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 78 毫秒
1.
Methyltin compounds (MeSn) which do not originate from man–made pollution are common in estuaries and particularly in salt marshes containing the marsh grass Spartina alterniflora. This study reports the results of experiments in which estuarine water containing S. alterniflora leaves is spiked with inorganic tin, and estuarine water alone is spiked with inorganic tin and MeSn. When decaying leaves are present, inorganic tin concentrations in the water decrease and there is a 10-fold increase in inorganic tin concentration in the leaves. This biosorption follows pseudo–first–order kinetics. MeSn3+ and Me2Sn2+ occur occasionally in the water. The Me2Sn2+ concentration decreases with time and the Me3Sn2+ concentration increases with time in S. alterniflora leaves. The results of estuarine water amended with inorganic tin and MeSn in the absence of leaves are quite different. The overall inorganic tin concentration decreases significantly during the experiment, the MeSn3+ concentration is approximately constant, and concentrations of Me2Sn2+ and Me3Sn+ increase. This means that net methylation of inorganic tin has occurred. We conclude that decaying S. alterniflora is likely to be important in the cycling of tin in salt marshes.  相似文献   

2.
This work describes a study of the underpotential deposition (UPD) of Sn2+ on a polycrystalline gold disc electrode using cyclic voltammetry (CV) and chronocoulometry (CC). Sn2+ ions showed well-defined peaks from UPD and UPD stripping (UPD-S) in 1 mol/L HCl solutions, while bulk deposition (BD) and BD stripping (BD-S) of the ions were also observed. The measured UPD shifts, EUPD, between the UPD-S and the BD-S peaks were more than 200 mV. The UPD charge and the surface coverage of tin were measured by CC. A new method for determining Sn2+ was therefore developed, based on the excellent electrochemical properties of the Au/Sn UPD system. A plot of the UPD-DPASV (differential pulse anodic stripping voltammetry) signal versus the Sn(II) concentration was obtained for [Sn(II)] of 1.98×10–7 to 3.64×10–5 M. The method developed here has been applied to determine the tin in a tin plate sample.  相似文献   

3.
The reaction Cu2+ + Cl? ? Cu+ + 12 Cl2 has been studied in three different solvents—LiCl—KCl (70–30 % mol), eutectic LiCl–KCl (58–42 % mol) and LiCl–CsCl (55–45 % mol) at different temperatures by visible and near i.r. spectrophotometry. Equilibrium constants are calculated. The standard potential of the couple Cu2+/Cu+ with reference to the standard potential of Cl2/2Cl?, as well as the thermodynamic quantities ΔH and ΔS in the range 400–600 °C, have been deduced.  相似文献   

4.
The 119Sn Mössbauer spectra of polycrystalline NiTiO3 samples impregnated with a solution containing 0.3 at % Sn4+ are evidence that annealing in H2 converts tin into the state with the electron density |Ψ(0)|2 on 119Sn nuclei corresponding to “Sn3+” ions. The stabilization of tin atoms in such an untypical formal oxidation state occurs at a depth of no more than 2–3 nm from the surface of titanate crystallites. It was revealed that the Sn3+ ions are not subjected to spin polarization even at temperatures considerably lower than the Néel temperature of NiTiO3, which can be explained by their location in the Ni2+ positions. The formation of Sn3+ prevents the further reduction of tin to the divalent state and, hence, precludes localization of 119Sn probe cations in positions at the interface.  相似文献   

5.
Electrochemically reducing CO2 into fuels using renewable electricity is a contemporary global challenge that requires significant advances in catalyst design. Photodeposition techniques were used to screen ternary alloys of Cu‐Zn‐Sn, which includes brass and bronze, for the electrocatalytic reduction of CO2 to CO and formate. This analysis identified Cu0.2Zn0.4Sn0.4 and Cu0.2Sn0.8 to be capable of reaching Faradaic efficiencies of >80 % for CO and formate formation, respectively, and capable of achieving partial current densities of 3 mA cm−2 at an overpotential of merely 200 mV.  相似文献   

6.
The present work refers to high-temperature drop calorimetric measurements on liquid Al–Cu, Al–Sn, and Al–Cu–Sn alloys. The binary systems have been investigated at 973 K, up to 40 at.% Cu in case of Al–Cu, and over the entire concentrational range in case of Al–Sn. Measurements in the ternary Al–Cu–Sn system were performed along the following cross-sections: xAl/xCu = 1:1, xAl/xSn = 1:1, xCu/xSn = 7:3, xCu/xSn = 1:1, and xCu/xSn = 3:7 at 1273 K. Experimental data were used to find ternary interaction parameters by applying the Redlich–Kister–Muggianu model for substitutional solutions, and a full set of parameters describing the concentration dependence of the enthalpy of mixing was derived. From these, the isoenthalpy curves were constructed for 1273 K. The ternary system shows an exothermic enthalpy minimum of approx. ?18,000 J/mol in the Al–Cu binary and a maximum of approx. 4000 J/mol in the Al–Sn binary system. The Al–Cu–Sn system is characterized by considerable repulsive ternary interactions as shown by the positive ternary interaction parameters.  相似文献   

7.
Lithium reaction mechanism in amorphous tin composite oxide SnB0.6P0.4O2.9 is characterized by X-ray diffraction, 119Sn Mössbauer spectroscopy and X-ray absorption fine structure. The analysis of the experimental data concerning SnB0.6P0.4O2.9 shows that SnII is highly ionic and is surrounded by three oxygen atoms. The detailed analysis of lithium insertion mechanism shows a complex reduction mechanism in two steps. During the first one the main reaction corresponds to a partial SnII reduction, involving a mixed valence system. It is accompanied by modifications of tin environments, while lithium acts as a network modifier inducing the formation of nonbridging oxygen atoms. The second step corresponds to LiSn alloying process, with the formation of LiSn bonds. It is worth noting the persistence of SnO interactions in this second step. The reversible part of the mechanism can be explained from the formation of small particles of lithium–tin alloys in strong interaction with the oxygen atoms in the glass matrix, while the vitreous support presents an interesting dispersal effect.  相似文献   

8.
The tetranuclear mixed-valent oxo-cluster [SnIISnIVO(O2CCF3)4]2 (1) has been prepared by reacting SnCl2 with AgO2CCF3 in a sealed ampoule at 90 °C. Alternatively, 1 was obtained by acidolysis of Ph3SnSnPh3 with trifluoroacetic acid in solution. The X-ray diffraction study of 1 revealed the presence of a SnIIOSn2IVOSnII core comprised of the penta-coordinated divalent and six-coordinated tetravalent tin atoms. The 119Sn NMR studies confirmed the stability of the cluster in solution and the presence of two different oxidation states of tin. An acidolysis of Ph3SnSnPh3 in the presence of [Cu2II(O2CCF3)4] followed by sublimation of the resulting product at 90 °C afforded the first trinuclear mixed metal Sn–Cu cluster [(C6H5)2Sn2IVCuIIO(O2CCF3)6] (2). The X-ray diffraction analysis of 2 revealed the presence of two phenyl groups attached to the six-coordinated tin(IV) atoms and the tetragonal pyramidal environment of the copper(II) atom. Both complexes have been obtained free of exogenous ligands.  相似文献   

9.
Na6Sn4Se11 · 22 H2O can be crystallised at –8 °C as yellow‐orange needles from the 1 : 2 H2O/CH3OH mother liquor of a superheated reaction mixture of NaOH(s), Sn and Se. The bicyclic [Sn4Se11]6– anion exhibits crystallographic C2 symmetry and is composed of corner‐bridged SnSe4 tetrahedra. Two opposite tin atoms of an Sn4Se4 8‐membered ring are linked by a common Se atom, thereby affording two 6‐membered boat‐shaped Sn3Se3 rings with a shared Sn–Se–Sn bridging unit. [Sn4Se11]6– thus represents the immediate precursor of the well‐known adamantane‐like [Sn4Se10]4– anion.  相似文献   

10.
The silicon–tin chemical bond has been investigated by a study of the SiSn diatomic molecule and a number of new polyatomic SixSny molecules. These species, formed in the vapor produced from silicon–tin mixtures at high temperature, were experimentally studied by using a Knudsen effusion mass spectrometric technique. The heteronuclear diatomic SiSn, together with the triatomic Si2Sn and SiSn2 and tetratomic Si3Sn, Si2Sn2, and SiSn3 species, were identified in the vapor and studied in the overall temperature range 1474–1944 K. The atomization energy of all the above molecules was determined for the first time (values in kJ mol?1): 233.0±7.8 (SiSn), 625.6±11.6 (Si2Sn), 550.2±10.7 (SiSn2), 1046.1±19.9 (Si3Sn), 955.2±26.8 (Si2Sn2), and 860.2±19.0 (SiSn3). In addition, a computational study of the ground and low‐lying excited electronic states of the newly identified molecules has been made. These electronic‐structure calculations were performed at the DFT‐B3LYP/cc‐pVTZ and CCSD(T)/cc‐pVTZ levels, and allowed the estimation of reliable molecular parameters and hence the thermal functions of the species under study. Computed atomization energies were also derived by taking into account spin–orbit corrections and extrapolation to the complete basis‐set limit. A comparison between experimental and theoretical results is presented. Revised values of (716.5±16) kJ mol?1 (Si3) and (440±20) kJ mol?1 (Sn3) are also proposed for the atomization energies of the Si3 and Sn3 molecules.  相似文献   

11.
Using the reduction of tin oxides with the elemental alkaline metals rubidium and cesium, stannide stannates have been synthesized which contain Zintl anions [Sn4]4— (i.e. Sn—I) and isolated oxostannate ions [SnO3]4— (i.e. Sn+II) together with further oxide ions for charge compensation. The crystal structures of the three compounds A23.6Sn7.4O13.2 = A23.6[Sn4][SnO3]3.4[O]3 (A = Rb 1a : monoclinic, P21/c, a = 2174.2(6), b = 1137.0(6), c = 2373.6(6) pm, β = 116.11(2)°, Z = 4, R1 = 0.056; A = Cs 1b : monoclinic, P21/c, a = 2042.6(6), b = 1185.4(3), c = 2481.1(7) pm, β = 97.06(2)°, Z = 4, R1 = 0.075) and Cs48Sn20O21 = Cs48[Sn4]4[SnO3]4[O]7[O2] ( 2 monoclinic, P2/c, a = 1701.8(3), b = 877.4(2), c = 4556.9(7) pm, β= 101.47(1)°, R1 = 0.093) have been determined on the basis of single crystal data. The transparency of the compounds allowed the recording of raman spectra of the anion [Sn4]4—. The 119Sn Moessbauer spectrum of the rubidium compound shows a singulet in good agreement with RbSn, overlapping a doublet caused by Sn2+ in the asymmetrical environment of the strongly electronegative oxygen ligands of SnO.  相似文献   

12.
Two types of 4f–3d thiostannates with general formula [Hen]2[Ln(en)4(CuSn3S9)] ? 0.5 en ( Ln1 ; Ln=La, 1 ; Ce, 2 ) and [Hen]4[Ln(en)4]2[Cu6Sn6S20] ? 3 en ( Ln2 ; Ln=Nd, 3 ; Gd, 4 ; Er, 5 ) were prepared by reactions of Ln2O3, Cu, Sn, and S in ethylenediamine (en) under solvothermal conditions between 160 and 190 °C. However, reactions performed in the range from 120 to 140 °C resulted in crystallization of [Sn2S6]4? compounds and CuS powder. In 1 and 2 , three SnS4 tetrahedra and one CuS3 triangle are joined by sharing sulfur atoms to form a novel [CuSn3S9]5? cluster that coordinates to the Ln3+ ion of [Ln(en)4]3+ (Ln=La, Ce) as a monodentate ligand. The [CuSn3S9]5? unit is the first thio‐based heterometallic adamantane‐like cluster coordinating to a lanthanide center. In 3 – 5 , six SnS4 tetrahedra and six CuS3 triangles are connected by sharing common sulfur atoms to form the ternary [Cu6Sn6S20]10? cluster, in which a Cu6 core is enclosed by two Sn3S10 fragments. The topological structure of the novel Cu6 core can be regarded as two Cu4 tetrahedra joined by a common edge. The Ln3+ ions in Ln1 and Ln2 are in nine‐ and eightfold coordination, respectively, which leads to the formation of the [CuSn3S9]5? and [Cu6Sn6S20]10? clusters under identical synthetic conditions. The syntheses of Ln1 and Ln2 show the influence of the lanthanide contraction on the quaternary Ln/Cu/Sn/S system in ethylenediamine. Compounds 1 – 5 exhibit bandgaps in the range of 2.09–2.48 eV depending on the two different types of clusters in the compounds. Compounds 1 , 3 , and 4 lost their organic components in the temperature range of 110–350 °C by multistep processes.  相似文献   

13.
CaO–P2O5 glasses with high concentrations of monovalent copper ions were prepared by a simple melt–quench method through CuO and SnO co‐doping. Spectroscopic characterization was carried out by optical absorption with the aim of analyzing the effects of Cu+ ions on the optical band‐gap energies, which were estimated on the basis of indirect–allowed transitions. The copper(I) content is estimated in the CuO/SnO‐containing glasses after the assessment of the concentration dependence of Cu2+ absorption in the visible region for CuO singly doped glasses. An exponential dependence of the change in optical band gaps (relative to the host) with Cu+ concentration is inferred up to about 10 mol %. However, the entire range is divided into two distinct linear regions that are characterized by different rates of change with respect to concentration: 1) below 5 mol %, where the linear dependence presents a relatively high magnitude of the slope; and 2) from 5–10 mol %, where a lower magnitude of the slope is manifested. With increasing concentration, the mean Cu+?Cu+ interionic distance decreases, thereby decreasing the sensitivity of monovalent copper for light absorption. The decrease in optical band‐gap energies is ultimately shown to follow a linear dependence with the interionic distance, suggesting the potential of the approach to gauge the concentration of monovalent copper straightforwardly in amorphous hosts.  相似文献   

14.
Radical anion salts of metal‐containing and metal‐free phthalocyanines [MPc(3?)].?, where M=CuII, NiII, H2, SnII, PbII, TiIVO, and VIVO ( 1 – 10 ) with tetraalkylammonium cations have been obtained as single crystals by phthalocyanine reduction with sodium fluorenone ketyl. Their formation is accompanied by the Pc ligand reduction and affects the molecular structure of metal phthalocyanine radical anions as well as their optical and magnetic properties. Radical anions are characterized by the alternation of short and long C?Nimine bonds in the Pc ligand owing to the disruption of its aromaticity. Salts 1 – 10 show new bands at 833–1041 nm in the NIR range, whereas the Q‐ and Soret bands are blue‐shifted by 0.13–0.25 eV (38‐92 nm) and 0.04–0.07 eV (4–13 nm), respectively. Radical anions with NiII, SnII, PbII, and TiIVO have S=1/2 spin state, whereas [CuIIPc(3?)].? and [VIVOPc(3?)].? containing paramagnetic CuII and VIVO have two S=1/2 spins per radical anion. Central metal atoms strongly affect EPR spectra of phthalocyanine radical anions. Instead of narrow EPR signals characteristic of metal‐free phthalocyanine radical anions [H2Pc(3?)].? (linewidth of 0.08–0.24 mT), broad EPR signals are manifested (linewidth of 2–70 mT) with g‐factors and linewidths that are strongly temperature‐dependent. Salt 11 containing the [NaIPc(2?)]? anions as well as previously studied [FeIPc(2?)]? and [CoIPc(2?)]? anions that are formed without reduction of the Pc ligand do not show changes in molecular structure or optical and magnetic properties characteristic of [MPc(3?)].? in 1 – 10 .  相似文献   

15.
Synthesis and Crystal Structure of the Mixed Valent Complex [Sn2I3(NPPh3)3] The mixed valent phosphoraneiminato complex [Sn2I3(NPPh3)3] ( 1 ) was prepared by the reaction of the tin(II) complex [SnI(NPPh3)]2 with sodium in tetrahydrofuran. 1 crystallizes with two formula units of THF to form yellow, moisture sensitive single crystals, which were characterized by a crystal structure determination. 1 · 2 THF: Space group P21/c, Z = 4, lattice dimensions at –80 °C: a = 1964.5(2), b = 1766.0(2), c = 2058.6(2) pm; β = 118.33(1)°, R = 0.052. 1 forms dimeric molecules in which the tin atoms are linked by two nitrogen atoms of two (NPPh3) groups to form a planar Sn2N2 four‐membered ring. The SnIV atom is additionally coordinated by a terminal iodine atom and by a terminal (NPPh3) group, whereas the SnII atom is additionally coordinated by two iodine atoms forming a ψ trigonal‐bipyramidal surrounding.  相似文献   

16.
An ethylene glycol (EG) solution containing Au(III) and Sn(IV) compounds, and conditions for the electrochemical deposition of Au–Sn alloy based on AuSn and Au5Sn intermetallics with total tin content of 30–55 at % are proposed. Fundamental difficulties of the deposition of alloys with high tin content, (including eutectic Au–Sn alloy) from aqueous electrolytes are revealed. It is determined via voltammetry that the simultaneous deposition of gold and tin from aqueous and EG electrolytes proceeds with the depolarization effect of both Au(III) and Sn(IV) as a result of the formation of the alloy, the increase in the rate of tin cathodic reduction being more noticeable in case of EG solution. Formation of SnCl2EG(H2O)2+ complex upon the dissolution of SnCl4 · 5H2O in glycol, the stability of the composition of tetracyanoaurate ions upon the dissolution of K[Au(CN)4], and the weakening of intermolecular interactions in EG with small amounts of water were revealed via IR spectroscopy. It is suggested that the depolarization effect is due not only to alloy formation, but also to the formation of SnCl2EG(H2O)2+ cations, their association with Au(CN)4- anions, and a change in the mechanism of Au(III) and Sn(IV) reduction.  相似文献   

17.
The magnesium transition metal stannides MgRuSn4 and MgxRh3Sn7—x (x = 0.98—1.55) were synthesized from the elements in glassy carbon crucibles in a water‐cooled sample chamber of a high‐frequency furnace. They were characterized by X‐ray diffraction on powders and single crystals. MgRuSn4 adopts an ordered PdGa5 type structure: I4/mcm, a = 674.7(1), c = 1118.1(2) pm, wR2 = 0.0506, 515 F2 values and 12 variable parameters. The ruthenium atoms have a square‐antiprismatic tin coordination with Ru—Sn distances of 284 pm. These [RuSn8/2] antiprisms are condensed via common faces forming two‐dimensional networks. The magnesium atoms fill square‐prismatic cavities between adjacent [RuSn4] layers with Mg—Sn distances of 299 pm. The rhodium based stannides MgxRh3Sn7—x crystallize with the cubic Ir3Ge7 type structure, space groupe Im3m. The structures of four single crystals with x = 0.98, 1.17, 1.36, and 1.55 have been refined from X‐ray diffractometer data. With increasing tin substitution the a lattice parameter decreases from 932.3(1) pm for x = 0.98 to 929.49(6) pm for x = 1.55. The rhodium atoms have a square antiprismatic tin/magnesium coordination. Mixed Sn/Mg occupancies have been observed for both tin sites but to a larger extend for the 12d Sn2 site. Chemical bonding in MgRuSn4 and MgxRh3Sn7—x is briefly discussed.  相似文献   

18.
This work presents the perspective of applying the laser desorption/ionization mass spectrometry (LDI MS) for characterization the anode film of the Ag60Cu26Zn14, Ag58.5Cu31.5Pd10, and Ag63Cu27In10 alloys (at high concentrations of chloride ions in solutions). The reference LDI mass spectra of anode films of pure Ag and Cu have been used for the identification of product corrosion. Knowing the clusters detected in the reference spectra lead to the facilitating identification of the LDI mass spectrum of the sample and reduces the analysis time. The LDI MS analysis of these alloys revealed that the predominant corrosion product are AgCl (from AgnCln+1?/+, n = 1–3), and CuCl (from “superhalogen” CumCln? clusters, m = 1–2, n = 2–6); it also revealed Cu2(OH)3Cl (from Cu2(OH)(H2O)2+) and Cu2O (from Cu(H2O)+, Cu2O doped with chlorine). These results are in accordance with the X-ray diffraction and Raman analysis. The LDI MS spectra of alloys contain the additional peaks formed due to the mutual influences of different metals in the alloys (AgCuCl3? (AgCl-CuCl2?), AgCu2Cl4? (AgCl-CuCl-CuCl2?), and Ag2CuCl4? (AgCl-AgCl-CuCl?), which is consistent with the identified corrosion products. It should be noted that the LDI MS suggest the presence of CuCl2, which can be interpreted as the corrosion products retained in the porous films of alloys, and not detected by the other methods due to a small amount. The future theoretical and experimental studies of metal clusters, significant for metallurgy, can contribute that the LDI MS is becoming a powerful analytical tool for characterization the metal surfaces.  相似文献   

19.
The electrochemical behavior of Ti(IV) and the electrodeposition of Zn-Ti alloys were investigated in a ZnCl2-urea (1:3 molar ratio) deep eutectic solvent containing 0.27 mol L?1 TiCl4. The electrochemical reduction of Ti(IV) to Ti was complicated by the formation of intermediate oxidation states of Ti(III) and Ti(II), as well as the precipitation of TiCl3. It was possible to prepare Zn-Ti alloys containing 5.8–16.7 at.% Ti. The composition and surface morphology of Zn-Ti alloys depended on deposition potential and temperature. The deposits could be indexed to a disordered hexagonal close-packed structure similar to pure Zn and were completely chloride-free. The current efficiency for the deposition of Zn-Ti alloys varied from 38.4 to 67.9 %.  相似文献   

20.
CuSn thin films were deposited by the radio‐frequency (RF) magnetron co‐sputtering method on Si(100) with Cu and Sn metal targets with various RF powers. The thickness of the films was fixed at 200 ± 10 nm. The synthesized CuSn thin films mainly consisted of Cu20Sn6 and Cu39Sn11 phases, which was revealed by an X‐ray diffraction (XRD) study. The high‐resolution Cu 2p XPS and Cu LMM Auger electron spectra indicate that metallic Cu oxidized to Cu+ and Cu2+ as the RF power on Cu target increased. The atomic ratios of Sn0 and Sn4+ decreased, while that of Sn2+ increased with increasing RF power on the Cu target. The polar surface free energy (SFE) component has a different tendency in comparison with the total SFE and the dispersive SFE component. The dispersive SFE component was the dominating contributing factor to the total SFE compared with the polar SFE. Copyright © 2016 John Wiley & Sons, Ltd.  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