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 140 毫秒
1.
2.
Although the mechanism for the transformation of carbon dioxide to formate with copper hydride is well understood, it is not clear how formic acid is ultimately released. Herein, we show how formic acid is formed in the decomposition of the copper formate clusters Cu(II)(HCOO)3 and Cu(II)2(HCOO)5. Infrared irradiation resonant with the antisymmetric C−O stretching mode activates the cluster, resulting in the release of formic acid and carbon dioxide. For the binary cluster, electronic structure calculations indicate that CO2 is eliminated first, through hydride transfer from formate to copper. Formic acid is released via proton-coupled electron transfer (PCET) to a second formate ligand, evidenced by close to zero partial charge and spin density at the hydrogen atom in the transition state. Concomitantly, the two copper centers are reduced from Cu(II) to Cu(I). Depending on the detailed situation, either PCET or hydrogen atom transfer (HAT) takes place.  相似文献   

3.
Electrochemical reduction of CO2 has the benefit of turning greenhouse gas emissions into useful resources. We performed a comparative study of the electrochemical reduction of CO2 on stepped Pb(211) and Sn(112) surfaces based on the results of density functional theory slab calculations. We mapped out the potential energy profiles for electrochemical reduction of CO2 to formate and other possible products on both surfaces. Our results show that the first step is the formation of the adsorbed formate(HCOO*) species through an Eley-Rideal mechanism. The formate species can be reduced to HCOO- through a oneelectron reduction in basic solution, which produces formic acid as the predominant product. The respective potentials of forming HCOO* are predicted to be -0.72 and -0.58 V on Pb and Sn. Higher overpotentials make other reaction pathways accessible, leading to different products. On Sn(112), CO and CH4 can be generated at -0.65 V following formate formation. In contrast, the limiting potential to access alternative reaction channels on Pb(211) is -1.33 V, significantly higher than that of Sn.  相似文献   

4.
The infrared and Raman spectra of the light blue modification of anhydrous copper(II) formate, Cu(HCOO)2, and copper(II) formate-d2, Cu(DCOO)2, are reported, as well as the Raman spectra of copper(II) formate tetrahydrate Cu(HCOO)2 · 4H2O and copper(II) formate tetrahydrate-d8 Cu(HCOO)2 · 4D2O over a wide range of temperatures. In the latter two compounds, the fundamental formate modes, active in the Raman spectra, showed splittings when the phase transition temperature was traversed. These low-temperature Raman spectra were interpreted in terms of a P21 space group and prove that the phase transition not only involves an ordering in the orientation of the water molecules, but also displacements of the heavy atoms. Only a limited number of weak translational modes of the water molecules could be identified in the Raman spectra of the copper(II) formate tetrahydrate, and it is not possible therefore to determine exactly how ordering affects the Raman-active lattice modes of these molecules.  相似文献   

5.
We report the synthesis, crystal structures, and spectral, thermal, and magnetic properties of a family of metal–organic perovskite ABX3, [C(NH2)3][MII(HCOO)3], in which A=C(NH2)3 is guanidinium, B=M is a divalent metal ion (Mn, Fe, Co, Ni, Cu, or Zn), and X is the formate HCOO?. The compounds could be synthesized by either diffusion or hydrothermal methods from water or water‐rich solutions depending on the metal. The five members (Mn, Fe, Co, Ni, and Zn) are isostructural and crystallize in the orthorhombic space group Pnna, while the Cu member in Pna21. In the perovskite structures, the octahedrally coordinated metal ions are connected by the antianti formate bridges, thus forming the anionic NaCl‐type [M(HCOO)3]? frameworks, with the guanidinium in the nearly cubic cavities of the frameworks. The Jahn–Teller effect of Cu2+ results in a distorted anionic Cu–formate framework that can be regarded as Cu–formate chains through short basal Cu? O bonds linked by the long axial Cu? O bonds. These materials show higher thermal stability than other metal–organic perovskite series of [AmineH][M(HCOO)3] templated by the organic monoammonium cations (AmineH+) as a result of the stronger hydrogen bonding between guanidinium and the formate of the framework. A magnetic study revealed that the five magnetic members (except Zn) display spin‐canted antiferromagnetism, with a Néel temperature of 8.8 (Mn), 10.0 (Fe), 14.2 (Co), 34.2 (Ni), and 4.6 K (Cu). In addition to the general spin‐canted antiferromagnetism, the Fe compound shows two isothermal transformations (a spin‐flop and a spin‐flip to the paramagnetic phase) within 50 kOe. The Co member possesses quite a large canting angle. The Cu member is a magnetic system with low dimensional character and shows slow magnetic relaxation that probably results from the domain dynamics.  相似文献   

6.
Self‐assembly of the [Mo(CN)7]4– anion and the Mn2+ ion in the aqueous solution containing ammonium formate results in a new coordination polymer, {(NH4)3[(H2O)Mn3(HCOO)][Mo(CN)7]2·4H2O}n. Single crystal X‐ray analysis revealed a very complicated three‐dimensional (3D) framework, where both the [Mo(CN)7]4– and the formate anions act as bridges between the MnII centers. Magnetic measurements revealed that this compound displays ferrimagnetic ordering below 70 K. Competing antiferromagnetic interactions between the spin carriers might lead to spin frustration and non‐linear alignment of the magnetic moments. Specifically, this compound is the first mixed [Mo(CN)7]4–/HCOO bridged molecule magnet.  相似文献   

7.
The study of the thermal dehydration of the compounds CuX2.nH2O, where X were formate, salicylate or phtalate anions, was performed, including the observations of structural changes of compounds during their decomposition. It is shown that the dehydration of copper(II) formate tetra- and dihydrates is accompanied by significant changes in the bonding of the formato groups and remaining water molecules. Two structurally different modifications of Cu(HCOO)2.2H2O were prepared, the structure differences are clearly demonstrated in their decomposition stoichiometry. The dehydration of copper formate hydrates was found to be controlled by chemical reaction on the phase boundary. The dehydration of the copper salicylate tetrahydrate and copper phtalate monohydrate is accompanied by the structural changes of the whole compounds, as well, however these processes are diffusion or nucleation controlled.
Zusammenfassung Es wurde die thermische Dehydratation der Verbindungen CuX2.nH2O mit dem Formiat-, Salicylat- und Phthalatanion als X unter gleichzeitiger Beobachtung eventueller Strukturveränderungen dieser Verbindungen während der Dehydratation untersucht. Es wird gezeigt, dass die Dehydratation von Kupfer(II)-formiat Tetrabzw. Dihydrat von einer eindeutigen Veränderung der Bindungsverhältnisse der Formiatgruppen und der verbleibenden Wassermoleküle begleitet wird. Es wurden zwei strukturell verschiedene Modifikationen von Cu(HCOO)2.2H2O hergestellt, die Strukturunterschiede zeigen sich deutlich in der Stöchiometrie ihrer Zersetzung. Die Dehydratation von Kupferformiathydraten wird durch die chemische Reaktion an der Phasengrenze kontrolliert. Die Dehydratation von Kupfersalicylat Tetrahydrat und Kupferphthalat Monohydrat wird von einer Strukturveränderung des ganzen Moleküles begleitet, unabhängig davon, ob der Vorgang durch die Diffusion oder Kernbildung bestimmt wird.
  相似文献   

8.
A rapid, simple, and reliable competitive immunoassay was developed for measurement of lead ions Pb(II) in environmental samples. Avian antibodies were produced against Pb(II). Since lead ions are too small to elicit an immune response, the metal was coupled to protein carrier Bovine serum albumin (BSA) using a bifunctional chelator 1-(4-isothiocyanobenzyl) ethylenediamine N,N,N′,N′-tetra acetic acid (ITCBE). Poultry birds (layers) were immunised with this Pb(II)–ITCBE–BSA immunoconjugate and the avian antibodies (IgY) isolated from egg yolk recognised Pb(II)-ITCBE complexes as capture reagent and a Pb(II)–ITCBE conjugate of Alkaline phosphatase as an enzyme label. Antibody reaction was optimised for different concentrations of antigen and antibody dilutions. Cross reactivity with other metals were below 1% in competitive ELISA. The IC50 value of this avian antibody was 0.19?µg?mL?1. The detection range and the detection limit were 0.02–1000?µg?mL?1and 0.2?µg?mL?1, respectively.  相似文献   

9.
The existence of new polymorphic varieties of strontium formate (β, δ) and calcium formate (γ, δ) has been established by differential thermal analysis. In our experimental conditions the following species are detected: β-Sr(HCOO)2 below 235°C; δ-Sr(HCOO)2 above 235°C; γ-Ca(HCOO)2 between 150 and 300°C; and δ-Ca(HCOO)2 beyond 300°C. It is shown that some of the corresponding polymorphic transitions for these formates are possible only in the presence of water traces.  相似文献   

10.
Solubilities in the CuSO4 (CuCl2, Cu(NO3)2)–NaHCOO–H2O systems are studied at 25°C using the isothermal sections method. Crystallization regions of copper(II) formate mono- and dihydrate are elucidated. It is proved that copper(II) formate can be synthesized in CuAn2 + 2NaHCOO ? Cu(HCOO)2 + 2NaAn–H2O quaternary reciprocal systems using the conversion method.  相似文献   

11.
The systems Ba(HCOO)2-Ca(HCOO)2-H2O and Ba(HCOO)2-Mg(HCOO)2-H2O were studied at 25°C by the solubility method. The effect of the ionic radii of doubly charged metals on the formation of double salts containing barium formate was considered.  相似文献   

12.
The discovery of new nonlinear optical (NLO) crystals with excellent properties is in urgently demand because of their ability to generate coherent light. Herein, we report an unique NLO lead bromide formate, KCs2[Pb2Br5 (HCOO)2], which has been synthesized by a mix-solvothermal method. KCs2[Pb2Br5(HCOO)2] exhibits strong phase-matching second-harmonic generation (SHG) response (6.5×KDP), large birefringence (0.16@ 1064 nm), and a wide transparent window in most visible light and mid-IR region. Interestingly, KCs2[Pb2Br5(HCOO)2] features a polar 3D lead-bromide framework in which adjacent Pb−Br layers containing coplanar Pb6Br6 rings are not only parallel to each other, but also orient in the same direction. These oriented arrangements are responsible for the strong SHG response and large birefringence that are elucidated by both local dipole moment and theoretical calculations. This research provides a new strategy to explore subsequent NLO crystals.  相似文献   

13.
《Analytical letters》2012,45(9):1643-1655
Abstract

A Kinetic spectrophotometric method for the determination of ultra-trace amounts of lead(II) is described. This method is based on the catalytic action of this ion on the reduction of resazurin by sulfide. The course of the chemical reaction is followed spectrophotometrically by the measurement of reduction in absorbance of resazurin at 605 nm. The calibration range of Pb(II) is dependent on the sulfide concentration.

With this method 1 ng m1?1 lead can be detected. The method is used for determination of Pb(II) in KNO3 (Fluka) and in water.  相似文献   

14.
Reactions of meso‐bis[(diphenylphosphinomethyl)phenylphosphino]methane (dpmppm) with CuI species in the presence of NaBH4 afforded di‐ and tetranuclear copper hydride complexes, [Cu2(μ‐H)(μ‐dpmppm)2]X ( 1 ) and [Cu4(μ‐H)24‐H)(μ‐dpmppm)2]X ( 2 ) (X=BF4, PF6). Complex 1 undergoes facile insertion of CO2 (1 atm) at room temperature, leading to a formate‐bridged dicopper complex [Cu2(μ‐HCOO)(dpmppm)2]X ( 3 ). The experimental and DFT theoretical studies clearly demonstrate that CO2 insertion into the Cu2(μ‐H) unit occurred with the flexible dicopper platform. Complex 2 also undergoes CO2 insertion to give a formate‐bridged complex, [Cu4(μ‐HCOO)3(dpmppm)2]X, during which the square Cu4 framework opened up to a linear tetranuclear chain.  相似文献   

15.
Abstract

207Pb N.M.R. data are reported for a number of even chain length lead(II) carboxylates (soaps) at various temperatures. At room temperature, the solid lead(II) decanoate and tetradecanoate show similar spectra, with a single metal ion site, and modest shielding anisotropy. As the temperature is increased, the soaps (hexanoate to octadecanoate) all form a highly ordered smectic phase, which gives a very broad 207Pb signal of linewidth comparable to that of the solid phase. At higher temperatures, the hexanoate to dodecanoate soaps form a lamellar L α (smectic A) phase, whilst the longer chain length carboxylates melt directly to the liquid phase. Both the lamellar L α and liquid phase give fairly sharp, isotropic signals, whose chemical shifts and linewidths are strongly temperature dependent. Possible explanations for this effect include paramagnetic contributions to the shielding tensor from low-lying electronic states of Pb(II), and contributions to the observed signal from different coordination species produced in the lead(II) carboxylate system. Although there are discrete changes in chemical shift at the phase transition, the magnitudes observed in all the phases are similar, suggesting that there are no dramatic changes in the metal coordination environment.  相似文献   

16.
The three‐dimensional metal–organic framework poly[bis(dimethylammonium) [hexa‐μ2‐formato‐κ12O:O′‐aluminium(III)sodium(I)]], {(C6H8N)2[AlNa(HCOO)6]}n, was obtained serendipitously and has been characterized by X‐ray diffraction. The product has arisen as a result of a hydrolysis reaction of dimethylformamide (DMF) and contains dimethylammonium (DMA) cations included in structural voids formed by a three‐dimensional [AlNa(HCOO)6] network. This study provides evidence that, in the presence of traces of aluminium, DMF stored in a glass bottle can be hydrolysed to formate and dimethylamine with simultaneous extraction of Na+ cations from the glass. It also demonstrates that care must be taken regarding the metal and water content when DMF is not freshly distilled, since the hydrolysis of amide can occur.  相似文献   

17.
The cobalt‐formate coordination polymers {[Co(bpyph)(HCOO)2]8}n ( 1 ) (bpyph = 1,4‐bis(2‐(4‐pyridyl)ethenyl)benzene) and {[Co(HCONH2)2(HCOO)2]}n ( 2 ) have been prepared by interaction of Co(NO3)2 · 6 H2O in formamide solution with generation of formate anion by hydrolysis of the solvent. Coordination polymer 1 reveals an unprecedented example of “molecular multi‐rod cable” architecture, in which eight single “molecular wires” {[Co(bpyph)]}n are interlinked by bridging formate anions to give infinite octameric chains. The formate groups adopt mono‐, and bi‐ and tridentate bridging and chelate modes of coordination (Co–O 1.966–2.134 Å). The coordination geometry around the cobalt atoms is essentially dominated by the demands for most effective packing of parallel situated polycyclic aromatic ligands, with extensive CH…π, or edge‐to‐face stacking interactions within the single octameric chain as well as between the closest neighbours (C…C separations within this stack are ca. 3.50 Å).  相似文献   

18.
The 3D framework [Mn3(CH3COO)2(HCOO)4]n · nDMF ( 1 ) was obtained from the assembly of MnII ions with acetate and the in‐situ generated formate ligands. It features Mn‐centered MnMn4 tetrahedral nodes, each of which is linked to another four ones by sharing the apexes to form the 3D framework of 1 . Each of the acetate and formate ligands behaves as a synsyn:anti bridge to link two apical MnII ions and the central MnII ion. The magnetic measurement of 1 revealed the coexistence of spin‐canted antiferromagnetism and metamagnetism. It represents a typical example to synergistically use two kinds of carboxylate ligands to construct metal‐organic frameworks, as well as to tune the structure and magnetic properties of the aimed complex.  相似文献   

19.
A series of lead(II) coordination polymers containing [N(CN)2]? (DCA) or [Au(CN)2]? bridging ligands and substituted terpyridine (terpy) ancillary ligands ([Pb(DCA)2] ( 1 ), [Pb(terpy)(DCA)2] ( 2 ), [Pb(terpy){Au(CN)2}2] ( 3 ), [Pb(4′‐chloro‐terpy){Au(CN)2}2] ( 4 ) and [Pb(4′‐bromo‐terpy)(μ‐OH2)0.5{Au(CN)2}2] ( 5 )) was spectroscopically examined by solid‐state 207Pb MAS NMR spectroscopy in order to characterise the structural and electronic changes associated with lead(II) lone‐pair activity. Two new compounds, 2 and [Pb(4′‐hydroxy‐terpy){Au(CN)2}2] ( 6 ), were prepared and structurally characterised. The series displays contrasting coordination environments, bridging ligands with differing basicities and structural and electronic effects that occur with various substitutions on the terpyridine ligand (for the [Au(CN)2]? polymers). 207Pb NMR spectra show an increase in both isotropic chemical shift and span (Ω) with increasing ligand basicity (from δiso=?3090 ppm and Ω=389 ppm for 1 (the least basic) to δiso=?1553 ppm and Ω=2238 ppm for 3 (the most basic)). The trends observed in 207Pb NMR data correlate with the coordination sphere anisotropy through comparison and quantification of the Pb? N bond lengths about the lead centre. Density functional theory calculations confirm that the more basic ligands result in greater p‐orbital character and show a strong correlation to the 207Pb NMR chemical shift parameters. Preliminary trends suggest that 207Pb NMR chemical shift anisotropy relates to the measured birefringence, given the established correlations with structure and lone‐pair activity.  相似文献   

20.
In this work, an all the way green analytical procedure based on a silicon oxide film-solid phase extraction system is proposed for lead traces determination. From the synthesis of a solid phase extraction (SPE) system and throughout the metal preconcentration and determination only aqueous media were employed. Characterisation of the film was carried out by Scanning Electron Microscopy and Attenuated Total Reflectance-Fourier Transform Infrared Spectroscopy. Preconcentration conditions, prior to Pb(II) Electrothermal Atomic Absorption Spectrometry (ETAAS) determination, such as adsorption and desorption time, pH and temperature, were optimised. Langmuir, Freundlich and Dubinin-Radushkevich isotherm models were analysed along with the evaluation of adsorption energy and standard free energy (ΔG 0). The greatest adsorption was obtained with incubation at pH 7 and 37°C. By using a small volume of 0.5?mol?L?1 HNO3 (0.5?mL) lead was desorbed from the silicon oxide film after 2?h incubation, generating low amount of waste. The films showed better adjustment for the Langmuir model (R2 ?=?0.989). The detection limit (3.29σ) for Pb(II) was 0.228?µg?L?1. The developed procedure is 10-fold more sensitive in comparison to direct ETAAS determination. Recovery values from soft tap-water and soft well-water were above 95%. When hard water was analysed, Pb(II) adsorption was found to be interfered by Mg2+ and Ca2+. After five preconcentration cycles relative recovery was found not to decay below 90%, indicating that the silicon oxide film could be used for multiple lead determinations.  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