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 31 毫秒
1.
The influence of proline cis-trans isomerization on the kinetics of lysozyme unfolding was examined carefully according to the theory of Hagerman and Baldwin [(1976) Biochemistry 15, 1462–1473]. As a result, the kinetics of lysozyme unfolding was found to follow the two-state transition model well. The temperature dependencies of kuf and kf over a wide temperature range showed that ΔC = 0 and ΔC = ?6.7 kJ K?1 mol?1 in solutions of different concentrations of GuHCl. The data observed in solutions containing other denaturants also supported the conclusion that ΔC is nearly equal to zero. The activation enthalpies of unfolding (ΔH) were observed at various concentrations of several kinds of denaturants. They were independent of species and concentrations of denaturants ΔH = 200 kJ mol?1). These facts indicate that the aspect of interaction between protein and different kinds of solvent molecules varies only slightly during the unfolding to the transition state, that is, the transition state is at compact as the native one. Therefore, it is also suggested that ΔH of 200 kJ mol?1 is primarily required for the disruption of long-range interactions among different structural domains through a subtle conformational change. We compared the effects of several kinds of denaturants on the unfolding rate. The addition of PrOH more remarkably increases the unfolding rate than do other hydrophilic denaturants. This is probably because PrOH molecules can penetrate into the hydrophobic core of lysozyme, but hydrophilic reagents cannot because of the compactness of the transition state.  相似文献   

2.
3.
Wei Liu  Takashi Norisuye 《Biopolymers》1988,27(10):1641-1654
Weight-average molecular weights Mw, second virial coefficients, and z-average radii of gyration 〈S2〉 were determined by light scattering as a function of temperature T for four sodium salt samples of xanthan in 0.01M aqueous NaCl, in which the polysaccharide undergoes an order–disorder conformation change with increasing T. The data for 〈S2〉 and Mw at 25 and 80°C, the lowest and highest temperatures studied, confirmed the previous conclusion that the predominant conformation at the former T, i.e., in the ordered state, is a double helix, while that at the latter T, i.e., in the disordered state, is a dimerized coil expanded by electrostatic repulsions between charged groups of the polymer. As T was increased from 25 to 80°C, 〈S2〉 sigmoidally decreased or increased depending on the dimer's molecular weight. This temperature dependence of 〈S2〉 and that determined elsewhere for a high molecular weight sample were found to be described almost quantitatively by a simple dimer model in which the double helix melts from both ends, when the double-helical fraction in the dimer at a given T estimated previously from optical rotation data was used.  相似文献   

4.
Densities of solutions of several α-amino acids and peptides in 3 and 6m aqueous urea solvents have been determined at 298.15 K. These data have been used to evaluate the infinite-dilution apparent molar volumes of the solutes and the volume changes due to transfer (V ) of the α-amino acids and peptides at infinite dilution from water to aqueous urea solutions. The sign and magnitude of the V values have been rationalized in the framework of Friedman's cosphere-overlap model. The V values for the glycyl group (? CH2CONH? ) and alkyl side chains have been estimated.  相似文献   

5.
Small-angle x-ray scattering of poly(γ-methyl-L -glutamate), [Glu(OMe)]n, in m-cresol and in pyridine was measured to determine the mass per unit length, Mq, and the radius of gyration of the cross section, 〈S1/2. It was confirmed from the values of Mq that [Glu(OMe)]n exists in an α-helical conformation in these solvents. It was elucidated from the calculations on 〈S1/2 that the side chains come in moderately close contact with the main chain in these solvents. It was indicated from the analysis of the outer portion of the scattering curves that the side-chain conformation varied depending on the solvent.  相似文献   

6.
Four fundamental Raman lines were observed at 159, 111, 55 and 27 cm-1 corresponding to the I bound (I) in amyloses with DP from 20 to 100, regardless of the degree of polymerization of I and the excitation wavelength. The spectral resolution was based on the molar extinction coefficient and molar ellipticity spectra of I. Eight bands, named, S1, S2, ?, S8 from long to short wavelength, were isolated. These were found regardless of the DP. By a resonance excitation Raman study, the characteristics of S3 and S4, comprising the shoulder around 480 nm, were found to be different from those of S1 and S2, comprising the blue band. The assignment of the spectra was based on the electronic states of the monomeric I in the exciton-coupled dimeric unit. It was concluded that the blue band (S1,S2) belonged to the long-axis transitions and the shoulder band (S3,S4) to the short-axis ones on the monmeric coordinate system.  相似文献   

7.
The conformational transition of poly(L -agrignine) by binding with various mono-, di-, and polyvalent anions, especially with SO, was studied by CD measurements. The intramolecular random coil-to-α-helix conformational transition and the subsequent transition to the β-turn-like structure was caused by binding with SO. The binding data obtained from equilibrium dialysis experiments showed that the α-helical conformation of poly(L -arginine) is stabilized at a 1:3 stoichiometric ratio of bound SO to arginine residue; at higher free SO concentrations, the α-helix converts to the β-turn-like structure accompanied by a decrease in amount of bound SO. The same conformaitonal transition of poly(L -arginine) also occurred in the solutions of other divalent anions (SO, CO, and HPO) and polyvalent anions (P2O, P3O). Among the monovalent anions examined, CIO and dodecyl sulfate were effective in including α-helical conformation, while the other monovalent anions (OH?, Cl?, F?, H2PO, HCO and CIO) failed to induce poly(L -arginine) to assume the α-helical conformation. Thus, we noticed that, except for dodecyl sufate, the terahedral structure is common to the α-helix-forming anions. A well-defined model to the α-helical poly(L -arginine)/anion complex was proposed, in which both the binding stoichiometry of anions to the arginine residue and the tetrahedral structure of anions were taken into consideration. Based on these results, it was concluded that the tetrahedral-type anions stabilize the α-helical conformation of poly(L -arginine) by crosslinking between two guanidinium groups of nearby side chains on the same α-helix through the ringed structures stabilized by hydrogen bonds as well as by electrostatic interaction. Throughout the study it was noticed that the structural behavior of poly(L -arginine) toward anions is distinct from that of poly(L -lysine).  相似文献   

8.
The kinetics of binding of the cationic surfactant cetyltrimethyl ammonium bromide with the Na salt of carboxymethyl cellulose was studied by the electrometric method using cetyltrimetlyl ammonium+ (CTA+) ion-selective polyvinyl chloride membrane electrode. The binding process followed the first-order kinetics and occurred in three stages. Its affinity increased with increasing CTA bromide concentration and decreased with ionic strength. The activation process comprised moderate E and ΔH and negative ΔS for all three stages with a ΔH < TδS trend proving it to be entropy controlled. The ΔG values followed the trend ΔG < ΔG < ΔG (in accordance with k1 > k2 > k3). The enthalpies (ΔH) and entropies (ΔS) of activation followed a systematic and interdependent trend. The multiple-stage binding kinetics is grossly comparable with the kinetics of binding of proteins to solid surfaces. © 1995 John Wiley & Sons, Inc.  相似文献   

9.
Empirical force-field calculations and ir and 1H-nmr spectra indicate that five-membered (C5) and seven-membered (C) hydrogen-bonded rings are the preferred conformations of acetyl-L -Phe p-acetyl and p-valeryl anilides in nonpolar media. The C5/C ratio was found to be dependent on the dryness of the solute and the solvent. This fact and the results from conformational-energy calculations suggest that a molecule of water participates in the stabilization of the C conformation.  相似文献   

10.
We describe conditions which lead to complete helix formation of poly(I) in the presence of NH. Binding of NH is shown to be specific in the presence of Li+, which does not by itself support helix formation under these conditions. The NH–poly(I) complex is characterized by uv, CD, and ir spectroscopy. The CD spectrum is strikingly different from those of the Na+ or K+ complexes, the first extremum being changed from negative for the metal ions to positive for NH. A stereospecific model is proposed for the NH–poly(I) helix in which the N of NH is located on the axis of the four-stranded helix, midway between planar tetramers formed by the bases. The model is consistent with the tetrahedral symmetry of NH, the requirement for four acceptable hydrogen bonds, the observed stability of the helix, and the accepted geometry of the backbone.  相似文献   

11.
The kinetics of ethidium's intercalative binding to DNA packaged in bacteriophage T7 and two T7 deletion mutants have been determined, using enhancement of fluorescence to quantitate binding. At a constant ethidium concentration, the results can be described as first-order binding with two different rate constants, k (= k1 + k?1) and k (= k2 + k?2). The larger rate constant (k) was at least four orders of magnitude smaller than the comparable first-order forward rate constant for binding to DNA released from its capsid. At 25°C values of k decreased as the amount of DNA packaged per internal volume increased. This latter observation indicates that the rate of ethidium's binding to packaged T7 DNA is limited by an event that occurs inside of the DNA-containing region of T7, not by the crossing of T7 capsid's outer shell. Arrhenius plots of kM are biphasic, indicating a transition for packaged DNA at a temperature of 20°C. The data indicate that k s are limited by either sieving of ethidium during its passage through the packaged DNA or subsequent hindered intercalation.  相似文献   

12.
If the collagen triple helix is so built as to have one set of NH ? O hydrogen bonds of the type N3H3(A) ? O2(B), then it is possible to have a linkage between N1H1(B) and O1(A) through the intermediary of a water molecule with an oxygen O leading to the formation of the hydrogen bonds N1(B) ? O and O (A). In the same configuration, another water molecule with an oxygen O can link two earbonyl oxygens of chains A and B forming the hydrogen bonds O O1(A) and O O0 (B). The two water oxygens also become receptors at the same time for CH ? O hydrogen bonds. Thus, the neighboring chains in the triple helix are held together by secondary valence bond linkages occurring regularly sit intervals of about 3 Å along the length of the protofibril. The additional water molecules occur on the periphery of the proto-fibril and will contribute their full share towards stabilizing the structure in the solid state. In solution, they will be disturbed by the medium unless they are protected by long side groups. It appears that this type of two-bonded structure, in which one NH ? O bond is to a water molecule, can explain several observations on the stability and hydrogen exchange properties of collagen itself and related synthetic polypeptides. The nature of the water bonds and their strength are found to be better in the one-bonded structure proposed from Madras than in the one having the coordinates of Rich and Crick.  相似文献   

13.
The compositional buoyant densities, ρ;, of human γ-immunoglobulin, bovine serum mercaptalbumin, and egg albumin have been measured in CsCl solutions in the analytical ultracentrifuge as a function or pressure. Standard pressure coefficients, ψ0, and standard partial specific volumes of the solvated proteins, υ ,0, have been computed from these data. The ψ0 values obtained are strikingly different from each other and from the only other pressure coefficients which have been measured, those values obtained for nucleic acids and nucleoproteins. The ψ value for γ-immunoglobulin is negative, the first nonpositive value obtained, and suggests an unusual internal structure for this protein. The pressure coefficient of mercaptalbumin is not constant. A second-order relation is derived and utilized to interpret these data. The slope of the ρ(P) plot for egg albumin was constant and negative and yielded values of ψ0 which are about 20% as large as those reported for DNA. Evaluation of published isopiestic data for egg albumin in CsCl solutions provided the dependence of preferential hydration on water activity. This quantity, (dΓ′/da) as well as α, were found to be negative. The values of ψ0 and α were used to compute the effective density gradient from which the correct molecular weight of egg albumin was obtained. The apparent specific volume of egg albumin in a buoyant CsCl solution was measured using the Mettler-Paar densimeter.  相似文献   

14.
The low-temperature heat capacity of collagen (in the hydrated and dehydrated states) and the large entropy of collagen in the coiled state relative to the same protein in the helical state were investigated. The heat capacity for collagen in the solid state in the temperature range 4°–50° K changes proportionally to the square of temperature (CpT2). Above 50°K there is a linear dependence (CpT). The differences in the character of temperature dependence of heat capacity for the hydrated and dehydrated collagen show the importance of the specific interaction of water molecules with polypeptide chains of this protein. The peculiarities of the temperature dependence of the heat capacity difference (ΔCp) of hydrated denatured (random coiled) and hydrated native (helical) collagen are observed at 15°, 120°, and 240°K. These differences are caused by the varying degree of ordering of the hydrate water molecules in native and denatured collagen macromolecules. At all temperatures (4°–300°K) the entropy of the random coiled state is higher than that of collagen in the native state and at 298°K ΔS = ∫ (ΔCp/T)dT = 0.8 cal/100 g °K.  相似文献   

15.
Equilibrium unfolding (folding) studies reveal that the autoregulatory RNA pseudoknots derived from the bacteriophage T2 and T4 gene 32 mRNAs exhibit significant stabilization by increasing concentrations of divalent metal ions in solution. In this report, the apparent affinities of exchange inert trivalent Co(NH3) have been determined, relative to divalent Mg2+, for the folded, partially folded (Kf), and fully unfolded (Ku) conformations of these molecules. A general nonspecific, delocalized ion binding model was developed and applied to the analysis of the metal ion concentration dependence of individual two‐state unfolding transitions. Trivalent Co(NH3) was found to associate with the fully folded and partially unfolded pseudoknotted forms of these RNAs with a Kf of 5–8 × 104 M−1 in a background of 0.10 M K+, or 3‐ to 5‐fold larger than the Kf obtained for two model RNA hairpins and hairpin unfolding intermediates, and ≈ 40–50‐fold larger than Kf for Mg2+. The magnitude of Kf was found to be strongly dependent on the monovalent salt concentration in a manner qualitatively consistent with polyelectrolyte theory, with Kf reaching 1.2 × 105 M−1 in 50 mM K+. Two RNA hairpins were found to have affinities for Co(NH3) and Ru(NH3) of 1–2 ×104 M−1, or ≈ 15‐fold larger than the Kf of ∼ 1000 M−1 observed for Mg2+. Additionally, the Ku of 4,800 M−1 for the trivalent ligands is ≈ 8‐fold larger than the Ku of 600 M−1 observed for Mg2+. These findings suggest that the T2 and T4 gene 32 mRNA pseudoknots possess a site(s) for Mg2+ and Co(NH3) binding of significantly higher affinity than a “duplexlike” delocalized ion binding site that is strongly linked to the thermodynamic stability of these molecules. Imino proton perturbation nmr spectroscopy suggests that this site(s) lies near the base of the pseudoknot stem S2, near a patch of high negative electrostatic potential associated with the region where the single loop L1 adenosine crosses the major groove of stem S2. © 1999 John Wiley & Sons, Inc. Biopoly 50: 443–458, 1999  相似文献   

16.
Integral enthalpies of solution of several dipeptides and tripeptides in water at low concentrations have been determined at 25 and 35°C. These data have been used to derive the changes in heat capacity on dissolution at infinite dilution ΔC at 30°C. Limiting partial molal heat capacities ΔC have been determined by combining ΔC with Cp2 (heat capacity of pure solid peptides). Using the data on ω-amino acids and these peptides, the partial molal heat capacity of a peptide group ? CONH? was semiquantitatively estimated.  相似文献   

17.
Proton magnetic resonance spectra of model dipeptide molecules R1–C′1O1–N2H2–CHR2–C′2O2–N3H3–R3 in CCl4 solutions exhibit splited signals when investigating on mixtures of L and D enantiomers differing from the racemic composition. The major effect is observed on amide proton signals which are the ones most sensitive to the ratio of aggregation. The stereoselective dimerization of enantiomeric molecules in the so-called C5 conformational state is shown to be responsible for such a phenomenon, the intensity of which depends on the bulkiness of the side chain R2. A theoretical approach is proposed which gives predictions in close agreement with our own experimental findings.  相似文献   

18.
Effects of hydrophobic and electrostatic interactions on the self‐assembling process of the ionic‐complementary peptide EMK16‐II are investigated by atomic force microscopy imaging, circular dichroism spectra, light scattering, and chromatography. It is found that the hydrophobicity of the peptide promotes the aggregation in pure water even at a very low concentration, resulting in a much lower critical aggregation concentration than that of another peptide, EAK16‐II. The effect of anions in solution with different valences on electrostatic interactions is also important. Monovalent anions (Cl? and Ac?) with a proper concentration can facilitate the formation of peptide fibrils, with Cl? of smaller size being more effective than Ac? of larger size. However, only small amounts of fibrils, but plenty of large amorphous aggregates, are found when the peptide solution is incubated with multivalent anions, such as SO, C6H5O, and HPO. More importantly, by gel filtration chromatography, the citrate anion, which induces a similar effect on the self‐assembling process of EMK16‐II as that of SO and HPO, can interact with two or more positively charged residues of the peptide and reside in the amorphous aggregates. This implies a “salt bridge” effect of multivalent anions on the peptide self‐assembling process, which can interpret a previous puzzle why divalent cations inhibit the formation of ordered nanofibrils of the ionic‐complementary peptides. Thus, our results clarify the important effects of hydrophobic and electrostatic interactions on the self‐assembling process of the ionic‐complementary peptides. These are greatly helpful for us to understand the mechanism of peptides' self‐assembling process and protein folding and aggregation. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 318–329, 2010. This article was originally published online as an acceptedpreprint. The “Published Online” date corresponds to the preprintversion. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

19.
Diemer formation of poly(L -lysine HBr) in carbonate buffer at pD 10.5 was reported in our previous paper [Biopolymers(1981) 20 , 345–357]. The mechanism of the dimer formation was investigated employing carbon-13 and proton nmr. pD dependence of the 13C-nmr spectrum of poly(L -lysine HBr) in the presence of carbonate ion clearly shows that a complex formation between the CO ion and protonated ε-amino group is involved in the stabilization of the dimer form. The lifetime of the complex is longer than 10?2 s. A stoichiometric evaluation suggests that CO bridges two lysyl side chains. A molecular model of the dimer form designated as a single antiparallel β-ribbon was proposed and discussed in the light of hydrodynamic and ir spectroscopic properties reported earlier. Concentration change experiments indicate that CO binds not only to the dimer formation is inferred as stabilization of the single antiparallel β-ribbon as an intermediate structure in the conversion between the α-helix and β-sheet. The α-CH proton signal of the lysine residue located in the ordered structure (α-helix and β-form) was observed to be separate from that in the random-coil region.  相似文献   

20.
A Cabani  A Paci  V Rizzo 《Biopolymers》1976,15(1):113-129
Using the formalism of nearest-neighbor Ising model and assuming that the allowed states for a monomeric unity of a polypeptide chain in solutions containing strong acids are E (helix), C (coil), and CS (solvent-bonded coil), the partition function of the system was deduced analytically. Equations were obtained which permitted the prediction of the characteristic thermodynamic behavior of the helix–coil transition under these conditions. These equations were used to examine critically the possible correlations between experimental data obtained using different techniques. Particular attention was devoted to quantities called “transition enthalpies,” obtained from the slope of the transition curves at the point where the helix fraction is one-half (ΔH), or for measurements of the heat of solution of the polymer over the total range of solvent composition (ΔH), or from heat capacity measurements taken at various temperatures (ΔH). Literature data of ΔH(j = opt, sol, cal) for the system poly-γ-benzyl-L -glutamate in mixtures of dichloroacetic acid and 1,2-dichloroethane were carefully analyzed.  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