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.
Deamidation kinetics were measured for a model hexapeptide (L-Val-L-Tyr-L-Pro-L-Asn-Gly-L-Ala, 0.02 mg/mL) in aqueous solutions containing glycerol (0-50% w/w) and poly(vinyl pyrrolidone) (PVP, 0-20% w/w) at 37 degrees C and pH 10 to determine the effects of solution polarity and viscosity on reactivity. The observed pseudo-first order deamidation rate constants, k(obs), decreased markedly when the viscosity increased from 0.7 to 13 cp, but showed no significant change at viscosities >13 cp. Values of k(obs) also increased with increasing dielectric constant and decreasing refractive index. Molecular dynamics simulations indicated that the free energy associated with Asn side-chain motion is insensitive to changes in dielectric constant, suggesting that the observed dielectric constant dependence is instead related primarily to the height of the transition state energy barrier. An empirical model was proposed to describe the effects of the viscosity, refractive index and dielectric constant on k(obs). Analysis of the regression coefficients suggested that both permanent and induced dipoles of the medium affect the deamidation rate constant, but that solution viscosity is relatively unimportant in the range studied.  相似文献   

2.
B Gavish  M M Werber 《Biochemistry》1979,18(7):1269-1275
The effect of viscosity on the rate of catalysis of carboxypeptidase A has been tested. By use of the tripeptide carbobenzoxy-l-alanyl-l-alanyl-l-alanine [Z(L-Ala)3] as substrate, it was shown that most of the effect on the hydrolysis rate caused by the presence of 30 or 40% methanol or glycerol in aqueous solution can be ascribed to a contribution of viscosity to the catalytic rate constant, kcat. Arrhenius plots of kcat in 30 and 40% glycerol or methanol are linear and almost parallel. When the rate constants are "corrected" for the viscosity of various media, the difference between the various Arrhenius plots is considerably reduced; it vanishes, within experimental error, when the effect of the dielectric constant of the solutions is taken into account as well. It is proposed that the viscosity of the medium can influence the rate-limiting step of the enzymic reaction, which is the rate of transitions over the energy barrier preceding product formation. According to the suggested mechanism, the enzyme--substrate complex can overcome this energy barrier by viscosity-dependent structural fluctuations. The quantitative agreement between the theory and the experimental results suggests that (a) due to the temperature dependence of the viscosity of the solution, the potential energy barrier of the reaction is about 5 kcal/mol lower than the observed activation energy and (b) information about the structural flexibility of the complex can be obtained by kinetic measurements.  相似文献   

3.
Mo Y  Nishinari K 《Biorheology》2001,38(5-6):379-387
The extensional viscosity and the steady shear viscosity of sodium type hyaluronan (NaHA) in water with sodium chloride and/or sucrose and in DMSO solvent were measured. The extensional viscosities for HA in aqueous solution (0.05, 0.1, 0.3 w/v%) were constant at lower extensional rates, and then became strain thinning above a critical extensional rate. However, on adding sodium chloride, the extensional viscosity decreased and became strain thickening at higher extensional rates. Sodium ions shield the electrostatic repulsion between carboxyl residues of HA molecules and constrict the coil dimensions. The strain thickening of HA solution in the presence of sodium chloride at higher extension rates is due to the coil stretching. The addition of sucrose increased the extensional viscosity and shifted the critical extensional rate to lower strain rates. With increasing strain (shear) rates, extensional (shear) viscosities for HA aqueous solutions remained constant up to a critical extension (shear) rate; but they showed no plateau and decreased linearly in DMSO. It is clear that molecular interaction of HA in DMSO is stronger than that in aqueous solution. This should be attributed to the different conformations of HA in DMSO and in aqueous solutions.  相似文献   

4.
The stability of a 15-residue peptide has been investigated using CD spectroscopy and molecular simulation techniques. The sequence of the peptide was designed to include key features that are known to stabilize alpha-helices, including ion pairs, helix dipole capping, peptide bond capping, and aromatic interactions. The degree of helicity has been determined experimentally by CD in three solvents (aqueous buffer, methanol, and trifluoroethanol) and at two temperatures. Simulations of the peptide in the aqueous system have been performed over 500 ps at the same two temperatures using a fully explicit solvent model. Consistent with the CD data, the degree of helicity is decreased at the higher temperature. Our analysis of the simulation results has focused on competition between different side-chain/side-chain and side-chain/main-chain interactions, which can, in principle, stabilize the helix. The unfolding in aqueous solution occurs at the amino terminus because the side-chain interactions are insufficient to stabilize both the helix dipole and the peptide hydrogen bonds. Loss of capping of the peptide backbone leads to water insertion within the first peptide hydrogen bond and hence unfolding. In contrast, the carboxy terminus of the alpha-helix is stable in both simulations because the C-terminal lysine residue stabilizes the helix dipole, but at the expense of an ion pair.  相似文献   

5.
Although nectar feeding in insects has long been studied, the knowledge of the effect of nectar energy content on the ingestion dynamics separately from the viscosity of the fluid is very limited. To determine the effects of both factors on the feeding behavior of the hovering hawk moth Macroglossum stellatarum, we developed a method to independently manipulate sucrose concentrations and viscosity. The intake rate was analyzed as a function of sucrose concentration, the concentration at constant viscosity (kept constant by adding tylose, an inert polysaccharide), and of the different viscosities of a 30% weight/weight (w/w) sucrose solution (by adding different amounts of tylose). By increasing the concentration, and thus its viscosity, the solution intake rate (in microl s (-1)) decreased beyond a 20% w/w sucrose solution. For a 30% sucrose solution, the intake rate decreased with increasing viscosity. At constant viscosity, the solution intake rate decreased beyond a 30% w/w sucrose solution. However, if we considered the quantity of sucrose ingested per unit time (sucrose intake rate), the same fitted maximum was attained for both series in which the sucrose concentration changed (33.6% w/w). Results suggest that the gustatory input affects the dynamics of fluid ingestion separately from the viscosity.  相似文献   

6.
A simple and inexpensive aqueous two-phase system for the affinity partitioning of proteins is introduced. An aqueous solution consisting of maltodextrin (M100; molecular mass, 1800) and polyvinylpyrrolidone (PVP360; molecular mass, 360,000) formed two phases at 4 degrees C when the concentration of the polymers was 22.5% (w/w) and 4.0% (w/w), respectively. When the amino derivatives of chlorotriazine textile dyes or other azo textile dyes were added to the two-phase system they partitioned asymmetrically, favoring the upper, less dense, PVP360-rich phase. The association of the textile dyes with PVP360 did not prevent them from acting as affinity ligands for proteins. Three of the dyes screened increased the partition coefficient of purified lysozyme nearly 50-fold over a control containing no dye. Parameters such as pH, ionic strength, and dye concentration modulated the affinity-partitioning effect of the system. The partition coefficient of lysozyme in an egg white protein mixture increased severalfold as the total protein content of the system approached 4% (w/w), indicating that protein concentration is also important in determining the partitioning characteristics of this two-phase system. Proteins were efficiently freed of PVP360 and textile dye by recovery in a high-salt solution when another two-phase system was formed upon the addition of a solution of concentrated potassium phosphate to the isolated upper phase of a PVP360/M100/textile dye two-phase system. The affinity-partitioning system presented here allows one to screen large numbers of potentially useful protein ligands to optimize protein separation, followed by direct scaleup to a system size determined by the user.  相似文献   

7.
Abstract NMR and CD spectroscopy have been used to examine the conformation of the peptide, β(12-28), (VHHQKLVFFAEDVGSNK) in aqueous and 60% TFE/40% H(2)0 solution at pH 2.4. In 60% TFE solution, the peptide is helical as confirmed by the CD spectrum and by the pattern of the NOE cross peaks detected in the NOESY spectrum of the peptide. In aqueous solution, the peptide adopts a more extended and flexible conformation. Broadening of resonances at low temperature, temperature-dependent changes in the chemical shifts of several of the CH(α) resonances and the observation of a number of NOE contacts between the hydrophobic side-chain protons of the peptide are indicative of aggregation in aqueous solution. The behavior of β(12-28) in 60% TFE and in aqueous solution are consistent with the overall conformation and aggregation behavior reported for the larger peptide fragment, β(1-28) and the parent β-amyloid peptide.  相似文献   

8.
9.
Abstract

NMR and CD spectroscopy have been used to examine the conformation of the peptide, β(12–28), (VHHQKLVFFAEDVGSNK) in aqueous and 60% TFE/40% H20 solution at pH 2.4. In 60% TFE solution, the peptide is helical as confirmed by the CD spectrum and by the pattern of the NOE cross peaks detected in the NOESY spectrum of the peptide. In aqueous solution, the peptide adopts a more extended and flexible conformation. Broadening of resonances at low temperature, temperature-dependent changes in the chemical shifts of several of the CHα resonances and the observation of a number of NOE contacts between the hydrophobic side-chain protons of the peptide are indicative of aggregation in aqueous solution. The behavior of β(12–28) in 60% TFE and in aqueous solution are consistent with the overall conformation and aggregation behavior reported for the larger peptide fragment, β(1–28) and the parent β-amyloid peptide.  相似文献   

10.
The nonenzymatic rates of deamidation of Asn residues in a series of pentapeptides with the sequences VSNXV and VXNSV, where X is one of 10 different amino acids, were determined at neutral, alkaline, and acid pH values. The results demonstrate that in neutral and alkaline solutions the amino acid residue on the amino side of the Asn had little or no effect on the rate of deamidation regardless of its charge or size. The group on the carboxyl side of Asn affected the rate of deamidation significantly. Increasing size and branching in the side chain of this residue decreased the rate of deamidation by as much as 70-fold compared to glycine in the N-G sequence, which had the greatest rate of deamidation. In acidic solution, the rate of deamidation of the Asn residue was not affected by the amino acid sequence of the peptide. The products for each deamidation reaction were tested for the formation of isoAsp residues. In neutral and alkaline solutions, all products showed that the isoAsp:Asp peptide products were formed in about a 3:1 ratio. In acidic solution, the Asp peptide was the only deamidation product formed. All peptides in which a Ser residue follows the Asn residue were found to undergo a peptide cleavage reaction in neutral and alkaline solutions, yielding a tripeptide and a dipeptide. The rate of the cleavage reaction was about 10% of the rate of the deamidation pathway at neutral and alkaline pH values. The rates of deamidation of Asn residues in the peptides studied were not affected by ionic strength, and were not specific base catalyzed. General base catalysis was observed for small bases like ammonia. A model for the deamidation reaction is proposed to account for the observed effects.  相似文献   

11.
The influence of solvent viscosity on the kinetic parameters of the pyruvate reduction reaction catalyzed by lactate dehydrogenase has been investigated. The viscosity was adjusted by sucrose and glycerol solutions at concentrations from 0 to 44% and from 0 to 63%, respectively. The reaction rate decreased abruptly with an increase in viscosity. The study of different reaction stages (enzyme-substrate complex formation, catalysis, inhibitory complex decomposition, competitive inhibition by chlorine ions) revealed that the catalysis (and the related conformational changes) is the only stage (of the above mentioned) that depends markedly on the solvent viscosity. The reaction is insensitive to the changes in the dielectric properties of the solution induced by the addition of alcohols and dioxane. The observed power dependence of the rate constant on viscosity is explained in terms of Kramer's theory which considers the proton transition through the activation barrier to be a diffusion in the field of random forces. The influence of solvent viscosity on enzymic kinetics indicates a direct relation between solvent dynamics and relevant protein conformational movements.  相似文献   

12.
The configurational entropy of a beta-heptapeptide in solution at four different temperatures is calculated. The contributions of the backbone and of the side-chain atoms to the total peptide entropy are analyzed separately and the effective contribution to the entropy arising from correlations between these terms determined. The correlation between the backbone and side-chain atoms amounts to about 17% and is rather insensitive to the temperature. The correlation of motion within the backbone and within side-chains is much larger and decreases with temperature. As the peptide reversibly folds at higher temperatures, its change in entropy and enthalpy upon folding is analyzed. The change in entropy and enthalpy upon folding of the peptide alone cannot account for the observed change in free energy on folding of the peptide in solution. Enthalpic and entropic contributions of the solvent thus also play a key role. Proteins 2001;43:45-56.  相似文献   

13.
Block copolymers can adopt a wide range of morphologies in dilute aqueous solution. There is a significant amount of interest in the use of block copolymer vesicles for a number of applications. We show that a series of oligo(valine) and oligo(phenylalanine) peptides coupled to a methacrylic group can be prepared by conventional peptide coupling techniques. These can be successfully polymerized by atom transfer radical polymerization (ATRP) in hexafluoroisopropanol (HFIP) giving access to poly(ethylene oxide)- b-poly(side-chain peptides). Many of these polymers self-assemble to form vesicles using an organic to aqueous solvent exchange. One example with a divaline hydrophobic block gives a mixture of toroids and vesicles. Circular dichroism demonstrates that secondary structuring is observed in the hydrophobic region of the vesicle walls for the valine side-chain containing polymers.  相似文献   

14.
Fluorescence measurements and singlet singlet energy transfer experiments on endothelin-1 provide information on the conformation of this peptide in dilute aqueous solution. The tyrosine fluorescence quantum yield in the absence of transfer (in [Phe21]endothelin-1) is relatively large (Φtyr = 0.39), indicating the side-chain is oriented away from fluorescence quenching groups such as the two disulfide bonds of the peptide. The fluorescence emission maximum (λ = 351 nm) and quantum yield (Φtrp = 0.099) of tryptophan in endothelin-I suggests that this residue is fully accessible to the solvent and that the indole ring is not located near the fluorescence quenching histidinium moiety or the disulfide bonds.

Singlet-singlet fluorescence energy transfer measurements of the Tyr13/Trp21 intramolecular distance by both donor fluorescence quenching and relative enhancement of acceptor fluorescence yield a distance of about 12.8 ± 0.6 Å. Molecular modeling of a fully extended C-terminal hexapeptide indicates a Tyr13/Trp21 distance of about 25 Å. Thus, the C-terminal residues must bend back towards the bicyclic portion of the molecule.  相似文献   


15.
H. Hanssum  H. Rüterjans 《Biopolymers》1980,19(9):1571-1585
13C spin-lattice relaxation times of poly(L -lysine) have been obtained at 67.9 MHz in aqueous solution and in a mixed solvent (40% methanol/60% water). A concomitant determination of the conformation by CD permits the correlation of conformation and rotational diffusion of the polymer. The dependence on pH of the spin-lattice relaxation times of the 13Cα and the side-chain carbon resonances reflects the diffusional motion in the random-coil conformation, in the helix–coil transition, and in the conformation of the α-helix. In the mixed solvent the reorientational correlation time of the Cα-Hα vector increases from τ = 0.37 nsec (random coil) to τ = 12.0 nsec (α-helix). In aqueous solution the correlation time of this vector increases from τ = 0.33 nsec (random coil) to τ ? 11 nsec. The reorientation rates of the side-chain methylene groups in the two solvents are markedly different. The reorientation of all methylene groups is reduced in the mixed solvent.  相似文献   

16.
Asn deamidation was monitored in Ala-based octadecapeptides of varying alpha-helicity. Gly was substituted for Ala residues at positions 6 and 16 to create a peptide with less helicity. Ala --> Gly substitutions were made at three or more residues from the Asn to negate known primary sequence effects on deamidation rates. The extent of helicity and rate of Asn deamidation for alkaline aqueous solutions of each peptide was measured as a function of temperature by circular dichroism and reversed-phase high-performance liquid chromatography, respectively. The rate of deamidation in the peptides was inversely proportional to the extent of alpha-helicity. The results support the conclusion that Asn deamidation only occurs in the nonhelical population of conformers.  相似文献   

17.
S J Demarest  Y Hua  D P Raleigh 《Biochemistry》1999,38(22):7380-7387
There are a small number of peptides derived from proteins that have a propensity to adopt structure in aqueous solution which is similar to the structure they possess in the parent protein. There are far fewer examples of protein fragments which adopt stable nonnative structures in isolation. Understanding how nonnative interactions are involved in protein folding is crucial to our understanding of the topic. Here we show that a small, 11 amino acid peptide corresponding to residues 101-111 of the protein alpha-lactalbumin is remarkably structured in isolation in aqueous solution. The peptide has been characterized by 1H NMR, and 170 ROE-derived constraints were used to calculate a structure. The calculations yielded a single, high-resolution structure for residues 101-107 that is nonnative in both the backbone and side-chain conformations. In the pH 6.5 crystal structure, residues 101-105 are in an irregular turn-like conformation and residues 106-111 form an alpha-helix. In the pH 4.2 crystal structure, residues 101-105 form an alpha-helix, and residues 106-111 form a loopike structure. Both of these structures are significantly different from the conformation adopted by our peptide. The structure in the peptide model is primarily the result of local side-chain interactions that force the backbone to adopt a nonnative 310/turn-like structure in residues 103-106. The structure in aqueous solution was compared to the structure in 30% trifluoroethanol (TFE), and clear differences were observed. In particular, one of the side-chain interactions, a hydrophobic cluster involving residues 101-105, is different in the two solvents and residues 107-111 are considerably more ordered in 30% TFE. The implications of the nonnative structure for the folding of alpha-lactalbumin is discussed.  相似文献   

18.
Nectar is an essential resource for bumblebees and many other flower-visiting insects. The main constituents of nectar are sugars, which vary in both composition and concentration between plant species. We assessed the influence of sugar concentration, sugar solution viscosity and sugar solution composition on the imbibition and energy intake rate of bumblebees, Bombus impatiens Cresson (Hymenoptera: Apidae). To do this, we measured their rate of solution intake for 49 different sugar solution treatments, which varied in both sugar composition and concentration. In general, the imbibition rates of bumblebees were found to increase with increasing sugar concentration, probably due to their preference for high sugar concentrations, up to a concentration of 27% (w/w), at which point solutions reached a threshold viscosity of approximately 1.5–1.6 mPa.s. Above this threshold, the increasing viscosity of the solutions physically inhibited the imbibition rates of bees, and imbibition rate began to decrease as the concentration increased. Nevertheless, bumblebee energy intake rate increased with increasing concentration up to about 42–56%. Although we found that sugar solution composition had an impact on both imbibition and energy intake rate, its effect was not as straightforward as that of sugar concentration and viscosity.  相似文献   

19.
Tobi D  Elber R  Thirumalai D 《Biopolymers》2003,68(3):359-369
The conformational equilibrium of a blocked valine peptide in water and aqueous urea solution is studied using molecular dynamics simulations. Pair correlation functions indicate enhanced concentration of urea near the peptide. Stronger hydrogen bonding of urea-peptide compared to water-peptide is observed with preference for helical conformation. The potential of mean force, computed using umbrella sampling, shows only small differences between urea and water solvation that are difficult to quantify. The changes in solvent structure around the peptide are explained by favorable electrostatic interactions (hydrogen bonds) of urea with the peptide backbone. There is no evidence for significant changes in hydrophobic interactions in the two conformations of the peptide in urea solution. Our simulations suggest that urea denatures proteins by preferentially forming hydrogen bonds to the peptide backbone, reducing the barrier for exposing protein residues to the solvent, and reaching the unfolded state.  相似文献   

20.
Beta2-Microglobulin (beta2-m) is a major structural component of dialysis-related amyloid fibrils. Kozhukh et al. [J. Biol. Chem. 277 (2002) 1310] prepared a series of peptide fragments of beta2-m by the protease digestion and examined their ability to form amyloid fibrils in citrate buffer at pH 2.5. Among various peptides, a 22-residue K3 peptide corresponding to Ser20-Lys41 spontaneously formed amyloid fibrils in aqueous solution. This peptide also formed amyloid protofibrils in 20% (v/v) 2,2,2-trifluoroethanol (TFE). To investigate the influence of solvent conditions on fibril formation, we studied their structures by atomic force microscopy. In aqueous solution, fibrils had a diameter of 4 or 8 nm and tended to cluster each other. On the other hand, protofibrils in 20% (v/v) TFE had a diameter of 2 nm with no tendency of clustering. Intriguingly, when the K3 protofibrils were transferred from 20% (v/v) TFE to aqueous solution, some of them associated to form thicker fibrils with a diameter of 4-15 nm and a left-handed helical twist. TFE is a hydrophobic solvent, so that hydrophobic interactions between molecules may be weakened. The results suggest that the fibrils in aqueous conditions are formed by the cooperative association of protofibrils at the growing ends of the fibrils, in which hydrophobic interactions play a major role.  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