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.
Reversing-pulse electric birefringence (RPEB) was measured for the first time for four fractionated poly(γ-benzyl-L -glutamate), [Glu(OBzl)]n, samples in N,N-dimethylformamide (DMF) at 20°C and at 535 nm. The RPEB signal showed a deep minimum for each sample on reversal of an applied electric field. The profiles of the reverse-transient signal were analyzed by taking into account the polydispersity for the continuous distribution of molecular lengths. The best set of three quantities (lw, lw/ln, (βw)2/2γw), which determine a signal profile, was evaluated for each sample. By combining the experimental data of intrinsic viscosity and RPEB, the diameter of a cylinder, which is assumed for the [Glu(OBzl)]n helix, was found to be 17 Å. The value (βw)2/2γw, which is related to the ratio of weight-average permanent dipole moment μw, and electric polarizability anisotropy Δαw, was found to be in the range of 20–45. This indicates that the former contributes predominantly to electric field orientation, but the latter also should not be ignored. With the three parameters from the reverse portion, the rise and decay curves were regenerated theoretically in excellent agreement with experimental signals.  相似文献   

2.
The reversing-pulse electric birefringence (RPEB) technique was applied to the study of the temperature effect on the electrooptical and hydrodynamic properties of a fractionated [Glu(OBzl)]n sample, which is molecularly dissolved in cyclohexanone. The aim was to develop a standard analytical method for thermal denaturation and temperature-induced conformational transitions. The field-on reverse and steady-state signal, and the field-off decay signal, were measured at 535 nm and at a constant low field strength (ca. 3 kV/cm) over a wide temperature range (5–90°C). The steady-state birefringence and the relaxation time in the decay process were also measured at two constant temperatures (5 and 70°C) over a wide field strength range (E ≤ 20 kV/cm). By the combination of these two different sets of RPEB measurements, the unwanted effect of the high pulse field on polymer conformation at elevated temperatures could be minimized. Together with the density and viscosity of cyclohexanone measured between 5 and 95°C, the following quantities could be evaluated: the weight-average permanent dipole moment and polarizability anisotropy, the reduced optical anisotropy factor (Δg/n), the weight-average length, and the degree of polydispersity. All these quantities, except for Δg/n, were found to be almost independent of temperature (5–90°C) and concentration (1.54–4.27 mM).  相似文献   

3.
Poly(γ-benzyl-L -glutamate) having a terminal dimethylaminoanilide group as an electron donor (D) and a terminal 3,5-dinitrobenzoyl group as an electron acceptor (A) (A-[Glu(OBzl)]n-D) was synthesized by the N-carboxyanhydride method. Polymer samples were fractionated by gel chromatography and their number-average degrees of polymerization n were determined by the absorbances of the terminal chromophores. These polymers in chloroform and dimethylformamide solutions showed a charge-transfer (CT) absorption band around 455 nm, and the fraction of the polymer forming the CT complex was evaluated as a function of the chain length. CT absorption for short chains (n = 5 ~ 20) was attributed to intramolecular CT complex in which the A-[Glu(OBzl)]n-D chain takes cyclic conformations. An optimum chain length for the intramolecular CT was found to be n ? 10, where the [Glu(OBzl)]n chain may most easily bend back to form cyclic conformations. Stronger CT absorption observed for longer chains than n = 20 was shown to be intermolecular, and an intermolecular head-to-tail aggregation was found to be a cause of the strong CT interaction. All helical A-[Glu(OBzl)]n-D chains were found to form the head-to-tail dimers in chloroform solution.  相似文献   

4.
The hydrodynamic properties of α-helical poly(L -glutamic acid), (Glu)n in aqueous solutions and in mixtures of water with organic solvents have been interpreted in terms of the persistence length of the macromolecule. A modification of the method of Vitovskaya and Tsvetkov has been proposed in order to allow a more accurate determination of this parameter. The addition of an organic solvent increases strongly the rigidity of the helical conformation of (Glu)n. A comparison is made with some data of the literature of poly[N5-(3-hydroxy propyl)L -glutamine], [Gln(CH2)3OH]n, and poly(γ-benzyl-L -glutamate), [Glu(OBzl)]n.  相似文献   

5.
Poly(Glu(OBzl)-Gly)n, poly(Glu-Gly)n, poly(Gly)-(Glu(OBzl)-Gly), and poly(Gly-Glu-Gly) were synthesized from the pentachlorophenyl esters of the sequential monomer. Both of the polymers containing free glumatic-acid residues are soluble in water, as is the lower molecular weight fraction of the polytripeptides with the benzyl ester in place. Circular dichroism studies and infrared dichroism studies suggest that the 21 helix is favored for the polydipeptide with removal of the benzyl ester reducing the conformational integrity. The polytripeptide showed evidence of 31 helix in addition to the 21 form, depending on solvent. A rationale for the conformations observed is developed based on the bulkiness of the side-chain residues and conformational stabilization, in certain cases, by hydrophobic interactions between the benzyl ester groups.  相似文献   

6.
Nps-[Glu(OBzl)]6-NHEt has been prepared by coupling Nps-[Glu(OBzl)]2-OH with HCl,H-[Glu(OBzl)]4-NHEt by means of dicyclohexylcarbodiimide. The ir spectra of its nujol mull show that the hexapeptide has the β-structure of antiparallel chains. When it is dissolved in dioxane or ethylene dichloride, the hexapeptide consists of a mixture of the β-form and the solvated σ-form, but the β-form can exist only above a certain critical concentration. The critical concentration is about 0.4g dl?1 in dioxane and 0.08g dl?1 in ethylene dichloride, and the content of β-form increases with increasing concentration above it. The CD of the dioxane and ethylene dichloride solutions shows concentration dependence in visible and uv regions.  相似文献   

7.
C M Deber  P D Adawadkar 《Biopolymers》1979,18(10):2375-2396
We have synthesized and characterized a series of cation-binding cyclic octapeptides which may function as potential ionophoric substances. The materials contain varying degrees of hydrophobic character, which was controlled systematically through the incorporation of N-alkylglycine residues where N-alkyl = methyl, n-hexyl, cyclohexyl, or n-decyl. The peptides reported include cyclo(Phe-Sar-Gly-Sar)2, cyclo(Glu(OBzl)-Sar-Gly-Sar-Glu(OBzl)-Sar-Gly-(N-decyl)Gly), cyclo(Glu(OBzl)-Sar-Gly-(N-decyl)Gly)2, cyclo(Glu(OBzl)-Sar-Gly-(N-hexyl)Gly)2, cyclo(Glu(OBzl)-Sar-Gly-(N-cyclohexyl)Gly)2, and the corresponding free diacid forms of the Glu-containing compounds. Using 13C- and 1H-nmr spectra, we demonstrated that the mixture of cis/trans peptide bond-isomer conformers, characteristic of the free-peptide benzyl esters in solution, was converted to unique C2-symmetric, presumably all-trans conformers on complexation with calcium ions. Cation-transport experiments, using the thick-liquid model of transport in a Pressman cell, established that these compounds transport a variety of cations and that one peptide examined in detail, cyclo(Glu(OBzl)-Sar-Gly-(N-decyl)Gly)2 (selectivity Ca2+ > Na+ > K+ > Mn2+ > Cu2+ > Mg2+ > Co2+ > Zn2+), transports calcium about an order of magnitude more efficiently than magnesium.  相似文献   

8.
Wayne L. Mattice 《Biopolymers》1985,24(12):2231-2242
The intramolecular formation of multiple clusters of interacting helices has been characterized in a homopolymer. The configuration partition function permits the formation of clusters in which the number of interacting helices may be as large as the greatest integer in n/2, where n denotes the number of amino acid residues in the chain. The theoretical formulation has its origin in a recent [Mattice, W. L. & Scheraga, H. A. (1984) Biopolymers 23 , 1701–1724], tractable matrix expression for the configuration partition function for intramolecular antiparallel β-sheet formation. Reassignment of the expression for one of the n(n+3)/2 elements in the sparse statistical weight matrix, along with a simple change in notation, converts that treatment into a matrix formulation of the configuration partition function for a chain containing multiple clusters of interacting antiparallel helices. The five statistical weights used are δ, fl, w, and the Zimm-Bragg σ and s. Each tight bend that connects two interacting helices contributes a factor of δ, fl is used in the weight for larger loops between interacting helices, and w arises from helix–helix interaction. The influence of the helix–helix interaction is well illustrated by two helix–coil transitions in a chain with n = 156 and σ = 0.001. In the absence of helix–helix interaction, the transition occurs by the nucleation and subsequent elongation of a small number of helices. When helix–helix interaction is attractive, the transition can occur by a different mechanism. Formation of a single pair of interacting helices is followed by addition of new helices to the initial cluster. In the latter process, individual helices experience relatively little growth after they are formed.  相似文献   

9.
10.
The electric birefringence of poly(L -glutamic acid) (PLGA) in methanol–water mixtures has been measured by the use of the rectangular pulse technique at 25°C. The permanent dipole moment, the anisotropy of electrical polarizability, and the optical anisotropy factor of PLGA in solution were obtained from the dependence of the steady-state birefringence on the electric field strength. Further, the mean length of PLGA in solution was calculated by a parameter method developed for analyzing the decay curve of electric birefringence. The permanent dipole moment per unit length obtained from these studies was 2.96, 2.48, 2.30, 2.66 D/Å in pure methanol, 10, 30, and 50 vol-% water, respectively. The increase of water content caused the decrease of the mean length and broadened the length distribution of PLGA. These results are discussed in relation to the viscosity and the electrical conductivity of PLGA solutions.  相似文献   

11.
Electric birefringence measurements of suspensions of T3 and T7 bacteriophages in 10?2 M phosphate buffer, pH 6.9, show that there is a difference in their rotational diffusion coefficient. The values corrected to 25°C and water viscosity are D25,w = 4630 ± 130 sec?1 and D25,w = 5290 ± 260 sec?1 for T3 and T7, respectively. The value obtained from shell model calculations (according to Filson and Bloomfield) is D25,w = 4500 ± 600 sec?1. The apparent permanent dipole moments are 4.5 × 10?26 C·m and 1.7 × 10?26 C·m for T3 and T7, respectively. For both phage particles the intrinsic optical anisotropy is +7.2 × 10?3. It is shown that this anisotropy is mainly due to the DNA molecule inside the head of the phage. Its positive value means that there exists an excess orientation of the DNA helix perpendicular to the symmetry axis of the particle. For T7 an unexpectedly large increase of Δns and Ksp occurs at a glycerol concentration of about 30% (v/v). This increase is interpreted as being caused by a change of the shape of the particle and/or a change in the secondary structure of the DNA inside the head of the bacteriophage.  相似文献   

12.
A A Ribeiro  R Saltman  M Goodman 《Biopolymers》1985,24(12):2431-2447
The syntheses of three series of glutamate oligopeptides attached to a macromolecular solubilizing polyoxyethylene (POE) group Boc-[Glu(OMe)]n-OPOE, Ac-[Glu(OMe)]n-OPOE, pGlu-[Glu(OMe)]n?1-OPOE (n ? 1–7) and their various analogs specifically deuterated at individual α-CH positions using the liquid-phase method of peptide synthesis are described. It was shown that stepwise synthesis using the symmetrical anhydride gave homo-oligopeptides that are analytically pure. Fragment condensation methods using DCC-HOBt yield POE-peptides with POE-HOBt impurities but the peptide synthesis may be carried stoichiometrically with smaller quantities of amino acid derivatives. 360 MHz 1H-nmr conformational studies of these homo-oligopeptides in DMSO-d6 are presented. The α-deuterated peptides are shown to allow unequivocal homoligopeptide backbone NH assignments.  相似文献   

13.
Conformations of terminal peptide units of α-helical poly(γ-benzyl-L -glutamate), poly-[Glu(OBzl)], were examined by an induced circular dichroism (CD) of chromophores which were covalently attached to both ends of the chain. In chloroform, where the helical poly-[Glu(OBzl)] exists as a head-to-tail-type dimeric associate, the chromophores showed a strong CD induced by an asymmetric perturbation from the helical structure. The induced CD almost disappeared by an addition of a few percent of dichloroacetic acid, which has been reported as a powerful breaker of the associate. These results are explained by an incorporation of terminal peptide units into a helical conformation in the head-to-tail associate and a local unfolding of the terminal portions by the addition of acid. An induced CD of a charge-transfer complex between the two terminal chromophores was also observed and the structure of the helix–helix junction of the head-to-tail dimer is discussed.  相似文献   

14.
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   

15.
We first calculate the steady-state birefringence, expressed in the form of specific Kerr constant, Ksp, of rigid, bent-rod macromolecules. Equations are derived for Ksp as a function of the geometric and electro-optical parameters. We also consider flexibly hinged rods and evaluate Ksp for them by averaging over the angle between the two arms, ?. Next, we turn to the time decay of the electric birefringence. The decay function for rigid bent rods is a sum of three exponential terms, and a procedure for their calculation is indicated. We observe that single-exponential decays can be found for ? > 90° or ? < 60°, in spite of the high electro-optical and hydrodynamic anisotropy of the macromolecule. Special attention is paid to the case of rods with equal arms.  相似文献   

16.
The conformation and dilute solution properties of (2→1)-β-d-fructan in aqueous solution were studied by gel permeation chromatography, low-angle laser light-scattering photometry, viscometry, small-angle X-ray scattering and electron microscopy. Fractions covering a broad range of weight-average molecular weights (Mw) from 1.49 × 104 to 5.29 × 106 were obtained from a native sample by ultrasonic degradation and fractional precipitation. For Mw < 4 × 104, the intrinsic viscosity [η] varies with Mw0.71, indicating that the fructan chain behaves as a random coil expanded by an excluded-volume effect in this molecular weight region. For Mw > 105, [η] exhibits an unusually weak dependence on Mw and finally becomes almost independent of molecular weight. This behaviour is interpreted in terms of a globular conformation of the high-molecular-weight fructan molecules. Small-angle X-ray-scattering measurements and electron microscopic observations support this interpretation of the values of [η] observed.  相似文献   

17.
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   

18.
The electric birefringence of poly(L -glutamic acid) (PLGA) in dimethylsulfoxide (DMSO)–methanol mixtures has been measured by use of the rectangular pulse technique. The length distribution curve, the mean molecular length, and the mean apparent permanent dipole moment of PLGA in solution have been obtained from the decaycurve and field strength dependence of the steady-state birefringence according to the method developed for analyzing the electric birefringence of a polydisperse system. The length distribution curve exhibits one or two peaks. The length corresponding to a high peak and the mean length of PLGA undergo an abrupt change in the vicinity of 50 to 60 vol % DMSO at 30°C. Moreover, a sharp change of the Moffitt b0 parameter with the solvent composition is observed. These results provide evidence for the existence of a solvent-induced transition from a helical conformation (presumably α-helix) to another helical conformation with shorter length per amino acid residue. Further, the temperature dependence of the length distribution of PLGA in 50 vol % DMSO suggests the existence of a temperature-induced helix ? helix transition.  相似文献   

19.
The fractions obtained from the partially hydrolyzed branched Streptococcus salivarius levan were examined in solution. Sedimentation coefficients, S0, intrinsic viscosities, [η], weight-average molecular weights, M w, and radii of gyration were obtained from sedimentation velocity, viscosity, and light-scattering measurements. Double logarithmic plots of [η] vs M w and S0 vs M w each yielded two linear segments intersecting at M w ≈ 105. Hydrodynamic data suggest that fractions of M w > 105 behave as compact spheres, whereas for M w < 105, the particles are best characterized as linear random coils. Calculations based on theories of random coils and spheres support the above observations.  相似文献   

20.
Synthetic cyclic octapeptides of general structure cyclo[Glu(γOBzl)-Sar-Gly-(N-R)Gly]2 (R = n-hexyl and cyclohexyl) transport calcium ions selectively across organic phases and phospholipid membranes. We have now used proton nmr spectroscopy (360 MHz) to study the solution conformation(s) of their calcium complexes. When Ca(ClO4)2 was added to solutions of these peptides in CDCl3, nmr spectra of the resulting calcium complexes were characteristic of a single C2-symmetric conformer. From a Karplus-Bystrov analysis of vicinal coupling constants in both the peptide backbone and Glu side chain (treated as an ABCCMX spin system), in conjuction with model-building studies, a structure was proposed in which the calcium ion is bound in an octahedral-type complex by the four (coplanar) carbonyl groups of the (all-trans) Glu-Sar and Gly-(N-R)Gly peptide bonds. Occurrence of preferred rotamers about Glu side chain Cα–Cβ bonds indicated that restricted rotation in peptide side chains arises upon calcium binding.  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