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 375 毫秒
1.
The inducible acetyl-CoA:acetoacetate CoA-transferase of Escherichia coli catalyzes the transfer of CoA from acetyl-CoA to acetoacetate by a mechanism involving a covalent enzyme-CoA compound as a reaction intermediate. Acetyl-CoA + enzyme ? enzyme-CoA + Acetate Enzyme-CoA + acetoacetate ? acetoacetyl-CoA + enzyme These conclusions are based on the following data: 1) In the absence of acetoacetate, the maximal velocity of exchange of [14C]acetate into acetyl-CoA was comparable with maximal velocity of the complete reaction. 2) Incubation of the enzyme with NaBH4 after preincubation with an acyl-CoA substrate inactivated the enzyme by reduction of a glutamate residue in the β subunit of the CoA-transferase to α-amino-δ-hydroxyvaleric acid. Given the susceptibility of thioesters to borohydride reduction, the enzyme-CoA bond is a γ-glutamyl thiolester 3) Following incubation of the enzyme with a fluorescent derivative of acetyl-CoA, 1,N6-ethenoacetyl-CoA, etheno-CoA was bound to the CoA-transferase. Free etheno-CoA did not bind to the enzyme.  相似文献   

2.
The acetyl-CoA:acetoacetate CoA-transferase of Escherichia coli has the subunit structure α2β2 The enzyme contains six sulfhydryl groups, one per α chain and two per β chain, and no disulfides. The rates and extent of sulfhydryl group reactivity with 5,5′-dithiobis(2-nitrobenzoic acid) were compared in the free enzyme, the enzyme-CoA intermediate in the catalytic pathway, and a substrate analog-enzyme Michaelis complex. The analog used was acetylaminodesthio-CoA, a competitive inhibitor with respect to acetyl-CoA; the analog is not a substrate. The reactions were studied in the presence and absence of 10% glycerol. In the absence of glycerol, one sulfhydryl group reacted rapidly in the free enzyme and enzyme-CoA intermediate; relative to the free enzyme, the rate and number of subsequently reacting sulfhydryl groups were increased in the enzyme-CoA intermediate. In the presence of 10% glycerol, one sulfhydryl group reacted rapidly in the free enzyme, while two reacted rapidly in the enzyme-CoA compound; the rates and extents of subsequently reacting sulfhydryl groups were also enhanced in the enzyme-CoA compound. The data strongly suggested subunit interactions in the free enzyme and intermediate; glycerol abolished those interactions in the enzyme-CoA intermediate. In the absence of glycerol, sulfhydryl group reactivity in the Michaelis complex, enzyme-acetylaminodesthio-CoA, was similar to that in the free enzyme with one exception: One of the more slowly reacting sulfhydryl groups in the free enzyme reacted at a rate characteristic of the enzyme-CoA intermediate. The results obtained with N-ethylmaleimide were qualitatively similar. The fractional inactivation of the enzyme with N-ethylmaleimide as a function of sulfhydryl groups modified and the subunit location of those sulfhydryl groups indicated that the same sulfhydryl groups react in both enzyme species; however, those sulfhydryl groups reacted more rapidly in the enzyme-CoA compound. The data indicate both subunit interactions in the enzyme and characteristic conformational changes upon formation of an acyl-CoA-enzyme Michaelis complex and the enzyme-CoA intermediate.  相似文献   

3.
The acetyl-CoA:acetoacetate CoA-transferase of Escherichia coli was reversibly inactivated by pyridoxal 5′-phosphate. The residual activity of the enzyme was dependent on the concentration of the modifying reagent to a concentration of 5 mm. The maximum level of inactivation was 89%. Kinetic and equilibrium analyses of inactivation were consistent with a two-step process (Chen and Engel, 1975, Biochem. J.149, 619) in which the extent of inactivation was limited by the ratio of first-order rate constants for the reversible formation of an inactive Schiff base of pyridoxal 5′-phosphate and the enzyme from a noncovalent, dissociable complex of the enzyme and modifier. The calculated minimum residual activity was in close agreement with the experimentally determined value. The conclusion that the loss of catalytic activity resulted from modification of a lysine residue at the active site was based on the following data, (a) After incubation with 5 mm pyridoxal 5′-phosphate, 3.95 mol of the reagent was incorporated per mole of free enzyme with 89% loss of activity, while 2.75 mol of pyridoxal 5′-phosphate was incorporated into the enzyme-CoA intermediate with a loss of 10% of catalytic activity; the intermediate was formed in the presence of acetoacetyl-CoA; (b) acid hydrolysis of the modified, reduced enzyme-CoA intermediate yielded a single fluorescent compound that was identified as N6-pyridoxyllysine by chromatography in two solvent systems; (c) the enzyme was also protected from inactivation by saturating concentrations of free CoA and ADP but not by adenosine. The results suggested that a lysine residue is involved in the electrostatic binding of the pyrophosphate group of CoA. Carboxylic acid substrate did not protect the enzyme from inactivation.  相似文献   

4.
The rate of inactivation of succinyl-CoA:3-ketoacid coenzyme A transferase by thiol reagents is increased 3 to 100 times by very low concentrations of acyl-CoA substrates. The same maximum inactivation rate is found with acetoacetyl-CoA and succinyl-CoA. The enhanced rate of inactivation is caused by the stoichiometric formation of the enzyme-CoA intermediate and an accompanying conformation change of the enzyme. The inactivation rate provides a simple assay for the amount of enzyme present as the enzyme-CoA intermediate, using only catalytic concentrations of enzyme. This technique has been utilized to measure (a) a rate constant for hydrolysis of the enzyme-CoA intermediate of 0.10 min-1 at pH 8.1; (b) a stoichiometry of two active sites per enzyme molecule; and (c) the equilibrium constants for formation of the enzyme-CoA intermediate from dilute solutions of substrates (and hence for the overall reaction) by determining the ratio of [enzyme-CoA]/[enzyme] in the presence of a series of substrate "buffers" at different ratios of [RCOO-]/[RCOSCoA]. As the total concentration of acyl-CoA and carbosylate substrates is increased, the inactivation rate is decreased. This indicates that the Michaelis complexes are protected against inactivation.  相似文献   

5.
In this study, a propionate CoA-transferase (H16_A2718; EC 2.8.3.1) from Ralstonia eutropha H16 (Pct Re ) was characterized in detail. Glu342 was identified as catalytically active amino acid residue via site-directed mutagenesis. Activity of Pct Re was irreversibly lost after the treatment with NaBH4 in the presence of acetyl-CoA as it is shown for all CoA-transferases from class I, thereby confirming the formation of the covalent enzyme-CoA intermediate by Pct Re . In addition to already known CoA acceptors for Pct Re such as 3-hydroxypropionate, 3-hydroxybutyrate, acrylate, succinate, lactate, butyrate, crotonate and 4-hydroxybutyrate, it was found that glycolate, chloropropionate, acetoacetate, valerate, trans-2,3-pentenoate, isovalerate, hexanoate, octanoate and trans-2,3-octenoate formed also corresponding CoA-thioesters after incubation with acetyl-CoA and Pct Re . Isobutyrate was found to be preferentially used as CoA acceptor amongst other carboxylates tested in this study. In contrast, no products were detected with acetyl-CoA and formiate, bromopropionate, glycine, pyruvate, 2-hydroxybutyrate, malonate, fumarate, itaconate, β-alanine, γ-aminobutyrate, levulate, glutarate or adipate as potential CoA acceptor. Amongst CoA donors, butyryl-CoA, crotonyl-CoA, 3-hydroxybutyryl-CoA, isobutyryl-CoA, succinyl-CoA and valeryl-CoA apart from already known propionyl-CoA and acetyl-CoA could also donate CoA to acetate. The highest rate of the reaction was observed with 3-hydroxybutyryl-CoA (2.5 μmol mg?1 min?1). K m values for propionyl-CoA, acetyl-CoA, acetate and 3-hydroxybutyrate were 0.3, 0.6, 4.5 and 4.3 mM, respectively. The rather broad substrate range might be a good starting point for enzyme engineering approaches and for the application of Pct Re in biotechnological polyester production.  相似文献   

6.
Tammam SD  Rochet JC  Fraser ME 《Biochemistry》2007,46(38):10852-10863
Succinyl-CoA:3-ketoacid CoA transferase (SCOT) transfers CoA from succinyl-CoA to acetoacetate via a thioester intermediate with its active site glutamate residue, Glu 305. When CoA is linked to the enzyme, a cysteine residue can now be rapidly modified by 5,5'-dithiobis(2-nitrobenzoic acid), reflecting a conformational change of SCOT upon formation of the thioester. Since either Cys 28 or Cys 196 could be the target, each was mutated to Ser to distinguish between them. Like wild-type SCOT, the C196S mutant protein was modified rapidly in the presence of acyl-CoA substrates. In contrast, the C28S mutant protein was modified much more slowly under identical conditions, indicating that Cys 28 is the residue exposed on binding CoA. The specific activity of the C28S mutant protein was unexpectedly lower than that of wild-type SCOT. X-ray crystallography revealed that Ser adopts a different conformation than the native Cys. A chloride ion is bound to one of four active sites in the crystal structure of the C28S mutant protein, mimicking substrate, interacting with Lys 329, Asn 51, and Asn 52. On the basis of these results and the studies of the structurally similar CoA transferase from Escherichia coli, YdiF, bound to CoA, the conformational change in SCOT was deduced to be a domain rotation of 17 degrees coupled with movement of two loops: residues 321-329 that bury Cys 28 and interact with succinate or acetoacetate and residues 374-386 that interact with CoA. Modeling this conformational change has led to the proposal of a new mechanism for catalysis by SCOT.  相似文献   

7.
Synthesis of long-chain fatty alcohols in preputial glands of mice is catalyzed by an NADPH-dependent acyl coenzyme A (CoA) reductase located in microsomal membranes; sensitivity to trypsin digestion indicates that the reductase is on the cytoplasmic side of the membrane. Results with pyrazole and phenobarbital demonstrate the reaction is not catalyzed by a nonspecific alcohol dehydrogenase or an aldehyde reductase. Acyl-CoA reductase activity is sensitive to sulfhydryl and serine reagent modification, is stimulated by bovine serum albumin, and produces an aldehyde intermediate. The activity is extremely detergent sensitive and cannot be restored even after removal of the detergents. Phospholipase C or asolectin treatment does not release the acyl-CoA reductase from microsomal membranes, but causes a significant decrease in the activity recovered in the membrane pellet. Glycerol does not solubilize the reductase activity, nor does 3.0 m NaCl; however, the combination of glycerol and 3.0 m NaCl did release about 50% of the acyl-CoA reductase from the microsomal pellet. Substrate concentration curves obtained in the presence or absence of bovine serum albumin show significant differences in enzyme activities. The reductase is sensitive to the concentration of palmitoyl-CoA and is progressively inhibited at levels beyond the critical micellar concentration of the substrate. The apparent Km for acyl-CoA reductase is 14 μm; however, the maximum velocity varies with the concentration of albumin used. Expression of enzyme activity in delipidated microsomes requires specific phospholipids, which suggests that in vivo regulation of acyl-CoA reductase activity could be achieved through modifications in membrane lipid composition.  相似文献   

8.
The proteolytic activities of α-chymotrypsin, trypsin, pepsin, bromelain, and an extract from germinating pumpkin seeds (Cucurbita moschata) were determined by their ability to effect the release of 1-anilino-8-naphthalenesulfonate bound to internal hydrophobic sites in intact protein substrates. Casein, glyceraldehyde-3-P dehydrogenase, urease, catalase, pumpkin seed globulin, and bovine serum albumin enhanced the fluorescence of 1-anilino-8-naphthalenesulfonate sufficiently to be used as proteolytic substrates. Chymotrypsin, trypsin, pepsin, and bromelain exhibited activity against all or almost all of the protein substrates. The activity of 1 μg of α-chymotrypsin or trypsin and 100 ng of pepsin could be easily detected by this method of assay within 4 to 5 minutes depending upon the substrate. The enzyme extracted from 3-day germinated pumpkin seeds exhibited strong activity only against pumpkin seed globulin, weak activity against the globulins of squash and cucumber and casein, and no activity against the other protein substrates. Activity against pumpkin globulin was maximal at pH 7.4. When assayed by an increase in ninhydrin-positive products, the enzyme extract from pumpkin seeds also showed strong activity against pumpkin globulin and weak activity against casein. The 1-anilino-8-naphthalenesulfonate-fluorescence method was at least 20 times more sensitive than the ninhydrin method and was 10 to 20 times more rapid.  相似文献   

9.
The binding of 8-anilino-1-naphthalenesulfonate (ANS) to ciliary dynein ATPase leads to a marked increase in the dye's fluorescence intensity, accompanied by a blue shift in the observed fluorescence emission maximum. We found that dynein has 37 +/- 3 ANS binding sites and that experimentally applied ANS concentrations failed to alter enzyme activity. The fluorescence properties of the enzyme-dye complex were used to learn more about the binding characteristics of dynein substrates and effectors and to probe for possible conformational changes of the enzyme. The fluorescence of the dynein-ANS complex is increased by a number of substrates, including ATP, GTP, and UTP. The transfer of excitation energy from dynein chromophores to adsorbed ANS was also investigated. Our findings indicate that dynein appears to undergo a localized conformational change in its interaction with ATP. Native dynein was also found to be conformationally different from heat-activated or NEM-modified enzyme as evidenced by the emission and excitation spectra of the various enzyme-ANS complexes.  相似文献   

10.
Chemical group-transfer reactions by hydrolytic enzymes have considerable importance in biocatalytic synthesis and are exploited broadly in commercial-scale chemical production. Mechanistically, these reactions have in common the involvement of a covalent enzyme intermediate which is formed upon enzyme reaction with the donor substrate and is subsequently intercepted by a suitable acceptor. Here, we studied the glycosylation of glycerol from sucrose by sucrose phosphorylase (SucP) to clarify a peculiar, yet generally important characteristic of this reaction: partitioning between glycosylation of glycerol and hydrolysis depends on the type and the concentration of the donor substrate used (here: sucrose, α-d -glucose 1-phosphate (G1P)). We develop a kinetic framework to analyze the effect and provide evidence that, when G1P is used as donor substrate, hydrolysis occurs not only from the β-glucosyl-enzyme intermediate (E-Glc), but additionally from a noncovalent complex of E-Glc and substrate which unlike E-Glc is unreactive to glycerol. Depending on the relative rates of hydrolysis of free and substrate-bound E-Glc, inhibition (Leuconostoc mesenteroides SucP) or apparent activation (Bifidobacterium adolescentis SucP) is observed at high donor substrate concentration. At a G1P concentration that excludes the substrate-bound E-Glc, the transfer/hydrolysis ratio changes to a value consistent with reaction exclusively through E-Glc, independent of the donor substrate used. Collectively, these results give explanation for a kinetic behavior of SucP not previously accounted for, provide essential basis for design and optimization of the synthetic reaction, and establish a theoretical framework for the analysis of kinetically analogous group-transfer reactions by hydrolytic enzymes.  相似文献   

11.
Fumarate and ferricyanide increased the rate of amino acid uptake in anoxic cells of E. coli suspended in a glycerol medium. The stimulation correlated with the hydrophobicity of 16 amino acids transported by several carrier systems. Fumarate and ferricyanide increased membrane energization as measured by changes in transmembrane pH and electrical potential, or by quenching of fluorescence of 1-anilino-8-naphthalenesulfonate. The results suggest that a common, rate determining step among the amino acid transport systems is the transfer of the substrate from an aqueous to an apolar environment.  相似文献   

12.
PhaA from Ralstonia eutropha (RePhaA) is the first enzyme in the polyhydroxyalbutyrate (PHB) biosynthetic pathway and catalyzes the condensation of two molecules of acetyl-CoA to acetoacetyl-CoA. To investigate the molecular mechanism underlying PHB biosynthesis, we determined the crystal structures of the RePhaA protein in apo- and CoA-bound forms. The RePhaA structure adopts the type II biosynthetic thiolase fold forming a tetramer by means of dimerization of two dimers. The crystal structure of RePhaA in complex with CoA revealed that the enzyme contained a unique Phe219 residue, resulting that the ADP moiety binds in somewhat different position compared with that bound in other thiolase enzymes. Our study provides structural insight into the substrate specificity of RePhaA. Results indicate the presence of a small pocket near the Cys88 covalent catalytic residue leading to the possibility of the enzyme to accommodate acetyl-CoA as a sole substrate instead of larger acyl-CoA molecules such as propionyl-CoA. Furthermore, the roles of key residues involved in substrate binding and enzyme catalysis were confirmed by site-directed mutagenesis.  相似文献   

13.
The structural and functional properties of arginine kinase (AK) in alkaline conditions in the absence or presence of salt have been investigated. The conformational changes of AK during alkaline unfolding and salt-induced folding at alkaline pH were monitored using intrinsic fluorescence emission, binding of the fluorescence probe 1-anilino-8-naphthalenesulfonate and circular dichroism. The results for the alkaline unfolded enzyme showed that much lower pH (11.0) was required to cause the complete loss of AK activity than was required to cause an obvious conformational change of the enzyme. Compared with the completely unfolded state in 5 M urea, the high pH denatured enzyme had some residual secondary and tertiary structure even at pH 13.0. Increasing the ionic strength by adding salt at pH 12.75 resulted in the formation of a relatively compact tertiary structure and a little new secondary structure with hydrophobic surface enhancement. These results indicate that the partially folded state formed under alkaline conditions may have similarities to the molten globule state which is compact, but it has a poorly defined tertiary structure and a native-like secondary structure.  相似文献   

14.
Coenzyme A (CoA) transferase from Peptostreptococcus elsdenii has been purified and crystallized, and some of its properties have been established. The work was facilitated by a newly developed coupled and continuous spectrophotometric assay in which the disappearance of added acrylate could be followed at 245 nm. The rate-limiting conversion of acetyl- and beta-hydroxypropionyl CoA to acrylyl CoA by CoA transferase was followed by the non-rate-limiting conversion to beta-hydroxypropionyl CoA by excess crotonase. Thus, a small priming quantity of acetyl CoA served to generate acrylyl CoA, which, by hydration, generated beta-hydroxypropionyl CoA. This product then served to generate more acrylyl CoA in cyclic fashion. The net result was the CoA transferase-limited conversion of acrylate to beta-hydroxypropionate. The purified transferase has a molecular weight of 125,000 and is composed of two subunits of 63,000 each, as determined by disc gel electrophoresis. Short-chain-length monocarboxylic acids are substrates, whereas dicarboxylic or beta-ketocarboxylic acids are not. The reaction kinetics are typical of a ping-pong bi bi mechanism composed of two half reactions linked by a covalent enzyme intermediate. Incubation of the transferase with acetyl CoA in the absence of a fatty acid acceptor yielded a stable intermediate which, by absorption spectrophotometry, radioactivity measurements, reduction with borohydride, reactivity with hydroxylamine, and catalytic activity, was identified as an enzyme-CoA compound. Kinetic constants for CoA transferase are: final specific activity, 110 U/mg of protein corresponding to 1.38 X 10(4) mumol of acrylate activated per mumol of transferase; Km for acrylate, 1.2 X 10(-3) M; Km for acetyl CoA (beta-hydroxypropionyl CoA), 2.4 X 10(-5) M.  相似文献   

15.
The effects of carnitine on the metabolism of palmitoylcarnitine were studied by using isolated rat liver mitochondria. Particular attention was given to carnitine acyltransferase-mediated interactions between carnitine and the mitochondrial CoA pool. Carnitine concentrations less than 1.25mm resulted in an increased production of acetylcarnitine during palmitoylcarnitine oxidation. Despite this shunting of C2 units to acetylcarnitine formation, no change was observed in the rate of oxygen consumption or major product formation (citrate or acetoacetate). Further, no changes were observed in the mitochondrial content of acetyl-CoA, total acid-soluble CoA or acid-insoluble acyl-CoA. These observations support the concept, based on studies in vivo, that the carnitine/acylcarnitine pool is metabolically sluggish and the acyl-group flux low as compared with the CoA/acyl-CoA pool. Acid-insoluble acyl-CoA content was decreased and CoA content increased at carnitine concentrations greater than 1.25mm. When [14C]carnitine was used in the incubations, it was demonstrated that this resulted from acid-insoluble acylcarnitine formation from intramitochondrial acid-insoluble acyl-CoA mediated by carnitine palmitoyltransferase B. Again, the higher carnitine concentrations resulted in no changes in the rates of oxygen consumption or major product formation. The above effects of carnitine were observed whether citrate or acetoacetate was the major product of oxidation. In contrast, an increase in acetyl-CoA concentration was observed at high carnitine concentrations only when acetoacetate was the product. Since the rate of acetoacetate production was not changed, these higher acetyl-CoA concentrations suggest that a new steady state had been established to maintain acetoacetate-production rates. Since there was no change in acetyl-CoA concentration when citrate was the major product, a change in the activity of the pathway utilizing acetyl-CoA for ketone-body synthesis and the potential regulation of this pathway must be considered.  相似文献   

16.
Experiments were performed to define the relation between covalent binding of enzymes to β2-macroglobulin (α2M), the specific proteolysis of α2M subunits to 85K fragments, and the reactivity of the methylamine site on α2M. We studied the reaction of α2M with native trypsin, anhydrotrypsin, and two active lysyl-blocked derivatives, methyl-trypsin and dimethylmaleyl-trypsin, the last with reversibly modified amino groups that can be regenerated at low pH. The results were: (1) All enzymes tested reacted with α2M but only native trypsin formed covalent complexes (not dissociable by sodium dodecyl sulfate). Trypsin and the lysyl-blocked enzymes caused complete proteolysis of the α2M subunits, in agreement with previous studies. (2) The dimethyl-maleyl-trypsin became covalently bound to α2M only after removing the blocking groups of the bound enzyme, indicating that sequential proteolysis and covalent bond formation is possible. Under the conditions used for deblocking, there was no change in the covalent/noncovalent binding ratio of native trypsin, anhydrotrypsin, or the other lysyl-blocked derivative, methyl-trypsin. (3) Native trypsin or anhydrotrypsin displaced methyl- or dimethylmaleyl-trypsin from their α2M complexes but the newly bound enzymes with free amino groups did not form covalent bonds indicating that enzymes must remain in association with the inhibitor for the bond to form. (4) Methylamine reacts with noncovalent α2M complexes but not with covalent complexes. (5) Methylamine-treated α2M can still form complexes with trypsin but at a drastically reduced rate and only noncovalent complexes are formed. In summary, sequential proteolysis and covalent bond formation is possible under certain conditions, and there is a strong correlation between covalent binding and loss of methylamine reactivity. The latter observation is suggestive evidence for the identity of the covalent binding site of α2M and the putative thiol ester of the methylamine site. The enzyme lysyl amino groups, are likewise possible candidates for attacking nucleophile at that site.  相似文献   

17.
Human pregnancy zone protein (PZP) is a major pregnancy-associated plasma protein, strongly related to α2-macroglobulin (α2M). The proteinase binding reaction of PZP is investigated using chymotrypsin as a model enzyme. The time-course of the interaction is studied by measuring the change in intrinsic protein fluorescence of PZP-chymotrypsin reaction mixtures as a function of time after rapid mixing in a stopped-flow apparatus. Titrations show the changes of fluorescence at equilibrium to correspond with the formation of a chymotrypsin-PZP(tetramer) species. The kinetic results show the formation of the species to take place in an overall second-order process dependent on the concentrations of chymotrypsin and of PZP(dimers), k = 5 · 105M−1 ·s−1. Reactions of PZP-thiol groups do not give rise to fluorescence changes. The fluorescence changes most likely reflect the formation of an intermediate with intact thiol esters. Further analysis of the kinetic results suggests that the chymotrypsin-PZP(tetramer) intermediate is formed in two reaction steps: (1) initially native PZP(dimers) are cleaved at bait regions by enzyme molecules, and that is the rate determining reaction of the fluorescence changes; (2) association with another PZP(dimer) or PZP(dimer)-chymotrypsin complex in a very fast reaction that leads to the formation of 1:1-chymotrypsin-PZP(tetramer) intermediate, probably with intact thiol esters. The interactions studied apparently are established early in the path of the reaction and the fluorescence changes probably reflect noncovalent enzyme-PZP contacts, which are not changed when covalent binding occurs. Further, fluorescence changes are seen only in reactions of PZP with enzymes, not with methylamine.  相似文献   

18.
Phosphofructokinase was purified 1200-fold from extracts of Escherichia coli B. Kinetic studies of the enzyme were carried out in the presence of the fluorescent dye 8-anilino-1-naphthalenesulfonate (1,8-ANS). 1,8-ANS was competitive with ATP and an uncompetitive inhibitor with respect to fructose-6-P. These parabolic inhibitions were accounted for by assuming that at least two molecules of the inhibitor were responsible for decreasing the affinity of the enzyme for ATP. ADP and GDP are both positive effectors for E. coli Phosphofructokinase. Evidence is presented to indicate that 1,8-ANS binding decreases the affinity of a regulatory site for ADP but not the binding site for regulation by GDP.  相似文献   

19.
We have investigated the aggregation and amyloid fibril formation of bovine β-lactoglobulin variant A, with a focus on the early stages of aggregation. We used noncovalent labeling with thioflavin T and 1-anilino-8-naphthalenesulfonate to follow the conformational changes occurring in β-lactoglobulin during aggregation using time resolved luminescence. 1-Anilino-8-naphthalenesulfonate monitored the involvement of the hydrophobic core/calyx of β-lactoglobulin in the aggregation process. Thioflavin T luminescence monitored the formation of amyloid. The luminescence lifetime distributions of both probes showed changes that could be attributed to conformational changes occurring during and following aggregation. To correlate the luminescence measurements with the degree of aggregation and the morphology of the aggregates, we also measured dynamic light scattering and atomic force microscopy images. We evaluated the relative stability of the intermediates with an assay that is sensitive to aggregation reversibility. Our results suggest that initial aggregation during the first 5 days occurred with partial disruption of the characteristic calyx in β-lactoglobulin. As the globular aggregates grew from days 5 to 16, the calyx was completely disrupted and the globular aggregates became more stable. After this second phase of aggregation, conversion into a fibrillar form occurred, marking the growth phase, and still more changes in the luminescence signals were observed. Based on these observations, we propose a three-step process by which monomer is converted first into weakly associated aggregates, which rearrange into stable aggregates, which eventually convert into protofibrils that elongate in the growth phase.  相似文献   

20.
Using steady-state, polarized, and phase-modulation fluorometry, the dithiothreitol-induced denaturation of insulin and formation of its complex with alpha-crystallin in solution were studied. Prevention of the aggregation of insulin by alpha-crystallin is due to formation of chaperone complexes, i.e. interaction of chains of the denatured insulin with alpha-crystallin. The conformational changes in alpha-crystallin that occur during complex formation are rather small. It is unlikely that N-termini are directly involved in the complex formation. The 8-anilino-1-naphthalenesulfonate (ANS) is not sensitive to the complex formation. ANS emits mainly from alpha-crystallin monomers, dimers, and tetramers, but not from oligomers or aggregates. The possibility of highly sensitive detection of aggregates by light scattering using a spectrofluorometer with crossed monochromators is demonstrated.  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