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 609 毫秒
1.
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   

2.
N Murai  S Sugai 《Biopolymers》1974,13(6):1161-1171
In order to study the effect of side-chain length on the conformation of polypeptides, conformational changes of various ionic polypeptides with various lengths of side chain, poly-Nε-glutaryl-L -lysine (PGL), poly-Nδ-glutaryl-L -ornithine (PGO), poly-Nε-succinyl-L -lysine (PSL), and poly-Nδ-succinyl-L -ornithine (PGO), were investigated by ORD, potentiometric titration, and dilatometric measurements in aqueous solution. The results of optical rotation and potentiometric titration measurements indicate strongly that the α-helix stability increases in the sequence PSO < PSL ~ PGO < PGL, which corresponds to increased side-chain length. The volume change associated with the helix–coil transition also increased in the above sequence. This series of polymers seems to be more hydrophobic compared with poly-L -glutamic acid or poly-L -lysine, as suggested from the values of enthalpy and entropy changes for coil–helix transitions.  相似文献   

3.
H Yamamoto  T Hayakawa 《Biopolymers》1972,11(6):1259-1268
The solvent-and pH-induced conformational changes are examined in order to investigate the influence of benzyl group. Polymer was prepared via N?-benzyloxycarbonyl, N?-benzyl-Nα-carboxy-L -lysine anhydride. The resulting poly (N?-benzyloxycarbonyl, N?-benzyl-L -lysine) was obtained in high yield and had a high molecular weight. The protected polymer was removed into poly (N?-benzyl-L -lysine) by treating it with hydrogen bromide. From the results of the ORD and CD, the protected polymer has a righthanded α-helix, showing [m′]233 = –10,300, [θ]220 = –27,600 and [θ]207 = –25,100 in dioxane. The breakdown of the helical conformation is found to occur at 8% dichloroacetic acid in chloroform-dichloroacetic acid mixture. In the pH range 3.35–6.90, poly (N?-benzyl-L -lysine) is in a random coil structure. In the pH range 7.50–13.0, the polypeptide has a right-handed α-helix structure; [m′]233 = –12,000, [0]220 = –27,200, and [0]207 = –27,000. In comparison with poly-L -lysine, the coil-to-helix transition is observed at lower pH range in 50% n-propanol. Above pH 8 by heating, the α ? β transition of poly (N?-benzyl-L -lysine) is not observed in an aqueous media.  相似文献   

4.
K Suto  H Noda 《Biopolymers》1974,13(11):2391-2404
Measurements of the molecular weight of (Pro-Pro-Gly)n and (Pro-Pro-Gly)n(Ala-Pro-Gly)m(Pro-Pro-Gly)n, which were synthesized by the solid-phase method, revealed that they formed a trimer in an aqueous solution, and dissociated into single-stranded chains on warming. Accompanying the transition, a large decrease of optical rotation was observed, like the collagen–gelatin transition. The shape of the trimeric molecule was rodlike, and the dimensions were 12 Å in diameter and 2.8 Å per residue in length, regardless of the length of Ala-Pro-Gly sequences in a peptide chain. The data indicate that both Pro-Pro-Gly sequences and Ala-Pro-Gly sequences from the triple-helical structure similar to that of collagen in aqueous solution. All optical rotational dispersion (ORD) curves of solutions of the peptides were represented by a single-term Drude equation, and the Drude constant λc was 200 nm for all peptides regardless of the length of Ala-Pro-Gly sequences. The resemblance between the helical structure formed by Pro-Pro-Gly sequences and that by Ala-Pro-Gly sequences was also suggested by the formation of the hybrid triple helix from two kinds of peptide chains with different lengths of Ala-Pro-Gly sequences.  相似文献   

5.
Conformational studies of poly-L-alanine in water   总被引:5,自引:0,他引:5  
The conformational properties of poly-L -alanine have been examined in aqueous solutions in order to investigate the influence of hydrophobic interactions on the helix–random coil transition. Since water is a poor solvent for poly-L -alanine, water-soluble copolymers of the type (D , L -lysine)m–(L alanine)n-(D , L -lysine)m, having 10, 160, 450, and 1000 alanyl residues, respectively, in the central block, were synthezised. The optical rotatory dispersion of the samples was investigated in the range 190–500 mμ, and the rotation at 231 mμ was related to the α-helix content, θH, of the alanine section. In salt-free solutions, at neutral pH, the three large polymers show high θH values, which are greatly reduced when the temperature is increased from 5 to 80°C. No helicity was observed for the small (n = 10) polymer. By applying the Lifson-Roig theory, the following parameters were obtained for the transition of a residue from a coil to a helical state: ν = 0.012; ΔH = ?190 ± 40 cal./mole; ΔS = ?0.55 ± 0.12 e.u. Since ΔH and ΔS differ from the values expected for a process involving only the formation of a hydrogen bond, and in a manner predicted by theories for the influence of hydrophobic bonding on helix stability, it is concluded that a hydrophobic interaction is also involved. In the presence of salt (0.2M NaCl), or when the ε-amino groups of the lysyl residues are not protonated (pH = 12), the helical form of the two large polymers (n = 450 and n = 1000) is more stable than in water. Since the electrostatic repulsion between the lysine end blocks is greatly reduced under these conditions, the alanine helical sections fold back on themselves, and this conformation is stabilized by interchain hydrophobia bonds. This structure was predicted by the theory for the equilibrium between such interacting helices, non-interacting helices, and the random coil.  相似文献   

6.
Iwao Satake  Jen Tsi Yang 《Biopolymers》1975,14(9):1841-1846
The conformational phase diagram of poly(L -lysine) (4.6 × 10?4 M, residue) in sodium dodecyl sulfate (1.6 × 10?2 M) solution was constructed from circular dichroism results at various temperatures and pH's. Poly(L -lysine)–sodium dodecyl sulfate complexes undergo a β–helix transition upon raising the pH of the solution. The transition pH tends to shift downward at elevated temperatures. No helix–β transition can be detected for poly(L -lysine) in sodium dodecyl sulfate solution (pH > 11) even after 1-hr heating at 70°C. This is in marked contrast with uncharged poly(L -lysine) solution without sodium dodecyl sulfate, which is converted into the β-form upon mild heating of the solution above 50°C.  相似文献   

7.
In order to test theoretical predictions that poly-L -valine can exist in an α-helical conformation, water-soluble block copolymers of L -valine and D , L -lysine were prepared. By carrying out the synthesis on a resin support (with the use of N-carboxyanhydrides) contamination of the individual blocks by any unreacted monomer from the previous block was avoided. A single glycine residue was incorporated at the C-terminus of the chain for use in amino acid analyses. Using optical rotatory dispersion and circular dichroism criteria, about 50% of the short valine block of (D , L -lysine HCl)18-(L -valine)15-(D , L -lysine-HCl)16-glycine was found to be in the right-handed α-helical conformation in 98% aqueous methanol, in water, the polymer appears to be a dimer, with the valine block being involved in the formation of an intermolecular β-structure.  相似文献   

8.
The effect of the number of methylene groups in the side chains on the conformation of polypeptides is assessed for three poly(L -lysine) homologs with R = –(CH2)nNH2. Circular dichroism studies show a pH-induced helix–coil transition in 0.05 M KCl with midpoints at 9.6, 9.0, and 8.7 for n = 5, 6, and 7, respectively, as compared with 10.1 for (Lys)x (n = 4). Homologs with n = 6 and 7 could be partially helical even when the side groups are fully charged (with n = 7, the compound is highly aggregated above pH 9.1). Thus, the longer the number of methylene groups the more stable is the helical conformation of these homologs. Potentiometric titration of the n = 5 homolog gives a ΔG° of ?310 cal/mol (residue) for the uncharged coil-to-helix transition at 25°C. The corresponding ΔH° and ΔS° are ?1740 cal/mol (residue) and ?4.8 e.u./mol (residue). Unlike (Lys)x, the uncharged helix-to-β transition is slow and incomplete even after heating at 80°C for 1 hr. Addition of methanol enhances the helical formation in neutral solution with midpoints at 72, 52, and 27% methanol (v/v) for n = 5, 6, and 7, respectively [cf. 88% for (Lys)x]. Addition of sodium dodecyl sulfate induces a coil-to-helix transition for all three homologs in contrast with the β form of (Lys)x under similar conditions.  相似文献   

9.
M J Grourke  J H Gibbs 《Biopolymers》1971,10(5):795-808
The helix–coil transitions of aqueous solutions of poly-α-L -lysine (PLL), poly-α-L -ornithine (PLO), and poly(α,γ-L -diaminobutyric acid) (PLDBA) have been investigated as functions of pH at 25°C and of temperature at pH 11.75, where these polymers are uncharged; in the cases of the latter two polyamino acids, the transitions have also been studied as functions of apparent pH in methanol-water solution (50/50 by volume). The helix stability of the polypeptides is shown to be a direct function of the number of methylene groups on the side chain. From an analysis of potentiometric titration data, we find that the difference between the helix stability of PLL and that of PLO is due to a difference of about 1 e.u. in the ΔS° of the transition. Combining the “melting curves” obtained from optical rotatory dispersion studies with the potentiometric titration data permits evaluation of the initiation parameter Z (or 1/σ½) of the statistical mechanical theories for these transitions. The value obtained for Z in the case of uncharged aqueous PLO is ca. 35.  相似文献   

10.
A 13C-nmr study of the salt-induced helix–coil transition of the basic polypeptides poly(L -lysine) [(Lys)n], poly(L -arginine) [(Arg)n], and poly (L -ornithine) [(Orn)n] was performed to serve as a reference of the helical portion of histones and other proteins. As is the case with pH-induced helix–coil transition, the downfield displacement of the Cα and carbonyl carbon signals are observed in the helical state. The upfield shift of the Cβ signals, on the other hand, is noted in the salt-induced transition. Regardless of the differences in the side chains and also the salts used, very similar helix-induced chemical shifts are obtained for (Lys)n and (Arg)n. However, the displacement of the Cα, Cβ, and carbonyl carbons of (Orn)n in the presence of 4M NaClO4 is found to be almost 50% of that of (Lys)n and (Arg)n. This is explained by the fact that the maximum helical content is about 50%, consistent with the ORD result. Further, the motion of the backbone and side chains of the helical from was estimated by measuring the spin-lattice relaxation time (T1), nuclear Overhauser enhancement (NOE), and line width. In the case of (Lys)n, the motion of the side chains is charged very little in comparison with that of the random coil. Indicating that the aggregation of the salt-induced helix is small in contrast to that of the pH-induced helix. For (Arg)n, however, the precipitate of the helical polymers is mainly due to aggregation.  相似文献   

11.
A new representation of protein structure is obtained by the angular coordinate transformations ηi = (?i+1i)/2 and ξi = ?i+1i with careful mathematical attention to the cyclical boundary conditions of all of the variables involved. From published ?-ψ data it is possible to obtain a new η-ξ plot. As the angle ξi is varied from – 180° through 0° to + 180° in this plot, the local helicity of the polypeptide chain changes continuously and contiguously without sudden reversals in handedness. The variable, ηi, gives the torsional position of the ith peptide group. Some peptide groups in proteins, such as the second peptide residue in a type II β-turn, are nonhydrogen-bonded and can undergo considerable torsional oscillation. In such cases the η angle should be represented by a line whose length reflects the allowed dynamical variations in the peptide torsional position. Certain peptide residues in proteins may be able to undergo a complete torsional rotation of 360°. Such residues would be represented on the η-ξ plot as a straight line across the plot parallel to the abscissa. Other examples of the possible usefulness of this plot are also given.  相似文献   

12.
A strong Cotton effect, which practically govern the sign of the optical rotation at 589 nm ([M]d), was studied in phenyl 1-thio-α (and β)-d-glycopyranosides with our new chiroptical technique. The proposal optical rotatory dispersion (ORD) method, with calculations based on a one term Drude equation, showed the presence of a strong Cotton effect at 200–210 nm. Circular dichroism (CD), with accumulation technique, also gave the same Cotton effect. Agreement in these two methods suggests the usefulness of the proposed ORD calculation method. The rotational strengths and the signs were shown to reflect the anomeric configurations and conformations (α-anomer gave positive and β-anomer gave negative signs; axial gave strong and equatorial gave weak bands). This result is an extension of the ring oxygen helicity rule of alkyl and alkyl thioglycosides to phenyl 1-thioglycopyranosides, and probably to other aromatic glycosides.  相似文献   

13.
A polypeptide having the repealing sequence (Tyr-Ala-Glu)n was synthesized by the polymerization of the N-hydroxysuccinimide ester of O-benzyl-L -tyrosyl-L -alanyl-γ-benzyl-L -glutamate, followed by the removal of the benzyl groups by means of hydrogen bromide. The main fraction obtained on gel filtration had an average molecular weight of over 60, 000, corresponding to over 500 amino acid residues per polypcptide chain. The polymer is soluble in water above pH 5.5, and precipitates on lowering the pH. The x-ray powder photographs show features of an α-helical structure. The dependence of the ultraviolet absorption spectrum, the optical rotatory dispersion, and the fluorescence of poly(Tyr-Ala-Glu) on pH, in salt-free as well as in salt-containing aqueous solutions, was compared with the corresponding properties of a copolymer containing equimolar proportions of tyrosine, alanine, and glutamic acid in a random sequence. From these measurements it was concluded that poly(Tyr-Ala-Glu ) has a helical con formation at low pH and a random coil conformation at high pH, the transition taking place at pH 6 in the absence of salt and pH II in the presence of salt. Thus, in the range pH 7 to l0. random coil-to-helix transition can be achieved by merely increasing the ionic strength. A model is proposed for the structure of the helical poly peptide which accounts for the Stability of the helical conformation by assuming hydrogen bonding between the carboxylate group of the ith glutamic acid residue and the hydroxyl group of the (i + 4 )th tyrosine residue. The complex ORD of helical poly(Tyr-Ala-Glu) is explained as being due to a superposition of the ORD of an α-helix and that of a regular array of phenolic ehroniopholes originating from the immobilization of the aromatic rings in the specific structure of the polymer.  相似文献   

14.
The properties of apomyoglobin were examined in aqueous solutions and various helix- and random-coil-forming solvents by solvent perturbation, optical rotation, circular dichroism, and viscosity measurements. The solvent perturbation data obtained in neutral aqueous solutions suggest 25–40% exposure of the two tryptophyl residues and 50–60% exposure of the three tyrosyls. The estimates of burial of these groups are in the ranges expected for myoglobin based on its X-ray structure. In the helicogenic alcohols, methanol, ethanol, 2-chloroethanol, trifluoroethanol, and 1-propyl alcohol, as well as in acidic solutions, 8 M urea and 6M guanidine hydrochloride, essentially all the tryptophyl and tyrosyl residues are found to be exposed to solvent based on this method. Analysis of the ORD and CD data indicates that in the alcohols the α-helix content of apomyoglobin has in most cases changed from 58–59% to about 80–95%. Analysis of the intrinsic viscosity data based on the equations of Simha and Kirkwood and Auer indicates that the polypeptide chain in these solvents has the dimensions of fully extended α-helical rods, with lengths of 221–251 Å and mean diameters of 12.8–13.6 Å. It is concluded that apomyoglobin in the various alcohols must have an extended but somewhat irregular rodlike structure, having a few bend or irregular sequences between the α-helical segments due largely to the presence of the four proline residues, 37, 88, 100, and 120 in the amino acid sequence.  相似文献   

15.
P Y Chou  H A Scheraga 《Biopolymers》1971,10(4):657-680
The heat ΔH° for converting an uncharged lysine residue from a coil to an α-helical state in poly-L -lysine in 0.1N KCl has been determined calorimetrically to be ?1200 cal/mole at both 15°C and 25°C. Essentially the same value has been obtained for the conversion of an uncharged residue from a coil to a β-pleated sheet state. 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 ΔH°, the observed Calorimetric heat was corrected for the heat of breaking the sample cell, the heal of dilution of HCl, the heat of neutralization of OH? ion, and the heat of ionization of the ε-amino group in the random coil. The latter was obtained from similar Calorimetric measurements on poly-D ,L -lysine, which was shown to be a good model for the random coil form of poly-L -lysine. The measured transition heat was ~0.7 cal., which is only 7% of the total heat liberated when a 40 ml solution of 0.25% w/v poly-L -lysine is brought, from pH 11 to pH 7; nevertheless it could be determined with a precision of ±8%. The conformation of poly-L -lysine at pH 11 appears to be completely helical at 15°C, but a mixture of 90% α-helix, 5% β form, and 5% coil at 25°C. Since ΔH° ~ 0 for the α ? β conversion, the polymer behaves like one of 95% α-helix and 5% coil in the calorimeter at 25°C. At neutral pH, poly-L -lysine is an extended coil, like poly-D ,L -lysine.  相似文献   

16.
The conformation of chromatographically pure block oligopeptides (L -lysyl)m-(L -alanyl)n- L -tyrosyl-(L -alanyl)n-(L -lysyl)m with n = 3 and m = 6 or 3 is investigated. By circular dichroism it is shown that these peptides may exhibit a partially α-helical structure depending upon pH, ionic strength, solvent, and temprerature. An attempt is made to describe the helical content of these small peptides by utilizing the data obtained on high-molecular-weight poly(L -lysine). By measurement of the quantum yield and the decays of the peptides fluorescence, it is shown that, in aqueous solution, at neutral pH, the fluorescence of the peptides is quenched by interactions with the peptide carbonyl groups. The decays are multiexponential, which shows the presence of several conformations of the phenolic chromophore relative to the peptide chain. The addition of methanol, which induced the helix formation, decreases the quenching of the fluorescence and the multiexponential character of the decays. In presence of sodium hydroxide, which further increases the helical content of the peptides, a dynamic quenching occured that can be attributed to interactions between the phenol hydroxyl group of tyrosine (ith residue) and the ε-amino groups of the (i+4)th and (i -4)th lysyl residues.  相似文献   

17.
The epicuticular hydrocarbons of the larval, pupal and adult stages of the sugarcane borer Diatraea saccharalis Fabricius (Lepidoptera: Crambidae) are analysed. Dramatic changes are observed between the stages studied. Adult hydrocarbons are mostly saturated, with a predominance of 1–4 methyl‐branched straight carbon skeletons of 37–47 atoms; the major components are isomeric mixtures of internally branched trimethylderivatives of C39, C37 and C41 carbon backbones. By contrast, very small amounts of methyl‐branched components are detected in the pupae, although straight chain hydrocarbons of 23–35 carbons are the prevailing structures (70.7 ± 3.4%) with n‐C29 and n‐C27 as the major components. Unsaturated hydrocarbons (29.0 ± 3.5%) of similar chain lengths elute by gas chromatography of epicuticular extracts as complex mixtures of mono‐, di‐ and trienes; with the degree of unsaturation increasing with chain length. This is the first report of very long chain unsaturated hydrocarbons in cuticular extracts of a larval lepidopteran (93.3 ± 0.6% of the lipid components), with chain lengths in the range 37–53 carbons and up to four double bonds; the major component being C49:3, which co‐elutes with C49:4 and C49:2.  相似文献   

18.
A highly sensitive and selective HPLC-MS-MS method was developed for the determination of pramipexole in human plasma. The analytes, pramipexole and BHT-920 (internal standard), were extracted from plasma at basic pH with methyl tert.-butyl ether (MTBE). MTBE was evaporated to dryness and reconstituted in 100 μl of (95:5) methanol-water. Chromatographic separation was achieved on a Zorbax SB-CN column with a mobile phase of (15:5:80) water-0.1 M ammonium acetate—methanol. The analytes were detected utilizing HPLC in conjunction with atmospheric pressure chemical ionization (APCI) tandem mass spectrometry (MSM). The assay was linear in the concentration ranges of 50 to 5000 pg/ml. The analysis of pooled quality controls (150, 750, and 3000 pg/ml) demonstrated excellent precision with relative standard deviations (R.S.D.) (n=18) of 7.2%, 5.3% and 5.2%, respectively. The method is accurate with all intra-day (n=6) and overall (n=18) mean values being less than 11.7% from theoretical.  相似文献   

19.
The hagfishes of the genus Eptatretus (Myxinidae) from southern Africa are known from three poorly studied species: Eptatretus hexatrema, a common species from Namibia and South Africa; Eptatretus profundus, known only from the holotype collected off Cape Point (South Africa); and Eptatretus octatrema, known from two syntypes from the Agulhas Bank (South Africa). Taxonomic, morphological and distributional information about these three species are reviewed and updated based on the examination of additional specimens collected in South African waters. Eptatretus hexatrema differs from all congeners by having six pairs (rarely seven) of gill apertures arranged in a straight line, 3/2 multicusp pattern of teeth, total cusps 44–49, trunk pores 53–60, total pores 93–107, preventral length 45.1–57.4% TL, tail length 11.6–14.3% TL, tail depth 5.7–8.1% TL, and two bilaterally symmetrical nasal-sinus papillae. Eptatretus octatrema differs from all congeners by having usually eight (some specimens with seven) pairs of gill apertures arranged in a straight line, 3/2 multicusp pattern of teeth, 42–46 total cusps, 22–26 prebranchial pores, 63–68 trunk pores, 104–117 total pores, and two bilaterally symmetrical nasal-sinus papillae. Eptatretus profundus differs from all congeners by having five pairs of gill apertures arranged in a straight line, 3/2 multicusp pattern of teeth, total cusps 42–46, prebranchial pores 12–15, branchial pores 4–5, trunk pores 48–52, tail pores 15–17, total pores 81–86, and body depth at PCD 7.0–9.7% TL. An identification key for the hagfishes from southern Africa is provided and the conservation status of E. octatrema, a species considered to be Critically Endangered, is discussed in light of the new findings.  相似文献   

20.
It was recently found that some short peptides (including C- and S-peptide fragments of RNase A) can have considerable helicity in solution, 1–12 which was considered to be surprising. Does the observed helicity require a new explanation, or is it consistent with previous understanding? In this work we show that this helicity is consistent with the physical theory of secondary structure12–19 based on an extension of the conventional Zimm-Bragg model.20 Without any special modifications, this theory explains reasonably well almost all the experimentally observed dependencies of helicity on pH, temperature, and amino acid replacements. We conclude that the observed “general level” of helicity of C- and S-peptides (5–30% at room temperature and 10–50% near 0°C) is “normal” for short peptides consisting mainly of helix-forming and helix-indifferent residues. The helicity is modified by a multitude of weak specific side chain interactions, many of which are taken into account by the present theory;13–19 some discrepancies between the theory and experiment can be explained by weak side-chain-side chain interactions that were neglected. A reasonable coincidence of the theory with experiment suggests that it had been used to investigate the role of local interactions in the formation of α-helical “embryos” in unfolded protein chains.  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