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 interaction between poly(α,L -lysine) (DP = 180) and glutaraldehyde was investigated in dilute aqueous solution by measurement of the kinetics of proton release at constant pH and temperature and at various concentrations of the reaction components. Under various conditions, the release of protons at constant pH appeared kinetically to be composed of at least two steps: an initial zero-order reaction, followed by a slower reaction. At excess of polylysine amino groups, the pH optimum for the rates of reaction was at pH 9–10 (24–25°C). Under the conditions used and at pH 8, the initial rate of the second kinetic step was proportional to the glutaraldehyde concentration and was practically independent of polylysine concentration at pH 8 and 8.6, at an excess of amino groups. At pH values of 7, 8, and 8.6 the apparent overall energy of activation for the second kinetic step was 18–19 kcal/mole (temp. range 4–40°C). Comparing acetaldehyde with the difunctional glutaraldehyde, it was found that the rate of proton release was much smaller in the case of acetaldehyde. Comparing n-butylamine with the macromolecular polylysine at equal concentrations of amino groups, the rates of proton release were much smaller in the case of n-butylamine. Collagen in aqueous medium also interacted with glutaraldehyde in a manner analogous to polylysine, although the conditions were not quite comparable. In the case of collagen, the initial fast proton liberation step was relatively much larger than in the case of polylysine. A reaction scheme for the initial reaction steps is being proposed which includes primary complex formation between glutaraldehyde and polylysine. This dialdehyde–polyamino acid system is considered to serve as a model for tanning processes of hides and for fixation procedures.  相似文献   

2.
H Yamamoto  T Hayakawa 《Biopolymers》1972,11(6):1259-1268
The solvent-and pH-induced conformational changes are examined in order to investigate the influence of benzyl group. Polymer was prepared via N?-benzyloxycarbonyl, N?-benzyl-Nα-carboxy-L -lysine anhydride. The resulting poly (N?-benzyloxycarbonyl, N?-benzyl-L -lysine) was obtained in high yield and had a high molecular weight. The protected polymer was removed into poly (N?-benzyl-L -lysine) by treating it with hydrogen bromide. From the results of the ORD and CD, the protected polymer has a righthanded α-helix, showing [m′]233 = –10,300, [θ]220 = –27,600 and [θ]207 = –25,100 in dioxane. The breakdown of the helical conformation is found to occur at 8% dichloroacetic acid in chloroform-dichloroacetic acid mixture. In the pH range 3.35–6.90, poly (N?-benzyl-L -lysine) is in a random coil structure. In the pH range 7.50–13.0, the polypeptide has a right-handed α-helix structure; [m′]233 = –12,000, [0]220 = –27,200, and [0]207 = –27,000. In comparison with poly-L -lysine, the coil-to-helix transition is observed at lower pH range in 50% n-propanol. Above pH 8 by heating, the α ? β transition of poly (N?-benzyl-L -lysine) is not observed in an aqueous media.  相似文献   

3.
Circular dichroism spectra for acid-soluble calfskin collagen, gelatin, and poly(proline) II in solution have been extended into the vacuum ultraviolet region. The extended spectrum of gelatin reveals that the circular dichroism of this unordered polymer is more closely related to the spectrum of charged polypeptides than might be evident from near ultraviolet work. A short-wavelength band is found at about 172 nm, which corresponds in position, magnitude, and sign to a band recorded earlier for poly(L -glutamic acid) at pH 8.0. This band is observed in a helical structure for the first time in the vacuum ultraviolet circular dichroism and absorption spectra of poly(proline) II. Both circular dichroism and absorption spectra point to the assignement of this band as the nσ*. Neither the nσ* nor the expected positive lobe of the ππ* helix band is observed in the extended circular dichroism spectrum of collagen. We postulate that these two bands cancel here in analogy to the case of α-helical poly(L -glutamic acid).  相似文献   

4.
Iwao Satake  Jen Tsi Yang 《Biopolymers》1975,14(9):1841-1846
The conformational phase diagram of poly(L -lysine) (4.6 × 10?4 M, residue) in sodium dodecyl sulfate (1.6 × 10?2 M) solution was constructed from circular dichroism results at various temperatures and pH's. Poly(L -lysine)–sodium dodecyl sulfate complexes undergo a β–helix transition upon raising the pH of the solution. The transition pH tends to shift downward at elevated temperatures. No helix–β transition can be detected for poly(L -lysine) in sodium dodecyl sulfate solution (pH > 11) even after 1-hr heating at 70°C. This is in marked contrast with uncharged poly(L -lysine) solution without sodium dodecyl sulfate, which is converted into the β-form upon mild heating of the solution above 50°C.  相似文献   

5.
Poly(Nε-stearyl-L -lysine) and poly(Nε-pelargonyl-L -lysine) were synthesized both by polymerization of Nε-pelargonyl and Nε-stearyl-L -lysine NCA and by acylation of poly(L-lysine) with pelargonyl and stearyl chloride. This second route has proven to be very useful, since completely acylated polymers are obtained in almost quantitative yield, whereas the usual scheme of preparation of ε protected poly(L-lysine) cannot easily be applied due to solubility problems. Poly(Nεpelargonyl and stearyl-L -lysine) are soluble in alcohols containing linear aliphatic chains such as n-butanol and n-octanol and in mixtures of these alcohols with hydrocarbons such as n-hexane and n-heptane. Both polymers show an α-helical conformation in the above solvents, which can be disrupted upon addition of sulfuric acid. Also in the solid state, poly(Nε-stearyl-L -lysine) and poly(Nε-pelargonyl-L -lysine) show X-ray diffraction patterns typical of order structure.  相似文献   

6.
Glutaraldehyde has been used for several decades as an effective crosslinking agent for many applications including sample fixation for microscopy, enzyme and cell immobilization, and stabilization of protein crystals. Despite of its common use as a crosslinking agent, the mechanism and chemistry involved in glutaraldehyde crosslinking reaction is not yet fully understood. Here we describe feasibility study and results obtained from a new approach to investigate the process of protein crystals stabilization by glutaraldehyde crosslinking. It involves exposure of a model protein crystal (Lysozyme) to glutaraldehyde in alkaline or acidic pH for different incubation periods and reaction arrest by medium exchange with crystallization medium to remove unbound glutaraldehyde. The crystals were subsequently incubated in diluted buffer affecting dissolution of un-crosslinked crystals. Samples from the resulting solution were subjected to protein composition analysis by gel electrophoresis and mass spectroscopy while crosslinked, dissolution resistant crystals were subjected to high resolution X-ray structural analysis. Data from gel electrophoresis indicated that the crosslinking process starts at specific preferable crosslinking site by lysozyme dimer formation, for both acidic and alkaline pH values. These dimer formations were followed by trimer and tetramer formations leading eventually to dissolution resistant crystals. The crosslinking initiation site and the end products obtained from glutaraldehyde crosslinking in both pH ranges resulted from reactions between lysine residues of neighboring protein molecules and the polymeric form of glutaraldehyde. Reaction rate was much faster at alkaline pH. Different reaction end products, indicating different reaction mechanisms, were identified for crosslinking taking place under alkaline or acidic conditions.  相似文献   

7.
The Raman spectrum of poly-DL -alanine (PDLA) in the solid state is interpreted in terms of the disordered chain conformation, in analogy with the spectrum of mechanically deformed poly-L -alanine. The polymer is largely disordered with only a small α-helical content in the solid state. When PDLA is dissolved in water, the spectra suggest that short α-helical segments are formed upon dissolution. These helical regions might be stabilized by hydrophobic bonds between side-chain methyl groups. Addition of methanol to the aqueous PDLA solutions results in a Raman spectrum resembling that of solid PDLA. This result suggests that the methanol disrupts the helical regions by breaking the hydrophobic bonds. The Raman spectra of poly-DL -leucine (PDLL) and poly-L -leucine (PLL) are compared and only slight differences are observed in the amide I and III regions, indicating that PDLL does not have an appreciable disordered chain content. Significant differences are observed in the skeletal regions. The 931-cm?1 lines in the PLL and PDLL spectra are assigned to residues in α-helical segments of the preferred screw sense, i.e., L -residues in right-handed segments and D -residues in left-handed segments (in PDLL). On the other hand, the 890-cm?1 line in the spectrum of PDLL is assigned to residues not in the preferred helical sence, i.e., L -residues in left-handed segments and D -residues in right-handed ones. The Raman spectra of poly-DL -lysine and poly-L -lysine in salt-free water at pH 7.0 are compared. The Raman spectra of the two polymers are very similar. However, this does not negate the hypothesis of local order in poly-L -lysine because the distribution of the residues in poly-DL -lysine probably tends towards blocks, and the individual blocks may take up the 31 helix.  相似文献   

8.
The conformation of chromatographically pure block oligopeptides (L -lysyl)m-(L -alanyl)n- L -tyrosyl-(L -alanyl)n-(L -lysyl)m with n = 3 and m = 6 or 3 is investigated. By circular dichroism it is shown that these peptides may exhibit a partially α-helical structure depending upon pH, ionic strength, solvent, and temprerature. An attempt is made to describe the helical content of these small peptides by utilizing the data obtained on high-molecular-weight poly(L -lysine). By measurement of the quantum yield and the decays of the peptides fluorescence, it is shown that, in aqueous solution, at neutral pH, the fluorescence of the peptides is quenched by interactions with the peptide carbonyl groups. The decays are multiexponential, which shows the presence of several conformations of the phenolic chromophore relative to the peptide chain. The addition of methanol, which induced the helix formation, decreases the quenching of the fluorescence and the multiexponential character of the decays. In presence of sodium hydroxide, which further increases the helical content of the peptides, a dynamic quenching occured that can be attributed to interactions between the phenol hydroxyl group of tyrosine (ith residue) and the ε-amino groups of the (i+4)th and (i -4)th lysyl residues.  相似文献   

9.
Lysine oligopeptides. Preparation by ion-exchange chromatography   总被引:3,自引:0,他引:3  
The preparation of L -lysine peptides (Lysn, n = 2–14) from polyL -lysine is described. Fractionation by ion-exchange column chromatography of poly-L -lysine hydrolysates on a preparative scale resulted in 0.2–1.0 g quantities of individual members of the poly-L -lysine series. The peptides isolated proved to be analytically pure and the optical configuration was fully retained, as demonstrated by complete enzymic digestion. Peptides higher than n = 14 were also prepared. They consisted of oligolysine groups of narrow and accurately determined size distribution. Potentiometric titrations were used both to characterize the products and to demonstrate the characteristic dependence of the dissociation constants on size of the peptide.  相似文献   

10.
The present investigation was carried out to study the effects of methanolic extracts of Finger millet (Eleusine coracana) and Kodo millet (Paspalum scrobiculatum) on glycation and crosslinking of collagen. Tail tendons obtained from rats weighing 200-225 g were incubated with glucose (50 mM) and 3 mg of extracts of the above millets in methanol under physiological conditions of temperature and pH for 10 days. Early glycation was estimated by phenol-sulfuric acid method and the crosslinking was assessed by pepsin digestion, cyanogen bromide peptide map and viscosity measurements. Tendon collagen incubated with glucose (50 mM) showed 65% solubility on pepsin treatment; poor resolution of bands in the cyanogen bromide peptide map, and intrinsic viscosity of 0.84 dl/g. The collagen incubated with Finger millet and Kodo millet extracts inhibited glycation; 89% and 92% solubility in pepsin; good resolution of bands in the cyanogen bromide peptide map and intrinsic viscosity of 0.46 and 0.58 dl/g respectively. The study implicates the potential usefulness of the above millets in protection against glycation and crosslinking of collagen.  相似文献   

11.
The circular dichroism of Ac-(Ala)x-OMe and H-Lys-(Lys)x-OH with x = 1, 2, 3, and 4 has been measured in aqueous solutions. The oligomers with x = 4 show similar circular dichroism spectra in water when the lysyl amino groups are protonated, and they respond in similar fashion to heating and to sodium perchlorate. Both oligomers at 15°C exhibit a positive circular dichroism band at 217–218 nm, which is eliminated by the isothermal addition of 4 M sodium perchlorate or by heating. The positive circular dichroism of the lysine oligomer is also eliminated when the pH is elevated to deprotonate the amino groups. Positive circular dichroism is still observed for Ac-(Ala)4-OMe at elevated pH. Circular dichroism spectra have been estimated for poly(L -alanine) and poly(L -lysine) as statistical coils under the above conditions, based on the trends established with the oligomers. Poly(L -lysine) and poly(L -alanine) are predicted to exhibit similar circular dichroism behavior in aqueous solution so long as the lysyl amino groups are protonated. The circular dichroism of the statistical coil of poly(L -lysine), but not poly(L -alanine), is predicted to change when the pH is elevated sufficiently to deprotonate the lysyl amino groups. These results suggest that the unionized lysyl side chains participate in interactions that are not available to poly(L -alanine). Hydrophobic interactions may occur between the unionized lysyl side chains. Protonation of the lysyl amino groups is proposed to disrupt these interactions, causing poly(L -alanine) and protonated poly(L -lysine) to have similar circular dichroism properties.  相似文献   

12.
M J Grourke  J H Gibbs 《Biopolymers》1971,10(5):795-808
The helix–coil transitions of aqueous solutions of poly-α-L -lysine (PLL), poly-α-L -ornithine (PLO), and poly(α,γ-L -diaminobutyric acid) (PLDBA) have been investigated as functions of pH at 25°C and of temperature at pH 11.75, where these polymers are uncharged; in the cases of the latter two polyamino acids, the transitions have also been studied as functions of apparent pH in methanol-water solution (50/50 by volume). The helix stability of the polypeptides is shown to be a direct function of the number of methylene groups on the side chain. From an analysis of potentiometric titration data, we find that the difference between the helix stability of PLL and that of PLO is due to a difference of about 1 e.u. in the ΔS° of the transition. Combining the “melting curves” obtained from optical rotatory dispersion studies with the potentiometric titration data permits evaluation of the initiation parameter Z (or 1/σ½) of the statistical mechanical theories for these transitions. The value obtained for Z in the case of uncharged aqueous PLO is ca. 35.  相似文献   

13.
Absorption spectra and induced CD have been measured on aqueous solutions of water-soluble porphyrins with α-helical poly(L -glutamic acid) or α-helical poly (L -lysine) at different mixing ratios. For the former, porphyrin is porphine-meso-tetra (4-N-methylpyridinium) (TMpyP), and for the latter, it is porphine-meso-tetra (4-benzenesulfonate) (TPPS) or porphine-meso-tetra(4-benzoate) (TPPC). All the solutions of porphyrin-polypeptide systems show hypochromism in the Soret band and induced CD in the Soret region. The CD is characterized by a positive band at a shorter wavelength and a stronger negative band at a longer wavelength. The hypochromicity and the magnitude of molar ellipticities are much larger for the TPPS– and TPPC–poly (L -lysine) systems than for the TMpyP–poly (L -glutamic acid) system. Porphyrin ions bind to the α-helix electrostatically, and the two components of the Soret transition of porphyrin are subject to dissymmetric perturbation. TMpyP ions bind to the α-helix at isolated sites, while TPPS ions and TPPC ions are in pairs on the α-helix, that is, two ions bind consecutively and dissymmetrically. In the TMpyP–poly (L -glutamic acid) system a single CD band is associated with each of the two components of the Soret transition, and these are of opposite sign. In the TPPS– and TPPC–poly (L -lysine) systems, a pair of positive and negative CD bands is associated with each of the two components, thus giving apparently a single pair of CD bands with a shoulder, owing to partial overlapping.  相似文献   

14.
R M Santella  H J Li 《Biopolymers》1977,16(9):1879-1894
Poly(Lys48, His52), a random copolypeptide of L -lysine (48%) and L -histidine (52%), was used as a model protein for investigating the effects of protonation on the imidazole group of histidines on protein binding to DNA. The complexes formed between poly(Lys48, His52) and DNA were examined using absorbance, circular dichroism (CD), and thermal denaturation. Although increasing pH reduces the charges on histidine side chains in the model protein, the protein still binds the DNA with approximately one positive charge per negative charge in protein-bound regions. Nevertheless, CD and melting properties of poly(Lys48, His52)-DNA complexes still depend upon the solution pH which determines the protonation state of imidazole group of histidine side chains. At pH 7.0, the complexes show two characteristic melting bands with a tm (46–51°C) for free base pairs and a tm (94°C) for protein-bound base pairs. The tm of the complexes is reduced to 90°C at pH 9.2, although at this pH there is still one lysine per phosphate in protein-bound regions. Presumably, the presence of deprotonated histidine residues destabilizes the native structure of protein-bound DNA. The binding of this model protein to DNA causes a red shift of the crossover point and both a red shift and a reduction of the positive CD band of DNA near 275 nm. This phenomenon is similar to that caused by polylysine binding. These effects, however, are greatly diminished when histidine side chains in the model protein are deprotonated. The structure of already formed poly(Lys48, His52)·DNA complexes can be perturbed by changing the solution pH. However, the results suggest a readjustment of the complex to accommodate charge interactions rather than a full dissociation of the complex followed by reassociation between the model protein and DNA.  相似文献   

15.
In carbonate buffer at pH 10.5, a transparent solution of poly(L -lysine HBr) was obtained up to fairly high concentration of 3 g/dl at room temperature. The hydrodynamic behavior of the solution has been studied by sedimentation analyses and viscosity measurements. A dimer form for high concentrations and a monomer form for low concentrations were inferred. The dimer and monomer forms were assigned to a β-structure and α-helix, respectively, based on the CD and optical rotary dispersion spectra. Using CD spectroscopy, a reversible transition between α-helix and β-structure was observed as a function of either poly(L -lysine HBr) concentration or temperature. An aggregated form which was assigned to the antiparallel pleated sheet appeared at 50°C on the basis of its ir spectrum.  相似文献   

16.
The synthesis and some of the physical-chemical properties of tricopolymers of L -glutamic acid, L -lysine, and L -alanine are reported here. The molar ratios of the glutamyl: lysyl: alanyl residues were 1:1:X or 3:2:X, where the alanyl content X was increased in regular steps. The α-helix content calculated from the optical rotatory dispersion of the polypeptides is compared with a predicted helix content estimated from the composition of the polymers and the known behavior of the homopolypeptides at pH 3, 8, and 12. At pH 3 copolypeptides containing 20 mole-% or more alanine exhibit a helix content equal to the sum of their alanyl and glutamyl residue contents. At pH 8 the helix content equals the alanyl content when the latter was 40 mole-% or higher; at lower alanyl contents the electrostatic interaction between charged glutamyl and lysyl residues makes some contribution. At pH 12 the amount of helix observed is proportional to the mole ratio of alanine residues present in the polymer. The helix content of a tricopolymer containing 1:1:3 mole ratios of glutamyl: lysyl: alanyl residues was determined in solutions of lithium bromide and in urea solutions. Both reagents led to a decrease in helix content at pH 3 and 8 to a minimum of approximately 20% helix in 8M urea or 5.5M LiBr. The helix–random chain transition curves at pH 3 and 8 are parallel when the urea concentration is varied, but differ in shape when the lithium bromide concentration is varied at pH 3 and 8. The mode of action of these two “denaturing” reagents may thus be different. Heating the same tricopolypeptide at pH 3 or 8 from 5 to 80°C. also led to a helix–random chain transition centered at approximately 45°C.  相似文献   

17.
H Yamamoto  J T Yang 《Biopolymers》1974,13(6):1109-1116
Uncharged poly(Nε-methyl-L -lysine) (PMLL) and its isomer, poly(Nδ-ethyl-L -ornithine) (PELO), in alkaline solution (pH ca. 12) undergo a helix-to-β transition upon mild heating at 50°C or higher in a manner similar to that of poly(L -lysine) (PLL). The rate of conversion follows the order: PMLL < PELO < PLL. The helix can be regenerated upon cooling near zero degrees, for instance, after more than 12 hr at 2°C. At concentrations less than 0.02% the β form is intramolecular, but at higher concentrations both intra- and intermolecular β forms are generated. Poly(Nδ-methyl-L -ornithine) (PMLO), an isomer of PLL, behaves like poly(L -ornithine); uncharged PMLO in alkaline solution is partially helical and becomes disordered at elevated temperatures.  相似文献   

18.
Water-soluble block copolymers of the type (A)m-(B)n-(A)p, where (A)m,p was either poly(D ,L -lysine-α,β,β,γ,γ,δ,δ-d7) or poly(D ,L -lysine) and (B)n was either poly(L -alanine) or poly(L -phenylalanine), were synthesized for conformational studies by proton magnetic resonance spectroscopy. Analytical determination of the amount of the initiator fragment (n-hexylamine) at the C-terminus of the copolymers was used to obtain the number-average degrees of polymerization, DP n, and thereby, together with the amino acid composition, to establish the covalent structures of the polymers. The values of DP n were found to be much lower than those deduced from sedimentation equilibrium or form viscosity measurements. These deviations, which also are thought to have arisen in similar studies reported in the literature, are attributable to intermolecular aggregation; the relation of such aggregation to covalent structure (and its effect on the polymerization reaction) is discussed in terms of the conditions and mechanism of synthesis of block copolymers of amino acids.  相似文献   

19.
S Higuchi  M Tsuboi 《Biopolymers》1966,4(8):837-854
The optical density–temperature profile of double-stranded poly(A + U), triple stranded poly(A + 2U), and double-stranded RNA from rice dwarf virus in solutions with and without poly-L -lysine has been examined. When poly-L -lysine is added, more than one melting temperature Tm is observed for poly(A + U) and poly(A + 2U). One of them is considered to correspond to the melting of the polynucleotide molecule free from poly-L -lysine, and another to the melting of a polynucleotide–poly-L -lysine complex. For rice dwarf virus RNA, the Tm assignable to the complex is not found to be lower than 99°C. In every case, however, the hyperchromicity observed at the Tm of the free poly-nucleotide molecule is lowered linearly as the amount of poly-L -lysine added to the solution increases. This fact is taken as indicating that there is a stoichiometric complex formed. The stoichiometric ratio lysine/nucleotide in each complex is determined by examining the relation between the amount of poly-L -lysine added to the solution and the percentage of hyperchromicity remaining at Tm of the free polynucleotide molecule. The ratio is found to be 2/3 for all of the three complexes. A discussion is given on the molecular conformations of four types of polynucleotide–polylysine complex hitherto found: (A) double-stranded DNA plus poly-L -lysine in which the lyslne/nucleotide ratio is 1, (B) three-stranded RNA [poly(A + 2U)] plus poly-L -lysine in which the ratio is 2/3, (C) double-stranded RNA [poly (A + U) or rice dwarf virus RNA] plus poly-L -lysine in which the ratio is 2/3, and (D) double-stranded RNA [poly(I + C)] plus poly-L -lysine in which the ratio is 1/2.  相似文献   

20.
The interaction between the negatively charged phospholipid DPPG and positively charged poly(l-lysine) (PLL) of different lengths was studied by X-ray scattering in the SAXS and WAXS region. As a reference pure DPPG (Na salt) was investigated over a wide temperature range (−30 to 70°C). The phase behavior of DPPG in aqueous and in buffer/salt dispersions showed a metastable subgel phase at low temperatures and a recrystallization upon heating before reaching the liquid–crystalline phase. The presence of additional salt stabilizes the bilayer structure and decreases the recrystallization temperature. Large changes in the SAXS region are not connected with changes in chain packing. In DPPG/PLL samples, the PLL is inserted between adjacent headgroup layers and liberates counterions which give rise to a freezing point depression. In the complex with DPPG PLL form an α-helical secondary structure at pH 7 and temperatures below the gel to liquid–crystalline phase transition. This prevents DPPG from recrystallization and strongly increases the stacking order. The lamellar repeat distance is decreased and fixed by the helix conformation of PLL in the gel phase. PLL with n = 14 is too short to form helices and is squeezed out reversibly from the interbilayer space upon cooling by freezing of trapped water. In dispersions with longer PLLs (n > 400) at −20°C a 1D crystallization of PLL α-helices in the aqueous layer between the headgroups takes place. A structural model is presented for the lateral periodic complex, which is similar to the known cationic lipid/DNA complex. Dedicated to Prof. K. Arnold on the occasion of his 65th birthday.  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