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.
Complex formation between ATP (adenosine 5′-triphosphate) and tn2COIII(aq) (tn = trimethylenediamine) and resulting hydrolysis of the ATP to ADP (adenosine 5′-diphosphate), AMP (adenosine 5′-monophosphate), PPi (pyrophosphate), and Pi (orthophosphate) have been examined by means of 31P nmr. With ATP ~0.1 M and tn2CoIII(aq) up to 0.3 M, complex formation was promoted by equilibrating solutions for a period at pH 4, after which hydrolysis was allowed to proceed at each of several pHs in the range 5 to 9 prior to quenching by addition of strong base. With ATP 0.01 M and tn2CoIII(aq) up to 0.08 M, the above procedure was followed in some cases; in other experiments the pH of each ATP/tn2CoIII(aq) solution was adjusted immediately to a value in the range 5 to 9 with the remainder of the procedure as before. In most cases the hydrolysis was at 25°C, but temperature dependence was also examined. The integrals for the β-phosphorus resonance have been used to analyze for ATP in the quenched solutions; independent measurements of ATP by an enzyme/spectrophotometric method (Bergmeyer) gave similar results. Cobalt to ATP molar ratios up to 1 produce tn2CoIIIATP as the predominant ATP complex; this 1:1 complex shows no detectable acceleration in hydrolysis compared to free ATP. Cobalt to ATP molar ratios of ?1 lead to complexes of type (tn2CoIII)2ATP and (tn2CoIII)3ATP, which exhibit greatly enhanced reactivity towards ATP hydrolysis. At a 2:1 molar ratio (0.1 or 0.01 M ATP), the enhancement is rate is ~105 at pH 7 where the rate is a maximum (comparison for 25°C); at higher molar ratios the rate enhancements are even greater. The results support the view that effective metal ion catalysis of ATP hydrolysis requires formation of reactive species involving more than one metal ion per ATP.  相似文献   

2.
New solid complex compounds of La(III), Ce(III), Pr(III), Nd(III), Sm(III), Eu(III) and Gd(III) ions with morin were synthesized. The molecular formula of the complexes is Ln(C15H9O7)3 · nH2O, where Ln is the cation of lanthanide and n = 6 for La(III), Sm(III), Gd(III) or n = 8 for Ce(III), Pr(III), Nd(III) and Eu(III). Thermogravimetric studies and the values of dehydration enthalpy indicate that water occurring in the compounds is not present in the inner coordination sphere of the complex. The structure of the complexes was determined on the basis of UV-visible, IR, MS, 1H NMR and 13C NMR analyses. It was found that in binding the lanthanide ions the following groups of morin take part: 3OH and 4CO in the case of complexes of La, Pr, Nd, Sm and Eu, or 5OH and 4CO in the case of complexes of Ce and Gd. The complexes are five- and six-membered chelate compounds.  相似文献   

3.
The 1H and 13C nmr spectra of Co(NH3)5ImH3+ and the 1H nmr spectra of αCotrien(ImH)23+ and βCotrien(ImH)23+ are reported. The pKa values determined from the dependence of the chemical shift on pH are 10.0, 9.6, and 10.1, respectively. The range of the chemical shift between the acid and base forms is unusually small in the 1H nmr, 0.5–0.7 ppm for the C-2 H and about 0.25 ppm for the C-4 H and C-5 H. In the 13C nmr, C-2 and C-4 have large shifts to low field and C-5 a small shift to high field on deprotonation. The C-2 proton is not exchanged with solvent 2H under acidic or basic conditions, in marked contrast to the corresponding proton in both imidazole and 1-methylimidazole. These spectroscopic and chemical properties should be useful for the direct identification of metal-ion coordinated histidines in proteins.  相似文献   

4.
The kinetics of binding of Cu (II), Tb (III) and Fe(III) to ovotransferrin have been investigated using the stopped-flow technique. Rate constants for the second-order reaction, k +, were determined by monitoring the absorbance change upon formation of the metal-transferrin complex in time range of milliseconds to seconds. The N and C sites appeared to bind a particular metal ion with the same rate; thus, average formation rate constants k + (average) were 2.4 × 104 M–1 s–1 and 8.3 × 104 M–1 S –1 for Cu (II) and Tb (III) respectively. Site preference (N site for Cu (II) and C site for Tb (III)) is then mainly due to the difference in dissociation rate constant for the metals. Fe (III) binding from Fe-nitrilotriacetate complex to apo-ovotransferrin was found to be more rapid, giving an average formation rate constant k + (average) of 5 × 105 M–1 s–1, which was followed by a slow increase in absorbance at 465 nm. This slow process has an apparent rate constant in the range 3 s–1 to 0.5 s–1, depending upon the degree of Fe (III) saturation. The variation in the rate of the second phase is thought to reflect the difference in the rate of a conformational change for monoferric and diferric ovotransferrins. Monoferric ovotransferrin changes its conformation more rapidly (3.4s–1) than diferric ovotransferrin (0.52 s–1). A further absorbance decrease was observed over a period of several minutes; this could be assigned to release of NTA from the complex, as suggested by Honda et al. (1980).Abbreviations Tf ovotransferrin - NTA nitrilotriacetate Jichi Medical School, School of Nursing, Yakushiji 3311-159, Minamikawachi, Tochigi, 329-04 Japan  相似文献   

5.
《Inorganica chimica acta》1986,117(2):187-189
The isolation and characterization of nine polymeric complexes of the general formula [M(L)1.5S2]n (where M is the metal ion, L the ligand and S the solvent, C2H5OH) of La(III) and Ce(III), Pr(III), Nd(III), Sm(III), Gd(III), Tb(III), Dy(III), Ho(III) with.the biologically active compound embelin using elemental and thermal analysis, infrared and electronic spectral studies is reported.  相似文献   

6.
Efforts to delineate the interactions of neurotoxic Al(III) with low molecular mass substrates relevant to neurodegenerative processes, led to the investigation of the pH-specific synthetic chemistry of the binary Al(III)-[N-(phosphonomethyl) iminodiacetic acid] (Al-NTAP), Al(III)-[nitrilo-tris(methylene-phosphonic acid)] (Al-NTA3P), and Al(III)-[1-hydroxy ethylidene-1,1-diphosphonic acid] (Al-HEDP) systems, in correlation with solution speciation studies. Reaction of Al(NO3)3·9H2O with NTAP at pH 7.0 and 4.0 afforded the new species (CH6N3)4[Al2(C5H6NPO7)2(OH)2]·8H2O (1) and (NH4)2[Al2(C5H6NPO7)2(H2O)2]·4H2O (2), while reaction of Al(NO3)3·9H2O with NTA3P led to K8[Al2(C3H6NP3O9)2(OH)2]·20H2O (3). Complexes 13 were characterized by elemental analysis, FT-IR, 13C, 31P, 1H NMR (for 12 solid state and solution NMR where feasible), and X-ray crystallography. The structures of 13 reveal the presence of uniquely defined dinuclear complexes of octahedral Al(III) bound to fully deprotonated phosphonate ligands, water and hydroxo moieties. The aqueous solution speciation studies on the aforementioned binary systems project a clear picture of the binary Al(III)–(carboxy)phosphonate interactions and species under variable pH-conditions and specific Al(III):ligand stoichiometry. The concurrent solid state and solution work (a) exemplifies essential structural and chemical attributes of soluble aqueous species, reflecting well-defined interactions of Al(III) with phosphosubstrates and (b) strengthens the potential linkage of neurotoxic Al(III) chemical reactivity toward O,N-containing (carboxy)phosphate-rich cellular targets.  相似文献   

7.
A series of complexes of Au(III) with nucleosides and nucleotides and their methyl derivatives in different stoichiometry have been prepared. Ultraviolet, visible, ir, and nmr studies have been performed to determine the site of binding of these ligands with the metal ion. In (1:4) Au(III): guanosine complex, N7 is the binding site, whereas at 1:1 complex, a bidentate type of chelation through C6O and N7 is observed. C6-NH2 is favored over N1 as coordinating site at all stoichiometry in the adenosine complex. Inosine binds through N1 at r = 1. In cytidine, N1 is the binding site, whereas thymidine reacts only at high pH. In the case of nucleotides a bidentate type of chelation through the phosphate and the ring nitrogen occurs. The phosphate binding ability of Au(III) was further confirmed by studying the interaction of Au(III) with dimethyl phosphate—a conformational analog of the phosphate backbone in DNA chain.  相似文献   

8.
X-ray crystallography shows that Pt(NH3)2(CBDCA) is a square-planar complex with the dicarboxylate chelate ring in the boat conformation and a planar cyclobutane ring. 1H and 13C nmr studies suggest that rapid chelate ring flipping occurs in solution. The value of 195Pt nmr combined with 15N labeling as an informative new method of studying carboxylate coordination is illustrated. nmr results are also reported for the analogous ethylmalonate complex.  相似文献   

9.
Fructokinase (Fraction III) of Pea Seeds   总被引:5,自引:4,他引:1       下载免费PDF全文
A second fructokinase (EC 2.7.1.4) was obtained from pea seed (Pisum sativum L. var. Progress No. 9) extracts. The enzyme, termed fructokinase (fraction III), was specific for fructose and had little activity with glucose. With fructose concentrations above 0.25 millimolar, there was strong substrate inhibition at the optimum pH (8.0) and also at pH 6.6. The apparent Km values at pH 8.0 for fructose and glucose were 0.06 millimolar and 0.14 millimolar, respectively. The apparent Km for Mg adenosine 5′-triphosphate (MgATP) was 0.06 millimolar and excess MgATP was inhibitory. Mg2+ was essential for activity but the enzyme was inhibited by excess Mg2+ or ATP. Mg adenosine 5′-pyrophosphate was also inhibitory. Activity was stimulated by the addition of monovalent cations: of those tested K+, Rb+, and NH4+ were the most effective. The possible role of fructokinase (fraction III) is discussed.  相似文献   

10.
The binding behavior of lysozyme with Al(III) is described using luminol as a luminescence probe by flow injection–chemiluminescence (FI–CL) analysis. It was found that the CL intensity of the luminol–lysozyme reaction could be markedly enhanced by Al(III), and the increase in CL intensity was linear with the Al(III) concentration over the range 0.3–30.0 pg mL?1, with a detection limit of 0.1 pg mL?1 (3σ). Based on the interaction model of lysozyme with Al(III), lg[(I ? I0)/(2I0 ? I)] = lgK + nlg[M], the binding constant K = 6.84 × 106 L mol–1 and the number of binding sites (n) = 0.76. The relative standard deviations were 3.2, 2.4 and 2.0% for 10.0, 20.0 and 30.0 pg mL?1 Al(III) (n = 7), respectively. This new method was successfully applied to continuous, quantitative monitoring of picogram level Al(III) in human saliva following oral intake of compound aluminum hydroxide tablets. It was found that Al(III) in saliva reached a maximum of 101.2 ng mL?1 at 3.0 h. The absorption rate constant ka, elimination rate constant k and half‐life time t1/2 of Al(III) were 1.378 h?1, 0.264 h?1 and 2.624 h, respectively. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

11.
In weak acid medium, aluminum(III) can react with chlorophosphonazo III [CPA(III), H8L] to form a 1:1 coordination anion [Al(OH)(H4L)]2‐. At the same time, proteins such as bovine serum albumin (BSA), lysozyme (Lyso) and human serum albumin (HSA) existed as large cations with positive charges, which further combined with [Al(OH)(H4L)]2‐ to form a 1:4 chelate. This resulted in significant enhancement of resonance Rayleigh scattering (RRS), second‐order scattering (SOS) and frequency doubling scattering (FDS). In this study, we investigated the interaction between [Al(OH)(H4L)]2‐ and proteins, optimization of the reaction conditions and the spectral characteristics of RRS, SOS and FDS. The maximum RRS wavelengths of different protein systems were located at 357–370 nm. The maximum SOS and FDS wavelengths were located at 546 and 389 nm, respectively. The scattering intensities (ΔI) of the three methods were proportional to the concentration of the proteins, within certain ranges, and the detection limits of the most sensitive RRS method were 2.6–9.3 ng/mL. Moreover, the chelate reaction mechanism or the reasons for the enhancement of RRS were discussed through absorption spectra, fluorescence spectra and circular dichroism (CD) spectra. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

12.
The kinetics and mechanism of a linear trihydroxamic acid siderophore (deferriferrioxamine B, H4DFB+) ligand exchange with Al(H2O)63+ to form mono(deferriferrioxamine B)aluminum(III) (Al(H2O)4H3DFB)3+ have been investigated at 25 °C over the [H+] range 0.001−1.0 M and I = 2.0 M (HClO4/NaClO4) by 27Al NMR. Kinetic results are consistent with Al(H2O)4(H3DFB)3+ formation and dissociation proceeding through a parallel path mechanistic scheme involving Al(H2O)63+(k2/k−1) and Al(H2O)5(OH)2+(k2/k−2) where k1 = 0.13 M−1 s−1, k−1 = 8.7 × 10−3 M−1 s−1, k2 = 2.7 × 103 M−1 s−1, and k−2 = 9.6 × 10−4 s−1. Relative complex formation rates at Al(H2O)63+ and Al(H2O)5OH2+, and comparison with kinetic data for a series of synthetic hydroxamic acids, suggest that an interchange mechanism is operative. These results are also discussed in relation to kinetic data for the corresponding iron(III)-deferriferrioxamine B system.  相似文献   

13.
The guanine-rich sequence, specifically in DNA, telomeric DNA, is a potential target of anticancer drugs. In this work, a mononuclear Fe(III) complex containing two meloxicam ligands was synthesized as a G-quadruplex stabilizer. The interaction between the Fe(III) complex and G-quadruplex with sequence of 5′-G3(T2AG3)3-3′ (HTG21) was investigated using spectroscopic methods, molecular modeling, and polymerase chain reaction (PCR) assays. The spectroscopic methods of UV–vis, fluorescence, and circular dichroism showed that the metal complex can effectively induce and stabilize G-quadruplex structure in the G-rich 21-mer sequence. Also, the binding constant between the Fe(III) complex and G-quadruplex was measured by these methods and it was found to be 4.53(±0.30)?×?105 M?1). The PCR stop assay indicated that the Fe(III) complex inhibits DNA amplification. The cell viability assay showed that the complex has significant antitumor activities against Hela cells. According to the UV–vis results, the interaction of the Fe(III) complex with duplex DNA is an order of magnitude lower than G-quadruplex. Furthermore, the release of the complex incorporated in bovine serum albumin nanoparticles was also investigated in physiological conditions. The release of the complex followed a bi-phasic release pattern with high and low releasing rates at the first and second phases, respectively. Also, in order to obtain the binding mode of the Fe(III) complex with G-quadruplex, molecular modeling was performed. The molecular docking results showed that the Fe(III) complex was docked to the end-stacked of the G-quadruplex with a ππ interaction, created between the meloxicam ligand and the guanine bases of the G-quadruplex.  相似文献   

14.
《Inorganica chimica acta》1988,147(2):257-259
The solution equilibria of Alizarin Red S (NaH2- ARS, or 9,10-dihydro-3,4-dihydroxy-9,10-dioxo-2- anthracene sulphonic acid monosodium salt) with Al(III) and Ni(II) were investigated. A general procedure for speciation and the determination of equilibria was developed from work by Coleman. and Gampp. The results of the analyses gave a KH2 for NaH2ARS of 105.71. For the AI(III)/Na2- HARS equilibrium it was determined that a 1:2 complex was formed in solution with β2 = 1012.88. For the Ni(II)/NaH2ARS equilibrium two species exist, a 1:1 and a 1:2 complex with β1 = 105.35 and β2 = 1011.53.  相似文献   

15.
The toxic inorganic monomeric forms of aluminium (Al) that limit plant growth have been shown to be effectively detoxified by complexation with organic acid ligands released by breakdown of added organic materials. The binding capacity of these acids is dependent on the degree of dissociation of their carboxyl groups and their ability to form bonds with Al. 27Al NMR spectroscopy provides a non-invasive technique to study the bonding of Al with potential ligands without disturbing the equilibrium of the system. In single ligand systems containing oxalic acid, three 27Al resonance peaks were observed at 6.4, 11.4 and 16.0 ppm downfield from the Al3+ reference peak at 0 ppm. These were assigned to Alox, Alox2 and Alox3 complexes respectively and were observable at pH values down to 3.5. In the presence of the citrate ligand, two 27Al resonance peaks at 6.1 and 11.3 ppm, assigned respectively to the Alcit and Alcit2 complexes, were observed at pH 3.4. At pH 4.3 and an Al:citrate molar ratio of 1:2, the 6.1 ppm peak was not visible, and the second peak further downfield was split into two unresolved peaks at 10.8 and 12.4 ppm indicating the presence of two forms of the Alcit2 complex. Distribution of Al between the various species, based on integration of the resonance peaks and equilibrium calculations carried out using GEOCHEM, is discussed in light of the stability constants present in the database of GEOCHEM version (v.) 1.23 and GEOCHEM-PC v. 2.0. Large discrepancies between the computed values and the NMR measured values indicate the need to incorporate more recent literature values in the database for realistic equilibrium calculations in systems containing organic acid ligands. The potential of using quantitative 27Al NMR measurements to calculate stability constants is discussed.  相似文献   

16.
The reactions of chloroauric acid (HAuCl4) with inosine=ino, guanosine=guo, triacetylinosine=trino, triacetylguanosine=trguo, and cytidine=cyd were studied. Complexes of Au(III) and Au(I) with these nucleosides have been isolated from reactions at different pH values in aqueous and in methanolic solutions. The Au(I) complexes were obtained by reducing Au(III) with 1-ascorbic acid in aqueous solutions. All the isolated complexes were characterized by elemental analyses, conductivity measurements, IR, 1H nmr, and esr spectra. The Au(III) complexes correspond to the general formulae [Au(nucl)2Cl2] Cl, Au(nucl)Cl3, and Au(nucl-H+)Cl2, while the Au(I) complexes are of the Au(nucl)2Cl type, where nucl represents the above nucleosides. In the complex with the composition [AucydCl2]2 that was isolated from aqueous solutions, the Au atom is believed to be in the (II) oxidation state.Possible structures for all the isolated complexes based on the experimental data are proposed and discussed.  相似文献   

17.
The complex formation of curium(III) with adenosine 5′-triphosphate (ATP) was determined by time-resolved laser-induced fluorescence spectroscopy (TRLFS). The interaction between soluble species of curium(III) with ATP was studied at trace Cm(III) concentrations (3 × 10−7 M). The concentrations of ATP were varied between 6.0 × 10−7 and 1.5 × 10−4 M in the pH range of 1.5-7.0 using 0.154 M NaCl as background electrolyte.Three Cm-ATP species, MpHqLr, could be identified from the fluorescence emission spectra: (i) CmH2ATP+ with a peak maximum at 598.6 nm, (ii) CmHATP with a peak maximum at 600.3 nm, and (iii) CmATP with a peak maximum at 601.0 nm. The formation constants of these complexes were calculated from TRLFS measurements to be log β121 = 16.86 ± 0.09, log β111 = 13.23 ± 0.10, and log β101 = 8.19 ± 0.16. The hydrated Cm-ATP species showed fluorescence lifetimes between 88 and 96 μs; whereas the CmATP complex has a significantly longer fluorescence lifetime of 187 ± 7 μs.  相似文献   

18.
The rate of phosphate hydrolysis of ATP in the substitution-inert complex Co(NH3)4ATP-has been examined in the presence and absence of [Co(cyclen)(H2O)2]3+. The rate of hydrolysis of Co(NH3)4ATP- in the absence of [Co(cyclen)(H2O)2]3+ is essentially independent of pH in the range 6.0 to 9.0, and the rate constant is 2.6 × 10?5 sec ?1 at pH 9.0, 40°C, and 1.0 M ionic strength Rate constants for the hydrolysis of Co(NH3)4ATP- in the presence of [Co(cyclen)(H2O)2]3+ are sharply dependent upon pH in the same range. The rate constants at pH 8.0, 8.6, and 9.0 are 8, 63, and 95 times larger than the rate constant at pH 7.0. At pH 9 the rate constant is 1.2 × 10?3 sec?1 for 16 mM Co(NH3)4ATP- in the presence of 10 mM [Co(cyclen)(H2O)2]3+. The proposed mechanism for hydrolysis involves the coordination of a phosphate group of Co(NH3)4ATP- by [Co(cyclen)(H2O)2]3+ to form a dinuclear species, followed by internal attack of coordinated hydroxide on the phosphate chain.  相似文献   

19.
A series of new ternary lanthanide complexes Ln(TFNB)3L (where Ln = Eu, Sm, Nd, Er, Yb, TFNB = 4,4,4-trifluoro-1-(2-naphthyl)-1,3-butanedionate, L = 1-(4-carbazolylphenyl)-2-pyridinyl benzimidazole) have been synthesised. The photoluminescence properties and TGA of them are described in detail. The trifluorinated ligand TFNB displays excellent antenna effect to sensitize the Ln(III) ions to emit characteristic spectra. The carbazole-containing ligand L is testified to be an outstanding synergistic ligand. The luminescence properties investigated and the quantum efficiency measured in dichloromethane solution of Eu(TFNB)3L and Sm(TFNB)3L show that the carbazole moiety is good at absorbing energy to sensitize the metal-centered emitting states and can make the complexes more rigid, provide efficient shielding of the Ln(III) core towards external quenching compared with the reference complexes of Eu(TFNB)3(Pybm) and Sm(TFNB)3(Pybm) (Pybm = 2-(2-pyridine)-benzimidazole) which have no carbazole unit. The quantum efficiency of Eu(TFNB)3L in air-equilibrated CH2Cl2 solution is calculated to be 14.8% by using air-equilibrated aqueous [Ru(bpy)3]2+·2Cl solution as reference sample (Φstd = 2.8%).  相似文献   

20.
The unprecedented self-assembled formation of a two-dimensional salicylaldimine lanthanum coordination polymer is proved by the X-ray diffraction analysis of new lanthanum(III) nitrate complex containing the N,N′-bis(salicylidene)-1,4-butanediamine ligand L, which was prepared in situ through one-step template [2+1] Schiff base condensation of salicylaldehyde with 1,4-butanediamine in the presence of lanthanum(III) nitrate ion and characterized by spectroscopic (IR, ESI-MS, UV-Vis, and 1H NMR) data and microanalysis. The complex displays an infinite [La2L4(NO3)6] polymeric structure based on networks of ten-coordinated La(III) nodes linked by bridging L ligands. The coordination geometry around the lanthanum is a distorted bicapped dodecahedron. This polymer can be described as composed from the columns of dimers connected into chains, via the flexible chain parts of ligands.  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