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.
Prussian blue (PB) is a well-known archetype of mixed valency systems. In magnetic PB analogues {CxAy[B(CN)6]z}.nH2O (C alkali cation, A and B transition metal ions) and other metallic cyanometallates {Cx(AL)y[B(CN)8]z}.nH2O (L ligand), the presence of two valency states in the solid (either A-B, or A-A' or B-B') is crucial to get original magnetic properties: tunable high Curie temperature magnets; photomagnetic magnets; or photomagnetic high-spin molecules. We focus on a few mixed valency pairs: V(II)/V(III)/V(IV); Cr(II)/Cr(III); Fe(II)-Fe(III); Co(II)-Co(III); Cu(I)-Cu(II); and Mo(IV)/Mo(V), and discuss: (i) the control of the degree of mixed valency during the synthesis, (ii) the importance of mixed valency on the local and long-range structure and on the local and macroscopic magnetization, and (iii) the crucial role of the cyanide ligand to get these original systems and properties.  相似文献   

2.
New blue emitting mixed ligand iridium(III) complexes comprising one cyclometalating, two phosphines trans to each other such as Ir{(CF3)2Meppy}(PPhMe3)2(H)(L) [L = CI, NCMe, CN] [(CF3)2Meppy = 2-(3', 5'-bis-trifluoromethylphenyl)-4-methylpyridine] were synthesized and studied to tune the phosphorescence wavelength to the deep blue region and to enhance the luminescence efficiencies. To achieve deep blue emission, the trifluoromethyl group substituted on the phenyl ring and the methyl group substituted on the pyridyl ring increased HOMO-LUMO gap and achieved the hypsochromic shift. To gain insight into the factors responsible for the emission color change and the different luminescence efficiency, we investigate the electron-withdrawing capabilities of ancillary ligands using the DFT and TD-DFT calculations on the ground and excited states of the complexes. From these results, we discuss how the ancillary ligand influences the emission peak as well as the metal to ligand charge transfer (MLCT) transition efficiency. The maximum emission spectra of Ir{(CF3)2Meppy}(PPhMe3)2(H)(Cl), [Ir{(CF3),Meppy)(PPhMe3),(H)(NCMe)]+ and Ir{(CF3)2Meppy}(PPhMe3)2(H)(CN) were in the ranges of 441, 435, 434 nm, respectively.  相似文献   

3.
Iron oxide nanowhiskers with dimensions of approximately 2 × 20 nm were successfully synthesized by selectively heating an iron oleate complex. Such nanostructures resulted from the difference in the ligand coordination microenvironments of the Fe(III) oleate complex, according to our electronic structure calculations and thermogravimetric analysis. A ligand-directed growth mechanism was subsequently proposed to rationalize the growth process. The formation of the nanowhiskers provides a unique example of shape-controlled nanostructures, offering additional insights into nanoparticle synthesis.  相似文献   

4.
5.
Unknown and controversial aspects related to the mechanisms of hydrolysis of borane complexes and to the mechanisms of chemical vapor generation for trace element determination in strongly acidic media (0.01-10 M HCl) have been investigated and clarified. The overall hydrolysis rates of borane complexes (BH(4)(-), H(3)N-BH(3)) in the acidity range of 0.2-10 M HCl were several orders of magnitude lower than those predicted by kinetics laws and obtained in the pH range of 3.8-14. The decomposition of the borane complexes takes place stepwise and proceeds through the formation of hydroboron intermediates, L(x)()BH(4)(-)(x)()(n)() (x = 1, 2, 3), where L could be one or more species among the donor groups H(2)O, NH(3), OH(-), and Cl(-) and n is the charge of the hydroboron species (n = 0, +1, -1, depending on L). Some intermediates present surprisingly long lifetimes at elevated acidities and play a key role in determining both the overall hydrolysis rates of borane complexes and the reactivity of Hg(II), As(III), Sb(III), Bi(III), Se(IV), Te(IV), and Sn(IV) in chemical vapor generation for trace element determination. Atomic absorption experiments demonstrated that almost all trihydroboron species (LBH(3)(n)()), dihydroboron species (L(2)BH(2)(n)()), and monohydroboron species (L(3)BH(n)()) play an active role in the generation of elemental mercury and stibine. Some of these intermediates are inactive or play a marginal role in the generation of arsine, bismuthine, and hydrogen selenide. Hydrogen telluride is preferentially formed by those hydroboron species, which are stable in strongly acidic conditions, while the same species are unreactive in the generation of stannane. The collected experimental evidence is in agreement with the general reactivity of the elements in chemical vapor generation techniques and, together with other literature data, definitely rule out the hypothesis of "nascent hydrogen" as a possible mechanism of chemical vapor generation by borane complex derivatization.  相似文献   

6.
The effect of experimental conditions on UVL2 (1–2 mM) disproportionation was studied spectrophotometrically through the UIVL2 accumulation (L = P2W17O 61 10? ). In 1 M NaNO3 solution containing 0.01 M HAc and 0.01 M NaAc, the rate of UVL2 disappearance is described by the equation V = k 1[UVL2]. The k 1 value is almost constant with pH decreasing from 4.5 to 1.7, but increases with increasing acetate concentration; the presence of 1 mM UIVL2, U(VI), or L does not affect k 1. In the solutions of 0.1–1.0 M HClO4 (ionic strength 1), the reaction rate is described by the equation V = 2k 2[H+]2.5[UVL2]2. Probable disproportionation mechanism is discussed. The first stage is substitution of L by water molecules in the UIVL2 complex and appearance of the reactive U(V) complex with mixed coordination sphere.  相似文献   

7.
Despite the rapid progress in solar power conversion efficiency of archetype organic–inorganic hybrid perovskite CH3NH3PbI3‐based solar cells, the long‐term stability and toxicity of Pb remain the main challenges for the industrial deployment, leading to more uncertainties for global commercialization. The poor stabilities of CH3NH3PbI3‐based solar cells may not only be attributed to the organic molecules but also the halides themself, most of which exhibit intrinsic instability under moisture and light. As an alternative, the possibility of oxide perovskites for photovoltaic applications is explored here. The class of lead‐free stable oxide double perovskites A2M(III)M(V)O6 (A = Ca, Sr, Ba; M(III) = Sb3+ or Bi3+; M(V) = V5+, Nb5+, or Ta5+) is comprehensively explored with regard to their stability and their electronic and optical properties. Apart from the strong stability, this class of double perovskites exhibits direct bandgaps ranging from 0.3 to 3.8 eV. With proper B site alloying, the bandgap can be tuned within the range of 1.0–1.6 eV with optical absorptions as strong as CH3NH3PbI3, making them suitable for efficient single‐junction thin‐film solar cell application.  相似文献   

8.
Catalytic wet hydrogen peroxide oxidation of acid dye has been explored in this study. Manganese(III) complex of N,N'-ethylene bis(salicylidene-aminato) (salenH(2)) has been encapsulated in super cages of zeolite-Y by flexible ligand method. The catalyst has been characterized by FT-IR, XRD, TG/DTA and nitrogen adsorption studies. The effects of various parameters such as pH, catalyst and hydrogen peroxide concentration on the oxidation of dye were studied. The results indicate that after 20 min at 30 degrees C, 0.175 M H(2)O(2) and 3g/L catalyst, about 90% dye removal was obtained. These studies indicate that manganese-salen complex immobilized on zeolite framework can act as a good heterogeneous catalyst for removal of dye from wastewaters.  相似文献   

9.
9-Acridone-4-carboxylic acid has been established as an efficient Cr(III) fluorescent sensor. The binding of this ligand with Cr(III) is confirmed by FTIR, thermal and mass spectral analysis of the product. Based on this chelation assisted fluorescence quenching, a highly sensitive spectrofluorometric method is developed for trace level detection, estimation and speciation studies of chromium in DMF-water. The ligand has an excitation and emission maxima at 408 nm and 498.4 nm, respectively. The equilibrium binding constant of the ligand with Cr(III) is 8.1378 × 10(4) as calculated using Stern-Volmer equation. Up to 9 × 10(-6)mol L(-1) of [Cr(3+)], linearity has been observed. The interference of foreign ions has been found to be negligible.  相似文献   

10.
1.IntroductionThereisaconsiderableinterestinthechemistryofbarbituricandthiobarbituricacidderivativesbecauseoftheirabilitytofunctionaschelatingagentsutilizedfortheanalyticaldeterminationofalargenumberoftransitionmetalions[1~6].Theaimofthepresentstudyistoprepareandelucidatethestructureofthemetalchelatesformedbycombinationof5-(1-hydroxy-6-naphthylazo-3-sodiumsulphonate)thiobarbituricandbarbituricacidswithsometransitionmetalions(VO'+,Mn'+,Fe3+lCos+)Niz+andCu:+).2.ExperimentalAllthechemicalsused…  相似文献   

11.
A detailed chemisorption mechanism is proposed for the atomic layer deposition (ALD) of aluminium oxide nano layers using trimethyl aluminum (TMA) and water as precursors. Six possible chemisorption mechanisms, complete ligand exchange, partial ligand exchange, simple dissociation, complete dissociation via ligand exchange, complete dissociation and association, are proposed and related parameters like ligand to metal ratio (L/M), concentrations of metal atoms and methyl groups adsorbed are calculated and compared against reported values. The maximum number of methyl groups that can get attached on the surface is calculated in a different way which yields a more realistic value of 6·25 per nm2 substrate area. The dependence of the number of metal atoms adsorbed on OH concentration is explained clearly. It is proposed that a combination of complete ligand exchange and complete dissociation is the most probable chemisorption mechanism taking place at various OH concentrations.  相似文献   

12.
The interaction of An(IV) ions (An = Th, U, Np, Pu) with 2,6-pyridinedicarboxylic acid (2,6-PDCA) in solutions was studied by spectrophotometry. The electronic absorption spectra of the individual complex species An(PDC)2+, An(PDC)2, and An(PDC) 3 2? (PDC2? is 2,6-PDCA anion; An = U, Np, Pu) were obtained. At [2,6-PDCA] ? 3[An(IV)] + 0.01 M and [H+] ? 0.2 M, the prevalent An(IV) species are the complexes An(PDC) 3 2? . Their overall stability constant exceeds 1025 L3 mol?3 and increases in the series from Th(IV) to Pu(IV) by ~8 orders of magnitude. Very high stability of An(IV) complexes with 2,6-PDCA anions leads to significant shifts of the redox potentials of couples involving An(IV). In particular, large difference in the stability of An(III) and An(IV) complexes is responsible for the fact that Pu(III) in the presence of 2,6-PDCA is readily oxidized with atmospheric oxygen to Pu(IV).  相似文献   

13.
A heretofore-unavailable mixed Frobenius type series, in terms of affine-transformed x-y coordinate variables of the Eshelby–Stroh type, is introduced to develop a new eigenfunction expansion technique. This is used, in conjunction with separation of the z-variable, to derive three-dimensional mixed-mode I+II+III asymptotic displacement and stress fields in the vicinity of the front of a semi-infinite through-thickness $(111)[\bar{{1}}\bar{{1}}2]\times [1\bar{{1}}0]$ crack weakening an infinite diamond cubic mono-crystalline plate. Crack-face boundary conditions and those that are prescribed on the top and bottom (free, fixed or lubricated) surfaces of the diamond cubic mono-crystalline plate are exactly satisfied. Explicit expressions for the mixed mode I+II+III singular stresses in the vicinity of the front of the through-thickness crack, are presented. Most important mixed modes I+II+III response is elicited even though the far-field loading is only mode I or II or III or any combination thereof. Finally, atomistic modeling of cracks requires consideration of both the long range elastic interactions and the short range physico-chemical reactions, such as bond breaking. The Griffith-Irwin approach does not take the latter into account, and nano-structural details such as bond orientation must be accounted for. A new mixed-mode I+II+III crack deflection criterion elucidates the formation of steps and/or triangular ridges on the crack path. The planes of a multiply deflected crack are normal to the directions of broken bonds. Additionally, the mixed-mode (I+II+III) crack deflection and ridge formation are found to be strongly correlated with the elastic stiffness constants, ${c}^{\prime }_{14}$ and ${c}^{\prime }_{56}$ , of the diamond cubic single crystal concerned.  相似文献   

14.
Spectrofluorometric studies on the binding properties of 5-pyridino-2,8-dithia[9](2,9)-1,10-phenanthrolinephane (L) toward La3+, Sm3+, Gd3+, Yb3+, and Nd3+ in methanol solution revealed the occurrence of both 1:1 and 2:1 (ligand/metal) complexation with a stability order of Nd3+ > Yb3+ > Gd3+ > Sm3+ > La3+. Consequently, L was used as a suitable neutral ionophore for the preparation of a novel polymeric membrane-selective electrode for Nd3+ ion. The electrode exhibited a Nernstian response over a wide concentration range (1.0 x 10(-6)-1.0 x 10(-2) M) with a low limit of detection of 7.9 x 10(-7) M. The electrode possesses a fast response time of <5 s and can be used for at least 9 weeks without observing any considerable deviation. The proposed electrode revealed a very good selectivity for Nd3+ over a wide variety of alkali, alkaline earth, transition, and heavy metal ions, including members of the lanthanide family other than Nd3+. The potentiometric response of the electrode is independent of the pH of test solution in the pH range 4.0-6.5. The proposed electrode was successfully applied to the recovery of Nd3+ ion from tap water samples and, also, as an indicator electrode, in potentiometric titration of neodymium(III) ions.  相似文献   

15.
The luminescence intensity of trivalent lanthanides, especially terbium(III) and europium(III), is shown to be enhanced by coordination with the ligand 2,6-pyridinedicarboxylic acid (DPA). Further enhancement can be obtained by forming a columinescent complex aggregate with ions such as gadolinium(III), terbium(III), lanthanum(III), or yttrium(III), where the Y(III) complex shows the greatest enhancement. A surfactant, sodium dodecyl sulfate (SDS), can be added to the solution to enhance the luminescence intensity of many lanthanide-ligand complexes by segregating the complex from quenchers. In this study, SDS has little effect on the limit of detection, but it does act to extend the linear dynamic range to include higher concentrations. The combination of columinescent complex with surfactant resulted in decreased luminescence intensity coupled with an increase in the background due to the light scattered by the surfactant micelles. A mechanism for the enhancement of the lanthanide luminescent intensity by energy transfer has been described by Xu et al. This study differs in that a lanthanide ion excitation band represents the most efficient excitation path and not a ligand excitation band. The complexes examined in this study have advantages over those used in prior studies since they do not require surfactants to achieve low limits of detection (100 parts per quadrillion, ppq, for Eu(DPA)(3)(3)(-) and 60 ppq for Tb(DPA)(3)(3)(-)), and they exhibit longer linear dynamic ranges (from 4 to 6 decades) than other columinescent systems.  相似文献   

16.
At a weight fraction below 20%, Pu(IV) nitrate is isomorphically dissolved and homogeneously distributed in the uranyl nitrate hexahydrate phase, the solubility being independent of the HNO3 concentration in the mixed U-Pu melt up to the acid concentration of 7 M. In dissolving solid Pu(IV) nitrate in 1–6 M HNO3, Pu(IV) disproportionates to form Pu(III) and Pu(VI). With increasing HNO3 concentration to above 10 M, Pu(IV) passes into solution as Pu(NO3) 6 2? , and no disproportionation is observed.  相似文献   

17.
Electrospray ionization mass spectrometry (ESI-MS) is used to selectively detect analytes with a high affinity for metal ions. The detection method is based on the selective monitoring of a competing ligand at its specific m/z value that is released during the ligand-exchange reaction of a metal-ligand complex with analyte(s) eluting from a reversed-phase liquid chromatography column. The ligand-exchange reaction proceeds in a postcolumn reaction detection system placed prior to the inlet of the electrospray MS interface. The feasibility of metal affinity detection by ESI-MS is demonstrated using phosphorylated peptides and iron(III)methylcalcein blue as reactant, as a model system. Methylcalcein blue (MCB) released upon interaction with phosphorylated peptides is detected at m/z 278. The ligand-exchange detection is coupled to a C8 reversed-phase column to separate several nonphosphorylated enkephalins and the phosphorylated peptides pp60 c-src (P) and M2170. Detection limits of 2 microM were obtained for pp60 c-src (P) and M2170. The linearity of the detection method is tested in the range of 2-80 micromol/L phosphorylated compounds (r(2) = 0.9996), and a relative standard deviation of less than 8% (n = 3) for all MCB responses of the different concentrations of phosphorylated compounds was obtained. The presented method showed specificity for phosphorylated peptides and may prove a useful tool for studying other ligand-exchange reactions and metal-protein interactions.  相似文献   

18.
Guo H  Li Y  Zhao K  Ren Y  Wei C 《Journal of hazardous materials》2011,186(2-3):1847-1854
Synthetic siderite has been used as adsorbent for As(III) removal in this study. Effects of contact time, temperature, pH, co-existing anions on As(III) adsorption were intensively investigated. Adsorption mechanisms were also studied using the X-ray absorption technique. Results show that the maximum adsorption capacity is up to 9.98 mg g(-1) at 25°C at a siderite dosage of 2 g L(-1). Adsorption kinetics agrees with the Lagergren pseudo-second order model. Arsenic(III) adsorption can be better described by Langmuir isotherm model for As(III) adsorption at 55°C, indicating that the coverage of the adsorption sites is in the form of monolayer, although Freundlich isotherm yields a better fit to the experimental data at 25, 35 and 45°C. Thermodynamic study indicates that As(III) adsorption on the synthetic siderite is spontaneous and endothermic in nature. The adsorption capacity is enhanced with the increase in reaction temperature. The adsorption is independent on solution pH between 3.0 and 9.6. The presence of NO(3)(-), SO(4)(2-), PO(4)(3-) or SiO(3)(2-) with element concentrations less than 20 mg L(-1) does not have adverse effect on As(III) adsorption. XANES spectra indicate that As mainly occurs as As(V) in the As adsorbed-materials, and the fraction of oxidized As(III) increases with the decrease in As(III) concentration. The formation of Fe hydroxide minerals (such as lepidocrocite and goethite) followed by As(III) oxidation and adsorption is shown to be the main mechanism of As(III) removal by the synthetic siderite.  相似文献   

19.
Oxidation of glycolate ions with Na2S2O8 + RuCl3 mixture in 0.2 M NaOH was studied by spectrophotometry. Glycolate is oxidizd to oxalate at 20–70°C. The reaction of glycolate with persulfate follows the first-order rate law with respect to [S2O 8 2? ], weakly depends on the glycolate concentration, and accelerates with increasing the Ru(III) content from 2 × 10?5 to 1 × 10?4 M. Further increase in the Ru(III) concentration does not affect the reaction rate. Probable reaction mechanism was considered.  相似文献   

20.
A new procedure for the direct determination of picomolar levels of cobalt in seawater is presented. Cathodic stripping voltammetry is preceded by adsorptive accumulation of the cobalt-nioxime (cyclohexane-1,2-dione dioxime) complex from seawater containing 6 μM nioxime and 80 mM ammonia at pH 9.1, onto a hanging mercury drop electrode, followed by reduction of the adsorbed species. The reduction current is catalytically enhanced by the presence of 0.5 M nitrite. Optimized conditions for cobalt include a 30 s adsorption period at -0.7 V and a voltammetric scan using differential pulse modulation. According to the proposed reaction mechanism, dissolved Co(II) is oxidized to Co(III) upon addition of nioxime and high concentrations of ammonia and nitrite; a mixed Co(III)-ammonia-nitrite complex is adsorbed on the electrode surface; the Co(III) is reduced to Co(II) (complexed by nioxime) during the voltammetric scan, followed by its chemical reoxidation by the nitrite, initiating a catalytically enhanced current. A detection limit of 3 pM cobalt (at an adsorption period of 60 s) enables the detection of this metal in uncontaminated seawater using a very short adsorption time. UV digestion of seawater is essential, as part of the cobalt may occur strongly complexed by organic matter and rendered nonlabile. The method was applied successfully to the determination of the distribution of cobalt in the water column of the Mediterranean.  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