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.
2.
3.
Accumulation of expanded polyglutamine proteins is considered to be a major pathogenic biomarker of Huntington disease. We isolated SCAMP5 as a novel regulator of cellular accumulation of expanded polyglutamine track protein using cell-based aggregation assays. Ectopic expression of SCAMP5 augments the formation of ubiquitin-positive and detergent-resistant aggregates of mutant huntingtin (mtHTT). Expression of SCAMP5 is markedly increased in the striatum of Huntington disease patients and is induced in cultured striatal neurons by endoplasmic reticulum (ER) stress or by mtHTT. The increase of SCAMP5 impairs endocytosis, which in turn enhances mtHTT aggregation. On the contrary, down-regulation of SCAMP5 alleviates ER stress-induced mtHTT aggregation and endocytosis inhibition. Moreover, stereotactic injection into the striatum and intraperitoneal injection of tunicamycin significantly increase mtHTT aggregation in the striatum of R6/2 mice and in the cortex of N171-82Q mice, respectively. Taken together, these results suggest that exposure to ER stress increases SCAMP5 in the striatum, which positively regulates mtHTT aggregation via the endocytosis pathway.The expansion of CAG repeats (usually beyond a critical threshold of ∼37 glutamine repeats) encoding polyglutamine (polyQ)3 causes, to date, nine late-onset progressive neurodegenerative disorders (1, 2). Expanded polyQ-containing huntingtin is the main aggregate component in the affected neurons (3). Also, molecular chaperones, such as Hsp70, Hsp40/HDJ1 (dHDJ1), and chaperonin TRiC, perturb the aggregation of polyQ track protein and reduce polyQ track cytotoxicity in yeast and cell lines (46) and in Drosophila and mouse models (4, 7). Thus, it seems that HD pathology is closely correlated with the accumulation of insoluble aggregates of mutant huntingtin (mtHTT) containing expanded polyQ (2, 3, 8, 9).Endoplasmic reticulum (ER) stress is crucial in many biological responses and is generated by various signals, such as unfolded protein response, aberrant calcium regulation, oxidative stress, and inflammation (10, 11). ER stress response is generally considered an adaptive reaction of cells to environmental stress, serving as a survival signal (10). On the other hand, increasing evidence also strengthens the importance of ER stress in human diseases. A malfunction or excess of ER stress response caused by aging, genetic mutations, and environmental insults is implicated in human diseases, such as Alzheimer disease, Parkinson disease, diabetes mellitus, and inflammation (1216). mtHTT also induces ER stress at the early stage of HD, and pathogenic ER stress from an aging or stressful environment is severe at the late stage of HD (1719). However, the molecular event linking the aggregation of polyQ track protein to ER stress response is unknown.The ubiquitin/proteasome pathway, a major protein degradation system, is altered or impaired in the cell culture model of HD (2022). On the contrary, autophagy employing lysosomal degradation has been recently considered as a major clearance pathway of insoluble aggregates of polyQ track protein. Thus, inhibition of autophagy has been suggested to modulate the aggregate formation of mtHTT and to affect the toxicity of polyglutamine expansions in fly and mouse models of HD (2325). However, a key molecule controlling the aggregation and clearance of polyQ track proteins needs to be identified.To further our understanding of the regulation of polyQ track protein aggregation, we screened human full-length cDNAs and isolated SCAMP5 (secretory carrier membrane protein 5) as a modulator of polyQ track protein aggregation. SCAMP5 is up-regulated by mtHTT and ER stress and functions to inhibit endocytosis to increase mtHTT aggregation.  相似文献   

4.
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   

5.
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   

6.
The life cycle of protein kinase C (PKC) is tightly controlled by mechanisms that mature the enzyme, sustain the activation-competent enzyme, and degrade the enzyme. Here we show that a conserved PXXP motif (Kannan, N., Haste, N., Taylor, S. S., and Neuwald, A. F. (2007) Proc. Natl. Acad. Sci. U. S. A. 104, 1272–1277), in the C-terminal tail of AGC (c-AMP-dependent protein kinase/protein kinase G/protein kinase C) kinases, controls the processing phosphorylation of conventional and novel PKC isozymes, a required step in the maturation of the enzyme into a signaling-competent species. Mutation of both Pro-616 and Pro-619 to Ala in the conventional PKC βII abolishes the phosphorylation and activity of the kinase. Co-immunoprecipitation studies reveal that conventional and novel, but not atypical, PKC isozymes bind the chaperones Hsp90 and Cdc37 through a PXXP-dependent mechanism. Inhibitors of Hsp90 and Cdc37 significantly reduce the rate of processing phosphorylation of PKC. Of the two C-terminal sites processed by phosphorylation, the hydrophobic motif, but not the turn motif, is regulated by Hsp90. Overlay of purified Hsp90 onto a peptide array containing peptides covering the catalytic domain of PKC βII identified regions surrounding the PXXP segment, but not the PXXP motif itself, as major binding determinants for Hsp90. These Hsp90-binding regions, however, are tethered to the C-terminal tail via a “molecular clamp” formed between the PXXP motif and a conserved Tyr (Tyr-446) in the αE-helix. Disruption of the clamp by mutation of the Tyr to Ala recapitulates the phosphorylation defect of mutating the PXXP motif. These data are consistent with a model in which a molecular clamp created by the PXXP motif in the C-terminal tail and determinants in the αE-helix of the catalytic domain allows the chaperones Hsp90 and Cdc37 to bind newly synthesized PKC, a required event in the processing of PKC by phosphorylation.Protein kinases, which comprise ∼2% of the human genome, are key signal transducers that regulate a wide variety of cellular processes, such as growth, proliferation, and metabolism, through catalysis of specific phosphorylation events (1). By integrating signals from extracellular stimuli and transmitting them to targeted downstream substrates, protein kinases serve as a pivotal point of regulation within the cell. Deregulation and mutation of protein kinases play a causal role in human pathology, notably cancer, poising kinases as important targets for the design of therapeutics (25). Therefore, understanding the mechanisms that regulate protein kinases, such as those important for maturation and processing, would be critical for designing therapeutics that would maintain the correct functioning of signal transduction pathways.Heat shock proteins (Hsp),3 such as Hsp90, are ubiquitously expressed molecular chaperones that facilitate protein folding, regulate quality control, and guide protein turnover in an effort to maintain cellular homeostasis (68). Unlike other chaperones such as Hsp70, which non-specifically assists in folding of nascent polypeptide chains (7), Hsp90 works with a specific and discrete set of client proteins, particularly protein kinases (9). Many of the known client kinases of Hsp90, Src (10), Akt (11, 12), phosphoinositide dependent kinase-1 (PDK-1) (13), and ErbB2-/HER2 (14, 15), require the activity of Hsp90 to reach an activation-competent and mature state. Hsp90 is recruited to its kinase clients through interactions with cochaperones, such as Cdc37, which bridge the interaction between Hsp90 and the kinase client (16, 17); this mechanism is revealed in a structural analysis of the Cdc37-Cdk4-Hsp90 complex (18). Cdc37, originally identified in yeast (19), is a cochaperone specific to the kinome that not only assists Hsp90 function but can also recognize and stabilize clients independently of Hsp90 (20). By binding specific regions of the catalytic domains of these kinases, the Hsp90-Cdc37 complex utilizes ATP to promote and stabilize functional conformations of its clients (8, 16, 17, 21). Pharmacological inhibition of Hsp90 by ansamycin antibiotics such as 17-(allylamino)-17-demethoxygeldanamycin (17-AAG) leads to the destabilization and subsequent proteasomal degradation of its clients (6, 22). Recent studies have identified Hsp90 as a promising therapeutic target in cancer as levels of chaperones and activity of client kinases are frequently up-regulated (2325).The protein kinase C (PKC) family of Ser/Thr kinases serves as a paradigm of how conformation and processing by phosphorylation regulate activity, localization, and inter- and intramolecular interactions (26). The mammalian PKC family consists of 10 isozymes divided into three subclasses (conventional, novel, and atypical) based on their primary structure and second messenger mode of regulation. In the case of conventional PKC isozymes, newly synthesized enzyme is loosely engaged on the membrane in a conformation that exposes the activation loop for phosphorylation by the upstream kinase, PDK-1 (28). This phosphorylation triggers two sequential phosphorylations on the C terminus, one on the turn motif and one on the hydrophobic motif. Phosphorylation of the turn motif is required for phosphorylation of the hydrophobic motif and has recently been shown to depend on the mammalian target of rapamycin complex 2 (mTORC2), a complex consisting of the mammalian target of rapamycin (mTOR), Rictor, Sin1, and mLST8 (29, 30). Turn motif phosphorylation is rapidly followed by phosphorylation at the hydrophobic motif, a reaction that occurs by an intramolecular mechanism in vitro (31). Fully stey phosphorylated PKC is released into the cytosol in a closed conformation in which an autoinhibitory pseudosubstrate sequence occupies the substrate-binding cavity. Upon generation of the lipid second messenger diacylglycerol and elevation of intracellular Ca2+, conventional PKC isozymes (α, βI, βII, γ) translocate to membranes via their membrane-targeting C1 and C2 domains, where they adopt an open conformation in which the pseudosubstrate is expelled from the substrate-binding cavity, permitting phosphorylation of downstream substrates (32). Novel PKC isozymes (δ, ε, θ, and η) only respond to diacylglycerol, and their processing phosphorylations can occur through additional mechanisms (33, 34). Atypical PKC isozymes (ζ and ι/λ) do not respond to either Ca2+ or diacylglycerol but can also undergo regulation of their processing phosphorylations by external stimuli (35). In fact, atypical PKC isozymes contain a Glu at their hydrophobic motif site. Thus, although these three sites are conserved among PKC family members, additional layers of regulation generate specificity in how these kinases become signaling-competent enzymes.The mechanisms of regulation of the activity and signaling properties of PKC by lipid second messengers and phosphorylation have been well characterized; however, mounting evidence suggests that there are many other regulatory inputs for PKC function. Recent analysis of the evolutionary constraints acting on AGC kinase sequences have underscored the importance of the C-terminal tail as a critical regulatory module (36). Deletion mutants of the C-terminal tail have been shown to abrogate PKC activity (37). In addition to containing the key regulatory phosphorylation sites and docking the upstream kinase PDK-1, the C-terminal tail contains key conserved motifs, found within all AGC kinases, that facilitate ATP binding, promote substrate binding, and structure the catalytic core (36). One such motif comprises the segment PXXP; this motif makes key contacts with the catalytic core, where it is important for modulating movement of the catalytic domain (36). Although this PXXP motif is conserved in all the PKC isozymes, its functional role is unknown.In this study, we address the role of the conserved Pro residues in the PXXP motif of the C-terminal tail of PKC βII. We show that mutation of these two Pro to Ala (P616A and P619A) results in a kinase that is not processed by phosphorylation in cells and is thus inactive. Further analysis reveals that this mutant is not able to bind the chaperones Hsp90 and Cdc37, an event that is required for the processing of PKC by phosphorylation. Our peptide array data indicate that Hsp90 binds to regions of the catalytic core such as the αC-β4 loop and the αD-helix that serve as hinge points for C-helix movement (36). Structural analysis delineates that these hinge points are tethered to the C-terminal tail through a molecular clamp formed between the PXXP segment and AGC conserved residues in the αE-helix. Mutation of one of the “clamping” residues, a conserved Tyr (Tyr-446), recapitulates the defect resulting from mutation of the PXXP motif. Our data support a model in which the PXXP motif participates in an intramolecular clamp with determinants in the αE helix of the kinase core, by providing a recognition surface for Hsp90 to bind and facilitate the maturation of PKC, a required step in the processing of the enzyme.  相似文献   

7.
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   

8.
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   

9.
The RAD51 protein is a central player in homologous recombinational repair. The RAD51B protein is one of five RAD51 paralogs that function in the homologous recombinational repair pathway in higher eukaryotes. In the present study, we found that the human EVL (Ena/Vasp-like) protein, which is suggested to be involved in actin-remodeling processes, unexpectedly binds to the RAD51 and RAD51B proteins and stimulates the RAD51-mediated homologous pairing and strand exchange. The EVL knockdown cells impaired RAD51 assembly onto damaged DNA after ionizing radiation or mitomycin C treatment. The EVL protein alone promotes single-stranded DNA annealing, and the recombination activities of the EVL protein are further enhanced by the RAD51B protein. The expression of the EVL protein is not ubiquitous, but it is significantly expressed in breast cancer-derived MCF7 cells. These results suggest that the EVL protein is a novel recombination factor that may be required for repairing specific DNA lesions, and that may cause tumor malignancy by its inappropriate expression.Chromosomal DNA double strand breaks (DSBs)2 are potential inducers of chromosomal aberrations and tumorigenesis, and they are accurately repaired by the homologous recombinational repair (HRR) pathway, without base substitutions, deletions, and insertions (13). In the HRR pathway (4, 5), single-stranded DNA (ssDNA) tails are produced at the DSB sites. The RAD51 protein, a eukaryotic homologue of the bacterial RecA protein, binds to the ssDNA tail and forms a helical nucleoprotein filament. The RAD51-ssDNA filament then binds to the intact double-stranded DNA (dsDNA) to form a three-component complex, containing ssDNA, dsDNA, and the RAD51 protein. In this three-component complex, the RAD51 protein promotes recombination reactions, such as homologous pairing and strand exchange (69).The RAD51 protein requires auxiliary proteins to promote the homologous pairing and strand exchange reactions efficiently in cells (1012). In humans, the RAD52, RAD54, and RAD54B proteins directly interact with the RAD51 protein (1317) and stimulate the RAD51-mediated homologous pairing and/or strand exchange reactions in vitro (1821). The human RAD51AP1 protein, which directly binds to the RAD51 protein (22), was also found to stimulate RAD51-mediated homologous pairing in vitro (23, 24). The BRCA2 protein contains ssDNA-binding, dsDNA-binding, and RAD51-binding motifs (2533), and the Ustilago maydis BRCA2 ortholog, Brh2, reportedly stimulated RAD51-mediated strand exchange (34, 35). Most of these RAD51-interacting factors are known to be required for efficient RAD51 assembly onto DSB sites in cells treated with ionizing radiation (1012).The RAD51B (RAD51L1, Rec2) protein is a member of the RAD51 paralogs, which share about 20–30% amino acid sequence similarity with the RAD51 protein (3638). RAD51B-deficient cells are hypersensitive to DSB-inducing agents, such as cisplatin, mitomycin C (MMC), and γ-rays, indicating that the RAD51B protein is involved in the HRR pathway (3944). Genetic experiments revealed that RAD51B-deficient cells exhibited impaired RAD51 assembly onto DSB sites (39, 44), suggesting that the RAD51B protein functions in the early stage of the HRR pathway. Biochemical experiments also suggested that the RAD51B protein participates in the early to late stages of the HRR pathway (4547).In the present study, we found that the human EVL (Ena/Vasp-like) protein binds to the RAD51 and RAD51B proteins in a HeLa cell extract. The EVL protein is known to be involved in cytoplasmic actin remodeling (48) and is also overexpressed in breast cancer (49). Like the RAD51B knockdown cells, the EVL knockdown cells partially impaired RAD51 foci formation after DSB induction, suggesting that the EVL protein enhances RAD51 assembly onto DSB sites. The purified EVL protein preferentially bound to ssDNA and stimulated RAD51-mediated homologous pairing and strand exchange. The EVL protein also promoted the annealing of complementary strands. These recombination reactions that were stimulated or promoted by the EVL protein were further enhanced by the RAD51B protein. These results strongly suggested that the EVL protein is a novel factor that activates RAD51-mediated recombination reactions, probably with the RAD51B protein. We anticipate that, in addition to its involvement in cytoplasmic actin dynamics, the EVL protein may be required in homologous recombination for repairing specific DNA lesions, and it may cause tumor malignancy by inappropriate recombination enhanced by EVL overexpression in certain types of tumor cells.  相似文献   

10.
11.
12.
NHE5 is a brain-enriched Na+/H+ exchanger that dynamically shuttles between the plasma membrane and recycling endosomes, serving as a mechanism that acutely controls the local pH environment. In the current study we show that secretory carrier membrane proteins (SCAMPs), a group of tetraspanning integral membrane proteins that reside in multiple secretory and endocytic organelles, bind to NHE5 and co-localize predominantly in the recycling endosomes. In vitro protein-protein interaction assays revealed that NHE5 directly binds to the N- and C-terminal cytosolic extensions of SCAMP2. Heterologous expression of SCAMP2 but not SCAMP5 increased cell-surface abundance as well as transporter activity of NHE5 across the plasma membrane. Expression of a deletion mutant lacking the SCAMP2-specific N-terminal cytosolic domain, and a mini-gene encoding the N-terminal extension, reduced the transporter activity. Although both Arf6 and Rab11 positively regulate NHE5 cell-surface targeting and NHE5 activity across the plasma membrane, SCAMP2-mediated surface targeting of NHE5 was reversed by dominant-negative Arf6 but not by dominant-negative Rab11. Together, these results suggest that SCAMP2 regulates NHE5 transit through recycling endosomes and promotes its surface targeting in an Arf6-dependent manner.Neurons and glial cells in the central and peripheral nervous systems are especially sensitive to perturbations of pH (1). Many voltage- and ligand-gated ion channels that control membrane excitability are sensitive to changes in cellular pH (1-3). Neurotransmitter release and uptake are also influenced by cellular and organellar pH (4, 5). Moreover, the intra- and extracellular pH of both neurons and glia are modulated in a highly transient and localized manner by neuronal activity (6, 7). Thus, neurons and glia require sophisticated mechanisms to finely tune ion and pH homeostasis to maintain their normal functions.Na+/H+ exchangers (NHEs)3 were originally identified as a class of plasma membrane-bound ion transporters that exchange extracellular Na+ for intracellular H+, and thereby regulate cellular pH and volume. Since the discovery of NHE1 as the first mammalian NHE (8), eight additional isoforms (NHE2-9) that share 25-70% amino acid identity have been isolated in mammals (9, 10). NHE1-5 commonly exhibit transporter activity across the plasma membrane, whereas NHE6-9 are mostly found in organelle membranes and are believed to regulate organellar pH in most cell types at steady state (11). More recently, NHE10 was identified in human and mouse osteoclasts (12, 13). However, the cDNA encoding NHE10 shares only a low degree of sequence similarity with other known members of the NHE gene family, raising the possibility that this sodium-proton exchanger may belong to a separate gene family distantly related to NHE1-9 (see Ref. 9).NHE gene family members contain 12 putative transmembrane domains at the N terminus followed by a C-terminal cytosolic extension that plays a role in regulation of the transporter activity by protein-protein interactions and phosphorylation. NHEs have been shown to regulate the pH environment of synaptic nerve terminals and to regulate the release of neurotransmitters from multiple neuronal populations (14-16). The importance of NHEs in brain function is further exemplified by the findings that spontaneous or directed mutations of the ubiquitously expressed NHE1 gene lead to the progression of epileptic seizures, ataxia, and increased mortality in mice (17, 18). The progression of the disease phenotype is associated with loss of specific neuron populations and increased neuronal excitability. However, NHE1-null mice appear to develop normally until 2 weeks after birth when symptoms begin to appear. Therefore, other mechanisms may compensate for the loss of NHE1 during early development and play a protective role in the surviving neurons after the onset of the disease phenotype.NHE5 was identified as a unique member of the NHE gene family whose mRNA is expressed almost exclusively in the brain (19, 20), although more recent studies have suggested that NHE5 might be functional in other cell types such as sperm (21, 22) and osteosarcoma cells (23). Curiously, mutations found in several forms of congenital neurological disorders such as spinocerebellar ataxia type 4 (24-26) and autosomal dominant cerebellar ataxia (27-29) have been mapped to chromosome 16q22.1, a region containing NHE5. However, much remains unknown as to the molecular regulation of NHE5 and its role in brain function.Very few if any proteins work in isolation. Therefore identification and characterization of binding proteins often reveal novel functions and regulation mechanisms of the protein of interest. To begin to elucidate the biological role of NHE5, we have started to explore NHE5-binding proteins. Previously, β-arrestins, multifunctional scaffold proteins that play a key role in desensitization of G-protein-coupled receptors, were shown to directly bind to NHE5 and promote its endocytosis (30). This study demonstrated that NHE5 trafficking between endosomes and the plasma membrane is regulated by protein-protein interactions with scaffold proteins. More recently, we demonstrated that receptor for activated C-kinase 1 (RACK1), a scaffold protein that links signaling molecules such as activated protein kinase C, integrins, and Src kinase (31), directly interacts with and activates NHE5 via integrin-dependent and independent pathways (32). These results further indicate that NHE5 is partly associated with focal adhesions and that its targeting to the specialized microdomain of the plasma membrane may be regulated by various signaling pathways.Secretory carrier membrane proteins (SCAMPs) are a family of evolutionarily conserved tetra-spanning integral membrane proteins. SCAMPs are found in multiple organelles such as the Golgi apparatus, trans-Golgi network, recycling endosomes, synaptic vesicles, and the plasma membrane (33, 34) and have been shown to play a role in exocytosis (35-38) and endocytosis (39). Currently, five isoforms of SCAMP have been identified in mammals. The extended N terminus of SCAMP1-3 contain multiple Asn-Pro-Phe (NPF) repeats, which may allow these isoforms to participate in clathrin coat assembly and vesicle budding by binding to Eps15 homology (EH)-domain proteins (40, 41). Further, SCAMP2 was shown recently to bind to the small GTPase Arf6 (38), which is believed to participate in traffic between the recycling endosomes and the cell surface (42, 43). More recent studies have suggested that SCAMPs bind to organellar membrane type NHE7 (44) and the serotonin transporter SERT (45) and facilitate targeting of these integral membrane proteins to specific intracellular compartments. We show in the current study that SCAMP2 binds to NHE5, facilitates the cell-surface targeting of NHE5, and elevates Na+/H+ exchange activity at the plasma membrane, whereas expression of a SCAMP2 deletion mutant lacking the N-terminal domain containing the NPF repeats suppresses the effect. Further we show that this activity of SCAMP2 requires an active form of a small GTPase Arf6, but not Rab11. We propose a model in which SCAMPs bind to NHE5 in the endosomal compartment and control its cell-surface abundance via an Arf6-dependent pathway.  相似文献   

13.
Many G protein-coupled receptors (GPCRs) recycle after agonist-induced endocytosis by a sequence-dependent mechanism, which is distinct from default membrane flow and remains poorly understood. Efficient recycling of the β2-adrenergic receptor (β2AR) requires a C-terminal PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (PDZbd), an intact actin cytoskeleton, and is regulated by the endosomal protein Hrs (hepatocyte growth factor-regulated substrate). The PDZbd is thought to link receptors to actin through a series of protein interaction modules present in NHERF/EBP50 (Na+/H+ exchanger 3 regulatory factor/ezrin-binding phosphoprotein of 50 kDa) family and ERM (ezrin/radixin/moesin) family proteins. It is not known, however, if such actin connectivity is sufficient to recapitulate the natural features of sequence-dependent recycling. We addressed this question using a receptor fusion approach based on the sufficiency of the PDZbd to promote recycling when fused to a distinct GPCR, the δ-opioid receptor, which normally recycles inefficiently in HEK293 cells. Modular domains mediating actin connectivity promoted receptor recycling with similarly high efficiency as the PDZbd itself, and recycling promoted by all of the domains was actin-dependent. Regulation of receptor recycling by Hrs, however, was conferred only by the PDZbd and not by downstream interaction modules. These results suggest that actin connectivity is sufficient to mimic the core recycling activity of a GPCR-linked PDZbd but not its cellular regulation.G protein-coupled receptors (GPCRs)2 comprise the largest family of transmembrane signaling receptors expressed in animals and transduce a wide variety of physiological and pharmacological information. While these receptors share a common 7-transmembrane-spanning topology, structural differences between individual GPCR family members confer diverse functional and regulatory properties (1-4). A fundamental mechanism of GPCR regulation involves agonist-induced endocytosis of receptors via clathrin-coated pits (4). Regulated endocytosis can have multiple functional consequences, which are determined in part by the specificity with which internalized receptors traffic via divergent downstream membrane pathways (5-7).Trafficking of internalized GPCRs to lysosomes, a major pathway traversed by the δ-opioid receptor (δOR), contributes to proteolytic down-regulation of receptor number and produces a prolonged attenuation of subsequent cellular responsiveness to agonist (8, 9). Trafficking of internalized GPCRs via a rapid recycling pathway, a major route traversed by the β2-adrenergic receptor (β2AR), restores the complement of functional receptors present on the cell surface and promotes rapid recovery of cellular signaling responsiveness (6, 10, 11). When co-expressed in the same cells, the δOR and β2AR are efficiently sorted between these divergent downstream membrane pathways, highlighting the occurrence of specific molecular sorting of GPCRs after endocytosis (12).Recycling of various integral membrane proteins can occur by default, essentially by bulk membrane flow in the absence of lysosomal sorting determinants (13). There is increasing evidence that various GPCRs, such as the β2AR, require distinct cytoplasmic determinants to recycle efficiently (14). In addition to requiring a cytoplasmic sorting determinant, sequence-dependent recycling of the β2AR differs from default recycling in its dependence on an intact actin cytoskeleton and its regulation by the conserved endosomal sorting protein Hrs (hepatocyte growth factor receptor substrate) (11, 14). Compared with the present knowledge regarding protein complexes that mediate sorting of GPCRs to lysosomes (15, 16), however, relatively little is known about the biochemical basis of sequence-directed recycling or its regulation.The β2AR-derived recycling sequence conforms to a canonical PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (henceforth called PDZbd), and PDZ-mediated protein association(s) with this sequence appear to be primarily responsible for its endocytic sorting activity (17-20). Fusion of this sequence to the cytoplasmic tail of the δOR effectively re-routes endocytic trafficking of engineered receptors from lysosomal to recycling pathways, establishing the sufficiency of the PDZbd to function as a transplantable sorting determinant (18). The β2AR-derived PDZbd binds with relatively high specificity to the NHERF/EBP50 family of PDZ proteins (21, 22). A well-established biochemical function of NHERF/EBP50 family proteins is to associate integral membrane proteins with actin-associated cytoskeletal elements. This is achieved through a series of protein-interaction modules linking NHERF/EBP50 family proteins to ERM (ezrin-radixin-moesin) family proteins and, in turn, to actin filaments (23-26). Such indirect actin connectivity is known to mediate other effects on plasma membrane organization and function (23), however, and NHERF/EBP50 family proteins can bind to additional proteins potentially important for endocytic trafficking of receptors (23, 25). Thus it remains unclear if actin connectivity is itself sufficient to promote sequence-directed recycling of GPCRs and, if so, if such connectivity recapitulates the normal cellular regulation of sequence-dependent recycling. In the present study, we took advantage of the modular nature of protein connectivity proposed to mediate β2AR recycling (24, 26), and extended the opioid receptor fusion strategy used successfully for identifying diverse recycling sequences in GPCRs (27-29), to address these fundamental questions.Here we show that the recycling activity of the β2AR-derived PDZbd can be effectively bypassed by linking receptors to ERM family proteins in the absence of the PDZbd itself. Further, we establish that the protein connectivity network can be further simplified by fusing receptors to an interaction module that binds directly to actin filaments. We found that bypassing the PDZ-mediated interaction using either domain is sufficient to mimic the ability of the PDZbd to promote efficient, actin-dependent recycling of receptors. Hrs-dependent regulation, however, which is characteristic of sequence-dependent recycling of wild-type receptors, was recapitulated only by the fused PDZbd and not by the proposed downstream interaction modules. These results support a relatively simple architecture of protein connectivity that is sufficient to mimic the core recycling activity of the β2AR-derived PDZbd, but not its characteristic cellular regulation. Given that an increasing number of GPCRs have been shown to bind PDZ proteins that typically link directly or indirectly to cytoskeletal elements (17, 27, 30-32), the present results also suggest that actin connectivity may represent a common biochemical principle underlying sequence-dependent recycling of various GPCRs.  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.
The amyloid precursor protein (APP) is implied both in cell growth and differentiation and in neurodegenerative processes in Alzheimer disease. Regulated proteolysis of APP generates biologically active fragments such as the neuroprotective secreted ectodomain sAPPα and the neurotoxic β-amyloid peptide. Furthermore, it has been suggested that the intact transmembrane APP plays a signaling role, which might be important for both normal synaptic plasticity and neuronal dysfunction in dementia. To understand APP signaling, we tracked single molecules of APP using quantum dots and quantitated APP homodimerization using fluorescence lifetime imaging microscopy for the detection of Förster resonance energy transfer in living neuroblastoma cells. Using selective labeling with synthetic fluorophores, we show that the dimerization of APP is considerably higher at the plasma membrane than in intracellular membranes. Heparan sulfate significantly contributes to the almost complete dimerization of APP at the plasma membrane. Importantly, this technique for the first time structurally defines the initiation of APP signaling by binding of a relevant physiological extracellular ligand; our results indicate APP as receptor for neuroprotective sAPPα, as sAPPα binding disrupts APP dimers, and this disruption of APP dimers by sAPPα is necessary for the protection of neuroblastoma cells against starvation-induced cell death. Only cells expressing reversibly dimerized wild-type, but not covalently dimerized mutant APP are protected by sAPPα. These findings suggest a potentially beneficial effect of increasing sAPPα production or disrupting APP dimers for neuronal survival.The amyloid precursor protein (APP)4 is known both for its important role in the development and plasticity of the nervous system (16) and for its involvement in Alzheimer disease (AD) (7, 8). Despite intensive research efforts, the initial events that lead to the prevalent sporadic, i.e. non-familial, forms of AD are still unclear. Furthermore, although a higher gene dose of APP (9) or the presence of pathological APP mutations is sufficient to induce familial AD (for review, see Ref. 10), the exact pathological mechanism that is triggered by APP is still under debate.Some fragments of APP, such as the β-amyloid peptide (Aβ), are thought to contribute to synaptic dysfunction and neurotoxicity (11, 12). On the other hand, the α-secretase-derived extracellular fragment of APP (sAPPα), which is present at lower levels in AD patients than in controls (13), has been shown to be beneficial for memory function, to possess neuroprotective properties, and to counteract the effects of Aβ (1418).Signaling by transmembrane APP may directly contribute to neurodegeneration in AD (1924); however, the signal transduction pathway for transmembrane APP remains unknown, although several potential regulatory proteins, glycosaminoglycans, and metal ions are known to bind with high affinity to APP and sAPPα (25, 26). The most common form of signal transduction for single-pass transmembrane proteins is the ligand-induced perturbation of a monomer/dimer equilibrium. Indeed, the dimerization of transmembrane APP has been implied several times in the past. Several studies have investigated the effects of presumed dimer-breaking perturbations on biological read-outs, such as the production of Aβ (27, 28), but without directly measuring the APP aggregation state, or have investigated the aggregation state of APP subdomains, often reconstituted in cell-free systems (2732). Dimerization interfaces in both the extracellular and the transmembrane domain have been suggested.In the studies investigating the aggregation state of full-length APP, most of the employed methods, such as chemical cross-linking and co-immunoprecipitation, do not lend themselves readily to a rigorous quantitative analysis of the abundance of potentially instable dimers (31, 33), whereas in other cases the use of chimeras may have influenced the dimerization potential or precluded the search for a natural stimulus (23, 34). The only previously reported direct observation of APP dimerization by Förster resonance energy transfer (FRET) microscopy uses an assay in which the FRET efficiency varies with the level of overexpression (35). Therefore, a concentration-dependent FRET component due to nonspecific stochastic encounters cannot be excluded in this study.Most importantly, as none of the published procedures permitted the selective detection of APP dimers on the surface of live cells, where they would encounter ligands, they could not differentiate between subpopulations of APP. This may be one reason why no natural ligand of APP has ever been shown to signal via modulation of its monomer/dimer equilibrium.Another elusive goal is the identity of the receptor for neuroprotective sAPPα (3639). The ligand-dependent dimerization of sAPPα in solution (40) and its origination from transmembrane APP suggest that APP might serve as receptor for sAPPα, but this binding has never been experimentally shown.  相似文献   

16.
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   

17.
The cystic fibrosis transmembrane conductance regulator (CFTR) is a cAMP and protein kinase A (PKA)-regulated Cl channel in the apical membrane of epithelial cells. The metabolically regulated and adenosine monophosphate-stimulated kinase (AMPK) is colocalized with CFTR and attenuates its function. However, the sites for CFTR phosphorylation and the precise mechanism of inhibition of CFTR by AMPK remain obscure. We demonstrate that CFTR normally remains closed at baseline, but nevertheless, opens after inhibition of AMPK. AMPK phosphorylates CFTR in vitro at two essential serines (Ser737 and Ser768) in the R domain, formerly identified as “inhibitory” PKA sites. Replacement of both serines by alanines (i) reduced phosphorylation of the R domain, with Ser768 having dramatically greater impact, (ii) produced CFTR channels that were partially open in the absence of any stimulation, (iii) significantly augmented their activation by IBMX/forskolin, and (iv) eliminated CFTR inhibition post AMPK activation. Attenuation of CFTR by AMPK activation was detectable in the absence of cAMP-dependent stimulation but disappeared in maximally stimulated oocytes. Our data also suggest that AMP is produced by local phosphodiesterases in close proximity to CFTR. Thus we propose that CFTR channels are kept closed in nonstimulated epithelia with high baseline AMPK activity but CFTR may be basally active in tissues with lowered endogenous AMPK activity.The cystic fibrosis transmembrane regulator (CFTR)2 gene is mutated in patients with cystic fibrosis. CFTR has an adapted ABC transporter structural motif thereby creating an anion channel at the apical surface of secretory epithelia (1). The consequent CFTR-mediated ion transport is tightly controlled by ATP binding and phosphorylation by protein kinase A (PKA). However, a number of other protein kinases including PKC, Ca2+/calmodulin-dependent kinase, and cGMP-dependent kinase also control the activity of CFTR (24). These kinases converge on the regulatory domain of CFTR that is unique not only within the large ABC transporter family but among all known sequences, and may be considered as a “phosphorylation control module” (3). Regulation of CFTR by an inhibitory kinase, the adenosine monophosphate-dependent kinase (AMPK), has been described recently but the regulatory sites within CFTR, the mechanism of regulation, and the physiological relevance have all remained obscure (58). Additionally, CFTR mutation is linked to inflammation and a lack of functional CFTR expression has itself been suggested to up-regulate AMPK activity in epithelial cells carrying the cystic fibrosis (CF) defect. Pharmacologic AMPK activation was shown to inhibit secretion of inflammatory mediators (9). Thus AMPK may play multiple roles in CF pathophysiology making the mechanism of interaction an important problem in biology.AMPK is a ubiquitous serine/threonine kinase that exists as a heterotrimer with a catalytic α subunit and regulatory β and γ subunits, each with multiple isoforms. In response to metabolic depletion and a consequent increase in the cellular AMP to ATP ratio, AMPK phosphorylates numerous proteins and activates catabolic pathways that generate ATP, whereas inhibiting cell growth, protein biosynthesis, and a number of other ATP-consuming processes, thereby operating as a cellular “low-fuel” sensor (10, 11). AMPK also controls signaling pathways involved in apoptosis, cell cycle, and tissue inflammation (12). Because AMPK is a cellular metabolic sensor that inhibits CFTR and limits cAMP activated Cl secretion, a coupling of membrane transport by CFTR to the cellular metabolism has been proposed (13). However, AMPK activity can also increase without detectable changes in the cytosolic AMP to ATP ratio, suggesting a contribution of additional AMP-independent signals for regulation of CFTR by AMPK (14). Drugs used to combat type 2 diabetes, such as phenformin and metformin, act in this manner to activate AMPK, AMP-independently. It is also likely that cytosolic AMP is compartmentalized depending on the distribution of AMP generating enzymes such as phosphodiesterases that convert cAMP to AMP. The concept of spatiotemporal control of cAMP signaling by anchored protein complexes is well established (15). CFTR is known to form such macromolecular complexes with a number of interacting partners (1618). For example, competitive interaction of EBP50-PKA and Shank2-PDE4D with CFTR has been demonstrated recently (19). In addition, Barnes and co-workers (20) demonstrated that phosphodiesterase 4D generates a cAMP diffusion barrier local to the apical membrane of the airway epithelium. It is therefore likely that activator pathways through cAMP and inhibitory AMP/AMPK signaling occur in a local CFTR-organized compartment. Here we explore the functional links between CFTR, inhibition of phosphodiesterases, and AMPK focusing on the effects of mutating putative AMPK targets within the R domain on CFTR function.  相似文献   

18.
19.
The protein kinase D (PKD) family of serine/threonine kinases, which can be activated by gastrointestinal hormones, consists of three distinct isoforms that modulate a variety of cellular processes including intracellular protein transport as well as constitutive and regulated secretion. Although isoform-specific functions have been identified in a variety of cell lines, the expression and function of PKD isoforms in normal, differentiated secretory tissues is unknown. Here, we demonstrate that PKD isoforms are differentially expressed in the exocrine and endocrine cells of the pancreas. Specifically, PKD3 is the predominant isoform expressed in exocrine cells of the mouse and human pancreas, whereas PKD1 and PKD2 are more abundantly expressed in the pancreatic islets. Within isolated mouse pancreatic acinar cells, PKD3 undergoes rapid membrane translocation, trans-activating phosphorylation, and kinase activation after gastrointestinal hormone or cholinergic stimulation. PKD phosphorylation in pancreatic acinar cells occurs viaaCa2+-independent, diacylglycerol- and protein kinase C-dependent mechanism. PKD phosphorylation can also be induced by physiologic concentrations of secretagogues and by in vivo stimulation of the pancreas. Furthermore, activation of PKD3 potentiates MEK/ERK/RSK (RSK, ribosomal S6 kinase) signaling and significantly enhances cholecystokinin-mediated pancreatic amylase secretion. These findings reveal a novel distinction between the exocrine and endocrine cells of the pancreas and further identify PKD3 as a signaling molecule that promotes hormone-stimulated amylase secretion.Protein kinase D (PKD),2 a serine/threonine kinase family with a catalytic domain homologous to the Ca2+/calmodulin-dependent kinase domain and two cysteine-rich phorbol ester binding domains similar to those of protein kinase C (PKC), is a physiologically important downstream mediator of diacylglycerol (DAG) signal transduction (1, 2). The mammalian PKDs include three members, PKD1, PKD2, and PKD3, which demonstrate different expression patterns and functions depending on the cell type and external signal stimuli. PKDs are ubiquitously expressed, but levels of individual isoforms vary with developmental stage and cell type (3). PKD proteins are reported to localize in the cytosol, Golgi, nucleus, and vesicle structures (4-9). Activation of PKDs results in a dynamic translocation among subcellular compartments (10, 11). Expression of multiple isoforms in different cell types and in different subcellular localizations suggests that individual PKD isoforms may serve specific functions. The majority of findings demonstrating the diverse expression patterns and functions of PKD have been described using established cell lines (4-9, 12). However, little is known about PKD isoform expression and function in normal differentiated cells and tissues.Recent functional studies have shown that PKD isoforms differentially regulate exocytic protein trafficking and cargo specificity (9, 12-14). Furthermore, PKD isoforms are differentially activated by oxidative stress signaling via PKCδ-mediated tyrosine phosphorylation (15). In each of these studies, PKD3 was found to have a regulatory mechanism or cellular function distinct from that of PKD1 and PKD2. Unlike the other two isoforms, PKD3 lacks the N terminus hydrophobic domain or the C terminus PDZ binding motif and contains divergent PH (pleckstrin homology) and C1 domains, which are important for regulating its catalytic activity (12, 16, 17). Current knowledge of the physiologic function of PKD3 is limited. It has been demonstrated using kinase-inactive mutants that PKD3 activity is required for basolateral exocytosis in Madin-Darby canine kidney cells (13). PKD3 has also been implicated in the epigenetic control of chromatin by regulating class II histone deacetylases in B lymphocytes (18). Furthermore, PKD3 was found to be a specific regulator of glucose transport in skeletal muscle cells (19).The exocrine pancreas is highly specialized for the synthesis, storage, and exocrine secretion of digestive enzymes and bicarbonate-rich fluid (20). More than 90% of the newly synthesized proteins in the pancreas is targeted to the secretory pathway (21). In addition, the pancreas contains a variety of endocrine cells localized to the islets which secrete peptide hormones. Numerous steps in the secretory pathway are modulated by DAG signaling, which promotes secretion by maintaining Golgi function and/or activating DAG receptor kinases such as PKCs, which are regulators of exocytic proteins (1, 22-25). PKD is also critical for DAG-mediated secretion, as it is recruited by DAG to the trans-Golgi network, where it phosphorylates the lipid kinase phosphatidylinositol 4-kinase to initiate the process of vesicle fission (9, 26). Gastrointestinal (GI) hormones such as cholecystokinin (CCK), gastrin, neurotensin (NT), and bombesin (BBS)/gastrin-releasing peptide are potent regulatory peptides that modulate pancreatic function (27, 28). They are known to activate PKDs to promote cell proliferation and survival in gut epithelial cells (29-32); however, the role of PKDs in modulating the secretory actions of GI hormones is unknown.Although the PKD isoforms have been reported to be expressed in secretory tissues such as salivary glands, adrenal glands, intestinal mucosa, and the pituitary (3, 5, 33), the role of PKD in the process of regulated secretion remains poorly understood. Previously, we demonstrated that PKD1 mediates NT peptide secretion from a pancreas-derived neuroendocrine cell line, BON, and that PKD1 activation is regulated by PKC and Rho/Rho kinase pathways (4); PKD1 and PKD2 isoforms are highly expressed in this endocrine cell line with little to no PKD3 expression, thus suggesting that PKD1/2 may be the predominant isoforms for endocrine secretion. The distribution and role of PKD isoforms in the pancreas, an organ with both exocrine and endocrine functions, is not known. Interestingly, we demonstrate that in both human and mouse pancreas, PKD3 is the predominant PKD isoform expressed in the exocrine acini, whereas PKD1 and PKD2 are more highly expressed in endocrine islets. PKD3 is catalytically activated by GI hormone stimulation of the pancreas, and its activation is dependent on CCK1/2 receptor binding and on DAG/PKC activity. PKD3 overexpression in mouse pancreatic acinar cells significantly increased CCK-mediated pancreatic amylase secretion, suggesting that PKD3, in concert with other signaling molecules, contributes to stimulated amylase secretion. Our findings reveal a distinct expression pattern in the exocrine and endocrine cells of the mouse and human pancreas and identify PKD3 as a novel DAG-activated mediator of the exocrine secretory process in response to GI hormone signaling.  相似文献   

20.
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   

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