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.
The structure of poly(aspartic acid) prepared by thermal polycondensation has been studied by means of nmr spectroscopy. The analysis of the 13C-nmr spectra of the polymer at various pH values and comparison with the spectrum of poly(α-L -aspartic acid) revealed that the polymer contained aspartic acid linked in α- and β-peptide bonds. The mole fraction of the β-peptide bonds has been found to be 0.8 ± 0.1. The significance of the results for the evolutionary theory of S. W. Fox is mentioned.  相似文献   

2.
The conformation of poly(α-L -aspartic acid) was investigated on a sample in which β-bonds were not detected. CD and ir spectroscopy showed that poly(α-L -aspartic acid) passes through a conformational change induced by changes of the degree of ionization that is accompanied by precipitation; the precipitate is probably highly helical. The change was also detected by potentiometric titration.  相似文献   

3.
The aim of this study was to generate a new type of nanoparticles made of quaternized chitosan (QCS) and poly (aspartic acid) and to evaluate their potential for the association and delivery of protein drugs. QCS and poly (aspartic acid) were processed to nanoparticles via the ionotropic gelation technique. The size, polydispersity, zeta potential, and morphology of the nanoparticles were characterized. Entrapment studies of the nanoparticles were conducted using bovine serum albumin (BSA) as a model protein. The effects of the pH value of nanoparticles with different QCS/poly (aspartic acid) ratios, QCS molecular weight (MW), poly (aspartic acid) concentration, and BSA concentration on the nanoparticle size, the nanoparticle yield, and BSA encapsulation were studied in detail. Suitably pH value of nanoparticles with different QCS/poly (aspartic acid) ratios, moderate QCS MW, optimal concentration ratio of poly (aspartic acid), and QCS favored more nanoparticles formed and higher BSA encapsulation efficiency. The release of BSA from nanoparticles was pH-dependent. Fast release occurred in 0.1 M phosphate buffer solution (PBS, pH 7.4), while the release was slow in 0.1 M HCl (pH 1.2). The results showed that the new QCS/poly (aspartic acid) nanoparticles have a promising potential in protein delivery system.  相似文献   

4.
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   

5.
V. Saudek 《Biopolymers》1981,20(8):1625-1633
Polypeptides of dicarboxylic amino acids having the monomer units linked in α- and ω-peptide bonds contain two kinds of carboxyls of different acidity. How well potentiometric titration can distinguish these two carboxyls and so characterize the nature of the peptide bonds is evaluated critically. An analysis of the equation describing the dependence of pH on the degree of neutralization based on neglecting the polymer effect and a discussion of the dissociation behavior of polyanions show that the method of evaluating experimental data found in the literature is incorrect. Nevertheless, if a conformational transition does not interfere, some useful and reliable information may be gained by this method; namely, an indication of the presence of two different peptide bonds, their mole ratio, and an approximate pK value for the carboxyl of the amino acid linked in the ω-peptide bond. The presence of two types of carboxyls complicates the evaluation of the titration curves in the conformation studies.  相似文献   

6.
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   

7.
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   

8.
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   

9.
Transient electric birefringence of poly(L -α,γ-diaminobutyric acid hydrochloride) in methanol/water mixtures has been measured over a wide range of field strengths and solvent compositions and at different polymer concentrations and temperatures. The molar ellipticity at 222 nm and the specific Kerr constant underwent an abrupt change between 75 and 80 vol % methanol at 25°C, accompanied by a solvent-induced helix–coil transition. Anomalous birefringence transients were observed between 78 and 80 vol% methanol above threshold field strengths. The double logarithmic plots of the steady-state specific birefringence versus the square of field strength for different solvent compositions and polymer concentrations could be superimposed by shifting them horizontally along the abscissa and vertically along the ordinate except for the range where anomalous transients were observed. The threshold field strength could be estimated from the point at which a downward deviation occurred. It increased with increasing polymer concentration and with increasing methanol content on the verge of the transition region. The results were interpreted as indicating that a conformational change from the charged helix to the charged coil is induced by high fields in this system, as in the case of poly(L -lysine hydrobromide) in methanol/water mixtures.  相似文献   

10.
The CD of aqueous solutions of poly(S-carboxymethyl-L -cysteine) and poly(S-carboxyethyl-L -cysteine) has been measured at different pH, and the pH-induced β-coil transition is observed by changes in residue ellipticity of dichroic bands around 200 and 225 nm. The residue ellipticity at 200 nm of the former polypeptide is twice as large as that of the latter, when the β-conformation is formed in solution. However, the β-conformation of the latter polypeptide is more stable against electrostatic repulsion than that of the former. The transition curve of poly(S-carboxymethyl-L -cysteine) has also been determined for different molecular weights. The curves were found to be completely coincident with one another if the degree of polymerization were higher than about 100. Such a transition curve is generally divided into three steps: initiation, cooperative formation, and rearrangement of hydrogen bonds. The cooperative step is very sharp, occurring at a constant pH. These steps become agglomerated into two or one when the polypeptide concentration or added salt concentration is increased.  相似文献   

11.
B. R. Malcolm 《Biopolymers》1983,22(1):319-322
Electron diffraction photographs of poly (α-aminoisobutyric acid) treated with dichloracetic acid are shown to be consistent with space group R3c. This is strong evidence for a hexagonal cell in which right- and left-handed 310-helices form a honeycomb structure. It appears that the dichloracetic acid molecules are located in the centers of the holes.  相似文献   

12.
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   

13.
Fourier transform ir attenuated total reflectance (FTIR ATR) spectra have been obtained to investigate the secondary structure of poly(γ -methyl L -glutamate) (PMLG) surfaces untreated and treated with formic acid in a quantitative manner. Curve analysis including Fourier self-deconvolution and the band fitting was applied to the ir spectra in the amide II region, revealing that the amide II band of those surfaces consists of five components. The essentially α-helical form in the PMLG surface layer transformed readily into the β-structure by the formic acid treatment, and the β-structure content increased with increasing time of the treatment. The content of random coil structure of treated PMLG was generally very little and/or negligible. The depth profile obtained by considering the depth of ir beam penetration indicated that the β-structure content also increased with approaching the outermost surface.  相似文献   

14.
Examination of beta-carbons coordinates of seryl, aspartyl and histidyl residues in active sites of alpha-chymotrypsin and subtilisin BPN' shows that a close geometrical arrangement can be obtained in an antiparellel beta-structure. Therefore some polypeptides incorporating serine, aspartic acid and histidine, poly (Gly-Ser-Asp-His-Ala-Pro) and poly [(Asp-Leu-AsP-Leu)10, (His-Leu-Ser-Leu)1], and expected to have some tendency to give rise to an antiparallel beta-conformation, have been prepared and studied. The second polymer only adopts a fairly well-defined beta-structure in aqueous solution. Catalytic activities of these products towards p-nitrophenyl acetate are not improved as compared to histidine. However, kinetic pK of histidine side-chain depends markedly upon the nature of the product, owing probably to a hydrophobic environment effect.  相似文献   

15.
The cupric complexes of poly(Nε-acetoacetyl-L -lysine), [Lys(Acac)]n′ poly(Nδ-acetoacetyl-L -ornithine), [Orn(Acac)]n′ and poly(Nγ-acetoacetyl-L -diaminobutyric acid), [A2bu-(Acac)]n, as well as of the model compound n-hexyl acetoacetamide, have been investigated by means of absorption, potentiometric, equilibrium dialysis, and CD measurements. While in the complex of the model compound, one chelating group is bound to one cupric ion, in the polymeric complexes two β-ketoamide groups are bound to Cu(II) under the same experimental conditions. The binding constant of cupric ions to the three polymers and the formation constant of the Cu(II)-nhexylacetoacetamide complex have been evluated. Investigation on the chiroptical properties of the three polymeric complexes shows that the peptide backbone does not undergo conformational transitions, remaining α-helical when up to 20% of the side chains are bound to Cu(II). The optical activity of the β-ketoamide chromophores is substantially affected by complex formation and is discussed in terms of asymmetric induction from the chiral backbone.  相似文献   

16.
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   

17.
The electron-diffraction pattern of an oriented film of poly(α-aminoisobutyric acid) in the 310-helical conformation has been analyzed. The conformation was obtained by a linked-atom least-squares refinement of average values from crystal structures. Specimens treated with dichloracetic acid, to improve their crystallinity, conform to space group R3c with a = 21.8 Å, c = 5.95 Å. The structure contains channels that can accommodate molecules of dichloracetic acid. One molecule of acid per six residues fills the channels, and the R-factor then is 34% using 23 reflections. Ir evidence is presented to show that the acid may hydrogen bond to the peptide groups. Some reflections occasionally observed on the diffraction photographs are attributed to a 15/4 α-helix. The significance of the results is considered in relation to Aib-containing peptides.  相似文献   

18.
Immobilized β-galactosidase gel was prepared using poly(vinylpyrrolidone) (PVP) under β-ray irradiation. In contrast to the gelation of N-vinylpyrrolidone monomer–enzyme solution, the gelation of PVP-β-galactosidase solution (PVP content: 10%) was almost completely uneffected by the dose rate and amount of phosphate present. PVP-enzyme solution was gelled by irradiation with 3.0 Mrad. The expressed activity of the PVP-enzyme gel was about 30% of the initial activity and added activity was almost totally entrapped. No leakage of enzyme from these gels could be detected. Leakage was, however, detected in the case of the gelation of PVP-enzyme solution containing more than 1% of enzyme protein. When the general properties of the gel were compared with those of the native enzyme, the gel proved to be slightly inferior to the native enzyme with respect to optimum temperature, heat stability, pH activity, and pH stability. Continuous hydrolysis of lactose in acid whey could be carried out at 50°C using a column packed with the gel and sawdust and the degree of hydrolysis was found to be almost, constant for 12 days. The merits of using PVP in the immobilization of enzymes include the simplicity of the procedure and the fact that the PVP-enzyme gel can be used in the food industry without anxiety because of its high degree of compatibility with living organisms.  相似文献   

19.
Sphingomonas sp. KT-1 hydrolyzes poly(aspartic acid) (PAA) containing alpha- and beta-amide units and has at least two different types of PAA hydrolases. The PAA hydrolase-1 hydrolyzes selectively beta-beta amide units in PAA. Molecular cloning of PAA hydrolase-1 from Sphingomonas sp. KT-1 has been carried out to characterize its gene products. Genetic analysis shows that the deduced amino acid sequence of PAA hydrolase-1 has a similarity with those of the catalytic domain of poly(3-hydroxybutyric acid) (PHB) depolymerases from Alcaligenes faecalis AE122 and Pseudomonas lemoignei. Site-specific mutation analysis indicates that (176)Ser is a part of a strictly conserved pentapeptide sequence (Gly-Xaa-Ser-Xaa-Gly), which is the lipase box, and plays as an active residue.  相似文献   

20.
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   

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