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.
Tumor-causing γ-herpesviruses have evolved elaborate mechanisms to deal with almost every aspect of host cell defense. In this issue of Cell Host & Microbe, Leidal et?al. (2012) report an oncogenic synergy between the latent KSHV proteins v-FLIP and v-cyclin during KSHV persistent infection that reshapes autophagy.  相似文献   

2.
Intracellular bacterial pathogens engage in a tug-of-war with innate host defenses. In this issue of Cell Host & Microbe, Mostowy et?al. (2010) identify a role for the septin family of cytoskeletal proteins in targeting intracellular Shigella to the autophagy pathway.  相似文献   

3.
Autophagy is now recognized as a cellular defense mechanism that can restrict the growth of Mycobacterium tuberculosis (Mtb). In this issue of Cell Host & Microbe, Kim et?al. (2012) demonstrate that antibiotics routinely used to treat Mtb infection elicit a host autophagy response critical for bacterial clearance.  相似文献   

4.
EMBO J (2013) 32: 2685–2696 doi:10.1038/emboj.2013.189; published online August232013Beclin 1 is a crucial regulator of autophagy. It forms a complex with ATG14L, VPS34 or the class III phosphatidylinositol 3-kinase AMBRA1 to control autophagosome formation. A study in this issue of The EMBO Journal (Xia et al, 2013) reports that WASH, a protein known to regulate endosomal sorting, hampers autophagy by inhibiting the AMBRA1-dependent polyubiquitylation of Beclin 1. This modification is required to promote VPS34 activity and to initiate autophagy during starvation.Macroautophagy (hereafter called autophagy) is a lysosomal degradation pathway for cytoplasmic components. Autophagy is initiated by the formation of a double-membrane bound vacuole, the autophagosome that sequesters the autophagic cargo and eventually fuses with the lysosomal compartment, leading to cargo degradation. Autophagy plays a major role in cytoplasm homeostasis by removing protein aggregates and controlling organelle quality, and it is stimulated to promote cell survival during stressful situations, such as nutrient starvation and microbial infection. Autophagosome formation is dependent on evolutionarily conserved Atg (Autophagy-related) proteins initially identified in yeast (Mizushima et al, 2011). They function in complexes or functional modules on a membrane known as the phagophore that matures into the autophagosome via several stages (initiation, elongation and sealing). Phosphatidylinositol 3-phosphate kinase complex I (PI3K complex I) plays a key role in the initiation step. In this complex, Beclin 1 (the mammalian homologue of yeast Atg6) interacts with ATG14L, AMBRA1, and class III PI3K or VPS34 (Cecconi and Levine, 2008; Figure 1). The activity of this complex, which produces the lipid phosphatidylinositol 3-phosphate (PI3P) to recruit Atg18 homologues (WIPIs), has to be tightly regulated to keep autophagy under control.Open in a separate windowFigure 1Role of WASH in autophagosome formation. WASH interacts with Beclin 1 via a sequence (121–221) that is not involved in endosomal sorting. In the absence of WASH, when autophagy is stimulated by nutrient deprivation, Beclin 1 interacts with VPS34 (and its adaptor VPS15), ATG14L and AMBRA1. In this complex, Beclin 1 is polyubiquitylated by AMBRA1, which acts as an E3 ligase. The activation of VPS34 produces PI3P at the phagophore to recruit WIPI proteins, and then triggers the machinery to elongate and seal the membrane, thus forming an autophagosome. The VAC domain of WASH involved in endosomal sorting is not required for autophagy regulation.In this issue of The EMBO Journal, Xia et al (2013) demonstrate that WASH (Wiskott-Aldrich syndrome protein (WASP) and SCAR homologue) regulates autophagosome formation in response to nutrient starvation by influencing the ubiquitylation of Beclin 1. WASH forms a complex with four other proteins and plays an essential role in endosomal sorting by promoting actin polymerization to facilitate segregation of endosomal proteins (Seaman et al, 2013). Using WASH−/− mice, Xia et al (2013) observe that WASH deficiency causes embryonic lethality (E7.5−E9.5) with massive apoptosis-independent cell death and the accumulation of autophagic structures. However, whether autophagy is instrumental in the cell death observed in WASH−/− embryos remains to be investigated. From a series of classical readouts, the authors conclude that WASH downregulates starvation-induced autophagy. Interestingly, WASH is detected on the autophagosomal membrane before and after closure of the autophagosome, but not on the autolysosomes (organelles formed when the autophagosome fuses with the lysosome). The authors first show that WASH has distinct roles in autophagy and in endosomal sorting, because deletion of the VCA domain of WASH that is required for its endosomal function and the knockdown of FAM21 (a protein that directs WASH to the endosomal membrane) do not influence autophagy. They then show that WASH interacts with the coil-coiled domain of Beclin 1. WASH depletion induces more VPS34 to interact with Beclin 1, leading to augmented formation of PtdIns3P and increased recruitment of WIPI-1 to the autophagosomal membrane. However, WASH does not influence the stability of Beclin 1, although it does regulate its K63-linked polyubiquitylation at position K437 during starvation (K63-linked polyubiquitylation suggests a regulatory role, whereas K48-linked polyubiquitylation is frequently associated with protein degradation). Overexpression of WASH reduces the ubiquitylation of Beclin 1, and the K437R Beclin 1 mutant that cannot be ubiquitylated has a low level of interaction with VPS34, which prevents the stimulation of autophagy by starvation (Figure 1). Recently, Beclin 1 has been shown to be ubiquitylated at position K117 during TLR4-induced autophagy (Shi and Kehrl, 2010). Interestingly, Xia et al (2013) show that the K117R Beclin 1 mutant is still ubiquitylated in response to nutrient starvation. Finally, the authors identify the substrate-specific E3 ubiquitin ligase for Beclin 1 in this context. They exclude TRAF6 and NEDD4, two E3 ligases reported to ubiquitylate Beclin 1 (Kuang et al, 2013). AMBRA1, which regulates autophagy in the PI3K complex I, is also known as DCAF3 and interacts with the Cullin 4–DDB1 E3 ligase complex (Jin et al, 2006). Following in vitro E3 ligase assays, Xia et al (2013) conclude that AMBRA1 is the substrate receptor for Beclin 1 ubiquitylation at position K437, adding Cullin4/DDB1 and AMBRA1 to the list of E3 ligases involved in autophagy (Kuang et al, 2013). The finding that WASH and AMBRA1 compete to regulate starvation-induced autophagy raises several questions: how do WASH and AMBRA1 influence one another''s binding to Beclin 1? WASH is evolutionarily conserved, although it is not present in yeast (Linardopoulou et al, 2007)—does it also reduce starvation-induced autophagy in non-mammalian cells? WASH has recently been reported to be required for the lysosomal digestion of autophagy cargo in Dictyostelium during starvation (King et al, 2013). Does this mean that the function of WASH in autophagy has evolved from controlling cargo degradation to forming autophagosomes? Xia et al (2013) do not report any association between WASH and the lysosomal compartment during autophagy; the role of WASH in autophagosome maturation calls for further investigation. Finally, a recent report demonstrates that AMBRA1 interacts with the E3 ligase TRAF6 to ubiquitylate ULK1 (the mammalian orthologue of the yeast Atg1) and to ensure its subsequent stabilization and function (Nazio et al, 2013). ULK1 is part of a complex that acts with PI3K complex 1 to regulate the initiation of autophagy (Mizushima et al, 2011). How are these two AMBRA1-dependent ubiquitylation processes coordinated to control autophagy? In conclusion, both the study reported here (Xia et al, 2013) and the recent study of Nazio et al (2013) clearly demonstrate that AMBRA1 plays a central role in regulating the initiation of autophagy.  相似文献   

5.
Autophagy is generally considered to be a cytoprotective response to stress, whether in the form of nutrient deprivation or the presence of dysfunctional organelles. He et?al. now show in Nature that exercise-induced autophagy is needed for some of the beneficial effects of exercise on metabolism (He et?al., 2012).  相似文献   

6.
Highlights? Antimycobacterial antibiotics activate autophagy in Mtb-infected host cells ? Autophagy activation depends on cellular and mitochondrial reactive oxygen species ? Host cell autophagy is essential for antimycobacterial drug action in infected macrophages and flies ? Antibiotic-induced autophagy dampened proinflammatory responses in infected macrophages  相似文献   

7.
Autophagy is a catabolic membrane-trafficking process whereby cells recycle cytosolic proteins and organelles under stress conditions or during development. This degradative process is mediated by autophagy-related (ATG) proteins that have been described in yeast, animals, and more recently in plants. In this study, we report the molecular characterization of autophagy in the unicellular green alga Chlamydomonas reinhardtii. We demonstrate that the ATG8 protein from Chlamydomonas (CrATG8) is functionally conserved and may be used as a molecular autophagy marker. Like yeast ATG8, CrATG8 is cleaved at the carboxyl-terminal conserved glycine and is associated with membranes in Chlamydomonas. Cell aging or different stresses such as nutrient limitation, oxidative stress, or the accumulation of misfolded proteins in the endoplasmic reticulum caused an increase in CrATG8 abundance as well as the detection of modified forms of this protein, both landmarks of autophagy activation. Furthermore, rapamycin-mediated inhibition of the Target of Rapamycin signaling pathway, a major regulator of autophagy in eukaryotes, results in identical effects on CrATG8 and a relocalization of this protein in Chlamydomonas cells similar to the one observed upon nutrient limitation. Thus, our findings indicate that Chlamydomonas cells may respond to stress conditions by inducing autophagy via Target of Rapamycin signaling modulation.Protein turnover is essential for the adaptation of cells to variable environmental conditions. Similar to other eukaryotes, plants have developed two distinct mechanisms to regulate protein degradation, a selective ubiquitin/26S proteasome pathway (Vierstra, 2009) and macroautophagy (hereafter referred to as autophagy), a nonselective membrane-trafficking process (Bassham, 2009). During autophagy, a large number of cytosolic components, including entire organelles, organelle fragments, and protein complexes, are enclosed in bulk within a double-membrane structure known as the autophagosome and delivered to the vacuole/lysosome for degradation to recycle needed nutrients or degrade toxic components (Xie and Klionsky, 2007; Nakatogawa et al., 2009). The autophagosomes appear to arise from isolation membranes usually observed in close proximity to the vacuole called the preautophagosomal structure (PAS). These membranes expand and fuse to encircle portions of the cytoplasm, generating an autophagosome that is targeted to the vacuole. The outer membrane of the autophagosome then fuses with the vacuole membrane, and the remaining vesicle, known as the autophagic body, is finally released to the vacuole for its degradation (Xie and Klionsky, 2007).The evolutionary conservation of autophagy among eukaryotes indicates that structural and regulatory components of this cellular process must be also conserved. Accordingly, a significant number of autophagy-related (ATG) genes that participate in autophagy regulation and autophagosome formation have been identified, initially through genetic approaches in yeast and subsequently in higher eukaryotes, including mammals, insects, protozoa, and plants (Bassham et al., 2006; Bassham, 2007; Meijer et al., 2007). In yeast, two protein conjugation systems composed of the ubiquitin-like proteins ATG8 and ATG12 and the three enzymes ATG3, ATG7, and ATG10 play an essential role in autophagosome formation and seem to be conserved through evolution (Geng and Klionsky, 2008). ATG8 becomes modified with the lipid molecule phosphatidylethanolamine (PE) by the consecutive action of the ATG7 and ATG3 enzymes in a process mechanistically similar to ubiquitination (Ichimura et al., 2000). Prior to this modification, ATG8 must be cleaved by the Cys protease ATG4 to expose a C-terminal Gly residue that is conjugated to PE (Kirisako et al., 2000; Kim et al., 2001). ATG12 becomes covalently attached to the ATG5 protein in a conjugation reaction that is catalyzed by ATG7 and ATG10 (Mizushima et al., 1998). ATG8-PE and ATG12-ATG5 conjugates localize to autophagy-related membranes and are required for the initiation and expansion of autophagosomal membrane and hemifusion of this membrane with the vacuolar membrane (Hanada et al., 2007; Nakatogawa et al., 2007, 2009; Fujita et al., 2008; Geng and Klionsky, 2008; Xie et al., 2008).Our understanding of the autophagy process has substantially increased with the development of specific markers for autophagy. In plants, two markers for autophagosomes have been described, the monodansylcadaverine dye and GFP-ATG8 fusion protein (Yoshimoto et al., 2004; Contento et al., 2005; Thompson et al., 2005). As in other species, binding of ATG8 to autophagosomes has been used to monitor autophagy in plants. In contrast to yeast, where a single ATG8 gene is present, plants appear to contain a small gene family with several ATG8 isoforms, suggesting that autophagy is more complex in these photosynthetic organisms. For example, Arabidopsis (Arabidopsis thaliana) and maize (Zea mays) encode nine and five ATG8 genes, respectively (Doelling et al., 2002; Hanaoka et al., 2002; Ketelaar et al., 2004; Chung et al., 2009). However, despite the high complexity of the ATG8-conjugating system in plants, important findings have been recently reported on the molecular characterization of autophagy using ATG8 as an autophagy marker in these organisms. The use of specific markers for autophagy in plants has revealed that this process is active at a basal level under normal growth and is induced upon nitrogen- or carbon-limiting conditions as well as in response to oxidative stress (Yoshimoto et al., 2004; Thompson et al., 2005; Xiong et al., 2005, 2007; Chung et al., 2009). Reverse genetic approaches have also been applied to a number of Arabidopsis ATG genes using T-DNA insertional mutants or RNA interference in order to investigate the physiological roles of autophagy in plants. The initial characterization of autophagy-deficient plants demonstrated that the ATG system is not essential under nutrient-rich conditions. However, a detailed analysis of these mutants indicated that autophagy is required for the proper response of the plant to nutrient limitation or pathogen infection. Plants lacking the ATG4, ATG5, ATG7, ATG9, or ATG10 gene display premature leaf senescence and are hypersensitive to nitrogen or carbon limitation (Doelling et al., 2002; Hanaoka et al., 2002; Yoshimoto et al., 2004; Thompson et al., 2005; Phillips et al., 2008). Arabidopsis plants with reduced levels of ATG18, which is required for autophagosome formation, are more sensitive to methyl viologen treatment and accumulate high levels of oxidized proteins, demonstrating that autophagic processes participate in the response of the plant to oxidative stress (Xiong et al., 2005, 2007). Plants deficient in the autophagy genes ATG6/Beclin1, ATG3, ATG7, and ATG9 exhibit unrestricted hypersensitive response lesions in response to pathogen infection (Liu et al., 2005; Hofius et al., 2009). These findings implicate autophagy as a prosurvival mechanism to restrict programmed cell death associated with the pathogen-induced hypersensitive response in plants. Arabidopsis ATG6 has also been shown to mediate pollen germination in a manner independent of autophagy (Fujiki et al., 2007).As mentioned above, autophagy is triggered among other factors by a reduction in the availability of nutrients. This starvation signal is transmitted to the autophagic machinery by important regulatory factors, including the Ser/Thr kinases Target of Rapamycin (TOR), ATG1, and SNF1 and the phosphatidylinositol 3-kinase ATG6/Beclin1 (Diaz-Troya et al., 2008b; Bassham, 2009; Cebollero and Reggiori, 2009). TOR has been identified as a negative regulator of autophagy in yeast, mammals, and fruit flies (Diaz-Troya et al., 2008b). The pharmacological inhibition of TOR by rapamycin leads to autophagy activation through a mechanism that requires the activation of the ATG1 kinase (Kamada et al., 2000). It has been recently demonstrated in mammals and fruit flies that a rapamycin-sensitive TOR signaling complex termed TORC1 directly phosphorylates and inhibits the ATG1 kinase and its regulatory protein ATG13 (Chang and Neufeld, 2009; Hosokawa et al., 2009; Jung et al., 2009). These regulatory proteins are conserved in plants, although except for ATG6 (Liu et al., 2005), there is no direct evidence for regulation of autophagy by these signaling pathways.The unicellular green alga Chlamydomonas reinhardtii has been used as a model for the study of important cellular and metabolic processes in photosynthetic organisms (Harris, 2001). More recently, Chlamydomonas has also been proposed as a useful system for the characterization of the TOR signaling pathway in photosynthetic eukaryotes based on the finding that, unlike plants, Chlamydomonas cell growth is sensitive to rapamycin (Crespo et al., 2005; Diaz-Troya et al., 2008a). Treatment of Chlamydomonas cells with rapamycin results in a pronounced increase of vacuole size that resembles autophagy-like processes (Crespo et al., 2005). However, a role of TOR in autophagy regulation could not be demonstrated due to the absence of an autophagy marker in Chlamydomonas. Actually, no studies have been reported on any autophagy-related protein in green algae, despite the high conservation of ATG genes in Chlamydomonas (Diaz-Troya et al., 2008b).This study reports the molecular and cellular characterization of autophagy in the green alga Chlamydomonas. We demonstrate that the ATG8 protein from Chlamydomonas (CrATG8) may be used as a specific autophagy marker. Nutrient limitation and cell aging trigger an autophagic response that can be traced as an increase at the level of CrATG8, the detection of modified forms of CrATG8, and a change in the cellular localization of this protein. Furthermore, we demonstrate that autophagy is inhibited by a rapamycin-sensitive TOR cascade in Chlamydomonas.  相似文献   

8.
9.
EMBO J 31 13, 2852–2868 (2012); published online May292012Together with the proteasome, autophagy is one of the major catabolic pathways of the cell. In response to cellular needs or environmental cues, this transport route targets specific structures for degradation into the mammalian lysosomes or the yeast and plant vacuoles. The mechanisms allowing exclusive autophagic elimination of unwanted structures are currently the object of intensive investigations. The emerging picture is that there is a series of autophagy receptors that determines the specificity of the different selective types of autophagy. How cargo binding and recognition is regulated by these receptors, however, is largely unknown. In their study, Motley et al (2012) have shed light into the molecular principles underlying the turnover of excess peroxisomes in the budding yeast Saccharomyces cerevisiae.Peroxisomes perform a series of crucial functions and their number is regulated in response to the metabolic demands of the cell. After proliferation and when no more required, a selective type of autophagy called pexophagy degrades superfluous peroxisomes (Manjithaya et al, 2010). This turnover allows the cell to save the energy required for the maintenance of excess organelles and to generate metabolites that can be used to carry out other functions. Like all selective types of autophagy, pexophagy relies on the conserved core of the autophagy-related (Atg) machinery, but also requires additional proteins that confer specificity of the pathway such as cargo selection and membrane dynamics (Manjithaya et al, 2010). It is still unclear, however, which peroxisomal protein allows the recognition of peroxisomes by the autophagosomes. Although Pex3 and Pex14 have previously been indicated as possible suspects (Bellu et al, 2001, 2002; Farre et al, 2008), their specific contribution to pexophagy was difficult to establish. Deletion of either PEX3 or PEX14, as well as most other PEX genes, leads to defects in peroxisome biogenesis, which makes the dissection of their contribution to peroxisome degradation very difficult to assess. Motley et al (2012) have elegantly exploited S. cerevisiae genetics to isolate pex3 alleles specifically impaired in pexophagy and could thus demonstrate that Pex3 (and not Pex14) mediates the selective engulfment of peroxisomes by autophagosomes. In support to this result, the authors have also identified a new protein, Atg36, which binds Pex3 (Figure 1). Importantly, Atg36 interacts with Atg11, an autophagy adaptor involved in numerous selective types of autophagy in yeast, thereby bringing peroxisomes to the site where autophagosomes will be generated and coordinating the activation of the Atg machinery at this location (Kim et al, 2001; Reggiori et al, 2005; Monastyrska et al, 2008). Atg36, however, is only present in S. cerevisiae and related yeasts. Methylotrophic yeasts, in contrast, appear to have a different protein with the same properties, Atg30 (Farre et al, 2008). It is unclear, however, whether Atg30 is the functional counterpart of Atg36 because these two proteins do not display similarities in their amino-acid sequence.Open in a separate windowFigure 1Schematic representation for a putative Pex3 checkpoint. The peroxisomal integral membrane protein Pex3 acts as a master regulator to determine peroxisome fate. Organelle abundance is regulated by formation of new organelles, and their subsequent segregation (inheritance) and degradation. A new paradigm has been uncovered, whereby Pex3 controls peroxisome abundance through the regulated binding to specific co-factors. At the endoplasmic reticulum (ER), together with Pex19, it initiates biogenesis of new peroxisomes. At the peroxisomal membrane, it ensures that both mother and daughter cells obtain the correct number of peroxisomes, whereas when the organelles become dispensable, Pex3 can initiate their selective degradation. To keep peroxisomes in the mother cell during cell division, Pex3 associates with Inp1 and tether peroxisomes to cortical actin patches. Under pexophagy-inducing conditions, Pex3 binds the newly identified pexophagy factor Atg36 and delivers peroxisomes to the site of autophagosome formation for subsequent degradation into the vacuole.While it is unmistakable that Atg36 (and Atg30) is essential for pexophagy, it remains unclear whether this protein is an autophagy receptor. This class of molecules has four characteristics (Kraft et al, 2010). First, each autophagy receptor binds a specific cargo. Second, they often interact with adaptor proteins, which function as scaffolds that bring the cargo–receptor complex in contact with the core Atg machinery to allow the specific sequestration of the cargo. Third, they possess at least one LC3-interacting region (LIR) motif that enables them to interact with the LC3/Atg8 pool present in the interior autophagomes and assures the hermetic enwrapping of the cargo into these vesicles. Fourth, autophagy receptors are degraded in the lysosome/vacuole together with the cargo that they bind to. While Atg36 (and Atg30) binds both the cargo and the adaptor protein Atg11, this protein does not appear to be turned over in the vacuole during pexophagy and a LIR motif has not been pinpointed yet. Consequently, it is unclear whether Atg36 is a new type of autophagy receptor or acts together with a not yet identified autophagy receptor involved in pexophagy.A very interesting concept emerging from the work of Motley et al (2012) is that a single protein, that is, Pex3, could be the central regulator of peroxisome homoeostasis (Figure 1). Pex3 is involved in peroxisome biogenesis, segregation and degradation (Bellu et al, 2002; Hoepfner et al, 2005; Farre et al, 2008; Munck et al, 2009; Ma et al, 2011). As a result, the cell could regulate peroxisome abundance by modulating Pex3 function and/or its array of interactions. In this context, it would be particularly interesting to determine whether Pex3 is also the decision maker of a quality control mechanism that eliminates peroxisomes when not correctly assembled and thus dysfunctional, or when not accurately distributed during cell division. Clearly, additional experiments are needed to understand how Pex3 regulates peroxisome homoeostasis, but this protein and this organelle could represent a convenient system to unveil the principles that regulate the steady-state level of other subcellular compartments.  相似文献   

10.
11.
Survival of blood stage malaria parasites requires extensive host cell remodeling, which is facilitated by secretion of parasite proteins via a dedicated protein export pathway. In a recent Cell paper, Bhattacharjee et?al., (2012) describe PI(3)P binding as one of the first steps in targeting parasite proteins to the host cell.  相似文献   

12.
An emerging paradigm in tumor metabolism is that catabolism in host cells "fuels" the anabolic growth of cancer cells via energy transfer. A study in Nature Medicine (Nieman et al., 2011) supports this; they show that triglyceride catabolism in adipocytes drives ovarian cancer metastasis by providing fatty acids as mitochondrial fuels.  相似文献   

13.
14.
15.
Autophagy has been shown to contribute to defense against intracellular bacteria and parasites. In comparison, the ability of such pathogens to manipulate host cell autophagy to their advantage has not been examined. Here we present evidence that infection by Toxoplasma gondii, an intracellular protozoan parasite, induces host cell autophagy in both HeLa cells and primary fibroblasts, via a mechanism dependent on host Atg5 but independent of host mammalian target of rapamycin suppression. Infection led to the conversion of LC3 to the autophagosome-associated form LC3-II, to the accumulation of LC3-containing vesicles near the parasitophorous vacuole, and to the relocalization toward the vacuole of structures labeled by the phosphatidylinositol 3-phosphate indicator YFP-2×FYVE. The autophagy regulator beclin 1 was concentrated in the vicinity of the parasitophorous vacuole in infected cells. Inhibitor studies indicated that parasite-induced autophagy is dependent on calcium signaling and on abscisic acid. At physiologically relevant amino acid levels, parasite growth became defective in Atg5-deficient cells, indicating a role for host cell autophagy in parasite recovery of host cell nutrients. A flow cytometric analysis of cell size as a function of parasite content revealed that autophagy-dependent parasite growth correlates with autophagy-dependent consumption of host cell mass that is dependent on parasite progression. These findings indicate a new role for autophagy as a pathway by which parasites may effectively compete with the host cell for limiting anabolic resources.Macroautophagy (hereafter referred to as autophagy) is a major catabolic process in which cytosolic constituents are sequestered within double-membraned vesicles (autophagosomes) and subsequently delivered to lysosomes for degradation. Current evidence indicates at least two distinct functions for this process. On the one hand, autophagy can be up-regulated under nutrient-limiting conditions to increase nutrient supply via recycling of the products of autophagic degradation, which may be exported from the lysosome (1). The up-regulation of autophagy upon starvation is thought to be mediated by the suppression of signaling through the mTOR pathway (2). On the other hand, autophagy can serve to maintain cellular homeostasis by facilitating the removal of damaged or deleterious elements, such as misfolded protein aggregates (3). An important example of the latter function is the role of autophagy in restricting the growth of intracellular pathogens, including both free bacteria that have escaped into host cytosol, such as group A Streptococcus, and pathogens, such as Mycobacterium tuberculosis, that reside in parasitophorous vacuoles in macrophages (4, 5). In macrophages infected with Toxoplasma gondii, fusion of the parasitophorous vacuole with lysosomes can be induced in an autophagy-dependent manner when host cell anti-parasitic function is activated via CD40 (6). Autophagy as a component of host defense may be up-regulated by inflammatory agents such as lipopolysaccharide (7) and interferon-γ (8).Although the clearance function of autophagy may enhance pathogen killing in host cells that have been activated to generate antimicrobial or antiparasitic function, in permissive host cells, in which the pathogen is less susceptible to sequestration by the autophagosome, autophagy may conceivably play a quite different role. Modulation of the balance between anabolic and catabolic processes may affect the outcome of competition between pathogen and host cell for limiting nutrients. In particular, the nutritive function of autophagy could favor pathogen expansion by providing greater access to host cell biomass. The intracellular apicomplexan parasite, T. gondii, is a suitable agent for the investigation of this hypothesis, because it has been shown to be highly dependent on its host cell for the supply of several nutrients, including amino acids (9), lipids (10), and purines (11). T. gondii replicates within a parasitophorous vacuole that, in permissive host cells, is protected from lysosomal fusion. Recent evidence indicates that in such permissive cells, in which the parasite can differentiate into bradyzoites associated with chronic infection, the pathogen is able to actively sequester host cell lysosome-derived vesicles, thereby potentially gaining access to their contents (12).The ability of intracellular parasites to regulate host cell autophagy has been little examined, and there is also little information with respect to the impact of these pathogens on host cell signals that potentially affect the autophagic pathway. In addition to mTOR, these include calcium ions, which have been implicated in autophagy induced by endoplasmic reticulum stress (13). In this study, we provide evidence that T. gondii induces host cell autophagy by a mechanism dependent on calcium but independent of mTOR and that it exploits the nutritive function of host autophagy to enhance its proliferation.  相似文献   

16.
During cell migration, chemoattractant-induced signaling pathways determine the direction of movement by controlling the spatiotemporal dynamics of cytoskeletal components. In this issue of Developmental Cell, Liu et?al. report that the target of rapamycin complex 2 (TORC2) controls cell polarity and chemotaxis through regulation of both F-actin and myosin II in migrating neutrophils.  相似文献   

17.
Growth factors and oncogenic kinases play important roles in stimulating cell growth during development and transformation. These processes have significant energetic and synthetic requirements and it is apparent that a central function of growth signals is to promote glucose metabolism to support these demands. Because metabolic pathways represent a fundamental aspect of cell proliferation and survival, there is considerable interest in targeting metabolism as a means to eliminate cancer. A challenge, however, is that molecular links between metabolic stress and cell death are poorly understood. Here we review current literature on how cells cope with metabolic stress and how autophagy, apoptosis, and necrosis are tightly linked to cell metabolism. Ultimately, understanding of the interplay between nutrients, autophagy, and cell death will be a key component in development of new treatment strategies to exploit the altered metabolism of cancer cells.Although single-celled organisms grow and proliferate based on nutrient availability, metazoan cells rely on growth factor input to promote nutrient uptake, regulate growth and proliferation, and survive (Raff 1992; Rathmell et al. 2000). Access and competition for these signals are critical in developmental patterning and to maintain homeostasis of mature tissues. Cells that do not receive proper growth factor signals typically atrophy, lose the ability to uptake and use extracellular nutrients, and instead induce the self-digestive process of autophagy as an intracellular energy source before ultimately undergoing programmed cell death. Cancer cells, in contrast, often become independent of extracellular growth signals by gaining mutations or expressing oncogenic kinases to drive intrinsic growth signals that mimic growth factor input, which can be the source of oncogene addiction. Growth factor input or oncogenic signals often drive highly elevated glucose uptake and metabolism (Rathmell et al. 2000; DeBerardinis et al. 2008; Michalek and Rathmell 2010). First described in cancer by Warburg in the 1920s, this highly glycolytic metabolic program is termed aerobic glycolysis and is a general feature of many nontransformed proliferative cells (Warburg 1956; DeBerardinis et al. 2008).Nutrient uptake and aerobic glycolysis induced by growth signals play key roles in cell survival (Vander Heiden et al. 2001). Manipulating cell metabolism as a means to promote the death of inappropriately dividing cells, therefore, is a promising new avenue to treat disease. Targeting the altered metabolism of cancer cells in particular is of great interest. It is still unclear at the molecular level, however, how inhibiting or modulating cell metabolism leads to apoptosis, and how these pathways may best be exploited (Dang et al. 2009; Wise and Thompson 2010).Growth factor or oncogenic kinases promote multiple metabolic pathways that are essential to prevent metabolic stress and may be targets in efforts to link metabolism and cell death (Vander Heiden et al. 2001). Decreased glucose metabolism on loss of growth signals leads to decreased ATP generation as well as loss in generation of many biosynthetic precursor molecules, including nucleic acids, fatty acids, and acetyl-CoA for acetylation (Zhao et al. 2007; Wellen et al. 2009; Coloff et al. 2011). Glucose is also important as a precursor for the hexosamine pathway, to allow proper glycosylation and protein folding in the endoplasmic reticulum (Dennis et al. 2009; Kaufman et al. 2010). If glucose metabolism remains insufficient or disrupted, the cells can switch to rely on mitochondrial oxidation of fatty acids and amino acids, which are energy rich but do not readily support cell growth and can lead to potentially dangerous levels of reactive oxygen species (Wellen and Thompson 2010). Amino acid deficiency can directly inhibit components of the signaling pathways downstream from growth factors and activate autophagy (Lynch 2001; Beugnet et al. 2003; Byfield et al. 2005; Nobukuni et al. 2005). Finally, hypoxia induces a specific pathway to increase nutrient uptake and metabolism via the hypoxia-inducible factor (HIF1/2α) that promotes adaptation to anaerobic conditions, but may lead to apoptosis if hypoxia is severe (Saikumar et al. 1998; Suzuki et al. 2001; Fulda and Debatin 2007).Typically a combination of metabolic stresses rather than loss of a single nutrient input occur at a given time (Degenhardt et al. 2006) and autophagy is activated to mitigate damage and provide nutrients for short-term survival (Bernales et al. 2006; Tracy et al. 2007; Altman et al. 2011; Guo et al. 2011). Autophagy is a cellular process of bulk cytoplasmic and organelle degradation common to nearly all eukaryotes. Unique double-membraned vesicles known as autophagosomes engulf cellular material and fuse with lysosomes to promote degradation of the contents (Kelekar 2005). Described in greater detail below, autophagy can reduce sources of stress, such as protein aggregates and damaged or dysfunctional intracellular organelles, and provide nutrients during times of transient and acute nutrient withdrawal.Despite the protective effects of autophagy, cells deprived of growth signals, nutrients, or oxygen for prolonged times will eventually succumb to cell death. Apoptosis is the initial death response on metabolic stress and is regulated by Bcl-2 family proteins. In healthy cells, antiapoptotic Bcl-2 family proteins, such as Bcl-2, Bcl-xl, and Mcl-1, bind and inhibit the multidomain proapoptotic proteins Bax and Bak (van Delft and Huang 2006; Walensky 2006; Chipuk et al. 2010). In metabolic stress, proapoptotic “BH3-only” proteins of the Bcl-2 family are induced or activated and bind to and inhibit the antiapoptotic Bcl-2 family proteins to allow activation of the proapoptotic Bax and Bak (Galonek and Hardwick 2006). The BH3-only proteins Bim, Bid, and Puma can also directly bind and activate Bax and Bak (Letai et al. 2002; Ren et al. 2010). Active Bax and Bak disrupt the outer mitochondrial membrane (termed mitochondrial outer-membrane permeabilization, or MOMP) and release several proapoptotic factors including cytochrome-C that activate the apoptosome that in turn activates effector caspases to cleave a variety of cellular proteins and drive apoptosis (Schafer and Kornbluth 2006). In cases in which these apoptotic pathways are suppressed, metabolic stress can instead lead to necrotic cell death (Jin et al. 2007).  相似文献   

18.
自噬是一个通过降解细胞组分如细胞器和蛋白质等以维持细胞存活和功能的重要的溶酶体途径。肝脏作为新陈代谢的中枢器官,肝脏高度依赖于自噬以发挥正常功能并防止疾病发展。肝细胞自噬的改变参与肝损伤,脂肪肝等肝病的病理变化,以自噬为靶点寻求治疗各种肝病的方法已成为热点研究领域,但自噬在肝脏蛋白质和脂质代谢中的作用极其机制尚不清楚。本文对自噬在肝脏蛋白质和脂质代谢中的作用的最新进展进行综述。  相似文献   

19.
Chloroplasts contain approximately 80% of total leaf nitrogen and represent a major source of recycled nitrogen during leaf senescence. While bulk degradation of the cytosol and organelles in plants is mediated by autophagy, its role in chloroplast catabolism is largely unknown. We investigated the effects of autophagy disruption on the number and size of chloroplasts during senescence. When leaves were individually darkened, senescence was promoted similarly in both wild-type Arabidopsis (Arabidopsis thaliana) and in an autophagy-defective mutant, atg4a4b-1. The number and size of chloroplasts decreased in darkened leaves of wild type, while the number remained constant and the size decrease was suppressed in atg4a4b-1. When leaves of transgenic plants expressing stroma-targeted DsRed were individually darkened, a large accumulation of fluorescence in the vacuolar lumen was observed. Chloroplasts exhibiting chlorophyll fluorescence, as well as Rubisco-containing bodies, were also observed in the vacuole. No accumulation of stroma-targeted DsRed, chloroplasts, or Rubisco-containing bodies was observed in the vacuoles of the autophagy-defective mutant. We have succeeded in demonstrating chloroplast autophagy in living cells and provide direct evidence of chloroplast transportation into the vacuole.Chloroplasts contain 75% to 80% of total leaf nitrogen mainly as proteins (Makino and Osmond, 1991). During leaf senescence, chloroplast proteins are gradually degraded as a major source of nitrogen for new growth (Wittenbach, 1978; Friedrich and Huffaker, 1980; Mae et al., 1984), correlating with a decline in photosynthetic activity, while chloroplasts gradually shrink and transform into gerontoplasts, characterized by the disintegration of the thylakoid membranes and accumulation of plastoglobuli (for a recent review, see Krupinska, 2006). Concomitantly, a decline in the cellular population of chloroplasts is also evident in many cases, for example, during natural (Kura-Hotta et al., 1990; Inada et al., 1998), dark-induced (Wittenbach et al., 1982), and nutrient-limited senescence (Mae et al., 1984; Ono et al., 1995), suggesting the existence of a whole chloroplast degradation system. Some electron microscopic studies have shown whole chloroplasts in the central vacuole, which is rich in lytic hydrolases (Wittenbach et al., 1982; Minamikawa et al., 2001). However, there is no direct evidence of chloroplasts moving into the vacuole in living cells and the mechanism of transport is not yet understood (Hörtensteiner and Feller, 2002; Krupinska, 2006).The most abundant chloroplast protein is Rubisco (EC 4.1.1.39), comprising approximately 50% of the soluble protein (Wittenbach, 1978). The amount of Rubisco decreases rapidly in the early phase of leaf senescence, although more slowly in the later phase (Friedrich and Huffaker, 1980; Mae et al., 1984). In contrast, the chloroplast number remains relatively constant, making it impossible to explain Rubisco loss solely by whole chloroplast degradation. However, the mechanism of intrachloroplastic Rubisco degradation is still unknown (for review, see Feller et al., 2008). Using immunoelectron microscopy, we previously demonstrated in naturally senescing wheat (Triticum aestivum) leaves that Rubisco is released from chloroplasts into the cytoplasm and transported to the vacuole for subsequent degradation in small spherical bodies, named Rubisco-containing bodies (RCBs; Chiba et al., 2003). Similar chloroplast-derived structures were also subsequently confirmed in senescent leaves of soybean (Glycine max) and/or Arabidopsis (Arabidopsis thaliana) by electron microscopy (Otegui et al., 2005), and recently in tobacco (Nicotiana tabacum) leaves by immunoelectron microscopy, although the authors gave them a different name, Rubisco vesicular bodies (Prins et al., 2008). RCBs have double membranes, which seem to be derived from the chloroplast envelope; thus, the RCB-mediated degradation of stromal proteins represents a potential mechanism for chloroplast shrinkage during senescence. We recently demonstrated that Rubisco and stroma-targeted fluorescent proteins can be mobilized to the vacuole by ATG-dependent autophagy via RCBs, using leaves treated with concanamycin A, a vacuolar H+-ATPase inhibitor (Ishida et al., 2008). To investigate further, we wished to observe chloroplast autophagy and degradation directly in living cells to determine whether autophagy is responsible for chloroplast shrinkage and whether it is involved in the vacuolar degradation of whole chloroplasts during leaf senescence.Autophagy is known to be a major system for the bulk degradation of intracellular proteins and organelles in the vacuole in yeast and plants, or the lysosome in animals (for detailed mechanisms, see reviews by Ohsumi, 2001; Levine and Klionsky, 2004; Thompson and Vierstra, 2005; Bassham et al., 2006). In those systems, a portion of the cytoplasm, including entire organelles, is engulfed in membrane-bound vesicles and delivered to the vacuole/lysosome. A recent genome-wide search confirmed that Arabidopsis has many genes homologous to the yeast autophagy genes (ATGs; Doelling et al., 2002; Hanaoka et al., 2002; for detailed functions of ATGs, see the reviews noted above). Using knockout mutants of ATGs and a monitoring system with an autophagy marker, GFP-ATG8, numerous studies have demonstrated the presence of the autophagy system in plants and its importance in several biological processes (Yoshimoto et al., 2004; Liu et al., 2005; Suzuki et al., 2005; Thompson et al., 2005; Xiong et al., 2005, 2007; Fujiki et al., 2007; Phillips et al., 2008). These articles suggest that autophagy plays an important role in nutrient recycling during senescence, especially in nutrient-starved plants. The atg mutants exhibited an accelerated loss of some chloroplast proteins, but not all, under nutrient-starved conditions and during senescence, suggesting that autophagy is not the sole mechanism for the degradation of chloroplast proteins; other, as yet unidentified systems must be responsible for the degradation of chloroplast contents when the ATG system is compromised (Levine and Klionsky, 2004; Bassham et al., 2006). However, it still remains likely that autophagy is responsible for the vacuolar degradation of chloroplasts in wild-type plants.Prolonged observation is generally required to follow leaf senescence events in naturally aging leaves and senescence-associated processes tend to become chaotic over time. To observe chloroplast degradation over a short period, and to draw clear conclusions, a suitable experimental model of leaf senescence is required. Weaver and Amasino (2001) reported that senescence is rapidly induced in individually darkened leaves (IDLs) of Arabidopsis, but retarded in plants subjected to full darkness. In addition, Keech et al. (2007) observed a significant decrease of both the number and size of chloroplasts in IDLs within 6 d.In this study, using IDLs as a senescence model, we aimed to investigate the involvement of autophagy in chloroplast degradation. We show direct evidence for the transport of whole chloroplasts and RCBs to the vacuole by autophagy.  相似文献   

20.
Levine B 《Cell》2005,120(2):159-162
The eukaryotic cell uses an evolutionarily conserved lysosomal pathway of self-digestion (autophagy) for survival when extracellular nutrients are limited. In this issue of Cell, new evidence indicates that autophagy is used to for survival when intracellular nutrients are limited by growth factor deprivation (Lum et al., 2005). Other recent studies indicate that the autophagy machinery is also used to degrade foreign microbial invaders (xenophagy).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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