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.
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.
Poly(Nε-acetoacetyl-L -lysine), poly(Nδ-acetoacetyl-L -ornithine) and poly(Nγ-acetoacetyl-L -diaminobutyric acid) form colored complexes with ferric ions in water/dioxane solutions. These complexes are soluble at pH values lower than 2.8 and show their maximum absorption at 257 nm in the uv and at 478 nm in the visible region; whereas the ferric complex of the model compound n-hexylacetoacetamide exhibits absorptions centered at 258 and 536 nm, respectively. It is shown that in the complex of the model compound one metal ion is bound per acetoacetamide group, while in the complexes of the three polymers two β-ketoamides side chains are bound per ferric ion under the same solvent, pH, concentration, and ionic strength conditions. The binding constants of ferric ions to the three polymers, and the formation constant of the ferric complex of the model compound are also evaluated.  相似文献   

3.
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   

4.
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   

5.
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   

6.
A 13C-nmr study of the salt-induced helix–coil transition of the basic polypeptides poly(L -lysine) [(Lys)n], poly(L -arginine) [(Arg)n], and poly (L -ornithine) [(Orn)n] was performed to serve as a reference of the helical portion of histones and other proteins. As is the case with pH-induced helix–coil transition, the downfield displacement of the Cα and carbonyl carbon signals are observed in the helical state. The upfield shift of the Cβ signals, on the other hand, is noted in the salt-induced transition. Regardless of the differences in the side chains and also the salts used, very similar helix-induced chemical shifts are obtained for (Lys)n and (Arg)n. However, the displacement of the Cα, Cβ, and carbonyl carbons of (Orn)n in the presence of 4M NaClO4 is found to be almost 50% of that of (Lys)n and (Arg)n. This is explained by the fact that the maximum helical content is about 50%, consistent with the ORD result. Further, the motion of the backbone and side chains of the helical from was estimated by measuring the spin-lattice relaxation time (T1), nuclear Overhauser enhancement (NOE), and line width. In the case of (Lys)n, the motion of the side chains is charged very little in comparison with that of the random coil. Indicating that the aggregation of the salt-induced helix is small in contrast to that of the pH-induced helix. For (Arg)n, however, the precipitate of the helical polymers is mainly due to aggregation.  相似文献   

7.
The interaction of CuCl2 with poly(S-carboxymethyl-L -cysteine) (poly[Cys(CH2COOH)]) and poly(S-carboxyethyl-L -cysteine) (poly[Cys(C2H4COOH)]) were studied by absorption spectra and circular dichroism (CD). On mixing CuCl2 with polypeptide solutions, absorption bands appeared at 320–325 nm in both polypeptides, and at 255–260 nm in the case of poly[Cys(CH2COOH)]. A stable bound species was formed in the case of poly[Cys(CH2COOH)], since the apparent molar absorption coefficient of the bound species did not depend on the mixing ratio. From the absorption data, it was inferred that Cu2+ ions were complexed with the side chains, most probably with sulfur atoms and carboxyl groups. Induced optical activities were observed for the two polypeptides. The CD spectra of poly[Cys(CH2COOH)] + CuCl2 gave simpler aspects than those of poly[Cys(C2H4COOH)] + CuCl2.  相似文献   

8.
M. Branca  M. E. Marini  B. Pispisa 《Biopolymers》1976,15(11):2219-2226
The binding process between sodium poly(L -glutamate) and trans-2,2′,2″,2?-tetrapyridyl-Fe(III) complex ions in aqueous solution at pH around 7 has been studied by means of equilibrium dialysis and optical measurements. The binding isotherm indicates the occurrence of a cooperative process, whereby bound molecules facilitate the association of additional molecules. According to circular dichroism (CD) data, this effect is coupled with that which sees a conformational change in the charged polypeptide upon progessive binding of complex counterions. All these features are discussed in the light of the structural characteristics of the interacting species. A stereochemical model of the association “complex” is proposed.  相似文献   

9.
The fixation of trans-(NH3)2Cl2 Pt(II) to poly(I)·poly(C) at low rb (< 0.05) leads to the formation of two complexed species. The major species (ca. 82% of bound platinum) involves coordination of platinum to a single hypoxanthine base, while the other species involves coordination of two hypoxanthine bases, which are either far apart on the same strand or on separate poly(I) strands, to the platinum. These same two species are found after reaction with poly(I), as are two other species throughout the entire rb range studied (rb = 0–0.30). The latter two species are assigned to trans-Pt bound to two bases on a poly(I) strand with (a) one or (b) two free bases between the two bound bases. These two species, (a) and (b), account for ca. 35% of the bound platinum, although the 1:1 species remains dominant (ca. 55%). These two additional species are observed at high rb (>0.075) after reaction with poly(I)·poly(C) but as very minor species. They are formed by reaction with melted poly(I) loops. Also at high rb, we have observed a shifted cytidine H5 resonance arising from interaction of trans-Pt with a melted loop of poly(C). Most probably, this arises from an intramolecular poly(I) to poly(C) crosslink. Results from the reaction of trans-Pt with poly(C) are presented for comparison.  相似文献   

10.
H Inoue  T Izumi 《Biopolymers》1976,15(4):797-812
The preferential binding of solvent components with a nonionic homopolypeptide, poly(N5-(3-hydroxypropyl)-L -glutamine), ([Gln((CH2)3OH)]n), has been determined in water/dioxane mixtures using differential refractometry. The degree of preferential binding was calculated from the difference between the refractive index increments of [Gln((CH2)3OH)]n obtained from experiments carried out under two conditions: experiments where the molality of dioxane was kept identical in both compartments of the differential cell, and experiments where the chemical potential was kept identical. The polypeptide was preferentially hydrated between 10 and 70 wt % of dioxane; the amount of preferential hydration per gram of the mixed solvent increases monotonically (with a plateau region between 40 and 60 wt %) with the dioxane concentration. A monotonic increase was also observed in the degree of helicity of the polypeptide. The absolute amounts of water and dioxane bound by [Gln((CH2)3OH)]n were investigated in the frozen state by the method of nuclear magnetic resonance. Hydration was measured using a mixed solvent, water/dioxane-d8; dioxane solvation was measured using a mixed solvent, dioxane/D2O. The polypeptide binds about 0.35 g of water per g of the polymer in aqueous solution, and hydration decreases gradually with an increase in dioxane concentration. On the other hand, the amount of dioxane solvation increases to 0.04 g per g of the polymer in the dioxane concentration range between 0 and 20 wt %, and then levels off. The rapid increase in solvation is observed before the conformational transition from random coil to α-helix occurs in [Gln((CH2)3OH)]n. The dependence of the preferential and absolute binding of solvent components to [Gln((CH2)3OH)]n on dioxane concentration and the conformational change in the homopolypeptide suggest that addition of dioxane to aqueous solutions induces lowering of water activity and that the helical structure of the polypeptide is enhanced by the formation of intrachain hydrogen bonds. The validity of the frozen method is also discussed.  相似文献   

11.
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   

12.
Y C Fu  H V Wart  H A Scheraga 《Biopolymers》1976,15(9):1795-1813
The enthalpy change associated with the isothermal pH-induced uncharged coil-to-helix transition ΔHh° in poly(L -ornithine) in 0.1 N KCl has been determnined calorimetrically to be ?1530 ± 210 and ?1270 ± 530 cal/mol at 10° and 25°C, respectively. Titration data provided information about the state of charge of the polymer in the calorimetric experiments, and optical rotatory dispersion data about its conformation. In order to compute ΔHh°, the observed calorimetric heat was corrected for the heat of breaking the sample cell, the heat of dilution of HCl, the heat of neutralization of the OH? ion, and the heat of ionization of the δ-amino group in the random coil. The latter was obtained from similar calorimetric measurements on poly(D ,L -ornithine). Since it was discovered that poly(L -ornithine) undergoes chain cleavage at high pH, the calorimetric measurements were carried out under conditions where no degradation occurred. From the thermally induced uncharged helix–coil transition curve for poly(L -ornithine) at pH 11.68 in 0.1 N KCl in the 0°–40°C region, the transition temperature Ttr and the quantity (?θh/?T)Ttr have been obtained. From these values, together with the measured values of ΔHh°, the changes in the standard free energy ΔGh° and entropy ΔGh°, associated with the uncharged coil-to-helix transition at 10°C have been calculated to be ?33 cal/mol and ?5.3 cal/mol deg, respectively. The value of the Zimm–Bragg helix–coil stability constant σ has been calculated to be 1.4 × 10?2 and the value of s calculated to be 1.06 at 10°C, and between 0.60 and 0.92 at 25°C.  相似文献   

13.
The binding of the methylmercury cation CH3Hg+ by poly(L -glutamic acid) (PGA) and by poly(L -lysine) (PLL) has been investigated by Raman spectroscopy. Coordination on the side-chain COO? and NH groups of these polypeptides gave characteristic ligand–Hg stretching modes at ca. 505 and 450 cm?1, respectively. Precipitation generally occurred upon formation of the complexes and changes of conformation were common. The solid complex obtained from PGA at pH 4.6 was found to have a mostly disordered conformation, which differed from the respective α-helical and β-sheet structures of the dissolved and precipitated uncomplexed polypeptide in the same conditions. An α-helical structure was generally adopted by the complex formed with PLL, even in pH and temperature conditions where the free polypeptide normally exists in another conformation. The addition of a stronger complexing agent, glutathione, to the PLL/CH3Hg+ complex caused a migration of the bound cations and a restoration of the polypeptide to its original state.  相似文献   

14.
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   

15.
Poly(Nε-trimethyl-L -lysine), [Lys(Me3)]n, and poly(Nδ-trimethyl-L -ornithine), [Orn(Me3)]n, in sodium dodecylsulfate do not assume the β-structure or α-helix, respectively, of their parent polymers. In 0.5M Ca(ClO4)2 both [Lys(Me3)]n and [Orn(Me3)]n are aggregated and display CD spectra indicative of a regular, perhaps helical, structure. For [Lys]n and [Lys(Me3)]n, the T1 of the α-hydrogens are 0.379 and 0.230 sec, respectively, indicating greater rigidity for [Lys(Me3)]n. The CD spectrum of [Lys(Me3)]n at pH 8 is more heat resistant than that of [Lys]n. It is suggested that apolar interactions are more important in the methylated polymers than in the parent polymers.  相似文献   

16.
Poly(L -arginine) assumes the α-helix in the presence of the tetrahedral-type anions or some polyanions by forming the “ringed-structure bridge” between guanidinium groups and anions which is stabilized by a pair of hydrogen bonds and electrostatic interaction [Ichimura, S., Mita, K. & Zama, M. (1978) Biopolymers 17 , 2769–2782; Mita, K., Ichimura, S. & Zama, M. (1978) Biopolymers 17 , 2783–2798]. This paper describes the parallel CD studies on the conformational effects on poly (L -homoarginine) of various mono-, di-, polyvalent anions and some polyanions, as well as alcohol and sodium dodecylsulfate. The random coil to α-helix transition of poly(L -homoarginine) occurred only in NaClO4 solution or in the presence of high content of ethanol or methanol. The divalent and polyvalent anions of the tetrahedral type (SO, HPO, and P2O), which are strong α-helix-forming agents for poly(L -arginine), failed to induce the α-helical conformation of poly(L -homoarginine). By complexing with poly(L -glutamic acid) or with polyacrylate, which is also a strong α-helix-forming agent for poly(L -arginine), poly(L -homoarginine) only partially formed the α-helical conformation. Monovalent anions (OH?, Cl?, F?, and H2PO) did not change poly(L -homoarginine) to the α-helix, and in the range of pH 2–11, the polypeptide remained in an unordered conformation. In sodium dodecylsulfate, poly(L -homoarginine) exhibited the remarkably enlarged CD spectrum of an extended conformation, while poly(L -arginine) forms the α-helix by interacting with the agent. Thus poly(L -homoarginine), compared with poly(L -arginine), has a much lower ability to form the α-helical conformation by interacting with anions. The stronger hydrophobicity of homoarginine residue in comparison with the arginine residue would provide unfavorable conditions to maintain the α-helical conformation.  相似文献   

17.
The covalent binding of trans-Pt (NH3)2Cl2 to the double-stranded poly(I)·poly(C) follows three types of reactions, depending on rb and the concentration of polynucleotide in the reaction mixture. At rb ? 0.1, the principal reaction is coordination to poly(I), giving rise to some destabilization of the double strand, as shown by uv and CD spectra, and a decrease in Tm values, giving rise to free loops of poly(C). At higher rb and low polynucleotide concentration, the free cytidine bases react with platinum bound on the complementary strand to form intramolecular (interstrand) crosslinks that restabilize the double-stranded structure. At high rb and high polynucleotide concentration, while the above reaction still occurs, the predominant one is the formation of intermolecular crosslinks. Under no conditions has strand separation been observed.  相似文献   

18.
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   

19.
Abstract

Thermodynamic parameters of melting process (δHm, Tm, δTm) of calf thymus DNA, poly(dA)poly(dT) and poly(d(A-C))·poly(d(G-T)) were determined in the presence of various concentrations of TOEPyP(4) and its Zn complex. The investigated porphyrins caused serious stabilization of calf thymus DNA and poly poly(dA)poly(dT), but not poly(d(A-C))poly(d(G-T)). It was shown that TOEpyp(4) revealed GC specificity, it increased Tm of satellite fraction by 24°C, but ZnTOEpyp(4), on the contrary, predominately bound with AT-rich sites and increased DNA main stage Tm by 18°C, and Tm of poly(dA)poly(dT) increased by 40 °C, in comparison with the same polymers without porphyrin. ZnTOEpyp(4) binds with DNA and poly(dA)poly(dT) in two modes—strong and weak ones. In the range of r from 0.005 to 0.08 both modes were fulfilled, and in the range of r from 0.165 to 0.25 only one mode—strong binding—took place. The weak binding is characterized with shifting of Tm by some grades, and for the strong binding Tm shifts by ~ 30–40°C. Invariability of ΔHm of DNA and poly(dA)poly(dT), and sharp increase of Tm in the range of r from 0.08 to 0.25 for thymus DNA and 0.01–0.2 for poly(dA)poly(dT) we interpret as entropic character of these complexes melting. It was suggested that this entropic character of melting is connected with forcing out of H2O molecules from AT sites by ZnTOEpyp(4) and with formation of outside stacking at the sites of binding. Four-fold decrease of calf thymus DNA melting range width ΔTm caused by increase of added ZnTO- Epyp(4) concentration is explained by rapprochement of AT and GC pairs thermal stability, and it is in agreement with a well-known dependence, according to which ΔT~TGC-TAT for DNA obtained from higher organisms (L. V. Berestetskaya, M. D. Frank-Kamenetskii, and Yu. S. Lazurkin. Biopolymers 13, 193–205 (1974)). Poly (d(A-C))poly(d(G-T)) in the presence of ZnTOEpyp(4) gives only one mode of weak binding. The conclusion is that binding of ZnTOEpyp(4) with DNA depends on its nucleotide sequence.  相似文献   

20.
Structures of Cu(II) complexes of pyridoxal Schiff bases with poly(L -lysine), poly(L -ornithine), and poly(L -α,γ-diaminobutyric acid) were investigated by absorption spectra, CD, and conformational analysis. Although the polypeptides retain their typical right-handed α-helical conformation, opposite Cotton effects were found for the poly(L -lysine) and poly(L -ornithine) polycomplexes in the whole range of wavelengths from 600 to 250 nm. As in the analogous derivatives of salicyladehyde, this effect seems to be due to a stereospecific binding of the square planar Cu(II)-bis-pyridoxylideneimine group to the α-helical matrix. Circular dichroism spectrum of poly(L -α,γ-diaminobutyric acid) polycomplex is similar to that found for poly(L -lysine) derivative, but indicates large tetrahedral distortion of the square-planar coordination of copper ion.  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