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 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   

2.
We studied the interactions of the substitution-inert inversion-labile complexes Fe(bipy) and Fe(phen) [and the inversion-stable complex Ru(bipy)] with DNA. The association of these complexes to DNA is mainly electrostatic, and Fe(phen) shows a more effective binding to DNA than the two bipyridyl complexes, possibly owing to a different binding mode. The interactions are enantioselective, leading to a Pfeiffer shift in the diastereomeric inversion equilibria and an excess of the Δ-enantiomer of Fe(phen) and Fe(bipy), which is directly monitorable through CD. The partition constants for the inversion equilibrium range from 1.3 to 2.0 for Fe(bipy) and Fe(phen), depending on ionic conditions. From flow LD information about the orientation of the complexes on DNA was obtained: it is consistent with a fit of the Δ-enantiomer in the major groove of the right-handed DNA helix. The mechanisms of interaction are discussed against equilibrium, spectroscopic, and kinetic data.  相似文献   

3.
4.
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   

5.
A comparison of the condensation of T4 phage DNA by spermidine and Co(NH3) at pH values between 5.1 and 10.2 has been made using quasielastic light scattering to determine translational diffusion coefficients and Stokes radii. Co(NH3) is more effective than spermidine in causing condensation at all pH, indicating that the differences observed in previous work were not due to pH effects, as might have been inferred from recent theories of intermolecular forces. The DNA particles collapsed with Co(NH3) are smaller than those obtained with spermidine. The hydrodynamic radius of spermide-collapsed structures decreases slightly with increasing pH, while the size of the Co(NH3)collapsed structures is almost independent of pH. These results confirm that there are specific ion effects in DNA condensation by oligocations, in addition to the dominant general polyelectrolyte effects.  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.
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   

8.
The high-resolution nmr spectrum of baker's yeast tRNA, a recently sequenced non-denaturable tRNA, has been compared with the spectra of the native and denatured conformers of the closely related species tRNA. Because of the presence of many common base pairs in the different tRNA's, it is possible to assign most of the low-field resonances to specific secondary-structure base pairs. A comparison of the observed positions of the various resonances with those predicted by a semiempirical ring-current shift theory shows a root-mean-square deviation of 0.14, 0.11, and 0.12 ppm for tRNA (native), and tRNA (denatured), respectively. These results support the ring-current shift theory currently used to interpret the low-field nmr spectra of the tRNA molecules. Differences between the predicted and observed positions of some resonances provide new evidence for higher order effects such as shifts from second nearest neighbors, anomalous shifts exerted by G·U base pairs, and tertiary-structure effects. A model that was previously proposed for the denatured conformer of tRNA is also supported.  相似文献   

9.
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   

10.
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   

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.
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   

13.
The binding of tris(2,2′-bipyridyl)ruthenium(II) cations [Ru(bpy)] with single- and double-stranded (ss and ds) DNA, and the polynucleotides poly(A), poly(C), poly(G), poly(I), poly(I) · poly(C), and poly(U), was studied in aqueous solution. Steady-state electrical conductivity measurements with the polynucleotides, ssDNA, and dsDNA reveal that approximately three nucleotides offer one binding site. This may be compared with the ratio [nucleotide]/[Mg2+] of 2.4 : 1 for dsDNA. After laser excitation (353 nm), the luminescence of Ru(bpy) bound to nucleic acids shows two decay components. The contribution of the fast component, which is interpreted as resulting from quenching processes of the absorbed ruthenium complex, exhibits a maximum with increasing [nucleotide]/[Ru(bpy)] at a ratio of about three to one. Bound Ru(bpy) can be released from the strand by addition of NaClO4 [half-concentration: 2.5 and ≤ 10 mM for poly(U) and dsDNA, respectively].  相似文献   

14.
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   

15.
G Spach  Y Trudelle  F Heitz 《Biopolymers》1983,22(1):403-407
The β- and β-helical structures do not appear consistent with the structural data and ion-transport properties of (Ala-Ala-Gly) or (Leu-Ser-Leu-Gly) oligomers. Oligoalanine derivatives also give rise to current fluctuations in bilayer lipid membranes. Bundles of molecules may explain the behavior of these various peptides in membranes.  相似文献   

16.
The starch–iodine blue complex formation does not involve negatively charged iodine species like I, I, or I; rather, neutral iodine units are involved. The heat of reaction is determined to be about ?110 kJ for every mole of I-I unit in the amylose helix, which suggests that the dissociation of I2 (binding energy 149 kJ/mol) does not take place during the complex formation. Quantum mechanical (INDO CI) calculations indicate that the linear as well as nonlinear polyiodine units, I6, with interiodine distance of 3.0 Å are responsible for characteristic absorbance bands of the starch–iodine complex. Based on our previous article [(1989) J. Polym. Sci. A 27 , 4161] and the present studies we identify (C6H10O5)16.5I6 to be the polymeric unit responsible for the characteristic blue color of the complex.  相似文献   

17.
Conformation and folding in histones H1 and H5   总被引:1,自引:0,他引:1  
Denatured histones H1 and H5 can be readily refolded on salt addition. Their digestion by trypsin leads to limit peptides of about 80 residues having the same nmr and CD spectra as those of the intact parent histones. Scanning microcalorimetry shows that (1) the folded structures of H1 and H5 are located entirely in their limit peptides; (2) both have values of the specific denaturation enthalpy typical for small globular proteins; and that (3) both exhibit a classic “2-state” transition (ΔH = ΔH). The heat-denaturation profiles of H5 measured using intrinsic and extrinsic Cotton effect and side-chain nmr peaks do not coincide at all. Only the intrinsic Cotton effects give a Tm and ΔH close to that from microcalorimetry. We conclude that these proteins exhibit large-scale side-chain motions that precede the macroscopic cooperative transition.  相似文献   

18.
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   

19.
Weili Qian  Samuel Krimm 《Biopolymers》1994,34(10):1377-1394
Infrared and Raman spectra have been obtained of crystalline glutathione and its deuterated derivative and interpreted by normal mode analysis. The force field consisted of our empirical force fields for the peptide group and NH and CO end groups, plus our ab initio force fields for the CH2SH and CH2COOH moieties. Observed bands are reproduced with an average error of 5 cm?1, demonstrating that the vibrational spectrum of such a complex molecule can be understood in great depth. © 1994 John Wiley & Sons, Inc.  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