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.
The accumulation of amyloid-β protein (Aβ) is an early event associated with synaptic and mitochondrial damage in Alzheimer''s disease (AD). Recent studies have implicated the filamentous actin (F-actin) severing protein, Cofilin, in synaptic remodeling, mitochondrial dysfunction, and AD pathogenesis. However, whether Cofilin is an essential component of the AD pathogenic process and how Aβ impinges its signals to Cofilin from the neuronal surface are unknown. In this study, we found that Aβ42 oligomers (Aβ42O, amyloid-β protein 1–42 oligomers) bind with high affinity to low or intermediate activation conformers of β1-integrin, resulting in the loss of surface β1-integrin and activation of Cofilin via Slingshot homology-1 (SSH1) activation. Specifically, conditional loss of β1-integrin prevented Aβ42O-induced Cofilin activation, and allosteric modulation or activation of β1-integrin significantly reduced Aβ42O binding to neurons while blocking Aβ42O-induced reactive oxygen species (ROS) production, mitochondrial dysfunction, depletion of F-actin/focal Vinculin, and apoptosis. Cofilin, in turn, was required for Aβ42O-induced loss of cell surface β1-integrin, disruption of F-actin/focal Talin–Vinculin, and depletion of F-actin-associated postsynaptic proteins. SSH1 reduction, which mitigated Cofilin activation, prevented Aβ42O-induced mitochondrial Cofilin translocation and apoptosis, while AD brain mitochondria contained significantly increased activated/oxidized Cofilin. In mechanistic support in vivo, AD mouse model (APP (amyloid precursor protein)/PS1) brains contained increased SSH1/Cofilin and decreased SSH1/14-3-3 complexes, indicative of SSH1–Cofilin activation via release of SSH1 from 14-3-3. Finally, genetic reduction in Cofilin rescued APP/Aβ-induced synaptic protein loss and gliosis in vivo as well as deficits in long-term potentiation (LTP) and contextual memory in APP/PS1 mice. These novel findings therefore implicate the essential involvement of the β1-integrin–SSH1–Cofilin pathway in mitochondrial and synaptic dysfunction in AD.The defining pathological hallmark of Alzheimer''s disease (AD) is the accumulation of amyloid-β protein (Aβ) in brain associated with tau pathology, synapse loss, cytoskeletal aberrations, mitochondrial dysfunction, and cognitive decline. Soluble oligomeric forms of Aβ are thought to be the most toxic species, resulting in synaptic loss and downstream neurotoxicity.1 An early and consistent impairment secondary to Aβ oligomer treatment in primary neurons is the shrinkage of dendritic spines2 involving the rearrangement of filamentous actin (F-actin) cytoskeleton in spines and loss of spine-associated proteins such as postsynaptic density-95 (PSD95) and Drebrin,3, 4 as well as impaired mitochondrial function.5, 6 Studies have implicated an involvement of the F-actin-severing protein Cofilin in Aβ-induced dendritic spine changes,3, 4 accumulation of Cofilin–Actin aggregates/rods in AD brains,7 and increased Cofilin activity in brains of AD patients.8 Cofilin normally functions as a key regulator of Actin dynamics that destabilizes F-actin. Cofilin is inactivated by phosphorylation on Ser3 by LIM kinase 1 (LIMK1), whereas its dephosphorylation by Slingshot homology-1 (SSH1) activates Cofilin.9, 10, 11 Upon oxidative stress and/or Ca2+ elevation,9, 12, 13 SSH1 is activated and active Cofilin becomes oxidized on cysteine residues, resulting in rapid mitochondrial translocation to promote mitochondria-mediated apoptosis and induction of Cofilin–Actin pathology.14, 15 Despite the circumstantial evidence for the involvement of Cofilin in AD pathogenesis, no direct evidence thus far has been presented.Heterodimeric integrins (α- and β-subunits) comprise major adhesion receptors that regulate multiple facets of cellular function, including adhesion, motility, survival, and synaptic plasticity.16 A primary function of integrins is to link the extracellular matrix to the F-actin cytoskeleton via structural scaffolding proteins such as Talin and Vinculin.17, 18 Among several proposed surface Aβ oligomer receptors such as PrPc (Prion protein (cellular))/mGluR5 (metabotropic glutamate receptor 5)19, 20 and paired immunoglobin-like receptor-B (Pir-B)8, it has been shown that α2/β1 and αv/β1-integrins are also required to mediate Aβ-induced apoptosis and impairment in LTP.21, 22 However, whether Aβ binds directly to integrins and how Aβ engagement alters downstream integrin function are unknown. In this study, we explored the mechanistic relationships among Aβ42O (amyloid-β protein 1–42 oligomers), β1-integrin, and Cofilin activities in vitro, HT22 cells, primary neurons, and genetically modified mice. Here we show that Aβ42O exhibits direct high-affinity binding β1-integrin, inducing its conformational alteration, loss of surface β1-integrin, and disruption of integrin-associated focal complexes (FCs), as well as mitochondrial and synaptic dysfunction via a pathway involving ROS-dependent activation of SSH1 and Cofilin.  相似文献   

2.
3.
4.
The use of β-lactam antibiotics has led to the evolution and global spread of a variety of resistance mechanisms, including β-lactamases, a group of enzymes that degrade the β-lactam ring. The evolution of increased β-lactam resistance was studied by exposing independent lineages of Salmonella typhimurium to progressive increases in cephalosporin concentration. Each lineage carried a β-lactamase gene (blaTEM-1) that provided very low resistance. In most lineages, the initial response to selection was an amplification of the blaTEM-1 gene copy number. Amplification was followed in some lineages by mutations (envZ, cpxA, or nmpC) that reduced expression of the uptake functions, the OmpC, OmpD, and OmpF porins. The initial resistance provided by blaTEM-1 amplification allowed the population to expand sufficiently to realize rare secondary point mutations. Mathematical modeling showed that amplification often is likely to be the initial response because events that duplicate or further amplify a gene are much more frequent than point mutations. These models show the importance of the population size to appearance of later point mutations. Transient gene amplification is likely to be a common initial mechanism and an intermediate in stable adaptive improvement. If later point mutations (allowed by amplification) provide sufficient adaptive improvement, the amplification may be lost.THE extensive use of β-lactam antibiotics has led to the evolution and spread of many chromosomal-, plasmid-, and transposon-borne resistance mechanisms (Livermore 1995; Weldhagen 2004). Prominent among these mechanisms is a class of enzymes, β-lactamases, that hydrolyze the β-lactam ring (Ambler 1980; Poole 2004). TEM-1 β-lactamase, encoded by the blaTEM-1 gene, hydrolyzes both penicillins and early cephalosporins (Matagne et al. 1990). As bacteria developed resistance, stable extended-spectrum cephalosporins (ESCs) were introduced, leading to evolution of TEM sequence variants with improved ESC hydrolysis (Petrosino et al. 1998). Resistance to β-lactams can also result from mutations that reduce levels of outer membrane proteins involved in uptake, altered target proteins (penicillin-binding proteins) to reduce β-lactam binding, or increased expression of efflux pumps that export the antibiotics (Poole 2004; Martínez-Martínez 2008; Zapun et al. 2008).Resistance to β-lactam antibiotics is linearly correlated with the lactamase level over a large range (Nordström et al. 1972) and resistance to β-lactam antibiotics can be provided by increasing enzyme levels. An early illustration of this process is the finding that Escherichia coli can develop ampicillin resistance by amplifying its ampC gene (Edlund and Normark 1981). Similar amplification has been observed in both eubacteria and eukaryotes (Craven and Neidle 2007; Wong et al. 2007) in response to various selective pressures, including antibiotics (Andersson and Hughes 2009; Sandegren and Andersson 2009). In an unselected bacterial population, the frequency of cells with a duplication of any specific chromosomal region ranges between 10−2 and 10−5 depending on the region (Anderson and Roth 1981), whereas a point mutation in that gene is expected to be carried by perhaps 1 cell in 107–108 (Hudson et al. 2002). Thus, the rate of duplication formation is ∼10−5/cell/division and further increases ∼0.01/cell/division (Pettersson et al. 2008) while the base substitution rate is ∼10−10/cell/division/base pair (Hudson et al. 2002). Thus, it is apparent that variants with an increased level of any enzyme activity are more likely to owe the increase to a gene copy number change than to a point mutation. Furthermore, because of the high intrinsic instability of tandem amplifications, haploid segregants are expected to take over the population when the selection pressure is released (Pettersson et al. 2008).To examine the importance of gene amplification in bacterial adaptation to cephalosporins, several independent Salmonella typhimurium lineages carrying the blaTEM-1 gene were allowed to develop resistance to progressively increased concentrations of cephalothin (a first-generation cephalosporin) and cefaclor (a second-generation cephalosporin). As these lineages developed resistance to higher antibiotic levels, amplification of the blaTEM-1 gene was the primary and most common resistance mechanism, which in some cases was followed by acquisition of rare point mutations that provided stable resistance.  相似文献   

5.
Heterotrimeric G proteins have been shown to transmit ultraviolet B (UV-B) signals in mammalian cells, but whether they also transmit UV-B signals in plant cells is not clear. In this paper, we report that 0.5 W m−2 UV-B induces stomatal closure in Arabidopsis (Arabidopsis thaliana) by eliciting a cascade of intracellular signaling events including Gα protein, hydrogen peroxide (H2O2), and nitric oxide (NO). UV-B triggered a significant increase in H2O2 or NO levels associated with stomatal closure in the wild type, but these effects were abolished in the single and double mutants of AtrbohD and AtrbohF or in the Nia1 mutants, respectively. Furthermore, we found that UV-B-mediated H2O2 and NO generation are regulated by GPA1, the Gα-subunit of heterotrimeric G proteins. UV-B-dependent H2O2 and NO accumulation were nullified in gpa1 knockout mutants but enhanced by overexpression of a constitutively active form of GPA1 (cGα). In addition, exogenously applied H2O2 or NO rescued the defect in UV-B-mediated stomatal closure in gpa1 mutants, whereas cGα AtrbohD/AtrbohF and cGα nia1 constructs exhibited a similar response to AtrbohD/AtrbohF and Nia1, respectively. Finally, we demonstrated that Gα activation of NO production depends on H2O2. The mutants of AtrbohD and AtrbohF had impaired NO generation in response to UV-B, but UV-B-induced H2O2 accumulation was not impaired in Nia1. Moreover, exogenously applied NO rescued the defect in UV-B-mediated stomatal closure in the mutants of AtrbohD and AtrbohF. These findings establish a signaling pathway leading to UV-B-induced stomatal closure that involves GPA1-dependent activation of H2O2 production and subsequent Nia1-dependent NO accumulation.Heterotrimeric G proteins, composed of α-, β-, and γ-subunits, are a key intracellular signaling molecule in both mammalian and plant systems. Classically, upon signal reception by a receptor coupled to the heterotrimer, the Gα-subunit separates from the Gβγ dimer, and either Gα or the Gβγ dimer can act as a functional unit and induce downstream signaling (Oldham and Hamm, 2008). In contrast to mammalian cells, where multiple α, β, and γ genes exist, there is only one prototypical Gα (GPA1), one Gβ (AGB1), and two known Gγ (AGG1 and AGG2) genes in Arabidopsis (Arabidopsis thaliana; Temple and Jones, 2007). Despite the comparative simplicity of players, G proteins have been shown to participate in multiple signaling pathways in Arabidopsis, including developmental processes, phytohormone responses, and responses to biotic and abiotic environmental signals such as pathogens, ozone, drought, and light (Assmann, 2005; Temple and Jones, 2007; Warpeha et al., 2007; Okamoto et al., 2009; Nilson and Assmann, 2010).Depletion of the stratospheric ozone layer results in increased levels of the sun’s UV-B radiation (280–315 nm) at the Earth’s surface. Although this influx of shortwave photons with high energy implies serious effects for all living organisms (Frohnmeyer and Staiger, 2003), UV-B is also a key environmental signal that initiates diverse responses in a range of organisms (Jansen and Bornman, 2012). Thus, understanding the mechanism of UV-B signal transduction in cells is very important. In recent years, significant progress has been made in identifying the molecular players and understanding the early mechanisms and functions of the UV-B perception and signaling pathway in plants. The perception of UV-B by UV RESISTANCE LOCUS8 (UVR8) followed by the interaction among UVR8, CONSTITUTIVELY PHOTOMORPHOGENIC1 (COP1), and ELONGATED HYPOCOTYL5 (HY5) has emerged as a primary mechanism of the UV-B response that is crucial for UV-B acclimation and tolerance (Rizzini et al., 2011; Christie et al., 2012; Heijde and Ulm, 2012; Jansen and Bornman, 2012). However, few of the molecular players involved in UV-B signal transduction are currently known. In mammalian cells, heterotrimeric G proteins have been shown to mediate various UV-B-induced cellular responses, such as secretion of heparin-binding epidermal growth factor (HB-EGF), activation of mitogen-activated protein kinases, cyclooxygenase2 expression, and apoptosis in human keratinocytes (Seo et al., 2004, 2007; Seo and Juhnn, 2010), suggesting that G proteins are important molecular players in UV-B signal transduction. However, at present, whether G proteins participate in the responses of plant cells to UV-B is not known.Stomata embedded in the epidermis of terrestrial plants are important for CO2 absorption and water transpiration and are possible points of entry for pathogens. Thus, the regulation of stomatal apertures is extremely important for the survival of plants. Phenotypic analyses of Arabidopsis mutants lacking the Gα- or Gβ-subunit show that these G proteins are involved in stomatal movement regulated by abscisic acid (ABA; Wang et al., 2001; Pandey and Assmann, 2004; Liu et al., 2007; Fan et al., 2008; Zhang et al., 2011), sphingosine-1-P (Coursol et al., 2003, 2005), phosphatidic acid (PA; Mishra et al., 2006), extracellular calmodulin (ExtCaM; Chen et al., 2004; Li et al., 2009), extracellular ATP (Hao et al., 2012), and the pathogen-associated molecular pattern flg22 (Zhang et al., 2008), suggesting that plant G proteins respond to various stimuli as key regulators of stomatal movement. On exposure to UV-B radiation, many plant species exhibit decreases in stomatal conductance and/or aperture under growth chamber, greenhouse, and field conditions (Musil and Wand, 1993; Nogués et al., 1999; Jansen and Noort, 2000). However, in some species, UV-B has been reported to induce either stomatal opening or stomatal closure, perhaps depending on the metabolic state of guard cells (Jansen and Noort, 2000). Furthermore, UV-B-inhibited photosynthesis is partially caused by stomatal limitation (He et al., 2004). Thus, understanding the mechanism of stomatal movement regulated by UV-B is extremely important for improving the resistance of plants to enhanced UV-B radiation, but, to date, it is poorly understood.Recently, compelling evidence emerged that hydrogen peroxide (H2O2) and nitric oxide (NO) function as signaling molecules in plants, mediating a range of responses to environmental stress including UV-B radiation (Neill et al., 2002; Qiao and Fan, 2008; Wilson et al., 2008). Increasing evidence also points to the role for H2O2 and NO as essential components in guard cell signaling. For example, both H2O2 and NO have been implicated in ABA-, salicylic acid (SA)-, ethylene-, ExtCaM-, and darkness-induced stomatal closure. Furthermore, several main cellular players in stomatal movement, such as mitogen-activated protein kinases, protein phosphatases, cytoskeleton, and ion channels, have already been identified as likely targets downstream of H2O2 or NO (Neill et al., 2008; Wang and Song, 2008; Huang et al., 2009; Li et al., 2009; Wilkins et al., 2011; Yemets et al., 2011). G protein signaling to the membrane-bound H2O2 synthetic enzyme, NADPH oxidase, has been implicated in the development of disease resistance and the apoptotic hypersensitive response in rice (Oryza sativa; Suharsono et al., 2002). Production of reactive oxygen species (ROS) in response to the air pollutant ozone is also impaired in a mutant lacking the Gα subunit (Joo et al., 2005). The heterotrimeric G proteins also participate in ROS metabolism in plant cells (Wei et al., 2008; Zhao et al., 2010). During stomatal movement, G proteins mediate H2O2 production induced by ABA (Zhang et al., 2011), ExtCaM (Chen et al., 2004; Li et al., 2009), and extracellular ATP (Hao et al., 2012) as well as NO production induced by ExtCaM in guard cells (Li et al., 2009). In addition, phospholipase Dα and its product PA, which interact with GPA1 during ABA inhibition of stomatal opening (Mishra et al., 2006), also promote ABA-induced ROS production (Zhang et al., 2009). These observations suggest that G proteins may be key regulators of H2O2 and NO production in plant cells, including guard cells. With regard to the stomatal movement regulated by UV-B radiation, our previous studies showed that H2O2 and NO generation are required for UV-B-induced stomatal closure (He et al., 2005, 2011a, 2011b). However, whether the UV-B-induced production of H2O2 and NO in guard cells is also regulated by G proteins remains unknown.In this study, we use Arabidopsis mutants (e.g. GPA1 null mutants gpa1-1 and gpa1-2; Nia1-2, Nia2-1, and Nia1-2/Nia2-5, which are defective in NO production; and AtrbohD, AtrbohF, and AtrbohD/AtrbohF, which are defective in producing H2O2) and pharmacological reagents to show that the G protein is involved in the regulation of UV-B-induced stomatal closure in Arabidopsis via sequential elucidation of H2O2 and NO, two key regulators of UV-B regulation of stomatal movements. Our results establish a linear signaling cascade in which the Gα protein transmits UV-B signals to elicit H2O2, which then elicits NO in guard cells to regulate UV-B-dependent stomatal closure.  相似文献   

6.
Yariv phenylglycosides [1,3,5-tri(p-glycosyloxyphenylazo)-2,4,6-trihydroxybenzene] are a group of chemical compounds that selectively bind to arabinogalactan proteins (AGPs), a type of plant proteoglycan. Yariv phenylglycosides are widely used as cytochemical reagents to perturb the molecular functions of AGPs as well as for the detection, quantification, purification, and staining of AGPs. However, the target structure in AGPs to which Yariv phenylglycosides bind has not been determined. Here, we identify the structural element of AGPs required for the interaction with Yariv phenylglycosides by stepwise trimming of the arabinogalactan moieties using combinations of specific glycoside hydrolases. Whereas the precipitation with Yariv phenylglycosides (Yariv reactivity) of radish (Raphanus sativus) root AGP was not reduced after enzyme treatment to remove α-l-arabinofuranosyl and β-glucuronosyl residues and β-1,6-galactan side chains, it was completely lost after degradation of the β-1,3-galactan main chains. In addition, Yariv reactivity of gum arabic, a commercial product of acacia (Acacia senegal) AGPs, increased rather than decreased during the repeated degradation of β-1,6-galactan side chains by Smith degradation. Among various oligosaccharides corresponding to partial structures of AGPs, β-1,3-galactooligosaccharides longer than β-1,3-galactoheptaose exhibited significant precipitation with Yariv in a radial diffusion assay on agar. A pull-down assay using oligosaccharides cross linked to hydrazine beads detected an interaction of β-1,3-galactooligosaccharides longer than β-1,3-galactopentaose with Yariv phenylglycoside. To the contrary, no interaction with Yariv was detected for β-1,6-galactooligosaccharides of any length. Therefore, we conclude that Yariv phenylglycosides should be considered specific binding reagents for β-1,3-galactan chains longer than five residues, and seven residues are sufficient for cross linking, leading to precipitation of the Yariv phenylglycosides.Arabinogalactan proteins (AGPs) are a type of plant proteoglycans consisting of a Hyp-rich core protein and large arabinogalactan (AG) moieties (Fincher et al., 1983; Nothnagel, 1997). Although there are many molecular species of AGP differentiated by their core proteins, the AG moieties commonly comprise β-1,3-galactan main chains and β-1,6-galactan side chains, to which l-Ara and other auxiliary sugars, such as GlcA, 4-O-methyl-GlcA, l-Fuc, l-Rha, and Xyl, are attached (Fincher et al., 1983; Nothnagel, 1997; Seifert and Roberts, 2007). A commercial product of AGPs prepared from the acacia (Acacia senegal) tree is known as gum arabic and utilized as a food stabilizer. In the Japanese herbal remedy Juzen-Taiho-To, AGs from Astragalus membranaceus are the active ingredient (Majewska-Sawka and Nothnagel, 2000; Kiyohara et al., 2002). In intact plants, AGPs are implicated in various physiological events and serve as extracellular constituents and signaling molecules. For instance, an AGP from stylar transmitting tissue attracts pollen tubes and stimulates their elongation in tobacco (Nicotiana tabacum; Cheung et al., 1995).Yariv phenylglycosides [1,3,5-tri(p-glycosyloxyphenylazo)-2,4,6-trihydroxybenzene] are a group of chemical compounds that were initially developed as carbohydrate antigens for the purification of anti-glycoside antibody and sugar-binding protein (Yariv et al., 1962, 1967a). It then turned out that Yariv phenylglycosides specifically precipitate AGPs (Yariv et al., 1967b; Jermyn and Yeow, 1975). The specific interaction of AGPs with Yariv phenylglycosides forming brown-red precipitate is called Yariv reactivity and has been recognized as an important criterion in the definition of AGPs, even though a number of AGPs do not exhibit Yariv reactivity. Nevertheless, the structure involved in the interaction with Yariv phenylglycoside is presumed to be conserved in many AGPs. The interaction of Yariv phenylglycosides with AGP depends on the glycosyl residues attached to the phenylazotrihydroxybenzene core. In particular, β-glucosyl Yariv phenylglycoside (β-Glc-Yariv) and β-galactosyl Yariv phenylglycoside (β-Gal-Yariv) bind to AGPs, whereas α-glucosyl Yariv and α-galactosyl Yariv (α-Gal-Yariv) do not bind to AGPs (Jermyn and Yeow, 1975; Larkin, 1977, 1978; Nothnagel and Lyon, 1986). Because of the specific interaction with the β-glycosyl Yariv phenylglycosides (β-Yarivs), AGPs were formerly called “β-lectins” (Jermyn and Yeow, 1975; Gleeson and Jermyn, 1979; Nothnagel and Lyon, 1986).The β-Yarivs are useful tools for staining, detection, and quantification of AGPs. Using β-Glc-Yariv, β-lectins were shown to exist in angiosperm, gymnosperm, fern, moss, and liverwort, illustrating the wide distribution of AGPs in the plant kingdom (Jermyn and Yeow, 1975; Clarke et al., 1978). In addition, β-Yarivs are also used as chemical reagents in the purification of AGPs. A nonclassical AGP, xylogen, which is a signaling molecule inducing the differentiation to tracheary elements, has been purified from the culture medium of zinnia (Zinnia elegans) cells by precipitation with β-Glc-Yariv (Motose et al., 2004). As the treatment with β-Yarivs causes the perturbation of various physiological processes in plants, β-Yarivs are reliable cytochemical reagents to explore AGP functions. Application of β-Yarivs to cultured cells of Arabidopsis (Arabidopsis thaliana) induced programmed cell death, demonstrating the involvement of AGPs in the determination of cell fate (Gao and Showalter, 1999). In tobacco cultured cells, the treatment with β-Yarivs has indicated a possible role of AGPs in the orientation of cortical microtubules and the polymerization of F-actin (Sardar et al., 2006).Although Yariv phenylglycosides have been extensively utilized in studies of AGPs over 40 years, the identification of the target structures on AGPs required for β-Yariv reactivity remains elusive (Nothnagel, 1997; Seifert and Roberts, 2007). It has been proposed that β-Yarivs bind to the Hyp-rich core protein, based on the observation that deglycosylation treatment with hydrogen fluoride did not abolish the Yariv reactivity of gum arabic and a tobacco AGP (Akiyama et al., 1984). To the contrary, other reports have asserted the importance of the carbohydrate moieties for Yariv reactivity (Komalavilas et al., 1991). However, with regard to the specific carbohydrate structure required for interaction with β-Yarivs, the results were not always consistent: neither α-l-arabinofuranosyl residues nor β-1,6-galactan side chains were found to be involved in Yariv reactivity of AGPs from Gladiolus spp., radish (Raphanus sativus), and grape (Vitis vinifera; Gleeson and Clarke, 1979; Tsumuraya et al., 1987; Saulnier et al., 1992); partial acid hydrolysis to remove α-l-arabinofuranosyl residues diminished Yariv reactivity of a rose (Rosa spp.) AGP (Komalavilas et al., 1991); and mugwort (Artemisia vulgaris) pollen O-glycans consisting of a β-1,6-galactan core and branched α-l-arabinofuranosyl side chains precipitated with β-Glc-Yariv (Léonard et al., 2005). Accordingly, it has also been suggested that Yariv reactivity depends on the overall physical and chemical properties rather than a specific structural feature of AGPs.In this study, we demonstrate that the peptide component of AGPs is not required for Yariv reactivity. By sequentially trimming the AG moieties of AGPs with sets of specific glycoside hydrolases, we show that β-Gal-Yariv binds to the β-1,3-galactan main chains of radish root AGP. We confirm that β-1,6-galactan side chains are not necessary for Yariv reactivity, we identify β-1,3-galactopentaose (β-1,3-Gal5) as the smallest carbohydrate structure to interact with β-Gal-Yariv, and we show that β-1,3-galactoheptaose (β-1,3-Gal7) or longer β-1,3-galactosyl chains are required for the formation of insoluble precipitate with Yariv phenylglycoside. Based on computational modeling, a possible interaction mechanism between β-Gal-Yariv and β-1,3-galactan is suggested.  相似文献   

7.
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   

8.
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   

9.
10.
Heterotrimeric G proteins, consisting of Gα, Gβ, and Gγ subunits, are a conserved signal transduction mechanism in eukaryotes. However, G protein subunit numbers in diploid plant genomes are greatly reduced as compared with animals and do not correlate with the diversity of functions and phenotypes in which heterotrimeric G proteins have been implicated. In addition to GPA1, the sole canonical Arabidopsis (Arabidopsis thaliana) Gα subunit, Arabidopsis has three related proteins: the extra-large GTP-binding proteins XLG1, XLG2, and XLG3. We demonstrate that the XLGs can bind Gβγ dimers (AGB1 plus a Gγ subunit: AGG1, AGG2, or AGG3) with differing specificity in yeast (Saccharomyces cerevisiae) three-hybrid assays. Our in silico structural analysis shows that XLG3 aligns closely to the crystal structure of GPA1, and XLG3 also competes with GPA1 for Gβγ binding in yeast. We observed interaction of the XLGs with all three Gβγ dimers at the plasma membrane in planta by bimolecular fluorescence complementation. Bioinformatic and localization studies identified and confirmed nuclear localization signals in XLG2 and XLG3 and a nuclear export signal in XLG3, which may facilitate intracellular shuttling. We found that tunicamycin, salt, and glucose hypersensitivity and increased stomatal density are agb1-specific phenotypes that are not observed in gpa1 mutants but are recapitulated in xlg mutants. Thus, XLG-Gβγ heterotrimers provide additional signaling modalities for tuning plant G protein responses and increase the repertoire of G protein heterotrimer combinations from three to 12. The potential for signal partitioning and competition between the XLGs and GPA1 is a new paradigm for plant-specific cell signaling.The classical heterotrimeric G protein consists of a GDP/GTP-binding Gα subunit with GTPase activity bound to an obligate dimer formed by Gβ and Gγ subunits. In the signaling paradigm largely elucidated from mammalian systems, the plasma membrane-associated heterotrimer contains Gα in its GDP-bound form. Upon receiving a molecular signal, typically transduced by a transmembrane protein (e.g. a G protein-coupled receptor), Gα exchanges GDP for GTP and dissociates from the Gβγ dimer. Both Gα and Gβγ interact with intracellular effectors to initiate downstream signaling cascades. The intrinsic GTPase activity of Gα restores Gα to the GDP-bound form, which binds Gβγ, thereby reconstituting the heterotrimer (McCudden et al., 2005; Oldham and Hamm, 2008).Signal transduction through a heterotrimeric G protein complex is an evolutionarily conserved eukaryotic mechanism common to metazoa and plants, although there are distinct differences in the functional intricacies between the evolutionary branches (Jones et al., 2011a, 2011b; Bradford et al., 2013). The numbers of each subunit encoded within genomes, and therefore the potential for combinatorial complexity within the heterotrimer, is one of the most striking differences between plants and animals. For example, the human genome encodes 23 Gα (encoded by 16 genes), five Gβ, and 12 Gγ subunits (Hurowitz et al., 2000; McCudden et al., 2005; Birnbaumer, 2007). The Arabidopsis (Arabidopsis thaliana) genome, however, only encodes one canonical Gα (GPA1; Ma et al., 1990), one Gβ (AGB1; Weiss et al., 1994), and three Gγ (AGG1, AGG2, and AGG3) subunits (Mason and Botella, 2000, 2001; Chakravorty et al., 2011), while the rice (Oryza sativa) genome encodes one Gα (Ishikawa et al., 1995), one Gβ (Ishikawa et al., 1996), and either four or five Gγ subunits (Kato et al., 2004; Chakravorty et al., 2011; Botella, 2012). As expected, genomes of polyploid plants have more copies due to genome duplication, with the soybean (Glycine max) genome encoding four Gα, four Gβ (Bisht et al., 2011), and 10 Gγ subunits (Choudhury et al., 2011). However, Arabidopsis heterotrimeric G proteins have been implicated in a surprisingly large number of phenotypes, which is seemingly contradictory given the relative scarcity of subunits. Arabidopsis G proteins have been implicated in cell division (Ullah et al., 2001; Chen et al., 2006) and morphological development in various tissues, including hypocotyls (Ullah et al., 2001, 2003), roots (Ullah et al., 2003; Chen et al., 2006; Li et al., 2012), leaves (Lease et al., 2001; Ullah et al., 2001), inflorescences (Ullah et al., 2003), and flowers and siliques (Lease et al., 2001), as well as in pathogen responses (Llorente et al., 2005; Trusov et al., 2006; Cheng et al., 2015), regulation of stomatal movement (Wang et al., 2001; Coursol et al., 2003; Fan et al., 2008) and development (Zhang et al., 2008; Nilson and Assmann, 2010), cell wall composition (Delgado-Cerezo et al., 2012), responses to various light stimuli (Warpeha et al., 2007; Botto et al., 2009), responses to multiple abiotic stimuli (Huang et al., 2006; Pandey et al., 2006; Trusov et al., 2007; Zhang et al., 2008; Colaneri et al., 2014), responses to various hormones during germination (Ullah et al., 2002), and postgermination development (Ullah et al., 2002; Pandey et al., 2006; Trusov et al., 2007). Since the Gγ subunit appeared to be the only subunit that provides diversity in heterotrimer composition in Arabidopsis, it was proposed that all functional specificity in heterotrimeric G protein signaling was provided by the Gγ subunit (Trusov et al., 2007; Chakravorty et al., 2011; Thung et al., 2012, 2013). This allowed for only three heterotrimer combinations to account for the wide range of G protein-associated phenotypes.In addition to the above typical G protein subunits, the plant kingdom contains a conserved protein family of extra-large GTP-binding proteins (XLGs). XLGs differ from typical Gα subunits in that they possess a long N-terminal extension of unknown function, but they are similar in that they all have a typical C-terminal Gα-like region, with five semiconserved G-box (G1–G5) motifs. The XLGs also possess the two sequence features that differentiate heterotrimeric G protein Gα subunits from monomeric G proteins: a helical region between the G1 and G2 motifs and an Asp/Glu-rich loop between the G3 and G4 motifs (Lee and Assmann, 1999; Ding et al., 2008; Heo et al., 2012). The Arabidopsis XLG family comprises XLG1, XLG2, and XLG3, and all three have demonstrated GTP-binding and GTPase activities, although they differ from GPA1 in exhibiting a much slower rate of GTP hydrolysis, with a Ca2+ cofactor requirement instead of an Mg2+ requirement, as for canonical Gα proteins (Heo et al., 2012). All three Arabidopsis XLGs were observed to be nuclear localized (Ding et al., 2008). Although much less is known about XLGs than canonical Gα subunits, XLG2 positively regulates resistance to the bacterial pathogen Pseudomonas syringae and was immunoprecipitated with AGB1 from tissue infected with P. syringae (Zhu et al., 2009). xlg3 mutants, like agb1 mutants, are impaired in root-waving and root-skewing responses (Pandey et al., 2008). During the preparation of this report, Maruta et al. (2015) further investigated XLG2, particularly focusing on the link between XLG2 and Gβγ in pathogen responses. Based on symptom progression in xlg mutants, they found that XLG2 is a positive regulator of resistance to both bacterial and fungal pathogens, with a minor contribution from XLG3 in resistance to Fusarium oxysporum. XLG2 and XLG3 are also positive regulators of reactive oxygen species (ROS) production in response to pathogen-associated molecular pattern elicitors. The resistance and pathogen-associated molecular pattern-induced ROS phenotypes of the agg1 agg2 and xlg2 xlg3 double mutants were not additive in an agg1 agg2 xlg2 xlg3 quadruple mutant, indicating that these two XLGs and the two Gγ subunits function in the same, rather than parallel, pathways. Unfortunately, the close proximity of XLG2 and AGB1 on chromosome 4 precluded the generation of an agb1 xlg2 double mutant; therefore, direct genetic evidence of XLG2 and AGB1 interaction is still lacking, but physical interactions between XLG2 and the Gβγ dimers were shown by yeast (Saccharomyces cerevisiae) three-hybrid and bimolecular fluorescence complementation (BiFC) assays (Maruta et al., 2015). Localization of all three XLGs was also reexamined, indicating that XLGs are capable of localizing to the plasma membrane in addition to the nucleus (Maruta et al., 2015).Interestingly, several other plant G protein-related phenotypes, in addition to pathogen resistance, have been observed only in Gβ and Gγ mutants, with opposite phenotypes observed in Gα (gpa1) mutants. Traditionally, the observation of opposite phenotypes in Gα versus Gβγ mutants in plants and other organisms has mechanistically been attributed to signaling mediated by free Gβγ, which increases in abundance in the absence of Gα. However, an intriguing alternative is that XLG proteins fulfill a Gα-like role in forming heterotrimeric complexes with Gβγ and function in non-GPA1-based G protein signaling processes. If XLGs function like Gα subunits, the corresponding increase in subunit diversity could potentially account for the diversity of G protein phenotypes. In light of this possibility, we assessed the heterotrimerization potential of all possible XLG and Gβγ dimer combinations, XLG localization and its regulation by Gβγ, and the effect of xlg mutation on selected known phenotypes associated with heterotrimeric G proteins. Our results provide compelling evidence for the formation of XLG-Gβγ heterotrimers and reveal that plant G protein signaling is substantially more complex than previously thought.  相似文献   

11.
12.
13.
14.
15.
16.
17.
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   

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

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