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 390 毫秒
1.
Previous studies have postulated the presence of two bile salt-binding sites regulating the activity of the pancreatic bile salt-dependent lipase. One of these sites, located in an N-terminal basic cluster, has been identified as the specific bile salt-binding site. Interaction of primary bile salts with this proximal site induces the formation of a micellar binding site from a pre-existing nonspecific or pre-micellar bile salt-binding site. Here we have investigated the functional significance of another basic cluster comprised of amino acid residues Arg(423), Lys(429), Arg(454), Arg(458), and Lys(462), distal from the catalytic site. For this purpose these residues were mutagenized in Ile or Ala residues. The mutagenized enzyme lost activity on both soluble and emulsified substrates in the presence of bile salts. However, in the absence of bile salts, the mutagenized enzyme displayed the same activity on soluble substrate as the wild-type recombinant enzyme. Consequently, the distal basic cluster may represent the nonspecific (or pre-micellar) bile salt-binding site susceptible to accommodate primary and secondary bile salts. According to the literature, tyrosine residue(s) should participate in this site. Therefore, two tyrosine residues, Tyr(427) and Tyr(453), associated with the distal basic cluster were also mutagenized. Each tyrosine substitution to serine did not inhibit the enzyme activity on soluble substrate, independently of the presence of primary or secondary bile salts. However, the enzyme activity on cholesteryl oleate solubilized in primary bile salt micelles was decreased by mutations substantiating that these residues are part of the nonspecific bile salt-binding site.  相似文献   

2.
Modification of arginine residues with 2,3-butanedione inhibits the carboxylic-ester hydrolase activity on soluble and emulsified substrates when assayed with bile salts. The alpha-dicarbonyl reagent modifies seven of the nineteen arginine residues present per enzyme molecule. Nevertheless the inactivation with butanedione is greatly diminished when the protein is in the presence of negatively charged micellar bile salt. In these conditions we observe the protection of one arginine residue by sodium taurodeoxycholate and of two arginine residues by sodium cholate. This suggests that the carboxylic-ester hydrolase from human pancreatic juice contains at least two arginine residues essential for the activation by bile salts. All our data confirm the presence of two bile-salt-binding sites on the enzyme in which one arginine per site is involved and plays the general role of an anionic binding site. This study provides evidence that arginine residues may play an essential role in the interaction between bile salts and protein.  相似文献   

3.
Through labeling with the sodium salt of the photolabile bile salt derivative (7,7-azo-3 alpha,12 alpha-dihydroxy-5 beta-[3 beta-3H]cholan-24-oyl)- 2-aminoethanesulfonic acid, a bile salt-binding polypeptide with an apparent molecular weight of 100,000 was identified in isolated canalicular but not basolateral (sinusoidal) rat liver plasma membranes. This labeled polypeptide was isolated from octyl glucoside-solubilized canalicular membranes by DEAE-cellulose and subsequent wheat germ lectin Sepharose chromatography. The purified protein still contained covalently incorporated radioactive bile salt derivative and exhibited a single band with an apparent molecular weight of 100,000 on sodium dodecyl sulfate-gels. Antibodies were raised in rabbits and their monospecificity toward this canalicular polypeptide demonstrated by immunoblot analysis. No cross-reactivity was found with basolateral membrane proteins. The antibodies inhibited taurocholate uptake into isolated canalicular but not basolateral membrane vesicles. In addition, the antibodies also decreased efflux of taurocholate from canalicular vesicles. If the canalicular bile salt-binding polypeptide was immunoprecipitated from Triton X-100-solubilized canalicular membranes and subsequently deglycosylated with trifluoromethanesulfonic acid, the apparent molecular weight was decreased from 100,000 to 48,000 (sodium dodecyl sulfate-polyacrylamide gel electrophoresis). These studies confirm previous results in intact liver tissue and strongly indicate that a canalicular specific glycoprotein with an apparent molecular weight of 100,000 is directly involved in canalicular excretion of bile salts.  相似文献   

4.
Tyrosine residues of the human pancreatic carboxylic-ester hydrolase (EC 3.1.1.1) (also referred to as cholesterol-ester hydrolase, EC 3.1.1.13) were nitrated in the ortho-position by the use of tetranitromethane. The specificity of the reaction has been verified and the inhibition observed was shown to be unrelated to the weak polymerization of the protein. Among the 27 tyrosines present in the enzyme, seven or eight were nitrated but only one residue, with a pK of 8.3, seems to be responsible for the loss of activity. This decrease in enzyme activity appears only in assays which were performed in the presence of bile salts, suggesting that of the two bile salt binding sites postulated on the enzyme, only one, referred to the as the 'unspecific site' (Lombardo, D. and Guy, O. (1980) Biochim. Act 611, 147-155), was modified. This is in agreement with the similar loss of enzyme activity observed on emulsified and soluble substrate. The most important result is the difference observed in experiments of the protective effects of bile salts. The protection with sodium taurodeoxycholate is independent of its critical micellar concentration, showing that monomers protect this site, whereas the protection observed in experiments with sodium cholate appears only for supramicellar concentrations of bile salt. Since this latter bile salt promotes the dimerization of the enzyme, we can conclude that a premicellar bile salt binding site (protected by monomers) is transformed in a functional micellar binding site (protected by micelles). This conformational transformation seems to be consecutive to the dimerization, as has been recently proposed.  相似文献   

5.
The properties of sodium fusidate micelles were determined by a spectral shift technique, surface tension measurements, and ultracentrifugal analysis. The critical micellar concentrations, mean molecular areas, and apparent aggregation numbers were estimated as a function of the concentration of counterion (0.001-1.0 m Na(+)) at 20 degrees C. The critical micellar concentrations were studied over a temperature range of 10 degrees C to 40 degrees C at one counterion concentration (0.001 m Na(+)), and from these data the standard thermo-dynamic functions of micellization were calculated. The ability of sodium fusidate solutions to solubilize the insoluble swelling amphiphiles, lecithin and monoolein, was investigated, and the results were compared with the solubilizing properties of sodium taurocholate. The critical micellar concentrations of sodium fusidate approximated those of sodium taurocholate. The values fell in the range of 1.44-4.56 mm, varying with the technique used, counterion concentration, and temperature. The percentage of counterions bound to fusidate micelles in water, calculated from the log critical micellar concentration-log Na(+) curve, was estimated to be negligible, which compares with sodium taurocholate micelles. The critical micellar concentration of sodium fusidate exhibited a minimum at 20 degrees C, a phenomenon observed with other ionic detergents and with bile salts. Micelle formation in sodium fusidate solutions was shown to be primarily entropy-driven at 10 degrees and 20 degrees C, whereas at 30 degrees and 40 degrees C the enthalpy factor predominated. From the surface tension measurements the molecular areas of sodium fusidate and sodium taurocholate were calculated. The mean molecular area of fusidate was 101 A(2), whereas sodium taurocholate possessed a molecular area of 88 A(2). It was demonstrated that the sodium fusidate molecule, like a bile salt molecule, lies with its longitudinal axis horizontal at an air-water interface. The apparent aggregation number of sodium fusidate micelles increased from 5 to 16 as the concentration of counterion increased from 0.01 to 0.60 m Na(+). These values are slightly larger than the corresponding aggregation numbers of sodium taurocholate micelles.  相似文献   

6.
A liquid membrane electrode is described which can measure the activity of biological detergents in solution. The results for two bile salts, sodium taurocholate and sodium taurodeoxycholate, are compared with those for a conventional detergent, sodium dodecyl sulfate. The activity of the bile anion is also measured in a mixed micellar solution (phosphatidylcholine added).  相似文献   

7.
Structural basis for bile salt inhibition of pancreatic phospholipase A2   总被引:1,自引:0,他引:1  
Bile salt interactions with phospholipid monolayers of fat emulsions are known to regulate the actions of gastrointestinal lipolytic enzymes in order to control the uptake of dietary fat. Specifically, on the lipid/aqueous interface of fat emulsions, the anionic portions of amphipathic bile salts have been thought to interact with and activate the enzyme group-IB phospholipase A2 (PLA2) derived from the pancreas. To explore this regulatory process, we have determined the crystal structures of the complexes of pancreatic PLA2 with the naturally occurring bile salts: cholate, glycocholate, taurocholate, glycochenodeoxycholate, and taurochenodeoxycholate. The five PLA2-bile salt complexes each result in a partly occluded active site, and the resulting ligand binding displays specific hydrogen bonding interactions and extensive hydrophobic packing. The amphipathic bile salts are bound to PLA2 with their polar hydroxyl and sulfate/carboxy groups oriented away from the enzyme's hydrophobic core. The impaired catalytic and interface binding functions implied by these structures provide a basis for the previous numerous observations of a biphasic dependence of the rate of PLA2 catalyzed hydrolysis of zwitterionic glycerophospholipids in the presence of bile salts. The rising or activation phase is consistent with enhanced binding and activation of the bound PLA2 by the bile salt induced anionic charge in a zwitterionic interface. The falling or inhibitory phase can be explained by the formation of a catalytically inert stoichiometric complex between PLA2 and any bile salts in which it forms a stable complex. The model provides new insight into the regulatory role that specific PLA2-bile salt interactions are likely to play in fat metabolism.  相似文献   

8.
Biochemical site of regulation of bile acid biosynthesis in the rat   总被引:15,自引:0,他引:15  
The production of bile salts by rat liver is regulated by a feedback mechanism, but it is not known which enzyme controls endogenous bile acid synthesis. In order to demonstrate the biochemical site of this control mechanism, bile fistula rats were infused intravenously with (14)C-labeled bile acid precursors, and bile acid biosynthesis was inhibited as required by intraduodenal infusion of sodium taurocholate. The infusion of taurocholate (11-14 mg/100 g of rat per hr) inhibited the incorporation of acetate-1-(14)C, mevalonolactone-2-(14)C, and cholesterol-4-(14)C into bile acids by approximately 90%. In contrast, the incorporation of 7alpha-hydroxycholesterol-4-(14)C into bile acids was reduced by less than 10% during taurocholate infusion. These results indicate that the regulation of bile acid biosynthesis is exerted via cholesterol 7alpha-hydroxylase provided that hepatic cholesterol synthesis is adequate.  相似文献   

9.
The effect of individual bile salts on alpha-amylase hydrolysis of Cibachron Blue starch was studied at pH 6.0. With sodium cholate, taurocholate and taurodeoxycholate, enzyme activity was increased to 150-160 percent of the control value, at a concentration of similar to 1 mmol/l bile salt. The increased activity extended up to 4 mmol/l. The bile salts sodium deoxycholate and taurochenodeoxycholate exerted activation and inhibition depending on the concentration. With deoxycholate (0.75 mmol/l), activation (150 percent) was evident, while inhibition was apparent above 2.5 mmol/l. With taurochenodeoxycholate maximum activity (135 percent) was observed at 0.25 mmol/l, while inhibition was evident above 1.5 mmol/l. Chenodeoxycholate and lithocholate exerted marked inhibition at concentrations as low as 0.5 mmol/l. Inhibition of alpha-amylase by chenodeoxycholate was competitive with both soluble and insoluble starch substrates. Since the pH of the jejunum is in the region of 6.0 the phenomenon of activation and inhibition of alpha-amylase by bile salts at this pH could be of physiological significance.  相似文献   

10.
The biliary excretion of bile salts, lysosomal acid phosphatase, and total proteins were studied in rats under different experimental conditions: during bile salt loss through a bile fistula and after loading with exogenous sodium taurocholate. The experimental models were suitable to demonstrate that variations in the excretion of bile salts were associated with those of acid phosphatase output. During bile salt depletion, acid phosphatase output showed a decrease parallel to that of bile salts. Following a single i.v. injection of sodium taurocholate and during its i.v. infusion, a rapid increase of acid phosphatase excretion in bile was seen. The patterns of enzyme outputs observed after administration of sodium taurocholate suggested a bulk discharge in bile of lysosomal contents. The profiles of protein output were similar to those of acid phosphatase suggesting an association between the secretory mechanism of these bile constituents. In contrast to sodium taurocholate, 4-methylumbelliferone, which also increases canalicular bile flow, did not produce changes in the excretory patterns of the bile components studied. Therefore, the results suggested a bile salt related secretion of acid phosphatase in the rat, which may involve protein secretion in bile.  相似文献   

11.
Nonspecific high affinity binding of bile salts to carboxylester lipases   总被引:1,自引:0,他引:1  
The interactions with bile salts of carboxylester lipases (EC 3.1.1.13) from human pancreatic juice and pig pancreas were characterized by physical methods. Bile salts cause a decrease in the fluorescence intensity of the proteins at the emission maximum of 333-335 nm. The concentration dependence of this decrease shows saturation behavior, is relatively nonspecific with respect to bile salt conjugation or the presence of the 7 alpha-hydroxyl group, and is consistent with a 1:1 interaction between enzyme and bile salt. Direct measurement of the binding of [3H]cholate by equilibrium dialysis supports the stoichiometry. Other detergents also bind, causing fluorescence changes, but with much lower affinities. Binding of taurocholate to the monomeric pig enzyme is enhanced by increasing ionic strength, indicating the predominance of hydrophobic interactions. In the range of pH 5.5-6.8, binding is pH-independent with dissociation constants of 3-20 microM. At higher pH, affinity is greatly reduced and the fluorescence spectrum changes, indicating the importance of a protonated group for efficient interaction. Occupancy of the bile salt binding site partially stabilizes the enzyme against inactivation by heat but not trypsin. However, circular dichroism spectra do not indicate that bile salt binding is accompanied by any change in secondary structure. The monomeric pig enzyme binds to the argon/water interface in the presence of bile salts and binding of taurocholate to diisopropylphosphoryl-enzyme is similar to that measured with native enzyme. These results suggest that surface binding and catalysis occur at sites distinct from the bile salt binding site of the enzyme. Stabilization of the monomeric pig enzyme against denaturation at high energy surfaces occurs concomitantly with occupancy of the bile salt binding site. Overall, the data suggest that an important role of bile salts in vivo is to stabilize these enzymes at lipid-water interfaces.  相似文献   

12.
The enzymatic basis for cod digestive lipolysis has been investigated. Lipase activity was found in aqueous extracts from pyloric caeca as well as in pancreatic tissue surrounding the caeca and the bile duct. A bile salt-dependent lipase (BSDL) was purified from either defatted powder of cod pyloric caeca or aqueous pancreatic extracts by combined affinity chromatography on cholate-Sepharose and gel filtration on Sephacryl S-200 HR. By SDS-PAGE analysis the molecular weight of purified cod BSDL was estimated to 60 kDa. The enzyme was totally dependent on bile salts for hydrolysis of insoluble fatty acid esters. Antiserum raised against purified cod BSDL reacted specifically with selected mammalian pancreatic BSDLs by Western blot analysis. Results presented in this paper strongly suggest that the bile salt-dependent lipase is the only pancreatic enzyme involved in lipid digestion in cod. The enzyme has been characterized and compared to human pancreatic BSDL with respect to substrate specificity, temperature- and pH-dependence and inhibitors. Both soluble and insoluble fatty acid esters were hydrolysed and the enzyme was 1,3-specific in hydrolysis of triolein. The enzyme was inhibited by di-isopropyl fluorophosphate and phenyl boronic acid, but not significantly by phenyl methyl sulfonyl fluoride. The cod BSDL is probably homologous to mammalian pancreatic BSDLs.  相似文献   

13.
Bilirubin, the yellow-orange tetrapyrrole pigment of jaundice, is essentially insoluble in pure water, but is much more soluble in solutions of bile salts such as sodium taurocholate. The biophysical chemistry of bilirubin in bile salt solutions is affected by changes in the pH of the solution in the range 5-9, suggesting that interactions with bile salt molecules and micelles may alter the acidity of the pigment. We have examined this possibility by determining the apparent pKa values for a series of carboxyl 13C-enriched model compounds, including the bilirubin analog mesobilirubin XIIIalpha, in solutions of sodium taurocholate and sodium taurodeoxycholate. Apparent pKa values were determined by 13C NMR titrations in dimethyl sulfoxide-water mixtures. The results show that the acidity of all compounds is decreased, or pKa increased, in micellar bile salt solution relative to pure water and that the effect is greatest for the larger, less water-soluble compounds. We have proposed a model to explain these results and discussed the implications of these findings for the biophysical chemistry of bilirubin in bile.  相似文献   

14.
The effects of the bile salts, sodium taurocholate (NaTc) and sodium cholate (NaCh), and toad bile gallbladder (bile) on short-circuit current (SCC) across isolated skin, and sodium taurocholate (NaTc) on isolated bladder of Bufo arenarum toads were tested. Sodium taurocholate (NaTc), sodium cholate (NaCh) and toad bile gallbladder (bile) promoted an increase in SCC, when added to the external side. The stimulatory effect was reversible after rinsing the preparation for 60 min. Implications on in vivo renal function of these results are discussed.  相似文献   

15.
Exposure of Salmonella enterica to sodium cholate, sodium deoxycholate, sodium chenodeoxycholate, sodium glycocholate, sodium taurocholate, or sodium glycochenodeoxycholate induces the SOS response, indicating that the DNA-damaging activity of bile resides in bile salts. Bile increases the frequency of GC --> AT transitions and induces the expression of genes belonging to the OxyR and SoxRS regulons, suggesting that bile salts may cause oxidative DNA damage. S. enterica mutants lacking both exonuclease III (XthA) and endonuclease IV (Nfo) are bile sensitive, indicating that S. enterica requires base excision repair (BER) to overcome DNA damage caused by bile salts. Bile resistance also requires DinB polymerase, suggesting the need of SOS-associated translesion DNA synthesis. Certain recombination functions are also required for bile resistance, and a key factor is the RecBCD enzyme. The extreme bile sensitivity of RecB-, RecC-, and RecA- RecD- mutants provides evidence that bile-induced damage may impair DNA replication.  相似文献   

16.
The effect of sulfate esterification of the 3 alpha- or 7 alpha-hydroxyl groups of taurochenodeoxycholate on calcium binding was studied at 0.154 M NaCl in the presence and absence of phosphatidylcholine using a calcium electrode. For comparison, similar studies were made with taurochenodeoxycholate, taurodeoxycholate, and taurocholate. No high affinity calcium binding was demonstrable for any of these bile salts in pre-micellar solutions. Taurine-conjugated bile salts have greater affinity for calcium when in a micellar form. At elevated bile salt concentrations, the calcium binding of unsulfated dihydroxy taurine conjugates was similar to that of the monosulfate esters of taurochenodeoxycholate. The presence of phosphatidylcholine decreased calcium binding of the unsulfated dihydroxy bile salts and slightly increased calcium binding by taurocholate. However, the addition of phosphatidylcholine to monosulfate esters of taurochenodeoxycholate results in large increments in calcium binding. The results indicate that increased calcium binding due to the presence of phosphatidylcholine in bile salt solutions depends, in part, on the hydrophilicity of the bile salt and that the interaction of monosulfate esters of taurochenodeoxycholate with phosphatidylcholine leads to the formation of a high affinity calcium binding site.  相似文献   

17.
Direct photoaffinity labeling of liver plasma membrane subfractions enriched in sinusoidal and canalicular membranes using [35S]adenosine 5'-O-(thiotriphosphate) ([35S]ATP gamma S) allows the identification of ATP-binding proteins in these domains. Comparative photoaffinity labeling with [35S]ATP gamma S and with the photolabile bile salt derivative (7,7-azo-3 alpha, 12 alpha-dihydroxy-5 beta-[3 beta-3H]-cholan-24-oyl-2'- aminoethanesulfonate followed by immunoprecipitation with a monoclonal antibody (Be 9.2) revealed the identity of the ATP-binding and the bile salt-binding canalicular membrane glycoprotein with the apparent Mr of 110,000 (gp110). The isoelectric point of this glycoprotein was 3.7. Transport of bile salt was studied in vesicles enriched in canalicular and sinusoidal liver membranes. Incubation of canalicular membrane vesicles with [3H] taurocholate in the presence of ATP resulted in an uptake of the bile salt into the vesicles which was sensitive to vanadate. ATP-dependent taurocholate transport was also observed in membrane vesicles from mutant rats deficient in the ATP-dependent transport of cysteinyl leukotrienes and related amphiphilic anions. Substrates of the P-glycoprotein (gp170), such as verapamil and doxorubicin, did not interfere with the ATP-dependent transport of taurocholate. Reconstitution of purified gp110 into liposomes resulted in an ATP-dependent uptake of [3H]taurocholate. These results demonstrate that gp110 functions as carrier in the ATP-dependent transport of bile salts from the hepatocyte into bile. This export carrier is distinct from hitherto characterized ATP-dependent transport systems.  相似文献   

18.
Hepatocyte growth factor (HGF) is a heparin-binding, multipotent growth factor that transduces a wide range of biological signals, including mitogenesis, motogenesis, and morphogenesis. Heparin or closely related heparan sulfate has profound effects on HGF signaling. A heparin-binding site in the N-terminal (N) domain of HGF was proposed on the basis of the clustering of surface positive charges [Zhou, H., Mazzulla, M. J., Kaufman, J. D., Stahl, S. J., Wingfield, P. T., Rubin, J. S., Bottaro, D. P., and Byrd, R. A. (1998) Structure 6, 109-116]. In the present study, we confirmed this binding site in a heparin titration experiment monitored by nuclear magnetic resonance spectroscopy, and we estimated the apparent dissociation constant (K(d)) of the heparin-protein complex by NMR and fluorescence techniques. The primary heparin-binding site is composed of Lys60, Lys62, and Arg73, with additional contributions from the adjacent Arg76, Lys78, and N-terminal basic residues. The K(d) of binding is in the micromolar range. A heparin disaccharide analogue, sucrose octasulfate, binds with similar affinity to the N domain and to a naturally occurring HGF isoform, NK1, at nearly the same region as in heparin binding. (15)N relaxation data indicate structural flexibility on a microsecond-to-millisecond time scale around the primary binding site in the N domain. This flexibility appears to be dramatically reduced by ligand binding. On the basis of the NK1 crystal structure, we propose a model in which heparin binds to the two primary binding sites and the N-terminal regions of the N domains and stabilizes an NK1 dimer.  相似文献   

19.
Pig pancreatic carboxylester lipase (cholesterol esterase, E.C. 3.1.1.13) was inactivated at a tributyrin/water interface. The apparent rate constant for inactivation increased with increase in the particle surface area of the tributyrin emulsion. The large energy of activation and entropy change for inactivation (33.7 Kcal.mol-1 and 35.8 cal.mol-1.deg-1, using 5 mM sonicated tributyrin at 37 degrees C, respectively) suggest that the observed inactivation reflects denaturation of the enzyme at the tributyrin/water interface. Bile salts protected the enzyme from irreversible inactivation at the tributyrin/water interface. The protection by bile salts was related both to their concentration and to the tributyrin concentration (substrate surface area). The protection by bile salts was not related to their concentration below or above their critical micellar concentration; the binding of bile salts to enzyme was probably the dominant protection factor. Similar stabilization was observed with other detergents such as Brij-35, Triton X-100, and sodium dodecyl sulfate. These results suggest that inactivation of carboxylester lipase occurs at a high-energy lipid-water interface and that an important role of bile salts in vivo is to stabilize carboxylester lipase at interfaces.  相似文献   

20.
The maximal micellar solubility, distribution and apparent monomer activity of cholesterol in taurine-conjugated cholate and chenodeoxycholate micellar solutions were studied to clarify the different modulating effect of these bile salt species on cholesterol uptake in an intestinal lumen. The maximal micellar solubility was significantly greater in taurochenodeoxycholate. The intermicellar cholesterol monomer concentration was not significantly different between the two kinds of micellar solution. However, the apparent cholesterol monomer activity determined using an artificial organic phase (polyethylene disc) was significantly higher in taurocholate than that in taurochenodeoxycholate. A linear relationship between the intermicellar cholesterol concentration and the apparent cholesterol monomer activity was found, with the slope depending upon the bile salt species. It is concluded that the difference in partitioning of cholesterol from taurocholate and taurochenodeoxycholate micelles into a fixed organic phase may contribute in part to the different regulating effects of these bile salts on the uptake of cholesterol in the intraluminal phase.  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