increased NSCLC cells sensitivity to TRAIL-mediated apoptosis at subtoxic concentrations, the mechanisms by which were accounted by the upregulation of death receptors and the downregulation of c-FLIP (cellular FLICE-like inhibitory protein). Further analysis revealed that death receptors expression by rotenone was regulated by p53, whereas c-FLIP downregulation was blocked by Bcl-XL overexpression. Rotenone triggered the mitochondria-derived reactive oxygen species (ROS) generation, which subsequently led to Bcl-XL downregulation and PUMA upregulation. As PUMA expression was regulated by p53, the PUMA, Bcl-XL and p53 in rotenone-treated cells form a positive feedback amplification loop to increase the apoptosis sensitivity. Mitochondria-derived ROS, however, promote the formation of this amplification loop. Collectively, we concluded that ROS generation, Bcl-XL and p53-mediated amplification mechanisms had an important role in the sensitization of NSCLC cells to TRAIL-mediated apoptosis by rotenone. The combined TRAIL and rotenone treatment may be appreciated as a useful approach for the therapy of NSCLC that warrants further investigation.Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) has emerged as a promising cancer therapeutic because it can selectively induce apoptosis in tumor cells in vitro, and most importantly, in vivo with little adverse effect on normal cells.1 However, a number of cancer cells are resistant to TRAIL, especially highly malignant tumors such as lung cancer.2, 3 Lung cancer, especially the non-small-cell lung carcinoma (NSCLC) constitutes a heavy threat to human life. Presently, the morbidity and mortality of NSCLC has markedly increased in the past decade,4 which highlights the need for more effective treatment strategies.TRAIL has been shown to interact with five receptors, including the death receptors 4 and 5 (DR4 and DR5), the decoy receptors DcR1 and DcR2, and osteoprotegerin.5 Ligation of TRAIL to DR4 or DR5 allows for the recruitment of Fas-associated protein with death domain (FADD), which leads to the formation of death-inducing signaling complex (DISC) and the subsequent activation of caspase-8/10.6 The effector caspase-3 is activated by caspase-8, which cleaves numerous regulatory and structural proteins resulting in cell apoptosis. Caspase-8 can also cleave the Bcl-2 inhibitory BH3-domain protein (Bid), which engages the intrinsic apoptotic pathway by binding to Bcl-2-associated X protein (Bax) and Bcl-2 homologous antagonist killer (BAK). The oligomerization between Bcl-2 and Bax promotes the release of cytochrome c from mitochondria to cytosol, and facilitates the formation of apoptosome and caspase-9 activation.7 Like caspase-8, caspase-9 can also activate caspase-3 and initiate cell apoptosis. Besides apoptosis-inducing molecules, several apoptosis-inhibitory proteins also exist and have function even when apoptosis program is initiated. For example, cellular FLICE-like inhibitory protein (c-FLIP) is able to suppress DISC formation and apoptosis induction by sequestering FADD.8, 9, 10, 11Until now, the recognized causes of TRAIL resistance include differential expression of death receptors, constitutively active AKT and NF-κB,12, 13 overexpression of c-FLIP and IAPs, mutations in Bax and BAK gene.2 Hence, resistance can be overcome by the use of sensitizing agents that modify the deregulated death receptor expression and/or apoptosis signaling pathways in cancer cells.5 Many sensitizing agents have been developed in a variety of tumor cell models.2 Although the clinical effectiveness of these agents needs further investigation, treatment of TRAIL-resistant tumor cells with sensitizing agents, especially the compounds with low molecular weight, as well as prolonged plasma half-life represents a promising trend for cancer therapy.Mitochondria emerge as intriguing targets for cancer therapy. Metabolic changes affecting mitochondria function inside cancer cells endow these cells with distinctive properties and survival advantage worthy of drug targeting, mitochondria-targeting drugs offer substantial promise as clinical treatment with minimal side effects.14, 15, 16 Rotenone is a potent inhibitor of NADH oxidoreductase in complex I, which demonstrates anti-neoplastic activity on a variety of cancer cells.17, 18, 19, 20, 21 However, the neurotoxicity of rotenone limits its potential application in cancer therapy. To avoid it, rotenone was effectively used in combination with other chemotherapeutic drugs to kill cancerous cells.22In our previous investigation, we found that rotenone was able to suppress membrane Na+,K+-ATPase activity and enhance ouabain-induced cancer cell death.23 Given these facts, we wonder whether rotenone may also be used as a sensitizing agent that can restore the susceptibility of NSCLC cells toward TRAIL-induced apoptosis, and increase the antitumor efficacy of TRAIL on NSCLC. To test this hypothesis, we initiated this study.  相似文献   

8.
Subversion of host cell apoptosis is an important survival strategy for viruses to ensure their own proliferation and survival. Certain viruses express proteins homologous in sequence, structure and function to mammalian pro-survival B-cell lymphoma 2 (Bcl-2) proteins, which prevent rapid clearance of infected host cells. In vaccinia virus (VV), the virulence factor F1L was shown to be a potent inhibitor of apoptosis that functions primarily be engaging pro-apoptotic Bim. Variola virus (VAR), the causative agent of smallpox, harbors a homolog of F1L of unknown function. We show that VAR F1L is a potent inhibitor of apoptosis, and unlike all other characterized anti-apoptotic Bcl-2 family members lacks affinity for the Bim Bcl-2 homology 3 (BH3) domain. Instead, VAR F1L engages Bid BH3 as well as Bak and Bax BH3 domains. Unlike its VV homolog, variola F1L only protects against Bax-mediated apoptosis in cellular assays. Crystal structures of variola F1L bound to Bid and Bak BH3 domains reveal that variola F1L forms a domain-swapped Bcl-2 fold, which accommodates Bid and Bak BH3 in the canonical Bcl-2-binding groove, in a manner similar to VV F1L. Despite the observed conservation of structure and sequence, variola F1L inhibits apoptosis using a startlingly different mechanism compared with its VV counterpart. Our results suggest that unlike during VV infection, Bim neutralization may not be required during VAR infection. As molecular determinants for the human-specific tropism of VAR remain essentially unknown, identification of a different mechanism of action and utilization of host factors used by a VAR virulence factor compared with its VV homolog suggest that studying VAR directly may be essential to understand its unique tropism.Variola virus (VAR), the causative agent of smallpox, is a member of the poxvirus family and belongs to the orthopoxviridae. Despite its successful eradication nearly 30 years ago, VAR remains an ongoing concern because of its potential use as a bioterrorism agent.1 The threat of intentional use of VAR coupled with the absence of an FDA-approved drug for the prevention or treatment of smallpox infection is cause for considerable interest in the development of small-molecule therapeutics against VAR. Current strategies for dealing with smallpox are based on vaccination using live vaccinia virus (VV),2, 3 a closely related member of the orthopoxvirus genus, which shares >90% sequence identity with VAR. Vaccination using live VV, however, can cause serious complications,4 underscoring the need for effective anti-viral treatments, particularly since anti-viral treatment may be a more efficacious strategy compared with vaccination.5 Recent strategies to target VAR for small-molecule therapeutics included the use of polymerase inhibitors,6 notably Cidofovir, inhibitors of extracellular virus formation7 and tyrosine kinase inhibitors including Gleevec.8, 9 Cidofovir is currently the only approved antiviral drug for the treatment of orthopoxviruses, although it is not approved for smallpox treatment. Other host–virus interactions have been identified that may be suitable drug targets10, 11 but currently require further investigation.Several poxvirus members other than VAR have been shown to rely on virulence factors that prevent premature host cell demise via programmed cell death or apoptosis,12, 13, 14, 15, 16 thus ensuring survival and proliferation. The B-cell lymphoma 2 (Bcl-2) protein family is a key mediator for maintaining cell survival or to drive apoptosis, thereby removing infected, damaged or unwanted cells,17 and sequence, structural and functional orthologs of Bcl-2 have been found in a number of poxviruses.18 Certain viral Bcl-2-like proteins were only identified as family members after their 3D structures were determined, owing to their complete lack of sequence identity to mammalian Bcl-2 proteins. This group of proteins include the myxoma virus M11L12 and VV F1L15 and N1L.19 Myxoma virus M11L was shown to adopt the classical Bcl-2 fold20, 21 that utilizes the canonical Bcl-2 homology 3 (BH3)-binding groove to engage BH3 ligands to exert its pro-survival effect. VV F1L also adopts a Bcl-2 fold, but unlike M11L it exists as a domain-swapped dimer,22, 23 whereas N1L also adopted a dimeric Bcl-2 fold but with a different dimeric arrangement.24, 25Although F1L from VAR has not previously been investigated, the VV homolog is well characterized. VV F1L has been shown to inhibit the mitochondrial pathway of apoptosis by replacing Mcl-126 and interacts with the isolated BH3 domains of Bim, Bax and Bak,23 which are bound in the canonical Bcl-2-binding groove.22 Furthermore, an F1L-deficient VV potently causes Bak/Bax-mediated apoptosis.15, 27 Functionally, VV F1L appears to rely primarily on neutralization of Bim in the context of a viral infection.22 Given the close similarity between VAR and VV, VAR may also rely on inhibition of host cell apoptosis for successful infection and proliferation. Disruption of VAR ability to inhibit apoptosis thus may constitute an attractive strategy for small-molecule-based intervention. To investigate this possibility, we performed a biochemical, structural and functional characterization of VAR F1L. Here we report that despite possessing a nearly identical 3D structure and sequence, VAR F1L inhibits apoptosis via a different mechanism compared with its homolog in VV.  相似文献   

9.
Caspase-3 is the best known executioner caspase in apoptosis. We generated caspase-3 knockout (C3KO) and knockdown human colorectal cancer cells, and found that they are unexpectedly sensitized to DNA-damaging agents including 5-fluorouracil (5-FU), etoposide, and camptothecin. C3KO xenograft tumors also displayed enhanced therapeutic response and cell death to 5-FU. C3KO cells showed intact apoptosis and activation of caspase-7 and -9, impaired processing of caspase-8, and induction of necrosis in response to DNA-damaging agents. This form of necrosis is associated with HMGB1 release and ROS production, and suppressed by genetic or pharmacological inhibition of RIP1, MLKL1, or caspase-8, but not inhibitors of pan-caspases or RIP3. 5-FU treatment led to the formation of a z-VAD-resistant pro-caspase-8/RIP1/FADD complex, which was strongly stabilized by caspase-3 KO. These data demonstrate a key role of caspase-3 in caspase-8 processing and suppression of DNA damage-induced necrosis, and provide a potentially novel way to chemosensitize cancer cells.Colorectal cancer is a major cancer killer in the United States and worldwide.1 Chemotherapeutic agents such as 5-fluorouracil (5-FU) and irinotecan (Camptosar) are commonly used in treating patients with colon cancer and other solid tumors. However, the 5-year survival of colon cancer patients with advanced diseases is <10% even with aggressive treatments.1 Most conventional chemotherapeutic agents cause DNA damage and trigger apoptosis,2 which is regulated by mitochondria-dependent intrinsic and death receptor-dependent extrinsic apoptotic pathways converging on the activation of executioner caspases-3 and -7.2 During transformation, neoplastic cells frequently become resistant to apoptosis via genetic and epigenetic mechanisms, driving accumulation of additional oncogenic events, and therapeutic resistance.3 Therefore, the exploration of alternative death pathways might provide new therapeutic options.Necrosis has long been viewed as an unregulated form of cell demise that promotes inflammation and tissue damage, whereas emerging evidence indicates that some forms of necrosis are programmed.4, 5 They can be initiated upon activation of the extended TNF-α receptor family at the cell surface, propagated through the receptor-interacting serine–threonine kinases, RIP1 and RIP3, and actively suppressed by apoptosis.6, 7, 8, 9 In mice, loss of caspase-8 leads to RIP3-dependent necrosis and embryonic lethality,10, 11 or intestinal inflammation involving TNF-α production.12, 13 In HT29 colon cancer cells, the addition of pan-caspase inhibitor z-VAD switches TNF-α and SMAC mimetic-induced apoptosis to RIP1/RIP3-dependent necrosis via downstream effector proteins mixed lineage kinase domain-like protein (MLKL) and phosphoglycerate mutase family member 5 (PGAM5).14, 15 Induction of programmed necrosis, or necroptosis, is stimuli- and cell type-dependent, and can also occur independent of either RIP1, RIP3,6, 16, 17 or both.18 The role and regulation of necrosis following DNA damage in relation to therapeutic outcomes has remained largely unexplored.8, 9In the current study, we report an unexpected function of caspase-3 in suppressing necrosis triggered by DNA-damaging agents in colon cancer cells. Caspase-3 knockout (C3KO) or knockdown (KD) colon cancer cells showed normal apoptotic response, but increased sensitivities to DNA-damaging agents in cell culture and in mice, at least in part, via RIP1-, and caspase-8-dependent necrosis. Our findings provide a potentially novel approach to chemosensitize cancer cells.  相似文献   

10.
The mitochondrial pathway of apoptosis is initiated by Bcl-2 homology region 3 (BH3)-only members of the Bcl-2 protein family. On upregulation or activation, certain BH3-only proteins can directly bind and activate Bak and Bax to induce conformation change, oligomerization and pore formation in mitochondria. BH3-only proteins, with the exception of Bid, are intrinsically disordered and therefore, functional studies often utilize peptides based on just their BH3 domains. However, these reagents do not possess the hydrophobic membrane targeting domains found on the native BH3-only molecule. To generate each BH3-only protein as a recombinant protein that could efficiently target mitochondria, we developed recombinant Bid chimeras in which the BH3 domain was replaced with that of other BH3-only proteins (Bim, Puma, Noxa, Bad, Bmf, Bik and Hrk). The chimeras were stable following purification, and each immunoprecipitated with full-length Bcl-xL according to the specificity reported for the related BH3 peptide. When tested for activation of Bak and Bax in mitochondrial permeabilization assays, Bid chimeras were ~1000-fold more effective than the related BH3 peptides. BH3 sequences from Bid and Bim were the strongest activators, followed by Puma, Hrk, Bmf and Bik, while Bad and Noxa were not activators. Notably, chimeras and peptides showed no apparent preference for activating Bak or Bax. In addition, within the BH3 domain, the h0 position recently found to be important for Bax activation, was important also for Bak activation. Together, our data with full-length proteins indicate that most BH3-only proteins can directly activate both Bak and Bax.The Bcl-2 family of proteins controls the mitochondrial pathway of apoptosis, a process often dysregulated in cancer and other diseases.1, 2, 3 Apoptotic triggers including DNA damage and oncogene activation cause the synthesis or activation of one or more pro-apoptotic Bcl-2 homology region 3 (BH3)-only proteins,1, 2, 3, 4 a subfamily that includes Bid, Bim, Puma, Noxa, Bad, Bik, Bmf and Hrk. These proteins then engage via their BH3 domain with other Bcl-2 family members. BH3-only proteins that can directly bind and activate the Bcl-2 effector proteins Bak or Bax are called ‘activators''.5 When Bak or Bax become activated and oligomerize in the mitochondrial outer membrane (MOM), the apoptotic ‘switch'' has flipped and the cell is committed to cell death. The prosurvival members (Bcl-2, Bcl-xL, Mcl-1, Bcl-w, Bfl-1/A1 and Bcl-B) inhibit apoptosis by specifically binding both the BH3-only proteins and activated Bak and Bax.6, 7, 8, 9, 10, 11 Thus, the cell''s complement of prosurvival proteins, Bak, and Bax, determines the sensitivity of that cell to each BH3-only protein, and by extension to each type of pro-apoptotic stimulus.A thorough understanding of BH3-only proteins is crucial for the development of cancer therapeutics such as the new class of anti-cancer molecules called BH3 mimetics that are showing significant promise in clinical trials.12, 13 The binding of BH3-only proteins to prosurvival proteins has been well-characterized and revealed significant preferences for engaging different members.6, 8, 9 How BH3-only proteins bind and activate Bak and Bax remains less understood for several reasons. First, generating stable recombinant BH3-only proteins is difficult because, except for Bid, they are intrinsically disordered14, 15, 16 and because most contain hydrophobic C-terminal membrane anchors.17 Thus, most in vitro studies of BH3-only proteins have used synthetic peptides corresponding to the BH3 domains, C-terminally truncated recombinant proteins or in vitro translated (IVT) proteins. Second, BH3-only reagents bind poorly to recombinant Bak and Bax in the absence of membranes, although detergents and liposomes may substitute for the MOM.18, 19, 20 Third, activation of Bak and Bax on mitochondria can be complicated by the presence of other proteins such as prosurvival proteins. Indeed, genetically altering BH3-only protein levels in mice resulted in complex phenotypes due to multiple interactions between family members, precluding firm conclusions as to which BH3-only proteins are direct activators.18, 21, 22Bid and Bim are direct activators according to a variety of approaches,5, 8, 9, 23, 24 and were recently proposed to be specific for Bak and Bax, respectively.25 Early studies using Noxa BH3 peptides5, 8 and IVT Noxa9 concluded that Noxa was not an activator. However, in more recent studies a Noxa BH3 peptide23 and purified recombinant NoxaΔC20 were found to be activators of both Bak and Bax. Puma has also been described as both an activator26, 27 and not an activator.8, 28 Du et al.23 analyzed the full panel of BH3 peptides and classified Bim as a strong activator, Bid, Noxa and Bmf as moderate activators, and Puma, Bik and Hrk as weak activators. The only BH3-only member that has never been described as an activator is Bad.While BH3 peptides and recombinant truncated BH3-only proteins have been useful for in vitro studies, new reagents that target mitochondria may better reflect the behavior of the parent proteins. As Bid is stable as a recombinant protein, we generated chimeras of Bid in which the BH3 domain of Bid was replaced with that of seven other BH3-only proteins. This is a similar approach to the Bim chimeras used for expression in cells18 and in mice.29 More recently, truncated Bid (tBid) chimeras containing the BH3 domains of Bim, Bak and Bax as well as those of the prosurvival proteins, have been generated as IVT proteins.11To compare the ability of BH3-only proteins to activate Bak and Bax in vitro, we incubated Bid chimeras and BH3 peptides with mitochondria containing either Bak or Bax. We found that the membrane-targeted Bid chimeras were much more potent activators than their related BH3 peptides, and that all BH3 domains except for Bad and Noxa were activators to some extent. We conclude that activation of Bak and Bax may be underestimated by studies using BH3 peptides, and that even BH3-only proteins such as Bik, Bmf and Hrk that are often considered unable to activate Bak or Bax, may act as activators under certain conditions.  相似文献   

11.
Neutral sphingomyelinase (nSMase) activation in response to environmental stress or inflammatory cytokine stimuli generates the second messenger ceramide, which mediates the stress-induced apoptosis. However, the signaling pathways and activation mechanism underlying this process have yet to be elucidated. Here we show that the phosphorylation of nSMase1 (sphingomyelin phosphodiesterase 2, SMPD2) by c-Jun N-terminal kinase (JNK) signaling stimulates ceramide generation and apoptosis and provide evidence for a signaling mechanism that integrates stress- and cytokine-activated apoptosis in vertebrate cells. An nSMase1 was identified as a JNK substrate, and the phosphorylation site responsible for its effects on stress and cytokine induction was Ser-270. In zebrafish cells, the substitution of Ser-270 for alanine blocked the phosphorylation and activation of nSMase1, whereas the substitution of Ser-270 for negatively charged glutamic acid mimicked the effect of phosphorylation. The JNK inhibitor SP600125 blocked the phosphorylation and activation of nSMase1, which in turn blocked ceramide signaling and apoptosis. A variety of stress conditions, including heat shock, UV exposure, hydrogen peroxide treatment, and anti-Fas antibody stimulation, led to the phosphorylation of nSMase1, activated nSMase1, and induced ceramide generation and apoptosis in zebrafish embryonic ZE and human Jurkat T cells. In addition, the depletion of MAPK8/9 or SMPD2 by RNAi knockdown decreased ceramide generation and stress- and cytokine-induced apoptosis in Jurkat cells. Therefore the phosphorylation of nSMase1 is a pivotal step in JNK signaling, which leads to ceramide generation and apoptosis under stress conditions and in response to cytokine stimulation. nSMase1 has a common central role in ceramide signaling during the stress and cytokine responses and apoptosis.The sphingomyelin pathway is initiated by the hydrolysis of sphingomyelin to generate the second messenger ceramide.1 Sphingomyelin hydrolysis is a major pathway for stress-induced ceramide generation. Neutral sphingomyelinase (nSMase) is activated by a variety of environmental stress conditions, such as heat shock,1, 2, 3 oxidative stress (hydrogen peroxide (H2O2), oxidized lipoproteins),1 ultraviolet (UV) radiation,1 chemotherapeutic agents,4 and β-amyloid peptides.5, 6 Cytokines, including tumor necrosis factor (TNF)-α,7, 8, 9 interleukin (IL)-1β,10 Fas ligand,11 and their associated proteins, also trigger the activation of nSMase.12 Membrane-bound Mg2+-dependent nSMase is considered to be a strong candidate for mediating the effects of stress and inflammatory cytokines on ceramide.3Among the four vertebrate nSMases, nSMase1 (SMPD2) was the first to be cloned and is localized in the endoplasmic reticulum (ER) and Golgi apparatus.13 Several studies have focused on the potential signaling roles of nSMase1, and some reports have suggested that nSMase1 is important for ceramide generation in response to stress.5, 6, 14, 15 In addition, nSMase1 is responsible for heat-induced apoptosis in zebrafish embryonic cultured (ZE) cells, and a loss-of-function study showed a reduction in ceramide generation, caspase-3 activation, and apoptosis in zebrafish embryos.16 However, nSMase1-knockout mice showed no lipid storage diseases or abnormalities in sphingomyelin metabolism.17 Therefore, the molecular mechanisms by which nSMase1 is activated have yet to be elucidated.Environmental stress and inflammatory cytokines1, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27 stimulate stress-activated protein kinase (SAPK)/c-Jun N-terminal kinase (JNK) signaling, which involves the sequential activation of members of the mitogen-activated protein kinase (MAPK) family, including MAPK/ERK kinase kinase (MEKK)1/MAPK kinase (MKK)4, and/or SAPK/ERK kinase (SEK)1/MKK7, JNK, and c-jun. Both the JNK and sphingomyelin signaling pathways coordinately mediate the induction of apoptosis.1 However, possible crosstalk between the JNK and sphingomyelin signaling pathways has not yet been characterized. Previously, we used SDS-PAGE to determine that nSMase1 polypeptides migrated at higher molecular masses,16 suggesting that the sphingomyelin signaling pathway might cause the production of a chemically modified phosphorylated nSMase1, which is stimulated under stressed conditions in ZE cells.16 Here, we demonstrate that JNK signaling results in the phosphorylation of Ser-270 of nSMase1, which initiates ceramide generation and apoptosis. We also provide evidence for a signaling mechanism that integrates cytokine- and stress-activated apoptosis in vertebrate cells. We studied stress-induced ceramide generation in two cell types: ZE cells and human leukemia Jurkat T-lymphoid cells. Stress-induced apoptosis has been investigated in these systems previously.16, 28  相似文献   

12.
Mitochondrial outer membrane permeabilization (MOMP) is a critical control point during apoptosis that results in the release of pro-apoptotic mitochondrial contents such as cytochrome c. MOMP is largely controlled by Bcl-2 family proteins such as Bax, which under various apoptotic stresses becomes activated and oligomerizes on the outer mitochondrial membrane. Bax oligomerization helps promote the diffusion of the mitochondrial contents into the cytoplasm activating the caspase cascade. In turn, Bax is regulated primarily by anti-apoptotic Bcl-2 proteins including Bcl-xL, which was recently shown to prevent Bax from accumulating at the mitochondria. However, the exact mechanisms by which Bcl-xL regulates Bax and thereby MOMP remain partially understood. In this study, we show that the small CHCH-domain-containing protein CHCHD2 binds to Bcl-xL and inhibits the mitochondrial accumulation and oligomerization of Bax. Our data show that in response to apoptotic stimuli, mitochondrial CHCHD2 decreases prior to MOMP. Furthermore, when CHCHD2 is absent from the mitochondria, the ability of Bcl-xL to inhibit Bax activation and to prevent apoptosis is attenuated, which results in increases in Bax oligomerization, MOMP and apoptosis. Collectively, our findings establish CHCHD2, a previously uncharacterized small mitochondrial protein with no known homology to the Bcl-2 family, as one of the negative regulators of mitochondria-mediated apoptosis.Apoptosis is a tightly regulated form of programmed cell death that is critical for proper embryonic development, tissue homeostasis and immune response. Aberrant regulation of apoptosis contributes to a wide range of ailments including autoimmune disorders, neurodegenerative diseases and cancer. Unlike necrotic cell death, apoptosis is a genetic program that is characterized by distinct morphological features such as membrane blebbing, chromatin condensation, DNA fragmentation and cell shrinkage.1 In vertebrates, apoptosis can occur through two pathways: extrinsic, or receptor-mediated apoptosis, and intrinsic, or mitochondria-mediated apoptosis. Intrinsic apoptosis is induced by cellular stressors such as DNA damage, which lead to mitochondrial outer membrane permeabilization (MOMP), cytochrome c release from the mitochondrial intermembrane space, activation of cysteine proteases (caspases) and induction of apoptosis. Once MOMP occurs, cell death is thought to be inevitable. Therefore, much research has been devoted to elucidating the mechanisms and signaling pathways that govern this critical regulatory point in apoptosis.MOMP is controlled largely by the B-cell lymphoma 2 (Bcl-2) family of proteins,2 all of which contain at least one of four BH (Bcl-2 homology) domains designated BH1–4. During apoptosis, the pro-apoptotic Bcl-2 proteins Bax and/or Bak become activated and oligomerize on the mitochondrial outer membrane3 increasing mitochondrial membrane permeabilization through a mechanism that is not entirely clear. Bax and Bak are activated by BH3-only Bcl-2 family proteins such as Bim, t-Bid and Puma.4, 5, 6, 7, 8, 9, 10, 11, 12, 13 Conversely, Bax and Bak are inhibited by pro-survival Bcl-2 family proteins such as Bcl-2, Mcl-1 and Bcl-xL.2, 14, 15, 16 Of the pro-survival Bcl-2 family proteins, Bcl-2 is found at the outer mitochondrial membrane, whereas Bcl-xL and Mcl-1 localize to the outer mitochondrial membrane and the mitochondrial matrix.17, 18 Matrix-localized Bcl-xL and Mcl-1 have been shown to promote mitochondrial respiration,19 suggesting that crosstalk exists between apoptotic pathways and other mitochondria-based biological events. Based on this recent discovery, one might reason that other mitochondrial proteins previously characterized as structural proteins or metabolism-associated enzymes could play an additional intermediate role in the regulation of apoptosis by interacting with Bcl-2 family proteins.We identified CHCHD2 in a mass spectrometry-based screen for binding partners of p32, a mitochondrial protein previously shown by our lab to bind and mediate the apoptotic effects of the tumor suppressor p14ARF.20 CHCHD2 was subsequently detected in independent screens for proteins that regulate cellular metabolism and migration;21, 22 however, the functions of CHCHD2 remain unknown. CHCHD2 is encoded by the chchd2 gene (coiled-coil helix coiled-coil helix domain-containing 2), which spans 4921 base pairs, contains 4 exons, and is located on human chromosome 7p11.2, a chromosomal region that is often amplified in glioblastomas.23 The protein encoded by the chchd2 gene is ubiquitously expressed24 and is relatively small, as it codes for only 151 amino acids. CHCHD2 is well-conserved among different species from humans to yeast, and mouse and human CHCHD2 share 87% amino acid sequence identity (Supplementary Figures S1A and S1B). CHCHD2 contains a C-terminal CHCH (coiled-coil helix coiled-coil helix) domain, which is characterized primarily by four cysteine residues spaced 10 amino acids apart from one another (CX(9)C motif).25 The function of the CHCH domain is not well understood, and the few characterized proteins that harbor this domain have diverse functions. Many CHCH domain-containing proteins localize to the mitochondrial inner membrane or the intermembrane space, including Cox12, Cox17, Cox19, Cox23, Mia40 (yeast homolog of human CHCHD4), CHCHD3 and CHCHD6. Cox17 and Cox19 aid in the assembly of the COX complex,26, 27 whereas Mia40/Tim40 has been shown to transport proteins into the mitochondrial intermembrane space.28, 29 Furthermore, CHCHD3 and CHCHD6 are essential for maintaining the integrity of mitochondrial cristae and thus mitochondrial function.30, 31, 32 Interestingly, a recent report has shown that CHCHD6 is regulated by DNA damage stress, and alterations in CHCHD6 expression affect the viability of breast cancer cells in response to genotoxic anticancer drugs.32Despite advances in our understanding of how MOMP and apoptosis are regulated by the Bcl-2 family of proteins, much remains unknown with respect to the mechanisms that lead to Bax activation and oligomerization particularly concerning the roles that mitochondria-associated proteins play in the process. In this study, we characterize the small, mitochondria-localized protein CHCHD2 as a novel regulator of Bax oligomerization and apoptosis. Furthermore, we show evidence that CHCHD2 binds to Bcl-xL at the mitochondria under unstressed conditions. In response to apoptotic stimuli, CHCHD2 decreases and loses its mitochondria localization, which is accompanied by decreased Bcl-xL–Bax interaction and increased Bax homo-oligomerization and Bax–Bak hetero-oligomerization. Collectively, our results suggest that CHCHD2 negatively regulates the apoptotic cascade upstream of Bax oligomerization.  相似文献   

13.
14.
15.
16.
Components of the death receptor-mediated pathways like caspase-8 have been identified in complexes at intracellular membranes to spatially restrict the processing of local targets. In this study, we report that the long isoform of the cellular FLICE-inhibitory protein (c-FLIPL), a well-known inhibitor of the extrinsic cell death initiator caspase-8, localizes at the endoplasmic reticulum (ER) and mitochondria-associated membranes (MAMs). ER morphology was disrupted and ER Ca2+-release as well as ER-mitochondria tethering was decreased in c-FLIP−/− mouse embryonic fibroblasts (MEFs). Mechanistically, c-FLIP ablation resulted in enhanced basal caspase-8 activation and in caspase-mediated processing of the ER-shaping protein reticulon-4 (RTN4) that was corrected by re-introduction of c-FLIPL and caspase inhibition, resulting in the recovery of a normal ER morphology and ER-mitochondria juxtaposition. Thus, the caspase-8 inhibitor c-FLIPL emerges as a component of the MAMs signaling platforms, where caspases appear to regulate ER morphology and ER-mitochondria crosstalk by impinging on ER-shaping proteins like the RTN4.Cellular FLICE inhibitory proteins (c-FLIP) inhibit death receptor (DR)-mediated apoptosis, by preventing caspase-8 activation.1 Among the three identified c-FLIP splicing forms,2, 3 c-FLIPS,R were described as cytosolic, whereas c-FLIPL was also observed in the nucleus. A pool of membrane-bound c-FLIPL was also described4 suggesting that caspase-8/c-FLIPL could re-distribute on stimulation, leading to a more subtle regulation of caspase-8 activity depending on substrates localization.5 Furthermore, caspase-8 itself and Fas-Associated Death Domain adaptor protein (FADD) were found or were shown to re-loca5lize in local complexes on ER6, 7, 8 and mitochondria,9, 10 mediating the exchange of signals between the two organelles.11, 12, 13 Several molecular platforms containing both membrane-bound proteins and cytosolic apoptosis modulators have been identified at the ER-mitochondria interface (the so-called mitochondria-associated membranes or MAMs),14 controlling ER-mitochondria anchorage as well as lipid metabolism, Ca2+ signaling and apoptosis.15 MAMs have been recently described as lipid raft-like domains that orient proteins to promote the ER-mitochondria juxtaposition;16 consequently, alterations in their composition may profoundly affect the physical and functional inter-organelle crosstalk. Furthermore, as mitochondrial and ER membranes are continuously and concertedly remodeled,17 it is not surprising that membrane-shaping proteins can also exert a function in regulating the ER-mitochondria coupling.12, 18 Different families of ER-shaping proteins control the organization of peripheral ER, which consists of sheet-like cisternae and tubules connected by three-way junctions.19 Among these, Reticulons (RTN) and Deleted in Polyposis locus 1 (DP1) proteins cause the ER membrane to curve and tubulate,20, 21 whereas the GTPases Atlastins (ATL) promote the branching of ER tubules;22 finally, ER sheet-enriched proteins such as the 63-kDa cytoskeleton-linking membrane protein (CLIMP63) control the width of ER cisternae, anchoring the organelle to microtubules and maintaining its spatial distribution.23, 24 Along with other components of the extrinsic apoptosis, here we described for the first time the enrichment of c-FLIPL at ER and ER-mitochondria interface. Furthermore, we observed that ER structure and tethering to mitochondria are impaired in cells lacking c-FLIP. Given the importance of membrane-shaping proteins and MAM complexes in regulating organelles structure and ER-mitochondria juxtaposition, we focused on the mechanism underlying this phenotype and we found that c-FLIPL deficiency induces the caspase-mediated processing of RTN4, thus affecting organelle shape and coupling to mitochondria. We therefore concluded that c-FLIPL is a novel regulator of ER morphology and ER-mitochondria crosstalk.  相似文献   

17.
Tumor necrosis factor α (TNFα) triggers necroptotic cell death through an intracellular signaling complex containing receptor-interacting protein kinase (RIPK) 1 and RIPK3, called the necrosome. RIPK1 phosphorylates RIPK3, which phosphorylates the pseudokinase mixed lineage kinase-domain-like (MLKL)—driving its oligomerization and membrane-disrupting necroptotic activity. Here, we show that TNF receptor-associated factor 2 (TRAF2)—previously implicated in apoptosis suppression—also inhibits necroptotic signaling by TNFα. TRAF2 disruption in mouse fibroblasts augmented TNFα–driven necrosome formation and RIPK3-MLKL association, promoting necroptosis. TRAF2 constitutively associated with MLKL, whereas TNFα reversed this via cylindromatosis-dependent TRAF2 deubiquitination. Ectopic interaction of TRAF2 and MLKL required the C-terminal portion but not the N-terminal, RING, or CIM region of TRAF2. Induced TRAF2 knockout (KO) in adult mice caused rapid lethality, in conjunction with increased hepatic necrosome assembly. By contrast, TRAF2 KO on a RIPK3 KO background caused delayed mortality, in concert with elevated intestinal caspase-8 protein and activity. Combined injection of TNFR1-Fc, Fas-Fc and DR5-Fc decoys prevented death upon TRAF2 KO. However, Fas-Fc and DR5-Fc were ineffective, whereas TNFR1-Fc and interferon α receptor (IFNAR1)-Fc were partially protective against lethality upon combined TRAF2 and RIPK3 KO. These results identify TRAF2 as an important biological suppressor of necroptosis in vitro and in vivo.Apoptotic cell death is mediated by caspases and has distinct morphological features, including membrane blebbing, cell shrinkage and nuclear fragmentation.1, 2, 3, 4 In contrast, necroptotic cell death is caspase-independent and is characterized by loss of membrane integrity, cell swelling and implosion.1, 2, 5 Nevertheless, necroptosis is a highly regulated process, requiring activation of RIPK1 and RIPK3, which form the core necrosome complex.1, 2, 5 Necrosome assembly can be induced via specific death receptors or toll-like receptors, among other modules.6, 7, 8, 9 The activated necrosome engages MLKL by RIPK3-mediated phosphorylation.6, 10, 11 MLKL then oligomerizes and binds to membrane phospholipids, forming pores that cause necroptotic cell death.10, 12, 13, 14, 15 Unchecked necroptosis disrupts embryonic development in mice and contributes to several human diseases.7, 8, 16, 17, 18, 19, 20, 21, 22The apoptotic mediators FADD, caspase-8 and cFLIP suppress necroptosis.19, 20, 21, 23, 24 Elimination of any of these genes in mice causes embryonic lethality, subverted by additional deletion of RIPK3 or MLKL.19, 20, 21, 25 Necroptosis is also regulated at the level of RIPK1. Whereas TNFα engagement of TNFR1 leads to K63-linked ubiquitination of RIPK1 by cellular inhibitor of apoptosis proteins (cIAPs) to promote nuclear factor (NF)-κB activation,26 necroptosis requires suppression or reversal of this modification to allow RIPK1 autophosphorylation and consequent RIPK3 activation.2, 23, 27, 28 CYLD promotes necroptotic signaling by deubiquitinating RIPK1, augmenting its interaction with RIPK3.29 Conversely, caspase-8-mediated CYLD cleavage inhibits necroptosis.24TRAF2 recruits cIAPs to the TNFα-TNFR1 signaling complex, facilitating NF-κB activation.30, 31, 32, 33 TRAF2 also supports K48-linked ubiquitination and proteasomal degradation of death-receptor-activated caspase-8, curbing apoptosis.34 TRAF2 KO mice display embryonic lethality; some survive through birth but have severe developmental and immune deficiencies and die prematurely.35, 36 Conditional TRAF2 KO leads to rapid intestinal inflammation and mortality.37 Furthermore, hepatic TRAF2 depletion augments apoptosis activation via Fas/CD95.34 TRAF2 attenuates necroptosis induction in vitro by the death ligands Apo2L/TRAIL and Fas/CD95L.38 However, it remains unclear whether TRAF2 regulates TNFα-induced necroptosis—and if so—how. Our present findings reveal that TRAF2 inhibits TNFα necroptotic signaling. Furthermore, our results establish TRAF2 as a biologically important necroptosis suppressor in vitro and in vivo and provide initial insight into the mechanisms underlying this function.  相似文献   

18.
19.
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   

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