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 46 毫秒
1.
The relationship between glycosylation, dimerization, and heparin affinity of lipoprotein lipase (LPL) was studied in 3T3-L1 adipocytes. Three forms of LPL subunits were found in normal cells; totally endo H-resistant (57 kDa), partially sensitive (54 kDa), and totally sensitive (51 kDa) forms. LPL in normal cells was active, dimeric, and showed high affinity for heparin. LPL in cells treated with tunicamycin, preventing the transfer of N-linked oligosaccharide chain, was unglycosylated (51 kDa) and inactive. LPL proteins were found as an aggregate, and had low affinity for heparin. After treatment with castanospermine, an inhibitor of ER glucosidase I, 80% of LPL activity was inhibited. Most of LPL proteins were totally endo H-sensitive, present as an aggregate, and had low affinity for heparin. LPL in cells treated with deoxymannojirimycin, an inhibitor of Golgi mannosidase I, was active, dimeric, and had high affinity for heparin as in normal cells. But LPL subunits were all endo H-sensitive. These results suggest that core glycosylation and subsequent removal of glucose residue is required, but processing after Golgi mannosidase I is not necessary for dimerization and acquisition of high heparin affinity of LPL.  相似文献   

2.
The relationship between maturation of lipoprotein lipase (LPL) and its translocation from the endoplasmic reticulum (ER) to the Golgi complex was determined by measuring lipolytic activity under conditions preventing transport of the enzyme from the ER to the Golgi compartment. In the presence of brefeldin A, a reagent that inhibits movement of proteins from the ER and causes the disassembly of the Golgi complex, pro-5 Chinese hamster ovary cells accumulated catalytically active LPL, while secretion of the enzyme was effectively blocked. LPL retained intracellularly by brefeldin A treatment possessed oligosaccharide chains that were processed to the complex form by the Golgi enzymes redistributed into the ER. At 16 degrees C, a condition disrupting protein transport to the cis-Golgi, the retained enzyme again remained catalytically active although the oligosaccharides remained in the high mannose form. Lastly, attachment of the specific ER retention signal KDEL (Lys-Asp-Glu-Leu) to the carboxyl terminus of LPL also resulted in intracellularly retained enzyme that was fully active. The importance of oligosaccharide processing for attainment of LPL catalytic activity in vitro was also determined. LPL was active and secreted when trimming of the mannose residues was inhibited by deoxymannojirimycin and when addition of complex sugars was blocked using Chinese hamster ovary mutants (lec1 and lec2), indicating that these processing events are not necessary for the expression of a functional enzyme. However, blocking glucose removal by glucosidase inhibitors (castanospermine and N-methyl-deoxynojirimycin) resulted in a significant reduction in LPL specific activity and secretion. Thus, glucose trimming of LPL oligosaccharides is essential for enzyme activation; however, further oligosaccharide processing or translocation of the enzyme to the cis-Golgi is not required for full expression of lipolytic activity in vitro.  相似文献   

3.
The mutations cld (combined lipase deficiency) and lec23 disrupt in a similar manner the expression of lipoprotein lipase (LPL). Whereas cld affects an unknown gene, lec23 abolishes the activity of alpha-glucosidase I, an enzyme essential for proper folding and assembly of nascent glycoproteins. The hypothesis that cld, like lec23, affects the folding/assembly of nascent LPL was confirmed by showing that in cell lines homozygous for these mutations (Cld and Lec23, respectively), the majority of LPL was inactive, displayed heterogeneous aggregation, and had a decreased affinity for heparin. While inactive LPL was retained in the ER, a small amount of LPL that had attained a native conformation was transported through the Golgi and secreted. Thus, Cld and Lec23 cells recognized and retained the majority of LPL as misfolded, maintaining the standard of quality control. Examination of candidate factors affecting protein maturation, such as glucose addition and trimming, proteins involved in lectin chaperone cycling, and other abundant ER chaperones, revealed that calnexin levels were dramatically reduced in livers from cld/cld mice; this finding was also confirmed in Cld cells.We conclude that cld may affect components in the ER, such as calnexin, that play a role in protein maturation. Whether the reduced calnexin levels per se contribute to the LPL deficiency awaits confirmation.  相似文献   

4.
Nordihydroguaiaretic acid (NDGA), an inhibitor of lipoxygenase, inhibits the secretion of proteins and causes the redistribution of resident Golgi proteins into the endoplasmic reticulum (ER). In this study, the effect of NDGA on lipoprotein lipase (LPL) secretion was investigated in 3T3-L1 adipocytes, and compared with those of brefeldin A (BFA), a well-known fungal metabolite that exhibits similar ER-Golgi redistribution. Both BFA and NDGA blocked secretions of LPL. In the presence of BFA, the active and dimeric LPL was accumulated in adipocytes. After endoglycosidase H (endo H) digestion, the proportion of LPL subunits with partially endo H-sensitive oligosaccharide was significantly increased with BFA. However, in the presence of NDGA, the cellular LPL became inactive, and only the endo H-sensitive fraction of the LPL subunit was observed. An increase of the aggregated forms was observed in the fractions of the sucrose-density gradient ultracentrifugation. These properties of LPL in the NDGA-treated cells were similar to those of LPL that is retained in ER, and the effects of NDGA could not be reversed by BFA. These results indicate that the inhibitory mechanism of NDGA on the LPL secretion is functionally different from the ER-Golgi redistribution that is induced by BFA.  相似文献   

5.
Lipoprotein lipase (LPL) plays a key role in lipid metabolism. Molecular modeling of dimeric LPL was carried out using insight ii based upon the crystal structures of human, porcine, and horse pancreatic lipase. The dimeric model reveals a saddle-shaped structure and the key heparin-binding residues in the amino-terminal domain located on the top of this saddle. The models of two dimeric conformations - a closed, inactive form and an open, active form - differ with respect to how surface-loop positions affect substrate access to the catalytic site. In the closed form, the surface loop covers the catalytic site, which becomes inaccessible to solvent. Large conformational changes in the open form, especially in the loop and carboxyl-terminal domain, allow substrate access to the active site. To dissect the structure-function relationships of the LPL carboxyl-terminal domain, several residues predicted by the model structure to be essential for the functions of heparin binding and substrate recognition were mutagenized. Arg405 plays an important role in heparin binding in the active dimer. Lys413/Lys414 or Lys414 regulates heparin affinity in both monomeric and dimeric forms. To evaluate the prediction that LPL forms a homodimer in a 'head-to-tail' orientation, two inactive LPL mutants - a catalytic site mutant (S132T) and a substrate-recognition mutant (W390A/W393A/W394A) - were cotransfected into COS7 cells. Lipase activity could be recovered only when heterodimerization occurred in a head-to-tail orientation. After cotransfection, 50% of the wild-type lipase activity was recovered, indicating that lipase activity is determined by the interaction between the catalytic site on one subunit and the substrate-recognition site on the other.  相似文献   

6.
Lipoprotein lipase (LPL) acts at the vascular endothelium. Earlier studies have shown that down-regulation of adipose tissue LPL during fasting is post-translational and involves a shift from active to inactive forms of the lipase. Studies in cell systems had indicated that during fasting LPL might be retained in the endoplasmic reticulum. We have now explored the relation between active/inactive and intra/extracellular forms of the lipase. Within adipocytes, neither LPL mass nor the distribution of LPL between active and inactive forms changed on fasting. Extracellular LPL mass also did not change significantly, but shifted from predominantly active to predominantly inactive. To explore if changes in secretion were compensated by changes in turnover, synthesis of new protein was blocked by cycloheximide. The rates at which intra- and extracellular LPL mass and activity decreased did not change on fasting. To further explore how LPL is distributed in the tissue, heparin (which detaches the enzyme from the endothelial surface) was injected. Tissue LPL activity decreased by about 10% in 2 min and by 50% in 1 h. Heparin released mainly the active form of the lipase. There was no change of LPL activity or mass within adipocytes. The fraction of extracellular LPL that heparin released and the time course were the same in fed and fasted rats, indicating that active, extracellular LPL was distributed in a similar way in the two nutritional states. This study suggests that the nutritional regulation of LPL in adipose tissue determines the activity state of extracellular LPL.  相似文献   

7.
The regulation of adipose tissue lipoprotein lipase (LPL) was examined in rats fed or fasted overnight, and was found to be controlled posttranslationally. LPL catalytic activity decreased by 50% after fasting while LPL mRNA levels and rates of synthesis increased nearly 2-fold; enzyme mass remained unchanged. The distribution of LPL within the endoplasmic reticulum (ER) and Golgi/post-Golgi secretory pathway was assessed by differentiating between LPL high mannose and complex forms. After fasting, the majority of LPL is in the high mannose ER form (65%, 0.97 micrograms/g wet weight tissue), whereas the LPL complex form comprises only 35% (or 0.52 micrograms/g). After refeeding, however, the Golgi-derived LPL complex form predominates (65%, 1.03 micrograms/g) over the high mannose ER form (35%, 0.55 micrograms/g). Kinetic analysis suggests that high mannose LPL disappears with a half-life of t0.5 = 40 min in both fed and fasted rats, indicating that the redistribution of LPL mass during feeding/fasting does not arise by differential retention within ER. Instead, the fractional catabolic rate of complex LPL within the Golgi/post-Golgi secretory compartment can be calculated to be 3.5-fold greater in fasting. In heart, changes in LPL activity in response to feeding/fasting are also not due to differences in mRNA levels or rates of synthesis. Based on these findings, a model of LPL posttranslational regulation is proposed and discussed.  相似文献   

8.
Lipoprotein lipase (LPL), an enzyme playing the central role in triglyceride metabolism, is a glycoprotein and a homodimer of identical subunits. Dimerization and proper processing of oligosaccharide chains are important maturation steps in post-translational regulation of enzyme activity. Indirect evidences suggest that dimerization of LPL occurs in endoplasmic reticulum (ER) or Golgi. In this study, we investigated the dimerization status of LPL in 3T3-L1 adipocytes, using sucrose density gradient ultracentrifugation and carbonyl cyanide m-chlorophenylhydrazone (CCCP), an inhibitor of ER-Golgi protein transport. In the presence of CCCP, no increase of cellular LPL activity was detected during 2 h of recovery period after the depletion of LPL with heparin and cycloheximide. Only endoglycosidase H (endo H)-sensitive subunits were found in CCCP-treated cells after endo H digestion, suggesting that inactive LPL was retained in ER. In the presence of castanospermine, an inhibitor of ER glucosidase I, LPL subunits of both control and CCCP-treated cells had same molecular weight, indicating that complete oligosaccharides were transferred to LPL subunits in the presence of CCCP. In sucrose density gradient ultracentrifugation, all the LPL protein synthesized in the presence of CCCP was found at the dimeric fractions as in control cells. Most of LPL protein in control cells showed high affinity for heparin, and there was no difference between the control and CCCP-treated cells. These results suggest that dimerization and acquisition of high affinity for heparin of LPL can occur in ER of CCCP-treated cells without acquisition of catalytic activity.  相似文献   

9.
Low density lipoprotein (LDL) and oxidized LDL are associated with collagen in the arterial intima, where the collagen is coated by the small proteoglycan decorin. When incubated in physiological ionic conditions, decorin-coated collagen bound only small amounts of native and oxidized LDL, the interaction being weak. When decorin-coated collagen was first allowed to bind lipoprotein lipase (LPL), binding of native and oxidized LDL increased dramatically (23- and 7-fold, respectively). This increase depended on strong interactions between LPL that was bound to the glycosaminoglycan chains of the collagen-bound decorin and native and oxidized LDL (kDa 12 and 5.9 nM, respectively). To distinguish between binding to monomeric (inactive) and dimeric (catalytically active) forms of LPL, affinity chromatography on heparin columns was conducted, which showed that native LDL bound to the monomeric LPL, whereas oxidized LDL, irrespective of the type of modification (Cu(2+), 2, 2'-azobis(2-amidinopropane)hydrochloride, hypochlorite, or soybean 15-lipoxygenase), bound preferably to dimeric LPL. However, catalytic activity of LPL was not required for binding to oxidized LDL. Finally, immunohistochemistry of atherosclerotic lesions of human coronary arteries revealed specific areas in which LDL, LPL, decorin, and collagen type I were present. The results suggest that LPL can retain LDL in atherosclerotic lesions along decorin-coated collagen fibers.  相似文献   

10.
Most missense mutations of the lipoprotein lipase (LPL) gene identified among LPL-deficient subjects cluster in a segment of the sequence that encodes the catalytic triad as well as functional elements involved in the activation of the lipase at lipid-water interfaces. Consequently, loss of activity may result either from direct alterations of such functional elements or from less specific effects on protein folding and stability. This issue was addressed by examining biochemical properties of four such variants (A176T, G188E, G195E, and S244T) in a heterologous expression system (COS-1 cells). Variant G195E (GGA----GAA) was previously unreported. In all instances, inactive enzyme was recovered in medium, albeit at reduced levels. Cellular synthesis and extracellular degradation were similar to those for wild type, suggesting that reduced secretion resulted from increased intracellular degradation. When cell extracts were subjected to heparin-Superose affinity chromatography followed by elution on a linear salt gradient, all variants exhibited a single, inactive, low affinity immunoreactive peak. By contrast, wild-type enzyme presented an additional, high affinity, active species, which we interpret as homodimeric enzyme. Substitution of the active-site serine (S132A) led to loss of activity but maintenance of the high affinity species. When large amounts of the G188E variant were applied to the column, small but significant amounts of high affinity, active enzyme were recovered. Systematic substitutions at residue 188 showed that only glycine could accommodate structural constraints at this position. We conclude that the mutations examined did not impart lipase deficiency by affecting specific functional elements of the enzyme. Rather, they appear to affect protein folding and stability, and thereby formation and maintenance of subunit assembly.  相似文献   

11.
Among three lipases in the lipase gene family, hepatic lipase (HL), lipoprotein lipase, and pancreatic lipase, HL exhibits the lowest intracellular specific activity (i.e. minimal amounts of catalytic activity accompanied by massive amounts of inactive lipase mass in the endoplasmic reticulum (ER)). In addition, HL has a distinctive sedimentation profile, where the inactive mass overlaps the region containing active dimeric HL and trails into progressively larger molecular forms. Eventually, at least half of the HL inactive mass in the ER reaches an active, dimeric conformation (t(1/2) = 2 h) and is rapidly secreted. The remaining inactive mass is degraded. HL maturation occurs in the ER and is strongly dependent on binding to calnexin in the early co-/post-translational stages. Later stages of HL maturation occur without calnexin assistance, although inactive HL at all stages appears to be associated in distinct complexes with other ER proteins. Thus, unlike other lipases in the gene family, HL maturation is the rate-limiting step in its secretion as a functional enzyme.  相似文献   

12.
13.
U Tatu  C Hammond    A Helenius 《The EMBO journal》1995,14(7):1340-1348
Influenza hemagglutinin (HA) was used to analyze the stepwise folding and oligomeric assembly of glycoproteins in the early secretory pathway of living cells. In addition to mature trimers, six distinct maturation intermediates were identified. Of these, all the incompletely oxidized forms were located in the endoplasmic reticulum (ER) and associated with calnexin, a membrane-bound, lectin-like ER chaperone. Once fully oxidized, the HA dissociated from calnexin as a monomer, which rapidly became resistant to dithiothreitol (DTT) reduction. Part of these extensively folded molecules moved as monomers into the intermediate compartment between the ER and the Golgi complex. Assembly of homotrimers occurred without calnexin-involvement within the ER and in the intermediate compartment. When anchor-free HA molecules were analyzed, it was found that they reach the DTT-resistant monomeric conformation but fail to trimerize. Taken together, the results provide a definition and intracellular localization of several intermediates in the conformational maturation of HA, including the immediate precursor for trimer assembly.  相似文献   

14.
15.
Lipoprotein lipase (LPL) is dependent on apolipoprotein CII (apoCII), a component of plasma lipoproteins, for function in vivo. The hydrophobic fluorescent probe 1,1'-bis(anilino)-4,4'-bis(naphthalene)-8,8'-disulfonate (bis-ANS) was found to be a potent inhibitor of LPL. ApoCII prevented the inhibition by bis-ANS, and was also able to restore the activity of inhibited LPL in a competitive manner, but only with triacylglycerols with acyl chains longer than three carbons. Studies of fluorescence and surface plasmon resonance indicated that LPL has an exposed hydrophobic site for binding of bis-ANS. The high affinity interaction was characterized by an equilibrium constant Kd of 0.10-0.26 microm and by a relatively high on rate constant kass = 2.0 x 10(4) m(-1) s(-1) and a slow off-rate with a dissociation rate constant kdiss = 1.2 x 10(-4) s(-1). The high affinity binding of bis-ANS did not influence interaction of LPL with heparin or with lipid/water interfaces and did not dissociate the active LPL dimer into monomers. Analysis of fragments of LPL after photoincorporation of bis-ANS indicated that the high affinity binding site was located in the middle part of the N-terminal folding domain. We propose that bis-ANS binds to an exposed hydrophobic area that is located close to the active site. This area may be the binding site for individual substrate molecules and also for apoCII.  相似文献   

16.
We have previously shown that transgenic expression of catalytically inactive lipoprotein lipase (LPL) in muscle (Mck-N-LPL) enhances triglyceride hydrolysis as well as whole particle lipoprotein and selective cholesterol ester uptake. In the current study, we have examined whether these functions can be performed by inactive LPL alone or require the presence of active LPL expressed in the same tissue. To study inactive LPL in the presence of active LPL in the same tissue, the Mck-N-LPL transgene was bred onto the heterozygous LPL-deficient (LPL1) background. At 18 h of age, Mck-N-LPL reduced triglycerides by 35% and markedly increased muscle lipid droplets. In adult mice, it reduced triglycerides by 40% and increased lipoprotein particle uptake into muscle by 60% and cholesterol ester uptake by 110%. To study inactive LPL alone, the Mck-N-LPL transgene was bred onto the LPL-deficient (LPL0) background. These mice die at approximately 24 h of age. At 18 h of age, in the absence of active LPL, inactive LPL expression did not diminish triglycerides nor did it result in the accumulation of muscle lipid droplets. To study inactive LPL in the absence of active LPL in the same tissue in adult animals, the Mck-N-LPL transgene was bred onto mice that only expressed active LPL in the heart (LPL0/He-LPL). In this case, Mck-N-LPL did not reduce triglycerides or increase the uptake of lipoprotein particles but did increase muscle uptake of chylomicron and very low density lipoprotein cholesterol ester by 40%. Thus, in the presence of active LPL in the same tissue, inactive LPL augments triglyceride hydrolysis and increases whole particle triglyceride-rich lipoprotein and selective cholesterol ester uptake. In the absence of active LPL in the same tissue, inactive LPL only mediates selective cholesterol ester uptake.  相似文献   

17.
Three forms of acetylcholinesterase (AChE) were detected in samples of the bivalve mollusc Mytilus galloprovincialis collected in sites of the Adriatic sea. Apart from the origin of the mussels, two spontaneously soluble (SS) AChE occur in the hemolymph and represent about 80% of total activity, perhaps hydrolyzing metabolism-borne choline esters. These hydrophilic enzymes (forms A and B) copurified by affinity chromatography (procainamide-Sepharose gel) and were separated by sucrose gradient centrifugation. They are, respectively, a globular tetramer (11.0-12.0 S) and a dimer (6.0-7.0 S) of catalytic subunits. The third form, also purified from tissue extracts by the same affinity matrix, proved to be an amphiphilic globular dimer (7.0 S) with a phosphatidylinositol tail giving cell membrane insertion, detergent (Triton X-100, Brij 96) interaction and self-aggregation. Such an AChE is likely functional in cholinergic synapses. All three AChE forms show a good substrate specificity and are inactive on butyrylthiocholine. Studies with inhibitors showed low inhibition by eserine and paraoxon, especially on SS forms, high sensitivity to 1,5-bis(4-allyldimethylammoniumphenyl)-pentan-3-one dibromide (BW284c51) and no inhibition with propoxur and diisopropylfluorophosphate (DFP). The ChE forms in M. galloprovincialis are possibly encoded by different genes. Some kinetic features of these enzymes suggest a genetic polymorphism.  相似文献   

18.
1. The isolation of NADP-linked malic enzyme (EC 1.1.1.40) from maize leaves is described, together with studies of its Mr and subunit composition. 2. The enzyme was purified to apparent homogeneity by affinity chromatography on N6-aminohexyl-2',5'-bisphosphoadenosine-agarose, gel filtration with Sephadex G-100 and ion-exchange chromatography on DEAE-Sephadex A-50. A purification of 140-fold with a 30% yield was obtained. 3. A detailed study of the Mr by several methods revealed the existence of different Mr forms in solution. 4. In the presence of dithiothreitol the enzyme appears to be present in triethanolamine buffer, pH 7.5, as a tetramer with a subunit Mr of 60,000 and an S20,w of 10.75 S. 5. In phosphate buffer, pH 7.0, it seems to be a dimer of Mr 120,000 with an S20,w of 7.95 S. 6. In the absence of dithiothreitol, lower-Mr forms were detected by sedimentation-equilibrium and sedimentation-velocity studies in triethanolamine buffer. 7. Results from gel filtration gave Mr values of about 340,000 in both buffers.  相似文献   

19.
The variant surface glycoprotein of African trypanosomes is released after overnight incubation of parasites at 4 degrees C in pH 5.5 phosphate glucose buffer and may be purified by Concanavalin A Sepharose affinity chromatography. The addition of proteinase inhibitors during the parasite incubation is necessary to prevent the proteolysis of the variant surface glycoprotein by the trypanosomal released proteinases. Using this procedure without the addition of proteinase inhibitors, the proteolytic activities, released from the bloodstream forms Trypanosoma brucei brucei variant AnTat 1.1, were separated by Concanavalin-A Sepharose affinity chromatography. The unretained material (F1) shows hydrolytic activity against the two synthetic substrates Z-Phe-Arg-AMC and Z-Arg-Arg-AMC, which is stimulated by dithiothreitol, but not inhibited by E-64, and characterized by an alkaline pH optimum and an estimated molecular mass of 80-100 kDa. The Michaelis constant for the substrates Z-Arg-Arg-AMC and Z-Phe-Arg-AMC was, respectively, 2.8 and 6.7 microM. The retained material eluted by addition of 1% methyl-alpha-D-mannopyranoside (F2) shows hydrolytic activity against the synthetic substrate Z-Phe-Arg-AMC, which is stimulated by dithiothreitol, inhibited by E-64, active between pH 6.0 and 8.0, and could be separated into two peaks of activity by HPLC, one peak of high molecular mass (greater than 70 kDa) and the other peak of lower molecular mass (30-70 kDa). By electrophoresis in gels containing gelatin as substrate, this fraction contains several proteins with gelatinolytic activity, whereas the unretained fraction F1 did not have any gelatinolytic activity.  相似文献   

20.
Tyrosinase, the rate-limiting enzyme in mammalian melanogenesis, is a copper-containing transmembrane glycoprotein. Tyrosinase undergoes a complex post-translational processing before reaching the melanosomal membrane. This processing involves N-glycosylation in several sites, including one located in the CuB copper binding site, movement from the endoplasmic reticulum (ER) to the Golgi, copper binding, and sorting to the melanosome. Aberrant processing is causally related to the depigmented phenotype of human melanomas. Moreover, some forms of albinism and several other pigmentary syndromes are considered ER retention diseases or trafficking defects. A critical step in tyrosinase maturation is the acquisition of an ER export-competent conformation recognized positively by the ER quality control system. However, the minimal structural requirements allowing exit from the ER to the Golgi have not yet been identified for tyrosinase or other melanosomal proteins. We addressed this question by analyzing the enzymatic activity and glycosylation pattern of mouse tyrosinase point mutants and chimeric constructs, where selected portions of tyrosinase were replaced by the homologous fragments of the highly similar tyrosinase-related protein 1. We show that a completely inactive tyrosinase point mutant lacking a critical histidine residue involved in copper binding is nevertheless able to exit from the ER and undergo further processing. Moreover, we demonstrate that tyrosinase displays at least two sites whose glycosylation is post-translational and most likely conformation-dependent and that a highly specific interaction involving the CuB site is essential not only for correct glycosylation but also for exit from the ER and enzymatic activity.  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