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 28 毫秒
1.
2.
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and peripheral neurons where it generates transient cation currents when extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic dendritic spines and has critical effects in neurological diseases associated with a reduced pH. However, knowledge of the proteins that interact with ASIC1a and influence its function is limited. Here, we show that α-actinin, which links membrane proteins to the actin cytoskeleton, associates with ASIC1a in brain and in cultured cells. The interaction depended on an α-actinin-binding site in the ASIC1a C terminus that was specific for ASIC1a versus other ASICs and for α-actinin-1 and -4. Co-expressing α-actinin-4 altered ASIC1a current density, pH sensitivity, desensitization rate, and recovery from desensitization. Moreover, reducing α-actinin expression altered acid-activated currents in hippocampal neurons. These findings suggest that α-actinins may link ASIC1a to a macromolecular complex in the postsynaptic membrane where it regulates ASIC1a activity.Acid-sensing ion channels (ASICs)2 are H+-gated members of the DEG/ENaC family (13). Members of this family contain cytosolic N and C termini, two transmembrane domains, and a large cysteine-rich extracellular domain. ASIC subunits combine as homo- or heterotrimers to form cation channels that are widely expressed in the central and peripheral nervous systems (14). In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have two splice forms, a and b. Central nervous system neurons express ASIC1a, ASIC2a, and ASIC2b (57). Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2, and half-maximal activation occurs at pH 6.5–6.8 (810). These channels desensitize in the continued presence of a low extracellular pH, and they can conduct Ca2+ (9, 1113). ASIC1a is required for acid-evoked currents in central nervous system neurons; disrupting the gene encoding ASIC1a eliminates H+-gated currents unless extracellular pH is reduced below pH 5.0 (5, 7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions and present in dendritic spines, the site of excitatory synapses (5, 14, 15). Consistent with this localization, ASIC1a null mice manifested deficits in hippocampal long term potentiation, learning, and memory, which suggested that ASIC1a is required for normal synaptic plasticity (5, 16). ASICs might be activated during neurotransmission when synaptic vesicles empty their acidic contents into the synaptic cleft or when neuronal activity lowers extracellular pH (1719). Ion channels, including those at the synapse often interact with multiple proteins in a macromolecular complex that incorporates regulators of their function (20, 21). For ASIC1a, only a few interacting proteins have been identified. Earlier work indicated that ASIC1a interacts with another postsynaptic scaffolding protein, PICK1 (15, 22, 23). ASIC1a also has been reported to interact with annexin II light chain p11 through its cytosolic N terminus to increase cell surface expression (24) and with Ca2+/calmodulin-dependent protein kinase II to phosphorylate the channel (25). However, whether ASIC1a interacts with additional proteins and with the cytoskeleton remain unknown. Moreover, it is not known whether such interactions alter ASIC1a function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic residues that might bind α-actinins. α-Actinins cluster membrane proteins and signaling molecules into macromolecular complexes and link membrane proteins to the actincytoskeleton (for review, Ref. 26). Four genes encode α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an N-terminal head domain that binds F-actin, a C-terminal region containing two EF-hand motifs, and a central rod domain containing four spectrin-like motifs (2628). The C-terminal portion of the rod segment appears to be crucial for binding to membrane proteins. The α-actinins assemble into antiparallel homodimers through interactions in their rod domain. α-Actinins-1, -2, and -4 are enriched in dendritic spines, concentrating at the postsynaptic membrane (2935). In the postsynaptic membrane of excitatory synapses, α-actinin connects the NMDA receptor to the actin cytoskeleton, and this interaction is key for Ca2+-dependent inhibition of NMDA receptors (3638). α-Actinins can also regulate the membrane trafficking and function of several cation channels, including L-type Ca2+ channels, K+ channels, and TRP channels (3941).To better understand the function of ASIC1a channels in macromolecular complexes, we asked if ASIC1a associates with α-actinins. We were interested in the α-actinins because they and ASIC1a, both, are present in dendritic spines, ASIC1a contains a potential α-actinin binding sequence, and the related epithelial Na+ channel (ENaC) interacts with the cytoskeleton (42, 43). Therefore, we hypothesized that α-actinin interacts structurally and functionally with ASIC1a.  相似文献   

3.
4.
We have investigated the possible biochemical basis for enhancements in NO production in endothelial cells that have been correlated with agonist- or shear stress-evoked phosphorylation at Ser-1179. We have found that a phosphomimetic substitution at Ser-1179 doubles maximal synthase activity, partially disinhibits cytochrome c reductase activity, and lowers the EC50(Ca2+) values for calmodulin binding and enzyme activation from the control values of 182 ± 2 and 422 ± 22 nm to 116 ± 2 and 300 ± 10 nm. These are similar to the effects of a phosphomimetic substitution at Ser-617 (Tran, Q. K., Leonard, J., Black, D. J., and Persechini, A. (2008) Biochemistry 47, 7557–7566). Although combining substitutions at Ser-617 and Ser-1179 has no additional effect on maximal synthase activity, cooperativity between the two substitutions completely disinhibits reductase activity and further reduces the EC50(Ca2+) values for calmodulin binding and enzyme activation to 77 ± 2 and 130 ± 5 nm. We have confirmed that specific Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 and phosphomimetic substitutions at these positions have similar functional effects. Changes in the biochemical properties of eNOS produced by combined phosphorylation at Ser-617 and Ser-1179 are predicted to substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm.The nitric-oxide synthases catalyze formation of NO and l-citrulline from l-arginine and O2, with NADPH as the electron donor (1). The role of NO generated by endothelial nitricoxide synthase (eNOS)2 in the regulation of smooth muscle tone is well established and was the first of several physiological roles for this small molecule that have so far been identified (2). The nitric-oxide synthases are homodimers of 130–160-kDa subunits. Each subunit contains a reductase and oxygenase domain (1). A significant difference between the reductase domains in eNOS and nNOS and the homologous P450 reductases is the presence of inserts in these synthase isoforms that appear to maintain them in their inactive states (3, 4). A calmodulin (CaM)-binding domain is located in the linker that connects the reductase and oxygenase domains, and the endothelial and neuronal synthases both require Ca2+ and exogenous CaM for activity (5, 6). When CaM is bound, it somehow counteracts the effects of the autoinhibitory insert(s) in the reductase. The high resolution structure for the complex between (Ca2+)4-CaM and the isolated CaM-binding domain from eNOS indicates that the C-ter and N-ter lobes of CaM, which each contain a pair of Ca2+-binding sites, enfold the domain, as has been observed in several other such CaM-peptide complexes (7). Consistent with this structure, investigations of CaM-dependent activation of the neuronal synthase suggest that both CaM lobes must participate (8, 9).Bovine eNOS can be phosphorylated in endothelial cells at Ser-116, Thr-497, Ser-617, Ser-635, and Ser-1179 (1012). There are equivalent phosphorylation sites in the human enzyme (1012). Phosphorylation of the bovine enzyme at Thr-497, which is located in the CaM-binding domain, blocks CaM binding and enzyme activation (7, 11, 13, 14). Ser-116 can be basally phosphorylated in cells (10, 11, 13, 15), and dephosphorylation of this site has been correlated with increased NO production (13, 15). However, it has also been reported that a phosphomimetic substitution at this position has no effect on enzyme activity measured in vitro (13). Ser-1179 is phosphorylated in response to a variety of stimuli, and this has been reliably correlated with enhanced NO production in cells (10, 11). Indeed, NO production is elevated in transgenic endothelium expressing an eNOS mutant containing an S1179D substitution, but not in tissue expressing an S1179A mutant (16). Shear stress or insulin treatment is correlated with Akt-catalyzed phosphorylation of Ser-1179 in endothelial cells, and this is correlated with increased NO production in the absence of extracellular Ca2+ (1719). Akt-catalyzed phosphorylation or an S1179D substitution has also been correlated with increased synthase activity in cell extracts at low intracellular free [Ca2+] (17). Increased NO production has also been observed in cells expressing an eNOS mutant containing an S617D substitution, and physiological stimuli such as shear-stress, bradykinin, VEGF, and ATP appear to stimulate Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 (12, 13, 20). Although S617D eNOS has been reported to have the same maximum activity in vitro as the wild type enzyme (20), in our hands an S617D substitution increases the maximal CaM-dependent synthase activity of purified mutant enzyme ∼2-fold, partially disinhibits reductase activity, and reduces the EC50(Ca2+) values for CaM binding and enzyme activation (21).In this report, we describe the effects of a phosphomimetic Asp substitution at Ser-1179 in eNOS on the Ca2+ dependence of CaM binding and CaM-dependent activation of reductase and synthase activities. We also describe the effects on these properties of combining this substitution with one at Ser-617. Finally, we demonstrate that Akt-catalyzed phosphorylation and Asp substitutions at Ser-617 and Ser-1179 have similar functional effects. Our results suggest that phosphorylation of eNOS at Ser-617 and Ser-1179 can substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm, while single phosphorylations at these sites produce smaller activity increases, and can do so only at higher free Ca2+ concentrations.  相似文献   

5.
6.
Members of the CLC gene family either function as chloride channels or as anion/proton exchangers. The plant AtClC-a uses the pH gradient across the vacuolar membrane to accumulate the nutrient in this organelle. When AtClC-a was expressed in Xenopus oocytes, it mediated exchange and less efficiently mediated Cl/H+ exchange. Mutating the “gating glutamate” Glu-203 to alanine resulted in an uncoupled anion conductance that was larger for Cl than . Replacing the “proton glutamate” Glu-270 by alanine abolished currents. These could be restored by the uncoupling E203A mutation. Whereas mammalian endosomal ClC-4 and ClC-5 mediate stoichiometrically coupled 2Cl/H+ exchange, their transport is largely uncoupled from protons. By contrast, the AtClC-a-mediated accumulation in plant vacuoles requires tight coupling. Comparison of AtClC-a and ClC-5 sequences identified a proline in AtClC-a that is replaced by serine in all mammalian CLC isoforms. When this proline was mutated to serine (P160S), Cl/H+ exchange of AtClC-a proceeded as efficiently as exchange, suggesting a role of this residue in exchange. Indeed, when the corresponding serine of ClC-5 was replaced by proline, this Cl/H+ exchanger gained efficient coupling. When inserted into the model Torpedo chloride channel ClC-0, the equivalent mutation increased nitrate relative to chloride conductance. Hence, proline in the CLC pore signature sequence is important for exchange and conductance both in plants and mammals. Gating and proton glutamates play similar roles in bacterial, plant, and mammalian CLC anion/proton exchangers.CLC proteins are found in all phyla from bacteria to humans and either mediate electrogenic anion/proton exchange or function as chloride channels (1). In mammals, the roles of plasma membrane CLC Cl channels include transepithelial transport (25) and control of muscle excitability (6), whereas vesicular CLC exchangers may facilitate endocytosis (7) and lysosomal function (810) by electrically shunting vesicular proton pump currents (11). In the plant Arabidopsis thaliana, there are seven CLC isoforms (AtClC-a–AtClC-g)2 (1215), which may mostly reside in intracellular membranes. AtClC-a uses the pH gradient across the vacuolar membrane to transport the nutrient nitrate into that organelle (16). This secondary active transport requires a tightly coupled exchange. Astonishingly, however, mammalian ClC-4 and -5 and bacterial EcClC-1 (one of the two CLC isoforms in Escherichia coli) display tightly coupled Cl/H+ exchange, but anion flux is largely uncoupled from H+ when is transported (1721). The lack of appropriate expression systems for plant CLC transporters (12) has so far impeded structure-function analysis that may shed light on the ability of AtClC-a to perform efficient exchange. This dearth of data contrasts with the extensive mutagenesis work performed with CLC proteins from animals and bacteria.The crystal structure of bacterial CLC homologues (22, 23) and the investigation of mutants (17, 1921, 2429) have yielded important insights into their structure and function. CLC proteins form dimers with two largely independent permeation pathways (22, 25, 30, 31). Each of the monomers displays two anion binding sites (22). A third binding site is observed when a certain key glutamate residue, which is located halfway in the permeation pathway of almost all CLC proteins, is mutated to alanine (23). Mutating this gating glutamate in CLC Cl channels strongly affects or even completely suppresses single pore gating (23), whereas CLC exchangers are transformed by such mutations into pure anion conductances that are not coupled to proton transport (17, 19, 20). Another key glutamate, located at the cytoplasmic surface of the CLC monomer, seems to be a hallmark of CLC anion/proton exchangers. Mutating this proton glutamate to nontitratable amino acids uncouples anion transport from protons in the bacterial EcClC-1 protein (27) but seems to abolish transport altogether in mammalian ClC-4 and -5 (21). In those latter proteins, anion transport could be restored by additionally introducing an uncoupling mutation at the gating glutamate (21).The functional complementation by AtClC-c and -d (12, 32) of growth phenotypes of a yeast strain deleted for the single yeast CLC Gef1 (33) suggested that these plant CLC proteins function in anion transport but could not reveal details of their biophysical properties. We report here the first functional expression of a plant CLC in animal cells. Expression of wild-type (WT) and mutant AtClC-a in Xenopus oocytes indicate a general role of gating and proton glutamate residues in anion/proton coupling across different isoforms and species. We identified a proline in the CLC signature sequence of AtClC-a that plays a crucial role in exchange. Mutating it to serine, the residue present in mammalian CLC proteins at this position, rendered AtClC-a Cl/H+ exchange as efficient as exchange. Conversely, changing the corresponding serine of ClC-5 to proline converted it into an efficient exchanger. When proline replaced the critical serine in Torpedo ClC-0, the relative conductance of this model Cl channel was drastically increased, and “fast” protopore gating was slowed.  相似文献   

7.
8.
9.
Protein kinase D (PKD) is a serine/threonine protein kinase rapidly activated by G protein-coupled receptor (GPCR) agonists via a protein kinase C (PKC)-dependent pathway. Recently, PKD has been implicated in the regulation of long term cellular activities, but little is known about the mechanism(s) of sustained PKD activation. Here, we show that cell treatment with the preferential PKC inhibitors GF 109203X or Gö 6983 blocked rapid (1–5-min) PKD activation induced by bombesin stimulation, but this inhibition was greatly diminished at later times of bombesin stimulation (e.g. 45 min). These results imply that GPCR-induced PKD activation is mediated by early PKC-dependent and late PKC-independent mechanisms. Western blot analysis with site-specific antibodies that detect the phosphorylated state of the activation loop residues Ser744 and Ser748 revealed striking PKC-independent phosphorylation of Ser748 as well as Ser744 phosphorylation that remained predominantly but not completely PKC-dependent at later times of bombesin or vasopressin stimulation (20–90 min). To determine the mechanisms involved, we examined activation loop phosphorylation in a set of PKD mutants, including kinase-deficient, constitutively activated, and PKD forms in which the activation loop residues were substituted for alanine. Our results show that PKC-dependent phosphorylation of the activation loop Ser744 and Ser748 is the primary mechanism involved in early phase PKD activation, whereas PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies identify a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.A rapid increase in the synthesis of lipid-derived second messengers with subsequent activation of protein phosphorylation cascades has emerged as a fundamental signal transduction mechanism triggered by multiple extracellular stimuli, including hormones, neurotransmitters, chemokines, and growth factors (1). Many of these agonists bind to G protein-coupled receptors (GPCRs),4 activate heterotrimeric G proteins and stimulate isoforms of the phospholipase C family, including β, γ, δ, and ε (reviewed in Refs. 1 and 2). Activated phospholipase Cs catalyze the hydrolysis of phosphatidylinositol 4,5-bisphosphate to produce the second messengers inositol 1,4,5-trisphosphate and diacylglycerol (DAG). Inositol 1,4,5-trisphosphate mobilizes Ca2+ from intracellular stores (3, 4) whereas DAG directly activates the classic (α, β, and γ) and novel (δ, ε, η, and θ) isoforms of PKC (57). Although it is increasingly recognized that each PKC isozyme has specific functions in vivo (58), the mechanisms by which PKC-mediated signals are propagated to critical downstream targets remain incompletely defined.PKD, also known initially as PKCμ (9, 10), and two recently identified serine protein kinases termed PKD2 (11) and PKCν/PKD3 (12, 13), which are similar in overall structure and primary amino acid sequence to PKD (14), constitute a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase group (15) and separate from the previously identified PKCs (14). Salient features of PKD structure include an N-terminal regulatory region containing a tandem repeat of cysteine-rich zinc finger-like motifs (termed the cysteine-rich domain) that confers high affinity binding to phorbol esters and DAG (9, 16, 17), followed by a pleckstrin homology (PH) domain that negatively regulates catalytic activity (18, 19). The C-terminal region of the PKDs contains its catalytic domain, which is distantly related to Ca2+-regulated kinases.In unstimulated cells, PKD is in a state of low kinase catalytic activity maintained by the N-terminal domain, which represses the catalytic activity of the enzyme by autoinhibition. Consistent with this model, deletions or single amino acid substitutions in the PH domain result in constitutive kinase activity (1820). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (21). In response to cellular stimuli, PKD is converted from a low activity form into a persistently active form that is retained during isolation from cells, as shown by in vitro kinase assays performed in the absence of lipid co-activators (21, 22). PKD activation has been demonstrated in response to engagement of specific GPCRs either by regulatory peptides (2330) or lysophosphatidic acid (27, 31, 32); signaling through Gq, G12, Gi, and Rho (27, 3134); activation of receptor tyrosine kinases, such as the platelet-derived growth factor receptor (23, 35, 36); cross-linking of B-cell receptor and T-cell receptor in B and T lymphocytes, respectively (3740); and oxidative stress (4144).Throughout 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. GF 109203X or Gö 6983) that do not directly inhibit PKD catalytic activity (21, 22), implying that PKD activation in intact cells is mediated, directly or indirectly, through PKCs. In line with this conclusion, cotransfection of PKD with active mutant forms of “novel” PKCs (PKCs δ, ε, η, and θ) resulted in robust PKD activation in the absence of cell stimulation (21, 4446). Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade in response to multiple GPCR agonists in a broad range of cell types, including normal and cancer cells (reviewed in Ref. 14). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as the activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (reviewed in Ref. 14). Collectively, these findings demonstrated the existence of rapidly activated PKC-PKD protein kinase cascade(s) and raised the possibility that some PKC-dependent biological responses involve PKD acting as a downstream effector.PKD has been reported recently to mediate several important cellular activities and processes, including signal transduction (30, 4749), chromatin modification (50), Golgi organization and function (51, 52), c-Jun function (47, 53, 54), NFκB-mediated gene expression (43, 55, 56), and cell survival, migration, and differentiation and DNA synthesis and proliferation (reviewed in Ref. 14). Thus, mounting evidence indicates that PKD has a remarkable diversity of both its signal generation and distribution and its potential for complex regulatory interactions with multiple downstream pathways, leading to multiple responses, including long term cellular events. Despite increasing recognition of its importance, very little is known about the mechanism(s) of sustained PKD activation as opposed to the well documented rapid, PKC-dependent PKD activation.The results presented here demonstrate that prolonged GPCR-induced PKD activation is mediated by sequential PKC-dependent and PKC-independent phases of regulation. We report here, for the first time, that PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies expand previous models of PKD regulation by identifying a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.  相似文献   

10.
Here we show that ouabain-induced cell growth regulation is intrinsically coupled to changes in the cellular amount of Na/K-ATPase via the phosphoinositide 3-kinase (PI3K)/Akt/mammalian target of rapamycin (mTOR) pathway. Ouabain increases the endocytosis and degradation of Na/K-ATPase in LLC-PK1, human breast (BT20), and prostate (DU145) cancer cells. However, ouabain stimulates the PI3K/Akt/mTOR pathway and consequently up-regulates the expression of Na/K-ATPase in LLC-PK1 but not BT20 and DU145 cells. This up-regulation is sufficient to replete the plasma membrane pool of Na/K-ATPase and to stimulate cell proliferation in LLC-PK1 cells. On the other hand, ouabain causes a gradual depletion of Na/K-ATPase and an increased expression of cell cycle inhibitor p21cip, which consequently inhibits cell proliferation in BT20 and DU145 cells. Consistently, we observe that small interfering RNA-mediated knockdown of Na/K-ATPase is sufficient to induce the expression of p21cip and slow the proliferation of LLC-PK1 cells. Moreover, this knockdown converts the growth stimulatory effect of ouabain to growth inhibition in LLC-PK1 cells. Mechanistically, both Src and caveolin-1 are required for ouabain-induced activation of Akt and up-regulation of Na/K-ATPase. Furthermore, inhibition of the PI3K/Akt/mTOR pathway by rapamycin completely blocks ouabain-induced expression of Na/K-ATPase and converts ouabain-induced growth stimulation to growth inhibition in LLC-PK1 cells. Taken together, we conclude that changes in the expression of Na/K-ATPase dictate the growth regulatory effects of ouabain on cells.The Na/K-ATPase, a member of P-type ATPase family, was discovered as an energy transducing ion pump. It transports Na+ and K+ across the cell membrane and maintains ion homeostasis in animal cells (1, 2). Recent studies indicate that the Na/K-ATPase is also an important receptor that can transduce ligand binding into the activation of protein kinase cascades (3). Specifically, the Na/K-ATPase interacts with Src, which provides at least two important cellular regulations (4, 5). First, association with Na/K-ATPase keeps Src in an inactive state. Thus, the Na/K-ATPase serves as a native negative Src regulator (4). Second, this interaction forms a functional receptor complex for cardiotonic steroids (CTS)3 (3), a group of well characterized ligands of the Na/K-ATPase. Cardiotonic steroids include cardenolides (e.g. ouabain) and bufadienolides (e.g. marinobufagenin) (6). Although CTS are known cardiac drugs, some of them have now been identified as endogenous steroid hormones (68). Binding of CTS to the receptor complex activates the Na/K-ATPase-associated Src. Subsequently, the activated Src transactivates other tyrosine kinases, and together they recruit and further phosphorylate multiple membrane and soluble proteins, which results in the activation of protein kinase cascades and the generation of second messengers (3, 4, 6). Ultimately, this chain of signaling events would alter cellular functions and cell growth in a cell-specific manner (5, 912). For instance, we and others have demonstrated that ouabain-induced activation of ERK and PI3K/Akt/mTOR pathways are responsible for cell growth stimulation in transformed cell lines, in primary cultures, as well as in vivo (1318).It has also been recognized for a long time that CTS inhibit cell growth in many cancer cells (1924). Of particular significance are studies that indicate the beneficial effects of CTS therapy in women with breast cancer (2529). Consistently, recent in vitro and in vivo studies have identified several new CTS compounds that exhibit anti-cancer activities (3032). Oleandrin, for example, is in clinical trials in the United States as an anti-cancer remedy for human cancers (31, 33). Although ouabain inhibits the pumping function of the Na/K-ATPase, it is important to note that the growth inhibitory effect of ouabain can occur at doses that neither cause significant changes in intracellular Na+ and K+ nor affect cell viability. Rather, much like its effect on cell growth stimulation, ouabain induces cell growth inhibition through the activation of protein kinases and the generation of second messengers (1923, 34). For example, a recent report showed that these nontoxic concentrations of ouabain stimulated Src, resulting in activation of the epidermal growth factor receptor/ERK pathway and induction of the expression of cell cycle inhibitor p21cip and cell growth arrest (34). Thus, it becomes important to understand the molecular mechanisms that govern different fates of cells in response to CTS stimulation.Prior studies have demonstrated that CTS induce endocytosis of the Na/K-ATPase and regulate its cellular expression via receptor-mediated signal transduction (35, 36). Because the Na/K-ATPase has both pumping and signaling functions, it is conceivable that changes in the amount of cellular Na/K-ATPase could have significant consequences on cell growth. Therefore, we have conducted the following experiments to reveal the role of cellular Na/K-ATPase in ouabain-induced cell growth regulation.  相似文献   

11.
PHLPP2 (PH domain leucine-rich repeat protein phosphatase 2) terminates Akt and protein kinase C (PKC) activity by specifically dephosphorylating these kinases at a key regulatory site, the hydrophobic motif (Ser-473 in Akt1). Here we identify a polymorphism that results in an amino acid change from a Leu to Ser at codon 1016 in the phosphatase domain of PHLPP2, which reduces phosphatase activity toward Akt both in vitro and in cells, in turn resulting in reduced apoptosis. Depletion of endogenous PHLPP2 variants in breast cancer cells revealed the Ser-1016 variant is less functional toward both Akt and PKC. In pair-matched high grade breast cancer samples we observed retention of only the Ser allele from heterozygous patients (identical results were observed in a pair-matched normal and tumor cell line). Thus, we have identified a functional polymorphism that impairs the activity of PHLPP2 and correlates with elevated Akt phosphorylation and increased PKC levels.Breast cancer is diagnosed in ∼180,000 women and is the cause of 40,000 deaths each year in the U.S.2 A prevalent underlying mechanism driving tumorigenesis is aberrant signal transduction pathways that result in constitutive activation of cell growth, proliferation, and survival pathways (2). A well characterized signal transduction pathway in breast cancer that promotes cellular survival, growth, and proliferation is the phosphatidylinositol 3-kinase/Akt pathway (3). This pathway is activated by a number of mechanisms, including gene amplification or gain of function mutations in upstream receptor protein-tyrosine kinases (4, 5), constitutive activation of hormone receptors (6), activating mutations in phosphatidylinositol 3-kinase and Akt (7, 8), and loss of function mutations in the regulatory phosphatase PTEN3 (phosphatase and tensin homolog on chromosome ten) (9). Thus, Akt is a major regulator of breast tumorigenesis.There are three isoforms of Akt present in humans. All three isoforms contain activating phosphorylation sites in the activation loop (Thr-308 in Akt1) and in the C-terminal hydrophobic motif (Ser-473 in Akt1) (10). Upon growth factor receptor stimulation, phosphatidylinositol 3-kinase becomes activated and phosphorylates the D3 position of, typically, phosphatidylinositol (4, 5) bisphosphate to generate phosphatidylinositol (3,4,5)-trisphosphate (11). This 3′-phosphorylated lipid recruits Akt to the plasma membrane by binding to its PH domain, resulting in conformational changes that allow access to the activation loop phosphorylation site (11). Constitutively bound phosphatidylinositol-dependent kinase-1 then phosphorylates Akt at Thr-308, accompanied by phosphorylation at Ser-473 resulting in a catalytically active kinase (12). Phosphorylation of Ser-473 depends on the mTORC2 complex (13-16). Signaling through this pathway is terminated by removal of the lipid second messenger phosphatidylinositol (3,4,5)-trisphosphate catalyzed by the phosphatase PTEN and by direct dephosphorylation of Akt by the recently-identified PHLPP family of phosphatases and protein phosphatase 2A-type phosphatases (17-20).The PHLPP family of phosphatases comprise three variants, the alternatively spliced PHLPP1α and PHLPP1β, and PHLPP2 (21). PHLPP1 and PHLPP2 specifically dephosphorylate the hydrophobic motif of specific Akt isozymes, thus decreasing Akt activity and promoting apoptosis (18, 19). PHLPP2 binds and dephosphorylates Akt1 and Akt3, whereas PHLPP1 binds and dephosphorylates Akt2 and Akt3 (18, 22). Their role in inactivating Akt suggests that both PHLPP1 and PHLPP2 could be potential tumor suppressors. Consistent with such a role, these phosphatases also dephosphorylate the hydrophobic motif of PKC, resulting in degradation of PKC. For this kinase, phosphorylation stabilizes the enzyme, so that the effect of depletion of the PHLPP phosphatases is to increase PKC protein levels (23). PKC is a well characterized oncogene, and loss of function of the PHLPP phosphatases could increase PKC protein levels and promote tumorigenesis (24). Providing further rationale that PHLPP2 could be a potential tumor suppressor, the phosphatase is located on chromosome 16q22.3, a region that encounters frequent loss of heterozygosity (LOH) in many primary and malignant breast tumors (25).Here we identify a non-synonymous polymorphism that results in an amino acid change from a Leu to a Ser at codon 1016 in the PP2C phosphatase domain of PHLPP2. Overexpression studies reveal the Ser-1016 variant has impaired phosphatase activity and is less effective at inducing apoptosis than the Leu-1016 variant. When comparing a pair-matched normal and breast cancer cell line or pair-matched normal and high grade tumor patient samples that are heterozygous, we observe preferential loss of the Leu allele in the tumor tissue or breast cancer cell line. This observation provides evidence that PHLPP2 could be one of the elusive tumor suppressor genes on chromosome 16q, and for heterozygous patients, loss of the more catalytically active Leu-1016 may promote breast tumorigenesis.  相似文献   

12.
Nitric oxide (NO) plays a crucial role in human physiology by regulating vascular tone and blood flow. The short life-span of NO in blood requires a mechanism to retain NO bioactivity in the circulation. Recent studies have suggested a mechanism involving the reduction of nitrite back to NO by deoxyhemoglobin in RBCs. A role for RBCs in transporting NO must, however, bypass the scavenging of NO in RBCs by hemoglobin. To understand how the nitrite reaction can deliver bioactive NO to the vasculature, we have studied the intermediates formed during the reaction. A reliable measure of the total concentration of heme-associated nitrite/NO intermediates formed was provided by combining filtration to measure free nitrite by chemiluminescence and electron paramagnetic resonance to measure the final product Hb(II)NO. By modifying the chemiluminescence method used to detect NO, we have been able to identify two intermediates: 1) a heme-associated nitrite complex that is released as NO in acid solution in the presence of ascorbate and 2) an intermediate that releases NO at neutral pH in the presence of ferricyanide when reacted with an Fe(III) ligand like azide. This species designated as “Hb(II)NO+ ⇆ Hb(III)NO” has properties of both isomeric forms resulting in a slower NO dissociation rate and much higher stability than Hb(III)NO, but provides a potential source for bioactive NO, which can be released from the RBC. This detailed analysis of the nitrite reaction with deoxyHb provides important insights into the mechanism for nitrite induced vasodilation by RBCs.Nitric oxide (NO), also known as the endothelium-derived relaxing factor, is an important messenger molecule involved in the regulation of vascular tone and blood flow (1). The primary source for the synthesis of NO in the circulatory system involves endothelial nitric-oxide synthase (2). This enzyme requires oxygen for the synthesis of NO and is, therefore, less effective in the microcirculation where hypoxic vasodilation regulates the delivery of oxygen. Because nitric oxide has a life-time in blood of <2 ms (3), a mechanism is required to allow for more distal and sustained effects of NO at the reduced oxygen pressures found in the microcirculation. Recent studies have suggested that the bioactivity of NO can be conserved in the blood by the uptake of NO and/or nitrite by red blood cells (RBCs)2 and its interaction with hemoglobin (47). However, any role for the red cell in transporting nitric oxide must be able to avoid the very efficient scavenging of nitric oxide by both oxyhemoglobin (oxyHb) and deoxyhemoglobin (deoxyHb) that destroy and trap NO, respectively, preventing a physiological role for RBC NO.In a series of studies, Stamler and co-workers (710) have hypothesized that NO can bypass this difficulty by being transferred to the β-93 thiol group of hemoglobin (Hb) forming S-nitrosylated hemoglobin (SNO-Hb) when partially heme nitrosylated hemoglobin (Hb(II)NO) is oxygenated. The allosteric quaternary conformational change of hemoglobin at low oxygen pressure destabilizes the β-93 nitrosylated thiol and results in the transfer of NO to membrane thiol groups facilitating the release of the NO to the plasma and the vasculature. However, the extremely low levels of SNO-Hb (11) found in human blood and its instability (12) as a result of intracellular reducing conditions within the RBCs do not support the SNO-Hb hypothesis as the major mechanism for NO transport (1113).The 2003 studies by Rifkind and Gladwin and their collaborators (4, 5, 14, 15) proposed an alternative mechanism that involved the reduction of nitrite, formed by the oxidation of NO, back to NO by a reaction with deoxyHb. Nitrite is present in the blood at fairly high levels (0.1–0.5 μmol/liter) (4, 1618), and it is much more stable than NO or S-nitrosothiols (6), making nitrite an ideal storage pool that can be converted to NO. However, the mechanism by which the NO produced in the red cell by nitrite reduction is exported without being trapped or destroyed is still unclear. Recent studies by Rifkind and co-workers (5, 13, 19) have suggested that the trapping of NO by deoxyHb and/or oxyHb can be bypassed by the formation of a metastable intermediate(s) that retains the NO in a state that is not quenched by reacting with oxyHb or deoxyHb.In this report, we quantitate the two intermediate species that are formed during the reduction of nitrite by deoxyHb when an excess of hemoglobin is present. We also demonstrate that one of the intermediate species designated as “Hb(II)NO+ ⇆ Hb(III)NO” has properties of Hb(II)NO+ and Hb(III)NO, respectively. This species has a slower NO dissociation rate and a much higher stability than Hb(III)NO. This intermediate is a potential source for bioactive NO that can be released from RBCs.  相似文献   

13.
14.
Tetrahydrobiopterin (BH4) is a key redox-active cofactor in endothelial isoform of NO synthase (eNOS) catalysis and is an important determinant of NO-dependent signaling pathways. BH4 oxidation is observed in vascular cells in the setting of the oxidative stress associated with diabetes. However, the relative roles of de novo BH4 synthesis and BH4 redox recycling in the regulation of eNOS bioactivity remain incompletely defined. We used small interference RNA (siRNA)-mediated “knockdown” GTP cyclohydrolase-1 (GTPCH1), the rate-limiting enzyme in BH4 biosynthesis, and dihydrofolate reductase (DHFR), an enzyme-recycling oxidized BH4 (7,8-dihydrobiopterin (BH2)), and studied the effects on eNOS regulation and biopterin metabolism in cultured aortic endothelial cells. Knockdown of either DHFR or GTPCH1 attenuated vascular endothelial growth factor (VEGF)-induced eNOS activity and NO production; these effects were recovered by supplementation with BH4. In contrast, supplementation with BH2 abolished VEGF-induced NO production. DHFR but not GTPCH1 knockdown increased reactive oxygen species (ROS) production. The increase in ROS production seen with siRNA-mediated DHFR knockdown was abolished either by simultaneous siRNA-mediated knockdown of eNOS or by supplementing with BH4. In contrast, addition of BH2 increased ROS production; this effect of BH2 was blocked by BH4 supplementation. DHFR but not GTPCH1 knockdown inhibited VEGF-induced dephosphorylation of eNOS at the inhibitory site serine 116; these effects were recovered by supplementation with BH4. These studies demonstrate a striking contrast in the pattern of eNOS regulation seen by the selective modulation of BH4 salvage/reduction versus de novo BH4 synthetic pathways. Our findings suggest that the depletion of BH4 is not sufficient to perturb NO signaling, but rather that concentration of intracellular BH2, as well as the relative concentrations of BH4 and BH2, together play a determining role in the redox regulation of eNOS-modulated endothelial responses.Regulation of endothelial nitric oxide (NO)2 production represents a critical mechanism for the modulation of vascular homeostasis. NO is released by endothelial cells in response to diverse humoral, neural, and mechanical stimuli (14). Endothelial cell-derived NO activates guanylate cyclase in vascular smooth muscle cells, leading to increased levels of cGMP and to smooth muscle relaxation. Blood platelets represent another key target for the actions of endothelium-derived NO (5): platelet aggregation is inhibited by NO-induced guanylate cyclase activation. Many other effects of NO have been identified in cultured vascular cells and in vascular tissues, including the regulation of apoptosis, cell adhesion, angiogenesis, thrombosis, vascular smooth muscle proliferation, and atherogenesis, among other cellular responses and (patho)physiological processes.The endothelial isoform of NO synthase (eNOS) is a membrane-associated homodimeric 135-kDa protein that is robustly expressed in endothelial cells (2, 4, 6, 7). Similar to all the mammalian NOS isoforms, eNOS functions as an obligate homodimer that includes a cysteine-complex Zn2+ (zinc-tetrathiolate) at the dimer interface (810). eNOS is a Ca2+/calmodulin-dependent enzyme that is activated in response to the stimulation of a variety of Ca2+-mobilizing cell surface receptors in vascular endothelium and in cardiac myocytes. The activity of eNOS is also regulated by phosphorylation at multiple sites (11) that are differentially modulated following the activation of cell surface receptors by agonists such as insulin and vascular endothelial growth factor (VEGF) (12). The phosphorylation of eNOS at Ser-1179 activates eNOS, but phosphorylation at Thr-497 or Ser-116 is associated with inhibition of eNOS activity (1317). eNOS is reversibly targeted to plasmalemmal caveolae as a consequence of the protein''s N-myristoylation and thiopalmitoylation. The generation of NO by eNOS requires several redox-active cofactors, including nicotinamide adenine dinucleotide phosphate (NADPH), flavin adenine dinucleotide (FAD), flavin mononucleotide (FMN), calmodulin, and tetrahydrobiopterin (BH4), which have key roles in the electron flow required for eNOS catalysis. If the flow of electrons within eNOS is disrupted, the enzyme is uncoupled from NO production and other redox-active products are generated, including hydrogen peroxide and superoxide anion radical (18, 19).In vascular disease states such as diabetes, endothelial dysfunction is characterized by a decrease in NO bioactivity and by a concomitant increase in superoxide formation, while eNOS mRNA and protein levels are maintained or even increased. “Uncoupled” eNOS generates reactive oxygen species (ROS), shifting the nitroso-redox balance and having adverse consequences in the vascular wall (20). Several enzymes expressed in vascular tissues contribute to the production and efficient degradation of ROS, and an enhanced activity of oxidant enzymes and/or reduced activity of antioxidant enzymes may cause oxidative stress. Various agonists, pathological conditions, and therapeutic interventions lead to modulated expression and function of oxidant and antioxidant enzymes. However, the intimate relationship between intracellular redox state, eNOS regulation, and NO bioavailability remains incompletely characterized.BH4 is a key redox-active cofactor for activity of all NOS enzymes (21). The exact role of BH4 in NOS catalysis is not yet completely defined, but this cofactor appears to facilitate electron transfer from the eNOS reductase domain and maintains the heme prosthetic group of the enzyme in its redox-active form (18, 22, 23). Moreover, BH4 promotes formation of active NOS homodimers (24) and inhibits the formation of hydrogen peroxide or superoxide by uncoupled eNOS (18, 19). It has been reported that the endothelial dysfunction associated with diabetes is accompanied a decrease in the abundance of bioactive BH4. Supplementation with BH4 has been shown to improve endothelial function in the models of diabetes and hypertension (25, 26, 27). Moreover, BH4 oxidation is seen in vascular cells in the setting of oxidative stress associated with diabetes (28) and hypertension (29).BH4 can be formed either by a de novo biosynthetic pathway or by a salvage pathway. Guanosine triphosphate cyclohydrolase-1 (GTPCH1) catalyzes the conversion of GTP to dihydroneopterin triphosphate. BH4 is generated by further steps catalyzed by 6-pyruvoyltetrahydropterin synthase and sepiapterin reductase (30). GTPCH1 appears to be the rate-limiting enzyme in BH4 biosynthesis; overexpression of GTPCH1 is sufficient to augment BH4 levels in cultured endothelial cells (31). On the other hand, dihydrofolate reductase (DHFR) catalyzes the regeneration of BH4 from its oxidized form, 7,8-dihydrobiopterin (BH2), in several cell types (30, 32). DHFR is mainly involved in folate metabolism and converts inactive BH2 back to BH4 and plays an important role in the metabolism of exogenously administered BH4. However, the relative contributions of endothelial GTPCH1 and DHFR to the modulation of eNOS-dependent pathways are incompletely understood.In these studies, we have used siRNA-mediated “knockdown” of GTPCH1 and DHFR to explore the relative roles of BH4 synthesis and recycling in the modulation of eNOS bioactivity, as well as in the regulation of NO-dependent signaling pathways in endothelial cells.  相似文献   

15.
16.
In Alzheimer disease (AD), frontotemporal dementia and parkinsonism linked to chromosome 17 (FTDP-17) and other tauopathies, tau accumulates and forms paired helical filaments (PHFs) in the brain. Tau isolated from PHFs is phosphorylated at a number of sites, migrates as ∼60-, 64-, and 68-kDa bands on SDS-gel, and does not promote microtubule assembly. Upon dephosphorylation, the PHF-tau migrates as ∼50–60-kDa bands on SDS-gels in a manner similar to tau that is isolated from normal brain and promotes microtubule assembly. The site(s) that inhibits microtubule assembly-promoting activity when phosphorylated in the diseased brain is not known. In this study, when tau was phosphorylated by Cdk5 in vitro, its mobility shifted from ∼60-kDa bands to ∼64- and 68-kDa bands in a time-dependent manner. This mobility shift correlated with phosphorylation at Ser202, and Ser202 phosphorylation inhibited tau microtubule-assembly promoting activity. When several tau point mutants were analyzed, G272V, P301L, V337M, and R406W mutations associated with FTDP-17, but not nonspecific mutations S214A and S262A, promoted Ser202 phosphorylation and mobility shift to a ∼68-kDa band. Furthermore, Ser202 phosphorylation inhibited the microtubule assembly-promoting activity of FTDP-17 mutants more than of WT. Our data indicate that FTDP-17 missense mutations, by promoting phosphorylation at Ser202, inhibit the microtubule assembly-promoting activity of tau in vitro, suggesting that Ser202 phosphorylation plays a major role in the development of NFT pathology in AD and related tauopathies.Neurofibrillary tangles (NFTs)4 and senile plaques are the two characteristic neuropathological lesions found in the brains of patients suffering from Alzheimer disease (AD). The major fibrous component of NFTs are paired helical filaments (PHFs) (for reviews see Refs. 13). Initially, PHFs were found to be composed of a protein component referred to as “A68” (4). Biochemical analysis reveled that A68 is identical to the microtubule-associated protein, tau (4, 5). Some characteristic features of tau isolated from PHFs (PHF-tau) are that it is abnormally hyperphosphorylated (phosphorylated on more sites than the normal brain tau), does not bind to microtubules, and does not promote microtubule assembly in vitro. Upon dephosphorylation, PHF-tau regains its ability to bind to and promote microtubule assembly (6, 7). Tau hyperphosphorylation is suggested to cause microtubule instability and PHF formation, leading to NFT pathology in the brain (13).PHF-tau is phosphorylated on at least 21 proline-directed and non-proline-directed sites (8, 9). The individual contribution of these sites in converting tau to PHFs is not entirely clear. However, some sites are only partially phosphorylated in PHFs (8), whereas phosphorylation on specific sites inhibits the microtubule assembly-promoting activity of tau (6, 10). These observations suggest that phosphorylation on a few sites may be responsible and sufficient for causing tau dysfunction in AD.Tau purified from the human brain migrates as ∼50–60-kDa bands on SDS-gel due to the presence of six isoforms that are phosphorylated to different extents (2). PHF-tau isolated from AD brain, on the other hand, displays ∼60-, 64-, and 68 kDa-bands on an SDS-gel (4, 5, 11). Studies have shown that ∼64- and 68-kDa tau bands (the authors have described the ∼68-kDa tau band as an ∼69-kDa band in these studies) are present only in brain areas affected by NFT degeneration (12, 13). Their amount is correlated with the NFT densities at the affected brain regions. Moreover, the increase in the amount of ∼64- and 68-kDa band tau in the brain correlated with a decline in the intellectual status of the patient. The ∼64- and 68-kDa tau bands were suggested to be the pathological marker of AD (12, 13). Biochemical analyses determined that ∼64- and 68-kDa bands are hyperphosphorylated tau, which upon dephosphorylation, migrated as normal tau on SDS-gel (4, 5, 11). Tau sites involved in the tau mobility shift to ∼64- and 68-kDa bands were suggested to have a role in AD pathology (12, 13). It is not known whether phosphorylation at all 21 PHF-sites is required for the tau mobility shift in AD. However, in vitro the tau mobility shift on SDS-gel is sensitive to phosphorylation only on some sites (6, 14). It is therefore possible that in the AD brain, phosphorylation on some sites also causes a tau mobility shift. Identification of such sites will significantly enhance our knowledge of how NFT pathology develops in the brain.PHFs are also the major component of NFTs found in the brains of patients suffering from a group of neurodegenerative disorders collectively called tauopathies (2, 11). These disorders include frontotemporal dementia and Parkinsonism linked to chromosome 17 (FTDP-17), corticobasal degeneration, progressive supranuclear palsy, and Pick disease. Each PHF-tau isolated from autopsied brains of patients suffering from various tauopathies is hyperphosphorylated, displays ∼60-, 64-, and 68-kDa bands on SDS-gel, and is incapable of binding to microtubules. Upon dephosphorylation, the above referenced PHF-tau migrates as a normal tau on SDS-gel, binds to microtubules, and promotes microtubule assembly (2, 11). These observations suggest that the mechanisms of NFT pathology in various tauopathies may be similar and the phosphorylation-dependent mobility shift of tau on SDS-gel may be an indicator of the disease. The tau gene is mutated in familial FTDP-17, and these mutations accelerate NFT pathology in the brain (1518). Understanding how FTDP-17 mutations promote tau phosphorylation can provide a better understanding of how NFT pathology develops in AD and various tauopathies. However, when expressed in CHO cells, G272V, R406W, V337M, and P301L tau mutations reduce tau phosphorylation (19, 20). In COS cells, although G272V, P301L, and V337M mutations do not show any significant affect, the R406W mutation caused a reduction in tau phosphorylation (21, 22). When expressed in SH-SY5Y cells subsequently differentiated into neurons, the R406W, P301L, and V337M mutations reduce tau phosphorylation (23). In contrast, in hippocampal neurons, R406W increases tau phosphorylation (24). When phosphorylated by recombinant GSK3β in vitro, the P301L and V337M mutations do not have any effect, and the R406W mutation inhibits phosphorylation (25). However, when incubated with rat brain extract, all of the G272V, P301L, V337M, and R406W mutations stimulate tau phosphorylation (26). The mechanism by which FTDP-17 mutations promote tau phosphorylation leading to development of NFT pathology has remained unclear.Cyclin-dependent protein kinase 5 (Cdk5) is one of the major kinases that phosphorylates tau in the brain (27, 28). In this study, to determine how FTDP-17 missense mutations affect tau phosphorylation, we phosphorylated four FTDP-17 tau mutants (G272V, P301L, V337M, and R406W) by Cdk5. We have found that phosphorylation of tau by Cdk5 causes a tau mobility shift to ∼64- and 68 kDa-bands. Although the mobility shift to a ∼64-kDa band is achieved by phosphorylation at Ser396/404 or Ser202, the mobility shift to a 68-kDa band occurs only in response to phosphorylation at Ser202. We show that in vitro, FTDP-17 missense mutations, by promoting phosphorylation at Ser202, enhance the mobility shift to ∼64- and 68-kDa bands and inhibit the microtubule assembly-promoting activity of tau. Our data suggest that Ser202 phosphorylation is the major event leading to NFT pathology in AD and related tauopathies.  相似文献   

17.
18.
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   

19.
Protein-disulfide isomerase (PDI), an endoplasmic reticulum (ER)-resident protein, is primarily known as a catalyst of oxidative protein folding but also has a protein unfolding activity. We showed previously that PDI unfolds the cholera toxin A1 (CTA1) polypeptide to facilitate the ER-to-cytosol retrotranslocation of the toxin during intoxication. We now provide insight into the mechanism of this unfoldase activity. PDI includes two redox-active (a and a′) and two redox-inactive (b and b′) thioredoxin-like domains, a linker (x), and a C-terminal domain (c) arranged as abbxac. Using recombinant PDI fragments, we show that binding of CTA1 by the continuous PDIbbxa′ fragment is necessary and sufficient to trigger unfolding. The specific linear arrangement of bbxa′ and the type a domain (aversus a) C-terminal to bbx are additional determinants of activity. These data suggest a general mechanism for the unfoldase activity of PDI: the concurrent and specific binding of bbxa′ to particular regions along the CTA1 molecule triggers its unfolding. Furthermore, we show the bb′ domains of PDI are indispensable to the unfolding reaction, whereas the function of its a′ domain can be substituted partially by the a′ domain from ERp57 (abbxac) or ERp72 (ca°abbxa′), PDI-like proteins that do not unfold CTA1 normally. However, the bb′ domains of PDI were insufficient to convert full-length ERp57 into an unfoldase because the a domain of ERp57 inhibited toxin binding. Thus, we propose that generating an unfoldase from thioredoxin-like domains requires the bb′(x) domains of PDI followed by an a′ domain but not preceded by an inhibitory a domain.Protein-disulfide isomerase (PDI)2 is a multifunctional protein that resides in the endoplasmic reticulum (ER) lumen of all eukaryotic cells (reviewed in Ref. 1). Mammalian PDI was first identified as a catalyst of oxidative protein folding (2), but it is now also known to mediate viral infection (3, 4), antigen processing (5), collagen assembly (6), and ER-associated degradation (79). To participate in this variety of cellular processes, PDI performs multiple activities. For example, during oxidative protein folding, PDI catalyzes the oxidation and isomerization of disulfide bonds and induces conformational changes in non-native polypeptides (10). Independently of redox chemistry, PDI is a molecular chaperone, binding polypeptides to prevent their aggregation (1113). PDI also acts as a structural subunit of the prolyl 4-hydroxylase (P4H) and microsomal triglyceride transfer protein complexes; however, this function is similar to its chaperone activity (1419). In contrast to its protein folding activities, PDI unfolds the catalytic A1 polypeptide of cholera toxin (CTA1) in preparation for the retrotranslocation of the toxin from the ER lumen into the cytosol (8, 20).Cholera toxin (CT) is a pathogenic factor that causes secretory diarrhea in animals (reviewed in Ref. 21). The holotoxin includes a single catalytic A subunit (CTA) and a homopentameric B subunit (CTB) joined noncovalently (22). Upon secretion from the bacterium Vibrio cholerae, CTA is cleaved into the A1 and A2 polypeptides, which are joined by a disulfide bond and noncovalent interactions (22, 23). To intoxicate a cell, CTB binds the ganglioside GM1 on the surface of the cell, and the holotoxin is transported in a retrograde manner to the ER lumen (24). In the ER, CTA is reduced to generate CTA1, and PDI unfolds and dissociates CTA1 from the holotoxin (20). The unfolded toxin is subsequently transported across the ER membrane (25, 26). Upon reaching the cytosol, CTA1 refolds and induces toxicity (27, 28).We showed previously that PDI acts as a redox-dependent chaperone to unfold CTA1 (20). In the reduced state of PDI, it binds and unfolds the toxin. Subsequent oxidation of PDI by ER oxidase 1 causes PDI to release unfolded CTA1 (25). Aside from this information, nothing is known about the mechanism of the unfolding activity of PDI.PDI is a modular protein comprising two a-type thioredoxin-like domains (a and a′), two b-type thioredoxin-like domains (b and b′), a flexible linker (x), and an extended C-terminal domain (c) arranged as abbxac (2931). The a-type domains are characterized by the presence of the catalytic sequence CXXC and are therefore redox-active, whereas the b-type domains lack this sequence and are redox-inactive (32). The thioredoxin-like domains of PDI differ from each other in primary structure despite having a common fold. The crystal structure of yeast PDI shows the bb′ domains form a rigid base from which the a-type domains extend like flexible arms (33, 34). This base is thought to be the core of a substrate-binding groove formed by all four thioredoxin-like domains (30, 33, 35).To understand the mechanism of the unfoldase activity of PDI, we analyzed the contribution of each domain to the ability of PDI to bind and unfold CTA1 using recombinant PDI fragments. Unfolded CTA1 was detected by an established in vitro trypsin sensitivity assay that relies on tryptic cleavage sites hidden in the folded toxin to be exposed in the unfolded toxin (20). Because CTA1 likely mimics a misfolded host cell protein for its recognition and unfolding by PDI (22, 36, 37), this study has implications for how PDI unfolds endogenous misfolded proteins in preparation for their retrotranslocation and subsequent ER-associated degradation.There are nearly 20 mammalian PDI-like proteins, characterized by the presence of one or more thioredoxin-like domains and ER localization (reviewed in Refs. 38, 39). We previously demonstrated that two PDI-like proteins, ERp72 and ERp57, do not facilitate CTA1 retrotranslocation (8). In contrast to PDI, ERp72 retains CTA1 in the ER and either stabilizes its native conformation or renders it more compact (8). To understand how these structurally homologous proteins are functionally unique, we tested whether the various thioredoxin-like domains of ERp57 and ERp72 could functionally replace the corresponding PDI domains to unfold CTA1. Thus, in addition to suggesting a general mechanism for the unfoldase activity of PDI, our data indicate functional similarities and differences among thioredoxin-like domains of PDI family proteins.  相似文献   

20.
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