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.
Surface-enhanced Raman scattering (SERS) of 4,4′-azopyridine (AZPY) on silver foil substrate was measured under 1064 nm excitation lines. Density-functional theory (DFT) methods were used to calculate the structure and vibrational spectra of models such as Ag–AZPY, Ag4–AZPY and Ag6–AZPY complexes with B3LYP/6-31++G(d,p)(C,H,N)/Lanl2dz(Ag) basis set. The Raman bands of AZPY were identified on the ground of analog computation of potential energy distribution. The calculated spectra of Ag4–AZPY and Ag6–AZPY models were much approximated to the experimental results than that of Ag–AZPY model. The DFT results showed that the angles between two pyridyl rings keep 0° from AZPY to Ag–AZPY, Ag4–AZPY and Ag6–AZPY model. The energy gaps between the HOMO and LUMO changed from 363 to 1140 nm for AZPY-Ag complexes according to the DFT results. An conclusion was conceived that chemical enhancement mechanism may play an important role in the SERS of AZPY on silver substrate.  相似文献   

2.
The oxalic acid electrooxidation kinetics is studied at polycrystalline electrodes of vacancy-disordered Au* and Pd* obtained by anodically modifying in 0.5 M H2SO4 the surface of Ag–Au, Cu–Au, Ag–Pd, and Cu–Pd alloys containing no less than 50 at. % of the noble metal. It is found that a qualitative correlation exists between variations in thermodynamic and electrocatalytic activities of both Au* and Pd* following a change in the surface concentration of vacancies in alloys of constant composition and in the volume concentration of the electronegative component in alloys with approximately identical concentrations of vacancies. Such a correlation may be due to the existence of some factor through which an anodically-modified alloy affects the electrocatalytic process. The factor, presumably the alloy's electronic structure, is more generic than the alloy's volume composition, the nature of components, and the vacancy concentration.  相似文献   

3.
We studied on the function of the metal in the sulfated zirconia(SO42–/ZrO2) catalyst for the isomerization reaction of light paraffins. The addition of Pt to the SO42–/ZrO2 carrier could keep the high catalytic activity. The improvement in this isomerization activity is because Pt promotes removal of the coke precursor deposited on the catalyst surface. Though this catalytic function was observed in other transition metals, such as Pd, Ru, Ni, Rh and W, Pt exhibited the highest effect among them. It was further found that the Pd/SO42–/ZrO2–Al2O3 catalyst possessed a catalytic function for desulfurization of sulfur-containing light naphtha in addition to the skeletal isomerization. The sulfur tolerance of catalyst depended on the method of adding Pd, and the catalyst prepared by impregnation of the SO42–/ZrO2–Al2O3 with an aqueous solution of Pd exhibited the highest sulfur tolerance.Further, we investigated the improvement in sulfur tolerance of the Pt/SO42–/ZrO2–Al2O3 catalyst by impregnation of Pd. The results of EPMA analysis indicated that this catalyst was a hybrid-type one (Pt/SO42–/ZrO2–Pd/Al2O3) in which Pt/SO42–/ZrO2 particles and Pd/Al2O3 particles adjoined closely. This hybrid catalyst possessed a very high sulfur tolerance to the raw light naphtha that was obtained from the atmospheric distillation apparatus, although this light naphtha contained much sulfur. We assume that such a high sulfur tolerance in the hybrid catalyst is brought about by the isomerization function of Pt/SO42–/ZrO2 particles and the hydrodesulfurization function of Pd/Al2O3 particles. Besides, since the hybrid catalyst also provides high catalytic activity in the isomerization of HDS light naphtha, we suggest that the Pd/Al2O3 particles supply atomic hydrogen to the Pt/SO42–/ZrO2 particles by homolytic dissociation of gaseous hydrogen and also enhance the sulfur tolerance of Pt/SO42–/ZrO2 particles. Finally, we also propose the most suitable location of Pd and Pt in the metal-supported SO42–/ZrO2–Al2O3 catalyst.  相似文献   

4.
采用硼氢化钠还原的方法合成了碳纳米管负载的钯基纳米催化剂(Pd/CNT,Pd7Ag3/CNT,Pd7Sn2/CNT,Pd7Ag1Sn2/CNT,Pd7Ag2Sn2/CNT和Pd7Ag3Sn2/CNT)。通过XRD,TEM和XPS对其进行了表征,结果表明,相比Pd/CNT和Pd-Ag(或Pd-Sn)催化剂的纳米颗粒,Pd-Ag-Sn催化剂展现出了更小的平均颗粒尺寸(2.3 nm)。此外,还通过循环伏安(CV)和计时电流法(CA)测试了这些催化剂对甲酸氧化的电活性,在酸碱介质中,Pd-Ag-Sn/CNT对甲酸氧化都表现出了更高的电流密度。其中,Pd7Ag2Sn2/CNT催化剂在酸碱介质中的电流密度分别是108.8和211.3 mA·cm-2,相应的Pd质量电流密度高达1 364和2 640 mA·mg-1,远远高于商业Pd/C,表明Pd-Ag-Sn/CNT催化剂对甲酸氧化表现出了极好的电催化活性。  相似文献   

5.
The kinetics of tetraamminecopper(II)-catalysed oxidation of SO2– 3 to SO2– 4 in ammonia buffers and in a nitrogen atmosphere obeys the rate law: –d[SIV]/dt = k 2[CuII][SO3 2–][NH3]–1. There is spectrophotometric evidence for the formation of the intermediate complex [Cu(NH3)3(SO3)] in a pre-equilibrium.  相似文献   

6.
The reaction of meso-tetraphenylporphine (H2TPP) with Ag(OAc) or KAuCl4in boiling acetic acid affords AgIITPP and (Cl)AuIIITPP complexes. The complexes are purified by column chromatography and identified by thin layer chromatography and IR and electronic absorption spectroscopy. The transfer of a proton to the porphyrin macrocycle and dissociation of the complexes via the metal–nitrogen bonds in concentrated H2SO4at different temperatures and H2SO4concentrations are studied by spectrophotometric and kinetic methods. The formation of the stable ion-molecular H-associate of the metalloporphyrin with a doubly-charged metal cation is found for the first time for silver(II) tetraphenylporphine. Gold(III) tetraphenylporphine exists in a sulfuric acid solution in the monomolecular form. The numeric values of true rates and activation parameters of the complex dissociation are determined. The stability, state in solution, and mechanism of dissociation of the silver(II) and gold(III) tetraphenylporphine complexes are determined by the metal electronic configuration of the complex cation and, especially, by the contribution of the component to the donor–acceptor M–N interaction.  相似文献   

7.
Zusammenfassung Eine Absorptionsmethode zur direkten Bestimmung von Cu, Ag und Pd in Pb-Sn-Legierungen wurde erarbeitet. Für die Bestimmung vorteilhafte Absorptionslinien (Cu 324,7 nm, Ag 338,3 nm und Pd 276,3 nm) wurden ermittelt und auf diesen intermetallische Interferenzen von 0–500 ppm Cu, Ag und Pd sowie von 10000ppm Pb und Sn gemessen. 1g der Legierung wird in 10 ml HNO3 (D 1,52) und 2 ml 50 % HF gelöst. Die Lösung wird in einer Acetylen-Luft-Flamme analysiert. Die Methode erlaubt Gehalte bis zu 3·10–4 % Cu, 5·10–4 % Ag und 8·10–4% Pd (zweifache Blindwerthöhe) mit ausreichender Genauigkeit zu bestimmen.
Rapid determination of Cu, Ag and Pd in PbSn-tinning baths by atomic-absorption spectroscopy
An atomic absorption method for direct determination of Cu, Ag and Pd in Pb-Sn-alloys has been developed. For the determination favourable absorption lines (Cu 324.7 nm, Ag 338.3 nm and Pd 276.3 nm) were found out. Intermetallic interferences of 0–500 ppm of Cu, Ag and Pd as also 10000 ppm of Pb and Sn were measured on these lines. 1 g of the alloy is dissolved in 10 ml HNO3 (d 1.52) and 2 ml of 50 % HF. The solution is analyzed in an acetylene-air flame. This method allows the analyzing of contents up to 3×10–4 % Cu, 5·10–4% Ag and 8×10–4% Pd (twofold blank) with sufficient precision.
  相似文献   

8.
The kinetics of the silver(I)-catalysed autoxidation of SO3 2– into SO4 2– in ammonia–ammonium nitrate buffer obeyed the rate law:R obs=k1 k2 K[AgI]T[SO3 2-}][O2] / ([NH3]+K[SO3 2-])(k1+k2[O2])The values of k 1, k 2/k –1 and K were found to be 1.2l mol–1 s–1, 5.3 × 102 l mol–1 and 0.6 respectively at 30 °C. Two alternative free radical mechanisms have been proposed.  相似文献   

9.
A rapid and reliable method is developed for the determination of manganese based on oxidation of the divalent cation with a known excess of KBrO3 to the tetravalent state. The unreacted oxidant as well as Mn(IV) are then reduced with H2SO3 to Br? and Mn(II). The resulting Br? is titrated with Ag2SO4 using silver metal as the indicator electrode. K2SO4 is added to increase sensitivity and establish equilibrium in the vicinity of the end point. Fe(III) when present is also reduced with SO2 to the divalent state, which can be reoxidized with Br2. The equivalent amount of bromide is again titrated with Ag2SO4. The method provides for the simultaneous determination of Mn and Fe and finds application to some ores and steels.  相似文献   

10.
The cyanide oxidation on vitreous carbon (VC), stainless steel 304 (SS 304) and titanium (Ti) was investigated through a voltammetric study of cyanide solutions also containing copper ions. Results showed that cyanide oxidation occurs by means of a catalytic mechanism involving adsorbed species as CN, Cu(CN)43– or Cu(CN)42– depending on the electrode material. It was observed that on VC, the adsorption of Cu(CN)43– controlled the oxidation rate. Instead, for SS 304 and Ti, the adsorption of CN controlled the global process. However, in all cases, the adsorption of Cu(CN)43– on the electrode surface was required for the catalytic oxidation of CN. Voltammetric experiments for solutions containing cyanide oxidation products, such as cyanogen (CN)2 and cyanate (CNO), confirmed that the adsorbed species mentioned above controlled the catalytic oxidation of CN depending on the electrode material. A voltammetric identification of the oxidation products showed that cyanogen, (CN)2 tended to adosorb on VC, while the formation of cyanate, CNO predominated on SS 304.  相似文献   

11.
The morphology and surface roughness of silver deposits formed by cementation in 0.5M H2SO4 solution containing 0.5M CuSO4 was investigated at various temperatures. The influence of O2 on the morphology of deposited Ag on the Cu surface was studied in solutions containing 20 or 100 mg/dm3 initial Ag+. Surface‐height‐distribution diagrams were calculated from scanning‐electron‐microscopic (SEM) images. For the lower Ag+ concentration, the formation of granular deposits occurred in the presence of O2. In contrast, under anaerobic conditions, rather flat deposits with tiny Ag crystals were observed. For the higher Ag+ concentration, the presence of O2 did not significantly affect the morphology of the Ag deposit, but increasing temperature resulted in more‐compact and denser dendrites. Differences in the Ag‐deposit morphology and surface roughness were attributed to a different mechanism in the absence of O2. Under anaerobic conditions, a competitive reaction between Ag+ and Cu+ occurs in bulk solution, which consumes additional Ag+ ions. The SEM images and, especially, distribution diagrams of the surface height provided useful information on the formation and expansion of anodic sites on the Cu surface at various temperatures.  相似文献   

12.
The oxidation by ozone of a suspension of silver or silver oxide in an aqueous solution of sodium hydroxide is described. It has been shown that the oxidation proceeds in two steps:AgO3→Ag2OO3→AgO.The experimental results are in good agreement with a mechanism of dissolution and precipitation. The silver (II) oxide obtained has remarkable properties of stability in alkaline solution and of reducibility to metallic silver. These special properties are probably due to the large size of the particles.  相似文献   

13.
Two silver(I) complexes of triethyl betaine (Et3N+CH2COO, Et3BET) have been prepared and characterized by X-ray crystallography. Both complexes, [Ag2(Et3BET)2 (NO3)2] (1) and [Ag2(Et3BET)2]n (ClO4)2n (2), contain centrosymmetric bis-carboxylato-bridged Ag2(carboxylato-O,O′)2 dimers (Ag---O = 2.16–2.23 Å). The dimeric unit in 1 is bound to a chelating nitrato group [Ag---O = 2.524(3), 2.619(3) Å] at each axial site, resulting in a discrete molecule. In 2 the dimers are extended into a stair-like cationic chain via the coordination of each metal centre by a carboxylato oxygen atom [Ag---O =2.565(5) Å] from an adjacent unit. The intra-dimer Ag… Ag distance is 2.928(1) Å for 1 and 2.856(2) Å for 2.  相似文献   

14.
Poly[aniline(AN)‐co‐5‐sulfo‐2‐anisidine(SA)] nanograins with rough and porous structure demonstrate ultrastrong adsorption and highly efficient recovery of silver ions. The effects of five key factors—AN/SA ratio, AgI concentration, sorption time, ultrasonic treatment, and coexisting ions—on AgI adsorbability were optimized, and AN/SA (50/50) copolymer nanograins were found to exhibit much stronger AgI adsorption than polyaniline and all other reported sorbents. The maximal AgI sorption capacity of up to 2034 mg g?1 (18.86 mmol g?1) is the highest thus far and also much higher than the maximal Hg‐ion sorption capacity (10.28 mmol g?1). Especially at ≤2 mM AgI, the nanosorbents exhibit ≥99.98 % adsorptivity, and thus achieve almost complete AgI sorption. The sorption fits the Langmuir isotherm well and follows pseudo‐second‐order kinetics. Studies by IR, UV/Vis, X‐ray diffraction, polarizing microscopy, centrifugation, thermogravimetry, and conductivity techniques showed that AgI sorption occurs by a redox mechanism mainly involving reduction of AgI to separable silver nanocrystals, chelation between AgI and ? NH? /? N?/? NH2/ ? SO3H/? OCH3, and ion exchange between AgI and H+ on ? SO3?H+. Competitive sorption of AgI with coexisting Hg, Pb, Cu, Fe, Al, K, and Na ions was systematically investigated. In particular, the copolymer nanoparticles bearing many functional groups on their rough and porous surface can be directly used to recover and separate precious silver nanocrystals from practical AgI wastewaters containing Fe, Al, K, and Na ions from Kodak Studio. The nanograins have great application potential in the noble metals industry, resource reuse, wastewater treatment, and functional hybrid nanocomposites.  相似文献   

15.
The distribution coefficients (DC) for HgCl 4 2– , Hg(SO4) 2 2– , Hg(NO3) 4 2– , Ag+, Ag(SCN) 2 and Ag(NH3) 2 + between aqueous solutions and Dowex A-1 were measured in varying hydrogen ion concentrations. The DC of Ag+ in the NO 3 media was very low (4 to 6). The DC for the Ag(SCN) 2 complex decreased as pH increased. The Ag(NH3) 2 + complex had a constant DC of about 65 from pH 8 and above. The trend observed for three mercury complexes in HCl, H2SO4 and HNO3 was similar; the DC decreased steadily from 0.1M to 6M. The HgCl 4 2– complex had the highest DC (9000) while the Hg(NO3) 4 2– complex had the lowest DC (2000).  相似文献   

16.
Summary Some copper(II) complexes of the types: Cu(HPPK)-(PPK)X, Cu(HMPK)(MPK)X (where HPPK = syn-phenyl-2-pyridylketoxime, HMPK = syn-methyl-2-pyridylketoxime and X = Cl, Br, I, NO3 , SCN or SeCN) Cu(HPPK)2SO4 3 H2O and Cu(HMPK)2SO4 · 3 H2O were synthesized and characterized by analysis, magnetic susceptibility, e.s.r., reflectance and i.r. spectral measurements. The spectral data suggest that Cu(HPPK)(PPK)X and Cu(HMPK)(MPK)X containcis square-coplanar [Cu(HPPK)(PPK)]+ and [Cu(HMPK)(MPK)]+ units respectively, linked by weakly coordinated anions, giving infinite polymeric highly distorted octahedral chain structures, whereas Cu(HPPK)2SO4 · 3H2O and Cu(HMPK)2SO4 · 3 H2O have acis distorted octahedral structure containing two ligand molecules of ketoxime and a bidentate sulphate group. The polycrystalline e.s.r. spectra suggest a distorted octahedral stereochemistry for the CuII ion involving a ground-state. By using e.s.r. and reflectance spectral data, the orbital reduction parameters, k11 and k1 were calculated and interpreted in terms of molecular orbital coefficients.  相似文献   

17.
Silver platinum binary alloys with compositions between about Ag2Pt98 and Ag95Pt5 at < 400 °C have largely not been observed in bulk due to the large immiscibility between these two metals. We present in this paper that Ag–Pt alloy nanostructures can be made in a broad composition range. The formation of Ag–Pt nanostructures is studied by powder X-ray diffraction (PXRD) and energy-dispersive X-ray (EDX). Our results indicate that lattice parameter changes almost linearly with composition in these Ag–Pt nanomaterials. In another word, lattice parameter and composition relationship follows the Vegard's law, which is a strong indication for the formation of metal alloys. Our transmission electron microscopy (TEM) study shows that the silver-rich Ag–Pt alloy nanostructures have spherical shape, while the platinum-rich ones possess wire-like morphology. The stability and crystal phase are investigated by annealing the alloy nanostructures directly or on carbon supports.  相似文献   

18.
The oxidation of americium in HNO3, H2SO4 and HClO4 solutions by a mixture of potassium persulfate with silver salt in the presence of potassium phosphotungstate has been investigated. The influence of acid and its concentration, of (NH4)2S2O3, K10P2W17O61 and silver salt on Am(III) oxidation rate, yield and stability of Am(IV) and Am(VI), has been studied. The complexation of Am(III), Am(IV) and Am(VI) with phosphotungstate ions has been investigated. It has been established that Am(III) and Am(IV) form ML2 complexes and their apparent stability constants have been estimated. The oxidation mechanism is discussed. A method for preparing of Am(IV) in 0.1–6M HNO3, O.1–3M H2SO4, 0.1–1M HClO4 solutions is proposed. The oxidation of Am(III) to Am(IV) by KBrO3 and K2Cr2O7 in HNO3, H2SO4, HClO4 solutions in the presence of K10P2W17O61 has been investigated.  相似文献   

19.
Ag16B4O10 has been obtained as a coarse crystalline material via hydrothermal synthesis, and was characterized by X-ray single crystal and powder diffraction, conductivity and magnetic susceptibility measurements, as well as by DFT based theoretical analyses. Neither composition nor crystal structure nor valence electron counts can be fully rationalized by applying known bonding schemes. While the rare cage anion (B4O10)8− is electron precise, and reflects standard bonding properties, the silver ion substructure necessarily has to accommodate eight excess electrons per formula unit, (Ag+)16(B3+)4(O2−)10 × 8e, rendering the compound sub-valent with respect to silver. However, the phenomena commonly associated with sub-valence metal (partial) structures are not perceptible in this case. Experimentally, the compound has been found to be semiconducting and diamagnetic, ruling out the presence of itinerant electrons; hence the excess electrons have to localize pairwise. However, no pairwise contractions of silver atoms are realized in the structure, thus excluding formation of 2e–2c bonds. Rather, cluster-like aggregates of an approximately tetrahedral shape exist where the Ag–Ag separations are significantly smaller than in elemental silver. The number of these subunits per formula is four, thus matching the required number of sites for pairwise nesting of eight excess electrons. This scenario has been corroborated by computational analyses of the densities of states and electron localization function (ELF), which clearly indicate the presence of an attractor within the shrunken tetrahedral voids in the silver substructure. However, one bonding electron pair of s and p type skeleton electrons per cluster unit is extremely low, and the significant propensity to form and the thermal stability of the title compound suggest d10–d10 bonding interactions to strengthen the inter-cluster bonding in a synergistic fashion. With the present state of knowledge, such a particular bonding pattern appears to be a singular feature of the oxide chemistry of silver; however, as indicated by analogous findings in related silver oxides, it is evolving as a general one.

Ag16B4O10, obtained via hydrothermal synthesis, displays an unprecedented bonding scheme, hosting excess electrons localized pairwise in cluster-like silver subunits.  相似文献   

20.
The oxidation of the [Fe(CO)4]2– dianion with Ag+ salts occurs through a particularinner-sphere mechanism, which involves an intermediate cascade of silver clusters stabilized by Fe(CO)4 ligands. The last detectable Ag-Fe cluster of the sequence is the [Ag13{-Fe(CO)4}8]3– trianion, which has been selectively obtained by using ca. 1.7 equivalents of Ag+ per mole of [Fe(CO)4]2–. The [Ag13{-Fe(CO)4}8]3–- trianion has been isolated in a crystalline state with several quaternary cations, and has been characterized by X-ray diffraction studies of its bis(triphenylphosphine)iminium salt. [N(PPh3)2]3 [Ag13{ 3-Fe(CO)4}8]·2(CH3)2CO, monoclinic, space group P21 (No.4),a = 16.284(2) Å,b =18.767(5) Å,c = 25.905(4) Å, = 90.46(1)°,V = 7916(3) Å3,Z = 2,R = 0.0324. The molecular structure of the anion consists of a centered cuboctahedron of silver atoms with the triangular faces capped by Fe(CO)4 units. Chemical reduction of ( Ag13{ 3-Fe(CO)4}8]3– affords the corresponding [Ag13{ 3-Fe(CO)4)8]4–, which in turn gives [Ag13{ 3-Fe(CO)4)8]5– and [Ag6{ 3-Fe(CO)4}4] upon further reduction. Electrochemical investigations confirm the reversibility of the [Ag13{ 3-Fe(CO)4}8]3–/4– redox change. Furthermore, in spite of some electrode poisoning effects, evidence of the existence of the [Ag13{ 3-Fe(CO)4}8]5– pentaanion was obtained. The yet structurally uncharacterized [Ag6{ 3-Fe(CO)4)4]2– dianion is quantitatively obtained by reaction of [Fe(CO)4]2– with ca. 1.5 equivalents of Ag+ or by addition of one equivalent of Ag+ to solutions of the [Ag5{Fe(CO)4}4]3– trianion. All attempts to isolate its quaternary salts as crystalline materials failed owing to formation of amorphous insoluble precipitates. The above series of 3-Fe(CO)4 octa-capped cuboctahedral Ag13 clusters can be envisioned as the Ag+ . Ag and Ag cryptates of the [Ag12{}3-Fe(CO)4}8]4– cryptand. respectively.Dedicated to Prof L. F. Dahl on his 65th birthday.  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