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 31 毫秒
1.
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   

2.
Intrinsic apoptosis involves BH3-only protein activation of Bax/Bak-mediated mitochondrial outer membrane permeabilization (MOMP). Consequently, cytochrome c is released from the mitochondria to activate caspases, and Smac (second mitochondria-derived activator of caspases) to inhibit XIAP-mediated caspase suppression. Dysfunctional mitochondria can be targeted for lysosomal degradation via autophagy (mitophagy), or directly through mitochondria-derived vesicle transport. However, the extent of autophagy and lysosomal interactions with apoptotic mitochondria remains largely unknown. We describe here a novel pathway of endolysosomal processing of mitochondria, activated in response to canonical BH3-only proteins and mitochondrial depolarization. We report that expression of canonical BH3-only proteins, tBid, BimEL, Bik, Bad, and mitophagy receptor mutants of atypical BH3-only proteins, Bnip3 and Bnip3L/Nix, leads to prominent relocalization of endolysosomes into inner mitochondrial compartments, in a manner independent of mitophagy. As an upstream regulator, we identified the XIAP E3 ligase. In response to mitochondrial depolarization, XIAP actuates Bax-mediated MOMP, even in the absence of BH3-only protein signaling. Subsequently, in an E3 ligase-dependent manner, XIAP rapidly localizes inside all the mitochondria, and XIAP-mediated mitochondrial ubiquitylation catalyses interactions of Rab membrane targeting components Rabex-5 and Rep-1 (RFP-tagged Rab escort protein-1), and Rab5- and Rab7-positive endolysosomes, at and within mitochondrial membrane compartments. While XIAP-mediated MOMP permits delayed cytochrome c release, within the mitochondria XIAP selectively signals lysosome- and proteasome-associated degradation of its inhibitor Smac. These findings suggest a general mechanism to lower the mitochondrial apoptotic potential via intramitochondrial degradation of Smac.The intrinsic mitochondrial apoptotic pathway is required for efficient chemotherapeutic killing of cancer cells,1 and is initiated through BH3-only protein activation of Bax/Bak-mediated mitochondrial outer membrane permeabilization (MOMP). MOMP releases cytochrome c to activate effector caspases.2 Conversely, inhibitor of apoptosis protein (IAP) family members suppress initiator and effector caspases via direct binding and E3 ligase activities.3, 4, 5 Consequently, MOMP-induced release of Smac (second mitochondria-derived activator of caspases) from the mitochondria, to inhibit XIAP (X-chromosome-linked IAP)-mediated caspase suppression, can be required for apoptosis.6Autophagy, a lysosomal degradative mechanism undergoing extensive crosstalk with cell death and survival pathways,7 degrades damaged mitochondria in a process termed mitophagy.8, 9 Damaged mitochondria are targeted by lysosomal degradation through the recruitment of autophagy receptors to the outer mitochondrial membrane (OMM),8 or via delivery of mitochondrial-derived vesicles (MDVs) directly to the lysosome.10 The E3 ubiquitin ligase Parkin targets and ubiquitylates mitochondria, mediating both MDV degradation11 and autophagy receptor-dependent mitophagy.12, 13 Alternatively, Fundc114 and atypical BH3-only Nip family proteins Bnip3 and Bnip3L/Nix localize to the OMM and act as mitophagy receptors via their LC3-interacting region (LIR).15, 16, 17, 18 While targeting of damaged mitochondria suggests that mitophagy may counter apoptotic mitochondria, mitophagy occurs progressively over days,12, 14, 16, 17, 19 a rate that is likely insufficient to alter intracellular propagation of mitochondrial apoptosis, which can occur within minutes.20, 21 Indeed, Bnip3- and Fundc1-induced mitophagy have no direct effect on apoptosis,14, 18 and we determined that Bnip3-mediated mitophagy was cytoprotective if activated before apoptosis.17 While MDV delivery of mitochondria to lysosomes operates at a higher rate, minutes to hours,10 this process is regulated by Parkin and restricted to specific mitochondrial components.11 Overall, for most intrinsic apoptosis scenarios it remains unknown whether lysosomal processing of mitochondria influences their capacity to activate or enhance apoptosis.Here, we used high-resolution imaging to evaluate the behavior of apoptosis, autophagy, lysosomal and ubiquitylation pathways in response to canonical (tBid, BimEL, Bik, Bad) and atypical (Bnip3, Bnip3L/Nix) BH3-only protein expression. We report that, in parallel to intrinsic apoptosis signaling, canonical BH3-only proteins induce the recruitment of endolysosomal machinery, in the absence of mitophagy. We determined that mitochondrial depolarization rapidly translocates the caspase inhibitor XIAP to the mitochondria. There, XIAP actuates MOMP within all mitochondria, concomitant with ubiquitylation at the OMM and inside OMM-bound regions, and triggers ubiquitin-dependent recruitment of Rab5 and its binding partners, as well as late endosomes into the mitochondria. Consequently, in a manner dependent on lysosome- and proteasome-activities, XIAP degrades its inhibitor Smac. We propose that in response to bioenergetic stress, the functional integration between lysosomes and mitochondria, mediated by XIAP and independent of autophagy, offers a novel mechanism to modulate mitochondrial apoptosis.  相似文献   

3.
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   

4.
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   

5.
In non-apoptotic cells, Bak constitutively resides in the mitochondrial outer membrane. In contrast, Bax is in a dynamic equilibrium between the cytosol and mitochondria, and is commonly predominant in the cytosol. In response to an apoptotic stimulus, Bax and Bak change conformation, leading to Bax accumulation at mitochondria and Bak/Bax oligomerization to form a pore in the mitochondrial outer membrane that is responsible for cell death. Using blue native-PAGE to investigate how Bax oligomerizes in the mitochondrial outer membrane, we observed that, like Bak, a proportion of Bax that constitutively resides at mitochondria associates with voltage-dependent anion channel (VDAC)2 prior to an apoptotic stimulus. During apoptosis, Bax dissociates from VDAC2 and homo-oligomerizes to form high molecular weight oligomers. In cells that lack VDAC2, constitutive mitochondrial localization of Bax and Bak was impaired, suggesting that VDAC2 has a role in Bax and Bak import to, or stability at, the mitochondrial outer membrane. However, following an apoptotic stimulus, Bak and Bax retained the ability to accumulate at VDAC2-deficient mitochondria and to mediate cell death. Silencing of Bak in VDAC2-deficient cells indicated that Bax required either VDAC2 or Bak in order to translocate to and oligomerize at the mitochondrial outer membrane to efficiently mediate apoptosis. In contrast, efficient Bak homo-oligomerization at the mitochondrial outer membrane and its pro-apoptotic function required neither VDAC2 nor Bax. Even a C-terminal mutant of Bax (S184L) that localizes to mitochondria did not constitutively target mitochondria deficient in VDAC2, but was recruited to mitochondria following an apoptotic stimulus dependent on Bak or upon over-expression of Bcl-xL. Together, our data suggest that Bax localizes to the mitochondrial outer membrane via alternate mechanisms, either constitutively via an interaction with VDAC2 or after activation via interaction with Bcl-2 family proteins.Bax and Bak are the key effectors of the intrinsic apoptotic pathway initiated in response to diverse stimuli including anoikis, DNA damage and growth factor withdrawal.1 Both proteins are normally dormant in healthy cells, but upon reception of an apoptotic stimulus, they undergo conformation change that allows their self-association to form pores in the mitochondrial outer membrane (MOM).2, 3, 4, 5, 6, 7 The consequence of disruption of the MOM is twofold; it impairs the ability of mitochondria to generate ATP by oxidative phosphorylation and it allows the release of intermembrane proteins including cytochrome c that agonizes caspases that dismantle the cell.Bak and Bax share significant structural homology in their inactive states and have conserved mechanism of conformation change and oligomerization.3, 8, 9, 10 Further, genetic studies reveal that Bak and Bax perform at least partially overlapping function, with deficiency in both necessary to perturb apoptosis during embryonic development and in response to toxic insult.1, 11 However, whether Bak and Bax are regulated similarly is unclear. Whereas Bak is constitutively anchored in the MOM via its hydrophobic C-terminal transmembrane domain, Bax is predominantly cytosolic in the majority of non-apoptotic cells.12 Recent evidence indicates that Bax is in a dynamic equilibrium between cytosol and mitochondria and is constantly trafficked away from the MOM in non-apoptotic cells.13, 14 In response to apoptotic stress this ‘retrotranslocation'' is disrupted causing Bax to accumulate at mitochondria; a hallmark of most apoptotic cells. The mechanism governing the dynamic distribution of Bax in healthy and apoptotic cells is unclear with interactions with pro-survival proteins debated.13, 14Voltage-dependent anion channels (VDACs) are the major channels responsible for ion passage across the MOM. Studies have also implicated an additional role for the VDACs in the regulation of Bak or Bax apoptotic function or potentially even constituting a component of the Bak/Bax apoptotic pore.15, 16, 17, 18 However, these studies have provided contrasting findings relating to whether VDACs might positively or negatively regulate Bak/Bax apoptotic function.We used blue native-PAGE (BN-PAGE) to investigate how Bax oligomerizes in the MOM during apoptosis. We observed that VDAC2 is a determinant of the constitutive association of both Bax and Bak with the MOM. The defect in Bax mitochondrial localization can be bypassed by Bak-dependent recruitment during apoptosis. Thus, our data suggest that mitochondrial localization of Bax occurs via distinct mechanisms in healthy and apoptotic cells and that either VDAC2 or Bak is required for the efficient translocation of Bax and hence for the oligomerization at the MOM and Bax apoptotic function.  相似文献   

6.
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   

7.
8.
Parkinson''s disease (PD) is the second most prevalent neurodegenerative disorder, affecting 1–3% of the population over 65. Mutations in the ubiquitin E3 ligase parkin are the most common cause of autosomal recessive PD. The parkin protein possesses potent cell-protective properties and has been mechanistically linked to both the regulation of apoptosis and the turnover of damaged mitochondria. Here, we explored these two functions of parkin and the relative scale of these processes in various cell types. While biochemical analyses and subcellular fractionation were sufficient to observe robust parkin-dependent mitophagy in immortalized cells, higher resolution techniques appear to be required for primary culture systems. These approaches, however, did affirm a critical role for parkin in the regulation of apoptosis in primary cultured neurons and all other cells studied. Our prior work demonstrated that parkin-dependent ubiquitination of endogenous Bax inhibits its mitochondrial translocation and can account for the anti-apoptotic effects of parkin. Having found a central role for parkin in the regulation of apoptosis, we further investigated the parkin-Bax interaction. We observed that the BH3 domain of Bax is critical for its recognition by parkin, and identified two lysines that are crucial for parkin-dependent regulation of Bax translocation. Last, a disease-linked mutation in parkin failed to influence Bax translocation to mitochondria after apoptotic stress. Taken together, our data suggest that regulation of apoptosis by the inhibition of Bax translocation is a prevalent physiological function of parkin regardless of the kind of cell stress, preventing overt cell death and supporting cell viability during mitochondrial injury and repair.Loss-of-function mutations in the ubiquitin E3 ligase parkin are the most common cause of autosomal recessive Parkinson''s disease (PD).1 Multiple functions have been ascribed to parkin, most notably the inhibition of apoptosis2, 3, 4, 5, 6, 7 and the induction of autophagic mitochondrial turnover (mitophagy).8, 9 However, the relative scale of these effects mediated by endogenous parkin and whether these processes can occur concomitantly or are mutually exclusive, is not known.Bax is a primary effector of cell death that translocates from the cytosol to the mitochondria upon stress, where it facilitates cytochrome c release and the subsequent caspase cascade.10 We previously identified Bax as a parkin substrate, and found that the anti-apoptotic effects of parkin can be directly linked to the parkin-dependent ubiquitination of Bax and inhibition of its mitochondrial translocation.3 Recent corroborative evidence showed that primary cultured neurons from parkin knock-out (KO) mice accumulate greater levels of activated Bax at the mitochondria than wild-type (WT) neurons after apoptotic stimulation,11 while a separate report showed the parkin-dependent ubiquitination of Bax during mitophagy.12In addition to its anti-apoptotic function, parkin facilitates a depolarization-induced and autophagy-dependent turnover of mitochondria. This process is robustly observed in immortalized cell lines expressing human parkin, where exposure to the mitochondrial depolarizing agent carbonyl cyanide 3-chlorophenylhydrazone (CCCP) causes rapid recruitment of parkin from the cytosol to the mitochondrial outer membrane and a coordinated proteasome and autophagosome-mediated turnover of the entire organelle.8, 13, 14, 15 Examination of this process in primary neuronal cultures with endogenous parkin expression, however, has been challenging,16, 17, 18, 19 and a cooperative role for inhibition of mitochondria-dependent cell death has not been investigated in the context of mitophagy.In this study, we sought further insight into the biological functions of parkin across multiple cell types. Our data showed that whole-cell biochemical techniques were not sufficient to observe the participation of endogenous parkin in mitochondrial turnover but were able to confirm the parkin-dependent regulation of apoptosis. Further examination of the parkin-dependent regulation of apoptosis identified two specific lysines of Bax that are critical for recognition and inhibition of its translocation to the mitochondria by parkin. In addition, the BH3 domain of Bax was critical for its association with parkin. Importantly, we observed parkin-dependent mitophagy and inhibition of apoptotic Bax translocation in the same cell culture systems, suggesting that these two pathways coexist and likely cooperate within neurons. Taken together, our data indicate that the parkin-dependent regulation of Bax is critical for cell survival, irrespective of the nature of cell stress involved.  相似文献   

9.
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   

10.
Mcl-1 is an antiapoptotic member of the Bcl-2 family frequently upregulated in non-small cell lung carcinoma (NSCLC). We now report the physiological significance of an interaction between Mcl-1 and the mitochondrial outer membrane-localized voltage-dependent anion channel (VDAC) in NSCLC cell lines. Mcl-1 bound with high affinity to VDAC1 and 3 isoforms but only very weakly to VDAC2 and binding was disrupted by peptides based on the VDAC1 sequence. In A549 cells, reducing Mcl-1 expression levels or application of VDAC-based peptides limited Ca2+ uptake into the mitochondrial matrix, the consequence of which was to inhibit reactive oxygen species (ROS) generation. In A549, H1299 and H460 cells, both Mcl-1 knockdown and VDAC-based peptides attenuated cell migration without affecting cell proliferation. Migration was rescued in Mcl-1 knockdown cells by experimentally restoring ROS levels, consistent with a model in which ROS production drives increased migration. These data suggest that an interaction between Mcl-1 and VDAC promotes lung cancer cell migration by a mechanism that involves Ca2+-dependent ROS production.The Bcl-2 proteins are a family of molecules comprised of both pro- and antiapoptotic members essential for the regulation of apoptotic cell death. In the classical paradigm, the antiapoptotic proteins Bcl-2, Bcl-xL and Mcl-1, inhibit cell death during receipt of apoptotic stimuli by binding and sequestering the proapoptotic members.1 It is now appreciated, however, that in the absence of apoptotic stimuli, Bcl-2 proteins have numerous non-canonical interactions that influence diverse cellular functions, although the precise mechanisms are poorly understood.2 Since antiapoptotic Bcl-2 family members are frequently upregulated in cancer, determining if and how these non-canonical interactions confer survival or other advantages to the cancer cell, will be an important step toward identifying new therapeutic targets. One such interaction is with the outer mitochondrial membrane-localized voltage-dependent anion channel (VDAC), a porin channel with three isoforms that serves as a major diffusion pathway for ions and metabolites,3 and whose gating properties are affected by either Bcl-2 or Bcl-xL binding.4, 5, 6We recently identified an important role for Bcl-xL/VDAC interactions in the regulation of mitochondrial [Ca2+].7 Moving Ca2+ from the cytoplasm to the mitochondrial matrix requires transfer across the outer membrane by VDAC3,8 and across the inner membrane by the Ca2+ uniporter.9 Our studies showed that Bcl-xL interacts with VDAC to facilitate Ca2+ uptake into the mitochondrial matrix. It is not known if other Bcl-2 family members, particularly Bcl-2 and Mcl-1, which are also known VDAC binding partners impart the same physiological regulation on mitochondrial [Ca2+]. Furthermore, the specific physiological consequences and significance of this regulation remain to be determined.Increased production and reduced scavenging of reactive oxygen species (ROS) is frequently observed in cancer cells.10 While excessive ROS levels are toxic, sub-lethal production serves an important signaling function, particularly in cancers, were ROS promote cell proliferation, migration and invasion.11, 12, 13, 14, 15 A primary source of ROS are the mitochondria, and a number of mitochondrial signaling pathways are known to be remodeled and contribute to elevated ROS in cancer cells, including those involved in regulating the electron transport chain (ETC) function and metabolic activity.11,16, 17, 18 It is recognized that upregulation of antiapoptotic Bcl-2 proteins are also associated with a pro-oxidant intracellular environment.19, 20, 21, 22 Mechanistically, they are thought to act at the level of the mitochondria to affect the respiratory chain and increase production of ROS. Since matrix [Ca2+] is an important regulator of mitochondrial metabolism,23,24 and as such, contributes to the regulation of mitochondrial ROS production,25 we reasoned that antiapoptotic Mcl-1/VDAC interactions could promote ROS generation by facilitating matrix Ca2+ uptake.Understanding non-canonical roles of Mcl-1 is an important step toward identifying novel therapeutic targets, particularly in cancers where it is highly expressed, such as in non-small cell lung cancer (NSCLC).26,27 Therefore, we hypothesized that Mcl-1 binding to VDAC promotes mitochondrial Ca2+ uptake and ROS production in NSCLC cells and that this is essential in maintaining the cancer cell phenotype. To test this, we assessed the biochemical interaction between Mcl-1 and VDAC and examined the effects of manipulating Mcl-1 expression levels and Mcl-1/VDAC interactions on mitochondrial Ca2+ uptake, ROS generation and NSCLC cell proliferation and migration.  相似文献   

11.
Bak and Bax mediate apoptotic cell death by oligomerizing and forming a pore in the mitochondrial outer membrane. Both proteins anchor to the outer membrane via a C-terminal transmembrane domain, although its topology within the apoptotic pore is not known. Cysteine-scanning mutagenesis and hydrophilic labeling confirmed that in healthy mitochondria the Bak α9 segment traverses the outer membrane, with 11 central residues shielded from labeling. After pore formation those residues remained shielded, indicating that α9 does not line a pore. Bak (and Bax) activation allowed linkage of α9 to neighboring α9 segments, identifying an α9:α9 interface in Bak (and Bax) oligomers. Although the linkage pattern along α9 indicated a preferred packing surface, there was no evidence of a dimerization motif. Rather, the interface was invoked in part by Bak conformation change and in part by BH3:groove dimerization. The α9:α9 interaction may constitute a secondary interface in Bak oligomers, as it could link BH3:groove dimers to high-order oligomers. Moreover, as high-order oligomers were generated when α9:α9 linkage in the membrane was combined with α6:α6 linkage on the membrane surface, the α6-α9 region in oligomerized Bak is flexible. These findings provide the first view of Bak carboxy terminus (C terminus) membrane topology within the apoptotic pore.Mitochondrial permeabilization during apoptosis is regulated by the Bcl-2 family of proteins.1, 2, 3 Although the Bcl-2 homology 3 (BH3)-only members such as Bid and Bim trigger apoptosis by binding to other family members, the prosurvival members block apoptosis by sequestering their pro-apoptotic relatives. Two remaining members, Bak and Bax, form the apoptotic pore within the mitochondrial outer membrane (MOM).Bak and Bax are globular proteins comprising nine α-helices.4, 5 They are activated by BH3-only proteins binding to the α2–α5 surface groove,6, 7, 8, 9, 10, 11, 12 or for Bax, to the α1/α6 ‘rear pocket''.13 Binding triggers dissociation of the latch domain (α6–α8) from the core domain (α2–α5), together with exposure of N-terminal epitopes and the BH3 domain.6, 7, 14, 15, 16 The exposed BH3 domain then binds to the hydrophobic groove in another Bak or Bax molecule to generate symmetric homodimers.6, 7, 14, 17, 18 In addition to dimerizing, parts of activated Bak and Bax associate with the lipid bilayer.19 In Bax, the α5 and α6 helices may insert into the MOM,20 although recent studies indicate that they lie in-plane on the membrane surface, with the hydrophobic α5 sandwiched between the membrane and a BH3:groove dimer interface.7, 21, 22, 23 The dimers can be linked via cysteine residues placed in α6,18, 24, 25 and more recently via cysteine residues in either α3 or α5,6, 21 allowing detection of the higher-order oligomers associated with pore formation.26, 27 However, whether these interactions are required for high-order oligomers and pore formation remains unclear.Like most Bcl-2 members, Bak and Bax are targeted to the MOM via a hydrophobic C-terminal region. The C terminus targets Bak to the MOM in healthy cells,28 whereas the Bax C terminus is either exposed29 or sequestered within the hydrophobic groove until apoptotic signals trigger Bax translocation.5, 30, 31 The hydrophobic stretch is important, as substituting polar or charged residues decreased targeting of Bak and Bax.10, 32 Mitochondrial targeting is also controlled by basic residues at the far C termini,32, 33, 34 and by interaction with VDAC235, 36 via the Bak and Bax C termini.37, 38 Retrotranslocation of Bak and Bax was also altered by swapping the C termini.39The membrane topology of the Bak and Bax C termini before and after apoptosis has not been examined directly, due in part to difficulty in reconstituting oligomers of full-length Bak in artificial membranes. Nor is it known whether the C termini contribute to pore formation by promoting oligomerization or disturbing the membrane. To address these questions synthetic peptides based on the Bak and Bax C termini have been studied in model membranes. The peptides adopt a predominantly α-helical secondary structure,40, 41, 42, 43 with orientation affected by lipid composition.42, 44, 45 The peptides could also permeabilize lipid vesicles,41, 43, 46, 47 suggesting that the C termini in full-length Bak and Bax may contribute to pore formation.Here we examined the membrane topology of the C termini within full-length Bak and Bax in the MOM, both before and after apoptotic pore formation. After pore formation the α9 helices of Bak (and of Bax) became juxtaposed but did not line the surface of a pore. The α9:α9 interaction occurred after Bak activation and conformation change, but was promoted by formation of BH3:groove dimers. Combining linkage at more than one interface indicated that the Bak α9:α9 interface can link BH3:groove dimers to high-order oligomers, and moreover, that the α6–α9 region is flexible in oligomerized Bak.  相似文献   

12.
13.
14.
Well-balanced mitochondrial fission and fusion processes are essential for nervous system development. Loss of function of the main mitochondrial fission mediator, dynamin-related protein 1 (Drp1), is lethal early during embryonic development or around birth, but the role of mitochondrial fission in adult neurons remains unclear. Here we show that inducible Drp1 ablation in neurons of the adult mouse forebrain results in progressive, neuronal subtype-specific alterations of mitochondrial morphology in the hippocampus that are marginally responsive to antioxidant treatment. Furthermore, DRP1 loss affects synaptic transmission and memory function. Although these changes culminate in hippocampal atrophy, they are not sufficient to cause neuronal cell death within 10 weeks of genetic Drp1 ablation. Collectively, our in vivo observations clarify the role of mitochondrial fission in neurons, demonstrating that Drp1 ablation in adult forebrain neurons compromises critical neuronal functions without causing overt neurodegeneration.In addition to their crucial importance in energy conversion, mitochondria serve many other housekeeping functions, including calcium buffering, amino-acid and steroid biosynthesis as well as fatty acids beta-oxidation and regulation of cell death. During the past decade, it has become increasingly clear that processes regulating mitochondrial morphology and ultrastructure are influenced by specific cellular requirements upon which mitochondria, in a precisely regulated manner, undergo fusion and division events.1 Maintaining this balance is especially important for highly energy-consuming, polarized cells such as neurons, where single organellar units sprouting from the mitochondrial network are transported along the cytoskeleton into dendrites and spines to meet local energy requirements.2 In addition, elaborate quality-control mechanisms also rely on mitochondrial dynamics: whereas defective organelles are sequestered by fission, enabling their removal from the mitochondrial network,3, 4 fusion supports qualitative homogeneity of the syncytium through complementation.5Mitochondrial fusion and fission are mediated by large GTPases of the dynamin superfamily.6 The outer mitochondrial membrane mitofusins 1 (MFN1) and 2 (MFN2) tether mitochondrial membranes by homodimer or heterodimer formation,7 thereby initiating fusion of the organelles, a process that also involves the inner mitochondrial membrane-associated GTPase Optic Atrophy 1.8 In addition, MFN2 also mediates contacts between mitochondria and endoplasmic reticulum.9 The only known mammalian mitochondrial fission protein, Dynamin-Related Protein 1 (Drp1), translocates upon dephosphorylation by calcineurin10 to fission sites where it binds to mitochondrial fission factor.11 Drp1 translocation is preceded by ER membranes wrapping around mitochondria to constrict the organelles,12 thereby facilitating the formation of multimeric Drp1 complexes that, upon GTP hydrolysis, further tighten to complete the process of mitochondrial fission.13Genetic evidence in mice and humans indicates that mitochondrial dynamics are crucially important in neurons: in humans, a sporadic dominant-negative DRP1 mutation caused a lethal syndromic defect with abnormal brain development;14 similarly, constitutive Drp1 knockout in the mouse brain leads to lethal neurodevelopmental defects.15, 16 Although the crucial role of Drp1 during brain development is undisputed, studies on Drp1 function in postmitotic (adult) neurons are scarce; likewise, Drp1 ablation studies in primary cultures have so far failed to yield a conclusive picture. In vitro, Drp1 ablation is reported to lead to a super-elongated neuroprotective17, 18, 19, 20, 21, 22, 23, 24 or an aggregated mitochondrial phenotype associated with neurodegeneration.15, 16, 25, 26, 27 These discrepancies are probably due to different experimental conditions: neuronal health is indeed influenced by the onset and duration of Drp1 inhibition, which varies considerably among the cited reports,28 and different types of neuronal cultures studied display different sensitivity to Drp1 inhibition. In vivo, Drp1 ablation in Purkinje cells results in oxidative stress and neurodegeneration,29 demonstrating that Drp1 is essential for postmitotic neurons'' health. In contrast, transient pharmacological Drp1 inhibition is neuroprotective in several mouse ischemia models, indicating that temporarily blocking mitochondrial fission holds therapeutic potential.30, 31, 32To elucidate the consequences of blocked mitochondrial fission in the central nervous system in vivo, we bypassed the critical role of Drp1 during brain development by generating Drp1flx/flx mice15 expressing tamoxifen-inducible Cre recombinase under the control of the CaMKIIα promoter.33 Upon induced Drp1 deletion in postmitotic adult mouse forebrain neurons, mice develop progressive, neuronal subtype-specific alterations in mitochondrial shape and distribution in the absence of overt neurodegeneration. In addition, respiratory capacity, ATP content, synaptic reserve pool vesicle recruitment as well as spatial working memory are impaired, demonstrating that severely dysregulated mitochondrial dynamics can compromise critical neuronal functions in vivo without causing neuronal cell death.  相似文献   

15.
16.
Tumor heterogeneity is in part determined by the existence of cancer stem cells (CSCs) and more differentiated tumor cells. CSCs are considered to be the tumorigenic root of cancers and suggested to be chemotherapy resistant. Here we exploited an assay that allowed us to measure chemotherapy-induced cell death in CSCs and differentiated tumor cells simultaneously. This confirmed that CSCs are selectively resistant to conventional chemotherapy, which we revealed is determined by decreased mitochondrial priming. In agreement, lowering the anti-apoptotic threshold using ABT-737 and WEHI-539 was sufficient to enhance chemotherapy efficacy, whereas ABT-199 failed to sensitize CSCs. Our data therefore point to a crucial role of BCLXL in protecting CSCs from chemotherapy and suggest that BH3 mimetics, in combination with chemotherapy, can be an efficient way to target chemotherapy-resistant CSCs.Colorectal cancer is the third most common cancer worldwide.1, 2 Patients with advanced stage colorectal cancer are routinely treated with 5-fluorouracil (5-FU), leucovorin and oxaliplatin (FOLFOX), or with 5-FU, leucovorin and irinotecan (FOLFIRI), often in combination with targeted agents such as anti-VEGF or anti-EGFR at metastatic disease.3, 4, 5, 6 Despite this intensive treatment, therapy is still insufficiently effective and chemotherapy resistance occurs frequently. Although still speculative, it has been suggested that unequal sensitivity to chemotherapy is due to an intratumoral heterogeneity that is orchestrated by cancer stem cells (CSCs) that can self-renew and give rise to more differentiated progeny.7, 8 When isolated from patients, CSCs efficiently form tumors upon xenotransplantation into mice which resemble the primary tumor from which they originated.9, 10, 11 In addition, many xenotransplantation studies have emphasized the importance of CSCs for tumor growth.9, 10, 11, 12 Colon CSCs were originally isolated from primary human colorectal tumor specimens using CD133 as cell surface marker and shown to be highly tumorigenic when compared with the non-CSCs population within a tumor.9, 10 Later, other cell surface markers as well as the activity of the Wnt pathway have been used to isolate colon CSCs from tumors.12, 13 Spheroid cultures have been established from human primary colorectal tumors that selectively enrich for the growth of colon CSCs,11, 12 although it is important to realize that these spheres also contain more differentiated tumor cells.12 In agreement, we have shown that the Wnt activity reporter that directs the expression of enhanced green fluorescent protein (TOP-GFP) allows for the separation of CSCs from more differentiated progeny in the spheroid cultures.12CSCs are suggested to be responsible for tumor recurrence after initial therapy, as they are considered to be selectively resistant to therapy.11, 14 Conventional chemotherapy induces, among others, DNA damage and subsequent activation of the mitochondrial cell death pathway, which is regulated by a balance between pro- and anti-apoptotic BCL2 family members.15 Upon activation of apoptosis, pro-apoptotic BH3 molecules are activated and these may perturb the balance in favor of apoptosis initiated by mitochondrial outer membrane polarization (MOMP), release of cytochrome c and subsequent activation of a caspase cascade.The apoptotic balance of cancer cells can be measured with the use of a functional assay called BH3 profiling.16 BH3 profiling is a method to determine the apoptotic ‘priming'' level of a cell by exposing mitochondria to standardized amounts of roughly 20-mer peptides derived from the alpha-helical BH3 domains of BH3-only proteins and determining the rate of mitochondrial depolarization. Using this approach, priming was measured in various cancers and compared with normal tissues.17, 18 In all cancer types tested, the mitochondrial priming correlated well with the observed clinical response to chemotherapy. That is, cancers that are highly primed are more chemosensitive, whereas chemoresistant tumor cells and normal tissues are poorly primed.17, 18 This suggests that increasing mitochondrial priming can potentially increase chemosensitivity, which can be achieved by directly inhibiting the anti-apoptotic BCL2 family members.18 To this end, small-molecule inhibitors, so-called BH3 mimetics, have been developed. ABT-737 and the highly related ABT-263 both inhibit BCL2, BCLXL and BCLW19, 20, 21 and were shown to be effective in killing cancer cells in vitro and in vivo21 with a preference for BCL2.19, 22 As BCL2 protein expression is often upregulated in hematopoietic cancers, it represents a promising target, which was supported by high efficacy of these BH3 mimetics in animal experiments.21 However, in vivo efficacy is limited due to thrombocytopenia, which relates to a dependence of platelets on BCLXL for survival.23, 24 To overcome this toxicity, a BCL2-specific compound, ABT-199, was developed.25 Souers et al.25 showed that inhibition of BCL2 using ABT-199 blocks tumor growth of acute lymphoblastic leukemia cells in xenografts. In addition to the single compound effects of ABT-199, combination with rituximab inhibited growth of non-Hodgkin''s lymphoma, mantle cell lymphoma and acute lymphoblastic leukemia tumor cells growth in vivo.25 Moreover, highly effective tumor lysis was observed in all three patients with chronic lymphocytic leukemia that were treated with ABT-199.25 More recently, a BCLXL-specific compound, WEHI-539, was discovered using high-throughput chemical screening.26As the apoptotic balance appears a useful target for the treatment of cancers and CSCs have been suggested to resist therapy selectively, we set out to analyze whether specifically colon CSCs are resistant to therapy and whether this is due to an enhanced anti-apoptotic threshold, specific to CSCs. To study chemosensitivity, we developed a robust single cell-based analysis in which we can measure apoptosis simultaneously in CSCs and their differentiated progeny. Utilizing this system we showed that colon CSCs and not their differentiated progeny are resistant to chemotherapeutic compounds and that this was due to the fact that these cells are less primed to mitochondrial death. Furthermore, inhibition of anti-apoptotic BCLXL molecule with either ABT-737 or WEHI-539, but not ABT-199, breaks this resistance and sensitizes the CSCs to chemotherapy.  相似文献   

17.
18.
Evidence indicates that nitrosative stress and mitochondrial dysfunction participate in the pathogenesis of Alzheimer''s disease (AD). Amyloid beta (Aβ) and peroxynitrite induce mitochondrial fragmentation and neuronal cell death by abnormal activation of dynamin-related protein 1 (DRP1), a large GTPase that regulates mitochondrial fission. The exact mechanisms of mitochondrial fragmentation and DRP1 overactivation in AD remain unknown; however, DRP1 serine 616 (S616) phosphorylation is likely involved. Although it is clear that nitrosative stress caused by peroxynitrite has a role in AD, effective antioxidant therapies are lacking. Cerium oxide nanoparticles, or nanoceria, switch between their Ce3+ and Ce4+ states and are able to scavenge superoxide anions, hydrogen peroxide and peroxynitrite. Therefore, nanoceria might protect against neurodegeneration. Here we report that nanoceria are internalized by neurons and accumulate at the mitochondrial outer membrane and plasma membrane. Furthermore, nanoceria reduce levels of reactive nitrogen species and protein tyrosine nitration in neurons exposed to peroxynitrite. Importantly, nanoceria reduce endogenous peroxynitrite and Aβ-induced mitochondrial fragmentation, DRP1 S616 hyperphosphorylation and neuronal cell death.Nitric oxide (NO) is a neurotransmitter and neuromodulator required for learning and memory.1 NO is generated by NO synthases, a group of enzymes that produce NO from L-arginine. In addition to its normal role in physiology, NO is implicated in pathophysiology. When overproduced, NO combines with superoxide anions (O2·), byproducts of aerobic metabolism and mitochondrial oxidative phosphorylation, to form peroxynitrite anions (ONOO) that are highly reactive and neurotoxic. Accumulation of these reactive oxygen species (ROS) and reactive nitrogen species (RNS), known as oxidative and nitrosative stress, respectively, is a common feature of aging, neurodegeneration and Alzheimer''s disease (AD).1Nitrosative stress caused by peroxynitrite has a critical role in the etiology and pathogenesis of AD.2, 3, 4, 5, 6, 7 Peroxynitrite is implicated in the formation of the two hallmarks of AD, Aβ aggregates and neurofibrillary tangles containing hyperphosphorylated Tau protein.1, 4, 7 In addition, peroxynitrite promotes the nitrotyrosination of presenilin 1, the catalytic subunit of the γ-secretase complex, which shifts production of Aβ to amyloid beta (Aβ)42 and increases the Aβ42/Aβ40 ratio, ultimately resulting in an increased propensity for aggregation and neurotoxicity.5 Furthermore, nitration of Aβ tyrosine 10 enhances its aggregation.6 Peroxynitrite can also modify enzymes, such as triosephosphate isomerase,4 and activate kinases, including Jun amino-terminal kinase and p38 mitogen-activated protein kinase, which enhance neuronal cell death.8, 9 Moreover, peroxynitrite can trigger the release of free metals such as Zn2+ from intracellular stores with consequent inhibition of mitochondrial function and enhancement of neuronal cell death.10, 11, 12 Finally, peroxynitrite can irreversibly inhibit complexes I and IV of the mitochondrial respiratory chain.11, 13Because mitochondria have a critical role in neurons as energy producers to fuel vital processes such as synaptic transmission and axonal transport,14 and mitochondrial dysfunction is a well-documented and early event in AD,15 it is important to consider how peroxynitrite and nitrosative stress affect mitochondria. Although the ultimate cause of mitochondrial dysfunction in AD remains unclear, an imbalance in mitochondrial fission and fusion is one possibility.1, 14, 16, 17, 18 Notably, peroxynitrite, N-methyl D-aspartate (NMDA) receptor activation and Aβ can induce mitochondrial fragmentation by activating mitochondrial fission and/or inhibiting fusion.16 Mitochondrial fission and fusion is regulated by large GTPases of the dynamin family, including dynamin-related protein 1 (DRP1) that is required for mitochondrial division,19 and inhibition of mitochondrial division by overexpression of the GTPase-defective DRP1K38A mutant provides protection against peroxynitrite-, NMDA- and Aβ-induced mitochondrial fragmentation and neuronal cell death.16The exact mechanism of peroxynitrite-induced mitochondrial fragmentation remains unclear. A recent report suggested that S-nitrosylation of DRP1 at cysteine 644 increases DRP1 activity and is the cause of peroxynitrite-induced mitochondrial fragmentation in AD;20 however, the work remains controversial, suggesting that alternative pathways might be involved.21 For example, peroxynitrite also causes rapid DRP1 S616 phosphorylation that promotes its translocation to mitochondria and organelle division.21, 22 In mitotic cells, DRP1 S616 phosphorylation is mediated by Cdk1/cyclinB1 and synchronizes mitochondrial division with cell division.23 Interestingly, DRP1 is S616 hyperphosphorylated in AD brains, suggesting that this event might contribute to mitochondrial fragmentation in the disease.21, 22 A recent report indicates that Cdk5/p35 is responsible for DRP1 S616 phosphorylation,24 and notably aberrant Cdk5/p35/p25 signaling is associated with AD pathogenesis.25 Thus, we explored here the possible role of DRP1 S616 hyperphosphorylation in Aβ- and peroxynitrite-mediated mitochondrial fragmentation.Under normal conditions, accumulated mitochondrial superoxide anions and hydrogen peroxide (H2O2) can be neutralized by superoxide dismutase (SOD) and catalase. Nitrosative stress in aging and AD might be explained by a loss of antioxidant enzymes. Previous studies suggest that expression of SOD subtypes is decreased in the human AD brain.26, 27 Furthermore, SOD1 deletion in a mouse model of AD increased the burden of amyloid plaques.26 By contrast, overexpression of SOD2 in a mouse model of AD decreased the Aβ42/Aβ40 ratio and alleviated memory deficits.28, 29 There is currently a lack of antioxidants that can effectively quench superoxide anions, H2O2 or peroxynitrite and provide lasting effects. Cerium is a rare earth element and cerium oxide (CeO2) nanoparticles, or nanoceria, shuttle between their 3+ or 4+ states. Oxidation of Ce4+ to Ce3+ causes oxygen vacancies and defects on the surface of the crystalline lattice structure of the nanoparticles, generating a cage for redox reactions to occur.30 Accordingly, nanoceria mimic the catalytic activities of antioxidant enzymes, such as SOD31, 32 and catalase,33 and are able to neutralize peroxynitrite.34 Because of these antioxidant properties, we hypothesized that nanoceria could detoxify peroxynitrite and protect against Aβ-induced DRP1 S616 hyperphosphorylation, mitochondrial fragmentation and neuronal cell death.  相似文献   

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.
Mutations in the leucine-rich repeat kinase 2 (LRRK2) gene have been associated with Parkinson''s disease, and its inhibition opens potential new therapeutic options. Among the drug inhibitors of both wild-type and mutant LRRK2 forms is the 2-arylmethyloxy-5-subtitutent-N-arylbenzamide GSK257815A. Using the well-established dopaminergic cell culture model SH-SY5Y, we have investigated the effects of GSK2578215A on crucial neurodegenerative features such as mitochondrial dynamics and autophagy. GSK2578215A induces mitochondrial fragmentation of an early step preceding autophagy. This increase in autophagosome results from inhibition of fusion rather than increases in synthesis. The observed effects were shared with LRRK2-IN-1, a well-described, structurally distinct kinase inhibitor compound or when knocking down LRRK2 expression using siRNA. Studies using the drug mitochondrial division inhibitor 1 indicated that translocation of the dynamin-related protein-1 has a relevant role in this process. In addition, autophagic inhibitors revealed the participation of autophagy as a cytoprotective response by removing damaged mitochondria. GSK2578215A induced oxidative stress as evidenced by the accumulation of 4-hydroxy-2-nonenal in SH-SY5Y cells. The mitochondrial-targeted reactive oxygen species scavenger MitoQ positioned these species as second messengers between mitochondrial morphologic alterations and autophagy. Altogether, our results demonstrated the relevance of LRRK2 in mitochondrial-activated pathways mediating in autophagy and cell fate, crucial features in neurodegenerative diseases.Nowadays, Parkinson''s disease (PD) constitutes the main motor disorder and the second neurodegenerative disease after Alzheimer''s disease. Etiology of PD remains unknown, but both environmental and genetic factors have been implicated. Among the genes associated with PD is the leucine-rich repeat kinase 2 (LRRK2, PARK8, OMIM 607060) encoding gene encoded by PARK8. Indeed, LRRK2 mutations have been described in a substantial number of idiopathic late-onset PD patients without a known family history of the disease.1, 2, 3The physiologic function remains unknown. It localizes in the cytosol as well as in specific membrane subdomains, including mitochondria, autophagosomes and autolysosomes,4 and interacts with a whole array of proteins, including both α- and β-tubulin,5, 6 tau,7 α-synuclein8 and F-actin.9 LRRK2 gene mutations, including the most common G2019S,3 are associated with increases in toxic putative kinase activity.1, 10 LRRK2 overexpression is toxic to cultured cells,11, 12 and LRRK2 loss did not cause neurodegenerative changes (for a review see Tong and Shen13). However, LRRK2 transgenic mice lack obvious PD-like behavioral phenotypes.14 LRRK2-associated PD patients show degeneration of dopaminergic neurons in the substantia nigra.15 Data from our own group and others have associated mitochondrial apoptotical pathways with PD,16, 17, 18 and, in this context, LRRK2 mutant-mediated toxicity could be due to mitochondria-dependent apoptosis.19 There is considerable evidence for impaired mitochondrial function and morphology in both early-onset, autosomal recessive inherited PD and late-onset sporadic PD.Mitochondrial dynamics include several mechanisms, such as fission, fusion and mitophagy.20, 21 Altered fission/fusion dynamics might be a common pathogenic pathway of neurodegenerative diseases. It is well documented that mitochondrial dynamics constitute a relevant issue in some experimental neurodegenerative models.20, 22, 23, 24, 25 Mitochondrial dynamics is tightly regulated by cellular pathways including those participated by the dynamin-related protein-1 (Drp1). Drp1 mostly locates in the cytoplasm, but is stimulated after fission stimuli to migrate to the mitochondria. Once there, Drp1 forms ring-like structures, which wrap around the scission site to constrict the mitochondrial membrane resulting in mitochondrial fission.26, 27 Interestingly, a functional interaction between PD-associated LRRK2 and members of the dynamin GTPase superfamily has been described.28Macroautophagy (hereafter referred to as autophagy) is an active cellular response, which functions in the intracellular degradation system of cellular debris such as damaged organelles. Whether autophagy promotes cell death or enhances survival is still controversial.29, 30 It requires the formation of autophagosomes where cellular content is to be degraded by the action of lysosomal enzymatic content. Autophagosome formation is regulated by an orderly action of >30 autophagy-related (Atg) proteins. Among them is the microtubule-associated protein 1A/1B-light chain 3 (LC3), a homolog of Apg8p, which is essential for autophagy in yeast and is associated with autophagosome membranes.31 Interestingly, these vesicles are mostly highly mobile in the cytoplasm.32 Wild-type and mutant LRRK2 expression has been related to autophagy.4, 33, 34, 35, 36 Reactive oxygen species (ROS) function as relevant second messengers after several stimuli, including mitochondrial disruption. Exacerbated ROS increases might result in overactivation of antioxidant systems and yield harmful oxidative stress. Among oxidative stress hallmarks is the accumulation of α,β-unsaturated hydroxyalkenal 4-hydroxy-2-nonenal (4-HNE), whose accumulation has been reported in PD post-mortem patient brains,37, 38 thus giving a significant relevance to ROS in the pathogenesis of PD.All these results indicate LRRK2 as a promising pharmacologic target in PD treatment.39 Several LRRK2 inhibitor drugs have been synthetized, such as the potent and highly selective 2-arylmethyloxy-5-substitutent-N-arylbenzamide (GSK2578215A). GSK2578215A exhibits biochemical IC50s of 10.9 nM against wild-type LRRK2, and possesses a high ratio of brain to plasma distribution.40 This study provides key insights into the mechanisms downstream of LRRK2 inhibition, and spreads light onto an underexplored, yet potentially tractable therapeutic target for treating LRRK2-associated PD. We demonstrate how inhibition of this kinase results in the activation of cellular death pathways such as the mitochondrial fission machinery, and how cells reply by activating a protective autophagic response. Our results show the presence of oxidative stress hallmarks, thus pointing to a key function for ROS, placed downstream of mitochondrial fission.  相似文献   

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