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.
The ability of rat hepatocytes to degrade internalized surface-bound 125I-asialoorosomucoid (ASOR) was determined by measuring the appearance of acid-soluble radioactivity at 37 degrees C. The degradation kinetics were biphasic in cells previously equilibrated at 37 degrees C for 1 h or cultured for 24 h. Degradation began immediately and was linear for at least 20 min after which the rate increased to a steady state value 3-4 times greater than the initial rate. We previously showed that hepatocytes have two functionally distinct populations of galactosyl receptors that mediate ligand dissociation by two kinetically different pathways (Weigel, P. H., Clarke, B. L., and Oka, J. A. (1986) Biochem. Biophys. Res. Commun. 140, 43-50). The activity of one receptor population, designated State 2 galactosyl receptors, can be reversibly modulated by incubating cells between 22 and 37 degrees C and is not expressed on the surface of freshly isolated cells. When 125I-ASOR was prebound to freshly isolated cells at 4 degrees C and degradation was assessed subsequently at 37 degrees C, the kinetics were monophasic, not biphasic. Degradation of the surface-bound 125I-ASOR began immediately and was greater than 90% complete by 6 h. Freshly isolated cells were incubated at temperatures between 22 and 37 degrees C, chilled to 4 degrees C, allowed to pre-bind 125I-ASOR, and then incubated at 37 degrees C. As the State 2 galactosyl receptor population increased, the kinetics of degradation became progressively more biphasic and the rate of the delayed degradation process increased. This effect could be reversed in cells in culture or in suspension by down-modulating surface receptor activity at temperatures below 37 degrees C; only the degradation process appearing after a 20-min lag was affected. Degradation in both pathways is an apparent first order process with identical rate constants (kappa = 0.006 min-1, t1/2 = 116 min). We conclude that there are two separate pathways by which asialoglycoproteins are degraded. The major "classic" pathway mediated by State 2 galactosyl receptors occurs after a 20-min lag and the minor pathway mediated by State 1 galactosyl receptors begins immediately with no detectable lag.  相似文献   

2.
Isolated rat hepatocytes were pulse-labelled with [35S]methionine at 37 degrees C and subsequently incubated (chased) for different periods of time at different temperatures (37-16 degrees C). The time courses for the secretion of [35S]methionine-labelled albumin and haptoglobin were determined by quantitative immunoprecipitation of the detergent-solubilized cells and of the chase media. Both proteins appeared in the chase medium only after a lag period, the length of which increased markedly with decreasing chase temperature: from about 10 and 20 min at 37 degrees C to about 60 and 120 min at 20 degrees C for albumin and haptoglobin respectively. The rates at which the proteins were externalized after the lag period were also strongly affected by temperature, the half-time for secretion being 20 min at 37 degrees C and 200 min at 20 degrees C for albumin; at 16 degrees C no secretion could be detected after incubation for 270 min. Analysis by subcellular fractionation showed that part of the lag occurred in the endoplasmic reticulum and that the rate of transfer to the Golgi complex was very temperature-dependent. The maximum amount of the two pulse-labelled proteins in Golgi fractions prepared from cells after different times of chase decreased with decreasing incubation temperatures, indicating that the transport from the Golgi complex to the cell surface was less affected by low temperatures than was the transport from the endoplasmic reticulum to the Golgi complex.  相似文献   

3.
Rat liver parenchymal cell binding, uptake, and proteolytic degradation of rat 125I-labeled high density lipoprotein (HDL) subfraction, HDL3 (1.10 less than d less than 1.210 g/ml), in which apo-A-I is the major polypeptide, were investigated. Structural and metabolic integrity of the isolated cells was verified by trypan blue exclusion, low lactic dehydrogenase leakage, expected morphology, and gluconeogenesis from lactate and pyruvate. 125I-labeled HDL3 was incubated with 10 X 10(6) cells at 37 degrees and 4 degrees in albumin and Krebs-Henseleit bicarbonate buffer, pH 7.4. Binding and uptake were determined by radioactivity in washed cells. Proteolytic degradation was determined by trichloroacetic acid-soluble radioactivity in the incubation medium. At 37 degrees, maximum HDL3 binding (Bmax) and uptake occurred at 30 min with a Bmax of 31 ng/mg dry weight of cells. The apparent dissociation constant of the HDL3 receptor system (Kd) was 60 X 10(-8) M, based on Mr = 28,000 of apo-A-I, the predominant rat HDL3 protein. Proteolytic degradation showed a 15-min lag and then constant proteolysis. After 2 hours 5.8% of incubated 125I-labeled HDL3 was degraded. Sixty per cent of cell radioactivity at 37 degrees was trypsin-releasable. At 37 degrees, 125I-labeled HDL3 was incubated with cells in the presence of varying concentrations of native (cold) HDL3, very low density lipoproteins, and low density lipoproteins. Incubation with native HDL3 resulted in greatest inhibition of 125I-labeled HDL3 binding, uptake, and proteolytic degradation. When 125I-labeled HDL3 was preincubated with increasing amounts of HDL3 antiserum, binding and uptake by cells were decreased to complete inhibition. Cell binding, uptake, and proteolytic degradation of 125I-labeled HDL3 were markedly diminished at 4 degrees. Less than 1 mM chloroquine enhanced 125I-labeled HDL3 proteolysis but at 5 mM or greater, chloroquine inhibited proteolysis with 125I-labeled HDL3 accumulation in cells. L-[U-14C]Lysine-labeled HDL3 was bound, taken up, and degraded by cells as effectively as 125I-labeled HDL3. These data suggest that liver cell binding, uptake, and proteolytic degradation of rat HDL3 are actively performed and linked in the sequence:binding, then uptake, and finally proteolytic degradation. Furthermore, there may be a specific HDL3 (lipoprotein A) receptor of recognition site(s) on the plasma membrane. Finally, our data further support our previous reports of the important role of liver lysosomes in proteolytic degradation of HDL3.  相似文献   

4.
Receptor sites for insulin on GH3 cells were characterized. Uptake of 125I-labeled insulin by the cells was dependent upon time and temperature, with apparent steady-states reached by 120, 20 and 10 min at 4, 23 and 37 degrees C, respectively. The binding sites were sensitive to trypsin, suggesting that the receptors contain protein. Insulin competed with 125I-labeled insulin for binding sites, with half-maximal competition observed at 5 nM insulin. Neither adrenocorticotropic hormone nor growth hormone competed for 125I-labeled insulin binding sites. 125I-labeled insulin binding was reversible, and saturable with respect to hormone concentration. 125I-labeled insulin was degraded at both 4 and 37 degrees C by GH3 cells, but not by medium conditioned by these cells. After a 5 min incubation at 37 degrees C, products of 125I-labeled insulin degradation could be recovered from the cells but were not detected extracellularly. Extending the time of incubation resulted in the recovery of fragments of 125I-labeled insulin from both cells and the medium. Native insulin inhibited most of the degradation of 125I-labeled insulin suggesting that degradation resulted, in part, from a saturable process. At steady-state, degradation products of 125I-labeled insulin, as well as intact hormone, were recovered from GH3 cells. After 30 min incubation at 37 degrees C, 80% of the cell-bound radioactivity was not extractable from GH3, cells with acetic acid.  相似文献   

5.
Using anticholeragen antibodies and 125I-protein A, we developed a specific and quantitative assay for measuring choleragen on the surfaces of cultured cells. When neuroblastoma cells containing bound toxin were incubated at 37 degrees C, surface toxin disappeared with a half-life of approximately 2 h and a significant loss was detected by 10 min. When cells were incubated with 125I-choleragen in order to measure toxin degradation, cell-associated radioactivity disappeared with time and a corresponding amount of TCA-soluble label appeared in the culture medium with a half-life of 4-6 h. No degradation was detected until 45 min. Although there was a lag of 15 min before bound choleragen activated adenylate cyclase, the enzyme became maximally activated between 45 and 60 min. Similar results were obtained with Friend erythroleukemia cells. Internalization, degradation, and activation all were blocked when the cells were maintained at 4 degrees C. At 22 degrees C, internalization and activation occurred, albeit at a slower rate, whereas degradation was effectively inhibited. These results indicated that choleragen does not have to be degraded by intact cells in order for it to activate adenylate cyclase. Some internalization of the toxin, however, appears to precede the activation process.  相似文献   

6.
The effect of solvent on the rate of leuprolide degradation and on the structure of the degradation products was explored. Leuprolide solutions (370 mg/mL) were prepared in water and dimethyl sulfoxide (DMSO) for delivery in DUROS osmotic implants. Both solvent systems demonstrated better than 90% stability after 1 year at 37 degrees C, where the DMSO formulation afforded better stability than the aqueous formulation and was used in subsequent clinical trials. The rate of leuprolide degradation in DMSO was also observed to accelerate with increasing moisture content, indicating that the aprotic solvent minimized chemical degradation. Interestingly, leuprolide degradation products varied with formulation vehicle. The proportions of leuprolide degradation products observed to form in water and DMSO at 37 degrees C were hydrolysis > aggregation > isomerization > oxidation and aggregation > oxidation > hydrolysis > isomerization, respectively. Specifically, more N-terminal hydrolysis and acetylation were observed under aqueous conditions, and increased Trp oxidation and Ser beta-elimination were seen under non-aqueous conditions. Furthermore, the major chemical degradation pathway changed with temperature in the DMSO formulation (decreasing oxidation with increasing temperature), but not in the aqueous formulation.  相似文献   

7.
We have developed conditions for studying the binding, uptake, degradation and transport of 125I-labelled IgG by yolk sac in vitro. Specific binding to tissue at 4 degrees C and to paraformaldehyde-treated tissue at 37 degrees C was time- and temperature-dependent and showed saturation kinetics (Kd,4 degrees C = 2.9 X 10(-6) M, Kd,37 degrees C = 5.3 X 10(-6) M). Uptake was studied at 37 degrees C using untreated tissue (K uptake = 13.3 X 10(-6) M) and was inhibited by preincubation with metabolic poisons but not with cycloheximide. Tissue that had been incubated with 125I-labelled IgG at 37 degrees C released radiolabelled degradation products and intact 125I-labelled IgG into the medium. Experiments with paraformaldehyde-treated and untreated tissue showed that release of intact 125I-labelled IgG was mostly the result of ligand dissociation from surface binding sites. However, more 125I-labelled IgG was released from untreated tissue than could be accounted for solely by loss of surface-bound ligand and the difference was presumed to reflect uptake, transport and exocytosis of 125I-labelled IgG. Degradation of 125I-labelled IgG was inhibited by leupeptin and lysosomotropic amines. These drugs had no detectable effect on 125I-labelled IgG release. The results suggest that degradation and transport of IgG are not intimately related and are consistent with a previously proposed model for IgG transport via coated vesicles which do not fuse with lysosomes and for non-selective uptake into another class of vesicle which does fuse with lysosomes.  相似文献   

8.
Degradative processing of internalized insulin in isolated adipocytes   总被引:9,自引:0,他引:9  
Based on the distribution of 125I-insulin between the cell surface and the cell interior, it was found that insulin rapidly binds (t 1/2 = 0.4 min) to surface receptors at 37 degrees C, and after an initial lag period of about 1 min, accumulates intracellularly until steady state is reached (t 1/2 = 3.5 min). At this time about 40% of the total cell-associated 125I-insulin resides in the cell interior reflecting a dynamic equilibrium between the rate of insulin endocytosis and the rate at which internalized insulin is processed and extruded from cells. Since this percentage decreased to 15% at 16 degrees C, it appears that internalization is more temperative-sensitive than the intracellular processing of insulin. When 125I-insulin was preloaded into the cell interior, it was found that internalized insulin was rapidly released to the medium at 37 degrees C (t 1/2 = 6.5 min) and consisted of both degraded products and intact insulin (as assessed by trichloroacetic acid precipitability and column chromatography). Since 75% of internalized insulin was ultimately degraded, and 25% was released intact, this indicates that degradation is the predominant pathway. To determine when incoming insulin enters a degradative compartment, cells were continually exposed to 125I-insulin and the composition of insulin in the cell interior over time was assessed. After 2 min all endocytosed insulin was intact, between 2-3 min degradation products began accumulating intracellularly, and by 15 min equilibrium was reached with 20% of internalized insulin consisting of degraded products. Degraded insulin was then released from the cell interior within 4-5 min after endocytotic uptake, since this was the earliest time chloroquine was found to inhibit the release of degradation products. Moreover, the final release of degraded insulin was not inhibitable by the energy depleter dinitrophenol. Thus, within the degradative pathway, insulin enters lysosomes by 2.5-3 min and is released to the medium by simple diffusion after an additional 1.5-2 min.  相似文献   

9.
Semen, collected from active honey bee drones, was centrifuged to separate spermatozoa from seminal plasma. Enzymes were then extracted from spermatozoa with detergent buffer mixture at 37°C for 1 h. The following hydrolytic-enzyme activities were detected in seminal plasma and/or detergent extracts of sperm: protease (azocoll hydrolysis), hexosaminidase (N-acetyl-β-D-glucosaminidase), hyaluronidase, aryl sulfatase A and B.  相似文献   

10.
125I-labelled human serum low density lipoproteins (LDL) were incubated with cultured mouse peritoneal macrophages at 37 degrees C, with the following study of cellular uptake and 125I-LDL degradation by measuring the content of TCA-soluble products of LDL hydrolysis in the cultural medium. It was shown that limited pepsin proteolysis of LDL (10%) led to a more effective LDL uptake and degradation by macrophages. The data suggest that enzyme-induced modification of LDL may increase their atherogenicity.  相似文献   

11.
Thyrotropin releasing hormone (TRH) causes phosphatidylinositol bisphosphate hydrolysis to form inositol trisphosphate and diacylglycerol. Since diacylglycerol activates protein kinase C (Ca2+/phospholipid-dependent enzyme), this enzyme may be involved in mediating the physiological response to TRH. Activation of protein kinase C leads to phosphorylation of receptors for epidermal growth factor (EGF) and decreased EGF affinity. The present study examined the effect of TRH on EGF binding to intact GH4C1 rat pituitary tumor cells to test whether TRH activates protein kinase C. Cells were incubated with TRH at 37 degrees C and specific 125I-EGF binding was then measured at 4 degrees C. 125I-EGF binding was decreased by a 10-min treatment with 0.1-100 nM TRH to 30-40% of control in a dose-dependent manner. 125I-EGF binding was not altered if cells were incubated at 4 degrees C, although TRH receptors were saturated or in a variant pituitary cell line without TRH receptors. TRH (10 min at 37 degrees C) decreased EGF receptor affinity but caused little change in receptor density, 125I-EGF internalization, or degradation. When cells were incubated continuously with TRH, there was a recovery of 125I-EGF binding after 24 h. Incubation with the protein kinase C activating phorbol ester TPA caused an immediate (less than 10 min) profound (greater than 85%) decrease in 125I-EGF binding followed by partial recovery at 24 h. Maximally effective doses of TRH and TPA decreased EGF receptor affinity with half-times of 3 min. EGF treatment (5 min) caused an increase in the tyrosine phosphate content of several proteins; prior incubation with TRH resulted in a small decline in the EGF response. GH4C1 cells were incubated with 500 nM TPA for 24 h in order to down-regulate protein kinase C. Protein kinase C depletion was confirmed by immunoblots and the effects of TRH and TPA on 125I-EGF binding were tested. TRH and TPA were both much less effective in cells pretreated with phorbol esters. TRH increased cytoplasmic pH measured with an intracellularly trapped pH sensitive dye after mild acidification with nigericin. This TRH response is presumed to be the result of protein kinase C-mediated activation of the amiloride-sensitive Na+/H+ exchanger and was blunted in protein kinase C-depleted cells. All of these results are consistent with the view that TRH acts rapidly in the intact cell to activate protein kinase C and that a consequence of this activation is EGF receptor phosphorylation and Na+/H+ exchanger activation.  相似文献   

12.
Internalization and degradation of insulin by human erythrocytes were studied. Erythrocytes were incubated with 125I-insulin at 4 degrees C, 15 degrees C, and 37 degrees C for varying time intervals. These erythrocytes were then subjected to a low pH wash to release bound insulin followed by TCA precipitation. After 4, 22, and 24 hours of insulin binding at 4 degrees C, 92 to 95% of the bound 125I-insulin was dissociable and 92 to 98% of the extractable insulin was undegraded. After 3.5 hours of incubation at 15 degrees, 82% of the bound insulin was dissociable and 60% of this was intact. However, after 60, 90, 120, and 180 minutes of incubation at 37 degrees C, only 42, 34, 24, and 37%, respectively, of the bound insulin was dissociable. The undissociated insulin in the 37 degrees C studies was considered to be intracellular. With increasing time of incubation at 37 degrees C, the extractability of cell bound insulin and the proportion of undegraded dissociable insulin were decreased. When 125I-insulin binding was 95% blocked by preincubating the erythrocytes with anti-insulin receptor antibody, 95% of the degradation of 125I-insulin was also blocked. These studies indicate that mature human erythrocytes degrade internalized insulin and this process is time, temperature, and insulin receptor dependent.  相似文献   

13.
We studied binding and degradation of labeled platelet thrombospondin (TSP) by normal and variant bovine aorta endothelial (BAE) cells. [125I]-labeled TSP bound to cells at 37 degrees C in a specific, saturable, and time-dependent fashion. Incubation of cell monolayers with fluoresceinated TSP resulted in punctate cellular staining, but no staining of the extracellular matrix. Heparin, fucoidan, chondroitin sulfate, platelet factor 4, beta-thromboglobulin, unlabeled TSP, and serum derived from whole blood all competed for binding of [125I]TSP. [125I]TSP was degraded to TCA-soluble radioactivity, which appeared in the medium after a 60-90-min lag. Degradation was inhibited to the same extent as binding by increasing concentrations of heparin, fucoidan, platelet factor 4, or whole blood serum. Normal BAE cells bound and degraded less [125I]TSP than variant BAE cells. The dissociation constants (Kds) for binding and the constants for degradation (Kms) for degradation by the two cell strains, however, were similar (30-50 nM). The inhibitory effects of heparin and platelet factor 4 were lost when the two inhibitors were present in a 1:1 (wt/wt) ratio. Treatment of suspended cells with trypsin or heparitinase caused less binding of TSP. These results indicate that there is a specific receptor for TSP on endothelial cells which mediates binding and degradation. This receptor may be a heparan sulfate proteoglycan.  相似文献   

14.
S Kanamura 《Histochemistry》1975,41(3):257-263
The effect of glutaraldehyde fixation on glucose 6-phosphatase activity in mouse liver was investigated. After transparenchymal perfusion with 2% glutaraldehyde for 1.5 minutes, the activity of the recovered enzyme was higher than those reported for acid phosphatase and aryl sulfatase activities after fixation under similar condition, and an abundant deposition of reaction product was observed in hepatocytes. Subsequent immersion in the same fixative solution for 30 minutes after 4 degrees C resulted in only a slight decrease in the activity. However, the activity was almost completely destroyed after 3 hours of immersion fixation at 4 degrees C following the perfusion. Therefore, the enzyme can be said to be aldehyde-sensitive when a long fixation time is used, but not aldehyde-sensitive during a short fixation time.  相似文献   

15.
Diacytosis of 125I-asialoorosomucoid by rat hepatocytes was studied by preincubating the cells with the labelled ligand at 37 degrees C for 30 min or 18 degrees C for 2 h, washing free of cell surface receptor-bound tracer at 4 degrees C and then reincubating at 37 degrees C. The cells preloaded at 37 degrees C released a maximum of 18% of the total intracellular ligand as undegraded molecules after 1 h of incubation with an apparent first-order rate constant of 0.018 min-1 (t1/2 = 39 min). When the preloaded cells were incubated in the presence of 100 micrograms/ml unlabelled asialoorosomucoid or 5 mM ethylene glycol bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, the amount of the released ligand increased to 32 and 37%, respectively, without apparent change in kinetics, indicating that these agents prevented rebinding of the released ligand. In the presence of 5 microM colchicine, 20 microM cytochalasin B, 20 microM chloroquine, 10 mM NH4Cl, 10 microM monensin or 20 microM leupeptin, degradation of the preloaded ligand was inhibited, whereas the release of the ligand was either slightly increased or unchanged. Similar effects of leupeptin, colchicine and asialoorosomucoid were observed with cells preloaded at 18 degrees C. These results indicate that diacytosis of 125I-asialoorosomucoid occurs from a prelysosomal compartment via a route insensitive to inhibition by the inhibitors of ligand degradation.  相似文献   

16.
Uptake and processing of human platelet factor 4 by hepatocytes   总被引:1,自引:0,他引:1  
We previously demonstrated rapid clearance of human platelet factor 4 (PF4) from rabbit and rat blood, its accumulation in the liver, and elimination of PF4 degradation products in urine. The purpose of the present experiments was to characterize interaction of PF4 with cultured rat hepatocytes. 125I-PF4 was taken up by hepatocytes reaching maximum at 180 min. The association of 125I-PF4 with hepatocytes was two times greater at 37 degrees C than at 4 degrees C. At 37 degrees C degradation of 125I-PF4 by hepatocytes was also observed as indicated by the increase of 125I-PF4 radioactivity soluble in 6% trichloroacetic acid. By contrast, no uptake of 125I-beta-thromboglobulin antigen was observed. Autoradiography demonstrated that short incubation (5-20 min) of 125I-PF4 with hepatocytes results in the association of 125I-radioactivity with cell membranes while after longer incubation (60 min) radioactivity was also localized in the endosomes. Heparin inhibited binding and uptake of 125I-PF4 radioactivity by hepatocytes. We propose that part of PF4 released in the circulating blood by activated platelets is bound to the surface of hepatocytes and that it is further processed by these cells.  相似文献   

17.
The endocytosis and intracellular transport of mannose-6-phosphate conjugated to bovine serum albumin (Man-6-P:BSA) by mouse T-lymphoma cells were investigated in detail using several methods of analysis, both morphological and biochemical. Man-6-P:BSA was labeled with fluorescein or 125I and used to locate both surface and intracellular Man-6-P binding sites by light or electron microscopy, respectively. Incubation of cells with either fluorescent- or 125I-labeled Man-6-P:BSA at 0 degree C revealed a uniform distribution of the Man-6-P binding sites over the cell surface. Competition experiments indicate that the Man-6-P:BSA binding sites on the cell surface are the same receptors that can recognize lysosomal hydrolases. After as little as 1 min incubation at 37 degrees C, endocytosis of Man-6-P binding sites was clearly observed to occur through regions of the plasma membrane and via vesicles that also bound anticlathrin antibody. After a 5-15-min incubation of cells at 37 degrees C, the internalized ligand was detected first in the cis region of the Golgi apparatus and then in the Golgi stacks using both autoradiography and immunocytochemistry to visualize the ligand. The appearance of Man-6-P:BSA in the Golgi region after 15-30 min was confirmed by subcellular fractionation, which demonstrated an accumulation of Man-6-P:BSA in light membrane fractions that corresponded with the Golgi fractions. After a 30-min incubation at 37 degrees C, the internalized Man-6-P binding sites were localized primarily in lysosomal structures whose membrane but not lumen co-stained for acid phosphatase. These results demonstrate a temporal participation of clathrin-containing coated vesicles during the initial endocytosis of Man-6-P binding sites and that one step in the Man-6-P:BSA transport pathway between plasma membrane and the lysosomal structure can involve a transit through the Golgi stacks.  相似文献   

18.
We have used isolated perfused rat livers to examine the intracellular processing of 125I-epidermal growth factor (EGF) and to determine where in the endocytic pathway the hydrolases which degrade EGF are acting. Following uptake of 125I-EGF at 37 or 16 degrees C, subcellular fractions enriched in endosomes and lysosomes were isolated, and their 125I-EGF content was examined by reverse-phase high performance liquid chromatography. Three forms of EGF processed at their carboxyl termini are generated in endosomes. At 37 degrees C, EGF is first processed in early endosomes by a carboxypeptidase B-like protease and is further processed in late endosomes by a trypsin-like protease and then a carboxypeptidase B-like protease. At 16 degrees C, entry of EGF into late endosomes is slowed, and only the first processed form is generated over 60 min. Longer perfusions (180 min) at 16 degrees C result in some processing (7%) by proteases found in late endosomes. EGF-horseradish peroxidase cytochemistry confirmed that the additional processing detected at 180 min correlated with movement of EGF from tubulovesicular to multivesicular endosomes. These results, combined with in vitro incubations of EGF in isolated endosomal and lysosomal fractions, suggest that different proteases are active at selective points in the endocytic pathway and that the full complement of proteases needed for complete degradation of EGF is active only in lysosomes.  相似文献   

19.
B E Bihain  F T Yen 《Biochemistry》1992,31(19):4628-4636
This paper describes a mechanism for degradation of low-density lipoprotein (LDL) in fibroblasts unable to synthesize the LDL receptor. In this cell line, long-chain free fatty acids (FFA) activated 125I-LDL uptake; unsaturated FFA were the most efficient. The first step of this pathway was the binding of LDL apoB to a single class of sites on the plasma membrane and was reversible in the presence of greater than or equal to 10 mM suramin. Binding equilibrium was achieved after a 60-90-min incubation at 37 degrees C with 1 mM oleate; under these conditions, the apparent Kd for 125I-LDL binding was 12.3 micrograms/mL. Both cholesterol-rich (LDL and beta-VLDL) and triglyceride-rich (VLDL) lipoproteins, but not apoE-free HDL, efficiently competed with 125I-LDL for this FFA-induced binding site. After LDL bound to the cell surface, they were internalized and delivered to lysosomes; chloroquine inhibited subsequent proteolysis of LDL and thereby increased the cellular content of the particles. A physiological oleate to albumin molar ratio, i.e., 1:1 (25 microM oleate and 2 mg/mL albumin), was sufficient to significantly (p less than 0.01) activate all three steps of this alternate pathway: for example, 644 +/- 217 (25 microM oleate) versus 33 +/- 57 (no oleate) ng of LDL/mg of cell protein was degraded after incubation (2 h, 37 degrees C) with 50 micrograms/mL 125I-LDL. We speculate that this pathway could contribute to the clearance of both chylomicron remnants and LDL.  相似文献   

20.
Rabbit alveolar macrophages rapidly internalize and degrade mannosylated bovine serum albumin (125I-mannose-BSA). Trichloroacetic acid-soluble degradation products appear in the cells as early as 6 min after uptake at 37 degrees C, and in the extracellular medium after 10 min. Incubation of endocytic vesicles containing this ligand in isotonic buffers at pH 7.4 + ATP resulted in intravesicular proteolysis, which was inhibited by monensin, nigericin, or ammonium chloride. At pH 5.0, degradation proceeded rapidly and was abolished by lysis of the vesicles with 0.1% Triton X-100. Readdition of lysosomes to the incubation mixture did not increase the rate of prelysosomal degradation. Proteolysis of 125I-mannose-BSA was optimal at pH 4.5, and inhibited by low concentrations of the cathepsin D inhibitor pepstatin A. After subcellular fractionation of the macrophages on Percoll gradients, 125I-mannose-BSA sedimented with prelysosomal vesicles and was not transported to secondary lysosomes. Addition of pepstatin A to extracellular medium during internalization of prebound 125I-mannose-BSA partially inhibited degradation of ligand, and resulted in transfer of undegraded 125I-mannose-BSA to lysosomes after 20 min. Using 125I-bovine serum albumin as a substrate for the protease in the presence of 0.1% Triton X-100, we have shown that as much as 36% of the total pepstatin A-sensitive activity sediments with nonlysosomal membranes. After intraendosomal iodination using lactoperoxidase, a labeled protease was isolated by affinity chromatography on pepstatin-agarose. The labeled protease, which had a subunit size of 46 kDa, was detected in endocytic vesicles after 5 min of internalization. These results suggest that a cathepsin D-like protease is responsible for the degradation of 125I-mannose-BSA in macrophages, and that this ligand is degraded in a prelysosomal vesicle.  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