首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
A new role is reported for CP12, a highly unfolded and flexible protein, mainly known for its redox function with A4 glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Both reduced and oxidized CP12 can prevent the in vitro thermal inactivation and aggregation of GAPDH from Chlamydomonas reinhardtii. This mechanism is thus not redox-dependent. The protection is specific to CP12, because other proteins, such as bovine serum albumin, thioredoxin, and a general chaperone, Hsp33, do not fully prevent denaturation of GAPDH. Furthermore, CP12 acts as a specific chaperone, since it does not protect other proteins, such as catalase, alcohol dehydrogenase, or lysozyme. The interaction between CP12 and GAPDH is necessary to prevent the aggregation and inactivation, since the mutant C66S that does not form any complex with GAPDH cannot accomplish this protection. Unlike the C66S mutant, the C23S mutant that lacks the N-terminal bridge is partially able to protect and to slow down the inactivation and aggregation. Tryptic digestion coupled to mass spectrometry confirmed that the S-loop of GAPDH is the interaction site with CP12. Thus, CP12 not only has a redox function but also behaves as a specific “chaperone-like protein” for GAPDH, although a stable and not transitory interaction is observed. This new function of CP12 may explain why it is also present in complexes involving A2B2 GAPDHs that possess a regulatory C-terminal extension (GapB subunit) and therefore do not require CP12 to be redox-regulated.CP12 is a small 8.2-kDa protein present in the chloroplasts of most photosynthetic organisms, including cyanobacteria (1, 2), higher plants (3), the diatom Asterionella formosa (4, 5), and green (1) and red algae (6). It allows the formation of a supramolecular complex between phosphoribulokinase (EC 2.7.1.19) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH),3 two key enzymes of the Calvin cycle pathway, and was recently shown to interact with fructose bisphosphate aldolase, another enzyme of the Calvin cycle pathway (7). The phosphoribulokinase·GAPDH·CP12 complex has been extensively studied in Chlamydomonas reinhardtii (8, 9) and in Arabidopsis thaliana (10, 11). In the green alga C. reinhardtii, the interaction between CP12 and GAPDH is strong (8). GAPDH may exist as a homotetramer composed of four GapA subunits (A4) in higher plants, cyanobacteria, and green and red algae (6, 12), but in higher plants, it can also exist as a heterotetramer (A2B2), composed of two subunits, GapA and GapB (13, 14). GapB, up to now, has exclusively been found in Streptophyta, but recently two prasinophycean green algae, Ostreococcus tauri and Ostreococcus lucimarinus, were also shown to possess a GapB gene, whereas CP12 is missing (15). The GapB subunit is similar to the GapA subunit but has a C-terminal extension containing two redox-regulated cysteine residues (16). Thus, although the A4 GAPDHs lack these regulatory cysteine residues (13, 14, 1720), they are also redox-regulated through its interaction with CP12, since the C terminus of this small protein resembles the C-terminal extension of the GapB subunit. The regulatory cysteine residues for GapA are thus supplied by CP12, as is well documented in the literature (1, 8, 11, 16).CP12 belongs to the family of intrinsically unstructured proteins (IUPs) (2126). The amino acid composition of these proteins causes them to have no or few secondary structures. Their total or partial lack of structure and their high flexibility allow them to be molecular adaptors (27, 28). They are often able to bind to several partners and are involved in most cellular functions (29, 30). Recently, some IUPs have been described in photosynthetic organisms (31, 32).There are many functional categories of IUPs (22, 33). They can be, for instance, involved in permanent binding and have (i) a scavenger role, neutralizing or storing small ligands; (ii) an assembler role by forming complexes; and (iii) an effector role by modulating the activity of a partner molecule (33). These functions are not exclusive; thus, CP12 can form a stable complex with GAPDH, regulating its redox properties (8, 34, 35), and can also bind a metal ion (36, 37). IUPs can also bind transiently to partners, and some of them have been found to possess a chaperone activity (31, 38). This chaperone function was first shown for α-synuclein (39) and for α-casein (40), which are fully disordered. The amino acid composition of IUPs is less hydrophobic than those of soluble proteins; hence, they lack hydrophobic cores and do not become insoluble when heated. Since CP12 belongs to this family, we tested if it was resistant to heat treatment and finally, since it is tightly bound to GAPDH, if it could prevent aggregation of its partner, GAPDH, an enzyme well known for its tendency to aggregate (4144) and consequently a substrate commonly used in chaperone studies (45, 46).Unlike chaperones, which form transient, dynamic complexes with their protein substrates through hydrophobic interactions (47, 48), CP12 forms a stable complex with GAPDH. The interaction involves the C-terminal part of the protein and the presence of negatively charged residues on CP12 (35). However, only a site-directed mutagenesis has been performed to characterize the interaction site on GAPDH. Although the mutation could have an indirect effect, the residue Arg-197 was shown to be a good candidate for the interaction site (49).In this report, we accordingly used proteolysis experiments coupled with mass spectrometry to detect which regions of GAPDH are protected by its association with CP12. To conclude, the aim of this report was to characterize a chaperone function of CP12 that had never been described before and to map the interaction site on GAPDH using an approach that does not involve site-directed mutagenesis.  相似文献   

2.
3.
4.
The antioxidant N-acetyl-l-cysteine prevented the autophagy-dependent delivery of mitochondria to the vacuoles, as examined by fluorescence microscopy of mitochondria-targeted green fluorescent protein, transmission electron microscopy, and Western blot analysis of mitochondrial proteins. The effect of N-acetyl-l-cysteine was specific to mitochondrial autophagy (mitophagy). Indeed, autophagy-dependent activation of alkaline phosphatase and the presence of hallmarks of non-selective microautophagy were not altered by N-acetyl-l-cysteine. The effect of N-acetyl-l-cysteine was not related to its scavenging properties, but rather to its fueling effect of the glutathione pool. As a matter of fact, the decrease of the glutathione pool induced by chemical or genetical manipulation did stimulate mitophagy but not general autophagy. Conversely, the addition of a cell-permeable form of glutathione inhibited mitophagy. Inhibition of glutathione synthesis had no effect in the strain Δuth1, which is deficient in selective mitochondrial degradation. These data show that mitophagy can be regulated independently of general autophagy, and that its implementation may depend on the cellular redox status.Autophagy is a major pathway for the lysosomal/vacuolar delivery of long-lived proteins and organelles, where they are degraded and recycled. Autophagy plays a crucial role in differentiation and cellular response to stress and is conserved in eukaryotic cells from yeast to mammals (1, 2). The main form of autophagy, macroautophagy, involves the non-selective sequestration of large portions of the cytoplasm into double-membrane structures termed autophagosomes, and their delivery to the vacuole/lysosome for degradation. Another process, microautophagy, involves the direct sequestration of parts of the cytoplasm by vacuole/lysosomes. The two processes coexist in yeast cells but their extent may depend on different factors including metabolic state: for example, we have observed that nitrogen-starved lactate-grown yeast cells develop microautophagy, whereas nitrogen-starved glucose-grown cells preferentially develop macroautophagy (3).Both macroautophagy and microautophagy are essentially non-selective, in the way that autophagosomes and vacuole invaginations do not appear to discriminate the sequestered material. However, selective forms of autophagy have been observed (4) that target namely peroxisomes (5, 6), chromatin (7, 8), endoplasmic reticulum (9), ribosomes (10), and mitochondria (3, 1113). Although non-selective autophagy plays an essential role in survival by nitrogen starvation, by providing amino acids to the cell, selective autophagy is more likely to have a function in the maintenance of cellular structures, both under normal conditions as a “housecleaning” process, and under stress conditions by eliminating altered organelles and macromolecular structures (1416). Selective autophagy targeting mitochondria, termed mitophagy, may be particularly relevant to stress conditions. The mitochondrial respiratory chain is both the main site and target of ROS4 production (17). Consequently, the maintenance of a pool of healthy mitochondria is a crucial challenge for the cells. The progressive accumulation of altered mitochondria (18) caused by the loss of efficiency of the maintenance process (degradation/biogenesis de novo) is often considered as a major cause of cellular aging (1923). In mammalian cells, autophagic removal of mitochondria has been shown to be triggered following induction/blockade of apoptosis (23), suggesting that autophagy of mitochondria was required for cell survival following mitochondria injury (14). Consistent with this idea, a direct alteration of mitochondrial permeability properties has been shown to induce mitochondrial autophagy (13, 24, 25). Furthermore, inactivation of catalase induced the autophagic elimination of altered mitochondria (26). In the yeast Saccharomyces cerevisiae, the alteration of F0F1-ATPase biogenesis in a conditional mutant has been shown to trigger autophagy (27). Alterations of mitochondrial ion homeostasis caused by the inactivation of the K+/H+ exchanger was shown to cause both autophagy and mitophagy (28). We have reported that treatment of cells with rapamycin induced early ROS production and mitochondrial lipid oxidation that could be inhibited by the hydrophobic antioxidant resveratrol (29). Furthermore, resveratrol treatment impaired autophagic degradation of both cytosolic and mitochondrial proteins and delayed rapamycin-induced cell death, suggesting that mitochondrial oxidation events may play a crucial role in the regulation of autophagy. This existence of regulation of autophagy by ROS has received molecular support in HeLa cells (30): these authors showed that starvation stimulated ROS production, namely H2O2, which was essential for autophagy. Furthermore, they identified the cysteine protease hsAtg4 as a direct target for oxidation by H2O2. This provided a possible connection between the mitochondrial status and regulation of autophagy.Investigations of mitochondrial autophagy in nitrogen-starved lactate-grown yeast cells have established the existence of two distinct processes: the first one occurring very early, is selective for mitochondria and is dependent on the presence of the mitochondrial protein Uth1p; the second one occurring later, is not selective for mitochondria, is not dependent on Uth1p, and is a form of bulk microautophagy (3). The absence of the selective process in the Δuth1 mutant strongly delays and decreases mitochondrial protein degradation (3, 12). The putative protein phosphatase Aup1p has been also shown to be essential in inducing mitophagy (31). Additionally several Atg proteins were shown to be involved in vacuolar sequestration of mitochondrial GFP (3, 12, 32, 33). Recently, the protein Atg11p, which had been already identified as an essential protein for selective autophagy has also been reported as being essential for mitophagy (33).The question remains as to identify of the signals that trigger selective mitophagy. It is particularly intriguing that selective mitophagy is activated very early after the shift to a nitrogen-deprived medium (3). Furthermore, selective mitophagy is very active on lactate-grown cells (with fully differentiated mitochondria) but is nearly absent in glucose-grown cells (3). In the present paper, we investigated the relationships between the redox status of the cells and selective mitophagy, namely by manipulating glutathione. Our results support the view that redox imbalance is a trigger for the selective elimination of mitochondria.  相似文献   

5.
6.
Ryanodine receptors (RyR) are Ca2+ channels that mediate Ca2+ release from intracellular stores in response to diverse intracellular signals. In RINm5F insulinoma cells, caffeine, and 4-chloro-m-cresol (4CmC), agonists of RyR, stimulated Ca2+ entry that was independent of store-operated Ca2+ entry, and blocked by prior incubation with a concentration of ryanodine that inactivates RyR. Patch-clamp recording identified small numbers of large-conductance (γK = 169 pS) cation channels that were activated by caffeine, 4CmC or low concentrations of ryanodine. Similar channels were detected in rat pancreatic β-cells. In RINm5F cells, the channels were blocked by cytosolic, but not extracellular, ruthenium red. Subcellular fractionation showed that type 3 IP3 receptors (IP3R3) were expressed predominantly in endoplasmic reticulum, whereas RyR2 were present also in plasma membrane fractions. Using RNAi selectively to reduce expression of RyR1, RyR2, or IP3R3, we showed that RyR2 mediates both the Ca2+ entry and the plasma membrane currents evoked by agonists of RyR. We conclude that small numbers of RyR2 are selectively expressed in the plasma membrane of RINm5F pancreatic β-cells, where they mediate Ca2+ entry.Ryanodine receptors (RyR)3 and inositol 1,4,5-trisphosphate receptors (IP3R) (1, 2) are the archetypal intracellular Ca2+ channels. Both are widely expressed, although RyR are more restricted in their expression than IP3R (3, 4). In common with many cells, pancreatic β-cells and insulin-secreting cell lines express both IP3R (predominantly IP3R3) (5, 6) and RyR (predominantly RyR2) (7). Both RyR and IP3R are expressed mostly within membranes of the endoplasmic (ER), where they mediate release of Ca2+. Functional RyR are also expressed in the secretory vesicles (8, 9) or, and perhaps more likely, in the endosomes of β-cells (10). Despite earlier suggestions (11), IP3R are probably not present in the secretory vesicles of β-cells (8, 12, 13).All three subtypes of IP3R are stimulated by IP3 with Ca2+ (1), and the three subtypes of RyR are each directly regulated by Ca2+. However, RyR differ in whether their most important physiological stimulus is depolarization of the plasma membrane (RyR1), Ca2+ (RyR2) or additional intracellular messengers like cyclic ADP-ribose. The latter stimulates both Ca2+ release and insulin secretion in β-cells (8, 14). The activities of both families of intracellular Ca2+ channels are also modulated by many additional signals that act directly or via phosphorylation (15, 16). Although they commonly mediate release of Ca2+ from the ER, both IP3R and RyR select rather poorly between Ca2+ and other cations (permeability ratio, PCa/PK ∼7) (1, 17). This may allow electrogenic Ca2+ release from the ER to be rapidly compensated by uptake of K+ (18), and where RyR or IP3R are expressed in other membranes it may allow them to affect membrane potential.Both Ca2+ entry and release of Ca2+ from intracellular stores contribute to the oscillatory increases in cytosolic Ca2+ concentration ([Ca2+]i) that stimulate exocytosis of insulin-containing vesicles in pancreatic β-cells (7). Glucose rapidly equilibrates across the plasma membrane (PM) of β-cells and its oxidative metabolism by mitochondria increases the cytosolic ATP/ADP ratio, causing KATP channels to close (19). This allows an unidentified leak current to depolarize the PM (20) and activate voltage-gated Ca2+ channels, predominantly L-type Ca2+ channels (21). The resulting Ca2+ entry is amplified by Ca2+-induced Ca2+ release from intracellular stores (7), triggering exocytotic release of insulin-containing dense-core vesicles (22). The importance of this sequence is clear from the widespread use of sulfonylurea drugs, which close KATP channels, in the treatment of type 2 diabetes. Ca2+ uptake by mitochondria beneath the PM further stimulates ATP production, amplifying the initial response to glucose and perhaps thereby contributing to the sustained phase of insulin release (23). However, neither the increase in [Ca2+]i nor the insulin release evoked by glucose or other nutrients is entirely dependent on Ca2+ entry (7, 24) or closure of KATP channels (25). This suggests that glucose metabolism may also more directly activate RyR (7, 26) and/or IP3R (27) to cause release of Ca2+ from intracellular stores. A change in the ATP/ADP ratio is one means whereby nutrient metabolism may be linked to opening of intracellular Ca2+ channels because both RyR (28) and IP3R (1) are stimulated by ATP.The other major physiological regulators of insulin release are the incretins: glucagon-like peptide-1 and glucose-dependent insulinotropic hormone (29). These hormones, released by cells in the small intestine, stimulate synthesis of cAMP in β-cells and thereby potentiate glucose-evoked insulin release (30). These pathways are also targets of drugs used successfully to treat type 2 diabetes (29). The responses of β-cells to cAMP involve both cAMP-dependent protein kinase and epacs (exchange factors activated by cAMP) (31, 32). The effects of the latter are, at least partly, due to release of Ca2+ from intracellular stores via RyR (3335) and perhaps also via IP3R (36). The interplays between Ca2+ and cAMP signaling generate oscillatory changes in the concentrations of both messengers (37). RyR and IP3R are thus implicated in mediating responses to each of the major physiological regulators of insulin secretion: glucose and incretins.Here we report that in addition to expression in intracellular stores, which probably include both the ER and secretory vesicles and/or endosomes, functional RyR2 are also expressed in small numbers in the PM of RINm5F insulinoma cells and rat pancreatic β-cells.  相似文献   

7.
8.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

9.
10.
11.
Thimet oligopeptidase (EC 3.4.24.15; EP24.15) is an intracellular enzyme that has been proposed to metabolize peptides within cells, thereby affecting antigen presentation and G protein-coupled receptor signal transduction. However, only a small number of intracellular substrates of EP24.15 have been reported previously. Here we have identified over 100 peptides in human embryonic kidney 293 (HEK293) cells that are derived from intracellular proteins; many but not all of these peptides are substrates or products of EP24.15. First, cellular peptides were extracted from HEK293 cells and incubated in vitro with purified EP24.15. Then the peptides were labeled with isotopic tags and analyzed by mass spectrometry to obtain quantitative data on the extent of cleavage. A related series of experiments tested the effect of overexpression of EP24.15 on the cellular levels of peptides in HEK293 cells. Finally, synthetic peptides that corresponded to 10 of the cellular peptides were incubated with purified EP24.15 in vitro, and the cleavage was monitored by high pressure liquid chromatography and mass spectrometry. Many of the EP24.15 substrates identified by these approaches are 9–11 amino acids in length, supporting the proposal that EP24.15 can function in the degradation of peptides that could be used for antigen presentation. However, EP24.15 also converts some peptides into products that are 8–10 amino acids, thus contributing to the formation of peptides for antigen presentation. In addition, the intracellular peptides described here are potential candidates to regulate protein interactions within cells.Intracellular protein turnover is a crucial step for cell functioning, and if this process is impaired, the elevated levels of aged proteins usually lead to the formation of intracellular insoluble aggregates that can cause severe pathologies (1). In mammalian cells, most proteins destined for degradation are initially tagged with a polyubiquitin chain in an energy-dependent process and then digested to small peptides by the 26 S proteasome, a large proteolytic complex involved in the regulation of cell division, gene expression, and other key processes (2, 3). In eukaryotes, 30–90% of newly synthesized proteins may be degraded by proteasomes within minutes of synthesis (3, 4). In addition to proteasomes, other extralysosomal proteolytic systems have been reported (5, 6). The proteasome cleaves proteins into peptides that are typically 2–20 amino acids in length (7). In most cases, these peptides are thought to be rapidly hydrolyzed into amino acids by aminopeptidases (810). However, some intracellular peptides escape complete degradation and are imported into the endoplasmic reticulum where they associate with major histocompatibility complex class I (MHC-I)3 molecules and traffic to the cell surface for presentation to the immune system (1012). Additionally, based on the fact that free peptides added to the intracellular milieu can regulate cellular functions mediated by protein interactions such as gene regulation, metabolism, cell signaling, and protein targeting (13, 14), intracellular peptides generated by proteasomes that escape degradation have been suggested to play a role in regulating protein interactions (15). Indeed, oligopeptides isolated from rat brain tissue using the catalytically inactive EP24.15 (EC 3.4.24.15) were introduced into Chinese hamster ovarian-S and HEK293 cells and were found capable of altering G protein-coupled receptor signal transduction (16). Moreover, EP24.15 overexpression itself changed both angiotensin II and isoproterenol signal transduction, suggesting a physiological function for its intracellular substrates/products (16).EP24.15 is a zinc-dependent peptidase of the metallopeptidase M3 family that contains the HEXXH motif (17). This enzyme was first described as a neuropeptide-degrading enzyme present in the soluble fraction of brain homogenates (18). Whereas EP24.15 can be secreted (19, 20), its predominant location in the cytosol and nucleus suggests that the primary function of this enzyme is not the extracellular degradation of neuropeptides and hormones (21, 22). EP24.15 was shown in vivo to participate in antigen presentation through MHC-I (2325) and in vitro to bind (26) or degrade (27) some MHC-I associated peptides. EP24.15 has also been shown in vitro to degrade peptides containing 5–17 amino acids produced after proteasome digestion of β-casein (28). EP24.15 shows substrate size restriction to peptides containing from 5 to 17 amino acids because of its catalytic center that is located in a deep channel (29). Despite the size restriction, EP24.15 has a broad substrate specificity (30), probably because a significant portion of the enzyme-binding site is lined with potentially flexible loops that allow reorganization of the active site following substrate binding (29). Recently, it has also been suggested that certain substrates may be cleaved by an open form of EP24.15 (31). This characteristic is supported by the ability of EP24.15 to accommodate different amino acid residues at subsites S4 to S3′, which even includes the uncommon post-proline cleavage (30). Such biochemical and structural features make EP24.15 a versatile enzyme to degrade structurally unrelated oligopeptides.Previously, brain peptides that bound to catalytically inactive EP24.15 were isolated and identified using mass spectrometry (22). The majority of peptides captured by the inactive enzyme were intracellular protein fragments that efficiently interacted with EP24.15; the smallest peptide isolated in these assays contained 5 and the largest 17 amino acids (15, 16, 22, 32), which is within the size range previously reported for natural and synthetic substrates of EP24.15 (18, 30, 33, 34). Interestingly, the peptides released by the proteasome are in the same size range of EP24.15 competitive inhibitors/substrates (7, 35, 36). Taken altogether, these data suggest that in the intracellular environment EP24.15 could further cleave proteasome-generated peptides unrelated to MHC-I antigen presentation (15).Although the mutated inactive enzyme “capture” assay was successful in identifying several cellular protein fragments that were substrates for EP24.15, it also found some interacting peptides that were not substrates. In this study, we used several approaches to directly screen for cellular peptides that were cleaved by EP24.15. The first approach involved the extraction of cellular peptides from the HEK293 cell line, incubation in vitro with purified EP24.15, labeling with isotopic tags, and analysis by mass spectrometry to obtain quantitative data on the extent of cleavage. The second approach examined the effect of EP24.15 overexpression on the cellular levels of peptides in the HEK293 cell line. The third set of experiments tested synthetic peptides with purified EP24.15 in vitro, and examined cleavage by high pressure liquid chromatography and mass spectrometry. Collectively, these studies have identified a large number of intracellular peptides, including those that likely represent the endogenous substrates and products of EP24.15, and this original information contributes to a better understanding of the function of this enzyme in vivo.  相似文献   

12.
Rheb G-protein plays critical roles in the TSC/Rheb/mTOR signaling pathway by activating mTORC1. The activation of mTORC1 by Rheb can be faithfully reproduced in vitro by using mTORC1 immunoprecipitated by the use of anti-raptor antibody from mammalian cells starved for nutrients. The low in vitro kinase activity against 4E-BP1 of this mTORC1 preparation is dramatically increased by the addition of recombinant Rheb. On the other hand, the addition of Rheb does not activate mTORC2 immunoprecipitated from mammalian cells by the use of anti-rictor antibody. The activation of mTORC1 is specific to Rheb, because other G-proteins such as KRas, RalA/B, and Cdc42 did not activate mTORC1. Both Rheb1 and Rheb2 activate mTORC1. In addition, the activation is dependent on the presence of bound GTP. We also find that the effector domain of Rheb is required for the mTORC1 activation. FKBP38, a recently proposed mediator of Rheb action, appears not to be involved in the Rheb-dependent activation of mTORC1 in vitro, because the preparation of mTORC1 that is devoid of FKBP38 is still activated by Rheb. The addition of Rheb results in a significant increase of binding of the substrate protein 4E-BP1 to mTORC1. PRAS40, a TOR signaling (TOS) motif-containing protein that competes with the binding of 4EBP1 to mTORC1, inhibits Rheb-induced activation of mTORC1. A preparation of mTORC1 that is devoid of raptor is not activated by Rheb. Rheb does not induce autophosphorylation of mTOR. These results suggest that Rheb induces alteration in the binding of 4E-BP1 with mTORC1 to regulate mTORC1 activation.Rheb defines a unique member of the Ras superfamily G-proteins (1). We have shown that Rheb proteins are conserved and are found from yeast to human (2). Although yeast and fruit fly have one Rheb, mouse and human have two Rheb proteins termed Rheb1 (or simply Rheb) and Rheb2 (RhebL1) (2). Structurally, these proteins contain G1-G5 boxes, short stretches of amino acids that define the function of the Ras superfamily G-proteins including guanine nucleotide binding (1, 3, 4). Rheb proteins have a conserved arginine at residue 15 that corresponds to residue 12 of Ras (1). The effector domain required for the binding with downstream effectors encompasses the G2 box and its adjacent sequences (1, 5). Structural analysis by x-ray crystallography further shows that the effector domain is exposed to solvent, is located close to the phosphates of GTP especially at residues 35–38, and undergoes conformational change during GTP/GDP exchange (6). In addition, all Rheb proteins end with the CAAX (C is cysteine, A is an aliphatic amino acid, and X is the C-terminal amino acid) motif that signals farnesylation. In fact, we as well as others have shown that these proteins are farnesylated (79).Rheb plays critical roles in the TSC/Rheb/mTOR signaling, a signaling pathway that plays central roles in regulating protein synthesis and growth in response to nutrient, energy, and growth conditions (1014). Rheb is down-regulated by a TSC1·TSC2 complex that acts as a GTPase-activating protein for Rheb (1519). Recent studies established that the GAP domain of TSC2 defines the functional domain for the down-regulation of Rheb (20). Mutations in the Tsc1 or Tsc2 gene lead to tuberous sclerosis whose symptoms include the appearance of benign tumors called hamartomas at different parts of the body as well as neurological symptoms (21, 22). Overexpression of Rheb results in constitutive activation of mTOR even in the absence of nutrients (15, 16). Two mTOR complexes, mTORC1 and mTORC2, have been identified (23, 24). Whereas mTORC1 is involved in protein synthesis activation mediated by S6K and 4EBP1, mTORC2 is involved in the phosphorylation of Akt in response to insulin. It has been suggested that Rheb is involved in the activation of mTORC1 but not mTORC2 (25).Although Rheb is clearly involved in the activation of mTOR, the mechanism of activation has not been established. We as well as others have suggested a model that involves the interaction of Rheb with the TOR complex (2628). Rheb activation of mTOR kinase activity using immunoprecipitated mTORC1 was reported (29). Rheb has been shown to interact with mTOR (27, 30), and this may involve direct interaction of Rheb with the kinase domain of mTOR (27). However, this Rheb/mTOR interaction is a weak interaction and is not dependent on the presence of GTP bound to Rheb (27, 28). Recently, a different model proposing that FKBP38 (FK506-binding protein 38) mediates the activation of mTORC1 by Rheb was proposed (31, 32). In this model, FKBP38 binds mTOR and negatively regulates mTOR activity, and this negative regulation is blocked by the binding of Rheb to FKBP38. However, recent reports dispute this idea (33).To further characterize Rheb activation of mTOR, we have utilized an in vitro system that reproduces activation of mTORC1 by the addition of recombinant Rheb. We used mTORC1 immunoprecipitated from nutrient-starved cells using anti-raptor antibody and have shown that its kinase activity against 4E-BP1 is dramatically increased by the addition of recombinant Rheb. Importantly, the activation of mTORC1 is specific to Rheb and is dependent on the presence of bound GTP as well as an intact effector domain. FKBP38 is not detected in our preparation and further investigation suggests that FKBP38 is not an essential component for the activation of mTORC1 by Rheb. Our study revealed that Rheb enhances the binding of a substrate 4E-BP1 with mTORC1 rather than increasing the kinase activity of mTOR.  相似文献   

13.
14.
15.
16.
Intersectin-short (intersectin-s) is a multimodule scaffolding protein functioning in constitutive and regulated forms of endocytosis in non-neuronal cells and in synaptic vesicle (SV) recycling at the neuromuscular junction of Drosophila and Caenorhabditis elegans. In vertebrates, alternative splicing generates a second isoform, intersectin-long (intersectin-l), that contains additional modular domains providing a guanine nucleotide exchange factor activity for Cdc42. In mammals, intersectin-s is expressed in multiple tissues and cells, including glia, but excluded from neurons, whereas intersectin-l is a neuron-specific isoform. Thus, intersectin-I may regulate multiple forms of endocytosis in mammalian neurons, including SV endocytosis. We now report, however, that intersectin-l is localized to somatodendritic regions of cultured hippocampal neurons, with some juxtanuclear accumulation, but is excluded from synaptophysin-labeled axon terminals. Consistently, intersectin-l knockdown (KD) does not affect SV recycling. Instead intersectin-l co-localizes with clathrin heavy chain and adaptor protein 2 in the somatodendritic region of neurons, and its KD reduces the rate of transferrin endocytosis. The protein also co-localizes with F-actin at dendritic spines, and intersectin-l KD disrupts spine maturation during development. Our data indicate that intersectin-l is indeed an important regulator of constitutive endocytosis and neuronal development but that it is not a prominent player in the regulated endocytosis of SVs.Clathrin-mediated endocytosis (CME)4 is a major mechanism by which cells take up nutrients, control the surface levels of multiple proteins, including ion channels and transporters, and regulate the coupling of signaling receptors to downstream signaling cascades (1-5). In neurons, CME takes on additional specialized roles; it is an important process regulating synaptic vesicle (SV) availability through endocytosis and recycling of SV membranes (6, 7), it shapes synaptic plasticity (8-10), and it is crucial in maintaining synaptic membranes and membrane structure (11).Numerous endocytic accessory proteins participate in CME, interacting with each other and with core components of the endocytic machinery such as clathrin heavy chain (CHC) and adaptor protein-2 (AP-2) through specific modules and peptide motifs (12). One such module is the Eps15 homology domain that binds to proteins bearing NPF motifs (13, 14). Another is the Src homology 3 (SH3) domain, which binds to proline-rich domains in protein partners (15). Intersectin is a multimodule scaffolding protein that interacts with a wide range of proteins, including several involved in CME (16). Intersectin has two N-terminal Eps15 homology domains that are responsible for binding to epsin, SCAMP1, and numb (17-19), a central coil-coiled domain that interacts with Eps15 and SNAP-23 and -25 (17, 20, 21), and five SH3 domains in its C-terminal region that interact with multiple proline-rich domain proteins, including synaptojanin, dynamin, N-WASP, CdGAP, and mSOS (16, 22-25). The rich binding capability of intersectin has linked it to various functions from CME (17, 26, 27) and signaling (22, 28, 29) to mitogenesis (30, 31) and regulation of the actin cytoskeleton (23).Intersectin functions in SV recycling at the neuromuscular junction of Drosophila and C. elegans where it acts as a scaffold, regulating the synaptic levels of endocytic accessory proteins (21, 32-34). In vertebrates, the intersectin gene is subject to alternative splicing, and a longer isoform (intersectin-l) is generated that is expressed exclusively in neurons (26, 28, 35, 36). This isoform has all the binding modules of its short (intersectin-s) counterpart but also has additional domains: a DH and a PH domain that provide guanine nucleotide exchange factor (GEF) activity specific for Cdc42 (23, 37) and a C2 domain at the C terminus. Through its GEF activity and binding to actin regulatory proteins, including N-WASP, intersectin-l has been implicated in actin regulation and the development of dendritic spines (19, 23, 24). In addition, because the rest of the binding modules are shared between intersectin-s and -l, it is generally thought that the two intersectin isoforms have the same endocytic functions. In particular, given the well defined role for the invertebrate orthologs of intersectin-s in SV endocytosis, it is thought that intersectin-l performs this role in mammalian neurons, which lack intersectin-s. Defining the complement of intersectin functional activities in mammalian neurons is particularly relevant given that the protein is involved in the pathophysiology of Down syndrome (DS). Specifically, the intersectin gene is localized on chromosome 21q22.2 and is overexpressed in DS brains (38). Interestingly, alterations in endosomal pathways are a hallmark of DS neurons and neurons from the partial trisomy 16 mouse, Ts65Dn, a model for DS (39, 40). Thus, an endocytic trafficking defect may contribute to the DS disease process.Here, the functional roles of intersectin-l were studied in cultured hippocampal neurons. We find that intersectin-l is localized to the somatodendritic regions of neurons, where it co-localizes with CHC and AP-2 and regulates the uptake of transferrin. Intersectin-l also co-localizes with actin at dendritic spines and disrupting intersectin-l function alters dendritic spine development. In contrast, intersectin-l is absent from presynaptic terminals and has little or no role in SV recycling.  相似文献   

17.
18.
19.
20.
Bacillus cereus and other Gram-positive bacteria elaborate pili via a sortase D-catalyzed transpeptidation mechanism from major and minor pilin precursor substrates. After cleavage of the LPXTG sorting signal of the major pilin, BcpA, sortase D forms an amide bond between the C-terminal threonine and the amino group of lysine within the YPKN motif of another BcpA subunit. Pilus assembly terminates upon sortase A cleavage of the BcpA sorting signal, resulting in a covalent bond between BcpA and the cell wall cross-bridge. Here, we show that the IPNTG sorting signal of BcpB, the minor pilin, is cleaved by sortase D but not by sortase A. The C-terminal threonine of BcpB is amide-linked to the YPKN motif of BcpA, thereby positioning BcpB at the tip of pili. Thus, unique attributes of the sorting signals of minor pilins provide Gram-positive bacteria with a universal mechanism ordering assembly of pili.Sortases catalyze transpeptidation reactions to assemble proteins in the envelope of Gram-positive bacteria (1). Secreted proteins require a C-terminal sorting signal for sortase recognition such that sortase cleaves the substrate at a short peptide motif and forms a thioester-linked intermediate to its active site cysteine (24). Nucleophilic attack by an amino group within the bacterial envelope resolves the thioester intermediate, generating an amide bond tethering surface proteins at their C terminus onto Gram-positive bacteria (5). Four classes of sortases can be distinguished on the basis of sequence homology and substrate recognition (6, 7). Sortase A cleaves secreted protein at LPXTG sorting signals and recognizes the amino group of lipid II peptidoglycan precursors as a nucleophile (8, 9). Sortase B cleaves protein substrates at NPQTN sorting signals (10). This enzyme immobilizes proteins within fully assembled cell walls, utilizing the cell wall cross-bridge as a nucleophile (11). Sortase C cuts LPNTA sorting signals and anchors proteins to the peptidoglycan cross-bridges in sporulating bacteria (12, 13). Finally, sortase D catalyzes transpeptidation reactions in the assembly of pili (14, 15). Sortase D recognizes the amino group of lysine residues within the YPKN motif of pilin subunits as nucleophiles (16). The resultant sortase D-catalyzed amide bond links adjacent pilin subunits to grow the pilus fiber (16, 17).Pili of Gram-positive bacteria comprised either two or three different pilin subunits synthesized as cytoplasmic precursors with N-terminal signal peptides and C-terminal sorting signals (P1 precursors) (14, 18). After translocation across the plasma membrane, P2 precursor species arise from removal of the signal peptide from P1 precursors by a signal peptidase (16). Bacillus cereus pili are composed of two subunits; that is, the major pilin, BcpA, and the minor pilin, BcpB (15). In contrast to BcpA, which is deposited throughout the pilus, BcpB is found at fiber tip (15). Sortase D cleaves the BcpA LPXTG motif sorting signal between the threonine and glycine residues to form an amide bond to the ε-amino group of the lysine within the YPKN motif of adjacent BcpA subunits (16). However, sortase A also cleaves BcpA precursors, which are subsequently linked to the side chain amino group of meso-diaminopimelic acid within lipid II (19). The latter reaction serves to terminate fiber elongation, immobilizing BcpA pili in the cell wall envelope (19).The conservation of sortase D, the YPKN motif, and C-terminal sorting signal in major pilin subunits suggest a universal pilus assembly mechanism among Gram-positive bacteria (14, 20). However, the molecular mechanism whereby bacilli deposit BcpB, the minor pilin, at the tip of BcpA pili is not known. Although the BcpB precursor harbors an N-terminal signal peptide and a C-terminal IPNTG sorting signal, it lacks the YPKN pilin motif of the major subunit (15). Furthermore, the substrate properties of the BcpB IPNTG sorting signal for the four classes of sortases expressed by bacilli has yet to be established.  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