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.
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   

2.
Na+ and K+ homeostasis are crucial for plant growth and development. Two HKT transporter/channel classes have been characterized that mediate either Na+ transport or Na+ and K+ transport when expressed in Xenopus laevis oocytes and yeast. However, the Na+/K+ selectivities of the K+-permeable HKT transporters have not yet been studied in plant cells. One study expressing 5′ untranslated region-modified HKT constructs in yeast has questioned the relevance of cation selectivities found in heterologous systems for selectivity predictions in plant cells. Therefore, here we analyze two highly homologous rice (Oryza sativa) HKT transporters in plant cells, OsHKT2;1 and OsHKT2;2, that show differential K+ permeabilities in heterologous systems. Upon stable expression in cultured tobacco (Nicotiana tabacum) Bright-Yellow 2 cells, OsHKT2;1 mediated Na+ uptake, but little Rb+ uptake, consistent with earlier studies and new findings presented here in oocytes. In contrast, OsHKT2;2 mediated Na+-K+ cotransport in plant cells such that extracellular K+ stimulated OsHKT2;2-mediated Na+ influx and vice versa. Furthermore, at millimolar Na+ concentrations, OsHKT2;2 mediated Na+ influx into plant cells without adding extracellular K+. This study shows that the Na+/K+ selectivities of these HKT transporters in plant cells coincide closely with the selectivities in oocytes and yeast. In addition, the presence of external K+ and Ca2+ down-regulated OsHKT2;1-mediated Na+ influx in two plant systems, Bright-Yellow 2 cells and intact rice roots, and also in Xenopus oocytes. Moreover, OsHKT transporter selectivities in plant cells are shown to depend on the imposed cationic conditions, supporting the model that HKT transporters are multi-ion pores.Intracellular Na+ and K+ homeostasis play vital roles in growth and development of higher plants (Clarkson and Hanson, 1980). Low cytosolic Na+ and high K+/Na+ ratios aid in maintaining an osmotic and biochemical equilibrium in plant cells. Na+ and K+ influx and efflux across membranes require the function of transmembrane Na+ and K+ transporters/channels. Several Na+-permeable transporters have been characterized in plants (Zhu, 2001; Horie and Schroeder, 2004; Apse and Blumwald, 2007). Na+/H+ antiporters mediate sequestration of Na+ into vacuoles under salt stress conditions in plants (Blumwald and Poole, 1985, 1987; Sze et al., 1999). Na+ (cation)/H+ antiporters are encoded by six AtNHX genes in Arabidopsis (Arabidopsis thaliana; Apse et al., 1999; Gaxiola et al., 1999; Yokoi et al., 2002; Aharon et al., 2003). A distinct Na+/H+ antiporter, Salt Overly Sensitive1, mediates Na+/H+ exchange at the plasma membrane and mediates cellular Na+ extrusion (Shi et al., 2000, 2002; Zhu, 2001; Ward et al., 2003). Electrophysiological analyses reveal that voltage-independent channels, also named nonselective cation channels, mediate Na+ influx into roots under high external Na+ concentrations (Amtmann et al., 1997; Tyerman et al., 1997; Buschmann et al., 2000; Davenport and Tester, 2000); however, the underlying genes remain unknown.Potassium is the most abundant cation in plants and an essential nutrient for plant growth. The Arabidopsis genome includes 13 genes encoding KUP/HAK/KT transporters (Quintero and Blatt, 1997; Santa-María et al., 1997; Fu and Luan, 1998; Kim et al., 1998), and 17 genes have been identified encoding this family of transporters in rice (Oryza sativa ‘Nipponbare’; Bañuelos et al., 2002). Several KUP/HAK/KT transporters have been characterized as mediating K+ uptake across the plasma membrane of plant cells (Rigas et al., 2001; Bañuelos et al., 2002; Gierth et al., 2005).Ionic balance, especially the Na+/K+ ratio, is a key factor of salt tolerance in plants (Niu et al., 1995; Maathuis and Amtmann, 1999; Shabala, 2000; Mäser et al., 2002a; Tester and Davenport, 2003; Horie et al., 2006; Apse and Blumwald, 2007; Chen et al., 2007; Gierth and Mäser, 2007). Salinity stress is a major problem for agricultural productivity of crops worldwide (Greenway and Munns, 1980; Zhu, 2001). The Arabidopsis AtHKT1;1 transporter plays a key role in salt tolerance of plants by mediating Na+ exclusion from leaves (Mäser et al., 2002a; Berthomieu et al., 2003; Gong et al., 2004; Sunarpi et al., 2005; Rus et al., 2006; Davenport et al., 2007; Horie et al., 2009). athkt1;1 mutations cause leaf chlorosis and elevated Na+ accumulation in leaves under salt stress conditions in Arabidopsis (Mäser et al., 2002a; Berthomieu et al., 2003; Gong et al., 2004; Sunarpi et al., 2005). AtHKT1;1 and its homolog in rice, OsHKT1;5 (SKC1), mediate leaf Na+ exclusion by removing Na+ from the xylem sap to protect plants from salinity stress (Ren et al., 2005; Sunarpi et al., 2005; Horie et al., 2006, 2009; Davenport et al., 2007).The land plant HKT gene family is divided into two classes based on their nucleic acid sequences and protein structures (Mäser et al., 2002b; Platten et al., 2006). Class 1 HKT transporters have a Ser residue at a selectivity filter position in the first pore loop, which is replaced by a Gly in all but one known class 2 HKT transporter (Horie et al., 2001; Mäser et al., 2002b; Garciadeblás et al., 2003). While the Arabidopsis genome includes only one HKT gene, AtHKT1;1 (Uozumi et al., 2000), seven full-length OsHKT genes were found in the japonica rice cv Nipponbare genome (Garciadeblás et al., 2003). Members of class 1 HKT transporters, AtHKT1;1 and SKC1/OsHKT1;5, have a relatively higher Na+-to-K+ selectivity in Xenopus laevis oocytes and yeast than class 2 HKT transporters (Uozumi et al., 2000; Horie et al., 2001; Mäser et al., 2002b; Ren et al., 2005). The first identified plant HKT transporter, TaHKT2;1 from wheat (Triticum aestivum), is a class 2 HKT transporter (Schachtman and Schroeder, 1994). TaHKT2;1 was found to mediate Na+-K+ cotransport and Na+ influx at high Na+ concentrations in heterologous expression systems (Rubio et al., 1995, 1999; Gassmann et al., 1996; Mäser et al., 2002b). Thus, class 1 HKT transporters have been characterized as Na+-preferring transporters with a smaller K+ permeability (Fairbairn et al., 2000; Uozumi et al., 2000; Su et al., 2003; Jabnoune et al., 2009), whereas class 2 HKT transporters function as Na+-K+ cotransporters or channels (Gassmann et al., 1996; Corratgé et al., 2007). In addition, at millimolar Na+ concentrations, class 2 HKT transporters were found to mediate Na+ influx, without adding external K+ in Xenopus oocytes and yeast (Rubio et al., 1995, 1999; Gassmann et al., 1996; Horie et al., 2001). However, the differential cation transport selectivities of the two types of HKT transporters have not yet been analyzed and compared in plant cells.A study of the barley (Hordeum vulgare) and wheat class 2 transporters has suggested that the transport properties of HvHKT2;1 and TaHKT2;1 expressed in yeast are variable, depending on the constructs from which the transporter is expressed, and have led to questioning of the K+ transport activity of HKT transporters characterized in Xenopus oocytes and yeast (Haro et al., 2005). It was further proposed that the 5′ translation initiation of HKT proteins in yeast at nonconventional (non-ATG) sites affects the transporter selectivities of HKT transporters (Haro et al., 2005), although direct evidence for this has not yet been presented. However, recent research has shown a K+ permeability of OsHKT2;1 but not of OsHKT1;1 and OsHKT1;3 in Xenopus oocytes. These three OsHKT transporters show overlapping and also distinctive expression patterns in rice (Jabnoune et al., 2009).The report of Haro et al. (2005) has opened a central question addressed in this study: are the Na+/K+ transport selectivities of plant HKT transporters characterized in heterologous systems of physiological relevance in plant cells, or do they exhibit strong differences in the cation transport selectivities in these nonplant versus plant systems? To address this question, we analyzed the Na+/K+ transport selectivities of the OsHKT2;1 and OsHKT2;2 transporters expressed in cultured tobacco (Nicotiana tabacum ‘Bright-Yellow 2’ [BY2]) cells. OsHKT2;1 and OsHKT2;2 are two highly homologous HKT transporters from indica rice cv Pokkali, sharing 91% amino acid and 93% cDNA sequence identity (Horie et al., 2001). OsHKT2;1 mediates mainly Na+ uptake, which correlates with the presence of a Ser residue in the first pore loop of OsHKT2;1 (Horie et al., 2001, 2007; Mäser et al., 2002b; Garciadeblás et al., 2003). In contrast, OsHKT2;2 mediates Na+-K+ cotransport in Xenopus oocytes and yeast (Horie et al., 2001). Furthermore, at millimolar Na+ concentrations, OsHKT2;2 mediates Na+ influx in the absence of added K+ (Horie et al., 2001). Recent research on oshkt2;1 loss-of-function mutant alleles has revealed that OsHKT2;1 from japonica rice mediates a large Na+ influx component into K+-starved roots, thus compensating for lack of K+ availability (Horie et al., 2007). But the detailed Na+/K+ selectivities of Gly-containing, predicted K+-transporting class 2 HKT transporters have not yet been analyzed in plant cells.Here, we have generated stable OsHKT2;1- and OsHKT2;2-expressing tobacco BY2 cell lines and characterized the cell lines by ion content measurements and tracer influx studies to directly analyze unidirectional fluxes (Epstein et al., 1963). These analyses showed that OsHKT2;1 exhibits Na+ uptake activity in plant BY2 cells in the absence of added K+, but little K+ (Rb+), influx activity. In contrast, OsHKT2;2 was found to function as a Na+-K+ cotransporter/channel in plant BY2 cells, showing K+-stimulated Na+ influx and Na+-stimulated K+ (Rb+) influx. The differential K+ selectivities of the two OsHKT2 transporters were consistently reproduced by voltage clamp experiments using Xenopus oocytes here, as reported previously (Horie et al., 2001). OsHKT2;2 was also found to mediate K+-independent Na+ influx at millimolar external Na+ concentrations. These findings demonstrate that the cation selectivities of OsHKT2;1 and OsHKT2;2 in plant cells are consistent with past findings obtained from heterologous expression analyses under similar ionic conditions (Horie et al., 2001; Garciadeblás et al., 2003; Tholema et al., 2005). Furthermore, the shift in OsHKT2;2 Na+-K+ selectivity depending on ionic editions is consistent with the model that HKT transporters/channels are multi-ion pores (Gassmann et al., 1996; Corratgé et al., 2007). Classical studies of ion channels have shown that ion channels, in which multiple ions can occupy the pore at the same time, can change their relative selectivities depending on the ionic conditions (Hille, 2001). Moreover, the presence of external K+ and Ca2+ was found here to down-regulate OsHKT2;1-mediated Na+ influx both in tobacco BY2 cells and in rice roots. The inhibitory effect of external K+ on OsHKT2;1-mediated Na+ influx into intact rice roots, however, showed a distinct difference in comparison with that of BY2 cells, which indicates a possible posttranslational regulation of OsHKT2;1 in K+-starved rice roots.  相似文献   

3.
Although cellular prion protein (PrPc) has been suggested to have physiological roles in neurogenesis and angiogenesis, the pathophysiological relevance of both processes remain unknown. To elucidate the role of PrPc in post-ischemic brain remodeling, we herein exposed PrPc wild type (WT), PrPc knockout (PrP−/−) and PrPc overexpressing (PrP+/+) mice to focal cerebral ischemia followed by up to 28 days reperfusion. Improved neurological recovery and sustained neuroprotection lasting over the observation period of 4 weeks were observed in ischemic PrP+/+ mice compared with WT mice. This observation was associated with increased neurogenesis and angiogenesis, whereas increased neurological deficits and brain injury were noted in ischemic PrP−/− mice. Proteasome activity and oxidative stress were increased in ischemic brain tissue of PrP−/− mice. Pharmacological proteasome inhibition reversed the exacerbation of brain injury induced by PrP−/−, indicating that proteasome inhibition mediates the neuroprotective effects of PrPc. Notably, reduced proteasome activity and oxidative stress in ischemic brain tissue of PrP+/+ mice were associated with an increased abundance of hypoxia-inducible factor 1α and PACAP-38, which are known stimulants of neural progenitor cell (NPC) migration and trafficking. To elucidate effects of PrPc on intracerebral NPC homing, we intravenously infused GFP+ NPCs in ischemic WT, PrP−/− and PrP+/+ mice, showing that brain accumulation of GFP+ NPCs was greatly reduced in PrP−/− mice, but increased in PrP+/+ animals. Our results suggest that PrPc induces post-ischemic long-term neuroprotection, neurogenesis and angiogenesis in the ischemic brain by inhibiting proteasome activity.Endogenous neurogenesis persists in the adult rodent brain within distinct niches such as the subventricular zone (SVZ) of the lateral ventricles,1, 2, 3, 4 which host astrocyte-like neural stem cells and neural progenitor cells (NPCs). Focal cerebral ischemia stimulates neurogenesis, and NPCs proliferate and migrate towards the site of lesion where they eventually differentiate.5, 6, 7 In light of low differentiation rates and high cell death rates of new-born cells,6, 8, 9 post-stroke neurogenesis is scarce.10Cellular prion protein (PrPc) is a glycoprotein that is attached to cell membranes by means of a glycosylphosphatidylinositol anchor.11 Although PrPc is ubiquitously expressed, it is most abundant within the central nervous system. Conversion into its misfolded isoform PrPsc causes neurodegenerative diseases such as Creutzfeldt-Jacob disease.11, 12 While a large body of studies analyzed the role of PrPsc in the context of transmissible spongiform encephalopathies, little is known about the physiological role of PrPc. Studies performed during both ontogenesis and adulthood suggest that PrPc regulates neuronal proliferation and differentiation, synaptic plasticity and angiogenesis.13, 14, 15, 16, 17, 18 The role of these processes under pathophysiological conditions, however, is largely unknown.Previous reports suggested a role of PrPc in post-ischemic neuroprotection.19, 20, 21, 22, 23, 24 Thus, PrPc was found to be overexpressed in ischemic brain tissue.19, 20, 21, 22, 23, 24 PrPc deficiency aggravated ischemic brain injury, possibly via enhanced ERK-1/2 activation and reduced phosphorylation of Akt, thus ultimately culminating in increased caspase-3 activity,21, 24 whereas PrPc overexpression protected against ischemia.19, 20, 21, 22, 23, 24 Nevertheless, these studies focused on acute injury processes with a maximal observation period of 3 days, leaving the biological role of PrPc in post-stroke neurogenesis and angiogenesis unanswered. To clarify the role of PrPc in the post-acute ischemic brain, we herein exposed PrPc wild type (WT), PrPc knockout (PrP−/−) and PrPc overexpressing (PrP+/+) mice to focal cerebral ischemia induced by intraluminal middle cerebral artery (MCA) occlusion, evaluating effects of PrPc on neurological recovery, ischemic injury, neurogenesis and angiogenesis, as well as the homing and efficacy of exogenously delivered NPCs.  相似文献   

4.
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   

5.
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   

6.
Overgrowth of white adipose tissue (WAT) in obesity occurs as a result of adipocyte hypertrophy and hyperplasia. Expansion and renewal of adipocytes relies on proliferation and differentiation of white adipocyte progenitors (WAP); however, the requirement of WAP for obesity development has not been proven. Here, we investigate whether depletion of WAP can be used to prevent WAT expansion. We test this approach by using a hunter-killer peptide designed to induce apoptosis selectively in WAP. We show that targeted WAP cytoablation results in a long-term WAT growth suppression despite increased caloric intake in a mouse diet-induced obesity model. Our data indicate that WAP depletion results in a compensatory population of adipose tissue with beige adipocytes. Consistent with reported thermogenic capacity of beige adipose tissue, WAP-depleted mice display increased energy expenditure. We conclude that targeting of white adipocyte progenitors could be developed as a strategy to sustained modulation of WAT metabolic activity.Obesity, a medical condition predisposing to diabetes, cardiovascular diseases, cancer, and complicating other life-threatening diseases, is becoming an increasingly important social problem.1, 2, 3 Development of pharmacological approaches to reduction of body fat has remained a daunting task.4 Approved obesity treatments typically produce only moderate and temporary effects.2,5 White adipocytes are the differentiated cells of white adipose tissue (WAT) that store triglycerides in lipid droplets.6,7 In contrast, adipocytes of brown adipose tissue (BAT) dissipate excess energy through adaptive thermogenesis. Under certain conditions, white adipocytes can become partially replaced with brown-like ‘beige'' (‘brite'') adipocytes that simulate the thermogenic function of BAT adipocytes.7,8 Obesity develops in the context of positive energy balance as a result of hypertrophy and hyperplasia of white adipocytes.9Expansion and renewal of the white adipocyte pool in WAT continues in adulthood.10,11 This process is believed to rely on proliferation and self-renewal of mesenchymal precursor cells12 that we term white adipocyte progenitors (WAPs). WAPs reside within the population of adipose stromal cells (ASCs)13 and are functionally similar to bone marrow mesenchymal stem cells (MSCs).14, 15, 16 ASCs can be isolated from the stromal/vascular fraction (SVF) of WAT based on negativity for hematopoietic (CD45) and endothelial (CD31) markers.17,18 ASCs support vascularization as mural/adventitial cells secreting angiogenic factors5,19 and, unlike bone marrow MSCs, express CD34.19,20 WAPs have been identified within the ASC population based on expression of mesenchymal markers, such as platelet-derived growth factor receptor-β (PDGFRβ, aka CD140b) and pericyte markers.17,18 Recently, a distinct ASC progenitor population capable of differentiating into both white and brown adipocytes has been identified in WAT based on PDGFRα (CD140a) expression and lack of PDGFRβ expression.21,22 The physiological relevance of the two precursor populations residing in WAT has not been explored.We have previously established an approach to isolate peptide ligands binding to receptors selectively expressed on the surface of cell populations of interest.23, 24, 25, 26, 27 Such cell-targeted peptides can be used for targeted delivery of experimental therapeutic agents in vivo. A number of ‘hunter-killer'' peptides28 composed of a cell-homing domain binding to a surface marker and of KLAKLAK2 (sequence KLAKLAKKLAKLAK), a moiety inducing apoptosis upon receptor-mediated internalization, has been described by our group.26,29 Such bimodal peptides have been used for depletion of malignant cells and organ-specific endothelial cells in preclinical animal models.26,30,31 Recently, we isolated a cyclic peptide WAT7 (amino acid sequence CSWKYWFGEC) based on its specific binding to ASCs.20 We identified Δ-decorin (ΔDCN), a proteolytic cleavage fragment of decorin, as the WAT7 receptor specifically expressed on the surface of CD34+PDGFRβ+CD31-CD45- WAPs and absent on MSCs in other organs.20Here, we investigated whether WAPs are required for obesity development in adulthood. By designing a new hunter-killer peptide that directs KLAKLAK2 to WAPs through WAT7/ΔDCN interaction, we depleted WAP in the mouse diet-induced obesity model. We demonstrate that WAP depletion suppresses WAT growth. We show that, in response to WAP deficiency, WAT becomes populated with beige adipocytes. Consistent with the reported thermogenic function of beige adipocytes,32,33 the observed WAT remodeling is associated with increased energy expenditure. We identify a population of PDGFRα-positive, PDGFRβ-negative ASCs reported recently22 as a population surviving WAP depletion and responsible for WAT browning.  相似文献   

7.
Heme oxygenase-1 (HO-1), an inducible enzyme up-regulated in Alzheimer''s disease, catabolises heme to biliverdin, Fe2+ and carbon monoxide (CO). CO can protect neurones from oxidative stress-induced apoptosis by inhibiting Kv2.1 channels, which mediates cellular K+ efflux as an early step in the apoptotic cascade. Since apoptosis contributes to the neuronal loss associated with amyloid β peptide (Aβ) toxicity in AD, we investigated the protective effects of HO-1 and CO against Aβ1-42 toxicity in SH-SY5Y cells, employing cells stably transfected with empty vector or expressing the cellular prion protein, PrPc, and rat primary hippocampal neurons. Aβ1-42 (containing protofibrils) caused a concentration-dependent decrease in cell viability, attributable at least in part to induction of apoptosis, with the PrPc-expressing cells showing greater susceptibility to Aβ1-42 toxicity. Pharmacological induction or genetic over-expression of HO-1 significantly ameliorated the effects of Aβ1-42. The CO-donor CORM-2 protected cells against Aβ1-42 toxicity in a concentration-dependent manner. Electrophysiological studies revealed no differences in the outward current pre- and post-Aβ1-42 treatment suggesting that K+ channel activity is unaffected in these cells. Instead, Aβ toxicity was reduced by the L-type Ca2+ channel blocker nifedipine, and by the CaMKKII inhibitor, STO-609. Aβ also activated the downstream kinase, AMP-dependent protein kinase (AMPK). CO prevented this activation of AMPK. Our findings indicate that HO-1 protects against Aβ toxicity via production of CO. Protection does not arise from inhibition of apoptosis-associated K+ efflux, but rather by inhibition of AMPK activation, which has been recently implicated in the toxic effects of Aβ. These data provide a novel, beneficial effect of CO which adds to its growing potential as a therapeutic agent.Amongst the earliest of events leading to neuronal loss in Alzheimer''s disease (AD) is the loss of functional synapses,1, 2, 3 apparent long before deposition of amyloid β peptide (Aβ)-containing plaques.4 Although other parts of the neurone (e.g. the axon or soma) appear intact, their health at this early stage of disease progression is not clear. However, neurones ultimately die in AD and there is clear evidence that numerous events indicative of apoptosis occur even at early stages of disease progression.5, 6, 7, 8 Thus, targeting of apoptotic mechanisms may be of therapeutic value in AD as well as in other neurodegenerative disorders. Furthermore, apoptosis is established as a mechanism of neuronal loss following other types of pathological stresses including ischemia associated with stroke,9 which can predispose individuals to the development of AD.10, 11, 12Apoptosis is strongly influenced by intracellular K+ levels13 which regulate caspase activation, mitochondrial membrane potential and volume, osmolarity and cell volume.13, 14 K+ loss via K+ channels is a key early stage in apoptosis,15, 16, 17, 18, 19 and K+ channel inhibitors can protect against apoptosis triggered by numerous insults including oxidative stress.20, 21 Evidence suggests a particularly important role for the voltage-gated channel Kv2.1 in this process: expression of dominant negative Kv2.1 constructs (thus lacking functional Kv2.1 channels) protects against oxidant-induced apoptosis, and over-expression of Kv2.1 increases susceptibility to apoptosis.22, 23 Pro-apoptotic agents cause a rapid increase in the surface expression of Kv2.1 channels,24 but whether or not this occurs in AD remains to be determined. Alternative pathways recently reported to promote cell death include activation of the AMP-dependent protein kinase (AMP kinase) which can act either as a Tau kinase25 or to inhibit the mTOR pathway26 and thus contribute to neurodegeneration.Heme oxygenases (HO) are enzymes widely distributed throughout the body. In the central nervous system, HO-2 is constitutively expressed in neurones and astrocytes, while HO-1 is inducible in both cell types.27, 28, 29, 30 Both HO-1 and HO-2 break down heme to liberate biliverdin, ferrous iron (Fe2+) and carbon monoxide (CO). This catalysis is of biological significance since it is crucial to iron and bile metabolism, and also generates a highly effective antioxidant in bilirubin (from biliverdin via bilirubin reductase). Numerous stimuli can induce HO-1 gene expression,31 including oxidative stress32 and Aβ peptides.33 Importantly, HO-1 is strikingly up-regulated in AD patients, a finding considered indicative of oxidative stress.27, 34, 35 Induction of HO-1 is clearly a neuroprotective response (although in some cases can exert detrimental effects27). However, there is growing evidence that CO can be neuroprotective, for example against the damage of focal ischemia.36 Our recent studies have demonstrated that CO provides protection against oxidant-induced apoptosis by selectively inhibiting Kv2.1.23, 37 In the present study, we have investigated whether HO-1, or its product CO, can provide protection against Aβ-induced toxicity in the human neuroblastoma, SH-SY5Y, and in rat primary hippocampal neurones, and whether this involves regulation of K+ channels. We show that both HO-1 and CO protect cells against the toxicity of protofibrillar Aβ1-42 but that protection does not arise from inhibition of apoptosis-associated K+ efflux, but rather by inhibition of AMPK activation.  相似文献   

8.
9.
During virus infection and autoimmune disease, inflammatory dendritic cells (iDCs) differentiate from blood monocytes and infiltrate infected tissue. Following acute infection with hepatotropic viruses, iDCs are essential for re-stimulating virus-specific CD8+ T cells and therefore contribute to virus control. Here we used the lymphocytic choriomeningitis virus (LCMV) model system to identify novel signals, which influence the recruitment and activation of iDCs in the liver. We observed that intrinsic expression of Toso (Faim3, FcμR) influenced the differentiation and activation of iDCs in vivo and DCs in vitro. Lack of iDCs in Toso-deficient (Toso–/–) mice reduced CD8+ T-cell function in the liver and resulted in virus persistence. Furthermore, Toso–/– DCs failed to induce autoimmune diabetes in the rat insulin promoter-glycoprotein (RIP-GP) autoimmune diabetes model. In conclusion, we found that Toso has an essential role in the differentiation and maturation of iDCs, a process that is required for the control of persistence-prone virus infection.More than 500 million people worldwide suffer from chronic infections with hepatitis B or hepatitis C viruses.1 Although both viruses are poorly cytopathic, persistence of either virus can lead to chronic liver inflammation and potentially cause liversteatosis, liver cirrhosis, end-stage liver failure or hepatocellular carcinoma. Virus-specific CD8+ T cells are a major determinant governing the outcome of viral hepatitis due to their antiviral activity against virus-infected hepatocytes.2, 3, 4, 5 However, during prolonged infection, virus-specific CD8+ T cells are exhausted, resulting in their loss of function and consequently virus persistence.1, 6 Regulators influencing CD8+ T-cell function during chronic virus infection still remain ill defined.Inflammatory dendritic cells (iDCs) can develop from a subset of monocytes recruited to the site of inflammation.7, 8 This monocyte subset is characterized by the expression of CD115+/Ly6Chi/CCR2+.7 iDCs express CD11c, CD11b, and Ly6C.9, 10, 11 IDCs that exhibit tumor necrosis factor (TNF)-α production and inducible nitric oxide synthase (iNOS) were named TNF-α and iNOS producing DCs (Tip-DCs). iDCs contribute to the elimination of pathogens following bacterial infection.12, 13, 14 During infection with influenza virus, iDCs enhance CD8+ T-cell immunopathology, but have limited impact on viral replication.11, 15 According to recent observations, chronic activation of toll-like receptor 9 leads to intrahepatic myeloid-cell aggregates (iMATE).16 These aggregates, which contain iDCs, are essential for T-cell activation and therefore participate in virus control.16 Co-stimulatory signals from either direct cell contact or from cytokines in combination with continued antigen contact in iMATEs lead to proliferation and activation of virus-specific T cells.16 These observations suggest that infiltration of professional antigen-presenting cells into target organs is important for the maintenance of strong antiviral cytotoxic CD8+ T-cell activity. Factors regulating iDC infiltration into the liver remain poorly understood.Toso is a membrane protein whose extracellular domain has homology to the immunoglobulin variable (IgV) domains. The cytoplasmic region has partial homology to the FAST kinase (Fas-activated serine/threonine kinase).17 Toso is expressed on B cells and activated T cells17 and is overexpressed in B-cell lymphomas.18, 19 Expression of Toso can influence survival of macrophages.20 Originally, Toso was described as an inhibitor of FAS signaling.17, 21 More recently, a role of Toso in IgM binding and TNFR signaling was also demonstrated22, 23, 24 and consistently, Toso-deficient animals are protected from lipopolysaccharide (LPS)-induced septic shock.24, 25 Recently, we identified a role of Toso in the activation of granulocytes, monocytes, and DCs.26, 27, 28 During infection with Listeria, the expression of Toso regulated granulocyte function.26, 27 The role of Toso in the function of monocytes and other myeloid cells still remains to be further elucidated.In this study, we investigated the role of Toso during chronic viral infection by using the murine lymphocytic choriomeningitis virus (LCMV). We report that Toso promotes the differentiation and maturation of iDCs at virus-infected sites, which were essential for effector CD8+ T-cell function and in accelerating the control of the virus. We further tested the role of Toso in the rat insulin promoter-glycoprotein (RIP-GP) autoimmune diabetes model and found that Toso was required to trigger diabetes in RIP-GP mice. Taken together, we have identified an essential role of Toso in the differentiation and maturation of iDCs, which is essential for the control of persistence-prone virus infection and triggering of autoimmune disease.  相似文献   

10.
Transforming growth factor-β1 (TGF-β1) is an important regulator of fibrogenesis in heart disease. In many other cellular systems, TGF-β1 may also induce autophagy, but a link between its fibrogenic and autophagic effects is unknown. Thus we tested whether or not TGF-β1-induced autophagy has a regulatory function on fibrosis in human atrial myofibroblasts (hATMyofbs). Primary hATMyofbs were treated with TGF-β1 to assess for fibrogenic and autophagic responses. Using immunoblotting, immunofluorescence and transmission electron microscopic analyses, we found that TGF-β1 promoted collagen type Iα2 and fibronectin synthesis in hATMyofbs and that this was paralleled by an increase in autophagic activation in these cells. Pharmacological inhibition of autophagy by bafilomycin-A1 and 3-methyladenine decreased the fibrotic response in hATMyofb cells. ATG7 knockdown in hATMyofbs and ATG5 knockout (mouse embryonic fibroblast) fibroblasts decreased the fibrotic effect of TGF-β1 in experimental versus control cells. Furthermore, using a coronary artery ligation model of myocardial infarction in rats, we observed increases in the levels of protein markers of fibrosis, autophagy and Smad2 phosphorylation in whole scar tissue lysates. Immunohistochemistry for LC3β indicated the localization of punctate LC3β with vimentin (a mesenchymal-derived cell marker), ED-A fibronectin and phosphorylated Smad2. These results support the hypothesis that TGF-β1-induced autophagy is required for the fibrogenic response in hATMyofbs.Interstitial fibrosis is common to many cardiovascular disease etiologies including myocardial infarction (MI),1 diabetic cardiomyopathy2 and hypertension.3 Fibrosis may arise due to maladaptive cardiac remodeling following injury and is a complex process resulting from activation of signaling pathways, such as TGF-β1.4 TGF-β1 signaling has broad-ranging effects that may affect cell growth, differentiation and the production of extracellular matrix (ECM) proteins.5, 6 Elevated TGF-β1 is observed in post-MI rat heart7 and is associated with fibroblast-to-myofibroblast phenoconversion and concomitant activation of canonical Smad signaling.8 The result is a proliferation of myofibroblasts, which then leads to inappropriate deposition of fibrillar collagens, impaired cardiac function and, ultimately, heart failure.9, 10Autophagy is necessary for cellular homeostasis and is involved in organelle and protein turnover.11, 12, 13, 14 Autophagy aids in cell survival by providing primary materials, for example, amino acids and fatty acids for anabolic pathways during starvation conditions.15, 16 Alternatively, autophagy may be associated with apoptosis through autodigestive cellular processes, cellular infection with pathogens or extracellular stimuli.17, 18, 19, 20 The overall control of cardiac fibrosis is likely due to the complex functioning of an array of regulatory factors, but to date, there is little evidence linking autophagy with fibrogenesis in cardiac tissue.11, 12, 13, 14, 15, 16, 17, 18, 21, 22Recent studies have demonstrated that TGF-β1 may not only promote autophagy in mouse fibroblasts and human tubular epithelial kidney cells15, 23, 24 but can also inhibit this process in fibroblasts extracted from human patients with idiopathic pulmonary fibrosis.25 Moreover, it has recently been reported that autophagy can negatively15 and positively25, 26, 27 regulate the fibrotic process in different model cell systems. In this study, we have explored the putative link between autophagy and TGF-β1-induced fibrogenesis in human atrial myofibroblasts (hATMyofbs) and in a model of MI rat heart.  相似文献   

11.
Gliosis of retinal Müller glial cells may have both beneficial and detrimental effects on neurons. To investigate the role of purinergic signaling in ischemia-induced reactive gliosis, transient retinal ischemia was evoked by elevation of the intraocular pressure in wild-type (Wt) mice and in mice deficient in the glia-specific nucleotide receptor P2Y1 (P2Y1 receptor-deficient (P2Y1R-KO)). While control retinae of P2Y1R-KO mice displayed reduced cell numbers in the ganglion cell and inner nuclear layers, ischemia induced apoptotic death of cells in all retinal layers in both, Wt and P2Y1R-KO mice, but the damage especially on photoreceptors was more pronounced in retinae of P2Y1R-KO mice. In contrast, gene expression profiling and histological data suggest an increased survival of amacrine cells in the postischemic retina of P2Y1R-KO mice. Interestingly, measuring the ischemia-induced downregulation of inwardly rectifying potassium channel (Kir)-mediated K+ currents as an indicator, reactive Müller cell gliosis was found to be weaker in P2Y1R-KO (current amplitude decreased by 18%) than in Wt mice (decrease by 68%). The inner retina harbors those neurons generating action potentials, which strongly rely on an intact ion homeostasis. This may explain why especially these cells appear to benefit from the preserved Kir4.1 expression in Müller cells, which should allow them to keep up their function in the context of spatial buffering of potassium. Especially under ischemic conditions, maintenance of this Müller cell function may dampen cytotoxic neuronal hyperexcitation and subsequent neuronal cell loss. In sum, we found that purinergic signaling modulates the gliotic activation pattern of Müller glia and lack of P2Y1 has janus-faced effects. In the end, the differential effects of a disrupted P2Y1 signaling onto neuronal survival in the ischemic retina call the putative therapeutical use of P2Y1-antagonists into question.Glial cells are crucially involved in the maintenance of neuronal activity in nervous tissues.1 The homeostasis of the extracellular space is regulated by various glial functions including spatial K+ buffering, cell volume regulation and uptake of neurotransmitters.2, 3, 4 Activation of membrane receptors and ion channels is critically implicated in mediating the neuron-supportive glial functions. The dominant K+ conductance of glial cells mediates spatial K+ buffering and is important for the very negative membrane potential of these cells, thereby supporting electrogenic membrane transporters.5 Alterations in glial function are characteristic for pathological processes of the nervous system.6 Reactive gliosis may have beneficial and detrimental effects and is considered as an attempt to maintain neuronal function, protecting the tissue from further destruction, and to initiate tissue regeneration.7, 8 However, reactive gliosis may cause secondary neuronal damage as major neuron-supportive functions of glial cells get lost.6Gliotic alterations of Müller cells, the dominant macroglia of the vertebrate retina, have been observed in various models of retinal diseases.9, 10 A prominent feature of Müller cell gliosis is the downregulation of the inwardly rectifying K+ conductance mediated by inwardly rectifying K+ (Kir) channels.9 It has been demonstrated in astrocytes that downregulation or conditional knockout of Kir4.1 results in an impairment of glial glutamate (Glu) uptake.11, 12 In addition, it has been suggested that autocrine/paracrine purinergic signaling may have a causative role in the development of reactive gliosis in brain and retina.13, 14 Müller cells express different subtypes of P2 nucleotide receptors including P2Y1 and P2Y4.15, 16 P2Y1 receptors have been demonstrated to be functionally expressed by Müller cells and microglial cells, rather than by neurons.15, 16, 17, 18Retinal ischemia, a characteristic of various important human blinding diseases including diabetic retinopathy, results in neuronal degeneration and reactive gliosis.19, 20 The reduced K+ permeability of Müller cell membranes is associated with an impaired cell volume regulation under hypoosmotic stress after high intraocular pressure (HIOP)-induced ischemia.21 It has been observed that tandem-pore domain K+ channels may fulfill certain functions under conditions where Kir channels are downregulated or lacking.22, 23 A malfunctional Müller cell volume regulation was also found after deletion of P2Y1 in the mouse retina.16 It has been suggested that impaired glial K+ buffering and cell volume regulation may contribute to neuronal degeneration in the ischemic retina by inducing neuronal hyperexcitation and Glu-induced cell death.14 In order to determine whether endogenous purinergic signaling is implicated in mediating and/or protecting from neuronal degeneration, we investigated the effects of HIOP-induced ischemia in the retinae of P2Y1-deficient mice.  相似文献   

12.
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   

13.
Cisplatin (cis-diaminedichloroplatinum-II) is an extensively used chemotherapeutic agent, and one of its most adverse effects is ototoxicity. A number of studies have demonstrated that these effects are related to oxidative stress and DNA damage. However, the precise mechanism underlying cisplatin-associated ototoxicity is still unclear. The cofactor nicotinamide adenine dinucleotide (NAD+) has emerged as a key regulator of cellular energy metabolism and homeostasis. Here, we demonstrate for the first time that, in cisplatin-mediated ototoxicity, the levels and activities of SIRT1 are suppressed by the reduction of intracellular NAD+ levels. We provide evidence that the decrease in SIRT1 activity and expression facilitated by increasing poly(ADP-ribose) transferase (PARP)-1 activation and microRNA-34a through p53 activation aggravates cisplatin-mediated ototoxicity. Moreover, we show that the induction of cellular NAD+ levels using β-lapachone (β-Lap), whose intracellular target is NQO1, prevents the toxic effects of cisplatin through the regulation of PARP-1 and SIRT1 activity. These results suggest that direct modulation of cellular NAD+ levels by pharmacological agents could be a promising therapeutic approach for protection from cisplatin-induced ototoxicity.Cisplatin (cis-diamminedichloroplatinum (II)) is a chemotherapeutic agent extensively used to treat a variety of solid tumors in the head and neck, bladder, lung, ovaries, testicles, and uterus.1 However, progressive irreversible side effects of cisplatin, including nephrotoxicity and ototoxicity, greatly impair the patient''s quality of life and frequently result in the need to lower the dosage during treatment or discontinuation of the treatment. Cisplatin ototoxicity primarily occurs in the cochlea and is generally caused by apoptotic damage to the outer hair cells (OHCs), spiral ganglion cells, and the marginal cells of the stria vascularis. In recent years, studies have demonstrated that cisplatin ototoxicity is also closely related to the damage of cochlear tissue by increased production of reactive oxygen species (ROS) and accompanied by the depletion of antioxidant substances and increased lipid peroxidation.2, 3 ROSs, particularly the hydroxyl radical, have a critical role in cisplatin-induced p53 activation through DNA damage.4 Although it is not easy to differentiate the cause from the consequence, a positive feedback loop between inflammatory cytokines and oxidative stress that worsen the cochlear damage is considered as one of the major mechanisms that facilitate cisplatin-induced hearing impairment.5 Interestingly, p53 and NF-κB have been described as key mediators of cisplatin-induced toxicity because of their involvement in oxidative stress, DNA damage, and inflammation through a mutual feedback process of ‘cause and effect.''6, 7 In addition, activities of p53 and NF-κB could be regulated by post-translational modifications, including phosphorylation and acetylation. Recent studies have reported that acetylated p53 and NF-κB are correlated with cisplatin-induced toxicity. Furthermore, acetylation of p53 and NF-κB is critically involved in cisplatin-induced renal injury.8, 9Cellular nicotinamide adenine dinucleotide (NAD+) and NADH levels have been shown to be important mediators of energy metabolism and cellular homeostasis.10, 11 As NAD+ acts as a cofactor for various enzymes, including sirtuins (SIRTs), poly(ADP-ribose) transferases (PARPs), and cyclic ADP (cADP)-ribose synthases,12, 13 the regulation of NAD+ level may have therapeutic benefits through its effect on NAD+-dependent enzymes. SIRTs, NAD+-dependent protein deacetylases, are present as seven homologs of Sir2 (SIRT1-7) that show differential subcellular localizations in mammals.11 Among these, nuclear SIRT1 is activated under energy stress conditions, such as fasting, exercise, or low glucose availability. 14 SIRT1 has a key role in metabolism, development, stress response, neurogenesis, hormone responses, and apoptosis15, 16 by deacetylation of substrates, such as NF-κB, FOXO, p53, and histones.17, 18, 19PARPs, the most abundant ADP-ribosyl transferases, also use NAD+ to generate large amounts of poly(ADP-ribose) (PAR), which facilitate the recruitment of DNA repair factors. In particular, PARP-1 is a DNA damage sensor that can be activated in response to DNA damage by various pathophysiological conditions, including oxidative stress and inflammatory injury. However, excessive hyperactivation of PARP-1 causes the depletion of intracellular NAD+ and ATP levels, which eventually leads to cell death.20, 21 PARP-1 activation is also known as one of the important pathogenic mechanisms in cisplatin-induced toxicity.22, 23A cytosolic antioxidant flavoprotein NADH:quinone oxidoreductase 1 (NQO1) catalyzes the reduction of quinones to hydroquinones by utilizing NADH as an electron donor, which consequently increases intracellular NAD+ levels.24, 25 In addition, accumulation evidence suggests that NQO1 has a role in other biological activities, including anti-inflammatory processes, scavenging of superoxide anion radicals, and stabilization of p53 and other tumor suppressor proteins.26, 27, 28 Several substrates of NQO1 enzyme, including mitomycin C, RH1, AZQ, Coenzyme Q10, and idebenone, have been identified,29, 30 of which β-lapachone (3,4-dihydro-2,2-dimethyl-2H-naphtho[1,2-b]pyran-5,6-dione; β-Lap) is recently well studied as a substrate of NQO1.31, 32 β-Lap was first isolated from the bark of the lapacho tree and reported to inhibit tumor cell line growth.33 However, recent reports indicate that the conversion of NADH to NAD+ by NQO1 and β-Lap has beneficial effects on several characteristics of metabolic syndrome, for example, prevention of health decline in aged mice, amelioration of obesity or hypertension, prevention of arterial restenosis, and protection against salt-induced renal injury.34, 35, 36, 37, 38 Furthermore, we recently have demonstrated that conversion of NADH to NAD+ by NQO1 and β-Lap suppresses cisplatin-induced acute kidney injury by downregulating potential damage mediators such as oxidative stress and inflammatory responses.9Although a link between NAD+-dependent molecular events and cellular metabolism is evident, it remains unclear whether modulation of NAD+ levels has an impact on cisplatin-induced hearing impairment. Therefore, herein we investigated the role of NAD+ metabolism on cisplatin-induced cochlear dysfunction, and the effect of increased levels of intracellular NAD+ facilitated by β-Lap on cisplatin-induced hearing impairment with a particular interest in NAD+-dependent enzymatic pathways including SIRTs and PARPs.  相似文献   

14.
Extracellular calmodulin (ExtCaM) regulates stomatal movement by eliciting a cascade of intracellular signaling events including heterotrimeric G protein, hydrogen peroxide (H2O2), and Ca2+. However, the ExtCaM-mediated guard cell signaling pathway remains poorly understood. In this report, we show that Arabidopsis (Arabidopsis thaliana) NITRIC OXIDE ASSOCIATED1 (AtNOA1)-dependent nitric oxide (NO) accumulation plays a crucial role in ExtCaM-induced stomatal closure. ExtCaM triggered a significant increase in NO levels associated with stomatal closure in the wild type, but both effects were abolished in the Atnoa1 mutant. Furthermore, we found that ExtCaM-mediated NO generation is regulated by GPA1, the Gα-subunit of heterotrimeric G protein. The ExtCaM-dependent NO accumulation was nullified in gpa1 knockout mutants but enhanced by overexpression of a constitutively active form of GPA1 (cGα). In addition, cGα Atnoa1 and gpa1-2 Atnoa1 double mutants exhibited a similar response as did Atnoa1. The defect in gpa1 was rescued by overexpression of AtNOA1. Finally, we demonstrated that G protein activation of NO production depends on H2O2. Reduced H2O2 levels in guard cells blocked the stomatal response of cGα lines, whereas exogenously applied H2O2 rescued the defect in ExtCaM-mediated stomatal closure in gpa1 mutants. Moreover, the atrbohD/F mutant, which lacks the NADPH oxidase activity in guard cells, had impaired NO generation in response to ExtCaM, and H2O2-induced stomatal closure and NO accumulation were greatly impaired in Atnoa1. These findings have established a signaling pathway leading to ExtCaM-induced stomatal closure, which involves GPA1-dependent activation of H2O2 production and subsequent AtNOA1-dependent NO accumulation.Plant guard cells control opening and closure of the stomata in response to phytohormones (e.g. abscisic acid [ABA]) and various environmental signals such as light and temperature, thereby regulating gas exchange for photosynthesis and water status via transpiration (Schroeder et al., 2001). Cytosolic calcium ([Ca2+]i) has been shown to be a key second messenger that changes in response to multiple stimuli in guard cells (McAinsh et al., 1995; Grabov and Blatt, 1998; Wood et al., 2000). A large proportion of Ca2+ is localized in extracellular space. It has been shown that external Ca2+ concentration ([Ca2+]o) promotes stomatal closure and induces oscillation in [Ca2+]i in guard cells (MacRobbie, 1992; McAinsh et al., 1995; Allen et al., 2001). However, how the guard cells perceive [Ca2+]o concentration and convert [Ca2+]o changes into [Ca2+]i changes was not understood until a calcium-sensing receptor (CAS) in the plasma membrane of guard cells in Arabidopsis (Arabidopsis thaliana) was identified (Han et al., 2003). The external Ca2+ (Ca2+o)-induced [Ca2+]i increase is abolished in CAS antisense lines (Han et al., 2003). Both [Ca2+]o and [Ca2+]i show diurnal oscillation that is determined by stomatal conductance, whereas the amplitude of [Ca2+]i oscillation is reduced in CAS antisense lines (Tang et al., 2007). The reduced amplitude of [Ca2+]i diurnal oscillation in response to Ca2+o treatment suggests the potential existence of other [Ca2+]o sensor(s) that may transmit [Ca2+]o information into the [Ca2+]i response in coordination with CAS. Extracellular calmodulin (ExtCaM) could be such an additional [Ca2+]o sensor.Calmodulin is a well-known Ca2+ sensor that is activated upon binding of Ca2+. It has been shown that calmodulin exists not only intracellularly but also extracellularly in many plant species (Biro et al., 1984; Sun et al., 1994, 1995; Cui et al., 2005). ExtCaM has been implicated in several important biological functions, such as the promotion of cell proliferation, pollen germination, and tube growth (Sun et al., 1994, 1995; Ma and Sun, 1997; Ma et al., 1999; Cui et al., 2005; Shang et al., 2005). ExtCaM is found in the cell wall of guard cells in Vicia faba and in the epidermis of Arabidopsis by immunogold labeling/electron microscopy and western-blot analyses, respectively, and the endogenous CaM in the extracellular space has been shown to regulate stomatal movements (Chen et al., 2003; Xiao et al., 2004). Under natural conditions, once the activity of ExtCaM has been inhibited by its membrane-impermeable antagonist W7-agrose or CaM antibody, stomatal opening under light is enhanced and stomatal closure in darkness is inhibited in V. faba and Arabidopsis (Chen et al., 2003; Xiao et al., 2004). [Ca2+]i and cytosolic hydrogen peroxide (H2O2) changes, two events involved in ExtCaM-regulated stomatal movement (Chen et al., 2004), are likely regulated by light/darkness (Chen and Gallie, 2004; Tang et al., 2007), suggesting that ExtCaM plays an important physiological role in the regulation of stomatal diurnal rhythm. Calmodulin-binding proteins have been found in the protoplast of suspension-cultured Arabidopsis cells, supporting the idea that ExtCaM functions as a peptide-signaling molecule (Cui et al., 2005). Furthermore, ExtCaM triggers [Ca2+]i elevation in guard cells of V. faba and Arabidopsis and in lily (Lilium daviddi) pollen (Chen et al., 2004; Xiao et al., 2004; Shang et al., 2005). These observations support the notion that ExtCaM could be a potential [Ca2+]o sensor for external calcium, and this external calcium sensing could subsequently regulate the [Ca2+]i level through a signaling cascade.It is interesting that ExtCaM and ABA induce some parallel changes in second messengers in guard cell signaling. Our previous studies show that ExtCaM induces [Ca2+]i increase and H2O2 generation through the Gα-subunit (GPA1) of a heterotrimeric G protein, and increased H2O2 further elevates [Ca2+]i (Chen et al., 2004). G protein, Ca2+, and H2O2 are well-known second messengers in ABA-induced guard cell signaling (McAinsh et al., 1995; Grabov and Blatt, 1998; Pei et al., 2000; Wang et al., 2001; Zhang et al., 2001; Liu et al., 2007). However, the signaling cascade triggered by ExtCaM in guard cells is poorly understood. New ABA signaling components in guard cells could provide a clue in the study of the molecular mechanism of ExtCaM guard cell signaling.Recently, nitric oxide (NO) has been shown to serve as an important signal molecule involved in many aspects of developmental processes, including floral transition, root growth, root gravitropism, adventitious root formation, xylogenesis, seed germination, and orientation of pollen tube growth (Beligni and Lamattina, 2000; Pagnussat et al., 2002; He et al., 2004; Prado et al., 2004; Gabaldón et al., 2005; Stohr and Stremlau, 2006). Increasing evidence points to a role for NO as an essential component in ABA signaling in guard cells (Garcia-Mata and Lamattina, 2001, 2002; Neill et al., 2002). It has been shown that nitrate reductase (NR) reduces nitrite to NO, and the nia1, nia2 NR-deficient mutant in Arabidopsis showed reduced ABA induction of stomatal closure (Desikan et al., 2002; Bright et al., 2006). Although animal nitric oxide synthase (NOS) activity has been detected in plants and inhibitors of mammalian NOS impair NO production in plants (Barroso et al., 1999; Corpas et al., 2001), the gene(s) encoding NOS in plants is still not clear. AtNOS1 in Arabidopsis was initially reported to encode a protein containing NOS activity (Guo et al., 2003). However, recent studies have raised critical questions regarding the nature of AtNOS1 and suggested that AtNOS1 appears not to encode a NOS (Crawford et al., 2006; Zemojtel et al., 2006). However, the originally described Atnos1 mutant is deficient in NO accumulation (Crawford et al., 2006). Consequently, AtNOS1 was renamed AtNOA1 (for NITRIC OXIDE ASSOCIATED1; Crawford et al., 2006). Therefore, the Atnoa1 mutant provides a useful tool for dissecting the function of NO in plants. At present, the molecules that regulate NO generation in ABA-mediated guard cell signaling are not clear. Evidence suggests that H2O2, a second messenger important for the regulation of many developmental processes and stomatal movement (Pei et al., 2000; Zhang et al., 2001; Coelho et al., 2002; Demidchik et al., 2003; Kwak et al., 2003), regulates NO generation in guard cells (Lum et al., 2002; He et al., 2005; Bright et al., 2006).Given the parallel signaling events induced by ABA and ExtCaM, we investigated whether NO is involved in the regulation of ExtCaM-induced stomatal closure in Arabidopsis and whether it is linked to G protein and H2O2, two key regulators of both ExtCaM and ABA regulation of stomatal movements. Using Arabidopsis mutants (e.g. GPA1 null mutants, the NO-producing mutant Atnoa1, and the guard cell H2O2 synthetic enzymatic mutant atrbohD/F) combined with pharmacological analysis, we present compelling evidence to establish a linear functional relationship between Gα, H2O2, and NO in ExtCaM guard cell signaling.  相似文献   

15.
Using the scanning ion-selective electrode technique, fluxes of H+, Na+, and Cl were investigated in roots and derived protoplasts of salt-tolerant Populus euphratica and salt-sensitive Populus popularis 35-44 (P. popularis). Compared to P. popularis, P. euphratica roots exhibited a higher capacity to extrude Na+ after a short-term exposure to 50 mm NaCl (24 h) and a long term in a saline environment of 100 mm NaCl (15 d). Root protoplasts, isolated from the long-term-stressed P. euphratica roots, had an enhanced Na+ efflux and a correspondingly increased H+ influx, especially at an acidic pH of 5.5. However, the NaCl-induced Na+/H+ exchange in root tissues and cells was inhibited by amiloride (a Na+/H+ antiporter inhibitor) or sodium orthovanadate (a plasma membrane H+-ATPase inhibitor). These results indicate that the Na+ extrusion in stressed P. euphratica roots is the result of an active Na+/H+ antiport across the plasma membrane. In comparison, the Na+/H+ antiport system in salt-stressed P. popularis roots was insufficient to exclude Na+ at both the tissue and cellular levels. Moreover, salt-treated P. euphratica roots retained a higher capacity for Cl exclusion than P. popularis, especially during a long term in high salinity. The pattern of NaCl-induced fluxes of H+, Na+, and Cl differs from that caused by isomotic mannitol in P. euphratica roots, suggesting that NaCl-induced alternations of root ion fluxes are mainly the result of ion-specific effects.Soil salinity causes increasingly agricultural and environmental problems on a worldwide scale, especially in arid areas. When plant roots are subjected to saline environments with high NaCl content, external Na+ and Cl establish a large electrochemical gradient favoring the passive entry of salt ions through a variety of cation and anion channels and/or transporters in the plasma membrane (PM; Blumwald et al., 2000; Hasegawa et al., 2000; White and Broadley, 2001; Roberts, 2006; Demidchik and Maathuis, 2007). The entry and accumulation of toxic ions lead to disruption of ion homeostasis and finally cause secondary stress, e.g. oxidative bursts (Zhu, 2001, 2003). Accordingly, the maintenance of low salt concentration in the cytosol is of great importance for salt adaptation of plants (Greenway and Munns, 1980; Munns and Tester, 2008).Active Na+ extrusion to the apoplast or external environment is essential for sustaining Na+ homeostasis in the cytosol (Blumwald et al., 2000; Tester and Davenport, 2003; Zhu, 2003; Apse and Blumwald, 2007). PM Na+/H+ antiporters have been widely considered to play a crucial role in active Na+ extrusion under saline conditions (Shi et al., 2000, 2002; Qiu et al., 2002; Martínez-Atienza et al., 2007). NaCl-induced activity of PM Na+/H+ antiporter has been reported in crop species, tomato (Solanum lycopersicum; Wilson and Shannon, 1995), Arabidopsis (Arabidopsis thaliana; Qiu et al., 2002, 2003), and rice (Oryza sativa; Martínez-Atienza et al., 2007). Furthermore, overexpression of the Na+/H+ antiporter gene AtSOS1 decreases the accumulation of Na+ in transgenic Arabidopsis under NaCl stress (Shi et al., 2003). These PM Na+/H+ antiporters depend on electrochemical H+ gradients, which are generated by PM H+-ATPase (Blumwald et al., 2000; Zhu, 2003). Using an ion-selective microelectrode, Shabala and a colleague suggested the involvement of PM H+-ATPase in the Na+/H+ antiport according to H+ kinetics on salt shock (Shabala, 2000; Shabala and Newman, 2000). Therefore, the NaCl-induced H+ pumping is fundamental to Na+/H+ exchange and salinity tolerance (Ayala et al., 1996; Vitart et al., 2001; Chen et al., 2007; Gévaudant et al., 2007). However, the active Na+/H+ antiport across PM and the contribution to salt exclusion have been rarely investigated in tree species.Munns and Tester (2008) claimed that Cl toxicity is more important than Na+ toxicity in some woody species, e.g. citrus. Similarly, we have noticed that the inability to restrict Cl uptake contributes to the NaCl-induced salt damage in salt-sensitive poplar (Populus spp.) species, in addition to toxicity of excess Na+ (Chen et al., 2001, 2002, 2003). The differences in Cl tolerance exhibited by plants are usually related to the ability to restrict Cl transport to the aerial part (Greenway and Munns, 1980; White and Broadley, 2001). Excluding Cl from the xylem seems to be an effective mechanism for lotus to cope with the interactive effect of salt and water logging (Teakle et al., 2007). An influx of Cl, immediately after addition of NaCl, was observed in bean (Vicia faba) mesophyll tissue (Shabala, 2000). The Cl flux response to salt shock is helpful to reveal the rapid adjustments of plants to salinity. However, Cl fluxes in salt-adapted roots, which are necessary to clarify plant adaptations to long durations of salinity, have not been examined.Populus euphratica has been widely considered as a model plant to elucidate physiological and molecular mechanisms of salt tolerance in woody species (Chen et al., 2001, 2002, 2003; Gu et al., 2004; Ottow et al., 2005a, 2005b; Junghans et al., 2006; Wang et al., 2007, 2008; Wu et al., 2007; Zhang et al., 2007). Comparative studies have shown that salt-stressed P. euphratica seedlings accumulate less Na+ and Cl in root and shoot tissues than salt-sensitive poplar species (Chen et al., 2001, 2002). It is suggested that the greater capacity to exclude NaCl in P. euphratica is likely the result of salt uptake and transport restriction in roots (Chen et al., 2002, 2003). However, this needs further investigations, e.g. by electrophysiology, to clarify.In this study, we used a noninvasive ion flux technique to measure the tissue and cellular fluxes of H+, Na+, and Cl in roots of the salt-tolerant P. euphratica and salt-sensitive P. popularis 35-44. The aim was to compare the NaCl-induced alternations of ion fluxes in poplar species differing in salt tolerance. Furthermore, we examined the effects of pH, salt shock, and PM transport inhibitors on Na+ and H+ fluxes in root-derived protoplasts of the salt-tolerant species, P. euphratica, which exhibited an evident Na+ exclusion under saline conditions.  相似文献   

16.
17.
18.
19.
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   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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