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(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   

2.
Sequential polypeptides with the repeating units L -glutamyl-L -DOPA, L -DOPA-L -glutamyl-L -DOPA, L -glutamyl-L -glutamyl-L -DOPA, L -DOPA-L -DOPA-L -glutamyl-L -DOPA, and L -glutamyl-L -glutamyl-L -glutamyl-L -DOPA have been synthesized by solution polymerization of the p-nitrophenyl esters of the corresponding di-, tri-, and tetrapeptides. The O, O′-dimethyl and γ-methyl groups were used to protect side chains of L -DOPA and L -glutamic acid. The monomers for the polytripeptides and polytetrapeptides were prepared by stepwise elongation, using the dicyclohexylcarbodiimide coupling method. Moderately high molecular weight sequential polypeptides were obtained. The protected groups of the side chain were removed simultaneously by use of boron tribromide in chloroform. Trimethylphosphate-soluble sequential polypeptides containing L -DOPA were obtained.  相似文献   

3.
Six different sequential polypeptides with the repeating units L -lysyl-L -DOPA, L -DOPA-L -lysine, L -lysyl-L -lysyl-L -DOPA, L -DOPA-L -DOPA-L -lysine, L -lysyl-L -lysyl-L -lysyl-L -DOPA, and L -DOPA-L -DOPA-L -DOPA-L -lysine have been synthesized by solution polymerization of the p-nitrophyenyl esters of the corresponding di-, tri-, and tetrapeptides. The O,O′-dimethyl and N-ε-2-chlorobenzyloxycarbonyl groups were used to protect side chains of L -DOPA and L -lysine. The monomers for the polytripeptides and polytetrapeptides were prepared by stepwise elongation, using the dicyclohexylcarbodiimide coupling method. Moderately high molecular weight sequential polypeptides were obtained. The addition of 1-hydroxybenzotriazole increased their molecular weights, but not so effectively. The protected groups of the side chains were removed simultaneously by use of boron tribromide in chloroform. Water-soluble sequential polypeptides containing L -DOPA were obtained.  相似文献   

4.
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   

5.
Laminins, a family of heterotrimeric proteins with cell adhesive/signaling properties, are characteristic components of basement membranes of vasculature and tissues. In the present study, permeabilized platelets were found to react with a monoclonal antibody to laminin γ1 chain by immunofluorescence. In Western blot analysis of platelet lysates, several monoclonal antibodies to γ1 and β1 laminin chains recognized 220- to 230-kDa polypeptides, under reducing conditions, and a structure with much slower electrophoretic mobility under nonreducing conditions. Immunoaffinity purification on a laminin β1 antibody–Sepharose column yielded polypeptides of 230, 220, 200, and 180 kDa from platelet lysates. In the purified material, mAbs to β1 and γ1 reacted with the two larger polypeptides, while affinity-purified rabbit antibodies to laminin α4 chain recognized the smallest polypeptide. Identity of the polypeptides was confirmed by microsequencing. One million platelets contained on average 1 ng of laminin (approximately 700 molecules per cell), of which 20–35% was secreted within minutes after stimulation with either thrombin or phorbol ester. Platelets adhered to plastic surfaces coated with the purified platelet laminin, and this process was largely inhibited by antibodies to β1 and α6 integrin chains. We conclude that platelets contain and, following activation, secrete laminin-8 (α4β1γ1) and that the cells adhere to the protein by using α6β1 integrin.  相似文献   

6.
In addition to being refractive proteins in the vertebrate lens, the two α-crystallin polypeptides (αA and αB) are also molecular chaperones that can protect proteins from thermal aggregation. The αB-crystallin polypeptide, a functional member of the small heat shock family, is expressed in many tissues in a developmentally regulated fashion, is stress-inducible, and is overexpressed in many degenerative diseases and some tumors indicating that it plays multiple roles. One possible clue to α-crystallin functions is the fact that both polypeptides are phosphorylated on serine residues by cAMP-dependent and cAMP-independent mechanisms. The cAMP-independent pathway is an autophosphorylation that has been demonstrated in vitro, depends on magnesium and requires cleavage of ATP. Disaggregation of αA-, but not αB-crystallin into tetramers results in an appreciable increase in autophosphorylation activity, reminiscent of other heat shock proteins, and suggests the possibility that changes in the aggregation state of αA-crystallin are involved in yet undiscovered signal transduction pathways. The α-crystallin polypeptides differ with respect to their abilities to undergo cAMP-dependent phosphorylation, with preference given to the αB-crystallin chain. These differences and complexities in α-crystallin phosphorylations, coupled with the differences in expression patterns of the two α-crystallin polypeptides, are consistent with the idea that each polypeptide has distinctive structural and metabolic roles.  相似文献   

7.
The conformation of several samples of poly(α,β-L -Asp) with a molar fraction of β-bonds ranging from 0.1 to 0.55 was investigated by means of ir and CD spectroscopy and potentiometric titration and compared with the results obtained previously with poly(α-L -Asp). All samples investigated underwent a conformational change induced by changes in their degree of ionization: unpronounced ir absorption of amide V at 650 cm?1 was shifted to 620 cm?1 and substantially increased on deionization; CD spectra changed with the degree of ionization, passing through an isosbestic point; and the pattern of the titration curves was more complex than that of a simple polyelectrolyte. The conformation developing with the decreasing degree of ionization may be considered to be α-helix, as deduced according to the analogous behavior of other polypeptides. The extent of the conformational change in the individual samples depends on the molar fraction of β-bonds: the higher it is, the lower is the helix-forming ability of the sample.  相似文献   

8.
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   

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.
Local conformation and overall conformation of poly(γ‐DL‐glutamic acid) (PγDLGA) and poly(γ‐L‐glutamic acid) (PγLGA) in aqueous solution was studied as a function of degree of ionization ε by 1H‐NMR, circular dichroism, and potentiometric titration. It was clarified that their local conformation is represented by random coil over an entire ε range and their overall conformation is represented by expanded random‐coil in a range of ε > ε*, where ε* is about 0.3, 0.35, 0.45, and 0.5 for added‐salt concentration of 0.02M, 0.05M, 0.1M, and 0.2M, respectively. In a range of ε < ε*, however, ε dependence of their overall conformation is significantly differentiated from each other. PγDLGA tends to aggregate intramolecularly and/or intermolecularly with decreasing ε, but PγLGA still behaves as expanded random‐coil. It is speculated that spatial arrangement of adjacent carboxyl groups along the backbone chain essentially affects the overall conformation of PγGA in acidic media. © 2015 Wiley Periodicals, Inc. Biopolymers 105: 191–198, 2016.  相似文献   

11.
Reduced viscosity and infrared spectra of low-molecular-weight poly-γ-benzyl-L -glutamate (which was prepared by polymerization of the N-carboxyanhydride with n-hexylamine initiation at [A]/[I] 3, 4, and 8) have been measured in various organic solvents. Infrared spectra indicate that the polypeptide molecules consist of a series of residues of two forms, the solvated σ-form and the hydrogen-bonded β-form, and relative abundance of the two forms depends on solvent species and polypeptide concentration. An approximate method is developed for estimating the content of β-structure from a single spectrum of dissolved polypeptide. The reduced viscosity of some solutions is scarcely dependent in polypeptide concentration, in which a single conformation is predominantly kept over the concentration range. In the other solutions the reduced viscosity displays a strong concentration dependence or some anomalous behavior. The observed viscosity behavior has been attributed to the changes in size and shape of aggregates, which are determined by the number of hydrogen bonds in the aggregate. This unusual behavior is exhibited by solutions of the polypeptides which have a moderate content of β-structure at a finite concentration. Both the content of β-structure and the extent of association increase in the following solvents, ranked in order of effectiveness: dimethylformamide, trifluoroethanol < trimethyl phosphate < chloroform < dioxane < ethylene dichloride < ethylene dibromide. Infrared spectra suggest that the conformation of the polypeptide in dichloroacetic acid differs from either the σ- or the β-conformation.  相似文献   

12.
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   

13.
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   

14.
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   

15.
The aggregation of poly(γ-benzyl-α,L -glutamate) and its enantiomer in toluene has been investigated by following the viscosity as a function of temperature, concentration, molecular weight, molecular-weight distribution, helix chirality, and shear rate. The temperature and concentration data for a 138,000-molecular-weight sample was fitted to an open, reversible end-to-end aggregation model. The aggregation numbers resulting from this fit were consistent with the sudden onset in non-Newtonian flow resulting from only a 0.2-wt% increase in concentration. The association equilibrium constant was then used to predict viscosity for comparison with other data, in particular, the effect of molecular weight and molecular-weight distribution. A mixture of right-and left-handed helices showed the aggregation was not chiral selective. The stiffness of end-to-end aggregated (hydrogen-bonded) molecules differed little from their covalent counterparts, at least below a molecular weight of ~106. We conclude that polybenzylglutamate aggregation in toluene can be described by an open end-to-end aggregation model.  相似文献   

16.
The present study was undertaken to examine the influence of the reversal of the sidechain sequential order on the conformation of branched polypeptides. At the same time, the influence of the optically active amino acid joined directly to the poly (L -Lys) backbone and the DL -Ala oligomer grafted as chain-terminating fragment were separately analyzed. Therefore two sets of polypeptides were synthesized corresponding to the general formula poly [Lys-(Xi,)] (XK) and poly[Lys-(DL -Alam-Xi)] (AXK) when X = Ala, D -Ala, Leu, D -Leu, Phe, D -Phe, Ile, Pro, Glu.,D -Glu, or His. For coupling amino acid X to polylysine, three types of active ester methods were compared: the use of pentafluorophenyl or pentachlorophenyl ester, and the effect of the addition of an equimolar amount of 1-hydroxybenzotriazole. After cleavage of protecting groups, AXK polypeptides were synthesized by grafting short oligo (DL -Ala) chains to XK by using N-carboxy-DL -Ala anhydride. The CD measurements performed in water solutions of various pH values and ionic strengths were used for classification of the polypeptide conformations as either ordered (helical) or unordered. Different from what was observed with the unsubstituted poly (L -Lys), poly[Lys-(Xi)] type polypeptides can adopt ordered structure even under nearly physiological conditions (pH 7.3, 0.2M NaCl). These data suggest that the introduction of amino acid residue with either (ar) alkyl side chain (Ala, Leu, Phe) or negatively charged side chain (Glu) promotes markedly the formation of ordered structure. Comparison of chiroptical properties of poly [Lys- (DL -Alam-Xi)] and of poly [Lys- (Xi)] reveals that side-chain interactions play an important role in the stabilization of ordered solution conformation of AXK type branched polypeptides. The results give rather conclusive evidence that not only hydrophobic interactions, but also ionic attraction, can be involved in the formation and stabilization of helical conformation of branched polypeptides. © 1993 John Wiley & Sons, Inc.  相似文献   

17.
A M Schwartz  G D Fasman 《Biochemistry》1977,16(10):2287-2299
Complexes of DNA with polypeptides composed of Lys, Ala, and Gly in both a sequential order, poly(L-lysine-L-alanine-glycine), and a statistical distribution, poly(L-lysine36-L-alanine28-glycine), were prepared using gradient dialysis. These polypeptide-DNA complexes were studied using ultraviolet absorption (UV) and circular dichroism (CD) to probe the conformation, binding, and melting behavior of DNA in the complex. Complexes with the sequential polypeptide showed no structural change in the DNA; however, the complexes with the random polypeptide yield CD spectra similar to phi DNA [Maniatis, T., Venable, Jr., J.S., and Lerman, L.S. (1974), J. Mol. Biol. 84, 37]. A second sequential polypeptide, poly(L-Lys-L-Ala-L-Pro)n, -DNA complex was also studied. It was found to exhibit pronounced structural changes as a function of ionic strength and poly-peptide-DNA ratio, more similar to the random sequence that the ordered sequence of the Lys, Ala, Gly polymer. Thus the importance of the composition and amino acid sequence in polypeptides which bind to DNA, even in such simple systems, is demonstrated. Evidence from thermal denaturation, employing simultaneous monitoring of CD and UV changes, supports a model in which specific polypeptides cause condensation of the DNA in the complex into an asymmetric tertiary structure. The relevance of these model systems to chromatin is discussed.  相似文献   

18.
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   

19.
The dependence of the dielectric constant and dielectric loss of aqueous solutions of poly-ε, N-succinyl-L -lysine on its degree of polymerization, degree of neutralization, concentration of the polymer, and counterion type was studied in a frequency range from 300 Hz to 5 MHz. Regardless of the conformation, a low-frequency dispersion in a frequency range lower than 10 kHz and a high-frequency dispersion in a range higher than 100 kHz were found. The large value of the dielectric increment, its nonlinear dependence on concentration, its remarkable dependence on counterion type, and its dependence on the degree of polymerization suggest that the low-frequency dispersion is mainly due to the polarization of loosely bound counterions. These data were found for both the helical and coiled forms. The rotational motion of the electric dipole on the molecule could not have been primarily responsible for these results. On the other hand, the high-frequency dispersions may be attributable to the Maxwell–Wagner-type effect. The results were compared with the dispersions of poly(L -glutamic acid), poly(L -lysine), and their salts reported previously.  相似文献   

20.
Poly(β–l–aspartate)s are known to take up helical conformations reminiscent of the α-helix of polypeptides. The isobuttyl, n-butyl, and 2-methoxyethyl esters have been examined by polarized ir spectroscopy in order to discriminate between the left ( 1L ) and right ( 2R ) -handed conformations, which are known to be compatible with the 13/4-helix adopted by these polyamides when crystallized in the hexagonal form. Dichroic ratios obtained from samples stretched in poly(ethylene oxide) together with orientation measurements made by x-ray diffraction were used to estimate the transition moment directions of amide A, I, and II bands with respect to the fiber axis. These were compared to those calculated by modeling simulations to conclude that the right-handed conformation consisting of 14-membered hydrogen-bonded rings is the correct model for the 13/4-helix. These results give definite support to earlier molecular mechanics calculations, which had shown that the 2R model is energetically favored over the 1L by about 2. 5 kcal/(mol residue). © 1995 John Wiley & Sons, Inc.  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