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.
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   

2.
3.
4.
RPE65 is a membrane-associated protein abundantly expressed in the retinal pigment epithelium, which converts all-trans-retinyl ester to 11-cis-retinol, a key step in the retinoid visual cycle. Although three cysteine residues (Cys-231, Cys-329, and Cys-330) were identified to be palmitylated in RPE65, recent studies showed that a triple mutant, with all three Cys replaced by an alanine residue, was still palmitylated and remained membrane-associated, suggesting that there are other yet to be identified palmitylated Cys residues in RPE65. Here we mapped the entire RPE65 using mass spectrometry analysis and demonstrated that a trypsin-digested RPE65 fragment (residues 98-118), which contains two Cys residues (Cys-106 and Cys-112), was singly palmitylated in both native bovine and recombinant human RPE65. To determine whether Cys-106 or Cys-112 is the palmitylation site, these Cys were separately replaced by alanine. Mass spectrometry analysis of purified wild-type RPE65 and C106A and C112A mutants showed that mutation of Cys-106 did not affect the palmitylation status of the fragment 98-118, whereas mutation of Cys-112 abolished palmitylation in this fragment. Subcellular fractionation and immunocytochemistry analyses both showed that mutation of Cys-112 dissociated RPE65 from the membrane, whereas the C106A mutant remained associated with the membrane. In vitro isomerohydrolase activity assay showed that C106A has an intact enzymatic activity similar to that of wtRPE65, whereas C112A lost its enzymatic activity. These results indicate that the newly identified Cys-112 palmitylation site is essential for the membrane association and activity of RPE65.Both rod and cone visual pigments in vertebrates require 11-cis-retinal as the chromophore. Isomerization of 11-cis-retinal to all-trans-retinal by a photon triggers the phototransduction cascade and initiates vision (1, 2). Recycling of 11-cis-retinal through the retinoid visual cycle is an essential process for the regeneration of visual pigments and for normal vision (3, 4). The key step in the visual cycle is to isomerize all-trans-retinyl ester to 11-cis-retinol in retinal pigment epithelium (RPE)2 (5, 6). This isomerization process is known to be catalyzed by an isomerohydrolase in the RPE. Several recent lines of evidence suggest that RPE65 is the isomerohydrolase in the visual cycle (7-9).RPE65 is a microsomal protein, abundantly expressed in the RPE (10-12). RPE65 knock-out (Rpe65-/-) mice showed a lack of 11-cis-retinoids, overaccumulation of all-trans-retinyl ester, impaired visual function, and early degeneration of cone photoreceptors (7-9). RPE65 is an iron(II)-dependent enzyme, in which an iron is coordinated by four conserved histidine (His) residues (His-180, -241, -313, and -527) based on molecular modeling using a crystal structure of apocarotenoid monooxygenase as a template (8, 13-15). RPE65 lacks any predicted transmembrane helix and is associated with the microsomal membrane (11). Previous studies have shown that membrane association of RPE65 is essential for its isomerohydrolase activity (7). However, the structural basis for its membrane association has not been revealed. An earlier study showed that three Cys residues (Cys-231, 329 and 330) in RPE65 were palmitylated, which were suggested to be responsible for its membrane association (16). However, triple mutations of all the three Cys residues did not completely dissociate RPE65 from the membrane (17, 18). Moreover, the triple Cys mutant remains palmitylated (17). These results suggested that either the site of palmitylation responsible for the membrane association of RPE65 had not yet been identified or other mechanisms, such as hydrophobic interactions, anchor the protein to cellular membranes (17, 19).In this study, we used the combination of mass spectrometric analysis and site-directed mutagenesis to identify the palmitylated site in RPE65. Moreover, we determined the role of this site in the membrane association and enzymatic activity of RPE65.  相似文献   

5.
The kinetochore, which consists of DNA sequence elements and structural proteins, is essential for high-fidelity chromosome transmission during cell division. In budding yeast, Sgt1 and Hsp90 help assemble the core kinetochore complex CBF3 by activating the CBF3 components Skp1 and Ctf13. In this study, we show that Sgt1 forms homodimers by performing in vitro and in vivo immunoprecipitation and analytical ultracentrifugation analyses. Analyses of the dimerization of Sgt1 deletion proteins showed that the Skp1-binding domain (amino acids 1–211) contains the Sgt1 homodimerization domain. Also, the Sgt1 mutant proteins that were unable to dimerize also did not bind Skp1, suggesting that Sgt1 dimerization is important for Sgt1-Skp1 binding. Restoring dimerization activity of a dimerization-deficient sgt1 mutant (sgt1-L31P) by using the CENP-B (centromere protein-B) dimerization domain suppressed the temperature sensitivity, the benomyl sensitivity, and the chromosome missegregation phenotype of sgt1-L31P. These results strongly suggest that Sgt1 dimerization is required for kinetochore assembly.Spindle microtubules are coupled to the centromeric region of the chromosome by a structural protein complex called the kinetochore (1, 2). The kinetochore is thought to generate a signal that arrests cells during mitosis when it is not properly attached to microtubules, thereby preventing aberrant chromosome transmission to the daughter cells, which can lead to tumorigenesis (3, 4). The kinetochore of the budding yeast Saccharomyces cerevisiae has been characterized thoroughly, genetically and biochemically; thus, its molecular structure is the most well detailed to date. More than 70 different proteins comprise the budding yeast kinetochore, and several of those are conserved in mammals (2).The budding yeast centromere DNA is a 125-bp region that contains three conserved regions, CDEI, CDEII, and CDEIII (5, 6). CDEI is bound by Cbf1 (79). CDEIII (25 bp) is essential for centromere function (10) and is the site where CBF3 binds to centromeric DNA. CBF3 contains four proteins: Ndc10, Cep3, Ctf13 (1118), and Skp1 (17, 18), all of which are essential for viability. Mutations in any of the four CBF3 proteins abolish the ability of CDEIII to bind to CBF3 (19, 20). All of the described kinetochore proteins, except the CDEI-binding Cbf1, localize to kinetochores dependent on the CBF3 complex (2). Therefore, the CBF3 complex is the fundamental structure of the kinetochore, and the mechanism of CBF3 assembly is of major interest.We previously isolated SGT1, the skp1-4 kinetochore-defective mutant dosage suppressor (21). Sgt1 and Skp1 activate Ctf13; thus, they are required for assembly of the CBF3 complex (21). The molecular chaperone Hsp90 is also required for the formation of the Skp1-Ctf13 complex (22). Sgt1 has two highly conserved motifs that are required for protein-protein interaction, the tetratricopeptide repeat (TPR)2 (21) and the CS (CHORD protein- and Sgt1-specific) motif. We and others (2326) have found that both domains are important for the interaction with Hsp90. The Sgt1-Hsp90 interaction is required for the assembly of the core kinetochore complex; this interaction is an initial step in kinetochore assembly (24, 26, 27) that is conserved between yeast and humans (28, 29).In this study, we further characterized the molecular mechanism of this assembly process. We found that Sgt1 forms dimers in vivo, and our results strongly suggest that Sgt1 dimerization is required for kinetochore assembly in budding yeast.  相似文献   

6.
Previously we have shown that tyrosine 718 of ASK1 when phosphorylated is critical for SOCS1 binding and SOCS1-mediated degradation of ASK1. However, the kinase and phosphatase responsible for phosphorylation and dephosphorylation of ASK1 at Tyr-718 are unknown. In this study, we identified JAK2 and SHP2 as a Tyr-718-specific kinase and phosphatase, respectively. Interferon-γ (IFN-γ) induced degradation of ASK1 in normal but not in SOCS1-KO endothelial cells (EC). IFN-γ-induced tyrosine phosphorylation of ASK1 at Tyr-718 was blocked by a JAK2-specific inhibitor. IFN-γ enhanced the association between JAK2 and ASK1, and the ASK1-JAK2 complex was labile and was stabilized by the proteasomal inhibitor MG132. Furthermore, JAK2, but not JAK1, directly bound to and phosphorylated ASK1 at Tyr-718, leading to an enhanced association of ASK1 with SOCS1 and subsequent ASK1 degradation. Next, we showed that overexpression of the SH2-containing protein-tyrosine phosphatase-2 (SHP2) augmented, whereas a phosphatase-inactive mutant of SHP2 inhibited, TNF-induced ASK1 dephosphorylation. SHP2 associated with ASK1 in response to tumor necrosis factor in EC. An SHP-2 substrate-trapping mutant formed a complex with tyrosine-phosphorylated ASK1, suggesting that ASK1 is a direct SHP2 substrate. Moreover, SHP2 wild type, but not a catalytically inactive mutant, dissociated SOCS1 from ASK1. IFN-γ-induced ASK1 Tyr(P)-718 was enhanced in mouse EC deficient in SHP2 (SHP2-KO). In contrast, tumor necrosis factor-induced dephosphorylation of ASK1 at Tyr(P)-718 and activation of ASK1-JNK signaling, as well as EC apoptosis, are significantly reduced in SHP2-KO EC. Our data suggest that JAK2-SOCS1 and SHP2 reciprocally regulate ASK1 phosphorylation and stability in response to cytokines.Myocardial infarction due to atherosclerosis of coronary arteries remains the leading cause of death in the United States. It has become clear that increases in inflammatory mediators represent a common pathogenic mechanism for atherosclerosis (1). The vascular cell that normally limits the inflammatory and atherosclerotic process is the EC.3 Proinflammatory stimuli induce EC dysfunction, which is characterized by an enhanced sensitivity of vascular cells to proinflammatory and proapoptotic stimuli. Studies from our laboratory and others have demonstrated that ASK1 (apoptosis signal-regulating kinase-1), a member of MAP3K family (2, 3), is an effector of inflammation in EC (48). Almost all inflammatory stimuli such as tumor necrosis factor-α (TNF), interleukin-1 (IL-1), and reactive oxygen species activate ASK1. Activated ASK1 subsequently recruits and activates its downstream target MAP2Ks (MKK3/7 and MKK4/7), which in turn activate MAPKs (JNK and p38). Studies from ASK1-deficient mice have also linked ASK1 to cardiovascular pathogenesis. ASK1 deletion in mice attenuated angiotensin II-induced cardiac hypertrophy and remodeling. Neointimal formation due to proliferation of smooth muscle cells in a cuff injury model is also attenuated by ASK1 deletion in mice (9, 10).Although the linkage of ASK1 to inflammation is very strong, the mechanism by which inflammatory stimuli, including TNF, activate ASK1 is not fully understood. The identification of proteins associated with ASK1 and their regulation on ASK1 have provided some insights into the mechanism for ASK1 activation. ASK1 is a 170-kDa protein that is composed of an inhibitory N-terminal domain, an internal kinase domain, and a C-terminal regulatory domain. One important regulatory mechanism of ASK1 activity is its Ser/Thr phosphorylation and dephosphorylation by kinases and phosphatases. ASK1 is basally phosphorylated at Ser-967 by an unidentified kinase, and 14-3-3 binds to this site and inhibits ASK1 activity (11, 12). TNF activates ASK1 in part by dissociating these cellular inhibitors from ASK1 (4, 7). Recently, we have identified PP2A as a phosphatase in TNF-induced dephosphorylation of ASK1 Ser(P)-967 (13). In addition to the 14-3-3-binding site, Ser(P)-967, ASK1 is phosphorylated at Ser-83 by Akt, leading to inhibition of ASK1 activity. In contrast, autophosphorylation of ASK1 at Thr-838 leads to oligomerization and activation (14). Phosphorylation of Thr-845 can be negatively regulated by the phosphatase PP5 (15). Similarly, we found that the ASK1 autophosphorylation at Thr-813 and Thr-842 also positively regulates ASK1 signaling (16).In contrast to Ser/Thr phosphorylation, regulation of ASK1 by tyrosine phosphorylation is less well understood. We have recently shown that ASK1 is phosphorylated at Tyr-718, and this phosphorylation is critical for the binding to suppressor of cytokine signaling-1 (SOCS1), a subunit of ubiquitin ligase responsible for ASK1 degradation (17). Tyrosine phosphorylation of ASK1 is up-regulated in response to growth factors and cytokines such as IFN-γ, whereas this phosphorylation can be down-regulated by TNF treatment, resulting in ASK1 dissociation from SOCS1. However, the kinase and phosphatase responsible for phosphorylation and dephosphorylation of ASK1 at Tyr-718 are not known.The cytoplasmic tyrosine kinase, JAK2, autophosphorylates in response to growth factors and cytokines, including IFN-γ. JAK2 then activates cytokine receptors and other cytoplasmic proteins such as the STATs by phosphorylating their key tyrosine residue. The JAK/STAT pathway can be regulated by SH2-containing protein-tyrosine phosphatases such as SHP2 (1820). SHP2 is ubiquitously expressed and composed of two SH2 domains on the N-terminal and C-terminal protein-tyrosine phosphatase (PTP) domain. The SH2 domain of SHP2 mediates the association with phosphotyrosine-containing proteins present on activated receptors as well as on activated JAKs and STATs; this association triggers activation of the tyrosine phosphatase domain and subsequent dephosphorylation of substrates. SHP2 signals downstream of receptor tyrosine kinases and cytokine receptors, and in most cases it serves to positively transduce signals from these receptors. In other instances SHP2 has been shown to exhibit inhibitory signaling properties by negatively regulating the JAK-STAT pathway (19).In this study, we demonstrate that the IFN-γ-activated kinase JAK2 and TNF-activated SHP2 are the tyrosine kinase and phosphatase for Tyr-718 on ASK1, respectively. The actions of both JAK2 and SHP2 affect protein turnover of ASK1 and thus regulate ASK1/JNK-dependent proinflammatory and proapoptotic pathways in EC.  相似文献   

7.
The presence of extensive reciprocal conformational freedom between the catalytic and the hemopexin-like domains of full-length matrix metalloproteinase-1 (MMP-1) is demonstrated by NMR and small angle x-ray scattering experiments. This finding is discussed in relation to the essentiality of the hemopexin-like domain for the collagenolytic activity of MMP-1. The conformational freedom experienced by the present system, having the shortest linker between the two domains, when compared with similar findings on MMP-12 and MMP-9 having longer and the longest linker within the family, respectively, suggests this type of conformational freedom to be a general property of all MMPs.Matrix metalloproteinases (MMP)2 are extracellular hydrolytic enzymes involved in a variety of processes including connective tissue cleavage and remodeling (13). All 23 members of the family are able to cleave simple peptides derived from connective tissue components such as collagen, gelatin, elastin, etc. A subset of MMPs is able to hydrolyze more resistant polymeric substrates, such as cross-linked elastin, and partially degraded collagen forms, such as gelatin and type IV collagens (4). Intact triple helical type I–III collagen is only attacked by collagenases MMP-1, MMP-8, and MMP-13 and by MMP-2 and MMP-14 (512). Although the detailed mechanism of cleavage of single chain peptides by MMP has been largely elucidated (1319), little is known about the process of hydrolysis of triple helical collagen. In fact, triple helical collagen cannot be accommodated in the substrate-binding groove of the catalytic site of MMPs (9).All MMPs (but MMP-7) in their active form are constituted by a catalytic domain (CAT) and a hemopexin-like domain (HPX) (2022). The CAT domain contains two zinc ions and one to three calcium ions. One zinc ion is at the catalytic site and is responsible for the activity, whereas the other metal ions have structural roles. The isolated CAT domains retain full catalytic activity toward simple peptides and single chain polymeric substrates such as elastin, whereas hydrolysis of triple helical collagen also requires the presence of the HPX domain (9, 2325). It has been shown that the isolated CAT domain regains a small fraction of the activity of the full-length (FL) protein when high amounts of either inactivated full-length proteins or isolated HPX domains are added to the assay solution (9). Finally, it has been shown that the presence of the HPX domain alone alters the CD spectrum of triple helical collagen in a way that suggests its partial unwinding (26, 27). It is tempting to speculate that full-length collagenases attack collagen by first locally unwinding the triple helical structure with the help of the HPX domain and then cleaving the resulting, exposed, single filaments (9, 28).Until 2007, three-dimensional structures of full-length MMPs had been reported only for collagenase MMP-1 (2931) and gelatinase MMP-2 (32). The structures of the two proteins are very similar and show a compact arrangement of the two domains, which are connected by a short linker (14 and 20 amino acids, respectively). It is difficult to envisage that rigid and compact molecules of this type can interact with triple helical collagen in a way that can lead to first unwinding and then cleavage of individual filaments. It has been recently suggested that such concerted action could occur much more easily if the two domains could enjoy at least a partial conformational independence (9). Slight differences in the reciprocal orientation of the CAT and HPX domains of MMP-1 in the presence (29) and absence (30, 31) of the prodomain were indeed taken as a hint that the two domains could experience relative mobility (29).Two recent solution studies have shown that conformational independence is indeed occurring in gelatinase MMP-9 (33) and elastase MMP-12 (34), whereas the x-ray structure of the latter (34) is only slightly less compact than those of MMP-1 (2931) and MMP-2 (32). Among MMPs, MMP-9 features an exceptionally long linker (68 amino acid) (33, 35), which in fact constitutes a small domain by itself (the O-glycosylated domain) (33), and therefore, this inspiring observation can hardly be taken as evidence that conformational freedom is a general characteristic of the two-domain MMPs. MMP-12 features a much more normal 16-amino acid linker, thereby making more probable a general functional role for this conformational freedom (34). However, both MMP-9 and MMP-12 retain their full catalytic activity against their substrates even when deprived of the HPX domain (9). Therefore, the question remains of whether conformational freedom is also a required characteristic for those MMPs that are only active as full-length proteins, i.e. collagenases. Interestingly, the three collagenases (MMP-1, MMP-8, and MMP-13) have the shortest linker (14 amino acids) among all MMPs. Demonstrating or negating the presence of conformational freedom in one of these collagenases would therefore constitute a significant step forward to formulate mechanistic hypotheses on their collagenolytic activity.Our recent studies on MMP-12 in solution (34) have shown that a combination of NMR relaxation studies and small angle x-ray scattering (SAXS) is enough to show the presence and the extent of the relative conformational freedom of the two domains of MMPs. Here we apply the same strategy to full-length MMP-1 and show that sizable conformational freedom is indeed experienced even by this prototypical collagenase, although somewhat less pronounced than that observed for MMP-12.  相似文献   

8.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  相似文献   

9.
It is known that platelet-activating factor (PAF) induces severe endothelial barrier leakiness, but the signaling mechanisms remain unclear. Here, using a wide range of biochemical and morphological approaches applied in both mouse models and cultured endothelial cells, we addressed the mechanisms of PAF-induced disruption of interendothelial junctions (IEJs) and of increased endothelial permeability. The formation of interendothelial gaps filled with filopodia and lamellipodia is the cellular event responsible for the disruption of endothelial barrier. We observed that PAF ligation of its receptor induced the activation of the Rho GTPase Rac1. Following PAF exposure, both Rac1 and its guanine nucleotide exchange factor Tiam1 were found associated with a membrane fraction from which they co-immunoprecipitated with PAF receptor. In the same time frame with Tiam1-Rac1 translocation, the junctional proteins ZO-1 and VE-cadherin were relocated from the IEJs, and formation of numerous interendothelial gaps was recorded. Notably, the response was independent of myosin light chain phosphorylation and thus distinct from other mediators, such as histamine and thrombin. The changes in actin status are driven by the PAF-induced localized actin polymerization as a consequence of Rac1 translocation and activation. Tiam1 was required for the activation of Rac1, actin polymerization, relocation of junctional associated proteins, and disruption of IEJs. Thus, PAF-induced IEJ disruption and increased endothelial permeability requires the activation of a Tiam1-Rac1 signaling module, suggesting a novel therapeutic target against increased vascular permeability associated with inflammatory diseases.The endothelial barrier is made up of endothelial cells (ECs)4 connected to each other by interendothelial junctions (IEJs) consisting of protein complexes organized as tight junctions (TJs) and adherens junctions (AJs). In addition, the focal adhesion complex located at the basal plasma membrane enables firm contact of ECs with the underlying basement membrane and also contributes to the barrier function (1-3). The glycocalyx, the endothelial monolayer, and the basement membrane all together constitute the vascular barrier.The structural integrity of the ECs along with their proper functionality are the two most important factors controlling the tightness of the endothelial barrier. Changes affecting these factors cause loss of barrier restrictiveness and leakiness. Therefore, defining and understanding the cellular and molecular mechanisms controlling these processes is of paramount importance. Increased width of IEJs in response to permeability-increasing mediators (4) regulates the magnitude of transendothelial exchange of fluid and solutes. Disruption of IEJs and the resultant barrier leakiness contribute to the genesis of diverse pathological conditions, such as inflammation (5), metastasis (6, 7), and uncontrolled angiogenesis (8, 9).Accumulated evidence demonstrated that IEJs changes are responsible for increased or decreased vascular permeability, and the generally accepted mechanism responsible for them was the myosin light chain (MLC)-mediated contraction of ECs (5, 10). However, published evidence showed that an increase in vascular permeability could be obtained without a direct involvement of any contractile mechanism (11-16).The main component of the vascular barrier, the ECs, has more than 10% of their total protein represented by actin (17), which under physiological salt concentrations subsists as monomers (G-actin) and assembled into filaments (F-actin). A large number of actin-interacting proteins may modulate the assembly, disassembly, and organization of G-actin and of actin filaments within a given cell type. Similar to the complexity of actin-interacting proteins found in other cell types, the ECs utilize their actin binding proteins to stabilize the endothelial monolayer in order to efficiently function as a selective barrier (11). In undisturbed ECs, the actin microfilaments are organized as different networks with distinctive functional and morphological characteristics: the peripheral filaments also known as peripheral dense band (PDB), the cytoplasmic fibers identified as stress fibers (SF), and the actin from the membrane cytoskeleton (18). The peripheral web, localized immediately under the membrane, is associated with (i) the luminal plasmalemma (on the apical side), (ii) the IEJ complexes on the lateral surfaces, and (iii) the focal adhesion complexes on the abluminal side (the basal part) of polarized ECs. The SF reside inside the endothelial cytoplasm and are believed to be directly connected with the plasmalemma proper on the luminal as well as on the abluminal side of the cell. As described, the endothelial actin cytoskeleton (specifically the SF) seems to be a stable structure helping the cells to remain flat under flow (19). It is also established that the actin fibers participate in correct localization of different junctional complexes while keeping them in place (20). However, it was suggested that the dynamic equilibrium between F- and G-actin might modulate the tightness of endothelial barrier in response to different challenges (13).Mediators effective at nanomolar concentrations or less that disrupt the endothelial barrier and increase vascular permeability include C2 toxin of Clostridium botulinum, vascular permeability factor, better known as vascular endothelial growth factor, and PAF (21). C2 toxin increases endothelial permeability by ribosylating monomeric G-actin at Arg-177 (22). This results in the impairment of actin polymerization (23), followed by rounding of ECs (16) and the disruption of junctional integrity. Vascular permeability factor was shown to open IEJs by redistribution of junctional proteins (24, 25) and by interfering with the equilibrium of actin pools (26). PAF (1-O-alkyl-2-acetyl-sn-glycero-3-phosphocoline), a naturally synthesized phospholipid is active at 10-10 m or less (27). PAF is synthesized by and acts on a variety of cell types, including platelets (28), neutrophils (29), monocytes (30), and ECs (31). PAF-mediated activation of ECs induced cell migration (32), angiogenesis (7), and vascular hyperpermeability (33) secondary to disassembly of IEJs (34). The effects of PAF on the endothelium are initiated through a G protein-coupled receptor (PAF-R) localized at the plasmalemma, in a large endosomal compartment inside the cell (34), and also in the nuclear membrane (35). In ECs, PAF-R was shown to signal through Gαq and downstream activation of phospholipase C isozymes (PLCβ3 and PLCγ1), and via cSrc (32, 36). Studies have shown that PAF challenge induced endothelial actin cytoskeletal rearrangement (37) and marked vascular leakiness (38); however, the signaling pathways have not been elucidated.Therefore, in the present study, we carried out a systematic analysis of PAF-induced morphological and biochemical changes of endothelial barrier in vivo and in cultured ECs. We found that the opening of endothelial barrier and the increased vascular leakiness induced by PAF are the result of a shift in actin pools without involvement of EC contraction, followed by a redistribution of tight junctional associated protein ZO-1 and adherens junctional protein VE-cadherin.  相似文献   

10.
The 26 S proteasome is a large proteolytic machine, which degrades most intracellular proteins. We found that thioredoxin, Txnl1/TRP32, binds to Rpn11, a subunit of the regulatory complex of the human 26 S proteasome. Txnl1 is abundant, metabolically stable, and widely expressed and is present in the cytoplasm and nucleus. Txnl1 has thioredoxin activity with a redox potential of about-250 mV. Mutant Txnl1 with one active site cysteine replaced by serine formed disulfide bonds to eEF1A1, a substrate-recruiting factor of the 26 S proteasome. eEF1A1 is therefore a likely physiological substrate. In response to knockdown of Txnl1, ubiquitin-protein conjugates were moderately stabilized. Hence, Txnl1 is the first example of a direct connection between protein reduction and proteolysis, two major intracellular protein quality control mechanisms.Degradation of proteins in eukaryotic cells plays a pivotal role in the regulation of several important processes, including cell division, antigen presentation, and signal transduction (1). Most intracellular proteins are degraded by the 26 S proteasome, a 2.5-MDa protease complex composed of more than 30 different subunits (2).To become degraded, proteins are typically first conjugated to a chain of ubiquitin moieties. This reaction is catalyzed by ubiquitin ligases. The ubiquitin chains lend the proteins affinity for the 26 S proteasome (3). For efficient degradation, certain ubiquitylated proteins are shuttled to the 26 S proteasome by substrate recruiting factors, such as Rad23, Dsk2, and eEF1A (4, 5).The 26 S proteasome is composed of two stable subcomplexes, the proteolytically active 20 S core and 19 S regulatory complexes, which bind to one or both ends of the cylindrical 20 S core particle (6). The regulatory complexes first recognize the ubiquitylated substrates (3), before the substrates are deubiquitylated (7, 8), unfolded (9, 10), and translocated into the 20 S particle for degradation.Although the 26 S proteasome has been known for more than 20 years (11), novel subunits and cofactors have been described recently (12, 13). Here we report another novel proteasome-associated protein, Txnl1 (thioredoxin-like protein 1), that associates directly with the proteasome subunit Rpn11. Txnl1 exhibits thioredoxin activity and targets eEF1A1 in vivo. Previous reports have shown that eEF1A1 transfers misfolded nascent proteins from the ribosome to the 26 S proteasome for degradation (5, 14, 15). Accordingly, ubiquitin-protein conjugates were stabilized upon knockdown of Txnl1 expression. Txnl1 therefore directly links protein reduction and proteolysis, two major intracellular protein quality control mechanisms.  相似文献   

11.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

12.
13.
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   

14.
15.
Macroautophagy is a vacuolar lysosomal catabolic pathway that is stimulated during periods of nutrient starvation to preserve cell integrity. Ceramide is a bioactive sphingolipid associated with a large range of cell processes. Here we show that short-chain ceramides (C2-ceramide and C6-ceramide) and stimulation of the de novo ceramide synthesis by tamoxifen induce the dissociation of the complex formed between the autophagy protein Beclin 1 and the anti-apoptotic protein Bcl-2. This dissociation is required for macroautophagy to be induced either in response to ceramide or to starvation. Three potential phosphorylation sites, Thr69, Ser70, and Ser87, located in the non-structural N-terminal loop of Bcl-2, play major roles in the dissociation of Bcl-2 from Beclin 1. We further show that activation of c-Jun N-terminal protein kinase 1 by ceramide is required both to phosphorylate Bcl-2 and to stimulate macroautophagy. These findings reveal a new aspect of sphingolipid signaling in up-regulating a major cell process involved in cell adaptation to stress.Macroautophagy (referred to below as “autophagy”) is a vacuolar, lysosomal degradation pathway for cytoplasmic constituents that is conserved in eukaryotic cells (13). Autophagy is initiated by the formation of a multimembrane-bound autophagosome that engulfs cytoplasmic proteins and organelles. The last stage in the process results in fusion with the lysosomal compartments, where the autophagic cargo undergoes degradation. Basal autophagy is important in controlling the quality of the cytoplasm by removing damaged organelles and protein aggregates. Inhibition of basal autophagy in the brain is deleterious, and leads to neurodegeneration in mouse models (4, 5). Stimulation of autophagy during periods of nutrient starvation is a physiological response present at birth and has been shown to provide energy in various tissues of newborn pups (6). In cultured cells, starvation-induced autophagy is an autonomous cell survival mechanism, which provides nutrients to maintain a metabolic rate and level of ATP compatible with cell survival (7). In addition, starvation-induced autophagy blocks the induction of apoptosis (8). In other contexts, such as drug treatment and a hypoxic environment, autophagy has also been shown to be cytoprotective in cancer cells (9, 10). However, autophagy is also part of cell death pathways in certain situations (11). Autophagy can be a player in apoptosis-independent type-2 cell death (type-1 cell death is apoptosis), also known as autophagic cell death. This situation has been shown to occur when the apoptotic machinery is crippled in mammalian cells (12, 13). Autophagy can also be part of the apoptotic program, for instance in tumor necrosis factor-α-induced cell death when NF-κB is inhibited (14), or in human immunodeficiency virus envelope-mediated cell death in bystander naive CD4 T cells (15). Moreover autophagy has recently been shown to be required for the externalization of phosphatidylserine, the eat-me signal for phagocytic cells, at the surface of apoptotic cells (16).The complex relationship between autophagy and apoptosis reflects the intertwined regulation of these processes (17, 18). Many signaling pathways involved in the regulation of autophagy also regulate apoptosis. This intertwining has recently been shown to occur at the level of the molecular machinery of autophagy. In fact the anti-apoptotic protein Bcl-2 has been shown to inhibit starvation-induced autophagy by interacting with the autophagy protein Beclin 1 (19). Beclin 1 is one of the Atg proteins conserved from yeast to humans (it is the mammalian orthologue of yeast Atg6) and is involved in autophagosome formation (20). Beclin 1 is a platform protein that interacts with several different partners, including hVps34 (class III phosphatidylinositol 3-kinase), which is responsible for the synthesis of phosphatidylinositol 3-phosphate. The production of this lipid is important for events associated with the nucleation of the isolation membrane before it elongates and closes to form autophagosomes in response to other Atg proteins, including the Atg12 and LC32 (microtubule-associated protein light chain 3 is the mammalian orthologue of the yeast Atg8) ubiquitin-like conjugation systems (3, 21). Various partners associated with the Beclin 1 complex modulate the activity of hVps34. For instance, Bcl-2 inhibits the activity of this enzyme, whereas UVRAG, Ambra-1, and Bif-1 all up-regulate it (22, 23).In view of the intertwining between autophagy and apoptosis, it is noteworthy that Beclin 1 belongs to the BH3-only family of proteins (2426). However, and unlike most of the proteins in this family, Beclin 1 is not able to trigger apoptosis when its expression is forced in cells (27). A BH3-mimetic drug, ABT-737, is able to dissociate the Beclin 1-Bcl-2 complex, and to trigger autophagy by mirroring the effect of starvation (25).The sphingolipids constitute a family of bioactive lipids (2832) of which several members, such as ceramide and sphingosine 1-phosphate, are signaling molecules. These molecules constitute a “sphingolipid rheostat” that determines the fate of the cell, because in many settings ceramide is pro-apoptotic and sphingosine 1-phosphate mitigates this apoptotic effect (31, 32). However, ceramide is also engaged in a wide variety of other cell processes, such as the formation of exosomes (33), differentiation, cell proliferation, and senescence (34). Recently we showed that both ceramide and sphingosine 1-phosphate are able to stimulate autophagy (35, 36). It has also been shown that ceramide triggers autophagy in a large panel of mammalian cells (3739). However, elucidation of the mechanism by which ceramide stimulates autophagy is still in its infancy. We have previously demonstrated that ceramide induces autophagy in breast and colon cancer cells by inhibiting the Class I phosphatidylinositol 3-phosphate/mTOR signaling pathway, which plays a central role in inhibiting autophagy (36). Inhibition of mTOR is another hallmark of starvation-induced autophagy (17). This finding led us to investigate the effect of ceramide on the Beclin 1-Bcl-2 complex. The results presented here show that ceramide is more potent than starvation in dissociating the Beclin 1-Bcl-2 complex (see Ref. 40). This dissociation is dependent on three phosphorylation sites (Thr69, Ser70, and Ser87) located in a non-structural loop of Bcl-2. Ceramide induces the c-Jun N-terminal kinase 1-dependent phosphorylation of Bcl-2. Expression of a dominant negative form of JNK1 blocks Bcl-2 phosphorylation, and thus the induction of autophagy by ceramide. These findings help to explain how autophagy is regulated by a major lipid second messenger.  相似文献   

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.
Galectins have been implicated in T cell homeostasis playing complementary pro-apoptotic roles. Here we show that galectin-8 (Gal-8) is a potent pro-apoptotic agent in Jurkat T cells inducing a complex phospholipase D/phosphatidic acid signaling pathway that has not been reported for any galectin before. Gal-8 increases phosphatidic signaling, which enhances the activity of both ERK1/2 and type 4 phosphodiesterases (PDE4), with a subsequent decrease in basal protein kinase A activity. Strikingly, rolipram inhibition of PDE4 decreases ERK1/2 activity. Thus Gal-8-induced PDE4 activation releases a negative influence of cAMP/protein kinase A on ERK1/2. The resulting strong ERK1/2 activation leads to expression of the death factor Fas ligand and caspase-mediated apoptosis. Several conditions that decrease ERK1/2 activity also decrease apoptosis, such as anti-Fas ligand blocking antibodies. In addition, experiments with freshly isolated human peripheral blood mononuclear cells, previously stimulated with anti-CD3 and anti-CD28, show that Gal-8 is pro-apoptotic on activated T cells, most likely on a subpopulation of them. Anti-Gal-8 autoantibodies from patients with systemic lupus erythematosus block the apoptotic effect of Gal-8. These results implicate Gal-8 as a novel T cell suppressive factor, which can be counterbalanced by function-blocking autoantibodies in autoimmunity.Glycan-binding proteins of the galectin family have been increasingly studied as regulators of the immune response and potential therapeutic agents for autoimmune disorders (1). To date, 15 galectins have been identified and classified according with the structural organization of their distinctive monomeric or dimeric carbohydrate recognition domain for β-galactosides (2, 3). Galectins are secreted by unconventional mechanisms and once outside the cells bind to and cross-link multiple glycoconjugates both at the cell surface and at the extracellular matrix, modulating processes as diverse as cell adhesion, migration, proliferation, differentiation, and apoptosis (410). Several galectins have been involved in T cell homeostasis because of their capability to kill thymocytes, activated T cells, and T cell lines (1116). Pro-apoptotic galectins might contribute to shape the T cell repertoire in the thymus by negative selection, restrict the immune response by eliminating activated T cells at the periphery (1), and help cancer cells to escape the immune system by eliminating cancer-infiltrating T cells (17). They have also a promising therapeutic potential to eliminate abnormally activated T cells and inflammatory cells (1). Studies on the mostly explored galectins, Gal-1, -3, and -9 (14, 15, 1820), as well as in Gal-2 (13), suggest immunosuppressive complementary roles inducing different pathways to apoptosis. Galectin-8 (Gal-8)4 is one of the most widely expressed galectins in human tissues (21, 22) and cancerous cells (23, 24). Depending on the cell context and mode of presentation, either as soluble stimulus or extracellular matrix, Gal-8 can promote cell adhesion, spreading, growth, and apoptosis (6, 7, 9, 10, 22, 25). Its role has been mostly studied in relation to tumor malignancy (23, 24). However, there is some evidence regarding a role for Gal-8 in T cell homeostasis and autoimmune or inflammatory disorders. For instance, the intrathymic expression and pro-apoptotic effect of Gal-8 upon CD4highCD8high thymocytes suggest a role for Gal-8 in shaping the T cell repertoire (16). Gal-8 could also modulate the inflammatory function of neutrophils (26), Moreover Gal-8-blocking agents have been detected in chronic autoimmune disorders (10, 27, 28). In rheumatoid arthritis, Gal-8 has an anti-inflammatory action, promoting apoptosis of synovial fluid cells, but can be counteracted by a specific rheumatoid version of CD44 (CD44vRA) (27). In systemic lupus erythematosus (SLE), a prototypic autoimmune disease, we recently described function-blocking autoantibodies against Gal-8 (10, 28). Thus it is important to define the role of Gal-8 and the influence of anti-Gal-8 autoantibodies in immune cells.In Jurkat T cells, we previously reported that Gal-8 interacts with specific integrins, such as α1β1, α3β1, and α5β1 but not α4β1, and as a matrix protein promotes cell adhesion and asymmetric spreading through activation of the extracellular signal-regulated kinases 1 and 2 (ERK1/2) (10). These early effects occur within 5–30 min. However, ERK1/2 signaling supports long term processes such as T cell survival or death, depending on the moment of the immune response. During T cell activation, ERK1/2 contributes to enhance the expression of interleukin-2 (IL-2) required for T cell clonal expansion (29). It also supports T cell survival against pro-apoptotic Fas ligand (FasL) produced by themselves and by other previously activated T cells (30, 31). Later on, ERK1/2 is required for activation-induced cell death, which controls the extension of the immune response by eliminating recently activated and restimulated T cells (32, 33). In activation-induced cell death, ERK1/2 signaling contributes to enhance the expression of FasL and its receptor Fas/CD95 (32, 33), which constitute a preponderant pro-apoptotic system in T cells (34). Here, we ask whether Gal-8 is able to modulate the intensity of ERK1/2 signaling enough to participate in long term processes involved in T cell homeostasis.The functional integration of ERK1/2 and PKA signaling (35) deserves special attention. cAMP/PKA signaling plays an immunosuppressive role in T cells (36) and is altered in SLE (37). Phosphodiesterases (PDEs) that degrade cAMP release the immunosuppressive action of cAMP/PKA during T cell activation (38, 39). PKA has been described to control the activity of ERK1/2 either positively or negatively in different cells and processes (35). A little explored integration among ERK1/2 and PKA occurs via phosphatidic acid (PA) and PDE signaling. Several stimuli activate phospholipase D (PLD) that hydrolyzes phosphatidylcholine into PA and choline. Such PLD-generated PA plays roles in signaling interacting with a variety of targeting proteins that bear PA-binding domains (40). In this way PA recruits Raf-1 to the plasma membrane (41). It is also converted by phosphatidic acid phosphohydrolase (PAP) activity into diacylglycerol (DAG), which among other functions, recruits and activates the GTPase Ras (42). Both Ras and Raf-1 are upstream elements of the ERK1/2 activation pathway (43). In addition, PA binds to and activates PDEs of the type 4 subfamily (PDE4s) leading to decreased cAMP levels and PKA down-regulation (44). The regulation and role of PA-mediated control of ERK1/2 and PKA remain relatively unknown in T cell homeostasis, because it is also unknown whether galectins stimulate the PLD/PA pathway.Here we found that Gal-8 induces apoptosis in Jurkat T cells by triggering cross-talk between PKA and ERK1/2 pathways mediated by PLD-generated PA. Our results for the first time show that a galectin increases the PA levels, down-regulates the cAMP/PKA system by enhancing rolipram-sensitive PDE activity, and induces an ERK1/2-dependent expression of the pro-apoptotic factor FasL. The enhanced PDE activity induced by Gal-8 is required for the activation of ERK1/2 that finally leads to apoptosis. Gal-8 also induces apoptosis in human peripheral blood mononuclear cells (PBMC), especially after activating T cells with anti-CD3/CD28. Therefore, Gal-8 shares with other galectins the property of killing activated T cells contributing to the T cell homeostasis. The pathway involves a particularly integrated signaling context, engaging PLD/PA, cAMP/PKA, and ERK1/2, which so far has not been reported for galectins. The pro-apoptotic function of Gal-8 also seems to be unique in its susceptibility to inhibition by anti-Gal-8 autoantibodies.  相似文献   

20.
The Ca2+ release-activated Ca2+ channel is a principal regulator of intracellular Ca2+ rise, which conducts various biological functions, including immune responses. This channel, involved in store-operated Ca2+ influx, is believed to be composed of at least two major components. Orai1 has a putative channel pore and locates in the plasma membrane, and STIM1 is a sensor for luminal Ca2+ store depletion in the endoplasmic reticulum membrane. Here we have purified the FLAG-fused Orai1 protein, determined its tetrameric stoichiometry, and reconstructed its three-dimensional structure at 21-Å resolution from 3681 automatically selected particle images, taken with an electron microscope. This first structural depiction of a member of the Orai family shows an elongated teardrop-shape 150Å in height and 95Å in width. Antibody decoration and volume estimation from the amino acid sequence indicate that the widest transmembrane domain is located between the round extracellular domain and the tapered cytoplasmic domain. The cytoplasmic length of 100Å is sufficient for direct association with STIM1. Orifices close to the extracellular and intracellular membrane surfaces of Orai1 seem to connect outside the molecule to large internal cavities.Ca2+ is an intracellular second messenger that plays important roles in various physiological functions such as immune response, muscle contraction, neurotransmitter release, and cell proliferation. Intracellular Ca2+ is mainly stored in the endoplasmic reticulum (ER).2 This ER system is distributed through the cytoplasm from around the nucleus to the cell periphery close to the plasma membrane. In non-excitable cells, the ER releases Ca2+ through the inositol 1,4,5-trisphosphate (IP3) receptor channel in response to various signals, and the Ca2+ store is depleted. Depletion of Ca2+ then induces Ca2+ influx from outside the cell to help in refilling the Ca2+ stores and to continue Ca2+ rise for several minutes in the cytoplasm (1, 2). This Ca2+ influx was first proposed by Putney (3) and was named store-operated Ca2+ influx. In the immune system, store-operated Ca2+ influx is mainly mediated by the Ca2+ release-activated Ca2+ (CRAC) current, which is a highly Ca2+-selective inwardly rectified current with low conductance (4, 5). Pathologically, the loss of CRAC current in T cells causes severe combined immunodeficiency (6) where many Ca2+ signal-dependent gene expressions, including cytokines, are interrupted (7). Therefore, CRAC current is necessary for T cell functions.Recently, Orai1 (also called CRACM1) and STIM1 have been physiologically characterized as essential components of the CRAC channel (812). They are separately located in the plasma membrane and in the ER membrane; co-expression of these proteins presents heterologous CRAC-like currents in various types of cells (10, 1315). Both of them are shown to be expressed ubiquitously in various tissues (1618). STIM1 senses Ca2+ depletion in the ER through its EF hand motif (19) and transmits a signal to Orai1 in the plasma membrane. Although Orai1 is proposed as a regulatory component for some transient receptor potential canonical channels (20, 21), it is believed from the mutation analyses to be the pore-forming subunit of the CRAC channel (8, 2224). In the steady state, both Orai1 and STIM1 molecules are dispersed in each membrane. When store depletion occurs, STIM1 proteins gather into clusters to form puncta in the ER membrane near the plasma membrane (11, 19). These clusters then trigger the clustering of Orai1 in the plasma membrane sites opposite the puncta (25, 26), and CRAC channels are activated (27).Orai1 has two homologous genes, Orai2 and Orai3 (8). They form the Orai family and have in common the four transmembrane (TM) segments with relatively large N and C termini. These termini are demonstrated to be in the cytoplasm, because both N- and C-terminally introduced tags are immunologically detected only in the membrane-permeabilized cells (8, 9). The subunit stoichiometry of Orai1 is as yet controversial: it is believed to be an oligomer, presumably a dimer or tetramer even in the steady state (16, 2830).Despite the accumulation of biochemical and electrophysiological data, structural information about Orai1 is limited due to difficulties in purification and crystallization. In this study, we have purified Orai1 in its tetrameric form and have reconstructed the three-dimensional structure from negatively stained electron microscopic (EM) images.  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