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 structure and conformation of HC-toxin   总被引:2,自引:0,他引:2  
Difference nuclear magnetic resonance studies and amino acid oxidase experiments establish the sequence and configuration of amino acids in HC-toxin as cyclo(L-Aoe-D-Pro-L-Ala-D-Ala). HC-toxin adopts the bis-γ-turn conformation in solution previously found for the cytostatic cyclic tetrapeptide chlamydocin.  相似文献   

2.
When injected into rats, a certain amount of mannitol is taken up by the liver and is associated with sedimentable structures. Isopycnic centrifugation in a sucrose gradient shows that a large part of mannitol is present in mitochondria, what remains is located in the lysosomes. The hypotonic release of mannitol present in organelles shows that the polyol is shared between mitochondria and lysosomes. The trapping of mannitol in lysosomes could result from the heterophagic or autophagic function of the lysosomes; the mechanism of its accumulation in mitochondria is still unexplained.  相似文献   

3.
Type I procollagen was thermally denatured and partially refolded by cooling to 20°C. The partially refolded protein was then used as a model system for testing proteolytic enzymes as probes for quantitative assay of fully aligned triple-helical molecules. Pepsin and chymotrypsin both digested fully denatured procollagen. However, digestion times of greater than 60 min were required, even with a large molar excess of the proteinases. These enzymes therefore are only useful for examining the folding of procollagen under conditions in which the process occurs at a slow rate. In contrast, trypsin cleaved the collagen domain of denatured procollagen within 2 min. Trypsin did not efficiently remove the precursor specific peptides, and therefore a mixture of chymotrypsin and trypsin was employed as an appropriate proteolytic probe for triple-helical conformation.  相似文献   

4.
Polyethylene glycol solutions (10% w/v) were used to detect the effect of mono- and divalent cations on some properties of thyroglobulin. It is shown that in presence of 10% w/v polyethylene glycol in 0.01 M Tris-HCl, pH 7.5, calcium (less than 0.05 M) modifies the solubility, the sedimentation rate, and the Stokes' radius of thyroglobulin, while monovalent cations up to 0.6 M do not effect any of these properties. These findings can be explained by an increase in molecular compactness of thyroglobulin. Furthermore, it was shown that a synthetic polymer, polyethylene glycol, could be used to detect conformational changes.  相似文献   

5.
Ionization of the phenolic group of N-acetyltyrosynamide has been studied using second-derivative spectroscopy. At pH 12.5 the second-derivative spectrum of the model compound revealed the presence of derivative bands in a spectral region (between 250 and 270 nm) where interference coming from other ultraviolet-absorbing chromophores is negligible. One of these peaks (260-nm peak) has been employed for the determination of tyrosyl groups in mixtures containing the aromatic amino acids.  相似文献   

6.
The performance of a branch and bound algorithm for molecular energy minimization is evaluated on a variety of test problems. Although not at present efficient enough for use in most practical situations, we show that it has distinct advantages over more conventional methods of global minimization. In addition, this study illustrates the technique on which the present algorithm is based, and the problems which must be overcome in developing an efficient algorithm based on similar principles.  相似文献   

7.
The interaction of chromium(III) with poly(dG-dC) inhibits the B to Z transition and results in the condensation of the polymer at high Cr/nucleotide ratios. At low Cr/nucleotide ratios chromium(III) enhanced the ability of ethanol to induce the B to Z transition of poly(dG-dC). The effects of chromium(III) on the conformation of DNA may be related to the carcinogenicity of chromium compounds.  相似文献   

8.
Paper 1 in this series1 outlined a method for identifying sets of backbone conformations of a polymer which are of a given helical type. This paper is concerned with fast methods for the calculation and mapping of contact-free structures in cases where the structure consists of two rigid molecular groups and where the position of each group varies by rotation about its own distinct axis. In particular, it is shown how such methods can be used to locate the neighbourhoods of crystal packing minima for a helical polymer chain, which could be generated by the method of Paper 1 and which is to be packed rigidly into a lattice. The method does not possess the flexibility of a full refinement procedure, but its advantages, apart from its conjectured speed, include an exactly known level of exhaustion of the examined variables and immunity to false packing minima and choice of initial starting position. Together, the methods may provide a means for checking the capability of refinement procedures to locate packing minima.  相似文献   

9.
The three-dimensional structure of the potent thyromimetic agent 3′-isopropyl- 3,5-diiodo-L-thyronine (iPr-T2) has been established by x-ray diffraction of single crystals of iPr-T2 hydrochloride. The molecular conformation is such that the β-ring 3′-isopropyl group is oriented in space proximal to the 3,5-diiodotyrosine α-ring, similar to the conformation adopted in the crystal structure of 3,5,3′-triiodo-L-thyronine.  相似文献   

10.
A new approach to the long-standing local minimum problem of molecular energy minimization is proposed. The approach relies upon a field of computer mathematics known as combinatorial optimization, together with methods of conformational analysis derived from distance geometry. The advantages over the usual numerical techniques of optimization are, first, that the algorithms derived are globally convergent, and second, that the mathematical problems involved are well-posed and suitable for study within the modern theory of computational complexity. In this paper we introduce the approach, and describe a computer program based on it.  相似文献   

11.
Various nucleoside di- and triphosphates have been compared with respect to their ability to protect rat brain hexokinase (ATP: d-hexose 6-phosphotransferase, EC 2.7.1.1) activity against inactivation by chymotrypsin, glutaraldehyde, heat, and 5,5′-dithiobis(2-nitrobenzoic) acid. ATP could be distinguished from other nucleoside triphosphates in these comparisons, which may be related to the specificity with which ATP is utilized as a substrate. All nucleoside derivatives examined provided substantial protection against two or more of the above inactivating agents, indicating relatively nonspecific binding of nucleotides by brain hexokinase, consistent with a similar lack of specificity in the inhibition of this enzyme by nucleoside derivatives. The fluorescence of 2-p-toluidinylnaphthalene-6-sulfonate (TNS) and of tetraiodofluorescein (TIF) was enhanced by binding to brain hexokinase. TNS binding was not affected by the presence of various relevant metabolites (Glc, glucose 6-phosphate, ATP), nor did TNS inhibit the enzyme. In contrast, substantial (approximately 70%) decreases in the fluorescence of bound TIF resulted from the addition of various nucleoside derivatives, and TIF served as a competitive inhibitor of brain hexokinase. These observations are consistent with the view that TIF binds to a nucleotide binding site of the enzyme. The inability of nucleotides to totally displace TIF was taken to indicate the existence of an additional TIF binding site (or sites) discrete from the catalytic site, and probably identical to the site(s) at which TNS binds with no effect on catalytic activity. The effects of saturating levels of ATP and ADP were not additive indicating that both compounds were displacing TIF from the same site i.e., a common nucleotide binding site. Glc, mannose, and 2-deoxyglucose greatly enhanced the ability of nucleotides to displace TIF, while fructose, galactose, and N-acetylglucosamine did not, indicating the existence of interactions between hexose and nucleotide binding sites; the hexoses themselves were not effective at displacing TIF. The enhanced binding of nucleotides in the presence of the first three hexoses but not the latter three can be directly correlated with the relative ability of these hexoses to induce specific conformational changes in the enzyme. The hexoses themselves were not effective at displacing TIF. Glucose 6-phosphate and 1,5-anhydroglucitol 6-phosphate could also displace TIF, and as with the nucleotides, a maximum of approximately 70% decrease in fluorescence was observed and the effectiveness of glucose 6-phosphate was enhanced in the presence of Glc. Other hexose 6-phosphates tested were not effective at displacing TIF. The specificity with which hexose 6-phosphates displaced TIF could be correlated with their ability to induce specific conformational change in the enzyme. The results are discussed as they relate to the kinetic mechanism and allosteric regulation by nucleotides that have been proposed for this enzyme.  相似文献   

12.
β-Alanyl-, acetimidoyl-, and carbamoyl-bradykinin [(IV), (V), and (VI), respectively] have been synthesized by the picolyl ester method. The first two analogs have very low smooth muscle contracting activity, but the carbamoyl derivative is as fully active as bradykinin itself.  相似文献   

13.
Castanospermine (1,6,7,8-tetrahydroxyoctahydroindolizine) is an indolizidine alkaloid that was isolated from the Australian plant, Castanospermum australe. This alkaloid was found to be a potent inhibitor of lysosomal alpha- and beta-glucosidases. In this report, the mechanism of inhibition of amyloglucosidase (an exo-1,4-alpha-glucosidase) and almond emulsin beta-glucosidase was examined. Castanospermine proved to be a competitive inhibitor of amyloglucosidase at both pH 4.5 and 6.0 when assayed with the p-nitrophenyl-alpha-D-glucoside. It was also a competitive inhibitor of almond emulsin beta-glucosidase at pH 6.5, but in this case previous studies had shown that inhibition was of the mixed type at pH 4.5 to 5.0. Th pH of the incubation mixture had a marked effect on the inhibition. Thus, in all cases, castanospermine was a much better inhibitor at pH 6.0 to 6.5 than it was at lower pH values. The pK for castanospermine was found to be 6.09, indicating that the alkaloid was probably more active in the unprotonated form. This was also suggested by the fact that the N-oxide of castanospermine, while still a competitive inhibitor, was 50 to 100 times less active than was castanospermine, and its activity was not markedly altered by pH. These results probably explain why castanospermine is a good inhibitor of the glycoprotein processing enzyme, glucosidase I, since this is a neutral enzyme.  相似文献   

14.
Synthetic [125I]-Tyr23, Phe2, Nle4-adrenocorticotropin (ACTH)-(1–38) ([125I]-ACTH analog) with full biological potency and near theoretical specific radioactivity (1800 ± 75 Ci/mmol) was used to investigate ACTH receptors on isolated rat adipocytes derived from 42-day-old rats. Binding to adipocytes was studied in the presence of 1% bovine serum albumin (BSA) as well as 4% BSA. The interaction of the [125I]-ACTH analog with adipocytes was highly specific, rapid, saturable, and reversible. Scatchard analysis of the binding data obtained in medium containing 1% BSA revealed a single class of binding sites with an apparent KD = 170 ± 11.9 pM. Competition experiments with unlabeled ACTH also yielded a comparable value for the apparent KD (143 ± 16.5 pm). The number of receptors per adipocyte was quite low (521–841/cell). The stimulation of lipolysis by ACTH was closely correlated with the binding, the apparent Km being 145–177 pm. At a concentration of 4% BSA in the incubation medium, the binding curve was shifted significantly to the right (apparent KD = 446 ± 77 pM) and the binding capacity was also significantly enhanced (1663 ± 208/cell) without any change in the apparent Km for glycerol release (187 ± 7.1 pm).  相似文献   

15.
F1ATPase from the Escherichia coli mutant of H+-ATPase, AN120 (uncA401), has less than 1% of the wild type activity and has been shown to be defective in the alpha subunit by in vitro reconstitution experiments. In the present study, the mutation site was located within a domain of the subunit by recombinant DNA technology. For this, a series of recombinant plasmids carrying various portions of the alpha subunit gene were constructed and used for genetic recombination with AN120. Analysis of the recombinants indicated that the mutation site could be located between amino acid residues 370 and 387. The biochemical properties of the mutant F1 were analyzed further using the fluorescent ATP analog DNS-ATP (2'-(5-dimethylaminonaphthalene-1-sulfonyl)-amino-2'-deoxy ATP). The single turnover process of E. coli F1ATPase proposed by Matsuoka et al. [(1982) J. Biochem. 92, 1383-1398.] was compared in the mutant and wild type F1's. Mutant F1 bound DNS-ATP and hydrolyzed it as efficiently as wild type F1. Results showed that binding of ATP to a low affinity site, possibly in the beta subunit, caused decrease of fluorescence of DNS-ATP in the wild type F1 and that this effect of ATP binding was inhibited by DCCD (dicyclohexyl carbodiimide). However, this effect was not inhibited by DCCD in the mutant F1, suggesting that in the proposed process some step(s) after ATP binding to the low affinity site differed in the mutant and wild F1's. When Pi was added to F1 bound to DNS-ATP or to aurovertin, a fluorescent probe capable of binding to the beta subunit, the opposite changes of fluorescence of these probes in the mutant and wild type F1's were observed, suggesting that the conformational change induced by phosphate binding was altered in the mutant F1. On the basis of the estimated mutation site and the biochemical properties of the mutant F1, the correlation of the domain of this site in the alpha subunit with the function of F1 ATPase is discussed.  相似文献   

16.
Phospholipase C-treated polymorphonuclear leucocytes were used to study the properties of NADPH oxidase activity of stimulated polymorphonuclear leucocytes.A comparison of the effects of phospholipase C treatment of whole leucocytes on the NADPH oxidase activity with other granule enzymes showed that the activities of β-glucuronidase and acid phosphatase were un-affected, whereas the NADPH oxidase activity was stimulated 4-fold and myeloperoxidase was inhibited about 30%.The distribution of NADPH oxidase activity among subcellular fractions of polymorphonuclear leucocyte homogenates was unaffected by phospholipase C whereas the other enzymes were released into the medium in soluble form; β-glucuronidase > acid phosphatase and myeloperoxidase.A number of solubilizing agents and procedures were tested for their ability to release NADPH oxidase activity from granules of phospholipase C-stimulated polymorphonuclear leucocytes. All procedures used caused appreciable release of granule protein but no release of NADPH oxidase activity. Most of the procedures used strongly inhibited the oxidase activity. These results indicate that the enzyme is tightly bound to granule structures and that the integrity of these structures is required for activity.Some of the solubilizing agents used (KCI, guanidium chloride) were very effective in solubilizing myeloperoxidase.The differential response of myeloperoxidase and NADPH oxidase to treatment with phospholipase C or solubilizing procedures suggests that the two activities are not due to the same enzyme. However, definite conclusion cannot be drawn because of the complex nature of myeloperoxidase.It was found necessary to lyse any erythrocytes present as contaminants of polymorphonuclear leucocytes preparations, since hemoglobin was converted to methemoglobin during the NADPH oxidase assay and methemoglobin exhibits appreciable NADPH oxidase activity.  相似文献   

17.
The membrane potential (average = ?52 mV) of a freely exposed electroplaque from a dissected prism of Torpedo marmorata electric organ is recorded with an intracellular glass microelectrode. The resting potential decreases with external potassium concentration. Acetylcholine (in the presence of O,O′-diethyl S-(β-diethylamino)ethyl phosphorothiolate), decamethonium, phenyltrimethyl-ammonium and carbamylcholine added to the bath cause a decrease of membrane potential, i.e. behave as agonists. Their effect is blocked in a competitive manner by d-tubocurarine, gallamine and hexamethonium, and in a non-competitive way by prilocaine; 1 μg Erabutoxin/ml completely abolishes the response to carbamylcholine. The apparent dissociation constants for seven cholinergic ligands are determined from the dose-response curves, and found to be closely related to those previously determined with Electrophorus electricus electroplaque with, however, a few differences. During these experiments it was noticed that potassium ions affect, in a differential manner, the response of T. marmorata electroplaque to carbamylcholine and decamethonium.  相似文献   

18.
Membrane preparations from Tetrahymena pyriformis catalyzed the acylations of glycerophosphate, isomeric monoacylglycerophosphate, and 1-acylglycerylphosphoryl-choline. Under the optimal conditions, glycerophosphate acyltransferase and 1-acylgly-cerophosphate acyltransferase used saturated and unsaturated acyl-CoA at comparable rates. The specificities of these acyltransferase systems for various acyl-CoAs as compared with the respective maximal velocities do not directly explain the fatty acid distribution in glycerophospholipids. However, the acylation of 2-acylglycerophosphate was highly selective for palmitate when the incubations were carried out in the presence of palmitoyl-CoA, oleoyl-CoA, 1-acylglycerophosphate, and 2-acylglycerophosphate. The 1-acylglycerylphosphorylcholine acyltransferase system showed relatively higher specificity for unsaturated acyl-CoA, which is consistent with the fatty acid pattern of phospholipids. Significant amounts of diglyceride and triglyceride were formed together with phosphatidic acid from acyl-CoA and glycerophosphate, indicating that the enzymes involved in triglyceride synthesis are closely associated with acyltransferase systems involved in phosphatidate synthesis in microsomes. These acyltransferase activities were found mainly in microsomes, and to a lesser extent, in pellicles, too. No significant difference was observed in the properties of acyltransferase systems in microsomes and pellicles.  相似文献   

19.
The molecular structure of N-tert-butyloxycarbonyl-D-leucyl-L-phenylalanyl ethanolamide (t-Boc-D-Leu-L-Phe-EA), a protected analogue of the C-terminal dipeptide of the membrane-active linear antibiotic gramicidin A, has been determined by X-ray diffraction. One of the two independent molecules in the asymmetric unit is characterized by a chain reversal stabilized by an intramolecular, three-centre, double hydrogen bonding. It represents the first experimental evidence for a beta-turn conformation fused with the oxy-analogue of an alpha-turn.  相似文献   

20.
Both uncomplexed subunits of the anthranilate synthetase-phosphoribosyltransferase enzyme complex from Salmonella typhimurium have an absolute requirement for divalent metal ions which can be satisfied by Mg2+, Mn2+, or Co2+. The metal ion kinetics for uncomplexed anthranilate synthetase give biphasic double-reciprocal plots and higher apparent Km values than those for anthranilate synthetase in the enzyme complex. In contrast, the apparent Km values for phosphoribosyltransferase are the same whether the enzyme is uncomplexed or complexed with anthranilate synthetase. This suggests that the metal ion sites on anthranilate synthetase, but not those on phosphoribosyltransferase, are altered upon formation of the enzyme complex. These results and the results of studies reported by others, suggest that complex formation between anthranilate synthetase and phosphoribosyltransferase leads to marked alterations at the active site of the former, but not the latter enzyme. Uncomplexed anthranilate synthetase can be stoichiometrically labeled with Co(III) under conditions which lead to inactivation of 75% of its activity. A comparison of the effects of anthranilate and tryptophan on phosphoribosyltransferase activity in the uncomplexed and complexed forms shows that anthranilate, but not tryptophan, inhibits the uncomplexed enzyme. The complexed phosphoribosyltransferase shows substrate inhibition by anthranilate binding to the phosphoribosyltransferase subunits. In contrast, in a tryptophan-hypersensitive variant complex, anthranilate inhibits phosphoribosyltransferase activity by acting on the anthranilate synthetase subunits. The data are interpreted to mean that there are two classes of binding sites for anthranilate, one on each type of subunit, which may participate in the regulation of anthranilate synthetase and phosphoribosyltransferase under different conditions.  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