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.
Poly(His-Ala-Glu) and poly(Lys-Ala-Glu) were examined by viscosity and potentiometric titration. These measurements were interpreted in terms of the hydrodynamic size of the above sequential polypeptides. Effects of polymer, size and concentration, and solution-salt concentration were demonstrated. Although the sequential polypeptides generally behave like polyampholytes, they do demonstrate some differences. These differences my be attributed to the ability of ionized side chains three residues apart to repel themselves, in the order His < Glu < Lys.  相似文献   

2.
Poly(Lys(Cbz)-Ala-Glu(OBzl)) was prepared by the self-condensation of Lys(Cbz)-Ala-Glu(OBzl)-ONSu in dimethylformamide. After deprotection of the side chains, the product was subjected to Sephadex G-50 chromatography. The molecular weight of unfractionated and fractionated poly(Lys-Ala-Glu) was calculated from a calibrated Sephadex G-50 column, spectrophotometrically from Dnp-(Lys-Ala-Glu), equilibrium centrifugation, and viscosity measurements. Approximately 21% of the unfractionated material was polymeric with the remaining 79% being cyclic and monomeric material. Treatment of polymer hydrolysate with L -amino acid and D -amino acid oxidase indicated poly(Lys-Ala-Glu) to be optically pure. The apparent pKa's of the two ionizable groups were 4.1 and 9.7.  相似文献   

3.
M. Rinaudo  A. Domard 《Biopolymers》1975,14(10):2035-2048
The preparation and characterization of α-L -glutamic acid oligomers with degree of polymerization (DP) up to 12 are described. The preparation of polymers with low DP corresponding to various A/I ratios (where A and I are monomer and initiator concentrations, respectively) with end groups blocked is given. The conditions of the fractionation, which separates the different oligomers by ion-exchange chromatography, are discussed. Finally, the isolation from salt solutions of the pure acidic form is given. Each polymer obtained for a given A/I is characterized at the end of the polymerization by its molecular-weight distribution. The average DP values calculated are compared to the A/I values; agreement is very good. Potentiometric behaviour during neutralization is obtained as a function of the degree of polymerization and the elaboration of the polyelectrolytic phenomenon is discussed.  相似文献   

4.
The synthesis of the tetrapeptide benzyloxycarbonyl(α-aminoisobutyryl-L -prolyl)2-methyl ester (Z-(Aib-Pro)2-OMe) and an analysis of its conformation in solution and the solid state are reported. Stepwise synthesis using dicyclohexylcarbodiimide leads to racemization at Pro(2). Evidence for the presence of diastereomeric tetrapeptides is obtained from 270-MHz1H-nmr and 67.89-MHz 13C-nmr. The all-L tetrapeptide is obtained by fractional crystallization from ethyl acetate. The NH of Aib(3) is shown to be involved in an intramo-lecular hydrogen bond by variable-temperature 1H-nmr and the solvent dependence of NH chemical shifts. The results are consistent with a β-turn conformation with Aib(1) and Pro(2) at the corners stabilized by a 4 → 1 hydrogen bond. The molecule crystallizes in the space group P212121, with a = 8.839, b = 14.938, and c = 22.015 Å. The structure has been refined to an R value of 0.051. The peptide backbone is all-trans, and a 4 → 1 hydrogen bond, between the CO group of the urethane moiety and Aib(3) NH, is observed. Aib(1) and Pro(2) occupy the corner positions of a type I β-turn with ? = ?55.4°, Ψ = ?31.3° for Aib(1) and ? = ?71.6°, Ψ = ?38° for Pro(2). The tertiary amide unit linking Pro(2) and Aib(3) is significantly distorted from planarity (Δω = 14.3°).  相似文献   

5.
M. Morcellet  C. Loucheux 《Biopolymers》1980,19(12):2177-2190
The preferential interaction of sodium poly(α-L -glutamate) and poly(α-L -glutamic acid) with the solvent components in water/2-chloroethanol mixtures has been determined using density-increment measurements. The degree of preferential interaction was deduced from the density increments at constant molality of 2-chloroethanol and at constant chemical potential of 2-chloroethanol. Sodium poly(α-L -glutamate) and poly(α-L -glutamic acid) are both preferentially hydrated in the whole range of solvent composition. A dehydration process occurs during the 2-chloroethanol-induced coil-to-helix transition of sodium poly(α-L -glutamate). This dehydration process was attributed to the release of some moles of water from the neighborhood of the peptide bond during the nucleation of the helix. After the conformational transition, sodium poly(α-L -glutamate) is solvated by one 2-chloroethanol molecule. The location of water and 2-chloroethanol molecules in the different parts of the residue (more polar and less polar portions) is also discussed.  相似文献   

6.
The synthesis of poly(ε-L -lysine) is described. This is a poly(ε-amino acid) in which the ε-amino group of lysine is condensed with the α-carboxyl group to produce a chain backbone that is a variant of the usual one seen in proteins and the side chain is the α-amino group. Conformational studies of poly(ε-L -lysine) and its t-butyloxycarbonyl derivative suggest the likelihood of a chain order that is formally similar to the antiparallel pleated-sheet conformation of proteins.  相似文献   

7.
X-ray diffraction and energy-minimization results are reported for poly(γ-phenethyl-L -glutamate). Orthorhombic unit-cell parameters of drawn fibers are a = 15.4 Å, b = 26.6 Å, c = 54.4 Å. Atomic coordinates are derived for an α-helix peptide conformation that corresponds to a calculated side-chain internal energy minimum. The side-chain conformation correlates well with the electron density projection; the side chains wrap around the α-helical main chain with the phenethyl ester group directed toward the N-terminus. The para-axis of the benzene ring is inclined at an angle nearly nearly normal to the helix axis. The x-ray structure factors calculated for this model, when compared to the 10 observed structure factors, yield a crystallographic reliability index of R = 0.23.  相似文献   

8.
By means of fluorescence, absorption, and acid-base titrations, it has been shows that there is a one-to-one correspondence between free carboxylates and bound acridine orange in the dye-polyacid complex. Contrary to expections, stacking coefficients for the dye were found to be virtually the same on binding to the helical or coiled polyacid, indicating a strong similarity in the binding sites for both forms of the polyelectrolyte.  相似文献   

9.
Poly(L -histidyl-L -alanyl-α-L -glutamic acid) has been prepared in order to test the acid–base catalytic ability of a carboxyl-imidazole hydrogen-bonded system. Two different blocked histidyl-alanyl-glutamic acid monomers were used in the polymerization step. The imidazole ring was blocked with either a dinitrophenyl or a t-butyloxycarbonyl group. The γ-carboxyl of glutamic acid was protected as the benzyl ester. Both the coupling reactions and the polymerization step were via the N-hydroxysuccinimide active ester method. Thiolysis removed the dinitrophenyl group, while hydrogen bromide removed the t-butyloxycarbonyl and the benzyl groups. The water-soluble unblocked polymers obtained were fractionated on Sephadex G-50 or Bio-Gel P30. Fractions within a range of average molecular weights of 2,000 to 25,000 were isolated. Enzymatic oxidation of the acid hydrolyzate of the polymers revealed that no detectable racemization had occurred.  相似文献   

10.
The formation of complex between the Cu2+ ion and poly(α-L -glutamic acid) [poly(Glu)] in 150 mM NaCl solutions was studied by uv–visible absorption and equilibrium dialysis methods at the mixing ratios of Glu residues to Cu2+, R, of 32, 16, and 8 and in the pH range 4–7. The results showed that more than 90% of Cu2+ ions bind to the poly(Glu) at pH > 4.9, but the bound Cu(II) begins to dissociate with a decrease in pH. The absorption spectra of bound Cu(II) varied with pH and R in a complicated manner. Three different component spectra were disclosed from the analysis of the pH dependence of the bound spectra. We concluded that poly(Glu)–Cu(II) complexes fall into three classes in the pH range 4–7, with the proportions of these complexes varying with both pH and R. The three complexes predominate either in the helix or extended-coil region, in the helix–coil transition region, or in the helix-aggregate region. The stability constant and binding mode of each Cu(II)–Glu complex were estimated from the dialysis data. With these results, the possible structure of each complex is discussed.  相似文献   

11.
Young Kee Kang    Byung Jin Byun 《Biopolymers》2012,97(12):1018-1025
The γ‐peptide β‐turn structures have been designed computationally by the combination of chirospecific γ 2 , 3 ‐residues of 2‐(aminomethyl)cyclohexanecarboxylic acid (γAmc6) with a cyclohexyl constraint on the Cα?Cβ bond using density functional methods in water. The chirospecific γAmc6 dipeptide with the (2S,3S)‐(2R,3R) configurations forms a stable turn structure in water, resembling a type II′ turn of α‐peptides, which can be used as a β‐turn motif in β‐hairpins of Ala‐based α‐peptides. The γAmc6 dipeptide with homochiral (2S,3S)‐(2S,3S) configurations but different cyclohexyl puckerings shows the capability to be incorporated into one of two β‐turn motifs of gramicidin S. The overall structure of this gramicidin S analogue is quite similar to the native gramicidin S with the same patterns and geometries of hydrogen bonds. Our calculated results and the recently observed results may imply the wider applicability of chirospecific γ‐peptides with a cyclohexyl constraint on the backbone to form various peptide foldamers. © 2012 Wiley Periodicals, Inc. Biopolymers 97:1018–1025, 2012.  相似文献   

12.
13.
The conformational preference of Cα,α-diphenylglycinc (Døg) and Cα,α-dibenzylglycine (Dbz) residues was assessed in selected derivatives and small peptides by conformational energy computations, ir absorption, 1H-nmr, and x-ray diffraction. Conformational energy computations on the two monopeptides strongly support the view that these Cα,α-symmetrically disubstituted glycines are conformationally restricted and that their minimum energy conformation falls in the fully extended (C5) region. The results of the theoretical analyses appear to be in agreement with the solution and crystal-state structural propensities of three derivatives and seven di-and tripeptides.  相似文献   

14.
Copolymers of γ-methyl D - and L -glutamates with various D /L ratios were prepared. Infrared absorption spectra of solid films were measured and sums of right- and left-handed helix contents were determined from intensities of amide V bands. Farultraviolet absorption spectra and optical rotatory dispersion of these copolymers in solutions are used to ascertain their helical character. Chain conformations of DL -copolypeptides are discussed.  相似文献   

15.
The conformational preferences of helix foldamers having different sizes of the H‐bonded pseudocycles have been studied for di‐ to octa‐γ2,3‐peptides based on 2‐(aminomethyl)cyclohexanecarboxylic acid (γAmc6) with a cyclohexyl constraint on the Cα–Cβ bond using density functional methods. The helical structures of the γAmc6 oligopeptides with homochiral configurations are known to be much stable than those with heterochiral configurations in the gas phase and in solution (chloroform and water). In particular, it is found that the (P/M)?2.514‐helices are most preferred in the gas phase and in chloroform, whereas the (P/M)?2.312‐helices become most populated in water due to the larger helix dipole moments. As the peptide sequence becomes longer, the helix propensities of 14‐ and 12‐helices are found to increase both in the gas phase and in solution. The γAmc6 peptides longer than octapeptide are expected to exist as a mixture of 12‐ and 14‐helices with the similar populations in water. The mean backbone torsion angles and helical parameters of the 14‐helix foldamers of γAmc6 oligopeptides are quite similar to those of 2‐aminocyclohexylacetic acid oligopeptides and γ2,3,4‐aminobutyric acid tetrapeptide in the solid state, despite the different substituents on the backbone. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 87–95, 2014.  相似文献   

16.
In the present paper we describe the synthesis, purification, single-crystal x-ray analysis, solution conformational characterization, and conformational energy calculations of the cyclic tetrapeptide cyclo- (β-Ala-L -Pro-β-Ala-L -Val). The peptide was synthesized by classical solution methods and the cyclization of the free tetrapeptide was accomplished in good yields in diluted methylene chloride solution using N,N-dicyclohexyl-carbodiimide. The compound crystallizes in the monoclinic space group P21 from ethanol with two independent molecules in the unit cell. All peptide bonds are trans. The nmr molecular conformation in the acetonitrile solution as well as that derived from the molecular dynamic simulation in vacuo is quite different from those observed in the solid state and is very similar to that previously observed for the parent compound cyclo-(β-Ala-L -Pro-β-Ala-L -Pro). © 1993 John Wiley & Sons, Inc.  相似文献   

17.
The peptide Boc-L-Val-ΔPhe-ΔPhe-L-Val-OCH3 was synthesized by the azlactone method in solution phase, and its crystal and molecular structures were determined by x-ray diffraction method. Single crystals were grown by slow evaporation from a methanol/water solution at 6°C. The crystals belong to an orthorhombic space group P212121 with a = 10.478 (6) Å, b = 13.953 (1), c = 24.347 (2) and Z = 4. The structure was determined by direct methods and refined by least squares procedure to an R value of 0.052. The structure consists of a peptide and a water molecule. The peptide adopts two overlapping β-turn conformations of Types II and I′ with torsion angles: ϕ1 = -54.8 (6), ψ1 = 130.5 (4), ϕ2 = 65.8 (5), ψ2 = 12.8 (6), ϕ3 = 79.4 (5), ψ3 = 3.9 (7)°. The conformation is stabilized by intramolecular hydrogen bonds involving Boc CO and NH of ΔPhe3 and CO of Val1 and NH of Val4. The molecules are tightly packed in the unit cell. The crystal structure is stabilized by hydrogen bonds involving NH of ΔPhe2 and CO of a symmetry related (x-½, ½ -y, -z) ΔPhe2. The solvent-water molecule forms two hydrogen bonds with peptide molecule involving NH of Val1 as an acceptor and another with CO of a symmetry related (1 -x, y-½, ½ -z) ΔPhe3 as a donor. These studies indicate that a tetrapeptide with two consecutive ΔPhe residues sequenced with valines on both ends adopts two overlapping β-turns of Types II and I′. © 1996 John Wiley & Sons, Inc.  相似文献   

18.
19.
20.
The hydrolysis of p-nitrophenyl acetate is catalyzed by imidazole, free in solution or as the side chain in poly(His-Ala-Glu). This is based on the observations that the reaction is first order in ester and first order in nonprotonated imidazole. Catalysis of p-nitrophenyl acetate hydrolysis is dependent on solvent conditions. The effect of low concentrations of ethanol, dioxane, and trifluoroethanol were investigated. As the concentration of organic solvent is increased, the second-order rate constant for imidazole catalysis decreases. The decrease, however, is greater for imidazole than for poly(His-Ala-Glu). In 2% trifluoroethanol/water solution, free imidazole has twice the catalytic activity of polymeric imidazole, while in 40% trifluoroethanol/water they have equal activity. Since under the latter solvent conditions poly(His-Ala-Glu) is partially α-helical, the relative improvement in polymeric–imidazole catalysis may be attributed to imidazole hydrogen-bonded to a carboxylate ion. With this assumption the carboxylate–imidazole hydrogen-bonded system has been calculated to have three times the base catalytic activity of imidazole.  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