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.
2.
3.
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   

4.
5.
The accumulation of mutant protein in intracellular aggregates is a common feature of neurodegenerative disease. In Huntington disease, mutant huntingtin leads to inclusion body (IB) formation and neuronal toxicity. Impairment of the ubiquitin-proteasome system (UPS) has been implicated in IB formation and Huntington disease pathogenesis. However, IBs form asynchronously in only a subset of cells with mutant huntingtin, and the relationship between IB formation and UPS function has been difficult to elucidate. Here, we applied single-cell longitudinal acquisition and analysis to monitor mutant huntingtin IB formation, UPS function, and neuronal toxicity. We found that proteasome inhibition is toxic to striatal neurons in a dose-dependent fashion. Before IB formation, the UPS is more impaired in neurons that go on to form IBs than in those that do not. After forming IBs, impairment is lower in neurons with IBs than in those without. These findings suggest IBs are a protective cellular response to mutant protein mediated in part by improving intracellular protein degradation.Huntington disease (HD)4 is a progressive incurable neurodegenerative disorder caused by the expansion of a polyglutamine (polyQ) stretch in the N-terminal end of the huntingtin (htt) protein above a threshold length of ∼36 (1). The deposition of polyQ-expanded aggregated mutant htt in inclusion bodies (IBs) is a hallmark of HD, and IBs are found in human post-mortem samples, transgenic mouse brain, and cell-culture models (2). The accumulation of ubiquitinated proteins in IBs has implicated the ubiquitin-proteasome system (UPS) in the pathogenesis of HD, amyotrophic lateral sclerosis, Parkinson disease, and polyQ-mediated disorders (3).The UPS is a major pathway of intracellular protein degradation. After a series of three reactions, each catalyzed by a different set of enzymes, ubiquitin, a 76-amino acid polypeptide, forms an isopeptide bond with the amino group of lysine residues on substrate proteins. Several lysine residues within ubiquitin are sites for more ubiquitin additions. Once a protein accumulates four or more ubiquitins, it is efficiently targeted to the proteasome for degradation. The proteasome binds polyubiquitinated substrates and hydrolyzes ubiquitin isopeptide bonds, releasing ubiquitin moieties before degrading substrate proteins through chymotrypsin-like, trypsin-like, and post-glutamyl peptidase activities (3).Increased polyubiquitin levels and changes in ubiquitin linkages accompany the accumulation of UPS substrates in the brains of HD patients and transgenic mice and in cellular HD models (4). UPS substrates accumulate throughout the cell in polyQ models, even before IB formation (5, 6). This has added to the confusion over whether polyQ expansion leads to toxicity through direct impairment of proteasomal degradation. Proteasomes have been reported to cleave polyQ stretches efficiently (7), inefficiently (8), or essentially not at all (9). In vivo, polyQ-dependent degeneration occurs with no detectable proteasome inhibition (10, 11) or is tightly linked to it (12, 13). The inability of some studies to detect UPS impairment in HD models may be due to the limited sensitivity of conventional approaches to identify cell-to-cell variations in UPS function.The relationship between IB formation and UPS function has been difficult to determine. Protein turnover in cells with IBs is evidently reduced and accompanied by the accumulation of cellular proteins (1416); HEK293 cells containing mutant htt IBs have a greater degree of UPS impairment than those without IBs (5). Proteasome subunits and heat shock proteins colocalize with IBs, but it is unclear if this colocalization facilitates protein delivery or unfolding at the mouth of active proteasomes, or if it harms proteasome function by sequestering essential cellular machinery (18). Some IBs are relatively static (8, 25), but the proteins in others are dynamically exchanged with cytoplasmic and nuclear pools (19, 20).UPS function is critical to cellular homeostasis. Deletion of one of the two inducible polyubiquitin genes in mice leads to lower intracellular ubiquitin levels in germ cells and hypothalamic neurons. These same populations undergo cell-cycle arrest and hypothalamic neurodegeneration, respectively (22, 23). Cell lines expressing mutant huntingtin accumulate ubiquitinated proteins and undergo cell-cycle arrest in G2/M (5). In neurons, UPS impairment may lead to cell death through an accumulation of signals for apoptosis, a decrease in NF-κB signaling, sensitization to other toxic stimuli, remodeling of synapses, retraction of neurites, or other unidentified mechanisms (24). The effect of UPS impairment depends on cell type and cell cycle, and the relationship between UPS impairment and striatal neuronal survival is largely unknown.Diffuse species of mutant htt induce IB formation and neuronal death in a protein concentration-dependent manner (2). IB formation delays neuronal death, suggesting that IB formation helps neurons cope with toxic diffuse mutant htt. Whether the effect of IB formation on survival is mediated through UPS function has been difficult to determine. IB formation and neuronal death occur asynchronously in overlapping but distinct subsets of neurons that express mutant htt. The observation that IB formation is not required for UPS impairment also complicates population analysis (6, 26).To explore this problem, we applied single-cell analysis. We tracked single neurons over their entire lifetimes, gaining spatial and temporal resolution while simultaneously monitoring IB formation, UPS inhibition, and neuronal toxicity.  相似文献   

6.
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   

7.
ATP-binding cassette (ABC) transporters transduce the free energy of ATP hydrolysis to power the mechanical work of substrate translocation across cell membranes. MsbA is an ABC transporter implicated in trafficking lipid A across the inner membrane of Escherichia coli. It has sequence similarity and overlapping substrate specificity with multidrug ABC transporters that export cytotoxic molecules in humans and prokaryotes. Despite rapid advances in structure determination of ABC efflux transporters, little is known regarding the location of substrate-binding sites in the transmembrane segment and the translocation pathway across the membrane. In this study, we have mapped residues proximal to the daunorubicin (DNR)-binding site in MsbA using site-specific, ATP-dependent quenching of DNR intrinsic fluorescence by spin labels. In the nucleotide-free MsbA intermediate, DNR-binding residues cluster at the cytoplasmic end of helices 3 and 6 at a site accessible from the membrane/water interface and extending into an aqueous chamber formed at the interface between the two transmembrane domains. Binding of a nonhydrolyzable ATP analog inverts the transporter to an outward-facing conformation and relieves DNR quenching by spin labels suggesting DNR exclusion from proximity to the spin labels. The simplest model consistent with our data has DNR entering near an elbow helix parallel to the water/membrane interface, partitioning into the open chamber, and then translocating toward the periplasm upon ATP binding.ATP-binding cassette (ABC)2 transporters transduce the energy of ATP hydrolysis to power the movement of a wide range of substrates across the cell membranes (1, 2). They constitute the largest family of prokaryotic transporters, import essential cell nutrients, flip lipids, and export toxic molecules (3). Forty eight human ABC transporters have been identified, including ABCB1, or P-glycoprotein, which is implicated in cross-resistance to drugs and cytotoxic molecules (4, 5). Inherited mutations in these proteins are linked to diseases such as cystic fibrosis, persistent hypoglycemia of infancy, and immune deficiency (6).The functional unit of an ABC transporter consists of four modules. Two highly conserved ABCs or nucleotide-binding domains (NBDs) bind and hydrolyze ATP to supply the active energy for transport (7). ABCs drive the mechanical work of proteins with diverse functions ranging from membrane transport to DNA repair (3, 5). Substrate specificity is determined by two transmembrane domains (TMDs) that also provide the translocation pathway across the bilayer (7). Bacterial ABC exporters are expressed as monomers, each consisting of one NBD and one TMD, that dimerize to form the active transporter (3). The number of transmembrane helices and their organization differ significantly between ABC importers and exporters reflecting the divergent structural and chemical nature of their substrates (1, 8, 9). Furthermore, ABC exporters bind substrates directly from the cytoplasm or bilayer inner leaflet and release them to the periplasm or bilayer outer leaflet (10, 11). In contrast, bacterial importers have their substrates delivered to the TMD by a dedicated high affinity substrate-binding protein (12).In Gram-negative bacteria, lipid A trafficking from its synthesis site on the inner membrane to its final destination in the outer membrane requires the ABC transporter MsbA (13). Although MsbA has not been directly shown to transport lipid A, suppression of MsbA activity leads to cytoplasmic accumulation of lipid A and inhibits bacterial growth strongly suggesting a role in translocation (14-16). In addition to this role in lipid A transport, MsbA shares sequence similarity with multidrug ABC transporters such as human ABCB1, LmrA of Lactococcus lactis, and Sav1866 of Staphylococcus aureus (16-19). ABCB1, a prototype of the ABC family, is a plasma membrane protein whose overexpression provides resistance to chemotherapeutic agents in cancer cells (1). LmrA and MsbA have overlapping substrate specificity with ABCB1 suggesting that both proteins can function as drug exporters (18, 20). Indeed, cells expressing MsbA confer resistance to erythromycin and ethidium bromide (21). MsbA can be photolabeled with the ABCB1/LmrA substrate azidopine and can transport Hoechst 33342 (H33342) across membrane vesicles in an energy-dependent manner (21).The structural mechanics of ABC exporters was revealed from comparison of the MsbA crystal structures in the apo- and nucleotide-bound states as well as from analysis by spin labeling EPR spectroscopy in liposomes (17, 19, 22, 23). The energy harnessed from ATP binding and hydrolysis drives a cycle of NBD association and dissociation that is transmitted to induce reorientation of the TMD from an inward- to outward-facing conformation (17, 19, 22). Large amplitude motion closes the cytoplasmic end of a chamber found at the interface between the two TMDs and opens it to the periplasm (23). These rearrangements lead to significant changes in chamber hydration, which may drive substrate translocation (22).Substrate binding must precede energy input, otherwise the cycle is futile, wasting the energy of ATP hydrolysis without substrate extrusion (7). Consistent with this model, ATP binding reduces ABCB1 substrate affinity, potentially through binding site occlusion (24-26). Furthermore, the TMD substrate-binding event signals the NBD to stimulate ATP hydrolysis increasing transport efficiency (1, 27, 28). However, there is a paucity of information regarding the location of substrate binding, the transport pathway, and the structural basis of substrate recognition by ABC exporters. In vitro studies of MsbA substrate specificity identify a broad range of substrates that stimulate ATPase activity (29). In addition to the putative physiological substrates lipid A and lipopolysaccharide (LPS), the ABCB1 substrates Ilmofosine, H33342, and verapamil differentially enhance ATP hydrolysis of MsbA (29, 30). Intrinsic MsbA tryptophan (Trp) fluorescence quenching by these putative substrate molecules provides further support of interaction (29).Extensive biochemical analysis of ABCB1 and LmrA provides a general model of substrate binding to ABC efflux exporters. This so-called “hydrophobic cleaner model” describes substrates binding from the inner leaflet of the bilayer and then translocating through the TMD (10, 31, 32). These studies also identified a large number of residues involved in substrate binding and selectivity (33). When these crucial residues are mapped onto the crystal structures of MsbA, a subset of homologous residues clusters to helices 3 and 6 lining the putative substrate pathway (34). Consistent with a role in substrate binding and specificity, simultaneous replacement of two serines (Ser-289 and Ser-290) in helix 6 of MsbA reduces binding and transport of ethidium and taxol, although H33342 and erythromycin interactions remain unaffected (34).The tendency of lipophilic substrates to partition into membranes confounds direct analysis of substrate interactions with ABC exporters (35, 36). Such partitioning may promote dynamic collisions with exposed Trp residues and nonspecific cross-linking in photo-affinity labeling experiments. In this study, we utilize a site-specific quenching approach to identify residues in the vicinity of the daunorubicin (DNR)-binding site (37). Although the data on DNR stimulation of ATP hydrolysis is inconclusive (20, 29, 30), the quenching of MsbA Trp fluorescence suggests a specific interaction. Spin labels were introduced along transmembrane helices 3, 4, and 6 of MsbA to assess their ATP-dependent quenching of DNR fluorescence. Residues that quench DNR cluster along the cytoplasmic end of helices 3 and 6 consistent with specific binding of DNR. Furthermore, many of these residues are not lipid-exposed but face the putative substrate chamber formed between the two TMDs. These residues are proximal to two Trps, which likely explains the previously reported quenching (29). Our results suggest DNR partitions to the membrane and then binds MsbA in a manner consistent with the hydrophobic cleaner model. Interpretation in the context of the crystal structures of MsbA identifies a putative translocation pathway through the transmembrane segment.  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 has recently been shown that the epithelial Na+ channel (ENaC) is compartmentalized in caveolin-rich lipid rafts and that pharmacological depletion of membrane cholesterol, which disrupts lipid raft formation, decreases the activity of ENaC. Here we show, for the first time, that a signature protein of caveolae, caveolin-1 (Cav-1), down-regulates the activity and membrane surface expression of ENaC. Physical interaction between ENaC and Cav-1 was also confirmed in a coimmunoprecipitation assay. We found that the effect of Cav-1 on ENaC requires the activity of Nedd4-2, a ubiquitin protein ligase of the Nedd4 family, which is known to induce ubiquitination and internalization of ENaC. The effect of Cav-1 on ENaC requires the proline-rich motifs at the C termini of the β- and γ-subunits of ENaC, the binding motifs that mediate interaction with Nedd4-2. Taken together, our data suggest that Cav-1 inhibits the activity of ENaC by decreasing expression of ENaC at the cell membrane via a mechanism that involves the promotion of Nedd4-2-dependent internalization of the channel.Amiloride-sensitive epithelial Na+ channels (ENaC)3 are membrane proteins that are expressed in salt-absorptive epithelia, including the distal collecting tubules of the kidney, the mucosa of the distal colon, the respiratory epithelium, and the excretory ducts of sweat and salivary glands (14). Na+ absorption via ENaC is critical to the normal regulation of Na+ and fluid homeostasis and is important for maintaining blood pressure (5) and the volume of fluid in the respiratory passages (6). Increased ENaC activity has been implicated in the salt-sensitive inherited form of hypertension, Liddle''s syndrome (7), and dehydration of the surface of the airway epithelium in the pathology associated with cystic fibrosis lung disease (8).Expression of ENaC at the cell membrane surface is regulated by the E3 ubiquitin protein ligase, Nedd4-2 (neural precursor cell expressed developmentally down-regulated protein 4) (9). Interaction between the WW domains of Nedd4-2 and the proline-rich PY motifs (PPPXY) on ENaC is essential for Nedd4-2 to exert a negative effect on the channel (10, 11). This interaction leads to ubiquitination-dependent internalization of ENaC (12, 13). Several regulators of ENaC exert their effects on the channel by modulating the action of Nedd4-2. For instance, serum and glucocorticoid-dependent protein kinase (14), protein kinase B (15), and G protein-coupled receptor kinase (16) up-regulate activity of ENaC by inhibiting Nedd4-2. Although the details of cellular mechanisms that underlie internalization of ENaC remain to be elucidated, the physiological significance of Nedd4-dependent internalization of the channel has been well established. For instance, heritable mutations that delete the cytosolic termini of the β-or γ-subunit of ENaC, which contain the proline-rich motifs, are known to cause hyperactivity of ENaC in the kidney (17) and increase cell surface expression of the channel (7, 18).The plasma membranes of most cell types contain lipid raft microdomains that are enriched with glycosphingolipid and cholesterol (19), that have distinctive biophysical properties, and that selectively include or exclude signaling molecules (20). These microdomains promote clustering of an array of integral membrane proteins in the membrane leaflets (21) and may be important for organizing cascades of signaling molecules (22, 23). Processes in which raft microdomains are involved include the intracellular transport of proteins and lipids to the cell membrane (24), the endocytotic retrieval of membrane proteins (25, 26), and signal transduction (27, 28). In addition, segregation of signaling molecules within lipid rafts may facilitate cross-talk between signal transduction pathways (29), a phenomenon that may be important in ensuring rapid and efficient integration of multiple cellular signaling events (30, 31). Of particular interest is the subpopulation of lipid rafts enriched with caveolin proteins. Caveolin-1 (Cav-1), a major caveolin isoform expressed in nonmuscle cells, has been identified as being involved in diverse cellular functions, such as vesicular transport, cholesterol homeostasis, and signal transduction (32). Cav-1 also regulates the activity and membrane expression of ion channels and transporters (28).In epithelia, the majority of lipid rafts exist at the apical membrane surface (22). Pools of ENaC (3336) and several proteins that regulate activity of ENaC, such as Nedd4 (37), protein kinase B (38), protein kinase C (39), Go (40), and the G protein-coupled receptor kinase (41), have been identified in detergent-insoluble and cholesterol-rich membrane fractions from a variety of cell types, consistent with localization of these proteins in lipid rafts. Furthermore, detergent-free buoyant density separation of lipid rafts has revealed the presence of Cav-1 with ENaC in the lipid raft-rich membrane fraction (35). The physiological role of lipid rafts in the regulation of ENaC has been the subject of many recent investigations. Most of these studies used a pharmacological agent, methyl-β-cyclodextrin (MβCD), to promote redistribution of proteins away from the cholesterol-enriched membrane domains. The results were, however, inconclusive. In some studies, MβCD treatment was found to inhibit open probability (42) or cell surface expression of ENaC (35), whereas others found no direct effect of MβCD on the channel (33, 43).Despite a number of studies into the role of lipid rafts on the regulation of ENaC, little is known about the physiological relevance of caveolins to the function of this ion channel. In the present study, we use gene interference and gene expression techniques to determine the role of Cav-1 in the regulation of ENaC activity. We provide evidence of the association of Cav-1 with ENaC and evidence that Cav-1 negatively regulates both activity and abundance of ENaC at the surface of epithelial cells. Importantly, we demonstrate, for the first time, that the mechanism by which Cav-1 regulates activity of ENaC involves the E3 ubiquitin protein ligase, Nedd4-2.  相似文献   

10.
11.
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   

12.
The chemokine receptor CXCR4 plays important roles in the immune and nervous systems. Abnormal expression of CXCR4 contributes to cancer and inflammatory and neurodegenerative disorders. Although ligand-dependent CXCR4 ubiquitination is known to accelerate CXCR4 degradation, little is known about counter mechanisms for receptor deubiquitination. CXCL12, a CXCR4 agonist, induces a time-dependent association of USP14 with CXCR4, or its C terminus, that is not mimicked by USP2A, USP4, or USP7, other members of the deubiquitination catalytic family. Co-localization of CXCR4 and USP14 also is time-dependent following CXCL12 stimulation. The physical interaction of CXCR4 and USP14 is paralleled by USP14-catalyzed deubiquitination of the receptor; knockdown of endogenous USP14 by RNA interference (RNAi) blocks CXCR4 deubiquitination, whereas overexpression of USP14 promotes CXCR4 deubiquitination. We also observed that ubiquitination of CXCR4 facilitated receptor degradation, whereas overexpression of USP14 or RNAi-induced knockdown of USP14 blocked CXCL12-mediated CXCR4 degradation. Most interestingly, CXCR4-mediated chemotactic cell migration was blocked by either overexpression or RNAi-mediated knockdown of USP14, implying that a CXCR4-ubiquitin cycle on the receptor, rather than a particular ubiquitinated state of the receptor, is critical for the ligand gradient sensing and directed motility required for chemokine-mediated chemotaxis. Our observation that a mutant of CXCR4, HA-3K/R CXCR4, which cannot be ubiquitinated and does not mediate a chemotactic response to CXCL12, indicates the importance of this covalent modification not only in marking receptors for degradation but also for permitting CXCR4-mediated signaling. Finally, the indistinguishable activation of ERK by wild typeor 3K/R-CXCR4 suggests that chemotaxis in response to CXCL12 may be independent of the ERK cascade.The CXCR4 (CXC chemokine receptor 4) is a member of the chemokine receptor family, which belongs to the superfamily of G protein-coupled receptors (GPCRs)2 (1). Its ligand, CXCL12, also known as SDF-1α, also binds to RDC1, another chemokine receptor that is being proposed to be renamed as CXCR7 (2). CXCR4 mediates CXCL12-induced migration of peripheral blood lymphocytes (3), CD34+ progenitor cells (4), and pre- and pro-B cell lines (5). CXCR4 also plays an important role in the development of the immune system, because mouse embryos lacking either expression of the CXCR4 receptor or of its CXCL12 ligand are embryonic lethal and also manifest abnormalities in B cell lymphopoiesis and bone marrow myelopoiesis (3, 6, 7). The altered cerebellar neuron migration in mice null for the CXCR4 receptor also suggests a role for this receptor in central nervous system development. Abnormal expression and/or function of CXCR4 have been implicated in a number of diseases, including human immunodeficiency virus infection (8), cardiovascular disease (9), allergic inflammatory disease (10), neuroinflammation (11), neurodegenerative diseases (12, 13), and cancers (14-24).Stimulation of CXCR4 triggers various intracellular signaling cascades (1, 14, 25-27), such as extracellular signal-regulated kinase (ERK), which likely contribute to CXCR4-induced cell proliferation, differentiation, and/or migration. Ligand stimulation of CXCR4 also induces endocytosis of these receptors, which are targeted to lysosomes for degradation through a pathway involving ubiquitination of the C-terminal lysine residues (28). CXCR4 ubiquitination can be catalyzed by a member of the HECT family of E3 ligases, AIP4 (atrophin-interacting protein 4) (29, 30). The ubiquitinated CXCR4 is delivered to the endosomal compartments via a regulated pathway involving several adaptor proteins (31).It has been noted that deubiquitination also regulates the fate and function of ubiquitin-conjugated proteins. Deubiquitinating enzymes, which catalyze the removal of ubiquitin from ubiquitin-conjugated proteins, represent the largest family of enzymes in the ubiquitin system, implying the possibility that substrate selectivity is even greater for these enzymes than for those that catalyze ubiquitin ligation. Little is known about the mechanisms of CXCR4 deubiquitination and their regulation by receptor ligands. A proteomics study revealed that the steady state level of USP14 was increased upon CXCL12 stimulation of target cells (32), and preliminary studies revealed that ligand stimulation led to enhanced association of USP14 with the CXCR4. The present studies were undertaken to ascertain the functional consequences of this interaction, the selectivity of CXCR4 for USP14, when compared with three other deubiquitinating enzymes, USP2a, USP4, and USP7, and the impact of modifying the ubiquitinated state of the receptor on CXCR4 turnover, CXCL12-evoked chemotaxis, and CXCL12-induced activation of ERK.  相似文献   

13.
Rapid protein kinase D (PKD) activation and phosphorylation via protein kinase C (PKC) have been extensively documented in many cell types cells stimulated by multiple stimuli. In contrast, little is known about the role and mechanism(s) of a recently identified sustained phase of PKD activation in response to G protein-coupled receptor agonists. To elucidate the role of biphasic PKD activation, we used Swiss 3T3 cells because PKD expression in these cells potently enhanced duration of ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. Cell treatment with the preferential PKC inhibitors GF109203X or Gö6983 profoundly inhibited PKD activation induced by bombesin stimulation for <15 min but did not prevent PKD catalytic activation induced by bombesin stimulation for longer times (>60 min). The existence of sequential PKC-dependent and PKC-independent PKD activation was demonstrated in 3T3 cells stimulated with various concentrations of bombesin (0.3–10 nm) or with vasopressin, a different Gq-coupled receptor agonist. To gain insight into the mechanisms involved, we determined the phosphorylation state of the activation loop residues Ser744 and Ser748. Transphosphorylation targeted Ser744, whereas autophosphorylation was the predominant mechanism for Ser748 in cells stimulated with Gq-coupled receptor agonists. We next determined which phase of PKD activation is responsible for promoting enhanced ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. We show, for the first time, that the PKC-independent phase of PKD activation mediates prolonged ERK signaling and progression to DNA synthesis in response to bombesin or vasopressin through a pathway that requires epidermal growth factor receptor-tyrosine kinase activity. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.The understanding of the mechanisms that control cell proliferation requires the identification of the molecular pathways that govern the transition of quiescent cells into the S phase of the cell cycle. In this context the activation and phosphorylation of protein kinase D (PKD),4 the founding member of a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase (CAMK) group and separate from the previously identified PKCs (for review, see Ref. 1), are attracting intense attention. In unstimulated cells, PKD is in a state of low catalytic (kinase) activity maintained by autoinhibition mediated by the N-terminal domain, a region containing a repeat of cysteinerich zinc finger-like motifs and a pleckstrin homology (PH) domain (14). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (57). In response to cellular stimuli (1), including phorbol esters, growth factors (e.g. PDGF), and G protein-coupled receptor (GPCR) agonists (6, 816) that signal through Gq, G12, Gi, and Rho (11, 1519), PKD is converted into a form with high catalytic activity, as shown by in vitro kinase assays performed in the absence of lipid co-activators (5, 20).During these studies multiple lines of evidence indicated that PKC activity is necessary for rapid PKD activation within intact cells. For example, rapid PKD activation was selectively and potently blocked by cell treatment with preferential PKC inhibitors (e.g. GF109203X or Gö6983) that do not directly inhibit PKD catalytic activity (5, 20), implying that PKD activation in intact cells is mediated directly or indirectly through PKCs. Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade induced by multiple GPCR agonists and other receptor ligands in a range of cell types (for review, see Ref. 1). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (1, 4, 7, 17, 21). Collectively, these findings demonstrated the existence of a rapidly activated PKC-PKD protein kinase cascade(s). In a recent study we found that the rapid PKC-dependent PKD activation was followed by a late, PKC-independent phase of catalytic activation and phosphorylation induced by stimulation of the bombesin Gq-coupled receptor ectopically expressed in COS-7 cells (22). This study raised the possibility that PKD mediates rapid biological responses downstream of PKCs, whereas, in striking contrast, PKD could mediate long term responses through PKC-independent pathways. Despite its potential importance for defining the role of PKC and PKD in signal transduction, this hypothesis has not been tested in any cell type.Accumulating evidence demonstrates that PKD plays an important role in several cellular processes and activities, including signal transduction (14, 2325), chromatin organization (26), Golgi function (27, 28), gene expression (2931), immune regulation (26), and cell survival, adhesion, motility, differentiation, DNA synthesis, and proliferation (for review, see Ref. 1). In Swiss 3T3 fibroblasts, a cell line used extensively as a model system to elucidate mechanisms of mitogenic signaling (3234), PKD expression potently enhances ERK activation, DNA synthesis, and cell proliferation induced by Gq-coupled receptor agonists (8, 14). Here, we used this model system to elucidate the role and mechanism(s) of biphasic PKD activation. First, we show that the Gq-coupled receptor agonists bombesin and vasopressin, in contrast to phorbol esters, specifically induce PKD activation through early PKC-dependent and late PKC-independent mechanisms in Swiss 3T3 cells. Subsequently, we demonstrate for the first time that the PKC-independent phase of PKD activation is responsible for promoting ERK signaling and progression to DNA synthesis through an epidermal growth factor receptor (EGFR)-dependent pathway. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.  相似文献   

14.
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   

15.
16.
Cell membranes predominantly consist of lamellar lipid bilayers. When studied in vitro, however, many membrane lipids can exhibit non-lamellar morphologies, often with cubic symmetries. An open issue is how lipid polymorphisms influence organelle and cell shape. Here, we used controlled dimerization of artificial membrane proteins in mammalian tissue culture cells to induce an expansion of the endoplasmic reticulum (ER) with cubic symmetry. Although this observation emphasizes ER architectural plasticity, we found that the changed ER membrane became sequestered into large autophagic vacuoles, positive for the autophagy protein LC3. Autophagy may be targeting irregular membrane shapes and/or aggregated protein. We suggest that membrane morphology can be controlled in cells.The observation that simple mixtures of amphiphilic (polar) lipids and water yield a rich flora of phase structures has opened a long-standing debate as to whether such membrane polymorphisms are relevant for living organisms (17). Lipid bilayers with planar geometry, termed lamellar symmetry, dominate the membrane structure of cells. However, this architecture comprises only a fraction of the structures seen with in vitro lipid-water systems (711). The propensity to form lamellar bilayers (a property exclusive to cylindrically shaped lipids) is flanked by a continuum of lipid structures that occur in a number of exotic and probably non-physiological non-bilayer configurations (3, 12). However, certain lipids, particularly those with smaller head groups and more bulky hydrocarbon chains, can adopt bilayered non-lamellar phases called cubic phases. Here the bilayer is curved everywhere in the form of saddle shapes corresponding to an energetically favorable minimal surface of zero mean curvature (1, 7). Because a substantial number of the lipids present in biological membranes, when studied as individual pure lipids, form cubic phases (13), cubic membranes have received particular interest in cell biology.Since the application of electron microscopy (EM)3 to the study of cell ultrastructure, unusual membrane morphologies have been reported for virtually every organelle (14, 15). However, interpretation of three-dimensional structures from two-dimensional electron micrographs is not easy (16). In seminal work, Landh (17) developed the method of direct template correlative matching, a technique that unequivocally assesses the presence of cubic membranes in biological specimens (16). Cubic phases adopt mathematically well defined three-dimensional configurations whose two-dimensional analogs have been derived (4, 17). In direct template correlative matching, electron micrographs are matched to these analogs. Cubic cell membrane geometries and in vitro cubic phases of purified lipid mixtures do differ in their lattice parameters; however, such deviations are thought to relate to differences in water activity and lipid to protein ratios (10, 14, 18). Direct template correlative matching has revealed thousands of examples of cellular cubic membranes in a broad survey of electron micrographs ranging from protozoa to human cells (14, 17) and, more recently, in the mitochondria of amoeba (19) and in subcellular membrane compartments associated with severe acute respiratory syndrome virus (20). Analysis of cellular cubic membranes has also been furthered by the development of EM tomography that confirmed the presence of cubic bilayers in the mitochondrial membranes of amoeba (21, 22).Although it is now clear that cubic membranes can exist in living cells, the generation of such architecture would appear tightly regulated, as evidenced by the dominance of lamellar bilayers in biology. In this light, we examined the capability and implications of generating cubic membranes in the endoplasmic reticulum (ER) of mammalian tissue culture cells. The ER is a spatially interconnected complex consisting of two domains, the nuclear envelope and the peripheral ER (2326). The nuclear envelope surrounds the nucleus and is composed of two continuous sheets of membranes, an inner and outer nuclear membrane connected to each other at nuclear pores. The peripheral ER constitutes a network of branching trijunctional tubules that are continuous with membrane sheet regions that occur in closer proximity to the nucleus. Recently it has been suggested that the classical morphological definition of rough ER (ribosome-studded) and smooth ER (ribosome-free) may correspond to sheet-like and tubular ER domains, respectively (27). The ER has a strong potential for cubic architectures, as demonstrated by the fact that the majority of cubic cell membranes in the EM record come from ER-derived structures (14, 17). Furthermore, ER cubic symmetries are an inducible class of organized smooth ER (OSER), a definition collectively referring to ordered smooth ER membranes (=stacked cisternae on the outer nuclear membrane, also called Karmelle (2830), packed sinusoidal ER (31), concentric membrane whorls (30, 3234), and arrays of crystalloid ER (3537)). Specifically, weak homotypic interactions between membrane proteins produce both a whorled and a sinusoidal OSER phenotype (38), the latter exhibiting a cubic symmetry (16, 39).We were able to produce OSER with cubic membrane morphology via induction of homo-dimerization of artificial membrane proteins. Interestingly, the resultant cubic membrane architecture was removed from the ER system by incorporation into large autophagic vacuoles. To assess whether these cubic symmetries were favored in the absence of cellular energy, we depleted ATP. To our surprise, the cells responded by forming large domains of tubulated membrane, suggesting that a cubic symmetry was not the preferred conformation of the system. Our results suggest that whereas the endoplasmic reticulum is capable of adopting cubic symmetries, both the inherent properties of the ER system and active cellular mechanisms, such as autophagy, can tightly control their appearance.  相似文献   

17.
18.
19.
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   

20.
Cytokinesis in bacteria depends upon the contractile Z ring, which is composed of dynamic polymers of the tubulin homolog FtsZ as well as other membrane-associated proteins such as FtsA, a homolog of actin that is required for membrane attachment of the Z ring and its subsequent constriction. Here we show that a previously characterized hypermorphic mutant FtsA (FtsA*) partially disassembled FtsZ polymers in vitro. This effect was strictly dependent on ATP or ADP binding to FtsA* and occurred at substoichiometric levels relative to FtsZ, similar to cellular levels. Nucleotide-bound FtsA* did not affect FtsZ GTPase activity or the critical concentration for FtsZ assembly but was able to disassemble preformed FtsZ polymers, suggesting that FtsA* acts on FtsZ polymers. Microscopic examination of the inhibited FtsZ polymers revealed a transition from long, straight polymers and polymer bundles to mainly short, curved protofilaments. These results indicate that a bacterial actin, when activated by adenine nucleotides, can modify the length distribution of bacterial tubulin polymers, analogous to the effects of actin-depolymerizing factor/cofilin on F-actin.Bacterial cell division requires a large number of proteins that colocalize to form a putative protein machine at the cell membrane (1). This machine, sometimes called the divisome, recruits enzymes to synthesize the septum cell wall and to initiate and coordinate the invagination of the cytoplasmic membrane (and in Gram-negative bacteria, the outer membrane). The most widely conserved and key protein for this process is FtsZ, a homolog of tubulin that forms a ring structure called the Z ring, which marks the site of septum formation (2, 3). Like tubulin, FtsZ assembles into filaments with GTP but does not form microtubules (4). The precise assembly state and conformation of these FtsZ filaments at the division ring is not clear, although recent electron tomography work suggests that the FtsZ ring consists of multiple short filaments tethered to the membrane at discrete junctures (5), which may represent points along the filaments bridged by membrane anchor proteins.In Escherichia coli, two of these anchor proteins are known. One of these, ZipA, is not well conserved but is an essential protein in E. coli. ZipA binds to the C-terminal tail of FtsZ (68), and purified ZipA promotes bundling of FtsZ filaments in vitro (9, 10). The other, FtsA, is also essential in E. coli and is more widely conserved among bacterial species. FtsA is a member of the HSP70/actin superfamily (11, 12), and like ZipA, it interacts with the C-terminal tail of FtsZ (7, 1315). FtsA can self-associate (16, 17) and bind ATP (12, 18), but reports of ATPase activity vary, with Bacillus subtilis FtsA having high activity (19) and Streptococcus pneumoniae FtsA exhibiting no detectable activity (20). There are no reports of any other in vitro activities of FtsA, including effects on FtsZ assembly.Understanding how FtsA affects FtsZ assembly is important because FtsA has a number of key activities in the cell. It is required for recruitment of a number of divisome proteins (21, 22) and helps to tether the Z ring to the membrane via a C-terminal membrane-targeting sequence (23). FtsA, like ZipA and other divisome proteins, is necessary to activate the contraction of the Z ring (24, 25). In E. coli, the FtsA:FtsZ ratio is crucial for proper cell division, with either too high or too low a ratio inhibiting septum formation (26, 27). This ratio is roughly 1:5, with ∼700 molecules of FtsA and 3200 molecules of FtsZ per cell (28), which works out to concentrations of 1–2 and 5–10 μm, respectively.Another interesting property of FtsA is that single residue alterations in the protein can result in significant enhancement of divisome activity. For example, the R286W mutation of FtsA, also called FtsA*, can substitute for the native FtsA and divide the cell. However, this mutant FtsA causes E. coli cells to divide at less than 80% of their normal length (29) and allows efficient division of E. coli cells in the absence of ZipA (30), indicating that it has gain-of-function activity. FtsA* and other hypermorphic mutations such as E124A and I143L can also increase division activity in cells lacking other essential divisome components (3133). The R286W and E124A mutants of FtsA also bypass the FtsA:FtsZ ratio rule, allowing cell division to occur at higher ratios than with WT2 FtsA. This may be because the altered FtsA proteins self-associate more readily than WT FtsA, which may cause different changes in FtsZ assembly state as compared with WT FtsA (17, 34).In this study, we use an in vitro system with purified FtsZ and a purified tagged version of FtsA* to elucidate the role of FtsA in activating constriction of the Z ring in vivo. We show that FtsA*, at physiological concentrations in the presence of ATP or ADP, has significant effects on the assembly of FtsZ filaments.  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