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 0 毫秒
1.
Cu (II) — poly (L-arginine) (PLA) complexes have been studied using potentiometric titrations, optical absorption and circular dichroism spectra. Three different complexes have been observed. The first one (complex I) is formed up to pH 8 and results from the coordination of two guanidinium groups to the metal ion. The second and third complexes (complexes IIA and IIB) are formed between pH 8 and 11, in different proportions which are dependent on PLA: Cu molar ratio. In these two complexes two guanidinium groups and two peptide nitrogens participate as ligands around the copper ion.  相似文献   

2.
Cu(II)-poly(L-lysine) complexes have been studied using potentiometric titrations, optical absorption and circular dichroism spectra. As in the Cu(II)-poly(L-arginine) system studied previously potentiometric and spectral data consistently show that two types of complexes are formed. The first formed below pH 7.6 contains two amine nitrogens and two oxygen from water molecules at the corners of a square in which the metal occupies the center. The second is obtained at pH above 7.6 when the oxygen atoms are replaced by two adjacent peptide nitrogens.  相似文献   

3.
E Patton  H E Auer 《Biopolymers》1975,14(4):849-869
Poly(L -tyrosine) [(L -Tyr)n] has been characterized in aqueous solution using circular dichroism (CD) and infrared (ir) spectroscopy, and ultracentrifugal analysis. Most of the experiments were carried out at 0.01% polymer or less to avoid the complications caused by precipitation previously encountered by others. This permitted us to study solutions of (L -Tyr)n at lower pH values than had been attained previously. Our results show that a transition to an antiparallel-β conformation occurs at pH 11.32 upon titration from higher pH. The β structure is intramolecular when first formed and aggregates with time or upon titration below pH 11. Ultracentrifugal analysis of the intramolecular β conformation shows that it is quite compact, with a frictional coefficient ratio, f/fmin, of 1.09. In addition to the β structure, a nonordered form of the polymer has been obtained below pH 11 by rapid titration of the ionized polyelectrolyte. This form is nonaggregated and was found to have an f/fmin of 1.01, and is therefore almost spherical. The aggregated β form was found to be thermodynamically more stable than the nonordered form at pH 10.7.  相似文献   

4.
Further direct evidences are given that a clear correlation exists between potentiometric and spectroscopic measurements in monitoring the poly(L-glutamic, acid) helix+ coil transtition. Specific Li+ ion poly (L -glUtamic acid) interactions have been observed, suresting that Li+ ions may exert a distinct destabilizing action on the helical conformation of the polyelectrolyte.  相似文献   

5.
The interaction between ionizable carboxyl groups and the conformation of poly-(glutamic acid) (PGA) in aqueous solution were investigated by the mechanical method. The dynamic rigidity of the PGA solution has a maximum value at the pH corresponding to about 50% neutralization point. This may be due to establishing of a maximum attractive force by proton/charge fluctuation between ionizable carboxyl groups at that pH. The dynamic viscosity has a sharp change in the region of pH 5.5–6.5. It is suggested that this behavior is due to the helix–coil transition.  相似文献   

6.
Aqueous gallium(III) citrate complexes have been studied in the 10(-2) M concentration range with extended X-ray absorption fine structure (EXAFS) and FTIR techniques. From EXAFS data, one mononuclear and one oligomeric species were identified at different Ga(III) to citrate ratios. The first shell of the mononuclear complex was found to be distorted, with average Ga-O bond lengths of 1.95 and 2.06 A, in agreement with the solid-state structure of Ga(Cit)2(3-) (Cit=citrate). Also the oligomeric species was found to have a distorted first shell, with average Ga-O bond lengths of 1.95 and 2.04 A. This complex was found to contain two Ga-Ga distances at 3.03 and 3.56 A, typical for edge and corner sharing GaO6 octahedra, respectively. The gallium(III) and aluminum(III) citrate systems were compared by means of FTIR, and were found to be analogous. The IR results suggest that the bond lengths derived from EXAFS for the 1:2 gallium(III) citrate complex also provide a good estimate of the corresponding distances in the mononuclear 1:1 complex. Direct coordination of citrate to the metal ions in the oligomeric gallium(III) citrate complex was indicated from both EXAFS and IR results, and this complex is stoichiometrically analogous to the Al3(H-1Cit)3(OH)(H2O)4- complex, which has been structurally determined. However, while the formation of the aluminum trimer has been shown to be slow, the gallium trimer was significantly more labile with a rate of formation indicated to be in the order of seconds or faster.  相似文献   

7.
Here we present a novel NMR method for the structure determination of calcium-calmodulin (Ca2+-CaM)-peptide complexes from a limited set of experimental restraints. A comparison of solved CaM-peptide structures reveals invariability in CaM’s backbone conformation and a structural plasticity in CaM’s domain orientation enabled by a flexible linker. Knowing this, the collection and analysis of an extensive set of NOESY spectra is redundant. Although RDCs can define CaM domain orientation in the complex, they lack the translational information required to position the domains on the bound peptide and highlight the necessity of intermolecular NOEs. Here we employ a specific isotope labeling strategy in which the role of methionine in CaM-peptide interactions is exploited to collect these critical NOEs. By 1H, 13C-labeling the methyl groups of deuterated methionine against a 2H, 12C background, we can acquire a 13C-edited NOESY characterized by simplified, easily analyzable spectra. Together with measured CaM backbone HN-N RDCs and intrapeptide NOE-based distances, these intermolecular NOEs provide restraints for a low temperature torsion-angle dynamics and simulated annealing protocol used to calculate the complex structure. We have applied our method to a CaM complex previously solved through X-ray crystallography: Ca2+-CaM bound to the CaM kinase I peptide (PDB code: 1MXE). The resulting structure has a backbone RMSD of 1.6 Å to that previously published. We have also used this test complex to investigate the importance of homologous model selection on the calculated outcome. In addition to having application for fast complex structure determination, this method can be used to determine the structures of difficult complexes characterized by chemical shift overlap and broad signals for which the traditional method based on the use of fully 13C, 15N-labeled CaM fails.  相似文献   

8.
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   

9.
The absorption and circular dichroism spectra of the 1:1 copolymer (L -Lys, L -Tyr)n have been investigated in aqueous solutions at pH ranging from 3 to 13. The spectral patterns indicate that the fully charged polympholyte assumes a nonperiodic conformation on the acid and basic sides of the isoelectric point. At pH ranging from 9.2 to 11.6, the polymer is largely ordered and takes mostly an antiparallel β-structure as is shown by the infrared spectra in D2O solutions. Moreover, the rotational strength of the La transition of tyrosyl is independent of the polymer conformation, whereas that of the Lb transition is strongly sensitive to it.  相似文献   

10.
Electronic absorption and emission spectra, along with lifetime measurements and vibrational spectra, are used to investigate the interaction between nitrate and trivalent europium ions in dilute solutions in anhydrous and aqueous acetonitrile. Upon addition of increasing quantities of nitrate, the complexes [Eu(NO3)n](3?n)+, with n = 1–5, form quantitatively in anhydrous acetonitrile. In solution, the pentanitrato species is not further solvated and its spectroscopic properties are similar to those of solid samples, indicating a similar structure with five bidentate nitrates bonded to the 10-coordinate Eu(III) ion. The lifetimes of the 5D0 level are 1.35(5) and 1.25(5) ms for Eu(NO3)3 and (Me4N)2Eu(NO3)5 0.05 M in CH3CN. The quantum yield of Eu(NO3)3 in CH3CN is 27.4%.The addition of small quantities of water to Eu(NO3)3 solutions does not result in the dissociation of the nitrate ions, provided Rw = [H2O]t/[Eu3+]t is smaller than 8; the apparent equilibrium rations for [Eu(NO3)3(H2O)n] are K3 = 40 ± 15 M?1 and K4 = 9 ± 3 M?1; K1 and K2 are too large to be determined. The formation of nitrato complexes is studied in mixtures containing increasing amounts of water and nitrate. Deconvolution of the different components of the 5D07F0 transition allows a semi-quantitative estimate of the relative concentration of the nitrato complexes. The total number of coordinated nitrate ions per europium ion can be determined on the basis of fluorescence lifetime measurements. The apparent equilibrium ratios for the formation of the mono- and dinitrato species amount to K1 = 23 ± 3, 15 ± 5 and 5 ± 1 for Rw = 44, 94 and 304, respectively, and to K2 = 17 ± 8 for Rw = 44 and 94.  相似文献   

11.
Two sequential polypeptides, poly(O-benzyl-L -Tyr-γ-benzyl-L -Glu-L -Ala-Gly) and poly(ε-benzyloxycarbonyl-L -Lys-L -Glu-L -Ala), were synthesized, the former by the pentachlorophenyl ester of the tetrapeptide monomer and the latter by the azide of the tripeptide monomer. After deprotection and dialysis, poly(L -Tyr-L -Glu-L -Ala-Gly) was obtained in 71% yield and had a molecular weight of 53,000. The circular dichroism spectra (CD) of the polymer at pH's 7.2, 10.5, and 11.8 and of oligomers and of the monomer at pH 7.2 indicated that the polymer exists in an α-helical conformation. After deprotection, poly(L -Lys-L -Glu-L -Ala) was obtained in 37% yield and had a molecular weight of 3000. The CD spectra of the polymer at pH 7.2 and 2.8, and of the monomer at pH 7.2, indicated that the polymer is in a randomly coiled configuration.  相似文献   

12.
Complexes of the general structure cis-[PtX(2)(hydrazide)(2)] and cis-[PtX(2)NH(3)(hydrazide)], where X=Cl(-), Br(-) and I(-), and hydrazide=cyclohexylcarboxylic acid hydrazide (chcah), cyclopentylcarboxylic acid hydrazide (cpcah), 3-aminocyclohexanspiro-5-hydantoin (achsh) and 3-aminocyclopentanspiro-5-hydantoin (acpsh), were investigated with respect to aqueous stability, DNA platination rates and cytotoxic activity on a panel of seven human cancer cell lines as well as a cisplatin-resistant cell line. Stabilities in aqueous solution, determined by RP-HPLC and UV-Vis methods, were highly dependent on the type of halide ligand, with stability decreasing in the order I(-)>Cl(-)>Br(-). Added chloride (100 mM) only stabilized the dichloro-Pt(II) complexes containing the hydrazide as part of a hydantoin ring (i.e., achsh). Platination of calf thymus DNA determined by AAS was most rapid with dichloro-Pt(II) complexes containing achsh ligand. The mixed-amine dichloro-Pt(II) complexes with either chcah or cpcah ligands also platinated DNA >80%, but at a slower rate, while dihydrazide dichloro-Pt(II) complexes with either chcah or cpcah ligands resulted in <25% DNA platination at 24 h. cis-[PtX(2)(hydrazide)(2)], where hydrazide=chcah or cpcah, were the most potent compounds (chcah>cpcah), but activity was independent of the halide ligand (I(-)=Cl(-)=Br(-)). These complexes showed no cross-resistance with cisplatin, but they also showed little differentiation in potency over the seven cell lines. Complexes with the hydantoin ligands achsh and acpsh were inactive in all cell lines. Thus, neither stability in aqueous media nor covalent binding to DNA are correlated with biological activity, suggesting that cis-dihydrazide Pt(II) complexes act by a unique mechanism of action.  相似文献   

13.
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   

14.
The complexes formed in the dimethylthallium(III) (Me2Tl+), glutathionate (EGC3−) and hydrogen ion system in aqueous solution at 37 °C and I = 150 mmol dm−3 (NaCl) have been characterised by means of glass-electrode potentiometry. Glutathione protonation constants were found to be 9.123 ± 0.007, 17.42 ± 0.01, 20.78 ± 0.02, and 22.93 ± 0.02. Formation constants for the complexes [(Me2Tl)EGCH] and [(·Me2Tl) EGC]2− were found to be 11.19 ± 0.03 and 2.39 ± 0.02, respectively. Particular attention has been paid to the evaluation of the effect of possible systematic errors on the constant values determined. Reliable standard deviation estimations have been made by applying a Monte Carlo calculation technique.  相似文献   

15.
The structure of noncrystalline sucrose in the amorphous, solid state and in aqueous solution was investigated. Differences of structure of amorphous solid samples, the quenched-melt, and freeze-dried sucrose, are revealed by differential thermal analysis (d.t.a.) and from the Fourier-transform infrared (F.t.-i.r.) spectra. Factor analysis of the F.t.-i.r. spectra of aqueous solutions of sucrose shows the existence of at least two forms of the sucrose molecule. Analysis of 13C-n.m.r. spectra of amorphous and crystalline sucrose reveals a sensitivity of the fructosyl moiety to the morphology of the sample.  相似文献   

16.
Spectroscopic characterization of poly(Glu-Ala)   总被引:3,自引:0,他引:3  
Infrared linear dichroism and ultraviolet circular dichroism speetroscopy have been used to distinguish four conformational forms of the ionizable sequential polypeptide poly(Glu-Ala). Two of these conformations, the α helix and the β form, were observed for the unionized polypeptide in solution. The α helix appeared immediately upon neutralization of the side-chain carboxyl functions, whereas the β form was observed after the neutralized solution had been standing for several days. The β form was also observed for films cast from either high or low pH solutions. Ionization of the glutamyl residues resulted in a circular dichroism spectrum which has previously been observed for charged homopolymers and appears to result from an extended helical conformation. Further, heating either the α helical or the charged extended helix resulted in a transition to a disordered chain. These results are consistent with the results of conformational calculations presented elsewhere.  相似文献   

17.
In aqueous solution, it was found that the amphiphilic copolymer poly(ethylene glycol)-b-poly(caprolactone) (PEG(5000)-b-PCL(4100)) formed different morphologies, including long rod-like, short rod-like, or spherical aggregates, when the copolymer concentration was increased. Nearly identical morphologies were observed with the addition of increasing amounts of PEG(2000)-distearoylphosphoethanolamine (PEG(2000)-DSPE) to the copolymer. The morphologies of the aggregates in solution were confirmed by negative stain transmission electron microscopy (TEM) and cryogenic-TEM (cryo-TEM). The critical micelle concentrations of the PEG(5000)-b-PCL(4100) copolymer, PEG(2000)-DSPE and a mixture of the two materials (PEG(5000)-b-PCL 4100/PEG(2000)-DSPE) were evaluated to determine the thermodynamic stability of the aggregates. Differential scanning calorimetry was performed to gain insight into the degree of mixing of PEG(5000)-b-PCL(4100) and PEG(2000)-DSPE. Overall, combining PEG(5000)-b-PCL(4100) and PEG(2000)-DSPE produced a single population of mixed micelles with rod-like or spherical morphologies depending on the material composition and concentration.  相似文献   

18.
Bis (difluoroboron - α - furilglyoximato) nickel (II), C20H12O8N4B2F4Ni, was prepared by cyclization of its hydrogen-bonded precursor with BF3·OEt2. The compound crystallizes in the space group P21/c with a = 11.162(2), b = 5.569(2), c = 19.527(3) Å, β = 100.08(1)°, U = 1195.1(3) Å3, and Z = 2. The structure was refined to an R value of 0.033 using 2371 unique reflections collected with a CAD4-SDP diffractometer system. Unlike the corresponding planar macrocyclic as well as hydrogen-bonded dimethylglyoximates, the title compound neither dimerizes not exhibits columnar stacked structure. The 14-member macrocycle is planar except the B atoms, and no metal-metal interactions are observed in this compound. The complexation and cyclization reactions were investigated using spectral data. The structure is compared with other macrocyclic complexes.  相似文献   

19.
Molecular complexes of the types (Urd)H(x)(PA) and (UMP)H(x)(PA) are formed in the uridine (Urd) or uridine 5'-monophosphate (UMP) plus spermidine or spermine systems, as shown by the results of equilibrium and spectral studies. Overall stability constants of the adducts and equilibrium constants of their formation have been determined. An increase in the efficiency of the reaction between the bioligands is observed with increasing length of the polyamine. The pH range of adduct formation is found to coincide with that in which the polyamine is protonated while uridine or its monophosphate is deprotonated. The -NH(x)(+) groups from PA and the N(3) atom of the purine base as well as phosphate groups from the nucleotides have been identified as the significant centres of non-covalent interactions. Compared to cytidine, the pH range of Urd adduct formation is shifted significantly higher due to differences in the protonation constants of the endocyclic N(3) donor atoms of particular nucleosides. Overall stability constants of the Cu(II) complexes with uridine and uridine 5'-monophosphate in ternary systems with spermidine or spermine have been determined. It has been found from spectral data that in the Cu(II) ternary complexes with nucleosides and polyamines the reaction of metallation involves mainly N(3) atoms from the pyrimidine bases, as well as the amine groups of PA. This unexpected type of interaction has been evidenced in the coordination mode of the complexes forming in the Cu-UMP systems including spermidine or spermine. Results of spectral and equilibrium studies indicate that the phosphate groups taking part in metallation are at the same time involved in non-covalent interaction with the protonated polyamine.  相似文献   

20.
Hu Y  Zhang L  Cao Y  Ge H  Jiang X  Yang C 《Biomacromolecules》2004,5(5):1756-1762
Poly(epsilon-caprolactone)-b-poly(ethylene glycol)-b-poly(epsilon-caprolactone) triblock copolymers were synthesized by the ring-opening polymerization of epsilon-caprolactone in the presence of hydroxyl-terminated poly(ethylene glycol) with different molecular weights, using stannous octoate catalyst. Micelles prepared by the precipitation method with these triblock copolymers exhibit a core-shell structure. The degradation behaviors of these core-shell micelles in aqueous solution were investigated by FT-IR, 1H NMR, GPC, DLS, TEM, and AFM. It was found that the degradation behavior of micelles in aqueous solution was quite different from that of bulk materials. The size of the micelles increased in the initial degradation stages and decreased gradually when the degradation period was extended. The caprolactone/ethylene oxide (CL/EO) ratio in micelles measured by NMR also shows an increase at the initial degradation stage and a decrease at later stages. The morphology of these micelles became more and more irregular during the degradation period. We explain the observed behavior by a two-stage degradation mechanism with interfacial erosion between the cores and the shells followed by core erosion.  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