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 140 毫秒
1.
The 1H-nmr studies were extensively carried out to elucidate preferred conformations of dipeptides CH3C*O—X—NHCH3, with X = Abu, nVal, and Val in various solvents. The vicinal 1H—1H coupling constants for the NH—CαH moiety and those around the Cα—Cβ bond in the articulated side chain provided the information regarding the average conformation of these molecules. The results indicate that transformation of skeletal conformations takes place in solution among conformers having similar dihedral angles, θ, in the Karplus expression.  相似文献   

2.
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   

3.
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   

4.
The solution structure of human salivary histatin 5 (D-S-H-A-K-R-H-H-G-Y-K-R-K-F-H-E-K-H-H-S-H-R-G-Y) was examined in water (pH 3.8) and dimethyl sulfoxide solutions using 500 MHz homo- and heteronuclear two-dimensional (2D) nmr. The resonance assignment of peptide backbone and side-chain protons was accomplished by 2D total correlated spectroscopy and nuclear Overhauser effect (NOE) spectroscopy. The high J values (≥7.4 Hz), absence of any characteristic NH-NH(i, i + 1) or CαH-CβH(i, i + 3) NOE connectivities, high dδ/dT values (≥0.004 ppm K−1) and the fast 1H/2H amide exchange suggest that histatin 5 molecules remain unstructured in aqueous solution at pH 3.8. In contrast, histatin 5 prefers largely α-helical conformation in dimethyl sulfoxide solution as evident from the J values (≤6.4 Hz), slow 1H/2H exchange, low dδ/dT values (≤0.003 ppm K−1) observed for amide resonances of residues 6–24, and the characteristic NH-NH(i, i + 1) and CαH-CβH(i, i +3) NOE connectivities. All backbone amide 15N-1H connectivities fall within 6 ppm on the 15N scale in the 2D heteronuclear single quantum correlated spectrum, and the restrained structure calculations using DIANA suggest the prevalence of α-helical conformations stabilized by 19 (5 → 1) intramolecular backbone amide hydrogen bonds in polar aprotic medium such as dimethyl sulfoxide. The interside-chain hydrogen bonding and salt-bridge type interactions that normally stabilize the helical structure of linear peptides in aqueous solutions are not observed. Histatin 5, unlike other naturally occurring antimicrobial polypeptides such as magainins, defensins, and tachyplesins, does not adopt amphiphilic structure, precluding its insertion into microbial membranes and formation of ion channels across membranes. Electrostatic (ionic type) and hydrogen bonding interactions of the positively charged and polar residues with the head groups of microbial membranes or with a membrane-bound receptor could be the initial step involved in the mechanism of antimicrobial activity of histatins. © 1998 John Wiley & Sons, Inc. Biopoly 45: 51–67, 1998  相似文献   

5.
K A Marx 《Biopolymers》1975,14(5):1103-1107
The magnitude of intramolecular DNA optical absorbance is presented as a function of [Na+] and GC composition of the DNA. The data are presented as a ratio of absorbances A0/A0 for the DNA at the denaturing temperature (Td) and renaturing temperature (Tr) under the given conditions. All ratios were determined for Tr corresponding to the temperature optimum (Tm ? Tr = 25°C) in DNA reassociation rate. This fact, coupled with the convenient A0/A0 ratio representation, permits the quick estimation of the magnitude of this optical effect in DNA reassociation reactions over a wide range of experimental conditions.  相似文献   

6.
The molecular theory of the previous paper in this series is extended to determine the effect of divalent metal ions on helix stability relative to coil at fixed ionic strength and nucleotide phosphate concentration. Specification of the state of condensed counterions, as well as their concentration, is essential for the solution of this problem, and it is assumed that they translate freely within a thin cylindrical shell close to the polynucleotide. As a function of divalent counterion concentration m the relative stability of the helix is highly nonlinear. Although the overall trend is that the helix stability increases with addition of divalent metal ion, there is a narrow concentration range for which it decreases slightly. The behavior of the relative stability as a function of m is determined by the translational degrees of freedom of the counterions, both univalent and divalent, both condensed and uncondensed. Detailed comparison of the theory with data is not given here, but it is pointed out that the calculated values of the relative stability are consistent with the order of magnitude of the observed effect Mg2+ on the melting temperature.  相似文献   

7.
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   

8.
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   

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.
Alan E. Tonelli 《Biopolymers》1978,17(5):1175-1179
Solution conformations of the cyclic pentapeptide plant-hormone malformin A, whose conformational freedom is constrained by an intramolecular disulfide bridge, are derived and presented here. The nmr and CD data of Ptak are used to place restrictions on the search for possible malformin A solution conformers of low energy. Only two distinct conformers were found to be consistent with Ptak's data. Both structures are characterized by an internally buried (solvent-shielded) D -Cys2 amide proton, a seven-membered (1–3)hydrogen bond between (N–H) and (O?C), and a disulfide bridge conformation with a P chirality as manifested in the nmr study by the temperature independence of the amide proton chemical shifts for the D -Cys2 and D -Leu4 residues and the negative sign of the long wavelength maximum in the CD spectrum, respectively. Inspection of space-filling molecular models of both structures indicates severe steric barriers to their rapid interconversion. Thus, it appears that only one of the two conformers may be present in solution. The difference in their calculated dipole moments (4.6 and 6.9D) suggests an experimental method for distinguishing between the two proposed solution structures.  相似文献   

11.
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   

12.
Intensity fluctuations of laser light scattered from filamentous viruses Pf1 [length L (Å) × diameter d (Å) = 20,000 × 90], M13 (9000 × 90), potato virus X (5150 × 130), and tobacco mosaic virus (3000 × 180) in sucrose density gradients were measured with a photon correlation spectrometer over a range of scattering angles from 15° to 120°. The experimental data can be approximated by two exponential decays, “slow” and “fast.” The slow decay rate constant t corresponds to the translational diffusion D of the virus, i.e., t = K2D, where K is the magnitude of the scattering vector. The amplitude of the slow component, i.e., translational diffusion, remains greater than that of the fast component, even at high KL. The fast decay rate constant t is also proportional to K2 for viruses such as Pf1, M13, and even potato virus X. In the companion paper, we shall attribute the amplitude enhancement of the translational diffusion to the coupling of its anisotropy to the rotational diffusion modes. In order to explain the excessive decay rates in the fast component, we need to consider the bending mode of rodlike viruses, especially in the longer viruses such as M13 and Pf1, in addition to the usually expected rotational diffusion modes.  相似文献   

13.
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   

14.
15.
The 1H, 13C, and 15N resonances of FKBP when bound to the immunosuppressant, ascomycin, were assigned using a computer-aided analysis of heteronuclear double and triple resonance three-dimensional nmr spectra of [U-15N] FKBP/ascomycin and [U-15N, 13C] FKBP/ascomycin. In addition, from a preliminary analysis of two heteronuclear four-dimensional data sets, 3J coupling constants, amide exchange data, and the differences between the Cα and Cβ chemical shifts of FKBP to random coil values, the secondary structure of FKBP when bound to ascomycin was determined. The secondary structure of FKBP when bound to ascomycin in solution closely resembled the x-ray structure of the FKBP/FK506 complex but differed in some aspects from the structure of uncomplexed FKBP in solution. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
Poly(L -arginine) assumes the α-helix in the presence of the tetrahedral-type anions or some polyanions by forming the “ringed-structure bridge” between guanidinium groups and anions which is stabilized by a pair of hydrogen bonds and electrostatic interaction [Ichimura, S., Mita, K. & Zama, M. (1978) Biopolymers 17 , 2769–2782; Mita, K., Ichimura, S. & Zama, M. (1978) Biopolymers 17 , 2783–2798]. This paper describes the parallel CD studies on the conformational effects on poly (L -homoarginine) of various mono-, di-, polyvalent anions and some polyanions, as well as alcohol and sodium dodecylsulfate. The random coil to α-helix transition of poly(L -homoarginine) occurred only in NaClO4 solution or in the presence of high content of ethanol or methanol. The divalent and polyvalent anions of the tetrahedral type (SO, HPO, and P2O), which are strong α-helix-forming agents for poly(L -arginine), failed to induce the α-helical conformation of poly(L -homoarginine). By complexing with poly(L -glutamic acid) or with polyacrylate, which is also a strong α-helix-forming agent for poly(L -arginine), poly(L -homoarginine) only partially formed the α-helical conformation. Monovalent anions (OH?, Cl?, F?, and H2PO) did not change poly(L -homoarginine) to the α-helix, and in the range of pH 2–11, the polypeptide remained in an unordered conformation. In sodium dodecylsulfate, poly(L -homoarginine) exhibited the remarkably enlarged CD spectrum of an extended conformation, while poly(L -arginine) forms the α-helix by interacting with the agent. Thus poly(L -homoarginine), compared with poly(L -arginine), has a much lower ability to form the α-helical conformation by interacting with anions. The stronger hydrophobicity of homoarginine residue in comparison with the arginine residue would provide unfavorable conditions to maintain the α-helical conformation.  相似文献   

17.
Molecular mechanics calculations have been used to determine the preferred physical association sites of the known alkylating agent dimethyl aziridinium ion (Az+) and a CH prototype test probe with B-form, tetrameric DNA sequences. Electrostatic interactions are most important in determining these preferential physical association sites. In turn, the intermolecular energy minima depend on the charge distribution assigned to the DNA sequence. However, for three reported DNA charge distributions, only two distinct sets of energy minima were obtained for the CH-like ion interacting with (G-C)4, (A-T)4, and [(G-C)·(A-T)]2 deoxyribonucleic acids. These minima correspond to physical association geometries in which the CH-like ion is near known alkylation sites. The results of the Az+ … [(G-C)·(A-T)]2 interaction are virtually identical to those found for the CH-like ion. Aqueous solvation energetics have little effect on the physical association of Az+ with [(G-C)·(A-T)]2.  相似文献   

18.
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   

19.
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   

20.
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   

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