首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
Massive Ca2+ influx into mitochondria is critically involved in cell death induction but it is unknown how this activates the organelle for cell destruction. Using multiple approaches including subcellular fractionation, FRET in intact cells, and in vitro reconstitutions, we show that mitochondrial Ca2+ influx prompts complex II of the respiratory chain to disintegrate, thereby releasing an enzymatically competent sub-complex that generates excessive reactive oxygen species (ROS) for cell death induction. This Ca2+-dependent dissociation of complex II is also observed in model membrane systems, but not when cardiolipin is replaced with a lipid devoid of Ca2+ binding. Cardiolipin is known to associate with complex II and upon Ca2+ binding coalesces into separate homotypic clusters. When complex II is deprived of this lipid, it disintegrates for ROS formation and cell death. Our results reveal Ca2+ binding to cardiolipin for complex II disintegration as a pivotal step for oxidative stress and cell death induction.Cellular calcium ion (Ca2+) overload is known to be of fundamental importance in pathological cell death induction for instance during brain ischemia, ischemia-reperfusion of the heart, and excitotoxicity of neurons.1 Upon entering the cytosol from the extracellular space, Ca2+ ions accumulate in mitochondria at very high levels. An alternative route into mitochondria, observed during many scenarios of cell death, as well as when therapeutically induced by anticancer agents, is through Ca2+ release from the ER. After crossing the ER-mitochondrial junction, the ion is taken up by the mitochondrial calcium uniporter.2, 3 The close apposition of the two organelles ensures that a very high Ca2+ concentration can be reached in mitochondria.4 The direct target of mitochondrial Ca2+ influx for cell death induction, however, is unknown.Cells deficient in complex II of the respiratory chain become resistant to many cell death signals.5 The ability of this complex to produce deleterious amounts of reactive oxygen species (ROS) has been recognized.6, 7 Initial in vitro experiments using blue native gels indicated that during cell death, the sub-complex SDHA/SDHB, which remains enzymatically active,8 is specifically released from the membrane-anchoring SDHC and SDHD complex II subunits.9 It can then remove electrons from the substrate succinate and transfer them to molecular oxygen to generate ROS for cell death induction.5, 9The principal lipid in the inner mitochondrial membrane that harbors the components of the respiratory chain, including complex II, is the diphosphatidylglycerol cardiolipin. This lipid is known to be involved in cell death, although its effects have been connected mostly with cellular sites different from its most prominent residence.10, 11, 12In this study, we investigated whether excessive Ca2+ influx into mitochondria can affect on the integrity of complex II and activate this complex for cell death.  相似文献   

3.
Mitochondrial disorders are a group of pathologies characterized by impairment of mitochondrial function mainly due to defects of the respiratory chain and consequent organellar energetics. This affects organs and tissues that require an efficient energy supply, such as brain and skeletal muscle. They are caused by mutations in both nuclear- and mitochondrial DNA (mtDNA)-encoded genes and their clinical manifestations show a great heterogeneity in terms of age of onset and severity, suggesting that patient-specific features are key determinants of the pathogenic process. In order to correlate the genetic defect to the clinical phenotype, we used a cell culture model consisting of fibroblasts derived from patients with different mutations in the mtDNA-encoded ND5 complex I subunit and with different severities of the illness. Interestingly, we found that cells from patients with the 13514A>G mutation, who manifested a relatively late onset and slower progression of the disease, display an increased autophagic flux when compared with fibroblasts from other patients or healthy donors. We characterized their mitochondrial phenotype by investigating organelle turnover, morphology, membrane potential and Ca2+ homeostasis, demonstrating that mitochondrial quality control through mitophagy is upregulated in 13514A>G cells. This is due to a specific downregulation of mitochondrial Ca2+ uptake that causes the stimulation of the autophagic machinery through the AMPK signaling axis. Genetic and pharmacological manipulation of mitochondrial Ca2+ homeostasis can revert this phenotype, but concurrently decreases cell viability. This indicates that the higher mitochondrial turnover in complex I deficient cells with this specific mutation is a pro-survival compensatory mechanism that could contribute to the mild clinical phenotype of this patient.Mitochondrial disorders include a wide range of pathological conditions characterized by defects in organelle homoeostasis and energy metabolism, in particular in the electron transport chain (ETC) complexes. They are mostly caused by mutations in nuclear- or mtDNA-encoded genes of the respiratory chain complexes leading to a variety of clinical manifestations, ranging from lesions in specific tissues, such as in Leber''s hereditary optic neuropathy, to complex multisystem syndromes, such as myoclonic epilepsy with ragged-red fibers, Leigh syndrome or the mitochondrial encephalomyopathy, lactic acidosis and stroke-like episodes syndrome (MELAS).1, 2 Despite the detailed knowledge of the molecular defects in these diseases, their pathogenesis remains poorly understood. The heterogeneity of signs and symptoms depends on the diversity of the genetic background and on patient-specific compensatory mechanisms. Several studies investigated the consequences of nuclear DNA mutations on intracellular organelle physiology and Ca2+ homeostasis.3, 4 Here we analyzed a cohort of patients with mutations in the mtDNA-encoded ND5 subunit of NADH dehydrogenase in order to correlate the clinical phenotype with relevant intracellular parameters involved in mitochondrial physiology, such as the rate of autophagy and mitophagy fluxes, mitochondrial Ca2+ dynamics, mitochondrial membrane potential and their functional relationship.Mitochondrial Ca2+ is a key regulator of organelle physiology, and impairment of cation homeostasis is a general feature of many pathological conditions, including mitochondrial diseases.5 In addition, Ca2+ uptake in this organelle has recently been demonstrated to be a fundamental regulator of autophagy.6, 7 Autophagy is involved in physiological organelle turnover and in the removal of damaged or non-functional mitochondria by autophagy (called ‘mitophagy'')8, 9, 10, 11 and is critical for organelle quality control. Given the pivotal role of mitochondrial Ca2+ in the adaptation of adenosine triphosphate (ATP) production to cellular energy demand, the recent identification of the channel responsible for Ca2+ entry into the organelle, the mitochondrial Ca2+ uniporter (MCU), is instrumental for the understanding of the regulation of mitochondrial Ca2+ transport in both physiological and pathological conditions. MCU was identified in 2011,12, 13 and in the following years, molecular insight on its complex regulatory mechanism was obtained. The pore region is composed of MCU, its isoform MCUb14 and essential MCU regulator (EMRE).15 The channel is gated by the Ca2+-sensitive proteins mitochondrial Ca2+ uptake 1 (MICU1) and MICU216, 17, 18, 19 and further regulated by the SLC25A23 protein.20 As to its cellular function, mitochondrial Ca2+ has been shown to stimulate ATP production by positive regulation of three key dehydrogenases of the tricarboxylic acid cycle21 and of the ETC.22 In parallel, unregulated and sustained organelle Ca2+ overload can also lead to the opening of the mitochondrial permeability transition pore,23, 24 with consequent dissipation of mitochondrial membrane potential (ΔΨmt), release of caspase cofactors and activation of the apoptotic cascade.5 Despite the significant molecular understanding of all these cellular processes, their role in the pathogenesis of mitochondrial diseases is still poorly understood. Here we investigated the interplay of these pathways and the possibility of their contribution to determine the severity of the pathology in a cellular model consisting of fibroblasts from patients carrying mutations in the mitochondrial ND5 gene.  相似文献   

4.
Mitochondrial alterations are critically involved in increased vulnerability to disease during aging. We investigated the contribution of mitochondria–sarcoplasmic reticulum (SR) communication in cardiomyocyte functional alterations during aging. Heart function (echocardiography) and ATP/phosphocreatine (NMR spectroscopy) were preserved in hearts from old mice (>20 months) with respect to young mice (5–6 months). Mitochondrial membrane potential and resting O2 consumption were similar in mitochondria from young and old hearts. However, maximal ADP-stimulated O2 consumption was specifically reduced in interfibrillar mitochondria from aged hearts. Second generation proteomics disclosed an increased mitochondrial protein oxidation in advanced age. Because energy production and oxidative status are regulated by mitochondrial Ca2+, we investigated the effect of age on mitochondrial Ca2+ uptake. Although no age-dependent differences were found in Ca2+ uptake kinetics in isolated mitochondria, mitochondrial Ca2+ uptake secondary to SR Ca2+ release was significantly reduced in cardiomyocytes from old hearts, and this effect was associated with decreased NAD(P)H regeneration and increased mitochondrial ROS upon increased contractile activity. Immunofluorescence and proximity ligation assay identified the defective communication between mitochondrial voltage-dependent anion channel and SR ryanodine receptor (RyR) in cardiomyocytes from aged hearts associated with altered Ca2+ handling. Age-dependent alterations in SR Ca2+ transfer to mitochondria and in Ca2+ handling could be reproduced in cardiomyoctes from young hearts after interorganelle disruption with colchicine, at concentrations that had no effect in aged cardiomyocytes or isolated mitochondria. Thus, defective SR–mitochondria communication underlies inefficient interorganelle Ca2+ exchange that contributes to energy demand/supply mistmach and oxidative stress in the aged heart.Age is the main independent risk factor for cardiovascular morbidity and mortality.1 It increases heart vulnerability to cardiac diseases as well as the severity of their clinical manifestations, and reduces the efficacy of cardioprotective interventions.2 At the cellular level, some of the structural and functional age-dependent changes resemble those of failing cardiac myocytes.3, 4 Specifically, disturbed Ca2+ homeostasis and excitation–contraction coupling,5 as well as deficient mitochondrial energetics6 and excessive ROS production,7 have been consistently reported in senescent cardiomyocytes. These subcellular alterations likely contribute to the reduced adaptive capacity to stress (exercise, β-adrenergic stimulation) and increased vulnerability to disease of the aged hearts.In cardiac cells, electrochemical coupling and metabolic adaptations are based upon the coordination between sarcoplasmic reticulum (SR) and mitochondria tightly interconnected forming an interface to support local ionic exchange and signal transduction in a beat-to-beat basis.8 This privileged interorganelle communication facilitates mitochondrial ATP transport for SR Ca2+ cycling and ensures energy replenishment by reciprocal Ca2+ and ADP exchange. Ca2+ is taken up by mitochondria using a low-affinity uniporter whose activity is driven by the elevated Ca2+ concentration in the microenvironment present around ryanodine receptors (RyR).9 Indeed, the kinetics of mitochondrial Ca2+ uptake is more dependent on the concentration of Ca2+ at the SR–mitochondria contact points than on bulk cytosolic Ca2+ concentration.8 Mitochondrial Ca2+ uptake allows energy supply–demand matching through the activation of Krebs cycle dehydrogenases and electron transport chain activity, and at the same time it regulates the regeneration of Krebs-coupled antioxidative defenses (NAD(P)H).10Defective SR–mitochondria cross talk has been causally linked to the abnormal mitochondrial Ca2+ uptake in failing hearts and may underlie their increased oxidative stress.11 Also, in diabetic cardiomyopathy, intracellular Ca2+ overload and depletion of energy stores appear to develop as a consequence of sequential SR–mitochondria dysfunction.12 Atrial fibrillation has been associated with an increased fusion of mitochondria and a subsequent increased colocalization of giant mitochondria with SR, a subcellular remodeling process that contributes to the perpetuation of the arrhythmia.13 Because mitochondria are highly dynamic structures, some molecular links have been proposed to provide a stable physical interorganelle bridge14, 15 while others appear to facilitate direct tunneling of Ca2+ and other signaling mediators.16 In the present study, we hypothesized that aging may negatively impact on mitochondria–SR communication by mechanisms involving defective Ca2+ transmission, and we identified reduced physical interaction between RyR and mitochondrial voltage-dependent anion channel (VDAC) as the main responsible of this effect.  相似文献   

5.
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   

6.
Globoid cell leukodystrophy (GLD) is a metabolic disease caused by mutations in the galactocerebrosidase (GALC) gene. GALC is a lysosomal enzyme whose function is to degrade galacto-lipids, including galactosyl-ceramide and galactosyl-sphingosine (psychosine, PSY). GALC loss of function causes progressive intracellular accumulation of PSY. It is widely held that PSY is the main trigger for the degeneration of myelinating cells and progressive white-matter loss. However, still little is known about the molecular mechanisms by which PSY imparts toxicity. Here, we address the role of calcium dynamics during PSY-induced cell death. Using the human oligodendrocyte cell line MO3.13, we report that cell death by PSY is accompanied by robust cytosolic and mitochondrial calcium (Ca2+) elevations, and by mitochondrial reactive oxygen species (ROS) production. Importantly, we demonstrate that the reduction of extracellular calcium content by the chelating agent ethylenediaminetetraacetic acid can decrease intra-mitochondrial ROS production and enhance cell viability. Antioxidant administration also reduces mitochondrial ROS production and cell loss, but this treatment does not synergize with Ca2+ chelation. Our results disclose novel intracellular pathways involved in PSY-induced death that may be exploited for therapeutic purposes to delay GLD onset and/or slow down its progression.Globoid cell leukodystrophy (GLD), also known as Krabbe disease, is a childhood leukodystrophy triggered by mutations in the galactocerebrosidase (GALC) gene; the physio-pathological hallmarks of GLD are progressive demyelination, reactive astrocytosis and microgliosis.1 GALC is a lysosomal enzyme essential for the normal catabolism of galacto-lipids, including galactosyl-ceramide and galactosyl-sphingosine (psychosine, PSY). GALC loss of function causes progressive accumulation of PSY, a cytotoxic metabolite that has been assumed as the main cause for GLD pathogenesis.2 PSY leads to Schwann cell and oligodendrocyte death, but still little is known about the molecular mechanisms by which PSY imparts toxicity. It has been demonstrated that PSY accumulates in cell membrane raft micro-domains, disrupting their architecture3 and inhibiting protein kinase C translocation to the plasma membrane.4 Recently, increased raft clustering was also reported in cultured dorsal root ganglion neurons prepared from the GLD murine model (i.e., the Twitcher mouse), and this was associated with the dysregulation of tyrosine kinase receptor A membrane recruitment and ligand-tyrosine kinase receptor A activated endocytosis.5 PSY induces p53-mediated apoptotic cell death,6 tumor necrosis factor-related apoptosis,6, 7 activation of secretory phospholipase A2,8 cytochrome C release from mitochondria and apoptosis activation via the caspase-9 pathway.9 Moreover, several authors found that peroxisomal β-oxidation was significantly inhibited and very long-chain fatty acid levels and reactive oxygen species (ROS) production were increased in PSY-treated cells.10, 11Calcium (Ca2+) is an essential ion for cell life, acting as a key second messenger in almost all cellular functions. It is well established that Ca2+ is one of the main second messengers involved in apoptotic cell death in neurons and in other cell types; sustained cytosolic Ca2+ increase can activate apoptosis.12 This can originate from extracellular influx or by release from intracellular stores like the endoplasmic reticulum.13 Importantly, mitochondria are also involved in Ca2+ homeostasis.14 Mitochondrial Ca2+ in basal conditions is maintained at low concentrations, but mitochondria are organelles that can take up high Ca2+ concentrations; indeed, different stimuli, such as nutrients, hormones or neurotransmitters that increase the cytoplasmic Ca2+ content also induce intra-mitochondrial Ca2+ increase.15 If this increase is relevant, ROS production increases and this is associated with mitochondrial membrane destruction, release of cytochrome C and apoptosis induction.16 During this process, pro-apoptotic Bcl2 family of proteins plays a crucial role by regulating the intracellular/mitochondrial Ca2+ content, and by inducing mitochondrial permeabilization, the essential step for cytochrome C release and caspase activation.12, 17It has been reported that some sphingolipid metabolites, such as ceramides and sphingosine, can play a crucial role in many steps of apoptosis induction as regulators of some Bcl2 family proteins, by increasing intracellular Ca2+ levels and inducing mitochondrial stress.18 However, these mechanisms have never been explored during PSY-induced cell death.In this article, we report on the role of intracellular Ca2+ dynamics during PSY-induced cell death in vitro. Using the human oligodendrocyte cell line MO3.13 and fluorescent probes, we measured Ca2+ variations in cytoplasm and mitochondria upon PSY administration until cell death. Moreover, we studied oxidative stress production in mitochondria by flow cytometry and time-lapse confocal fluorescence microscopy. Finally, in order to rescue cell viability in presence of PSY, we investigated the use of Ca2+ chelation in the extracellular medium, and its possible synergic effect with antioxidant treatment.  相似文献   

7.
Transient ischemia is a leading cause of cognitive dysfunction. Postischemic ROS generation and an increase in the cytosolic Zn2+ level ([Zn2+]c) are critical in delayed CA1 pyramidal neuronal death, but the underlying mechanisms are not fully understood. Here we investigated the role of ROS-sensitive TRPM2 (transient receptor potential melastatin-related 2) channel. Using in vivo and in vitro models of ischemia–reperfusion, we showed that genetic knockout of TRPM2 strongly prohibited the delayed increase in the [Zn2+]c, ROS generation, CA1 pyramidal neuronal death and postischemic memory impairment. Time-lapse imaging revealed that TRPM2 deficiency had no effect on the ischemia-induced increase in the [Zn2+]c but abolished the cytosolic Zn2+ accumulation during reperfusion as well as ROS-elicited increases in the [Zn2+]c. These results provide the first evidence to show a critical role for TRPM2 channel activation during reperfusion in the delayed increase in the [Zn2+]c and CA1 pyramidal neuronal death and identify TRPM2 as a key molecule signaling ROS generation to postischemic brain injury.Transient ischemia is a major cause of chronic neurological disabilities including memory impairment and cognitive dysfunctions in stroke survivors.1, 2 The underlying mechanisms are complicated and multiple, and remain not fully understood.3 It is well documented in rodents, non-human primates and humans that pyramidal neurons in the CA1 region of the hippocampus are particularly vulnerable and these neurons are demised after transient ischemia, commonly referred to as the delayed neuronal death.4 Studies using in vitro and in vivo models of transient ischemia have demonstrated that an increase in the [Zn2+]c or cytosolic Zn2+ accumulation is a critical factor.5, 6, 7, 8, 9, 10, 11 There is evidence supporting a role for ischemia-evoked release of vesicular Zn2+ at glutamatergic presynaptic terminals and subsequent entry into postsynaptic neurons via GluA2-lacking AMPA subtype glutamate receptors (AMPARs) to raise the [Zn2+]c.12, 13, 14, 15, 16 Upon reperfusion, while glutamate release returns to the preischemia level,17 Zn2+ can activate diverse ROS-generating machineries to generate excessive ROS as oxygen becomes available, which in turn elicits further Zn2+ accumulation during reperfusion.18, 19 ROS generation and cytosolic Zn2+ accumulation have a critical role in driving delayed CA1 pyramidal neuronal death,7, 12, 20, 21, 22 but the molecular mechanisms underlying such a vicious positive feedback during reperfusion remain poorly understood.Transient receptor potential melastatin-related 2 (TRPM2) forms non-selective cationic channels; their sensitivity to activation by ROS via a mechanism generating the channel activator ADP-ribose (ADPR) confers diverse cell types including hippocampal neurons with susceptibility to ROS-induced cell death, and thus TRPM2 acts as an important signaling molecule mediating ROS-induced adversities such as neurodegeneration.23, 24, 25, 26 Emergent evidence indeed supports the involvement of TRPM2 in transient ischemia-induced CA1 pyramidal neuronal death.27, 28, 29, 30 This has been attributed to the modulation of NMDA receptor-mediated signaling; despite that ROS-induced activation of the TRPM2 channels results in no change in the excitability of neurons from the wild-type (WT) mice, TRPM2 deficiency appeared to favor prosurvival synaptic Glu2A expression and inhibit prodeath extrasynaptic GluN2B expression.30 A recent study suggests that TRPM2 activation results in extracellular Zn2+ influx to elevate the [Zn2+]c.31 The present study, using TRPM2-deficient mice in conjunction with in vivo and in vitro models of transient global ischemia, provides compelling evidence to show ROS-induced TRPM2 activation during reperfusion as a crucial mechanism determining the delayed cytosolic Zn2+ accumulation, CA1 neuronal death and postischemic memory impairment.  相似文献   

8.
Autophagy is a cellular catabolic process needed for the degradation and recycling of protein aggregates and damaged organelles. Although Ca2+ is suggested to have an important role in cell survival, the ion channel(s) involved in autophagy have not been identified. Here we demonstrate that increase in intracellular Ca2+ via transient receptor potential canonical channel-1 (TRPC1) regulates autophagy, thereby preventing cell death in two morphologically distinct cells lines. The addition of DMOG or DFO, a cell permeable hypoxia-mimetic agents, or serum starvation, induces autophagy in both epithelial and neuronal cells. The induction of autophagy increases Ca2+ entry via the TRPC1 channel, which was inhibited by the addition of 2APB and SKF96365. Importantly, TRPC1-mediated Ca2+ entry resulted in increased expression of autophagic markers that prevented cell death. Furthermore, hypoxia-mediated autophagy also increased TRPC1, but not STIM1 or Orai1, expression. Silencing of TRPC1 or inhibition of autophagy by 3-methyladenine, but not TRPC3, attenuated hypoxia-induced increase in intracellular Ca2+ influx, decreased autophagy, and increased cell death. Furthermore, the primary salivary gland cells isolated from mice exposed to hypoxic conditions also showed increased expression of TRPC1 as well as increase in Ca2+ entry along with increased expression of autophagic markers. Altogether, we provide evidence for the involvement of Ca2+ influx via TRPC1 in regulating autophagy to protect against cell death.Autophagy is a cellular process responsible for the delivery of proteins or organelles to lysosomes for its degradation. Autophagy participates not only in maintaining cellular homeostasis, but also promotes cell survival during cellular stress situations.1, 2 The stress conditions including nutrient starvation, hypoxia conditions, invading microbes, and tumor formation, have been shown to induce autophagy that allows cell survival in these stressful or pathological situations.1 In addition, autophagy also recycles existing cytoplasmic components to generate the molecules that are required to sustain the most vital cellular functions.3 Till date, three forms of autophagy have been identified, which are designated as chaperone-mediated autophagy, microautophagy, and macroautophagy.4 Although the precise mechanism as to how autophagy is initiated is not well understood, many of the genes first identified in yeast that are involved in autophagy have orthologs in other eukaryotes including human homologs.5, 6 The presence of similar genes in all organisms suggests that autophagy might be a phenomenon that is evolutionally conserved that is essential for cell survival. In addition, since autophagy delivers a fresh pool of amino acids and other essential molecules to the cell, initiation of autophagy is highly beneficial particularly during nutritional stress situations or tissue remodeling during development and embryogenesis.6 Consequently, impaired or altered autophagy is often implicated in several pathologies, like neurodegenerative disorders and cancer,7, 8, 9 which again highlight its importance.Ca2+ has a vital role in the regulation of a large number of cellular processes such as cell proliferation, survival, migration, invasion, motility, and apoptosis.10, 11 To perform functions on such a broad spectrum, the cells have evolved multiple mechanisms regulating cellular Ca2+ levels, mainly by regulating the function of various Ca2+ channels present in different locations. Mitochondrial, ER, lysosomal, and cytosolic Ca2+ levels are regulated by Ca2+ permeable ion channels localized either on the membranes of the intracellular organelles or on the plasma membrane.10 The Ca2+ permeable channels, including families of TRPCs, Orais, voltage-gated, two-pore, mitochondrial Ca2+ uniporter, IP3, and ryanodine receptors have all been identified to contribute towards changes in intracellular Ca2+ ([Ca2+]i).10, 12, 13, 14 Channels of the TRPCs and Orai families have been related to several Ca2+-dependent physiological processes in various cell types, ranging from cell proliferation to contractility, to apoptosis under both physiological and pathological conditions.12 Moreover, it has been suggested that intracellular Ca2+ is one of the key regulators of autophagy;15 however, the possible role of Ca2+ in autophagy is still inconclusive. Many reports also suggest that Ca2+ inhibits autophagy,16, 17, 18 whereas others have indicated a stimulatory role for Ca2+ towards autophagy.19, 20, 21 Furthermore, the identity of the major Ca2+ channel(s) involved in autophagy is not known. Members of the TRPC family have been suggested as mediators of Ca2+ entry into cells. Activation of the G-protein (Gq/11–PLC pathway) leads to the generation of second messenger IP3.10, 22 IP3 binds to the IP3R, which initiates Ca2+ release from the ER stores, thereby facilitating stromal interacting molecule-1 (STIM1) to rearrange and activate Ca2+ entry via the store-operated channels.22 Two families of proteins (TRPCs and Orais) have been identified as potential candidates for SOC-mediated Ca2+ entry.12, 22 However, their role in autophagy has not yet been determined. Thus, here we investigated the role of Ca2+ entry channels (TRPCs and Orais) in autophagy and show that both hypoxia-mimetic and nutrient depression induces autophagy in two different cell lines. Furthermore, our data indicates that autophagy was dependent on TRPC1-mediated increase in intracellular Ca2+ levels, suggesting that TRPC1 has an important role in regulating autophagy and inhibiting cell death.  相似文献   

9.
The mitochondrial phosphate carrier (PiC) is critical for ATP synthesis by serving as the primary means for mitochondrial phosphate import across the inner membrane. In addition to its role in energy production, PiC is hypothesized to have a role in cell death as either a component or a regulator of the mitochondrial permeability transition pore (MPTP) complex. Here, we have generated a mouse model with inducible and cardiac-specific deletion of the Slc25a3 gene (PiC protein). Loss of PiC protein did not prevent MPTP opening, suggesting it is not a direct pore-forming component of this complex. However, Slc25a3 deletion in the heart blunted MPTP opening in response to Ca2+ challenge and led to a greater Ca2+ uptake capacity. This desensitization of MPTP opening due to loss or reduction in PiC protein attenuated cardiac ischemic-reperfusion injury, as well as partially protected cells in culture from Ca2+ overload induced death. Intriguingly, deletion of the Slc25a3 gene from the heart long-term resulted in profound hypertrophy with ventricular dilation and depressed cardiac function, all features that reflect the cardiomyopathy observed in humans with mutations in SLC25A3. Together, these results demonstrate that although the PiC is not a direct component of the MPTP, it can regulate its activity, suggesting a novel therapeutic target for reducing necrotic cell death. In addition, mice lacking Slc25a3 in the heart serve as a novel model of metabolic, mitochondrial-driven cardiomyopathy.The mitochondrial oxidative phosphorylation (OXPHOS) system is the primary source of cellular energy production. Defects in OXPHOS occur with a frequency of 1 in 5000 live births1 and underlie a wide range of mitochondrial disorders that often affect multiple organ systems and tissues with high oxidative energy demands, such as brain, skeletal muscle, and heart.2 Cardiac phenotypes associated with mitochondrial disease are diverse, and can range from cardiomyopathies to cardiac conduction defects.3, 4, 5The mitochondrial phosphate carrier (PiC) is a member of the solute carrier 25A family that has a critical role in OXPHOS, serving as the primary route for inorganic phosphate (Pi) import into the mitochondrial matrix.6, 7 PiC, together with the adenine nucleotide translocator (ANT) and the ATP synthase, forms the ATP synthasome whereby all of the metabolites needed to generate ATP are within one immediate microdomain.8, 9 The importance of PiC in facilitating energy production is highlighted by the profound disease phenotype observed in patients presenting with mutations in the skeletal muscle-specific isoform of this gene.10, 11 Such patients present with a multisystemic disorder characterized by muscle hypotonia, lactic acidosis, severe hypertrophic cardiomyopathy, and shortened lifespan.10, 11 Similarly, patients with SLC25A4 (ANT1 protein) deficiency present with cardiomyopathy,12 as do mice lacking the Slc25a4 gene,13 likely due to a similar molecular defect in the efficiency of ATP production within the mitochondria.In addition to its role in mitochondrial energy metabolism, PiC has been implicated in regulating cell death by serving either as a modulator or a direct component of the mitochondrial permeability transition pore (MPTP).14, 15, 16 The MPTP is a non-selective channel that forms in response to Ca2+ overload and oxidant stress that allows inner-membrane permeability to solutes up to 1500 Da in size, leading to loss of mitochondrial membrane potential, mitochondrial swelling and rupture, and eventually cell death through necrosis.15 Structurally, the MPTP complex has been proposed to be comprised of the ATP synthase17, 18 and to be regulated by cyclophilin D (CypD),19, 20 ANT,21 and the pro-apoptotic proteins Bax and Bak in the outer mitochondrial membrane.22, 23 PiC has also been suggested to be an inner-membrane component of the MPTP because it can form nonspecific channels in lipid membranes and because the MPTP is known to be activated by Pi.24, 25, 26, 27 Finally, PiC directly interacts with CypD in the mitochondrial matrix, which is a verified regulator and component of the MPTP.16 Saccharomyces cerevisiae lacking PiC have altered MPTP characteristics with a smaller pore size, suggesting it might directly participate in the mitochondrial permeability pore.28 However, partial reduction of PiC by siRNAs in cultured cells had no effect on mitochondrial permeability activity, suggesting that PiC is not required for MPTP function.27 Definitive genetic proof of PiC''s involvement in MPTP formation/function is currently lacking.In the present study, we tested the role of PiC in MPTP regulation and cell death in vivo using a mouse model with inducible cardiomyocyte-specific deletion of the Slc25a3 gene (encodes PiC). We found that cardiac mitochondria depleted of PiC were able to undergo permeability transition, suggesting that PiC is not a requisite component of the MPTP. However, the extent of Ca2+-induced MPTP opening was blunted, suggesting that PiC serves to regulate this activity. Furthermore, Slc25a3 deletion produced a unique mouse model of mitochondrial-driven hypertrophic cardiomyopathy that recapitulates features observed in human patients with phosphate carrier deficiency and metabolic cardiomyopathy.  相似文献   

10.
11.
Seizure activity has been proposed to result in the generation of reactive oxygen species (ROS), which then contribute to seizure-induced neuronal damage and eventually cell death. Although the mechanisms of seizure-induced ROS generation are unclear, mitochondria and cellular calcium overload have been proposed to have a crucial role. We aim to determine the sources of seizure-induced ROS and their contribution to seizure-induced cell death. Using live cell imaging techniques in glioneuronal cultures, we show that prolonged seizure-like activity increases ROS production in an NMDA receptor-dependent manner. Unexpectedly, however, mitochondria did not contribute to ROS production during seizure-like activity. ROS were generated primarily by NADPH oxidase and later by xanthine oxidase (XO) activity in a calcium-independent manner. This calcium-independent neuronal ROS production was accompanied by an increase in intracellular [Na+] through NMDA receptor activation. Inhibition of NADPH or XO markedly reduced seizure-like activity-induced neuronal apoptosis. These findings demonstrate a critical role for ROS in seizure-induced neuronal cell death and identify novel therapeutic targets.Reactive oxygen species (ROS) contribute to neuronal damage and have been linked to excitotoxicity.1, 2, 3, 4 An increase in ROS generation has also been identified in acute neurologic disease such as stroke,5,6 and recent evidence indicates that this may contribute to neuronal damage in seizures and epilepsy.7, 8, 9, 10 However, ROS measurements during seizure-like activity were predominantly performed in homogenates, extracellular fluids or brain regions with no clear demonstration of whether the ROS were of neuronal origin.9,11,12 Moreover, these studies lacked the necessary temporal resolution to determine accurately the evolution of ROS generation during and after prolonged seizure activity. Such obstacles can be overcome by live cell imaging of ROS, which has emerged as a powerful tool to study disease mechanisms.13If seizure activity induces ROS production in neurons, then a critical question is which sources of ROS production are triggered by such activity. Previous studies have suggested that mitochondria are the primary source of ROS generation in seizure models.8,14 However, there are alternative sources of ROS, in particular the enzymes NADPH oxidase and xanthine oxidase (XO). How these contribute to excitotoxicity during seizure activity is uncertain. That these enzymes may have an important role in seizure-induced ROS generation is suggested by two observations: (1) NMDA receptors have a pivotal role in seizure-induced neuronal damage15 and (2) direct pharmacologic activation of NMDA receptors can activate NADPH oxidase, increasing free radical production and consequently neuronal death.5,16,17 There is also burgeoning evidence of a role for NADPH oxidase activation in chronic brain pathology secondary to psychosocial stress, which leads to the development of neuropathologic alterations, and also in neurodegenerative disease.18,19Acute activation of NADPH oxidase in neurons has mainly been shown after direct pharmacologic activation of NMDA receptors via exposure to high levels of NMDA and this activation is calcium-dependent.16,17 More recently, activation of NADPH oxidase has been shown during seizure activity.9,20 These pathways also involved NMDA receptor activation and upregulation of NMDA receptor subunits NR1 and NR2B. Nonetheless, these studies used chemoconvulsant epilepsy models, which, in themselves, may have an impact on ROS generation. The mechanisms and relevance of activation of NADPH oxidase during seizure activity independent of chemoconvulsants is unclear, especially given the presence of alternative sources of ROS production. Moreover, XO may also represent a major potential source of ROS during periods of increased metabolism, such as that occuring during seizures. We have therefore asked whether NMDA receptor activation has a role in seizure-induced ROS production and which sources and mechanisms of ROS production are involved in its time course during seizure-like activity.Here, we demonstrate increased ROS generation during seizure-like activity. This is activity-dependent, but it is maintained by a Ca2+-independent pathway involving the activation of NMDA receptors, NADPH oxidase and XO at a later phase. Blocking NADPH oxidase and XO prevented seizure-induced neuronal cell death in vitro. We thus provide compelling evidence that these ROS-generating pathways are appropriate targets for preventing neuronal death in seizures.  相似文献   

12.
Pathological cardiac hypertrophy is a major risk factor associated with heart failure, a state concomitant with increased cell death. However, the mechanism governing progression of hypertrophy to apoptosis at the single-cell level remains elusive. Here, we demonstrate annexin A6 (Anxa6), a calcium (Ca2+)-dependent phospholipid-binding protein critically regulates the transition of chronic hypertrophied cardiomyocytes to apoptosis. Treatment of the H9c2(2-1) cardiomyocytes with hypertrophic agonists upregulates and relocalizes Anxa6 with increased cytosolic punctate appearance. Live cell imaging revealed that chronic exposure to hypertrophic agonists such as phenylephrine (PE) compromises the mitochondrial membrane potential (ΔΨm) and morphological dynamics. Such chronic hypertrophic induction also activated the caspases 9 and 3 and induced cleavage of the poly-(ADP-ribose) polymerase 1 (Parp1), which are the typical downstream events in the mitochondrial pathways of apoptosis. An increased rate of apoptosis was evident in the hypertrophied cardiomyocytes after 48–72 h of treatment with the hypertrophic agonists. Anxa6 was progressively associated with the mitochondrial fraction under chronic hypertrophic stimulation, and Anxa6 knockdown severely abrogated mitochondrial network and dynamics. Ectopically expressed Anxa6 protected the mitochondrial morphology and dynamics under PE treatment, and also increased the cellular susceptibility to apoptosis. Biochemical analysis showed that Anxa6 interacts with Parp1 and its 89 kDa cleaved product in a Ca2+-dependent manner through the N-terminal residues (1–28). Furthermore, expression of Anxa6S13E, a mutant dominant negative with respect to Parp1 binding, served as an enhancer of mitochondrial dynamics, even under chronic PE treatment. Chemical inhibition of Parp1 activity released the cellular vulnerability to apoptosis in Anxa6-expressing stable cell lines, thereby shifting the equilibrium away from cell death. Taken together, the present study depicts a dual regulatory function of Anxa6 that is crucial for balancing hypertrophy with apoptosis in cardiomyocytes.Complex machineries govern the life and death decisions in mammalian cells through a dynamic equilibrium, which is essential for physiological homeostasis.1 Such equilibrium is critical for cardiac myocytes because of their terminally differentiated states and low proliferative capacities. Stress response in cardiomyocytes often involves a switch between survival and cell death pathways.2, 3, 4 Cardiomyocyte hypertrophy is an adaptive response to stress, which may turn maladaptive and fatal,5 as evident in cardiovascular disorders that leads to heart failure.6 Hypertrophied phenotypes are also associated with a balance between cell growth and programmed cell death.7 These processes are aided by several patrolling proteins, which sense and operate to ameliorate the anomalies.8, 9 Understanding the dynamics of such signaling events is vital for the development of novel therapeutic strategies.Anxa6 belongs to the annexin family of calcium (Ca2+)/phospholipid-binding proteins.10 A major cardiac annexin,11 Anxa6 has diverse functions ranging from handling intracellular Ca2+ signaling, cholesterol transport,12 Ras inactivation13 and vesicular traffic.14 Anxa6 mostly functions as an intracellular scaffold.15 Although mice with targeted depletion of the Anxa6 gene remain viable,16 functional redundancies within the annexin family have been proposed to compensate for the loss of Anxa6 function.17, 18 A 10-fold overexpression of Anxa6 targeted to the heart developed cardiomyopathies in mice, whereas cardiomyocytes from Anxa6-knockout mice exhibited increased contractility and altered Ca2+ turnover.19, 20 Such contradictory findings may indicate participation of Anxa6 in counterbalancing signaling mechanisms. Moreover, end-stage heart failures have been reported to be associated with downregulation of Anxa6, and, in general, Anxa6 has compensatory roles in chronic pathological conditions.20, 21, 22 However, the function of differential Anxa6 expression or dynamics in chronic cardiomyocyte hypertrophy is poorly understood.We have reported the interactions of Anxa6 with the sarcomeric α-actinin and its role in cardiomyocyte contractility.23 Recently, we have characterized a role of Anxa6 in the antihypertrophic signaling via the regulation of atrial natriuretic peptide (ANP) secretion.24 The mechanistic spectrum of Anxa6 in the earlier study was limited to a short-term (24 h) exposure of H9c2 cardiomyocytes to the α1-adrenergic receptor agonist phenylephrine (PE). The dynamics of Anxa6 within this small window yielded valuable insight into the spatiotemporal regulation of hypertrophic signaling. Here, we extended the study to understand the dynamics of Anxa6 under chronic hypertrophic conditions. The mechanodeficient H9c2(2-1) cardiomyocyte line has been instrumental in our study to rule out the contributions of Anxa6 towards contractility,23 owing to its multidimensional scaffold activity and functional compensations.17, 18 The H9c2 cardiomyocytes have been extensively characterized and ARE an established animal origin-free model for studying signal-transduction pathways in cardiomyocytes, including hypertrophy.25, 26Adrenergic stimulation is crucial in compensatory and pathological cardiac hypertrophy, an early state that may proceed towards heart failure.27 Cardiac hypertrophy at advanced stages (chronic) is associated with mitochondrial dysfunction, which also contributes to cardiac decompensation.28 To explore the temporal events under chronic hypertrophy, we analyzed the effects of adrenergic induction on mitochondrial membrane potential (ΔΨm) and morphological dynamics, parameters that are directly correlated with mitochondrial dysfunction and programmed cell death.29, 30, 31 Anxa6 has been reported to be associated with mitochondria in some cell types.17, 32, 33 In the present study, we aim to understand the functions of Anxa6 under chronic hypertrophic conditions that may progress towards apoptosis.  相似文献   

13.
14.
15.
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   

16.
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   

17.
Glycogen synthase kinase-3β (GSK3β) is a multifunctional kinase whose inhibition is known to limit myocardial ischemia–reperfusion injury. However, the mechanism mediating this beneficial effect still remains unclear. Mitochondria and sarco/endoplasmic reticulum (SR/ER) are key players in cell death signaling. Their involvement in myocardial ischemia–reperfusion injury has gained recognition recently, but the underlying mechanisms are not yet well understood. We questioned here whether GSK3β might have a role in the Ca2+ transfer from SR/ER to mitochondria at reperfusion. We showed that a fraction of GSK3β protein is localized to the SR/ER and mitochondria-associated ER membranes (MAMs) in the heart, and that GSK3β specifically interacted with the inositol 1,4,5-trisphosphate receptors (IP3Rs) Ca2+ channeling complex in MAMs. We demonstrated that both pharmacological and genetic inhibition of GSK3β decreased protein interaction of IP3R with the Ca2+ channeling complex, impaired SR/ER Ca2+ release and reduced the histamine-stimulated Ca2+ exchange between SR/ER and mitochondria in cardiomyocytes. During hypoxia reoxygenation, cell death is associated with an increase of GSK3β activity and IP3R phosphorylation, which leads to enhanced transfer of Ca2+ from SR/ER to mitochondria. Inhibition of GSK3β at reperfusion reduced both IP3R phosphorylation and SR/ER Ca2+ release, which consequently diminished both cytosolic and mitochondrial Ca2+ concentrations, as well as sensitivity to apoptosis. We conclude that inhibition of GSK3β at reperfusion diminishes Ca2+ leak from IP3R at MAMs in the heart, which limits both cytosolic and mitochondrial Ca2+ overload and subsequent cell death.Glycogen synthase kinase-3 (GSK3) was originally identified as a phosphorylating kinase for glycogen synthase.1, 2 It has two isoforms, α and β, that possess strong homology in their kinase domains with, however, distinct functions.3 GSK3 is constitutively active but it can be inhibited by phosphorylation on serine 21 (Ser21) for GSK3α and Ser9 for GSK3β.4 In the heart, GSK3β has several important roles in cardiac hypertrophy5 and ischemia–reperfusion (IR) injury.6 Accumulating evidence indicates that phospho-Ser9-GSK3β-mediated cytoprotection is achieved by an increased threshold for permeability transition pore (PTP) opening.6, 7, 8, 9 The mechanism by which GSK3β delays PTP opening still remains unclear. It has been reported that GSK3β could interact with ANT at the inner mitochondrial membrane in the heart9 and/or to phosphorylate voltage-dependent anion channel (VDAC) and cyclophilin D (CypD) in cancer cells.10, 11 GSK3β also has other proposed mechanisms of action, including a poorly characterized role in calcium (Ca2+) homeostasis regulation12 and protein–protein interactions,9 as well as functions in different subcellular fractions such as the nucleus, cytosol and mitochondria.13Reperfusion is the most powerful intervention to salvage ischemic myocardium. However, it can also paradoxically lead to cardiomyocyte injury and death.14 One of the main actors of this lethal reperfusion injury is cellular Ca2+ overload,15 which results in part from excessive sarco/endoplasmic reticulum (SR/ER) Ca2+ release and Ca2+ influx through the plasma membrane (e.g. through L-type Ca2+channel and NCX (sodium-calcium exchanger)).16 Although ryanodine receptors (RyRs) are the major cardiac SR/ER Ca2+-release channels involved in excitation–contraction coupling (ECC)17 and ischemia–reperfusion (IR) injury,18 recent studies reported an increasing role for inositol 1,4,5-trisphosphate receptors (IP3Rs) Ca2+-release channels in the modulation of ECC and cell death.19, 20 Ca2+-handling proteins of ER and mitochondria are highly concentrated at mitochondria-associated ER membranes (MAMs), providing a direct and proper mitochondrial Ca2+ signaling, including VDAC, Grp75 and IP3R1.20, 21, 22Here, we provide evidence that, following IR, a fraction of cellular GSK3β is localized at the SR/ER and MAMs. At the MAMs interface, GSK3β can specifically interact and regulate the protein composition of the IP3R Ca2+ channeling complex and modulate Ca2+ transfer between SR/ER and mitochondria. These findings support a novel mechanism of action of GSK3β in cell death process during reperfusion injury.  相似文献   

18.
Excessive N-methyl-D-aspartate receptor (NMDAR) activation and the resulting activation of neuronal nitric oxide synthase (nNOS) cause neuronal injury. Homer1b/c facilitates NMDAR-PSD95-nNOS complex interactions, and Homer1a is a negative competitor of Homer1b/c. We report that Homer1a was both upregulated by and protected against NMDA-induced neuronal injury in vitro and in vivo. The neuroprotective activity of Homer1a was associated with NMDA-induced Ca2+ influx, oxidative stress and the resultant downstream signaling activation. Additionally, we found that Homer1a functionally regulated NMDAR channel properties in neurons, but did not regulate recombinant NR1/NR2B receptors in HEK293 cells. Furthermore, we found that Homer1a detached the physical links among NR2B, PSD95 and nNOS and reduced the membrane distribution of NMDAR. NMDA-induced neuronal injury was more severe in Homer1a homozygous knockout mice (KO, Homer1a−/−) when compared with NMDA-induced neuronal injury in wild-type mice (WT, Homer1a+/+). Additionally, Homer1a overexpression in the cortex of Homer1a−/− mice alleviated NMDA-induced neuronal injury. These findings suggest that Homer1a may be a key neuroprotective endogenous molecule that protects against NMDA-induced neuronal injury by disassembling NR2B-PSD95-nNOS complexes and reducing the membrane distribution of NMDARs.Glutamate (Glu) acts on glutamate receptors, such as the N-methyl-D-aspartate receptor (NMDAR), and leads to neuronal hyper-excitability and death in a dose-dependent manner.1 NMDAR activation induces Ca2+ influx and specifically activates neuronal nitric oxide synthase (nNOS) and downstream signaling pathways.2, 3, 4 Ca2+ influx is involved in glutamate-induced apoptosis caused by the activation of apoptosis-related signaling pathways, mitochondrial dysfunction and ROS induction.3, 4 Additionally, nNOS has been reported to contribute to NMDA-induced excitotoxicity.5, 6 Considering that direct NMDAR inhibition has not yet demonstrated favorable efficacy in most clinic trails and further considering the remarkable role of nNOS in NMDA-induced neuronal death,7 measures that can effectively protect neurons from NMDA-induced neuronal injury are urgently needed and represent a worthwhile research goal.Homer proteins belong to the postsynaptic density (PSD) family and consist of two major groups: the short-form Homer proteins (Homer1a and Ania3) and the long-form Homer proteins (Homer1b/c, Homer2 and Homer3).8 Homer1b/c has a conserved N-terminal Ena/VASP homology 1 domain and binds to group I metabotropic glutamate receptors (mGluRs), inositol triphosphate receptors and Shank family proteins.9, 10, 11, 12 Homer1b/c regulates surface receptor expression,13, 14 clustering,15 transient receptor potential family channels and mGluRs coupled to ion channels.10, 16, 17, 18, 19 Additionally, because of its C-terminal coiled-coil (CC) domains, Homer1b/c can self-multimerize, form multiprotein complexes and facilitate signal transduction to downstream pathways. Homer1a, which lacks the CC domain, is believed to compete with constitutive Homer1b/c and disrupt the association of multiple Homer1b/c complexes.Notably, Homer1b/c can interact with the Glu-induced Ca2+ influx pathway by binding to Shank, a NMDAR complex adaptor protein (NMDAR-PSD95-GKAP-Shank-Homer1b/c).12, 20 Furthermore, Homer1a also interacts with Shank, NMDA, nNOS and other Homer1b/c target proteins. Homer1a has a negative regulatory role by physically replacing certain target proteins, and is involved in the regulation of a variety of cellular and molecular functions in neurological diseases.21, 22, 23, 24, 25 Nevertheless, the mechanisms of action and associations between Homer1a and NMDA-induced neuronal injury have not yet been studied. Here, we aimed to investigate the possible neuroprotective effects of Homer1a and explore the mechanisms underlying Homer1a activity in NMDA-induced neuronal injury.  相似文献   

19.
Although cytosolic free Ca2+ mobilization induced by microbe/pathogen-associated molecular patterns is postulated to play a pivotal role in innate immunity in plants, the molecular links between Ca2+ and downstream defense responses still remain largely unknown. Calcineurin B-like proteins (CBLs) act as Ca2+ sensors to activate specific protein kinases, CBL-interacting protein kinases (CIPKs). We here identified two CIPKs, OsCIPK14 and OsCIPK15, rapidly induced by microbe-associated molecular patterns, including chitooligosaccharides and xylanase (Trichoderma viride/ethylene-inducing xylanase [TvX/EIX]), in rice (Oryza sativa). Although they are located on different chromosomes, they have over 95% nucleotide sequence identity, including the surrounding genomic region, suggesting that they are duplicated genes. OsCIPK14/15 interacted with several OsCBLs through the FISL/NAF motif in yeast cells and showed the strongest interaction with OsCBL4. The recombinant OsCIPK14/15 proteins showed Mn2+-dependent protein kinase activity, which was enhanced both by deletion of their FISL/NAF motifs and by combination with OsCBL4. OsCIPK14/15-RNAi transgenic cell lines showed reduced sensitivity to TvX/EIX for the induction of a wide range of defense responses, including hypersensitive cell death, mitochondrial dysfunction, phytoalexin biosynthesis, and pathogenesis-related gene expression. On the other hand, TvX/EIX-induced cell death was enhanced in OsCIPK15-overexpressing lines. Our results suggest that OsCIPK14/15 play a crucial role in the microbe-associated molecular pattern-induced defense signaling pathway in rice cultured cells.Calcium ions regulate diverse cellular processes in plants as a ubiquitous internal second messenger, conveying signals received at the cell surface to the inside of the cell through spatial and temporal concentration changes that are decoded by an array of Ca2+ sensors (Reddy, 2001; Sanders et al., 2002; Yang and Poovaiah, 2003). Several families of Ca2+ sensors have been identified in higher plants. The best known are calmodulins (CaMs) and CaM-related proteins, which typically contain four EF-hand domains for Ca2+ binding (Zielinski, 1998). Unlike mammals, which possess single molecular species of CaM, plants have at least three distinct molecular species of CaM playing diverse physiological functions and whose expression is differently regulated (Yamakawa et al., 2001; Luan et al., 2002; Karita et al., 2004; Takabatake et al., 2007). The second major class is exemplified by the Ca2+-dependent protein kinases, which contain CaM-like Ca2+-binding domains and a kinase domain in a single protein (Harmon et al., 2000). In addition, a new family of Ca2+ sensors was identified as calcineurin B-like (CBL) proteins, which consists of proteins similar to both the regulatory β-subunit of calcineurin and the neuronal Ca2+ sensor in animals (Liu and Zhu, 1998; Kudla et al., 1999).Unlike CaMs, which interact with a large variety of target proteins, CBLs specifically target a family of protein kinases referred to as CBL-interacting protein kinases (CIPKs) or SnRK3s (for sucrose nonfermenting 1-related protein kinases type 3), which are most similar to the SNF family protein kinases in yeast (Luan et al., 2002). A database search of the Arabidopsis (Arabidopsis thaliana) genome sequence revealed 10 CBL and 25 CIPK homologues (Luan et al., 2002). Expression patterns of these Ca2+ sensors and protein kinases suggest their diverse functions in different signaling processes, including light, hormone, sugar, and stress responses (Batistic and Kudla, 2004). AtCBL4/Salt Overly Sensitive3 (SOS3) and AtCIPK24/SOS2 have been shown to play a key role in Ca2+-mediated salt stress adaptation (Zhu, 2002). The CBL-CIPK system has been shown to be involved in signaling pathways of abscisic acid (Kim et al., 2003a), sugar (Gong et al., 2002a), gibberellins (Hwang et al., 2005), salicylic acid (Mahajan et al., 2006), and K+ channel regulation (Li et al., 2006; Lee et al., 2007; for review, see Luan, 2009; Batistic and Kudla, 2009). However, physiological functions of most of the family members still remain largely unknown.Plants respond to pathogen attack by activating a variety of defense responses, including the generation of reactive oxygen species (ROS), synthesis of phytoalexins, expression of pathogenesis-related (PR) genes, cell cycle arrest, and mitochondrial dysfunction followed by a form of hypersensitive cell death known as the hypersensitive response (Nürnberger and Scheel, 2001; Greenberg and Yao, 2004; Kadota et al., 2004b). Transient membrane potential changes and Ca2+ influx are involved at the initial stage of defense responses (Kuchitsu et al., 1993; Pugin et al., 1997; Blume et al., 2000; Kadota et al., 2004a). Many kinds of defense responses are prevented when Ca2+ influx is compromised by Ca2+ chelators (Nürnberger and Scheel, 2001; Lecourieux et al., 2002). Since complex spatiotemporal patterns of cytosolic free Ca2+ concentration have been suggested to play pivotal roles in defense signaling (Nürnberger and Scheel, 2001; Sanders et al., 2002), multiple Ca2+ sensor proteins and their effectors should function in the defense signaling pathways. Although possible involvement of some CaM isoforms (Heo et al., 1999; Yamakawa et al., 2001), Ca2+-dependent protein kinases (Romeis et al., 2000, 2001; Ludwig et al., 2005; Kobayashi et al., 2007; Yoshioka et al., 2009), as well as Ca2+ regulation of EF-hand-containing enzymes such as ROS-generating NADPH oxidase (Ogasawara et al., 2008) have been suggested, other Ca2+-regulated signaling components still remain to be identified. No CBLs or CIPKs have so far been implicated as signaling components in defense signaling.N-Acetylchitooligosaccharides, chitin fragments, are microbe-associated molecular patterns (MAMPs) that are recognized by plasma membrane receptors (Kaku et al., 2006; Miya et al., 2007) and induce a variety of defense responses, such as membrane depolarization (Kuchitsu et al., 1993; Kikuyama et al., 1997), ion fluxes (Kuchitsu et al., 1997), ROS production (Kuchitsu et al., 1995), phytoalexin biosynthesis (Yamada et al., 1993), and induction of PR genes (Nishizawa et al., 1999), without hypersensitive cell death in rice (Oryza sativa) cells. In contrast, a fungal proteinaceous elicitor, xylanase from Trichoderma viride (TvX)/ethylene-inducing xylanase (EIX), which is recognized by two putative plasma membrane receptors, LeEix1 and LeEix2 (Ron and Avni, 2004), triggers hypersensitive cell death along with different kinetics of ROS production and activation of a mitogen-activated protein kinase, OsMPK6, previously named as OsMPK2 or OsMAPK6, in rice cells (Kurusu et al., 2005). These two fungal MAMPs thus provide excellent model systems to study innate immunity in rice cells.This study identified two CIPKs involved in various MAMP-induced layers of defense responses, including PR gene expression, phytoalexin biosynthesis, mitochondrial dysfunction, and cell death, in rice. Molecular characterization of these CIPKs, including interaction with the putative Ca2+ sensors as well as their physiological functions, is discussed.  相似文献   

20.
The role of calcium-mediated signaling has been extensively studied in plant responses to abiotic stress signals. Calcineurin B-like proteins (CBLs) and CBL-interacting protein kinases (CIPKs) constitute a complex signaling network acting in diverse plant stress responses. Osmotic stress imposed by soil salinity and drought is a major abiotic stress that impedes plant growth and development and involves calcium-signaling processes. In this study, we report the functional analysis of CIPK21, an Arabidopsis (Arabidopsis thaliana) CBL-interacting protein kinase, ubiquitously expressed in plant tissues and up-regulated under multiple abiotic stress conditions. The growth of a loss-of-function mutant of CIPK21, cipk21, was hypersensitive to high salt and osmotic stress conditions. The calcium sensors CBL2 and CBL3 were found to physically interact with CIPK21 and target this kinase to the tonoplast. Moreover, preferential localization of CIPK21 to the tonoplast was detected under salt stress condition when coexpressed with CBL2 or CBL3. These findings suggest that CIPK21 mediates responses to salt stress condition in Arabidopsis, at least in part, by regulating ion and water homeostasis across the vacuolar membranes.Drought and salinity cause osmotic stress in plants and severely affect crop productivity throughout the world. Plants respond to osmotic stress by changing a number of cellular processes (Xiong et al., 1999; Xiong and Zhu, 2002; Bartels and Sunkar, 2005; Boudsocq and Lauriére, 2005). Some of these changes include activation of stress-responsive genes, regulation of membrane transport at both plasma membrane (PM) and vacuolar membrane (tonoplast) to maintain water and ionic homeostasis, and metabolic changes to produce compatible osmolytes such as Pro (Stewart and Lee, 1974; Krasensky and Jonak, 2012). It has been well established that a specific calcium (Ca2+) signature is generated in response to a particular environmental stimulus (Trewavas and Malhó, 1998; Scrase-Field and Knight, 2003; Luan, 2009; Kudla et al., 2010). The Ca2+ changes are primarily perceived by several Ca2+ sensors such as calmodulin (Reddy, 2001; Luan et al., 2002), Ca2+-dependent protein kinases (Harper and Harmon, 2005), calcineurin B-like proteins (CBLs; Luan et al., 2002; Batistič and Kudla, 2004; Pandey, 2008; Luan, 2009; Sanyal et al., 2015), and other Ca2+-binding proteins (Reddy, 2001; Shao et al., 2008) to initiate various cellular responses.Plant CBL-type Ca2+ sensors interact with and activate CBL-interacting protein kinases (CIPKs) that phosphorylate downstream components to transduce Ca2+ signals (Liu et al., 2000; Luan et al., 2002; Batistič and Kudla, 2004; Luan, 2009). In several plant species, multiple members have been identified in the CBL and CIPK family (Luan et al., 2002; Kolukisaoglu et al., 2004; Pandey, 2008; Batistič and Kudla, 2009; Weinl and Kudla, 2009; Pandey et al., 2014). Involvement of specific CBL-CIPK pair to decode a particular type of signal entails the alternative and selective complex formation leading to stimulus-response coupling (D’Angelo et al., 2006; Batistič et al., 2010).Several CBL and CIPK family members have been implicated in plant responses to drought, salinity, and osmotic stress based on genetic analysis of Arabidopsis (Arabidopsis thaliana) mutants (Zhu, 2002; Cheong et al., 2003, 2007; Kim et al., 2003; Pandey et al., 2004, 2008; D’Angelo et al., 2006; Qin et al., 2008; Tripathi et al., 2009; Held et al., 2011; Tang et al., 2012; Drerup et al., 2013; Eckert et al., 2014). A few CIPKs have also been functionally characterized by gain-of-function approach in crop plants such as rice (Oryza sativa), pea (Pisum sativum), and maize (Zea mays) and were found to be involved in osmotic stress responses (Mahajan et al., 2006; Xiang et al., 2007; Yang et al., 2008; Tripathi et al., 2009; Zhao et al., 2009; Cuéllar et al., 2010).In this report, we examined the role of the Arabidopsis CIPK21 gene in osmotic stress response by reverse genetic analysis. The loss-of-function mutant plants became hypersensitive to salt and mannitol stress conditions, suggesting that CIPK21 is involved in the regulation of osmotic stress response in Arabidopsis. These findings are further supported by an enhanced tonoplast targeting of the cytoplasmic CIPK21 through interaction with the vacuolar Ca2+ sensors CBL2 and CBL3 under salt stress condition.  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