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.
We have measured the VCD of polytyrosine in the amide I and II regions in dimethyl sulfoxide (DMSO) and in 80:20 mixtures of DMSO with trifluoroethanol, trifluoroacetic acid (TFA), and dichloroacetic acid and in 50:50 mixtures of DMSO and trimethyl phosphate (TMP). Additionally, VCD was obtained for deuterated polytyrosine in DMSO and DMSO:D2O, DMSO:TFA(d1), and DMSO:TMP mixtures as before. Amide A VCD was obtained in DMSO and DMSO:TMP mixtures. In the pure solvent, VCD of an opposite sign was seen as compared with that seen in the mixtures. The latter were characteristic in sign pattern and shape of right-handed α-helices for poly(L -tyrosine). The pure polytyrosine:DMSO results are similar to those of polylysine:D2O at neutral pH and poly(β-benzyl-aspartate) in DMSO and may be characteristic of random-coil VCD.  相似文献   

2.
This article reports on both (1) the precision and capability of a computerized multidimensional spectrophotometric system recently developed in our laboratory and (2) the high-resolution study of the helix–coil transition of poly(L-glutamic acid)[poly(Glu)], especially with regard to the discovery of an overlooked transition which is attributable to order–disorder rearrangement of the poly(Glu) side chain in the α-helical conformation. This study was made possible by the high performance of the system used. The simultaneous and continuous measurement of the circular dichroism, the absorbance and light-scattering intensity, and the pH titration curve of poly(Glu) in aqueous salt solution was carried out under continuous scanning of pH ranging from 8 to 2. Besides the well-known random coil to α-helix transition that occurs at about pH 5.5, a highly cooperative transition, which is indicated as a small but definite step in several spectral dimensions, is observed for the first time at pH 4.3. The transition is ascribed to an order–disorder conversion of the side chain on the α-helix backbone.  相似文献   

3.
The size and shape parameters of poly(L -tyrosine) in dimethyl formamide were investigated with fractionated samples of different molecular weight by small-angle X-ray scattering. The molecular weight, the radius of gyration of the molecule as a whole, the radius of gyration of the cross section, the mass per unit length, and the length of helix molecule were determined. The molecular conformations proposed by Applequist and Pao for poly(L -tyrosine) were compared with the experimental results obtained. It was concluded that poly(L -tyrosine) exists in a form of the right-handed α-helix in dimethyl formamide.  相似文献   

4.
High-molecular-weight poly(0,0′-dicarbobenzoxy-L -β-3,4-dihydroxyphenyl-α-alanine) was prepared by the N-carboxyanhydride method. From the results obtained by a study of the optical rotation, nuclear magnetic resonance, and solution infrared absorption, the conformation of poly(0,0′-dicarbobenzoxy-L -β-3,4-dihydroxyphenyl-α-alanine) depended greatly on the solvent taking a right-handed helix with [θ]225 = ?13,600 ~ ?18,900 in alkyl halides, a left-handed helix with [θ]228 = 22,100 ~ 24,800 in cyclic ethers or trimethylphosphate, and a random coil structure in dichloroacetic acid, trifluoroacetic acid, or hexafluoroacetone sesquihydrate. The polypeptide underwent a right-handed helix-coil transition in chloroform/dichloroacetic acid (or trifluoroacetic acid) mixed solvents and a left-handed helix-coil transition in dioxane/dichloroacetic acid (or trifluoroacetic acid) mixed solvents. The results were compared with those of poly(0-carbobenzoxy-L -tyrosine).  相似文献   

5.
γ-(l)-Menthyl L - and D -glutamates were prepared by a fusion reaction of N-phthalyl-L - and D -glutamic anhydrides with l-menthol, followed by hydrazinolysis. The monomers were polymerized to poly[γ-(l)-menthyl L - and D -glutamates] by the N-carboxyanhydride method. These polymers were soluble in many organic solvents, such as ethyl ether, chloroform, tetrahydrofuran, and n-hexane. From the results obtained by a study of the infrared absorption spectra, the x-ray photographs, the optical rotatory dispersions and the circular dichroisms, poly[γ-(l)-menthyl L -glutamate] was found to be a right-handed α-helix in the solid state and in solution. Similarly, poly[γ-(l)-menthyl D -glutamate] was a left-handed α-helix. The helix-coil transition of these polymers was observed in the vicinity of 40% dichloroacetic acid in a chloroform–dichloroacetic acid mixture.  相似文献   

6.
From the results of 13C-nmr measurement of poly(β-benzyl-L -aspartate) and its model compounds in dimethyl sulphoxide/deuterated chloroform mixtures, it was found that the side chain of poly(β-benzyl-L -aspartate) is solvated by dimethyl sulphoxide in the region more than dimethyl sulphoxide 20% (v/v), where the backbone maintains the α-helix. The chemical shift differences in the benzyl group carbons of poly(γ-benzyl-L -glutamate) (trifluoroacetic acid/deuterated chloroform) accompanied by the helix-coil transition, originate from the interaction between the ester group of the side chain and trifluoroacetic acid. The chemical shift difference in the ester carbon is similar. On the other hand, the chemical shift differences of the side-chain carbons in the alkyl portion (Cβ, Cγ) originate not only from the interaction between the ester group of the side chain and trifluoroacetic acid, but also from some other unknown factors. The chemical shift differences of the side-chain carbons of poly(β-benzyl-L -aspartate) originate from the interaction between the ester group of the side chain and trifluoroacetic acid.  相似文献   

7.
Y Suzuki  Y Inoue  R Chùjò 《Biopolymers》1975,14(6):1223-1230
The helix–coil conformational transition undergone by poly(γ-benzyl-L -glutamate) in solutions of trifluoroacetic acid and deuterated chloroform was studied by proton and carbon-13 nmr. The results indicate that in the case of the solvent-induced helix–coil transition, the side chain assumes a helical conformation before the backbone. In the thermally induced helix–coil transition, the results indicate the existence of an intermediate state, which is between the α-helix and random coil and is free from intramolecular hydrogen bonding.  相似文献   

8.
Nature of amino acid side chain and alpha-helix stability.   总被引:1,自引:0,他引:1  
In order to investigate the ability of neutral amino acids to support the α-helix conformation, the coil–helix transition of poly(L -lysine) and of lysine copolymers with these amino acids was studied in water/methanol using circular dichroism. The transtions were recorded at constant pH adding buffer to the methanol/water mixtures. With poly(L -lysine), experiments were performed at several constant pH's; the transition midpoint on the water (methanol) concentration scale was found to depend strongly upon pH; the helix stability region is shifted towards higher water concentrations, when the pH is increased. Copolymers of lysine and several neutral amino acids revealed the same effect in that increasing amounts of, for example, norleucine also shifted the transition midpoint to higher water concentrations. A series of copolymers containing L -lysine as the host and different hydrophobic amino acids were synthesized and the helix–coil transition in water/methanol was observed at constant pH. Different copolymers of equal composition showed significant differences with respect to the nature of the amino acid incorporated into polylysine. From these studies an α-helix-philic scale (in decreasing order): Leu, Nle, Ile, Ala, Phe, Val, Gly is deduced and discussed; the results obtained were compared with those of different procedures.  相似文献   

9.
The conformational transition of poly-L -tyrosine in 0.1M KCl was investigated by ORD and infrared spectroscopy, potentiometric titration, and sedimentation velocity experiments. It is shown that the fully ordered conformer is obtained by slow titration of the random coil with 0.1N HCl at 25°C. The charge-induced transition, at variance with other poly-α-amino acids, is completed in a narrow range of α. An aggregation process was detected both by potentiometric titration and sedimentation velocity. The polyamino acid aggregates around α = 0.7 at 25°C when the conformational transition is almost complete. Infrared spectra, in the region of the amide I band (1650 cm?1) showed that the transition is a random coil → antiparallel β one. Evidence exists that the form is of the intramolecular type. The foregoing interpretations of ORD and CD spectra in terms of the α-helix conformation are discussed.  相似文献   

10.
On the basis of the two-state model of a polyelectrolyte solution, the ion concentration in the polymer domain has been calculated by using the spherical Poisson–Boltzmann equation. The ion accumulation in the neighboring of the polyion influences, on different time scales, various electrical properties of the solution, in particular the low-frequency electrical conductivity and the high-frequency dielectric dispersion. These predictions have been compared with recent dielectric measurements on poly (L -glutamic acid) aqueous solutions during the conformational transition from the α-helix to random coil, and a satisfactory agreement has been found. This finding suggests that counterion distribution plays a different role in determining the electrical properties of charged polymer solutions, causing a electrophoretic contribution of the polymer domain to the electrical conductivity and influencing the high-frequency dielectric dispersion. © 1993 John Wiley & Sons, Inc.  相似文献   

11.
Poly-β-benzyl-L -aspartate (poly[Asp(OBzl)]) forms either a lefthanded α-helix, β-sheet, ω-helix, or random coil under appropriate conditions. In this paper the Raman spectra of the above poly[Asp(OBzl)] conformations are compared. The Raman active amide I line shifts from 1663 cm?1 to 1679 cm?1 upon thermal conversion of poly[Asp(OBzl)] from the α-helical to β-sheet conformation while an intense line appearing at 890 cm?1 in the spectrum of the α-helix decreases in intensity. The 890 cm?1 line also displays weak intensity when the polymer is dissolved in chloroform–dichloroacetic acid solution and therefore is converted to the random coil. This line probably arises from a skeletal vibration and is expected to be conformationally sensitive. Similar behavior in the intensity of skeletal vibrations is discussed for other polypeptides undergoing conformational transitions. The Raman spectra of two cross-β-sheet copolypeptides, poly(Ala-Gly) and poly(Ser-Gly), are examined. These sequential polypeptides are model compounds for the crystalline regions of Bombyx mori silk fibroin which forms an extensive β-sheet structure. The amide I, III, and skeletal vibrations appeared in the Raman spectra of these polypeptides at the frequencies and intensities associated with β-sheet homopolypeptides. Since the sequential copolypeptides are intermediate in complexity between the homopolypeptides and the proteins, these results indicate that Raman structure–frequency correlations obtained from homopolypeptide studies can now be applied to protein spectra with greater confidence. The perturbation scheme developed by Krimm and Abe for explaining the frequency splitting of the amide I vibrations in β-sheet polyglycine is applied to poly(L -valine), poly-(Ala-Gly), poly(Ser-Gly), and poly[Asp(OBzl)]. The value of the “unperturbed” frequency, V0, for poly[Asp(OBzl)] was significantly greater than the corresponding values for the other polypeptides. A structural origin for this difference may be displacement of adjacent hydrogen-bonded chains relative to the standard β-sheet conformation.  相似文献   

12.
A synthetic copolymer of L -glutamic acid and L -tyrosine (23:1) with molecular weight 17,000 was examined conformationally as a function of pH, using circular dichroism, difference spectrophotometry, fluorescence, potentiometic titration, and high-resolution nuclear magnetic resonance (220 MHz). A water-dioxan mixture (2:1) was used to avoid complications due to aggregation (which was shown by infrared spectroscopy to lead to the formation of β-structures). In the α-helical form the tyrosine residues generate a sizeable negative Cotton effect in the near ultraviolet; this is a consequence of perturbation of the chromophores by the helix, and not of tyrosine-tyrosine interactions, which are known to give rise in the right-handed α-helical state to positive Cotton effects. The pH profile of this Cotton effect is different from that of the peptide Cotton effects, which reflect the helix-random coil equilibrium. The data are interpreted in terms of preferential breakdown of the α-helix in the neighborhood of the tyrosine residues. An ultraviolet difference spectrum in the tyrosine absorption bands is generated at the low pH extreme of the conformational transition, the absorbance change being largely complete at a pH at which the other optical parameters have only begun to change. A possible explanation is the formation of a hydrogen bond between the phenolic hydroxyl and a carboxylate, the pK of which is lowered by the hydrogen bonding. An alternative explanation is the freezing of side-chain rotations at a pH below the onset of the helix-random coil transition, when the degree of side chain ionization approaches zero. Some support for the latter scheme comes from the splitting of side-chain methylene proton resonances, indicating partial immobilization, as well as small changes in chemical shift of tyrosine ring protons in the pH (or pD) region in which the difference spectrum appears.  相似文献   

13.
Binding of an azo dye, 4′-dimethyl amino azo benzene-4-carboxylic acid (DAAC) to poly(L -lysine) (PLL) in basic aqueous solutions at 20°C has been studied. The azo dye was found to bind to PLL when its side-chain amino groups are in the uncharged state. This was found to be a cooperative phenomenon, and the binding constant and cooperativity factor have been evaluated. The binding of the dye was found to result in a conformational transition of PLL from the α-helix to the β-sheet, which in turn helps in increased dye binding.  相似文献   

14.
Glycoprotein ovalbumin is an important protein to study helix/sheet transitions as it possess almost equal amount of α-helix and β-sheet. Conformational changes on ovalbumin at various concentrations of glyoxal, ethylene glycol (EG) and polyethylene glycol-400 (PEG-400) were investigated by fluorescence spectroscopy, circular dichroism, attenuated total reflection Fourier transform infra red spectroscopy, 8-anilino-1-naphthalenesulfonic acid and thioflavin T assay. A partially folded state of ovalbumin at 50 % v/v glyoxal was detected that preceded the onset of the aggregation process at the maximum concentration (90 % v/v) of this aldehyde. Aggregates of ovalbumin in the presence EG and PEG-400 were deduced at 70 and 80 % v/v respectively. Maximum aggregation of ovalbumin was observed at 80 % v/v PEG-400, followed by 70 % v/v EG and 90 % v/v glyoxal. Our study confirms that protein aggregation is influenced by the chemistry of organic solvent used thus follows an order of solvent effectiveness (PEG > EG > glyoxal) in inducing the transition. These results provide valuable information on the mechanisms involved in the pathogenesis of some conformational diseases. The α-helix to β-sheet conversion is a diagnostic feature of protein aggregation and has been considered as a general characteristic of amyloid fibrillogenesis in vitro.  相似文献   

15.
H Yamamoto  T Hayakawa  J T Yang 《Biopolymers》1974,13(6):1117-1125
Poly(Nδ-carbobenzoxy, Nδ-benzyl-L -ornithine) (PCBLO) was prepared by the standard NCA method. PCBLO was converted into poly(Nδ-benzyl-L -ornithine) (PBLO) through decarbobenzoxylation with hydrogen bromide. The monomer Nδ-benzyl-L -ornithine was synthesized by reacting L -ornithine with benzaldehyde, followed by hydrogenation. The conformation of the two polypeptides was studied by optical rotatory dispersion and circular dichroism. PCBLO forms a right-handed helix in helix-promoting solvents. In mixed solvents of chloroform and dichloroacetic acid (DCA) it undergoes a sharp helix–coil transition at 12% (v/v) DCA at 25°C, as compared with 36% for poly(Nδ-carbobenzoxy-L -ornithine) (PCLO). Like PCLO, the helix–coil transition is “inverse,” that is, high temperature favors the helical form. PBLO is soluble in water at pH below 7 and has a “coiled” conformation. In 88% (v/v) 1-propanol above pH (apparent) 9.6 it is completely helical. In 50% 1-propanol the transition pH (apparent) is about 7.4; this compares with a pHtr of about 10 for poly-L -ornithine in the same solvent.  相似文献   

16.
CD spectra have been obtained for poly(L -glutamic acid) and poly(L -aspartic acid) as functions of temperature and concentration of cationic detergents. Dodecylammonium chloride induces a coil–helix transition in fully ionized poly(L -glutamic acid). The interaction of the monomeric detergent with the polypeptide is responsible for the conformational transition. The detergent concentration required to produce the transition is independent of temperature. The CD of fully ionized poly(L -aspartic acid) is nearly unaffected by dodecylammonium chloride, in marked contrast to the situation found with poly(L -glutamic acid). However, these results do not imply that dodecylammonium chloride interacts differently with aspartyl and glutamyl residues. The observed results can be accounted for by the well-known fact that the glutamyl residue has a higher helix-forming tendency that the aspartyl residue. Cetyltrimethylammonium chloride destabilizes the helical form of poly(L -glutamic acid). This detergent presents an exception to the usual ability of ionic detergents to promote formation of ordered structures in oppositely charged homopolypeptides.  相似文献   

17.
A Kagemotto  R Fujishiro 《Biopolymers》1968,6(12):1753-1758
By combining the heat of dilution of a poly(γ-benzyl L -glutamate)–dichloroacetic acid–1, 2-dichloroethane solution with the corresponding heat of mixing of two solvents, the integrated heat of the coil–helix transition of poly(γ-benzyl L -glutamate) in the solution was estimated to be about 750 cal/mole.  相似文献   

18.
T Hayakawa  H Yamamoto  N Aoto 《Biopolymers》1972,11(1):185-196
Several β,N-alkyl L -asparagines were prepared from the phthalyl and benzyloxycarbonyl derivatives. High-molecular-weight poly-β,N-benzyl L -asparagine and poly-β,N-(1)-phenethyl L -asparagine were prepared from the corresponding N-carboxyanhy-drides. From the results obtained by a study of the infrared absorption spectra and the optical rotatory dispersion, poly-β-N-benzyl L -asparagine was found to be a random coil structure in dichloroacetic acid and the optical rotatory dispersion curves gradually changed into the left-handed α-helix structure when chloroform was added to the solution. The coil-to-helix transition was observed in the vicinity of 20% dichloroacetic acid in a dichloroacetic acid-chloroform mixture. Poly-β,N-(d), (l), and (d + l, 1:1)
  • 1 (d + l, 1:1): mixed polymer containing the same weighed poly-β,N-(d) and (l)-(1)-phenethyl L -asparagines.
  • -(1)-phenethyl L -asparagines showed a nearly constant specific rotation in the dichloroacetic acidchloroform solvent system. Poly-β,N-(dl)-(1)-phenethyl L -asparagine caused a gradual folding of the helix at dichloroacetic acid content of less than 20%.  相似文献   

    19.
    Poly(L -lysine) of various molecular weights between 2700 and 475,000 was spin-labeled. From the electron spin resonance spectra, the degree of freedom of the nitroxide was determined by calculation of the rotational correlation time as the poly(L -lysine) underwent the pH-induced random coil to α-helix conformational transition. In general, the rotational correlation time of the nitroxide increased as the pH was increased, indicating a more restricted environment for the spin label when poly(L -lysine) is deprotonated. For the high-molecular-weight poly(L -lysine) this corresponds to the formation of the α-helix and indicates that the side chain–side chain interaction and decreased segmental motion of the backbone (slightly) restricts the motion of the spin label. For the 2700-molecular-weight poly(L -lysine), previously shown not to assume a helical conformation at high pH, the increase in the rotational correlation time of the spin label indicates that the side chain–side chain interaction takes place after deprotonation but without helix formation. This may indicate that helix formation per se is not needed to produce the observed effect even with the high-molecular-weight polymers. The rotational correlation time of the spin label at a particular pH did not depend on the molecular weight of the poly(L -lysine) over the 200-fold range of molecular weights. This indicates that the rotational correlation time reflects the rotational mobility of the spin label in a localized environment and not the rotational diffusion of the entire macromolecule.  相似文献   

    20.
    Circular dichroism (CD) measurements were carried out on various copolymers of L -tryptophan and γ-ethyl L -glutamate in ethylene glycol monomethyl ether as the solvent. On increasing the L -tryptophan content of the copolymers a gradual change in the CD spectra was observed. The typical spectrum of the right-handedα-helix becomes more and more evident as the L -tryptophan content decreases. On the basis of these results we assumed that no conformational transition occurs on proceeding from pure poly (γ-ethyl L -glutamate) to pure poly-L -tryptophan in ethylene glycol monomethyl ether: therefore the conformation of poly-L -tryptophan should be that of a right-handed α-helix. Moreover we observed that the change in the CD spectra of the copolymers is gradual but not linear on increasing the tryptophan content. The deviations from linearity were attributed to interactions among side-chain chromophores whose contributions to the optical activity are not simply additive. An x-ray analysis carried out on oriented films of poly-L -tryptophan casted from solutions of the polymer in dimethylformamide shows conclusively that the solid-state conformation of the polymer is that, of an α-helix.  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