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 kinetic activation parameters (activation free energy, activation free enthalpy, and activation free entropy change) of the conformational change of alpha-chymotrypsin from an inactive to the active conformation were determined after a pH jump from pH 11.0 to pH 6.8 by the fluorescence stopped-flow method. The conformational change was followed by measuring changes in the protein fluorescence. For the bovine wild-type protein, the same kinetic parameters are obtained as in the study of proflavin binding. Several mutants were made with the goal to accelerate or decelerate this conformational transition. The inspiration for the choice of the mutants came from a previous modelling study done on the bovine wild-type chymotrypsin. The results of the fluorescence stopped flow experiments show that several mutants behaved as was expected based on the information provided by the modeling study on the wild-type variant. For some mutants our assumptions were not correct, and therefore additional modeling studies of the activation pathways of these mutant proteins are necessary to be able to explain the observed kinetic behavior.  相似文献   

2.
The peptide hormone bradykinin (BK) (Arg(1)-Pro(2)-Pro(3)-Gly(4)-Phe(5)-Ser(6)-Pro(7)-Phe(8)-Arg(9)) and its shorter homolog BK(1-5) (Arg(1)-Pro(2)-Pro(3)-Gly(4)-Phe(5)) were labeled with the extrinsic fluorescent probe ortho-aminobenzoic acid (Abz) bound to the N-terminal and amidated in the C-terminal carboxyl group (Abz-BK-NH(2) and Abz-BK(1-5)-NH(2)). The fragment des-Arg(9)-BK was synthesized with the Abz fluorescent probe attached to the 3-amino group of 2,3-amino propionic acid (DAP), which positioned the Abz group at the C-terminal side of BK sequence, constituting the peptide des-Arg(9)-BK-DAP(Abz)-NH(2). The spectral characteristics of the probe were similar in the three peptides, and their fluorescent properties were monitored to study the interaction of the peptides with anionic vesicles of dimyristoylphosphatidylglycerol (DMPG). Time-resolved fluorescence experiments showed that the fluorescence decay of the peptides was best described by double-exponential kinetics, with mean lifetimes values around 8.0 ns in buffer pH 7.4 that increased about 10% in the presence of DMPG vesicles. About a 10-fold increase, compared with the values in aqueous solution, was observed in the steady-state anisotropy in the presence of vesicles. A similar increase was also observed for the rotational correlation times obtained from time-resolved anisotropy decay profiles, and related to the overall tumbling of the peptides. Equilibrium binding constants for the peptide-lipid interaction were examined monitoring anisotropy values in titration experiments and the electrostatic effects were evaluated through Gouy-Chapman potential calculations. Without corrections for electrostatic effects, the labeled fragment Abz-BK(1-5)-NH(2) presented the major affinity for DMPG vesicles. Corrections for the changes in peptide concentration due to electrostatic interactions suggested higher affinity of the BK fragments to the hydrophobic phase of the bilayer.  相似文献   

3.
The interaction of the dyes oxonol V and oxonol VI with unilamellar dioleoylphosphatidylcholine vesicles was investigated using a fluorescence stopped-flow technique. On mixing with the vesicles, both dyes exhibit an increase in their fluorescence, which occurs in two phases. According to the dependence of the reciprocal relaxation time on vesicle concentration, the rapid phase appears to be due to a second-order binding of the dye to the lipid membrane, which is very close to being diffusion-controlled. The slow phase is almost independent of vesicle concentration, and it is suggested that this may be due to a change in dye conformation or position within the membrane, possibly diffusion across the membrane to the internal monolayer. The response times of the dyes to a rapid jump in the membrane potential has also been investigated. Oxonol VI was found to respond to the potential change in less than 1 s, whereas oxonol required several minutes. This has been attributed to lower mobility of oxonol V within the lipid membrane.  相似文献   

4.
The kinetics of the binding of MDL 27048 to tubulin have been studied by fluorescence stopped flow. The binding is accompanied by a fluorescence increase. The time course can be described by a sum of two exponentials, assumed to be due to the presence of two major tubulin isoforms. The observed rate constants depend in a nonlinear way on the concentration of MDL in pseudo-first-order conditions. This concentration dependence can be described by the presence of a fast equilibrium of low affinity, followed by an isomerization of the initial complex. The dissociation kinetics have been studied by displacement experiments, in which MTC was used as a competitive ligand. The reaction enthalpy change for the first binding equilibrium and the activation energies for the forward and reverse steps of the isomerization were determined from the temperature dependence. This was possible for the two tubulin isotype populations. The kinetics of the binding of MDL to tubulin are slowed down in the presence of 3',4',5'-trimethoxyacetophenone, a fast binding analog of the colchicine A-ring, but are not influenced by the binding of tropolone methyl ether, indicating that the binding site of MDL has the A-subsite in common with colchicine, but not the C-subsite.  相似文献   

5.
To better understand substrate recognition and catalysis by RNase III, we examined steady-state and pre-steady-state reaction kinetics, and changes in intrinsic enzyme fluorescence. The multiple turnover cleavage of a model RNA substrate shows a pre-steady-state burst of product formation followed by a slower phase, indicating that the steady-state reaction rate is not limited by substrate cleavage. RNase III catalyzed hydrolysis is slower at low pH, permitting the use of pre-steady-state kinetics to measure the dissociation constant for formation of the enzyme-substrate complex (K(d)=5.4(+/-0.6) nM), and the rate constant for phosphodiester bond cleavage (k(c)=1.160(+/-0.001) min(-1), pH 5.4). Isotope incorporation analysis shows that a single solvent oxygen atom is incorporated into the 5' phosphate of the RNA product, which demonstrates that the cleavage step is irreversible. Analysis of the pH dependence of the single turnover rate constant, k(c), fits best to a model for two or more titratable groups with pK(a) of ca 5.6, suggesting a role for conserved acidic residues in catalysis. Additionally, we find that k(c) is dependent on the pK(a) value of the hydrated divalent metal ion included in the reaction, providing evidence for participation of a metal ion hydroxide in catalysis, potentially in developing the nucleophile for the hydrolysis reaction. In order to assess whether conformational changes also contribute to the enzyme mechanism, we monitored intrinsic tryptophan fluorescence. During a single round of binding and cleavage by the enzyme we detect a biphasic change in fluorescence. The rate of the initial increase in fluorescence was dependent on substrate concentration yielding a second-order rate constant of 1.0(+/-0.1)x10(8) M(-1) s(-1), while the rate constant of the second phase was concentration independent (6.4(+/-0.8) s(-1); pH 7.3). These data, together with the unique dependence of each phase on divalent metal ion identity and pH, support the hypothesis that the two fluorescence transitions, which we attribute to conformational changes, correlate with substrate binding and catalysis.  相似文献   

6.
The ionization of fatty acids, fatty amines and N-acylamino acids incorporated in phosphatidylcholine single-walled vesicles has been measured. The guest molecules have been specifically enriched with 13C and titrated by using NMR spectroscopy. The apparent pKa of fatty acids in phosphatidylcholine bilayers is 7.2–7.4 and those of fatty amines are approx. 9.5. These pKa values depend on many different parameters related to the structure of the lipid/ solution interface, to the composition of the aqueous medium and to the localization of the ionizable groups. A special sensitivity to the ionic strength and to the surface charge has been found. A positive surface charge decreases the pKa value whereas a negative one increases it, the total range of variation being 2.5–3 units. In a qualitative macroscopic interpretation, it is proposed that pKa is essentially determined by the low polarity of the lipidic matrix.  相似文献   

7.
T. N. Kropacheva  J. Raap   《FEBS letters》1999,460(3):1219-504
The effect of a transmembrane potential on ion channel formation by zervamicin II (ZER-II) was studied in a vesicular model system. The dissipation of diffusion potential caused by addition of ZER-II to small phosphatidylcholine vesicles was monitored using fluorescent (Safranine T) and optical (Oxonol YI) probes. Cis-positive potentials facilitated channel formation, while at cis-negative potentials, ion fluxes were inhibited. A potential-independent behavior of ZER-II was observed at high peptide concentrations, most likely due to its membrane modifying property.  相似文献   

8.
A series of anthroyloxy fatty acid (AF) fluorescent probes, with the anthroyloxy group covalently linked at various positions along the alkyl chain, were studied in solvents exhibiting a wide range of polarity and hydrogen-bond donor (Hd) and acceptor (Ha) ability. These probes were sensitive to the solvent polarity as reflected by the Stokes' shift observed in steady state fluorescence. As determined by multi-linear regression analysis of the observed Stokes' shift and solvent parameters, such as orientation polarizability (Δf), Hd and Ha of the solvents, all the probes were sensitive to the Hd of solvents but were not affected by the Ha of solvents except the 2-AF. Due to the proximity of the polar headgroup to the fluorophore, it appears that some intramolecular hydrogen-bonding is present in 2-AF, an interaction that is sensitive to the pH of the solvent, but is less sensitive to the Hd and Ha of the solvents. Fluorescence lifetimes measured by the multi-frequency phase-modulation technique in mixtures of hexane and ethanol reflect a modified Stern-Volmer behavior suggesting the second solvent, ethanol, specifically interacts with the probe, in part through collisional quenching. Also, the lifetime data were sensitive to very low concentrations of the second solvent (0–0.1%, by vol.). The results from this study provide insight into the intrinsic differences between the different AF positions that must be taken into consideration while investigating the dynamics of lipid bilayer systems. Moreover, this study illustrates the utility and resolving power of lifetime based measurements needed for the interpretation of heterogeneous biophysical environments.  相似文献   

9.
The protein folding kinetics of hen egg white lysozyme (HEWL) was studied using experimental and bioinformatics tools. The structure of the transition state in the unfolding pathway of lysozyme was determined with stopped-flow kinetics using intact HEWL and its chemically modified derivative, in which six lysine residues have been modified. The overall consistency of φ-value (φ ≈ 1) indicates that lysine side chains interactions are subject to breaking in the structure of the transition state. Following experimental evidences, multiple sequence alignment of lysozyme family in vertebrates and exact structural examination of lysozyme, showed that the α-helix in the structure of lysozyme has critical role in the unfolding kinetics.  相似文献   

10.
Ye H  Qiu B  Lin Z  Chen G 《Luminescence》2011,26(5):336-341
The interaction between tamibarotene and bovine serum albumin (BSA) was studied using fluorescence quenching technique and ultraviolet–visible spectrophotometry. The results of experiments showed that tamibarotene could strongly quench the intrinsic fluorescence of BSA by a dynamic quenching mechanism. The apparent binding constant, number of binding site and corresponding thermodynamic parameters at different temperatures were calculated respectively, and the main interaction force between tamibarotene and BSA was proved to be hydrophobic force. Synchronous fluorescence spectra showed that tamibarotene changed the molecular conformation of BSA. When BSA concentration was 1.00 × 10?6 mol L?1, the quenched fluorescence ΔF had a good linear relationship with the concentration of tamibarotene in the range 1.00 × 10?6 to 12.00 × 10?6 mol L?1 with the detection limit of 6.52 × 10?7 mol L?1. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

11.
Phase modulation fluorescence spectroscopy was used to investigate the influence of cholesterol (0 to 50 mol%) on acyl chain dynamics in multilamellar vesicles of phosphatidylcholine. Four different phosphatidylcholines (DPPC, DOPC, POPC, and egg PC) and six different fluorescent probes (diphenylhexatriene and five anthroyloxy fatty acids) were employed. We found that: (1) Increased cholesterol content had only slight effects on fluorescence lifetimes of the six probes. (2) Increased cholesterol content increased the steady-state fluorescence anisotropy (r) of all the probes except 16-anthroyloxy palmitate (16-AP) in each of the four phosphatidylcholines. (3) Added cholesterol tended to limit the extent of probe rotation (as reflected by r, the infinite-time anisotropy) to a much greater extent than it altered the rate of probe rotation. (4) The tendency for cholesterol to order the structure of the bilayer was greatest in the proximal half of the acyl chains and diminished toward the center of the bilayer. (5) In some phosphatidylcholines the rotation rates of probes located near the bilayer center (diphenylhexatriene and 16-AP) were apparently increased by increasing levels of cholesterol. (6) In several respects dipalmitoylphosphatidylcholine (DPPC) vesicles responded differently to increased cholesterol than vesicles of the other three phosphatidylcholines. (7) A single second-order equation described the relationship between rand r for the five anthroyloxy fatty acid probes in the four different phosphatidylcholines over a wide range of cholesterol content. The data for diphenylhexatriene in the different phosphatidylcholines could not be fit by a single equation.  相似文献   

12.
The hydrophobic interaction between spin-labelled stearic acid and spectrin was studied by electron paramagnetic resonance (EPR) and fluorescence quenching. The results are quantitatively interpreted in terms of two types of binding site on spectrin. A comparison between the results of the EPR and fluorescence experiments show the drawback of the fluorescence method in binding studies.  相似文献   

13.
Summary When lecithin-cholesterol vesicles, containing the membrane-bound spin probe 3-doxyl-cholestane, were set in contact with mouse lymphocytes, the vesicles adsorbed to the cell and vesicle-membrane components were transferred to it. The spin probe was enzymatically reduced at the inside of the cell membrane. The spin-label method provided a means to determine quantitatively the extent of vesicles adsorption and vesicle-cell fusion by measuring the transfer of vesicles membrane material to the cell. This method, together with the reduction of spin label by the cell, allowed also a quantitative estimate of the extent of endocytosis during cell-liposome interaction.  相似文献   

14.
15.
We tested the ability of saturated n-monocarboxylic acids ranging from eight to 12 carbons in length to self-assemble into vesicles, and determined the minimal concentrations and chain lengths necessary to form stable bilayer membranes. Under defined conditions of pH and concentrations exceeding 150 mM, an unbranched monocarboxylic acid as short as eight carbons in length (n-octanoic acid) assembled into vesicular structures. Nonanoic acid (85 mM) formed stable vesicles at pH 7.0, the pK of the acid in bilayers, and was chosen for further testing. At pH 6 and below, the vesicles were unstable and the acid was present as droplets. At pH ranges of 8 and above clear solutions of micelles formed. However, addition of small amounts of an alcohol (nonanol) markedly stabilized the bilayers, and vesicles were present at significantly lower concentrations (∼20 mM) at pH ranges up to 11. The formation of vesicles near the pKa of the acids can be explained by the formation of stable RCOO…HOOCR hydrogen bond networks in the presence of both ionized and neutral acid functions. Similarly, the effects of alcohols at high pH suggests the formation of stable RCOO…HOR hydrogen bond networks when neutral RCOOH groups are absent. The vesicles provided a selective permeability barrier, as indicated by osmotic activity and ionic dye capture, and could encapsulate macromolecules such as DNA and a protein. When catalase was encapsulated in vesicles of decanoic acid and decanol, the enzyme was protected from degradation by protease, and could act as a catalyst for its substrate, hydrogen peroxide, which readily diffused across the membrane. We conclude that membranous vesicles produced by mixed short chain monocarboxylic acids and alcohols are useful models for testing the limits of stabilizing hydrophobic effects in membranes and for prebiotic membrane formation.  相似文献   

16.
The interaction of the specific sugar residue in ginsenosides with egg phosphatidylcholine vesicles was investigated by ESR spectrometry using phosphatidic acid spin-labeled at the polar head groups. Ginsenoside-Rc, which has an α-l-arabinofuranose residue and agglutinability toward egg yolk phosphatidylcholine vesicles (Fukuda, K. et al. (1985) Biochim. Biophys. Acta 820, 199–206), caused the restriction of the segmental motion of spin-labeled phosphatidic acid in egg phosphatidylcholine vesicles, indicating that the saponin interacted with the polar head groups of vesicles. Other ginsenosides-Rb2, Rb1, Rd and p-nitrophenyl glycoside derivatives which have less or no agglutinability were also investigated in the same manner. Only ginsenoside-Rb2 and p-nitrophenyl α-l-arabinofuranoside which have the specific sugar residue (arabinose) showed a strong interaction with the polar head groups of vesicles. To gain an insight into the mechanism of agglutination by ginsenoside-Rc, the interaction with the fatty acyl groups was also studied by using phosphatidylcholine spin-labeled at the fatty acyl groups. Ginsenoside-Rc increased the order parameter of the spin-labeled phosphatidylcholine, indicating that the saponin was inserted into lipid bilayers. In other saponins investigated, only ginsenoside-Rb2 interacted with the fatty acyl part of vesicles. The process of expression of agglutination by ginsenoside-Rc was discussed on the basis of the ESR studies.  相似文献   

17.
The interaction between fleroxacin (FLX) and pepsin was investigated by spectrofluorimetry. The effects of FLX on pepsin showed that the microenvironment of tryptophan residues and molecular conformation of pepsin were changed based on fluorescence quenching and synchronous fluorescence spectroscopy in combination with three‐dimensional fluorescence spectroscopy. Static quenching was suggested and it was proved that the fluorescence quenching of pepsin by FLX was related to the formation of a new complex and a non‐radiation energy transfer. The quenching constants KSV, binding constants K and binding sites n were calculated at different temperatures. The molecular interaction distance (r = 6.71) and energy transfer efficiency (E = 0.216) between pepsin and FLX were obtained according to the Forster mechanism of non‐radiation energy transfer. Hydrophobic and electrostatic interaction played a major role in FLX–pepsin association. In addition, the hydrophobic interaction and binding free energy were further tested by molecular modeling study. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

18.
The interaction of two types of vesicle systems was investigated: micrometer-sized, giant unilamellar vesicles (GUVs) formed from 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) and submicrometer-sized, large unilamellar vesicles (LUVs) formed from oleic acid and oleate, both in a buffered aqueous solution (pH 8.8). Individual POPC GUVs were transferred with a micropipette into a suspension of oleic acid/oleate LUVs, and the shape changes of the GUVs were monitored using optical microscopy. The behavior of POPC GUVs upon transfer into a 0.8 mM suspension of oleic acid, in which oleic acid/oleate forms vesicular bilayer structures, was qualitatively different from the behavior upon transfer into a 0.3 mM suspension of oleic acid/oleate, in which oleic acid/oleate is predominantly present in the form of monomers and possibly non-vesicular aggregates. In both cases, changes in vesicle morphology were observed within tens of seconds after the transfer. After an initial increase of the vesicle cross-section, the vesicle started to evaginate, spawning dozens of satellite vesicles connected to the mother vesicle with narrow necks or tethers. In 60% of the cases of transfer into a 0.8 mM oleic acid suspension, the evagination process reversed and proceeded to the point where the membrane formed invaginations. In some of these cases, several consecutive transitions between invaginated and evaginated shapes were observed. In the remaining 40% of the cases of transfer into the 0.8 mM oleic acid suspension and in all cases of vesicle transfer into the 0.3 mM oleic acid suspension, no invaginations nor subsequent evaginations were observed. An interpretation of the observed vesicle shape transformation on the basis of the bilayer-couple model is proposed, which takes into account uptake of oleic acid/oleate molecules by the POPC vesicles, oleic acid flip-flop processes and transient pore formation.  相似文献   

19.
Ortiz A  Cajal Y  Haro I  Reig F  Alsina MA 《Biopolymers》2000,53(6):455-466
The interaction of the multiple antigenic peptide MAP4VP3 with lipid membranes has been studied by spectroscopic techniques. MAP4VP3 is a multimeric peptide that corresponds to four units of the sequence 110-121 of the capsid protein VP3 of hepatitis A virus. In order to evaluate the electrostatic and hydrophobic components on the lipid-peptide interaction, small unilamelar vesicles of different compositions, including zwitterionic dipalmitoylphosphatidylcholine (DPPC), anionic dipalmitoylphosphatidylcholine/phatidylinositol (DPPC:PI 9:1), and cationic dipalmitoylphosphatidylcholine/stearylamine (DPPC:SA 9.5:0.5), were used as membrane models. Intrinsic tryptophan fluorescence changes and energy transfer experiments show that MAP4VP3 binds to all three types of vesicles with the same stoichiometry, indicating that the electrostatic component of the interaction is not important for binding of this anionic peptide. Steady-state polarization experiments with vesicles labeled with 1,6-diphenyl-1,3,5-hexatriene or with 1-anilino-8-naphtalene sulphonic acid indicate that MAP4VP3 induces a change in the packing of the bilayers, with a decrease in the fluidity of the lipids and an increase in the temperature of phase transition in all the vesicles. The percentage of lipid exposed to the bulk aqueous phase is around 60% in intact vesicles, and it does not change upon binding of MAP4VP3 to DPPC vesicles, indicating that the peptide does not alter the permeability of the membrane. An increase in the amount of lipid exposed to the aqueous phase in cationic vesicles indicates either lipid flip-flop or disruption of the vesicles. Binding to DPPC vesicles occurs without leakage of entrapped carboxyfluorescein, even at high mol fractions of peptide. However, a time-dependent leakage is seen with cationic DPPC/SA and anionic DPPC/PI vesicles, indicating that the peptide induces membrane destabilization and not lipid flip-flop. Resonance energy transfer experiments show that MAP4VP3 leakage from cationic vesicles is due to membrane fusion, whereas leakage from anionic vesicles is not accompanied by lipid mixing. Results show that MAP4VP3 interacts strongly with the lipid components of the membrane, and although binding is not of electrostatic nature, the bound form of the peptide has different activity depending on the membrane net charge; thus, it is membrane disruptive in cationic and anionic vesicles, whereas no destabilizing effect is seen in DPPC vesicles.  相似文献   

20.
The fluorescence decay kinetics of 1-methylpyrene in small unilamellar l-alpha-dimyristoylphosphatidylcholine vesicles above the phase transition temperature has been studied as a function of concentration and temperature. When the 1-methylpyrene/phospholipid ratio equals 1:2000 no excimer is observed and the fluorescence decay is monoexponential. When this ratio is equal to or higher than 1 200, excimer is observed and the monomer and excimer decays can be adequately described by two exponential terms. The deviation of the monomer decays from monoexponentiality cannot be described by a model where the diffusion-controlled excimer formation is time dependent. The observed decays are compatible with the excimer formation scheme which is valid in an isotropic medium. The activation energy of excimer formation is found to be 29-9 +/-1.4 kJ mol . The (apparent) excimer formation constant and the excimer lifetime at different temperatures have been determined. The diffusion coefficient associated with the excimer formation process varies between 2 x 10(-10) m(2)/s at 70 degrees C to 4 x 10(-11) m(2)/s at 25 degrees C.  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