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 resonance Raman spectra of K2(Ti(O2)(SO4)2)·5H2O and K2(Ti(O2)(C2O4)2)·3H2O are recorded. The results are consistent with the triangular structure of the peroxotitanium unit, Ti(O2), with C symmetry. The ν(OO), νs(TiO) and νas(TiO) are observed around 890, 610 and 535 cm−1, respectively. The resonance effects are shown to be associated with the 425 nm absorption band. This band is assigned to the O22− → Ti(IV) charge-transfer transition. The calculated force constant values for the O22− and TiO bonds are 320 and 275 N m−1, respectively.  相似文献   

2.
3.
We studied hydrated calcium oxalate and its ions at the restricted Hartree–Fock RHF/6-31G* level of theory. Performing a configurational search seems to improve the fit of the HF/6-31G* level to experimental data. The first solvation shell of calcium oxalate contains 13 water molecules, while the first solvation shell of oxalate ion is formed by 14 water molecules. The first solvation shell of Ca(II) is formed by six water molecules, while the second shell contains five. At 298.15 K, we estimate the asymptotic limits (infinite dilution) of the total standard enthalpies of hydration for Ca(II), oxalate ion and calcium oxalate as ?480.78, –302.78 and –312.73 kcal mol?1, resp. The dissociation of hydrated calcium oxalate is an endothermic process with an asymptotic limit of +470.84 kcal mol?1.
Figure
CaC2O4(H2O)16 and C2O4 2-(H2O)14  相似文献   

4.
5.
《Inorganica chimica acta》2001,312(1-2):188-196
The reaction of MoO3 and 2,4,6-tripyridyltriazine (tptz) in water at 180°C for 48 h and pH 5.5 produces (H2tptz)2[Mo8O26]·2H2O in 70% yield. The structure is constructed from δ-Mo8O26 4− clusters, H2tptz2+ and H3O+ cations linked through hydrogen bonding into a network. Crystal data: C18H16Mo4N6O14; monoclinic P21/n; a=10.2225(5) Å, b=14.0072(6) Å, c=18.1154(8) Å, β=93.896(1)°, V=2587.9(2) Å3, Z=4, Dcalc=2.372 g cm−3; R1=0.0271 based on 3212 reflections.  相似文献   

6.
The reaction of [(H2O)(NH3)5RuII]2+ with calf thymus and salmon sperm DNA has been studied over a wide fange of transition metal ion concentrations. Kinetic studies revealcd a biphasic reaction with an initial fairly rapid coordination of the metal ion being followed by slower reactions. Binding studies were done under pseudo-equilibrium conditions following completion of the initial rapid reaction. Spectra and HPLC of acid-hydrolyzed samples of [(NH3)5RuII]n-DNA prepared by incubation of [(H2O)(NH3)5RuII]2+ with DNA (where [PDNA] = 1.5 mM and reactant [RuII]/[PDNA] ratios were in the fange 0.1 to 0.3) followed by air oxidation showed the predominant binding site on helical DNA to be in the major groove at the N-7 of guanine. The equilibrium constant for [(H2O)(NH3)5RuII]2+ binding to the G7 site in helical CT DNA is 5.1 x 103. Differential pulse voltammetry exhibited a single peak at 48 mV, which is attributed to the reduction of Rum on the G7 sites.At [Run]/[PDNA] <0.5, Tm values for the DNA decreased linearly with increasing ruthenium concentration and an increase in the intensity of the 565 nm dG→ Ru(III) charge transfer band was noted upon melting. The UV and CD spectra of these samples indicated no extensive destacking or alteration in geometry (B family) compared to unsubstituted DNA. At [Run]/[PDNA]〉 0.5 or when single-stranded DNA was used, increased absorbance at 530 nm and 480 nm suggested additional binding to the exocyclic amine sites of adenine and cytosine residues. HPLC and individual spectrophotometric identification of the products derived from hydrolysis of these spec~es yielded both [(Gua)(NH3)5RuIII] and [(Ade)(NH3)5RuIII]. Earlier studies have established the cytidine and adenosine binding sites of [(NH3)5RuIII] to be at their exocyclic amines (C4 and A6). Coordination to these positions indicates disruption of the double helix since these amines are involved in hydrogen bonding on the interior of B-DNA.Agrose gel electrophoresis of superhelical pBR322 plasmid DNA after exposure to various complexes of [(nh3)5Ruiii] in the presence of a reductant and air generally revealed moderately efficient cleavage of the DNA, presumably due to the generation of hydroxyl radical via Fenton's chemistry. However, similar studies involving [(NH3)5RuIII] directly coordinated to the DNA showed no strand cutting above background. Polyacrylamide gel electrophoresis of a 381 bp, 3′-32P-labeled fragment of pBR322 plasmid DNA containing low levels of bound [(NH3)5 RuIII] further indicated negligible DNA cutting by the coordinated metal ion.  相似文献   

7.
Rates of free radical initiation were determined at 20°C in 10 mM phosphate buffer (pH 7.4) in the systems metmyoglobin (methemoglobin)–H2O2 using 2,2"-azino-bis-(3-ethylbenzthiazoline-6-sulfonic acid) as the diammonium salt (ABTS). The catalytic activity of MetMb was 2-3-fold higher than that of MetHb. The process can be described by the Michaelis–Menten equation, from which effective values of K m and V max were calculated. Comparative kinetic studies on the inhibition of ABTS oxidation were carried out using Trolox, propylgallate (PG), polydisulfide of gallic acid (poly(DSG)), polydisulfide of (2-amino-4-nitrophenol) (poly(ADSNP)), and its conjugate with human serum albumin (HSA–poly(ADSNP)). The inhibitors were characterized by inhibition constants K i and stoichiometric inhibition coefficients f (the number of radicals terminated by a single molecule of inhibitor). The minimum K i and the maximum f values were obtained for poly(DSG), and in the system of MetHb–H2O2–ABTS they were 0.08 M and 27.5, respectively. According to their antiradical activities, the inhibitors can be arranged as follows: poly(DSG) > poly(ADSNP) > PG > Trolox. PG, poly(DSG), poly(ADSNP), and its conjugate with HSA are suggested as calibrators, i.e., inhibition standards for evaluation of antioxidant status of biological fluids in possible test systems based on the free radical-generating pair MetMb–H2O2 with ABTS as the acceptor of the active radicals.  相似文献   

8.
Calcium boro fluoro zinc phosphate glasses modified using alkali oxide and doped with Nd3+ and Er3+ ions with the chemical composition of 69.5 (B2O3) + 10 (P2O5) + 10 (CaF2) + 5 (ZnO) + 5 (Na2O/Li2O/K2O) + 0.5 (Er2O3/Nd2O3) were prepared using a conventional melt quenching technique. The results of X-ray diffraction patterns indicated the amorphous nature of all the prepared glasses. The visible–near-infrared red (NIR) absorption spectra of these glasses were analyzed systematically. The NIR emission spectra of Er3+ and Nd3+:calcium boro fluoro zinc phosphate glasses showed prominent emission bands at 1536 nm (4I13/24I15/2) and 1069 nm (4F3/24I11/2) respectively with λexci = 514.5 nm (Ar+ laser) as the excitation source.  相似文献   

9.
Initial investigations into the possible roles of homocitric acid in the biosynthesis and function of the active site cofactor of nitrogenase resulted in the isolation and characterization of the dinuclear vanadium(V) species [K2(H2O)5][(VO2)2(R,S-C7H8O7)2]·H2O ( 1). Complex 1 represents the first synthetic structurally characterized transition metal homocitrate complex and may represent an early mobilized precursor in the biosynthesis of VFeco. Compound 1 was characterized by a variety of physical methods, including X-ray crystallography. Crystal data: space group P?* (#2), with a?=?10.292 (3)?Å, b?=?16.663 (3)?Å, c?=?8.343 (1)?Å, α?=?95.93 (1)°, β?=?105.74 (2)°, γ?=?90.86 (2)°, V?=?1386 (1)?Å3, and Z?=?2. The homocitrate ligand is coordinated to the vanadium(V) atoms in a bidentate fashion via the deprotonated bridging hydroxyl group and a carboxylate donor. This unique coordination mode accurately mimics the coordination of homocitrate to the cofactor of nitrogenase.  相似文献   

10.
UV—B辐射对小麦叶片H2O2代谢的影响   总被引:11,自引:1,他引:11  
研究了温室种植的小麦在0(CK)、8.82kJ/m^2(T1)和12.6kJ/m^2(T2)三种剂量的紫外线B(UV-B)辐射下H2O2含量的变化及其机理。UV-B辐射下H2O2、还原型抗坏血酸(AsA)和还原型谷胱甘肽(GSH)含量增加,抗坏血酸过氧化物酶(APx)和谷胱甘肽不原酶(GR)活性升高,脂肪酸不饱和度指数(IUFA)降低。SDS-PAGE谱图没有质上的差异,但凝胶着色深浅有变化。分析  相似文献   

11.
Radmer R  Ollinger O 《FEBS letters》1986,195(1-2):285-289
A modified mass spectrometer was used to determine whether the higher oxidation states of the photosynthetic O2-evolving system contain substrate water that is not freely exchangeable with the external medium. Our data indicated that the higher oxidation states contain no appreciable bound, non-exchangeable H2O. This suggests that H2O oxidation takes place via a rapid, concerted, all-or-none mechanism rather than by a mechanism involving stable, partially oxidized, H2O-derived intermediates. These findings set definite constraints on possible mechanisms of O2 evolution.  相似文献   

12.
Acetoacetate (AA) and 2-methylacetoacetate (MAA) are accumulated in metabolic disorders such as diabetes and isoleucinemia. Here we examine the mechanism of AA and MAA aerobic oxidation initiated by myoglobin (Mb)/H2O2. We propose a chemiluminescent route involving a dioxetanone intermediate whose thermolysis yields triplet α-dicarbonyl species (methylglyoxal and diacetyl). The observed ultraweak chemiluminescence increased linearly on raising the concentration of either Mb (10-500 μM) or AA (10-100 mM). Oxygen uptake studies revealed that MAA is almost a 100-fold more reactive than AA. EPR spin-trapping studies with MNP/MAA revealed the intermediacy of an α-carbon-centered radical and acetyl radical. The latter radical, probably derived from triplet diacetyl, is totally suppressed by sorbate, a well-known quencher of triplet carbonyls. Furthermore, an EPR signal assignable to MNP-AA adduct was observed and confirmed by isotope effects. Oxygen consumption and α-dicarbonyl yield were shown to be dependent on AA or MAA concentrations (1-50 mM) and on H2O2 or tert-butOOH added to the Mb-containing reaction mixtures. That ferrylMb is involved in a peroxidase cycle acting on the substrates is suggested by the reaction pH profiles and immunospin-trapping experiments. The generation of radicals and triplet dicarbonyl products by Mb/H2O2/β-ketoacids may contribute to the adverse health effects of ketogenic unbalance.  相似文献   

13.
《Inorganica chimica acta》1986,122(1):111-118
The title complex, prepared in 1 M NaOH, was crystallized from hot N,N-dimethylformamide/ ethanol solutions to give Na12[Ce(C6H2O2(SO3)2)4]· 9H2O·6DMF. The purple—brown crystals were examined by X-ray diffraction while inside quartz capillaries filled with DMF, (λmax 425 nm, ϵ 3664; λsh 520 nm, ϵ 2240) and belong to space group Pbca, Z=8 with a=21.846(4), b=17.348(2), c=43.103- (6) Å, V=16.335(7) Å3, Dc=1.693 gcmt−3, Do=1.725 g cmt−3. Diffractometer data were collected using Mo Kα radiation to 2θ=43o. For 7331 independent data with Fo2>3σ(Fo2) full matrix least squares refinement converged to unweighted and weighted R factors of 0.072 and 0.110, respectively, with a mixture of anisotropic and isotropic thermal parameters. The disordered DMF atom parameters were not refined. The structure consists of discrete monomeric Ce(C6H2S2O8)412− units with 12 Na+ counter cations and 10 H2O molecules (two with half occupancy), and 6 DMF molecules of solvation filling up spaces between cations and anions. Cerium(IV) is in a general position with a coordination polyhedron close to the trigonal-faced dodecahedron, D2d, with the angles between the two BAAB trapezoids of 2.3o and 3.7o. The average CeO(A) distance, 2.363(9) Å is longer than the average CeO(B) distance, 2.326(15)Å, with the reverse being true for one of the four tironato ligands. The average ring OCeO angle is 67.9(1)o. The cerium (IV) complex is found by cyclic voltammetry to undergo a quasi-reversible one-electron reduction (in strongly basic solution with excess tiron) with Ef=−497 mV vs. SCE, hence the ratio of the formation constants for tetrakis(tironato)cerate(IV) to that for tetrakis(tironato)cerate(III), KIV/KIII, is 1033. Characterization of other tiron salts is reported.  相似文献   

14.
15.
SDHD mutations are associated with human cancers but the mechanisms that may contribute to transformation are unknown. The hypothesis that mutations in SDHD increase levels of superoxide leading to genomic instability was tested using site-directed mutagenesis to generate a truncated SDHD cDNA that was expressed in Chinese hamster fibroblasts. Stable expression of mutant SDHD resulted in 2-fold increases in steady-state levels of superoxide that were accompanied by a significantly increased mutation rate as well as a 70-fold increase in mutation frequency at the hprt locus. Overexpression of MnSOD or treatment with polyethylene glycol conjugated (PEG)-catalase suppressed mutation frequency in SDHD mutant cells by 50% (P<0.05). Simultaneous treatment with PEG-catalase and PEG-SOD suppressed mutation frequency in SDHD mutant cells by 90% (P<0.0005). Finally, 95% depletion of glutathione using l-buthionine-[S,R]-sulfoximine (BSO) in SDHD mutant cells caused a 4-fold increase in mutation frequency (P<0.05). These results demonstrate that mutations in SDHD cause increased steady-state levels of superoxide which significantly contributed to increases in mutation rates and frequency mediated by superoxide and hydrogen peroxide. These results support the hypothesis that mutations in SDHD may contribute to carcinogenesis by increasing genomic instability mediated by increased steady-state levels of reactive oxygen species.  相似文献   

16.
ABSTRACT

The effect of (H2O)n (n?=?1–3) on the HNO2?+?HO → H2O?+?NO2 reaction has been investigated theoretically at the CCSD(T)/CBS//B3LYP/6-311?+?G(3df,2pd) level of theory, coupled with rate constant calculations by using variational transition state theory. Our results show that, when (H2O)n (n?=?1–3) was introduced into HNO2?+?HO → H2O?+?NO2 reaction, the product of the reaction did not change, but the potential energy surface became quite complex, yielding two kinds of reactions, namely HNO2···(H2O)n (n?=?1–3)?+?HO and HO···(H2O)n (n?=?1–3)?+?HNO2. In all catalysed reactions with (H2O)n (n?=?1–3), the former reaction type is favourable than the latter one with its effective rate constant respectively larger by 6–1 orders of magnitude than that of latter one. Within the temperature range of 240–320?K, the relative impacts on water monomer are much more obvious than dimer and trimer. However, the effective rate constant with water is larger by 658%–17% times of magnitude, showing that the positive water effect is obvious under atmospheric conditions.  相似文献   

17.
18.
19.
《Inorganica chimica acta》1986,122(2):161-168
The preparations of Pt(theophylline)2Cl2, K[Pt- (theophylline)Cl3], K[Pt(theobromine)Cl3]·H2O (1), trans-[Pt(isocaffeine)2Cl2]·H2O (2), and K(isocaffeinium)[PtCl4]·H2O (3) are reported.Crystals of 1 are monoclinic P21/n with a=7.641- (2), b=11.873(3), c=15.868(4) Å, β=90.80(2)°, Z=4. The structure was refined on 1443 reflections to R=0.028. In the planar [Pt(theobromine)Cl3] anion Pt-N(9)=2.016(6) Å, Pt-Cl=2.299(2), 2.289(2), and 2.303(2) Å. The imidazole ring is rotated away from the coordination plane by 79.8°. Symmetry related theobromine units pack parallel to each other with a mean inter-ring separation of 3.27 Å.Crystals of 2 are monoclinic P21/a with a=7.345- (2), b=20.021(5), c=8.031(2) Å, β=104.18(2)°, Z=2. The structure was refined on 1132 reflections to R=0.029. The Pt-N(7) distance is 2.003(3) Å and Pt-Cl=2.298(1) Å. The imidazole ring is rotated away from the PtCl2N2 plane by 76.8°. In this compound, the isocaffeine units do not stack, but form a staggered arrangement within the unit cell.Crystals of 3 are monoclinic P1/c with a= 7.382- (1), b=14.014(4), c=15.757(4) », β=92.30(2)°, Z=4. The structure was refined on 2057 reflections to R=0.032. The isocaffeine is protonated at N(7). The Pt-Cl distances in the PtCl42− anion range between 2.29–2.31 Å. The protonated isocaffeine cations and the PtCl42− anions form a very nearly parallel infinitely stacked arrangement with minimum interlayer atomic separations of 3.37 and 3.44 Å.  相似文献   

20.
The new d–f cyanido-bridged 1D assembly [Nd(pzam)3(H2O)Mo(CN)8] · H2O was prepared by self-assembly of pyrazine-2-carboxamide (pzam), Nd(NO3) · nH2O and (Bu3NH)3[Mo(CN)8] · 4H2O in acetonitrile. X-ray crystallographic studies indicate that the complex comprises chains of alternating, cyanido-bridged [Nd(pzam)3(H2O)]3+ and Mo(CN)8]3? fragments. The magneto-structural properties have been studied by field-dependent magnetization and specific heat measurements at low temperatures (?0.3 K). Below ≈10 K the Nd(III) moment is well approximated by an effective spin S = 1/2, with anisotropic g-tensor. The exchange coupling between the Nd(III) and the Mo(V) spins S = 1/2 along the structural chains is found to be ferromagnetic, with J/kB = 1.8 ± 0.2 K and approximately XY (planar) anisotropy. No evidence for 3D interchain magnetic ordering is found. A comparison with magneto-structural data of other cyanido-bridged complexes involving the Nd(III) ion is presented.  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