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 62 毫秒
1.
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   

2.
The conformational properties of ferric complexes of poly(Nε-acetoacetyl-L -lysine), poly(Nδ-acetoacetyl-L -ornithine), and poly(Nγ-acetoacetyl-L -diaminobutyric acid) were investigated in 1:1 water/dioxane by CD techniques. Optical activity was found in the visible and in the uv absorption region of the polymeric complexes. The conformation of the peptide backbone was always that of a right-handed α-helix, and was found independent of the degree of complexation, at least up to a degree of binding of 20%. In the absorption region of the side-chain chromophores the optical activity is substantially affected by complex formation. In all three cases a splitting of the ligand π → π* transition centered at 257 nm is observed. These data suggest a stereospecific complex formation. From the signs of the splitting it also appears that the chirality of the poly(Nδ-acetoacetyl-L -ornithine) complex is opposite that of the other two polymers.  相似文献   

3.
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   

4.
Silver ions binding to poly(U) and poly(I) produce highly ordered multistranded helices under conditions which would otherwise lead to random coils. Evidence for helicity comes from the hypochromicity and high ellipticity generated in the polymers by Ag+ binding, as well as from x-ray studies and from the cooperativity of the Ag+ complexing reaction. Continuous variation studies show that both polymers form 1:1 and 2:1 polymer–Ag+ complexes. Low pH favors the 1:1 complex with poly(U) and the 2:1 complex with poly(I); the reverse is true at high pH. Ag+ binding and proton-release experiments make it clear that at low pH, unprotonated electron-donor groups are complexed preferentially, but that at high pH, Ag+ readily displaces H+ from protonated groups. In poly(I) the unprotonated donor is N(7), leading at low pH to a 2:1 complex containing N(7)-Ag-N(7) bonds; at high pH, proton release from N(1) leads to a 1:1 complex containing N(1)-Ag-O bonds. In poly(U) there is no unprotonated donor; the low-pH 1:1 complex involves deprotonation of only one N(3) per bound Ag+, leading to N3-Ag-O bonding, while high pH causes deprotonation of two N(3) per Ag+ and a 2:1 N(3)-Ag-N(3) complex. Thus silver ions react with the nucleotide bases in chemically predictable ways, and the formation of different Ag–nucleotide bonds leads to different multiple-helix structures.  相似文献   

5.
Studies on poly(L-lysine50, L-tyrosine50)-DNA interaction   总被引:3,自引:0,他引:3  
R M Santella  H J Li 《Biopolymers》1974,13(9):1909-1926
Interaction between poly(Lys50, Tyr50) and DNA has been studied by absorption, circular dichroism (CD), and fluorescence spectroscopy and thermal denaturation in 0.001M Tris, pH 6.8. The binding of this copolypeptide to DNA results in an absorbance enhancement and fluorescence quenching on tyrosine. There is also an increase in the tyrosine CD at 230 nm. The CD of DNA above 250 nm is slightly shifted to the longer wavelength which is qualitatively similar to, but quantitatively much smaller than, that induced by polylysine binding. At physiological pH the poly(Lys50, Tyr50)–DNA complex is soluble until there is one lysine and one tyrosine per nucleotide in the complex. The same ratio of amino acid residues to nucleotide has also been observed in copolypeptide-bound regions of the complex. The addition of more poly(Lys50, Tyr50) to DNA yields a constant melting temperature, Tm′, for bound base pairs at 90°C which is close to that of polylysine-bound DNA under the same condition. The melting temperature, Tm, of free base pairs at about 60°C on the other hand, is increased by 10°C as more copolypeptide is bound to DNA. As the temperature is raised, both absorption and CD spectra of the complexes with high coverage are changed, suggesting structural alteration, perhaps deprotonation, on bound tyrosine. The results in this report also suggest that intercalation of tyrosine in DNA is unlikely to be the mode of binding.  相似文献   

6.
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   

7.
The helix–coil transitions of poly(Nε-methyl, Nε-carbobenzoxy-L -lysine), poly(Nδ-methyl, Nδ-carbobenzoxy-L -ornithine), and poly(Nδ-ethyl, Nδ-carbobenzoxy-L -ornithine) in chloroform–dichloroacetic acid and their corresponding decarbobenzoxylated polypeptides in alkaline solutions were followed by optical rotation measurements. The introduction of a methyl or an ethyl group to the side chains of the carbobenzoxy derivatives of poly(L -lysine) and poly(L -ornithine) appeared to weaken the helical conformation at 25°C. The thermodynamic quantities of the three water-soluble polypeptides were calculated from the data on potentiometric titrations at several temperatures. For uncharged coil-to-helix transition, ΔH = ?370 cal/mol and ΔS = ?1.1 eu/mol for poly(Nε-methyl-L -lysine), and ΔH = ?540 cal/mol and ΔS = ?1.6 eu/mol for poly(Nδ-ethyl-L -ornithine) (all on molar residue basis). The absolute values of ΔH and ΔS dropped in the region of pH-induced transition and eventually both quantities became positive. The initiation factor σ was about 2 × 10?3, which was essentially independent of temperature. For poly(Nδ-methyl-L -ornithine) the coil-to-helix transition was not complete even when the polymer was uncharged at high pH.  相似文献   

8.
The Raman spectra of poly(L -lysine) with various structures, ionized poly(L -glutamic acid), and deuterated N-methylacetamides have been observed using visible and the 257.3-nm laser lines as the light source. Most of the Raman bands with significantly enhanced intensities in the uv-excited spectra of the polymers have been assigned to the vibrations associated with the C?O and C–N stretching modes, the amide I, II, III, I′, II′, and III′, with reference to the results obtained for simple amide molecules including the deuterated N-methylacetamides. Several amide frequencies have been newly identified and the structures of the polymers have been discussed through the comparison of the Raman and ir amide frequencies.  相似文献   

9.
Arlette Garnier  Lucia Tosi 《Biopolymers》1975,14(11):2247-2262
The study of Cu(II)–poly(L -arginine) complexes by potentiometric titration, as well as by optical, circular dichroism, and infrared spectra, provides information about the nature of ligands and the coordination sphere around the metal ion. Three different complexes have been identified. The first, which is formed below pH 8, contains two guanidinium nitrogens and two water molecules at the corners of the coordination square. The constant of the overall process as determined by the Gregor method equals 2.0 ± 0.1 × 10?9. The two other complexes form between pH 8 and 10.5 and they contain two guanidinium and two peptide nitrogens as nearest ligands. One of them is a monomer and the other probably a dimer, which differ in the symmetry of the coordination sphere around the cupric ion. The optical spectra of the three complexes show an absorption band at 260 nm that we have assigned to a charge-transfer transition between a σ metal nitrogen (amine) molecular orbital and a dx2?y2 metal orbital. The spectra of the two complexes containing peptide nitrogens exhibit another absorption band at 320 nm, which we have assigned to a charge transfer from a π orbital of the amide group to the dx2?y2 metal orbital.  相似文献   

10.
H Yamamoto  T Hayakawa  J T Yang 《Biopolymers》1974,13(6):1117-1125
Poly(Nδ-carbobenzoxy, Nδ-benzyl-L -ornithine) (PCBLO) was prepared by the standard NCA method. PCBLO was converted into poly(Nδ-benzyl-L -ornithine) (PBLO) through decarbobenzoxylation with hydrogen bromide. The monomer Nδ-benzyl-L -ornithine was synthesized by reacting L -ornithine with benzaldehyde, followed by hydrogenation. The conformation of the two polypeptides was studied by optical rotatory dispersion and circular dichroism. PCBLO forms a right-handed helix in helix-promoting solvents. In mixed solvents of chloroform and dichloroacetic acid (DCA) it undergoes a sharp helix–coil transition at 12% (v/v) DCA at 25°C, as compared with 36% for poly(Nδ-carbobenzoxy-L -ornithine) (PCLO). Like PCLO, the helix–coil transition is “inverse,” that is, high temperature favors the helical form. PBLO is soluble in water at pH below 7 and has a “coiled” conformation. In 88% (v/v) 1-propanol above pH (apparent) 9.6 it is completely helical. In 50% 1-propanol the transition pH (apparent) is about 7.4; this compares with a pHtr of about 10 for poly-L -ornithine in the same solvent.  相似文献   

11.
Transient electric dichroism has been measured for the ferriheme–poly(L -lysine)[(Lys)n], ferroheme–(Lys)n, and ferroheme–(Lys)n–carbon monoxide (CO) solutions at pH 9–12. The Soret absorption maximum in electronic spectrum (λ), the reduced linear dichroism (ρ) at complete orientation and the calculated angle (?) between the porphyrin plane of a bound heme and the oriented polymer axis are determined for the following complexes: a ferriheme–(Lys)n complex at pH 9.5–10.5 (λ = 420 nm, ρ = 0.50, and ? = 19°), a ferroheme–(Lys)n complex at pH 9.5–10.2 (λ = 432 nm, ρ = 0.77, and ? = 0°), and a ferroheme–(Lys)n–CO complex at pH 9.25 (λ = 430 nm, ρ = 0.38, and ? = 24°). Based on the above data, the validity of the structures of heme–(Lys)n complexes proposed by other investigators is discussed.  相似文献   

12.
The absorption spectra and circular dichroism (CD) have been measured for aqueous solutions of acridine orange of a constant concentration, [D] = 5 × 10?5M, mixed with poly(S-carboxyethyl-L -cysteine) in various mixing ratios, [P]/[D], ranging from 330 to 11, at different pH. The absorption spectra of the dye–polymer solutions are hypochromic, and the main band is located at 470 nm, accompanying a shoulder at 500 nm. At alkaline pH, no CD is induced in the visible region. At neutral and acidic pH, where the polymer is in the β-conformation, CD is induced in the visible and near-uv regions. A pair of CD bands is located at the region around 450 nm, when the pH is around the neutrality, while it appears at the region around 500 nm at acidic pH. Thus, the optically active species of bound dye changes from dimer to monomer on lowering the pH. These species form dissymmetric arrays along a polypeptide chain. The fraction of bound dye forming dissymmetric sequences is not high, but most of bound dye is adsorbed randomly on the ionized carboxyl groups of polypeptide chain and gives rise to hypochromism only. A dissymmetric structure of dye–polymer complexes is presented, in which the polymer has the β-conformation and the dye cations, either dimeric or monomeric, bind to its side chains, in such a way that the longer axes of molecular planes of bound dye form a two-fold, right-handed helix along the extended polypeptide chain. A zeroth-order calculation of CD based on the coupled oscillator model leads to the result that each dissymmetric array of dye consists, on the average, of two dimeric or monomeric cations. This low number of bound cations in a dissymmetric array and the large fraction of randomly adsorbed dye suggest that the hydrophobic interaction of dye with the polymer is strong, so that dye cations are adsorbed sparsely on both sides of the extended polypeptide chain.  相似文献   

13.
The interaction of silver ions with poly(A) was studied by potentiometric titration, uv spectrophotometry, and stopped-flow spectroscopy. For 0 < rb < 0.5, where rb is moles of silver ion bound per mole of nucleotide base, there exists only one type of binding for poly(A). Using McGhee's theory, the binding parameters, such as intrinsic binding constant, number of sites per nucleotide, and cooperativity, were determined from the potentiometric titration data. Using the stopped-flow method, one relaxation time was observed in 0 < r0 < 0.5, where r0 is the moles of silver ions added per mole of nucleotide base. The concentration dependences of the relaxation time suggest that the binding of silver ions to poly(A) proceeds through the following mechanism: where M is free silver ions, P the free binding sites on poly(A), and C and C′ are two forms of the complex. The nature of the binding of silver ions to poly(A) is also discussed.  相似文献   

14.
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   

15.
Iwao Satake  Jen Tsi Yang 《Biopolymers》1976,15(11):2263-2275
The binding isotherms of sodium decyl sulfate to poly(L -ornithine), poly(D ,L -ornithine), and poly(L -lysine) at neutral pH were determined potentiometrically. The nature of a highly cooperative binding in all three cases suggests a micelle-like clustering of the surfactant ions onto the polypeptide side groups. The hydrophobic interaction between the nonpolar groups overshadows the coulombic interaction between the charged groups. The titration curves can be interpreted well by the Zimm–Bragg theory. The average cluster size of bound surfactant ions is sufficiently large to promote the β-structure of (L -Lys)n even at a very low binding ratio of surfactant to polypeptide residue, whereas the onset of the helical structure for (L -Orn)n begins after about 7 surfactant ions are bound to two turns of the helix. The CD results are consistent with this explanation.  相似文献   

16.
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   

17.
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   

18.
Shewanella putrefaciens is a bacterium implicated in oil pipeline corrosion and fish spoilage, and is one of very few isolated microorganisms able to use iron(III) as an electron acceptor. S. putrefaciens strain 200 produced a novel cyclic dihydroxamate siderophore, putrebactin, during aerobic growth. Putrebactin was determined to be 1,11-dihydroxy-1,6,11,16-tetraazacycloeicosane-2,5,12,15-tetrone, a cyclic dimer of succinyl-(N-hydroxyputrescine), by 1H and 13C NMR spectroscopy, fast atom bombardment and chemical ionization mass spectrometry, and X-ray crystallography. The protonation constants of putrebactin were determined to be 8.82 and 9.71. Potentiometric titration of the ferric complex revealed a sharp equivalence point at 3.0 equivalents of base per mole of Fe(III), consistent with loss of 3 protons per equivalent of bound ferric ion, while Job's method of continuous variation supported a shift from 1 : 1 to 3 : 2 complex stoichiometry as a function of pH. Putrebactin is similar in structure to two other siderophores, bisucaberin and alcaligin, produced by unrelated bacteria. Received: 21 May 1996 / Accepted: 9 November 1996  相似文献   

19.
Tamarind fruit shell (TFS) was converted to a cation exchanger (PGTFS-SP-COOH) having a carboxylate functional group at the chain end by grafting poly(hydroxyethylmethacrylate) onto TFS (a lignocellulosic residue) using potassium peroxydisulfate-sodium thiosulfate redox initiator, and in the presence of N, N ′-methylenebisacrylamide as a cross-linking agent, followed by functionalization. The chemical modification was investigated using Fourier transform infrared (FTIR), X-ray diffraction (XRD), and potentiometric titrations. The feasibility of PGTFS-SP-COOH for the removal of heavy metals such as U(VI), Cu(II), Zn(II), and Co(II) ions from aqueous solutions was investigated by batch process. The optimum pH range for the removal of meal ions was found to be 6.0. For all the metal ions, equilibrium was attained within 2 h. The kinetic and isotherm data, obtained at optimum pH value 6.0, could be fitted with pseudo-second-order equation and Sips isotherm model, respectively. The Sips maximum adsorption capacity for U(VI), Cu(II), Zn(II), and Co(II) ions at 30°C was found to be 100.79, 65.69, 65.97, and 58. 81 mg/g, respectively. Increase of ionic strength decreased the metal ion adsorption. Different wastewater samples were treated with PGTFS-SP-COOH to demonstrate its efficiency in removing metal ions from wastewater. The adsorbed metal ions on PGTFS-SP-COOH can be recovered by treating with 1.0 M NaCl + 0.5 M HCl for U(VI) ions and 0.2 M HCl for Cu(II), Co(II), and Zn(II) ions. Four adsorption/desorption cycles were performed without significant decrease in removal capacity. The results showed that PGTFS-SP-COOH developed in this study exhibited considerable adsorption potential for the removal of U(VI), Cu(II), Zn(II), and Co(II) ions from water and wastewaters.  相似文献   

20.
Formation (affinity) constants for 1:1 complexes of N-(2-acetamido)iminodiacetic acid (ADAH2) with Ca(II), Mg(II), Mn(II), Zn(II), Co(II), Ni(II), and Cu(II) have been determined. Probable structures of the various metal chelates existing in solution are discussed. Values for the deprotonation of the amide group in [Cu(ADA)] and subsequent hydroxo complex formation are also reported. The use of ADA as a buffer is considered in terms of metal buffers complexes which can be formed at physiological pH, i.e., at pH 7.0 there is essentially no free metal ion in 1:1 M2+ to ADA solutions.  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