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.
T Matsumoto  A Teramoto 《Biopolymers》1974,13(7):1347-1356
The Zimm–Bragg parameters s and σ were determined for poly(γ-benzyl L -glutamate) (PBLG) in m-cresol and in dimethylformamide (DMF) from ORD data as a function of molecular weight. It was found that, within the temperature range between 10 and 55°C and on the average, s = 1.61 ± 0.1 and √σ = 0.04 ± 0.01 in m-cresol and s = 1.65 ± 0.05 and √σ = 0.045 ± 0.015 in DMF. The values of s in m-cresol decreased with increasing temperature, while the values of σ in the same solvent increased. This result for s suggests that PBLG in m-cresol will undergo a thermal helix–coil transition of normal type. The parameters in DMF showed no appreciable trend to vary with temperature. Aside from the difference between the two solvents, our results are consistent with existing data for various conformation-dependent properties such as light-scattering radius, intrinsic viscosity, and dipole moment, each indicating that the polypeptide chain has some flexibility in helicogenic solvents.  相似文献   

2.
The course of the reversible helix formation of poly(γ-benzyl L -glutamate) (PBG) dissolved in a mixture of dichloroacetic acid (DCA) and 1,2-dichloroethane (EDC) was followed by measuring the heat capacity and the optical rotation of the system through the transition region. The results of these measurements indicate that the transition enthalpy ΔH the transition temperature Tc, and the Zimm-Bragg parameter σ depend considerably on the PBG concentration as well as on the composition of the solvent. For the standard state of infinite dilution, however, a linear extrapolation of the measured ΔH if values results in a standard value ΔH° = 950 cal./mole, independent of the solvent composition. The results of the calorimetric measurements are discussed in relationship to changes in optical rotation. Some peculiarities in the measured thermodynamic and optical properties in solutions with relatively high content of dichloroacetic acid are reported.  相似文献   

3.
The effects of deuteration and of changes in solvent composition on the thermo dynamics of the helix–coil transition have been studied by calorimetric and optical measurements in the poly-γ-benzyl-L -glutamate–dichloroacetic acid–1,2-dichloro-ethanc system. For a given solvent composition, deuteration of the polypeptide and of the acid lowers the transition temperature Tc, while an increase in the volume fraction of acid in the solvent raises Tc. A rise in Tc is accompanied by a decrease in both the van't Hoff and the calorimetric heats of transition, but at different rates. The result is a temperature dependency in the Zimm-Bragg cooperation parameter σ. Possible causes of this result and its implications are discussed.  相似文献   

4.
The helix-coil transition temperature Tc of poly(γ-benzyl L -glutamate) in binary solvent mixtures of dichloroacetic acid and 1,4-dichlorobutane, 1-chlorooctane, or 1-chlorododecane have been measured. A treatment is presented with which the transition enthalpy can be calculated from the observed dependence of Tc on solvent composition. Results are compared with previously obtained calorimetric data. The underlying assumptions of the calculation are discussed.  相似文献   

5.
In a study of A–B type block copolymers of γ benzyl-L -gultamate and β-benzyl-L -aspartate use has been made of the observations: (1) that for poly aspartate esters the chemical shifts of the α-CH and NH resonances are sensitive to the helix sense, (2) that in both helical and random coil conformations the same resonances of poly-γ-benzyl-L -glutamate are well separated from those of poly aspartates. Since the sense of poly-β-benzyl-L -aspartate is very sensitive to the inclusion of γ-benzyl-L -glutamate residues, the degree of overlap between the blocks can be studied by monitoring the helix sense of the aspartate. The ability of the NMR method to make separate observation of the two blocks removes the necessity of relying on an overall ORD parameter such as b0. The copolymers studied include those having lefthanded, righthanded, and mixed-sense aspartate, corresponding to differing degrees of overlap.  相似文献   

6.
L Ferrara  P A Temussi 《Biopolymers》1973,12(7):1451-1458
The coil to helix conformational transitions undergone by poly-γ-benzyl-L-glutamate in solutions of haloacetic acids and various cosolvents were studied by means of proton magnetic resonance. The results indicate a very small solvent dependence of the α-CH Helix–coil chemical shift difference. The helical stabilities of PBLG in different solvent mixtures were interpreted in terms of modifications of the “solvent structure.”  相似文献   

7.
The Poland–Fixman–Freire formalism was adapted for modeling of calorimetric DNA melting profiles, and applied to plasmid pBR 322 and long random sequences. We studied the influence of the difference (HGC?HAT) between the helix‐coil transition enthalpies of AT and GC base pairs on the calorimetric melting profile and on normalized calorimetric melting profile. A strong alteration of DNA calorimetrical profile with HGC?HAT was demonstrated. In contrast, there is a relatively slight change in the normalized profiles and in corresponding ordinary (optical) normalized differential melting curves (DMCs). For fixed HGC?HAT, the average relative deviation (S) between DMC and normalized calorimetric profile, and the difference between their melting temperatures (Tcal?Tm) are weakly dependent on peculiarities of the multipeak fine structure of DMCs. At the same time, both the deviation S and difference (Tcal?Tm) enlarge with the temperature melting range of the helix‐coil transition. It is shown that the local deviation between DMC and normalized calorimetric profile increases in regions of narrow peaks distant from the melting temperature.  相似文献   

8.
D Puett  A Ciferri 《Biopolymers》1971,10(3):547-564
We have studied the effect of polypeptide concentration on the helix–coil transition of poly(γ-benzyl L -glutamate) (PBLG) in both dichloroacetic acid (DCA) and DCA–chloroform (CHF) mixtures. In agreement with other reports, we find the van't Hoff transition enthalpy to be strongly dependent on PBLG concentration. Also, an apparent effect of polypeptide concentration was noted on the transition temperature; however, corrections for finite PBLG concentration on the mole fraction of DCA seem to remove this effect. In order to explain our data, as well as some calorimetric data in the literature, we consider the transition free energy and enthalpy as a sum of three partial terms. These represent the thermodynamic parameters associated with: (1) conformational changes of the polypeptide, e.g. formation or disruption of intramolecular hydrogen bonds; (2) binding by the strong acid to the nonhelical segments of the polypeptide; (3) an overall (weak) interaction of the polypeptide with the nonbound solvent giving rise to dilution parameters that are dependent on the polypeptide conformation. The latter effect is generally ignored, since it is assumed that solvent interactions, other than specific binding, are similar for both the helical and the nonhelical conformation. Striking effects of water (small amounts) and solution aging on the formation of PBLG helices was observed. Water, as expected, acts as a helicogenic solvent when combined with DCA. The processes occurring during solution aging are not known, although the net effect is to stabilize the helical conformation. Finally, we present some rather unique thermally induced transitions of concentrated PBLG (about 200 mg/ml) in DCA. At low temperatures the soluble randomly coiled conformation is present. Heating produces first an isotropic gel, followed at higher temperatures by an isotropic solution consisting of about 70% α-helicity.  相似文献   

9.
A Teramoto  T Norisuye 《Biopolymers》1972,11(8):1693-1700
For helix-coil transitions of polypeptide in binary mixtures consisting of helix-forming solvent and coil solvent, the transition enthalpy ΔH(T,x) has been found to depend significantly on temperature (T) and solvent composition (x). For such systems, calorimetric measurements may yield some averages of ΔH(T,x) which are no longer amenable to direct comparison with ΔH itself. Theoretical equations relating calorimetric data to ΔH(T,x) are derived and tested favorably with experimental data. It is demonstrated that the transition enthaply from heat capacity measurements is approximately equal to ΔHcfm, while those from heat of dilution and heat of solution measurements are equal to ΔHc. Here ΔHc denotes the value of ΔH at the transition point and fm represents the maximum helical content attained in a thermally induced transition. The discrepancies among calorimetric data are also discussed.  相似文献   

10.
In this paper two points are considered: the methods of evaluating the helical content θ and the calculation of the parameters of the transition from experimental data and its interpretation. The parameter ΔH obtained is in good agreement with the calorimetric one and v is found to be independent of temperature and solvent and in agreement with the ordinarily accepted value for poly(γ-benzyl-L -glutamate). The different methods of estimating θ are discussed for both polypeptides.  相似文献   

11.
Poly(ortho-, meta-, and para-γ-nitrobenzyl-L -glutamates) were studied by circular dichroism (CD) and optical rotatory dispersion (ORD) in two helicogenic solvents, hexafluoroisopropanol (HFIP) and dichloroethane (EDC), and two non-helicogenic solvents, dichloracetic acid (DCA) and trifluoroacetic acid (TFA). The corresponding glutamates were also studied in DCA and TFA. The symmetric nitrobenzylic chromophore is optically active when the polymers are in solution in DCA and TFA. The corresponding glutamates are also optically active under the same conditions. Thus, it was not possible to explain the origin of the optical activity of the side-chain chromophore when the polymer is in solution in a helicogenic solvent. Nevertheless, from a side-chain dichroic band, a helix–coil transition curve was determined and the stability of each poly(γ-nitrobenzyl-L -glutamate) given; this stability depends on the position of the nitro substituent on the aromatic ring.  相似文献   

12.
The various polymer–acid solvation possibilities occuring in the helix–coil transition process of polypeptides with polar side chains were systematically analyzed by infrared spectroscopy. The following samples have been considered: poly-γ-benzyl-L -glutamate (PBLG), alternating poly-γ-benzyl-D ,L -glutamate (PBD-LG), and poly-β-benzyl-L -as-partate (PBLA). The behavior of the amide A, I, II, and νC?O ester absorptions of each polymer dissolved in trifluoroacetic acid–chloroform mixtures was studied in depth. The classical assumptions concerning the interaction between a polypeptide and a proton donor solvent are discussed. This interaction was previously proposed in a theoretical model of helix–coil transition. For PBLG, the spectral characteristics of the cooperative transition are evidenced by the amide bands. These bands also show main chain–acid hydrogen bonding (I) Quantitative analysis of phenomenon (I) was performed in order to localize the “binding sites” of the polymer. In agreement with the theory, only the complexation of peptide units belonging to random coil and terminal helical regions were observed. However, in contrast to the theory in which the association constants KCO and KNH of these residues are generally kept equal, the present results have shown that the main binding site is the carbonyl group (KNH ? 0 or « KCO ). The behavior of the polar side chains of these polypeptides were analyzed during the transition. Similarly to the peptide backbone, they bind the acid by hydrogen bonding (II) Furthermore, this association is more important when the side chains are localized in the coiled regions than in the helical ones. This result suggests, by analogy with the main chain behavior, that the helix–coil transition theory should take into account two more association constants for polar side chains, namely k1 for the helical regions and k2 > k1 for the coiled ones.  相似文献   

13.
Thermodynamics of the B to Z transition in poly(dGdC)   总被引:1,自引:0,他引:1  
The thermodynamics of the B to Z transition in poly(dGdC) was examined by differential scanning calorimetry, temperature-dependent absorbance spectroscopy, and CD spectroscopy. In a buffer containing 1 mM Na cacodylate, 1 mM MgCl2, pH 6.3, the B to Z transition is centered at 76.4°C, and is characterized by ΔHcal = 2.02 kcal (mol base pair)?1 and a cooperative unit of 150 base pairs (bp). The tm of this transition is independent of both polynucleotide and Mg2+ concentrations. A second transition, with ΔHcal = 2.90 cal (mol bp)?1, follows the B to Z conversion, the tm of which is dependent upon both the polynucleotide and the Mg2+ concentrations. Turbidity changes are concomitant with the second transition, indicative of DNA aggregation. CD spectra recorded at a temperature above the second transition are similar to those reported for ψ(–)-DNA. Both the B to Z transition and the aggregation reaction are fully and rapidly reversible in calorimetric experiments. The helix to coil transition under these solution conditions is centered at 126°C, and is characterized by ΔHcal = 12.4 kcal (mol bp)?1 and a cooperative unit of 290 bp. In 5 mM MgCl2, a single transition is seen centered at 75.5°C, characterized by ΔHcal = 2.82 kcal (mol bp)?1 and a cooperative unit of 430 bp. This transition is not readily reversible in calorimetric experiments. Changes in turbidity are coincident with the transition, and CD spectra at a temperature just above the transition are characteristic of ψ(–)-DNA. A transition at 124.9°C is seen under these solution conditions, with ΔHcal = 10.0 kcal (mol bp)?1 and which requires a complex three-step reaction mechanism to approximate the experimental excess heat capacity curve. Our results provide a direct measure of the thermodynamics of the B to Z transition, and indicate that Z-DNA is an intermediate in the formation of the ψ-(–) aggregate under these solution conditions.  相似文献   

14.
Thermally induced helix–coil transitions of myosin rod, light meromyosin, and tropomyosin were studied by optical rotatory dispersion (ORD). Fractional helicity was calculated from both the Moffitt-Yang parameter, b0, and the corrected mean residue rotation [m′] at 231.4 nm. Between 3 and 30°C, [m′] increases linearly with a slope of 59/°C, whereas b0 is virtually constant, indicating apparently different thermal melting behavior. Poly(L -lysine) and poly(L -glutamic acid) in their helical forms and myoglobin also show a nearly linear temperature dependence of [m′]231.4. Muscle proteins in 6M guanidine hydrochloride and the random-coil forms of the homopolymers exhibit temperature-dependent values of [m′]231.4 and b0. We conclude from these observations that ORD properties of both α-helices and random-coil polypeptides have significant intrinsic temperature dependencies. A new method of estimating fractional helicity as a function of temperature is proposed.  相似文献   

15.
Y C Fu  H V Wart  H A Scheraga 《Biopolymers》1976,15(9):1795-1813
The enthalpy change associated with the isothermal pH-induced uncharged coil-to-helix transition ΔHh° in poly(L -ornithine) in 0.1 N KCl has been determnined calorimetrically to be ?1530 ± 210 and ?1270 ± 530 cal/mol at 10° and 25°C, respectively. Titration data provided information about the state of charge of the polymer in the calorimetric experiments, and optical rotatory dispersion data about its conformation. In order to compute ΔHh°, the observed calorimetric heat was corrected for the heat of breaking the sample cell, the heat of dilution of HCl, the heat of neutralization of the OH? ion, and the heat of ionization of the δ-amino group in the random coil. The latter was obtained from similar calorimetric measurements on poly(D ,L -ornithine). Since it was discovered that poly(L -ornithine) undergoes chain cleavage at high pH, the calorimetric measurements were carried out under conditions where no degradation occurred. From the thermally induced uncharged helix–coil transition curve for poly(L -ornithine) at pH 11.68 in 0.1 N KCl in the 0°–40°C region, the transition temperature Ttr and the quantity (?θh/?T)Ttr have been obtained. From these values, together with the measured values of ΔHh°, the changes in the standard free energy ΔGh° and entropy ΔGh°, associated with the uncharged coil-to-helix transition at 10°C have been calculated to be ?33 cal/mol and ?5.3 cal/mol deg, respectively. The value of the Zimm–Bragg helix–coil stability constant σ has been calculated to be 1.4 × 10?2 and the value of s calculated to be 1.06 at 10°C, and between 0.60 and 0.92 at 25°C.  相似文献   

16.
A Yaron  N Tal  A Berger 《Biopolymers》1972,11(12):2461-2481
The sequence-ordered copolymer poly-(Lys-Ala-Ala) was synthesized by polycondensation of the N-hydroxysuccinimide ester of ε,Z-Lys-Ala-Ala and deprotection of the polymerization product. A fraction of molecular weight 13,000 obtained by ion-exchange chromatography was investigated. The polymer is freely soluble in water at all pH values, and is completely digested by trypsin and elastase. From CD and ORD data it was concluded that in water at 1°C the ionized form (at pH 6.5) of the polymer is helical. On heating, helix-coil transition curves were obtained with a midpoint, Tm, depending on salt concentration. In salt-free water Tm = 12.3°C and in 0.2M NaCl Tm = 28.5°C. Adding MeOH, causes an increase in the helical content of the polymer (half helicity at 20% MeOH, without salt, at 29°C). Guanidine·HCl was shown to decrease the helicity. At 1°C half helicity. The nonionized polymer helix is more stable (Tm~90°C). At the high pH, at 60°C, when concentration of the polymer is higher than 1.9 × 10-2M, a precipitate is formed which redissolves on cooling with the original helicity. This does not occur in the presence of 50% MeOH. By comparison with polylysine it was concluded that replacing two-thirds of the lysine residues in polylysine by alanine leads to a polymer forming a more stable α-helix, when fully ionized. This is essentially due to the diminished coulombic repulsion. Uncharged lysine residues are comparable to alanine residues in their helix-forming tendency since the sequential polymer as well as one-third ionized polylysine are helical to approximately the same extent at room temperature.  相似文献   

17.
K Okita  A Teramoto  H Fujita 《Biopolymers》1970,9(6):717-738
A new procedure for evaluating u and σ characterizing σ-helix-forming polypeptides in solution was derived from Nagai's theory for the helix–coil transition of such polymers. Here u is the activity for helix formation from random coil, and σ is the helix initiation parameter. The necessary data are the helical content fN at fixed solvent and temperature as a function of N, where N is the degree of polymerization of the polypeptide sample. Such data were obtained from ORD measurements on a number of fractionated samples of poly-N5-(3-hydroxypropyl)-L -glutamine (PHPG) in mixtures of water and methanol covering the complete range of composition and at various termperatures (5–40°C). When analyzed in terms of the proposed procedure, they yielded values of σ which were in the range (3.2 ± 0.6) × 10?4, substantially independent of solvent composition and temperature. These values were much larger than those obtained recently for σ of poly(β-benzyl-L -aspartate) in m-cresol and in a mixture of chloroform and DCA. The data for [η] and s0 (limiting sedimentation coefficient) as functions of molecular weight indicated that the molecular shape of PHPG in pure methanol is essentially rodlike, whereas that in pure water is not entirely randomly coiled but rather may be regarded as an interrupted helix. These indications were consistent with the results from ORD measurements. When plotted against the corresponding values of fN, the values of [η] and [s0] for PHPG in mixtures of water and methanol of various compositions and temperatures formed smooth composite curves, and we attributed these phenomena to the fact that σ of PHPG was nearly constant under these solvent conditions. Here [s0] stands for a reduced limiting sedimentation coefficient which is equal to the inverse friction factor of the solute molecule.  相似文献   

18.
The helix–random-coil transition process for poly-γ-benzyl-L -glutamate (PBG) in solvent mixtures trifluoroacetic acid/deuterochloroform (TFA/CDCl3) at different temperatures has been studied by nmr. The chemical shift behavior of the α-CH resonances of the peptide chain and of the TFA carboxylic protons is reported.  相似文献   

19.
The helix–coil transition of poly-N5-(2-hydroxyethyl)L -glutamine (PHEG) in aqueous isopropanol was examined by means of optical rotatory dispersion (ORD) and intrinsic viscosity [η] measurements. The Zimm–Bragg parameters σ and s for the transition were determined from the ORD data as a function of molecular weight. It was found that the transition was characterized by a relatively low cooperativity; the values of \documentclass{article}\pagestyle{empty}\begin{document}$ \sqrt \sigma $\end{document} were in the range from 0.039 to 0.066, depending on the solvent composition. These σ values are much larger than those reported for other polypeptide–solvent systems. The transition enthalpy was negative and its magnitude varied with the solvent composition, with a maximum of 620 cal/mol at 40 wt% isopropanol. The curve of [η] versus helical content for a high-molecular-weight sample exhibited a very broad minimum, and this behavior was attributed to the low cooperativity of the transition.  相似文献   

20.
The collagen-like peptides (L -Pro-L -Pro-Gly)n and (L -Pro-L -Hyp-Gly)n with n = 5 and 10, were examined in terms of their triple helix ? coil transitions in aqueous and nonaqueous solvents. The peptides were soluble in 1,2-propanediol containing 3% acetic acid and they were found to form triple-helical structures in this solvent system. The water content of the solvent system and the amount of water bound to the peptides were assayed by equilibrating the solvent with molecular sieves and carrying out Karl Fischer titrations on the solvent phase. After the solvent was dehydrated, much less than one molecule of water per tripeptide unit was bound to the peptides. Since the peptides remained in a triple-helical conformation, the results indicated that water was not an essential component of the triple-helical structure. Comparison of peptides with the same chain length demonstrated that the presence of hydroxyproline increased the thermal stability of the triple helix even under anhydrous conditions. The results, therefore, did not support recent hypotheses that hydroxyproline stabilizes the triple helix of collagen and collagen-like peptides by a specific interaction with water molecules. Analysis of the thermal transition curves in several solvent systems showed that although the peptides containing hydroxyproline had tm values which were 18.6° to 32.7°C higher, the effect of hydroxyproline on ΔG was only 0.1 to 0.3 kcal per tripeptide unit at 25°C. The results suggested, therefore, that the influence of hydroxyproline on helical stability may be explained by intrinsic effects such as dipole–dipole interactions or by changes in the solvation of the peptides by alcohol, acetic acid, and water. A direct calorimetric measurement of the transition enthalpy for (L -Pro-L -Pro-Gly)n in 3% or 10% acetic acid gave a value of ?1.84 kcal per tripeptide unit for the coil-to-helix transition. From the value for enthalpy and from data on the effects of different chain lengths on the thermal transition, it was calculated that the apparent free energy for nucleation was +5 kcal/mol at 25°C (apparent nucleation parameter = 2 × 10?4 M?2). The value was dependent on solvent and on chemical modification of end groups.  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