首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
When intact rat liver Golgi vesicles were incubated with [acetyl-3H]acetyl coenzyme A, radioactivity was incorporated into the vesicles in a manner dependent upon temperature, time, protein, and acetyl-CoA concentration. The vesicles concentrated the label 121-fold relative to the medium within 20 min, suggesting an active transport mechanism operating in intact vesicles, and incorporated more than 50% of this label into acid-insoluble materials. This was supported by the finding that incorporation was markedly reduced by Triton X-100 at levels above its critical micellar concentration. While the intravesicular low molecular weight fraction was predominantly free acetate, acetate ions themselves were not permeant to the vesicles. Double-label experiments suggested that the transport process involved the entire acetyl-CoA molecule. This was further supported by the fact that coenzyme ASH, palmitoyl-CoA and butyryl-CoA were markedly inhibitory. Incorporation was optimal at 22 degrees C at pH 7.0, and was moderately stimulated by ATP. However, compounds known to abolish proton gradients or to inhibit the Golgi proton pump had no effect. The apparent Km for the utilization process was 0.61 microM with a Vmax of 21.3 pmol/mg of protein/min. Oligomycin and 4,4'-diisothiocyanostilbene-2,2'disulfonic acid were inhibitory, whereas CMP-NeuAc, UDP-GlcNAc, adenosine 3'-phosphate, 5'-phosphosulfate, atractylosides, tunicamycin, 2'5'-ADP, and 3',5'-ADP were not, showing that this transport process is distinct from other nucleotide transporters previously described in rat liver Golgi. 75-85% of the radioactivity incorporated was shown to be in O-acetylated sialic acids, by neuraminidase release, purification, and high pressure liquid chromatography. The majority of the neuraminidase-resistant radioactivity was released by alkaline hydroxylamine as [3H]acetylhydroxamate, but a significant fraction was resistant to this treatment. The nature of the non-sialic acid radioactivity remains unknown. The existence of this transport mechanism provides yet another level at which the O-acetylation of sialic acids could be regulated.  相似文献   

2.
Direct photoaffinity labeling with radioactively labeled adenosine 3'-phosphate 5'-phosphosulfate (PAPS) followed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and autoradiography was used to identify PAPS binding proteins in a Golgi membrane preparation of bovine adrenal medulla. [3'-32P]PAPS was synthesized from adenosine 5'-phosphosulfate (APS) and [gamma-32P]ATP using APS kinase prepared from yeast and was purified by reverse-phase ion pair high performance liquid chromatography. Upon irradiation with UV light, [3'-32P]PAPS, as well as [35S]PAPS under conditions which minimized sulfotransferase-catalyzed incorporation of 35SO4 from [35S]PAPS into proteins, bound selectively to a 34-kDa protein of the Golgi membrane preparation. PAPS binding to the 34-kDa protein was strongly inhibited by the presence of 50 microM atractyloside. The 34-kDa PAPS binding protein therefore appears to be similar to the mitochondrial ATP/ADP translocator with regard to both molecular weight and inhibition by atractyloside of adenine nucleotide binding. Photoaffinity labeling will be useful in the purification and functional identification of the 34-kDa protein.  相似文献   

3.
Chick embryo chondrocyte microsomes containing intact Golgi vesicles took up 3'-phosphoadenosine-5'-phospho[35S]sulfate ([35S]PAPS) in a time- and temperature-dependent, substrate-saturable manner. When [35S]PAPS and p-nitrophenyl-N-acetyl-beta-D-galactosaminide (pNP-GalNAc) were added to the incubation in the absence of detergent, the microsomes catalyzed the transfer of sulfate from [35S]PAPS to pNP-GalNAc to form pNP-GalNAc-6-35SO4. The apparent Km values for PAPS in the uptake and the pNP-GalNAc sulfation reactions were 2 X 10(-7) and 2 X 10(-6) M, respectively. The sulfation of pNP-GalNAc by the microsomal preparation was inhibited by detergent. The microsomal fraction also catalyzed the transfer of sulfate from [35S]PAPS to oligosaccharides prepared from chondroitin. However, in contrast to the sulfation of pNP-GalNAc, the rate of sulfation of these oligosaccharides was low in the absence of detergent and was markedly stimulated when detergent was added. Sulfation of pNP-GalNAc by the freeze-thawed microsomes was inhibited when the octasaccharide prepared from chondroitin was present in the reaction mixture. As the PAPS that had been internalized in the microsomal vesicles was consumed in the sulfation of pNP-GalNAc, more [35S]PAPS was taken up and the sulfated pNP-GalNAc was released from the vesicles. These observations suggest that pNP-GalNAc may serve as a model membrane-permeable substrate for study of the 6-sulfo-transferase reaction involved in sulfation of chondroitin sulfate in intact Golgi vesicles.  相似文献   

4.
A mixture of UDP-N-acetylglucosamine labeled with different radioisotopes in the uridine and glucosamine was used to show that the intact sugar nucleotide was translocated across the membrane of vesicles derived from rat liver rough endoplasmic reticulum (RER) and Golgi apparatus. Translocation was dependent on temperature, saturable at high concentrations of sugar nucleotide, and inhibited by treatment of vesicles with proteases, suggesting protein carrier mediated transport. Translocation of UDP-GlcNAc by RER-derived vesicles appeared to be specific since these vesicles were unable to translocate UDP-galactose, in contrast to those derived from the Golgi apparatus. Preliminary results suggest that the mechanism of UDP-GlcNAc translocation into RER-derived vesicles is via a coupled exchange with lumenal nucleoside monophosphate. This is similar to the recently postulated mechanism for translocation of sugar nucleotides into vesicles derived from the Golgi apparatus.  相似文献   

5.
3'-Phosphoadenosine 5'-phospho[35S]sulfate [( 35S]PAPS) specific binding properties of rat brain tissue were studied. [35S]PAPS specific binding was optimal at pH 5.8 in either Tris-maleate or potassium phosphate buffers. Association was maximal at low temperature, reaching equilibrium in 20 min. Dissociation was rapid, with a dissociation time of 80 s. Scatchard analysis of [35S]PAPS specific binding was consistent with a single site having a KD of 0.46 +/- 0.06 microM and a Bmax of 20.8 +/- 2.0 pmol/mg of protein. Low concentrations of Triton X-100 (0.025%) were effective in increasing the number of binding sites to a Bmax of 44.5 +/- 4.6 pmol/mg of protein without affecting the affinity. [35S]PAPS specific binding was enriched in crude synaptic membranes (P2) and microsomes (P3). Regional distribution of [35S]PAPS specific binding was quite homogeneous in all brain structures studied. The pharmacological profile of [35S]PAPS specific binding in rat brain microsomes was consistent with a membrane protein having a high selectivity for the 3'-O-phosphoryl group substitution on the ribose moiety. Thus, 3'-phosphoadenosine 5'-phosphate was more potent than 2'-phosphoadenosine 5'-phosphate in competing for [35S]PAPS specific binding. Adenosine 5'-phosphosulfate was a good inhibitor of [35S]PAPS specific binding. ATP and ADP were also good displacers. Dipyridamole, a highly selective marker for adenosine uptake sites, was ineffective. 4,4-Diisothiocyanostilbene-2,2-disulfonic acid, the chloride transporter inhibitor, showed an IC50 of 36 +/- 5.1 microM for inhibition of [35S]PAPS specific binding. 2,6-Dichloro-4-nitrophenol had a low selectivity in competing for the [35S]PAPS binding site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The kinetics of the forward tyrosyl protein sulfotransferase (TPS) reaction were examined using an assay based on the 35SO4 transfer from 3'-phosphoadenosine 5'-phospho(35S)sulfate [( 35S]PAPS) to tyrosyl residues of the non-sulfated cholecystokinin derivative, BocCCK-8(ns). TPS present in the microsomal membranes from rat cerebral cortex was used for these studies. Initial velocity measurements performed over a wide range of PAPS, BocCCK-8(ns), 3'-PAP and BocCCK-8(s) concentrations, indicated that the reaction follows an ordered mechanistic pathway. The KM value determined for BocCCK-8(ns) was 160 +/- 18 microM, and that for [35S]PAPS was 0.15 +/- 0.03 microM. 3'-Phosphoadenosine 5'-phosphate (3'-PAP) was found to be a product inhibitor with a Ki = 0.30 +/- 0.02 microM. BocCCK-8(s) produced an uncompetitive inhibition pattern on the TPS reaction. Adenosine 5'-phosphosulfate (APS) behaved as a competitive inhibitor versus PAPS with a Ki = 3.0 +/- 0.3 microM. ATP inhibited competitively the reaction when PAPS was the varied substrate with a Ki = 3.6 +/- 0.5 microM. The results of product and substrate inhibition studies and the patterns of dead end inhibition obtained with APS are best fit by an ordered Bi-Bi reaction mechanism where PAPS is the first substrate to bind and 3'-PAP is the last product to be released.  相似文献   

7.
Adenosine 3'-phosphate 5'-phosphosulfate (PAPS), the "active" sulfate donor for sulfated macromolecules, is synthesized in the cytosolic fraction of rat brains. This molecule is then translocated into the lumen of the Golgi apparatus so that it is available to the sulfotransferase enzymes. The protein responsible for the PAPS translocating activity has been solubilized from vesicles enriched in enzyme markers for the Golgi apparatus and reconstituted into liposomes. In reconstituted liposomes translocating activity has a pH optimum of 7.0 and activity was increased 3-fold by divalent cations, although EDTA produced no inhibition. The affinity of the reconstituted translocator for PAPS showed a Km of 1.2 mM with a Vmax of 14 pmol of PAPS translocated/min/mg of protein. Specificity of the translocator activity was tested with a number of nucleotide analogues and only 3',5'-adenosine diphosphate was a competitive inhibitor. Inhibitors of the mitochondrial ADP/ATP transporter and the red cell anion channel blocked transport of PAPS only at very high concentrations.  相似文献   

8.
Recent studies suggest that the major pathway for exit of HCO3- across the basolateral membrane of the proximal tubule cell is electrogenic Na+/HCO3- co-transport. We therefore evaluated the possible presence of Na+/HCO3- co-transport in basolateral membrane vesicles isolated from the rabbit renal cortex. Imposing an inward HCO3- gradient induced the transient uphill accumulation of Na+, and imposing an outward Na+ gradient caused HCO3- -dependent generation of an inside-acid pH gradient as monitored by quenching of acridine orange fluorescence, findings consistent with the presence of Na+/HCO3- co-transport. In the absence of other driving forces, generating an inside-positive membrane potential by imposing an inward K+ gradient in the presence of valinomycin caused net Na+ uptake via a HCO3- -dependent pathway, indicating that Na+/HCO3- co-transport is electrogenic and associated with a flow of negative charge. Imposing transmembrane Cl- gradients did not appreciably affect HCO3- gradient-stimulated Na+ influx, suggesting that Na+/HCO3- co-transport is not Cl- -dependent. The rate of HCO3- gradient-stimulated Na+ influx was a simple, saturable function of the Na+ concentration (Km = 9.7 mM, Vmax = 160 nmol/min/mg of protein), was inhibited by 4,4'-diisothiocyanostilbene-2,2'-disulfonic acid (I50 = 100 microM), but was inhibited less than 10% by up to 1 mM amiloride. We could not demonstrate a HCO3- -dependent or 4,4'-diisothiocyanostilbene-2,2'-disulfonic acid-sensitive component of Na+ influx in microvillus membrane vesicles. This study thus indicates the presence of a transport system mediating electrogenic Na+/HCO3- co-transport in basolateral, but not luminal, membrane vesicles isolated from the rabbit renal cortex. Analogous to the use of renal microvillus membrane vesicles to study Na+/H+ exchange, renal basolateral membrane vesicles may be a useful model system for examining the kinetics and possible regulation of Na+/HCO3- co-transport.  相似文献   

9.
Transport of organic cations by a renal epithelial cell line (OK)   总被引:1,自引:0,他引:1  
The goal of this study was to determine the mechanisms involved in the transport of the organic cation, tetraethylammonium (TEA), across the apical membrane of OK cells. [14C]TEA accumulated in OK cell monolayers reaching equilibrium in 2 h. The uptake of [14C]TEA at equilibrium was dependent upon temperature and was inhibited by sodium azide and by various organic cations, including N1-methylnicotinamide (NMN), mepiperphenidol, and cimetidine but not by the organic anion, p-aminohippuric acid. The initial uptake of [14C]TEA was characterized by a saturable process. The mean +/- S.D. Km was 27.8 +/- 2.6 microM and the Vmax was 414 +/- 26.5 pmol/mg protein/min. Both an accelerated efflux and influx of [14C]TEA in the presence of a trans-gradient of unlabeled TEA and NMN was observed, whereas a deaccelerated influx and efflux was observed in the presence of a trans-gradient of mepiperphenidol. The mechanism of interaction between NMN and TEA was examined. NMN significantly increased the apparent Km (mean +/- S.D.) of TEA to 82.8 +/- 16.4 microM (p less than 0.001), whereas the Vmax (mean +/- S.D.) was only slightly affected (478 +/- 72 pmol/mg protein/min) suggesting a competitive inhibition. The stimulatory effect of trans-gradients of NMN on TEA transport was due to an increase in the Vmax of TEA suggesting that NMN trans-stimulates TEA transport by increasing the turnover rate of the exchanger. In the presence of an inwardly directed proton gradient, the efflux at 30 s of [14C]TEA from the OK cell monolayers was significantly accelerated (p less than 0.05). Studies with the pH-sensitive fluorescent probe, 2',7'-bis(carboxyethyl)-5(6)-carboxyfluorescein, suggested that TEA could drive the countertransport of protons. In apical membrane vesicles prepared from OK cells, the uptake of [3H]NMN exhibited an apparent "overshoot phenomenon" in the presence of an initial outwardly directed proton gradient. Protons competitively inhibited TEA uptake suggesting that the proton/organic cation and the organic cation/organic cation self exchange mechanism are the same mechanism. This is the first report describing both TEA self-exchange and proton/TEA exchange in the apical membrane of a continuous cell line. OK cells are an excellent model for the study of organic cation transport across the apical membrane.  相似文献   

10.
[35S]Adenosine-5'-phosphosulfate (APS) binding to Penicillium chrysogenum APS kinase was measured by centrifugal ultrafiltration. APS did not bind to the free enzyme with a measurable affinity even at low ionic strength where substrate inhibition by APS is quite marked. However, APS bound with an apparent Kd of 0.54 microM in the presence of 5 mM MgADP. In the presence of 0.1 M (NH4)2SO4, Kd,app was increased to 2.1 +/- 0.7 microM. Bound [35S]APS was displaced by low concentrations of 3'-phosphoadenosine-5'-phosphosulfate (PAPS), or iso-(2') PAPS, or (less efficiently) by adenosine-3,5'-diphosphate (PAP) or adenosine-5'-monosulfate (AMS). The results support our conclusion that substrate inhibition of the fungal enzyme by APS results from the formation of a dead end E. MgADP.APS complex. That is, APS binds to the subsite vacated by PAPS in the compulsory (or predominately) ordered product release sequence (PAPS before MgADP). Radioligand displacement was used to verify the Kd for APS dissociation from E.MgADP.APS and to determine the Kd values for the dissociation of iso-PAPS (13 +/- 5 microM), PAP (4.8 mM), or AMS (5.2 mM) from their respective ternary enzyme.MgADP.ligand complexes. Incubation of the fungal enzyme with [gamma-32P]MgATP did not yield a phosphoenzyme that survives gel filtration or gel electrophoresis.  相似文献   

11.
Plasma membrane vesicles from a glucose-responsive insulinoma exhibited properties consistent with the presence of a membrane Na+/Ca2+ exchange. The exchange was rapid, reversible, and was dependent on the external Ca2+ concentration (Km = 4.1 +/- 1.1 microM). External Na+ inhibited the uptake in a dose-dependent manner (IC50 = 15 mM). Dissipation of the Na+ gradient by 10 microM monensin decreased Na+/Ca2+ exchange from 0.74 +/- 0.17 nmoles/mg protein/s to 0.11 +/- 0.05 nmoles/mg protein/s. Exchange was not influenced by veratridine, tetrodotoxin and ouabain, or by modifiers of cAMP. No effect was seen using the calcium channel blockers, nitrendipine or nifedipine. Glucose had no direct effect on Na+/Ca2+ exchange, while glyceraldehyde, glyceraldehyde-3-phosphate and dihydroxyacetone inhibited the exchange. Na+ induced efflux of calcium was seen in Ca2+ loaded vesicles and was half maximal at [Na+] of 11.1 +/- 0.75 mM. Ca2+ efflux was dependent on [Na+], with a Hill coefficient of 2.7 +/- 0.07 indicating that activation of Ca2+ release involves a minimum of three sites. The electrogenicity of this exchange was demonstrated using the lipophilic cation tetraphenylphosphonium [( 3H]-TPP), a membrane potential sensitive probe. [3H]-TPP uptake increased transiently during Na+/Ca2+ exchange indicating that the exchange generated a membrane potential. These results show that Na+/Ca2+ exchange operates in the beta cell and may be an important regulator of intracellular free Ca2+ concentrations.  相似文献   

12.
This paper presents data identifying adenosine 3',5'-diphosphate (3',5'-ADP) as the small heat-stable factor essential for the active steroid binding complex of the adrenocortical pregnenolone-binding protein (PBP). Factor activity obtained from the boiled supernatant of partially purified PBP was isolated by high performance liquid chromatography using weak anion-exchange and hydrophobic (C18) chromatography sequentially. The purified material retained characteristic factor activity and presented a UV spectrum identical to that for authentic 3',5'-ADP. Mass spectroscopic analysis of the isolated factor revealed an M-H ion of appropriate mass (m/z = 426) and a decomposition pattern for the M-H ion that was consistent with the structure of 3',5'-ADP. The studies presented here demonstrate that authentic 3',5'-ADP can categorically substitute for factor prepared from the soluble fraction of the guinea pig adrenal. Specifically, 3',5'-ADP potentiated ligand binding of partially purified native PBP and restored binding capacity to alkaline phosphatase-inactivated PBP in a dose-dependent manner. As is the case for adrenocortical factor activity, these effects were negated by pretreating the 3',5'-ADP with calf intestinal alkaline phosphatase. Other nucleotides similarly tested, including ADP isomers, were ineffective as factor substitutes. The sulfated form of 3',5'-ADP (i.e. 3'-phosphoadenosine 5'-phosphosulfate) demonstrated some potential for restoring binding capacity to phosphatase-inactivated PBP; however, this compound was clearly inhibitory rather than stimulatory for native PBP activity. Taken collectively, the data overwhelmingly demonstrate that 3',5'-ADP is in fact the molecule required by the PBP for high affinity steroid binding complex formation. It is not yet known whether 3',5'-ADP acts allosterically or contributes directly to the structure of the steroid binding site.  相似文献   

13.
The formation of the sulfate donor [35S]3'-phosphoadenosine 5'-phosphosulfate (PAPS) from inorganic [35S]sulfate was studied using a novel assay. The assay was based on the quantitative transfer of radioactivity from [35S]PAPS to beta-naphthol under the action of phenolsulfotransferase activity from rat brain cytosol, with the [35S]beta-naphthyl sulfate formed being isolated by polystyrene bead chromatography. This simple assay was validated by comparison of results with those derived from direct assay of [35S]PAPS isolated by either TLC or ion exchange chromatography. [35S]PAPS formation by a high-speed supernatant of rat cerebral cortex occurred with an optimal pH of approximately 7.6, varied linearly with time and protein concentration, and depended on the presence of Mg2+-ATP. The latter could not be replaced by other nucleotides such as GTP, UTP, or CTP, which at 1-5 mM concentrations inhibited the reaction. Mg2+ could not be replaced by Mn2+, which at micromolar concentrations inhibited the reaction. The apparent Km values of Mg2+-ATP (at 0.1 mM [35S]sulfate) and inorganic sulfate (at 5 mM Mg2+-ATP) were 2.7 and 0.2 mM, respectively. These kinetics parameters corresponded to those reported for purified ATP sulfurylase (EC 2.7.7.4), the enzyme responsible for the first step of PAPS synthesis in liver. The product of its reaction, [35S]adenosine 5'-phosphosulfate (APS), could not be detected after incubations, an observation implying that the action of APS kinase was not rate limiting in cerebral extracts tested under the selected experimental conditions. [35S]PAPS formation was detectable in cytosolic fractions from various brain regions, which displayed only limited differences in synthesizing activity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The mechanism of reversal of resistance to Vinca alkaloids by cyclosporins is unclear. We investigated the molecular mechanism of reversal of Vinca alkaloid resistance by cyclosporin A (CsA) and its nonimmunosuppressive analog O-acetyl C9(1) CsA (SDZ 33-243) in multidrug resistant DC-3F/VCRd-5L Chinese hamster cells. CsA at 3 microM increased vincristine (VCR) sensitivity and almost totally reversed VCR resistance. SDZ 33-243 at 1 microM reduced the IC50 for VCR in resistant cells from 62.0 to 0.00062 microM. CsA and SDZ 33-243 at 10 microM increased [3H]vinblastine (VBL) accumulation in DC-3F/VCRd-5L cells by 27- and 22-fold, respectively. At 10 microM, these compounds also increased [3H]VCR accumulation by 3.5- and 4.0-fold, respectively. [3H]VCR uptake by membrane vesicles from DC-3F/VCRd-5L cells showed high and low affinity components with Michaelis-Menten kinetics, and apparent Km values were 0.140 +/- 0.0523 and 24.8 +/- 6.67 microM, respectively. Kinetic analysis of [3H]VCR uptake in membrane vesicles in the presence of 0.2 microM CsA revealed that CsA competitively inhibited the high affinity [3H]VCR uptake with an apparent inhibition constant (Ki) of 0.126 +/- 0.0173 microM. In addition, CsA and SDZ 33-243 inhibited VBL photoaffinity labeling of P-glycoprotein in a dose-dependent manner, with half-maximum inhibition at 0.5 and 0.4 microM, respectively, compared with that of VBL at 0.6 microM. These data confirm that cyclosporins modulate Vinca alkaloid resistance at least partially through interaction with P-glycoprotein.  相似文献   

15.
This report concerns development of a cell-free system from rat liver to study transport of membrane constituents from the Golgi apparatus to the plasma membrane. Highly purified Golgi apparatus as donor and a mixture of sheets and vesicles as plasma membrane acceptor fractions were combined to analyze requirements for lipid and protein transport. In the reconstituted system, the Golgi apparatus donor was in suspension. To measure transfer, membrane constituents of the donor membranes were radiolabeled with [3H]acetate (lipids) or [3H]leucine (proteins). The plasma membrane vesicles were used as the acceptor and were unlabeled and immobilized on nitrocellulose for ease of recovery and analysis. The reconstituted cell-free transfer was dependent on temperature, but even at 37 degrees C, the amount of transfer did not increase with added ATP, was not specific for any particular membrane fraction or subfraction nor was it facilitated by cytosol. ATP was without effect both in the presence or absence of a cytosolic fraction capable of the support of cell-free transfer in other systems. In contrast to results with ATP, NADH added to the reconstituted system resulted in an increased amount of transfer. A further increase in transfer was obtained with NADH plus a mixture of ascorbate and dehydroascorbate to generate ascorbate free radical. The transfer of labeled membrane constituents from the Golgi apparatus to the plasma membrane supported by NADH plus ascorbate radical was stimulated by a cytosol fraction enriched in less than 10 kDa components. This was without effect in the absence of NADH/ascorbate radical or with ATP as the energy source. Specific transfer was inhibited by both N-ethylmaleimide and GTP gamma S. The findings point to the possibility of redox activities associated with the trans region of the Golgi apparatus as potentially involved in the transport of membrane vesicles from the Golgi apparatus to the cytoplasmic surface of the plasma membrane.  相似文献   

16.
Amiloride analogs with hydrophobic substitutions on the 5-amino nitrogen atom are relatively high affinity inhibitors of the plasma membrane Na(+)-H+ exchanger. We demonstrated that a high affinity-binding site for [3H]5-(N-methyl-N-isobutyl)amiloride ([3H]MIA) (Kd = 6.3 nM, Bmax = 1.2 pmol/mg of protein) is present in microvillus membrane vesicles but not in basolateral membrane vesicles isolated from rabbit renal cortex, in accord with the known membrane localization of the Na(+)-H+ exchanger in this tissue. The rank order potency for inhibition of microvillus membrane [3H]MIA binding by amiloride analogs was: MIA (I50 approximately 10 nM) greater than amiloride (I50 approximately 200 nM) greater than benzamil (I50 approximately 1200 nM). This correlated with a qualitatively similar rank order potency for inhibition of Na(+)-H+ exchange: MIA (I50 approximately 4 microM) greater than amiloride (I50 approximately 15 microM) greater than benzamil (I50 approximately 100 microM), but did not correlate with the rank order potency for inhibition of the organic cation-H+ exchanger in microvillus membrane vesicles: MIA approximately benzamil (I50 approximately 0.5 microM) greater than amiloride (I50 approximately 10 microM). However, tetraphenylammonium, an inhibitor of organic cation-H+ exchange, inhibited the rate of [3H]MIA binding without an effect on equilibrium [3H]MIA binding; the dissociation of bound [3H]MIA was inhibited by preloading the membrane vesicles with tetraphenylammonium. These findings indicated that high affinity [3H]MIA binding to renal microvillus membrane vesicles takes place at an internal site to which access is rate-limited by the tetraphenylammonium-sensitive organic cation transporter. Equilibrium [3H]MIA binding was inhibited by H+ but was unaffected by concentrations of Na+ or Li+ that saturate the external transport site of the Na(+)-H+ exchanger. Binding of MIA to its high affinity binding site had no effect on the rate of Na(+)-H+ exchange. This study suggests that the renal Na(+)-H+ exchanger has a high affinity internal binding site for amiloride analogs that is distinct from the external amiloride inhibitory site.  相似文献   

17.
Golgi membrane vesicles can be easily and very rapidly (within 10 min.) loaded with solutions of desired composition by centrifugation of the vesicles at high g force in an air-driven ultracentrifuge and subsequent resuspension of the vesicle pellet. This centrifugal/mechanical loading procedure does not destroy the integrity of these vesicles, as demonstrated by the ability of loaded vesicles to (i) retain their contents, (ii) maintain a K+ gradient when loaded with K+ ions, and (iii) exchange internal UMP for external [3H]UMP when loaded with UMP. When radiolabeled solutes are loaded into vesicles, the displaced internal volume can be measured using a rapid filtration assay. This simple and rapid technique of replacing the intravesicular contents of Golgi membrane vesicles should prove useful in studying transport across this membrane and may have a variety of other applications, such as intravesicular volume measurements, macromolecule and drug delivery protocols, and the study of membrane fusion events.  相似文献   

18.
Three methods were used to prepare microvillus membrane vesicles from each of six human placentas. Two of these incorporated an agitation stage to preferentially remove microvilli and either Ca2+ (Method 1) or Mg2+ (Method 2) aggregation of non-microvillus membrane. The third method involved homogenisation of the tissue followed by Mg2+ aggregation of non-microvillus membrane (Method 3). Enrichment of alkaline phosphatase activity (27.6 +/- 1.9, 25.3 +/- 2.7) and ouabain binding (5.9 +/- 2.6, 5.3 +/- 2.2, respectively) was similar in vesicles prepared by Methods 1 and 2, respectively. Method 3 vesicles showed a significantly (P less than 0.01) lower alkaline phosphatase enrichment (18.1 +/- 1.2), but ouabain binding enrichment (6.3 +/- 1.3) was not different and vesicle protein recovery (mg/g placenta) was 5-fold greater. Na+ uptake in the presence of an outwardly directed proton gradient was significantly inhibited in all microvillus membrane vesicles by amiloride (0.5 mM). However, the amiloride sensitive component of Na+ uptake was 3-6-fold greater in Method 3 vesicles than in Method 1 and 2 vesicles, and showed overshoot above equilibrium in the former but not the latter. Further experiments using the pH sensitive dye, 2',7'-bis(carboxyethyl)-5(6)-carboxyfluorescein suggested that the proton gradient dissipated faster from Method 1 than from Method 3 vesicles. Thus methodological differences can have a marked effect on transport processes in microvillus membrane vesicles prepared from the human placenta.  相似文献   

19.
The dependence on Na+, K+, and Cl- of uptake and accumulation of [3H]noradrenaline was studied in plasma membrane vesicles isolated from PC-12 pheochromocytoma cells. Plasma membrane vesicles accumulated [3H]noradrenaline when an inward-directed gradient for Na+ and an outward-directed gradient for K+ were imposed across the vesicle membrane. Under these conditions, initial rates of uptake of [3H]noradrenaline were saturable (Km = 0.14 microM) and inhibited by a series of substrates and inhibitors of "uptake". The IC50 values were positively correlated with those for inhibition of uptake into intact PC-12 cells. Uptake and accumulation of [3H]noradrenaline in plasma membrane vesicles were absolutely dependent on external Na+ and Cl-; they were dependent on an inwardly directed gradient for Na+ but less dependent on an inwardly directed gradient for Cl-. Internal K+ strongly enhanced uptake and accumulation of [3H]noradrenaline. Rb+, but not Li+, had the capacity to replace internal K+. Two explanations are proposed for this effect of internal K+: (a) creation of a K+ diffusion potential (inside negative) provides a driving force for inward transport, and/or (b) K+ increases the turnover rate by formation of a highly mobile potassium-carrier complex. A hypothetical scheme for the transport of noradrenaline is presented.  相似文献   

20.
Isolated intact rat liver Golgi vesicles utilize [acetyl-3H]coenzyme A to add 3H-O-acetyl esters to sialic acids of internally facing endogenous glycoproteins. During this reaction, [3H]acetate also accumulates in the vesicles, even though the vesicles are impermeant to free acetate. On the other hand, entry of intact AcCoA into the lumen of the vesicles could not be demonstrated, and permeabilization of the vesicles did not alter the reaction substantially (Diaz, S., Higa, H. H., Hayes, B. K., and Varki, A. (1989) J. Biol. Chem. 264, 19416-19426). When vesicles prelabeled with [acetyl-3H] coenzyme A are permeabilized with saponin, we can demonstrate a [3H]acetyl intermediate in the membrane that can transfer label to the 7- and 9-positions of exogenously added free N-acetylneuraminic acid but not to glucuronic acid or CMP-N-acetylneuraminic acid. This labeled acetyl intermediate represents a significant portion of the radioactivity incorporated into the membranes during the initial incubation and cannot be accounted for by nonspecifically "trapped" acetyl-CoA in the permeabilized vesicles. There was no evidence for involvement of acetylcarnitine or acetyl phosphate as an intermediate. The overall acetylation reaction appears to involve two steps. The first step (utilization of exogenous acetyl-CoA to form the acetyl intermediate) is inhibited by coenzyme A-SH (apparent Ki = 24-29 microM), whereas the second (transfer from the acetyl intermediate to sialic acid) is not affected by millimolar concentrations of the nucleotide. Studies with amino acid-modifying reagents indicate that 1 or more histidine residues are involved in the first step of the acetylation reaction. Diethylpyrocarbonate (which can react with both nonsubstituted and singly acetylated histidine residues) also blocks the second reaction, indicating that the acetyl intermediate on both sides of the membrane involves histidine residue(s). Taken together with data presented in the preceding paper, these results indicate that the acetylation of sialic acids in Golgi vesicles may occur by a transmembrane reaction, similar to that described for the acetylation of glucosamine in lysosomes (Bame, K. J., and Rome, L. H. (1985) J. Biol. Chem. 260, 11293-11299). However, several features of this Golgi reaction distinguish it from the lysosomal one, including the nature and kinetics of the reaction and the additional involvement of an essential lysine residue. The accumulation of free acetate in the lumen of the vesicles during the reaction may occur by abortive acetylation (viz. transfer of label from the acetyl intermediate to water). It is not clear if this is an artifact that occurs only in the in vitro reaction.  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