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 15 毫秒
1.
BACKGROUND INFORMATION: The c-Met-dependent, beta-actin-rich, blebbed pseudopodia of MSV-MDCK-INV (invasive Moloney-sarcoma-virus-transformed Madin-Darby canine kidney) cells are induced by Rho/ROCK (Rho kinase) activation, and are morphologically distinct from flat extended lamellipodia. RESULTS: Microtubules were shown to extend to these actin-rich pseudopodial domains, and microtubule depolymerization by nocodazole treatment resulted in progressive cellular blebbing, initiating in the pseudopodial domains and resulting in transient cellular rounding and blebbing after 30 min. The blebbing response was dependent on autocrine HGF (hepatocyte growth factor) activation of c-Met and prevented by inhibition of RhoA, ROCK and p38 MAPK (p38 mitogen-activated protein kinase), but not ERK (extracellular-signal-regulated kinase) or PI3K (phosphoinositide 3-kinase). Phospho-p38 MAPK was present in pseudopodia, localizing activation of this signalling pathway to this protrusive membrane structure. In serum-starved cells, LPA (lysophosphatidic acid) activation of RhoA induced p38 MAPK-dependent pseudopodial protrusions, and inhibition of p38 MAPK prevented pseudopodial protrusion and displacement of MSV-MDCK-INV cells. MSV-MDCK-INV cells exhibited intermittent blebbing and rounding, which may represent an integral part of their motile behaviour. CONCLUSIONS: The localized activation of an autocrine HGF/c-Met loop regulates Rho/ROCK activation of p38 MAPK signalling to stimulate both membrane blebbing and pseudopod formation.  相似文献   

2.
Interaction of glycolytic enzymes with F-actin is suggested to be a mechanism for compartmentation of the glycolytic pathway. Earlier work demonstrates that muscle F-actin strongly binds glycolytic enzymes, allowing for the general conclusion that "actin binds enzymes", which may be a generalized phenomenon. By taking actin from a lower form, such as yeast, which is more deviant from muscle actin than other higher animal forms, the generality of glycolytic enzyme interactions with actin and the cytoskeleton can be tested and compared with higher eukaryotes, e.g., rabbit muscle. Cosedimentation of rabbit skeletal muscle and yeast F-actin with muscle fructose-1,6-bisphosphate aldolase (aldolase) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) followed by Scatchard analysis revealed a biphasic binding, indicating high- and low-affinity domains. Muscle aldolase and GAPDH showed low-affinity for binding yeast F-actin, presumably because of fewer acidic residues at the N-terminus of yeast actin; this difference in affinity is also seen in Brownian dynamics computer simulations. Yeast GAPDH and aldolase showed low-affinity binding to yeast actin, which suggests that actin-glycolytic enzyme interactions may also occur in yeast although with lower affinity than in higher eukaryotes. The cosedimentation results were supported by viscometry results that revealed significant cross-linking at lower concentrations of rabbit muscle enzymes than yeast enzymes. Brownian dynamics simulations of yeast and muscle aldolase and GAPDH with yeast and muscle actin compared the relative association free energy. Yeast aldolase did not specifically bind to either yeast or muscle actin. Yeast GAPDH did bind to yeast actin although with a much lower affinity than when binding muscle actin. The binding of yeast enzymes to yeast actin was much less site specific and showed much lower affinities than in the case with muscle enzymes and muscle actin.  相似文献   

3.
The pseudopodial protrusions of Moloney sarcoma virus (MSV)-Madin-Darby canine kidney (MDCK)-invasive (INV) variant cells were purified on 1-microm pore polycarbonate filters that selectively allow passage of the pseudopodial domains but not the cell body. The purified pseudopodial fraction contains phosphotyrosinated proteins, including Met and FAK, and various signaling proteins, including Raf1, MEK1, ERK2, PKBalpha (Akt1), GSK3alpha, GSK3beta, Rb, and Stat3. Pseudopodial proteins identified by liquid chromatography tandem mass spectrometry included actin and actin-regulatory proteins (ERM, calpain, filamin, myosin, Sra-1, and IQGAP1), tubulin, vimentin, adhesion proteins (vinculin, talin, and beta1 integrin), glycolytic enzymes, proteins associated with protein translation, RNA translocation, and ubiquitin-mediated protein degradation, as well as protein chaperones (HSP90 and HSC70) and signaling proteins (RhoGDI and ROCK). Inhibitors of MEK1 (U0126) and HSP90 (geldanamycin) significantly reduced MSV-MDCK-INV cell motility and pseudopod expression, and geldanamycin treatment inhibited Met phosphorylation and induced the expression of actin stress fibers. ROCK inhibition did not inhibit cell motility but transformed the pseudopodial protrusions of MSV-MDCK-INV cells into extended lamellipodia. Dominant negative Rho disrupted pseudopod expression and, in serum-starved cells, L-alpha-lysophosphatidic acid (oleoyl) activation of Rho induced pseudopodial protrusions or, in the presence of the ROCK inhibitor, extended lamellipodia. RNA was localized to the actin-rich pseudopodial domains of MSV-MDCK-INV cells, but the extent of colocalization with dense actin ruffles was reduced in the extended lamellipodia formed upon ROCK inhibition. Rho/ROCK activation in epithelial tumor cells therefore regulates RNA translocation to a pseudopodial domain that contains proteins involved in signaling, cytoskeleton remodeling, cell adhesion, glycolysis, and protein translation and degradation.  相似文献   

4.
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) has previously been suggested to have almost absolute control over the glycolytic flux in Lactococcus lactis (B. Poolman, B. Bosman, J. Kiers, and W. N. Konings, J. Bacteriol. 169:5887-5890, 1987). Those studies were based on inhibitor titrations with iodoacetate, which specifically inhibits GAPDH, and the data suggested that it should be possible to increase the glycolytic flux by overproducing GAPDH activity. To test this hypothesis, we constructed a series of mutants with GAPDH activities from 14 to 210% of that of the reference strain MG1363. We found that the glycolytic flux was unchanged in the mutants overproducing GAPDH. Also, a decrease in the GAPDH activity had very little effect on the growth rate and the glycolytic flux until 25% activity was reached. Below this activity level, the glycolytic flux decreased proportionally with decreasing GAPDH activity. These data show that GAPDH activity has no control over the glycolytic flux (flux control coefficient = 0.0) at the wild-type enzyme level and that the enzyme is present in excess capacity by a factor of 3 to 4. The early experiments by Poolman and coworkers were performed with cells resuspended in buffer, i.e., nongrowing cells, and we therefore analyzed the control by GAPDH under similar conditions. We found that the glycolytic flux in resting cells was even more insensitive to changes in the GAPDH activity; in this case GAPDH was also present in a large excess and had no control over the glycolytic flux.  相似文献   

5.
In MDCK epithelial cells, cell contact at confluency initiates a protracted process of morphogenesis during which several proteins known to bind the cytoskeleton become progressively associated with the detergent-resistant cell fraction and distributed to their characteristic polarized domains. Using extraction protocols that identify this tight cytoskeletal linkage, here we show a similar but slower, time-dependent enrichment in the detergent resistant fraction of glyceraldehyde-3-phosphate dehydrogenase (GAPDH), a highly abundant glycolytic enzyme that is traditionally considered soluble. Similar enrichment did not occur for two other glycolytic enzymes, phosphoglycerate mutase or lactate dehydrogenase. Insoluble GAPDH was not homogeneously distributed in the cytoplasm but rather displayed several discrete patterns that varied within and among MDCK cells. It also localized prominently to a few nuclei in the phenotypically heterogeneous cells of late confluency cultures. Disruptors of cytoskeletal filaments were relatively ineffective in the postconfluent epithelial monolayers, although use of disrupting agents implicated actin as the cytoplasmic filament that tethers insoluble GAPDH. Catalytic activity could be demonstrated in the insoluble fraction of GAPDH from postconfluent cultures, but only after release by mechanical disruption of insoluble extracts. Treatment of postconfluent cells with agents that deplete ATP diminished the fraction of cytoskeletally associated GAPDH, and levels of insoluble GAPDH were restored with ATP repletion, suggesting that ATP levels may regulate cytoskeletal linkage and thereby local enzyme activity. We conclude that the highly abundant and ubiquitous enzyme GAPDH becomes progressively enriched in detergent stable subcellular compartments during the process of epithelial morphogenesis. The process that produces GAPDH compartments is slow, suggesting that epithelial cells just at confluency, when they are typically analyzed, have not yet maximized the organizational state that can be attained in monolayer culture.  相似文献   

6.
The multiple beta-actin rich pseudopodial protrusions of the invasive variant of Moloney sarcoma virus (MSV)-transformed epithelial MDCK cells (MSV-MDCK-INV) are strongly labeled for phosphotyrosine. Increased tyrosine phosphorylation among a number of proteins was detected in MSV-MDCK-INV cells relative to untransformed and MSV-transformed MDCK cells, especially for the hepatocyte growth factor receptor (HGF-R), otherwise known as c-met proto-oncogene. Cell surface expression of HGF-R was similar in the three cell lines, indicating that HGF-R is constitutively phosphorylated in MSV-MDCK-INV cells. Both the tyrosine kinase inhibitor herbimycin A and the HGFalpha antibody abolished HGF-R phosphorylation, induced retraction of pseudopodial protrusions, and promoted the establishment of cell-cell contacts as well as the apparition of numerous stabilizing stress fibers in MSV-MDCK-INV cells. Furthermore, anti-HGFalpha antibody abolished cell motility among MSV-MDCK-INV cells. Conditioned medium from MSV-MDCK-INV cells induced MDCK cell scattering, indicating that HGF is secreted by MSV-MDCK-INV cells. HGF titration followed by a subsequent washout of the antibodies led to renewed pseudopodial protrusion and cellular movement. We therefore show that activation of the tyrosine kinase activity of HGF-R/Met via an autocrine HGF loop is directly responsible for pseudopodial protrusion, thereby explaining the motile and invasive potential of this model epithelium-derived tumor cell line.  相似文献   

7.
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is a multifunctional protein as well as a classic glycolytic enzyme, and its pleiotropic functions are achieved by various post-translational modifications and the resulting translocations to intracellular compartments. In the present study, GAPDH in the plasma membrane of BeWo choriocarcinoma cells displayed a striking acidic shift in two-dimensional electrophoresis after cell-cell fusion induction by forskolin. This post-translational modification was deamidation of multiple glutaminyl residues, as determined by molecular mass measurement and tandem mass spectrometry of acidic GAPDH isoforms. Transglutaminase (TG) inhibitors prevented this acidic shift and reduced cell fusion. Knockdown of the TG2 gene by short hairpin RNA reproduced these effects of TG inhibitors. Various GAPDH mutants with replacement of different numbers (one to seven) of Gln by Glu were expressed in BeWo cells. These deamidated mutants reversed the suppressive effect of wild-type GAPDH overexpression on cell fusion. Interestingly, the mutants accumulated in the plasma membrane, and this accumulation was increased according to the number of Gln/Glu substitutions. Considering that GAPDH binds F-actin via an electrostatic interaction and that the cytoskeleton is rearranged in trophoblastic cell fusion, TG2-dependent GAPDH deamidation was suggested to participate in actin cytoskeletal remodeling.  相似文献   

8.
The in vitro interaction between the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and cytoskeletal elements is well documented. To verify this association within cells, the intracellular distribution of GAPDH under various metabolic conditions has been investigated in immunostained cells or cells expressing GAPDH as a GFP fusion protein. GAPDH was homogeneously distributed in the cytoplasm and no interaction of GAPDH with cytoskeletal elements, neither with microfilaments nor microtubules or intermediate filaments, was detectable. In living cells expressing GFP-GAPDH, stress fibres were excluded from the fluorescence. In contrast to proliferating cells, the cytoplasmic GAPDH of serum-depleted cells was not homogeneously distributed, but colocalised with stress fibres. The mechanism for stimulating this actin-binding affinity was independent of the NO-signalling pathway. The results support the idea of a specialised function for the interaction of GAPDH and cytoskeletal elements, rather than a general function, as e.g. microcompartmentalization of glycolytic enzymes.  相似文献   

9.
1.To study proteins transported with actin in axons, we pulse-labeled motoneurons in the chicken sciatic nerve with [35S]methionine and, 1–20 days later, isolated actin and its binding proteins by affinity chromatography of Triton soluble nerve extracts on DNase I–Sepharose. The DNase I-purified proteins were electrophoresed on two-dimensional gels and the specific activity of the radioactively labeled protein spots was estimated by fluorography.2.In addition to actin, which binds specifically to DNase I, a small number of other proteins were labeled, including established actin monomer binding proteins and a protein of 36 kDa and pI 8.5. On the basis of its molecular mass, pI, amino acid composition, and immunostaining, the unrecognized protein was identified as the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH).3.The high-affinity binding of GAPDH to actin was confirmed by incubation of Triton-soluble nerve extracts with either mouse anti-GAPDH (or antiactin) and indirect immunomagnetic separation with Dynabeads covalently linked to sheep anti-mouse antibody. Analysis by one-dimensional gel electrophoresis and immunoblotting showed that actin and GAPDH were the main proteins isolated by these methods.4.Analysis of labeled nerves at 12 and 20 days after pulse labeling showed that GAPDH and actin were transported at the same rate, i.e., 3–5 mm/day, which corresponds to slow component b of axonal transport. These proteins were not associated with rapidly transported proteins that accumulated proximal to a ligation 7 cm from the spinal cord 9 hr after injection of radioactivity.5.Our results indicate that GAPDH and actin are transported as a complex in axons and raise the possibility that GAPDH could act as a chaperone for monomeric actin, translocating it to intraaxonal sites for exchange with or assembly into actin filaments. Alternatively, actin could be involved in translocating and anchoring GAPDH to specialized sites in axons and nerve terminals that require a source of ATP by glycolysis.  相似文献   

10.
Microtubule-associated protein 1B, MAP1B, is a major cytoskeletal protein during brain development and one of the largest brain MAPs associated with microtubules and microfilaments. Here, we identified several proteins that bind to MAP1B via immunoprecipitation with a MAP1B-specific antibody, by one and two-dimensional gel electrophoresis and subsequent mass spectrometry identification of precipitated proteins. In addition to tubulin and actin, a variety of proteins were identified. Among these proteins were glyceraldehyde-3-phosphate dehydrogenase (GAPDH), heat shock protein 8, dihydropyrimidinase related proteins 2 and 3, protein-L-isoaspartate O-methyltransferase, beta-spectrin, and clathrin protein MKIAA0034, linking either directly or indirectly to MAP1B. In particular, GAPDH, a key glycolytic enzyme, was bound in large quantity to the heavy chain of MAP1B in adult brain tissue. In vitro binding studies confirmed a direct binding of GAPDH to MAP1B. In PC12 cells, GAPDH was found in cytoplasm and nuclei and partially co-localized with MAP1B. It disappeared from the cytoplasm under oxidative stress or after a disruption of cytoskeletal elements after colcemid or cytochalasin exposure. GAPDH may be essential in the local energy provision of cytoskeletal structures and MAP1B may help to keep this key enzyme close to the cytoskeleton.  相似文献   

11.
The putative tumor metastasis suppressor protein Nm23-H1 is a nucleoside diphosphate kinase that exhibits a novel protein kinase activity when bound to glyceraldehyde-3-phosphate dehydrogenase (GAPDH). In this study we show that the glycolytic enzyme phosphoglycerate mutase B (PGM) becomes phosphorylated in the presence of the Nm23-H1.GAPDH complex in vitro. Mutation of His-10 in PGM abolishes the Nm23-H1.GAPDH complex-induced phosphorylation. Nm23-H1, GAPDH, and PGM are known to co-localize as shown by free flow isoelectric focusing. In association with Nm23-H1 and GAPDH, PGM could be activated by dCTP, which is a substrate of Nm23-H1, in addition to the well known PGM activator 2,3-bisphosphoglycerate. A synthetic cell-penetrating peptide (PGMtide) encompassing the phosphorylated histidine and several residues from PGM (LIRHGE) promoted growth arrest of several tumor cell lines, whereas proliferation of tested non-tumor cells was not influenced. Analysis of metabolic activity of one of the tumor cell lines, MCF-7, indicated that PGMtide inhibited glycolytic flux, consistent with in vivo inhibition of PGM. The specificity of the observed effect was further determined experimentally by testing the effect of PGMtide on cells growing in the presence of pyruvate, which helps to compensate PGM inhibition in the glycolytic pathway. Thus, growth of MCF-7 cells was not arrested by PGMtide in the presence of pyruvate. The data presented here provide evidence that inhibition of PGM activity can be achieved by exogenous addition of a polypeptide, resulting in inhibition of glycolysis and cell growth arrest in cell culture.  相似文献   

12.
During heat shock, Xenopus laevis embryos exhibit an increase in the rate of accumulation of lactate and a loss of ATP relative to non-heat-shocked control embryos. These results suggest that heat shock stimulates a shift in energy metabolism to anaerobic glycolysis while at the same time causing an increase in the demand for ATP. We have evidence indicating that the embryo may meet such demands placed on it by increasing the levels of some glycolytic enzymes. In this report, we show that heat shock stimulates increases in the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase [( EC 1.2.1.12] GAPDH). The specific activity of GAPDH shows a significant increase after heat shock, which correlates with the accumulation of GAPDH in heat-shocked embryos as detected by immunoblotting. Increases in GAPDH-specific activity are variable, however, and are inversely proportional to the levels of specific activity in control embryos; i.e., constitutive enzyme activity. We further analyzed the heat-enhanced accumulation of GAPDH by electrophoretically separating GAPDH isozymes on nondenaturing polyacrylamide gels. Control embryos exhibit a single isozyme of GAPDH, whereas heat-shocked embryos exhibit two isozymes of GAPDH. When these isozymes are labeled with [35S]methionine, separated by nondenaturing gel electrophoresis, and analyzed by fluorography, a heat-shock protein is found to comigrate with the isozyme unique to the heat-shocked sample. Enzyme activity assays at different temperatures suggest that this isozyme has optimum enzymatic activity only at heat-shock temperatures. We have correlated a 35-kD heat-shock protein (hsp35) with GAPDH using the following evidence: this hsp comigrates with GAPDH on one-dimensional SDS polyacrylamide gels; heat-enhanced increases in GAPDH specific activity correlate with hsp35 synthesis; and hsp35 and GAPDH have similar peptide maps. This relationship also provides a compelling explanation for the restriction of hsp35 synthesis to the vegetal hemisphere cells of heat-shocked early gastrulae reported previously (Nickells, R. W., and L. W. Browder. 1985. Dev. Biol. 112:391-395).  相似文献   

13.
Photosynthetic glyceraldehyde-3-phosphate dehydrogenase (GAPDH) of Spinacia oleracea belongs to a wide group of GAPDHs found in most organisms displaying oxygenic photosynthesis, including cyanobacteria, green and red algae, and higher plants. As a major catalytic difference with respect to glycolytic GAPDH, photosynthetic GAPDH exhibits dual cofactor specificity toward pyridine nucleotides with a preference for NADP(H). Here we report the crystal structure of NAD-complexed recombinant A(4)-GAPDH (NAD-A(4)-GAPDH) from Spinacia oleracea, expressed in Escherichia coli. Its superimposition onto native A(4)-GAPDH complexed with NADP (NADP-A(4)-GAPDH) pinpoints specific conformational changes resulting from cofactor replacement. In photosynthetic NAD-A(4)-GAPDH, the side chain of Asp32 is oriented toward the coenzyme to interact with the adenine ribose diol, similar to glycolytic GAPDHs (NAD-specific). On the contrary, in NADP-A(4)-GAPDH Asp32 moves away to accommodate the additional 2'-phosphate group of the coenzyme and to minimize electrostatic repulsion. Asp32 rotation is allowed by the presence of the small residue Ala40, conserved in most photosynthetic GAPDHs, replacing bulky amino acid side chains in glycolytic GAPDHs. While in NADP-A(4)-GAPDH two amino acids, Thr33 and Ser188, are involved in hydrogen bonds with the 2'-phosphate group of NADP, in the NAD-complexed enzyme these interactions are lacking. The crystallographic structure of NAD-A(4)-GAPDH highlights that four residues, Thr33, Ala40, Ser188, and Ala187 (Leu, Leu, Pro, and Leu respectively, in glycolytic Bacillus stearothermophilus GAPDH sequence) are of primary importance for the dual cofactor specificity of photosynthetic GAPDH. These modifications seem to trace the minimum evolutionary route for a primitive NAD-specific GAPDH to be converted into the NADP-preferring enzyme of oxygenic photosynthetic organisms.  相似文献   

14.
Rab2 requires atypical protein kinase C iota/lambda (aPKC iota/lambda) to promote vesicle formation from vesicular tubular clusters (VTCs). The Rab2-generated vesicles are enriched in recycling proteins suggesting that the carriers are retrograde-directed and retrieve transport machinery back to the endoplasmic reticulum. These vesicles also contained the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH). We have previously established that GAPDH is required for membrane transport between the endoplasmic reticulum and the Golgi complex. Moreover, GAPDH is phosphorylated by aPKC iota/lambda and binds to the aPKC iota/lambda regulatory domain. In this study, we employed a combination of in vivo and in vitro assays and determined that GAPDH also interacts with Rab2. The site of GAPDH interaction was mapped to Rab2 residues 20-50. In addition to its glycolytic function, GAPDH has multiple intracellular roles. However, the function of GAPDH in the early secretory pathway is unknown. One possibility is that GAPDH ultimately provides energy in the form of ATP. To determine whether GAPDH catalytic activity was critical for transport in the early secretory pathway, a conservative substitution was made at Cys-149 located at the active site, and the mutant was biochemically characterized in a battery of assays. Although GAPDH (C149G) has no catalytic activity, Rab2 recruited the mutant protein to membranes in a quantitative binding assay. GAPDH (C149G) is phosphorylated by aPKC iota/lambda and binds directly to Rab2 when evaluated in an overlay binding assay. Importantly, VSV-G transport between the ER and Golgi complex is restored when an in vitro trafficking assay is performed with GAPDH-depleted cytosol and GAPDH (C149G). These data suggest that GAPDH imparts a unique function necessary for membrane trafficking from VTCs that does not require GAPDH glycolytic activity.  相似文献   

15.
The association of glycolytic enzymes with F-actin is proposed to be one mechanism by which these enzymes are compartmentalized, and, as a result, may possibly play important roles for: regulation of the glycolytic pathway, potential substrate channeling, and increasing glycolytic flux. Historically, in vitro experiments have shown that many enzyme/actin interactions are dependent on ionic strength. Herein, Brownian dynamics (BD) examines how ionic strength impacts the energetics of the association of F-actin with the glycolytic enzymes: lactate dehydrogenase (LDH), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), fructose-1,6-bisphosphate aldolase (aldolase), and triose phosphate isomerase (TPI). The BD simulations are steered by electrostatics calculated by Poisson-Boltzmann theory. The BD results confirm experimental observations that the degree of association diminishes as ionic strength increases but also suggest that these interactions are significant, at physiological ionic strengths. Furthermore, BD agrees with experiments that muscle LDH, aldolase, and GAPDH interact significantly with F-actin whereas TPI does not. BD indicates similarities in binding regions for aldolase and LDH among the different species investigated. Furthermore, the residues responsible for salt bridge formation in stable complexes persist as ionic strength increases. This suggests the importance of the residues determined for these binary complexes and specificity of the interactions. That these interactions are conserved across species, and there appears to be a general trend among the enzymes, support the importance of these enzyme-F-actin interactions in creating initial complexes critical for compartmentation.  相似文献   

16.
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) catalyzes the oxidative phosphorylation of glyceraldehyde 3-phosphate to 1,3-diphosphoglycerate, one of the precursors for glycolytic ATP biosynthesis. The enzyme contains an active site cysteine thiolate, which is critical for its catalytic function. As part of a continuing study of the interactions of quinones with biological systems, we have examined the susceptibility of GAPDH to inactivation by 9,10-phenanthrenequinone (9,10-PQ). In a previous study of quinone toxicity, this quinone, whose actions have been exclusively attributed to reactive oxygen species (ROS) generation, caused a reduction in the glycolytic activity of GAPDH under aerobic and anaerobic conditions, indicating indirect and possible direct actions on this enzyme. In this study, the effects of 9,10-PQ on GAPDH were examined in detail under aerobic and anaerobic conditions so that the role of oxygen could be distinguished from the direct effects of the quinone. The results indicate that, in the presence of the reducing agent DTT, GAPDH inhibition by 9,10-PQ under aerobic conditions was mostly indirect and comparable to the direct actions of exogenously-added H2O2 on this enzyme. GAPDH was also inhibited by 9,10-PQ anaerobically, but in a somewhat more complex manner. This quinone, which is not considered an electrophile, inhibited GAPDH in a time-dependent manner, consistent with irreversible modification and comparable to the electrophilic actions of 1,4-benzoquinone (1,4-BQ). Analysis of the anaerobic inactivation kinetics for the two quinones revealed comparable inactivation rate constants (k(inac)), but a much lower inhibitor binding constant (K(i)) for 1,4-BQ. Protection and thiol titration studies suggest that these quinones bind to the NAD+ binding site and modify the catalytic thiol from this site. Thus, 9,10-PQ inhibits GAPDH by two distinct mechanisms: through ROS generation that results in the oxidization of GAPDH thiols, and by an oxygen-independent mechanism that results in the modification of GAPDH catalytic thiols.  相似文献   

17.
Nitric oxide and nitric oxide-generating agents like 3-morpholinosydnonimine (SIN-1) stimulate the mono-ADP-ribosylation of a cytosolic, 39-kDa protein in various tissues. This protein was purified from human platelet cytosol by conventional and fast protein liquid chromatography techniques. N-terminal sequence analysis identified the isolated protein as the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Nitric oxide stimulates the auto-ADP-ribosylation of GAPDH in a time and concentration-dependent manner with maximal effects after about 60 min. Associated with ADP-ribosylation is a loss of enzymatic activity. NAD(+)-free enzyme is not inhibited by SIN-1, indicating the absolute requirement of NAD+ as the substrate of the ADP-ribosylation reaction. Inhibition of the glycolytic enzyme GAPDH may be relevant as a cytotoxic effect of NO complementary to its inhibitory actions on iron-sulfur enzymes like aconitase and electron transport proteins of the respiratory chain.  相似文献   

18.
Ceramide is a sphingolipid metabolite that induces cancer cell death. When C6-ceramide is encapsulated in a nanoliposome bilayer formulation, cell death is selectively induced in tumor models. However, the mechanism underlying this selectivity is unknown. As most tumors exhibit a preferential switch to glycolysis, as described in the “Warburg effect”, we hypothesize that ceramide nanoliposomes selectively target this glycolytic pathway in cancer. We utilize chronic lymphocytic leukemia (CLL) as a cancer model, which has an increased dependency on glycolysis. In CLL cells, we demonstrate that C6-ceramide nanoliposomes, but not control nanoliposomes, induce caspase 3/7-independent necrotic cell death. Nanoliposomal ceramide inhibits both the RNA and protein expression of GAPDH, an enzyme in the glycolytic pathway, which is overexpressed in CLL. To confirm that ceramide targets GAPDH, we demonstrate that downregulation of GAPDH potentiates the decrease in ATP after ceramide treatment and exogenous pyruvate treatment as well as GAPDH overexpression partially rescues ceramide-induced necrosis. Finally, an in vivo murine model of CLL shows that nanoliposomal C6-ceramide treatment elicits tumor regression, concomitant with GAPDH downregulation. We conclude that selective inhibition of the glycolytic pathway in CLL cells with nanoliposomal C6-ceramide could potentially be an effective therapy for leukemia by targeting the Warburg effect.  相似文献   

19.
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is a key enzyme of the glycolytic pathway. Recent studies have demonstrated an additional role in apoptosis: GAPDH is targeted to the nucleus during apoptotic signalling. This nuclear transport has also been observed in serum-depleted cells, but it is reversible in fibroblasts, in contrast to apoptotic-induced transport (Eur J Cell Biol 80 (2001) 419). Here, we analyse the serum depletion-induced transport processes of GAPDH in NIH 3T3 cells. Prolonged serum depletion did not cause cell death, nuclear fragmentation (hoechst staining) or a significant increase in DNA strand-breaks (comet assay). Using cells expressing green fluorescent protein (GFP)-tagged GAPDH allowed us to monitor its intracellular localisation by confocal laser scanning microscopy (CLSM). Treatment of cells with the exportin1 inhibitor leptomycin B (LMB) did not influence cytoplasmic localisation of GFP-GAPDH, indicating that nuclear targeting of GAPDH is not constitutive and may be altered via a serum-dependent regulatory export process. Suprisingly, the export of nuclear GFP-GAPDH after re-addition of serum to starved cells was not prevented by LMB. Thus, nuclear export of GAPDH upon serum depletion is not mediated by exportin1.  相似文献   

20.
This is the first report describing the glycolytic enzyme, glyceraldehyde-3-phosphate dehydrogenase (GAPDH), as a protein associated with the cell envelope of a gram-negative bacterium (Aeromonas hydrophila). Dose-dependent GAPDH activity was detected in whole bacterial cells from exponentially growing cultures, indicating that an active form of GAPDH is located outside the plasma membrane. This activity represents roughly 10–20% of total cell activity, and it is not reduced by pretreatment of the cells with trypsin. Assays with soluble GAPDH indicate that the activity measured in intact cells does not originate by rebinding to intact cells of cytosolic enzyme released following cell lysis. GAPDH activity levels detected in intact cells varied during the growth phase. The relationship between GAPDH activity and cell culture density was not linear, showing this activity as a major peak in the late-logarithmic phase (A600 = 1.1–1.3), and a decrease when cells entered the stationary phase. The late exponential growing cells showed a GAPDH activity 3 to 4-fold higher than early growing or stationary cells. No activity was detected in culture supernatants. Enzymatic and Western-immunoblotting analysis of subcellular fractions (cytosol, whole and outer membranes, and periplasm) showed that GAPDH is located in the cytosol, as expected, and also in the periplasm. These results place the periplasmic GAPDH of A. hydrophila into the family of multifunctional microbial cell wall-associated GAPDHs which retain their catalytic activity. This revised version was published online in June 2006 with corrections to the Cover Date.  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