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 31 毫秒
1.
The Rh target preparation for production of 103Pd was investigated by using a thick electrodeposition of rhodium metal on a copper backing. The electrodeposition experiments were performed in acidic sulfate media using RhCl3·3H2O, Rh2(SO4)3 (recovered from hydrochloric acid solution) and also in the commercially available Rhodex plating baths. For high current beam irradiation of a Rh target, the qualities of the deposit of the three baths were compared in terms of thermal shock, crack-free and morphology criteria. The quality of the plating obtained from a sulfate bath [Rh2(SO4)3] was comparable with the one obtained from commercially available Rhodex bath. The optimum conditions of the electrodepositions were as follows: 4.8 g rhodium [as Rh2(SO4)3], pH 2, DC current density of ca 8.5 mA·cm–2, 1% sulfamic acid (w/v) and temperature 40–60 °C.The authors would like to thank their colleagues at the VUB-Cyclotron department for help and assistance in preparation of the electrodeposition equipment and taking the SEM photomicrographs and also K. Aardaneh (NRCAM) for his assistance.  相似文献   

2.
Summary Rhodium electrodeposition on a copper substrate was investigated for 103Pd production. The electrodeposition was carried out by the commercially available Rhodex plating baths. The optimum conditions of the electrodeposition for complete depletion of Rh were: 6.2 g/l rhodium, DC current density of 12.83 mA . cm-2 and 60 °C temperature.  相似文献   

3.
Nitration of sulfate complexes of rhodium has been investigated by NMR 103Rh, 14N, 15N, and 17O NMR. At high pH, [Rh(NO2)6]3?, dimer [Rh2(μ-OH)2(NO2)8]4?, and trimer [Rh3(μ-OH)4(OH)(NO2)9]5? are the dominant species in solutions.  相似文献   

4.
The formation of rhodium(III) sulfate complexes under moderately rigorous temperature conditions was studied by 103Rh and 17O NMR spectroscopy. The complexes [Rh2(μ-SO4)2(H2O)8]2+, [Rh2(*μ-SO4)(H2O)8]4+, and [Rh3(μ-SO4)3(μ-OH)(H2O)10]2+ were found to be the most stable species in aged solutions.  相似文献   

5.
The catalytic hydrogenation of CO was studied over Mn- and/or Fe-promoted Rh/γ-Al2O3 catalysts. The catalysts were characterized by means of XRD, BET, H2-TPR·H2-TPD, XPS and DRIFTS. CO hydrogenation results showed that the doubly Mn- and Fe-promoted Rh/γ-Al2O3 catalysts exhibited superior catalytic activity and better ethanol selectivity. The DRIFTS results showed that Mn promoter stabilized the adsorbed CO on Rh+ and Fe stabilized adsorbed CO on Rh+ and Rh0, especially Rh0. The fact that doubly Mn- and Fe-promoted Rh/γ-Al2O3 owned more (Rhx0–Rhy+)–O–Fe3+·(Fe2+) active species was proposed to be a crucial factor accounting for its higher ethanol selectivity.  相似文献   

6.
《Electroanalysis》2004,16(19):1622-1627
The pH‐dependence of the stationary open‐circuit potential Ei=0st of rhodium electrode with a surface layer of anodically formed insoluble compounds has been studied in sulfate and phosphate solutions by means of cyclic voltammetry and chronopotentiometry. The range of potentials of the investigations performed has been confined to the region of rhodium electrochemical oxidation/reduction, i.e., 0.2<E<1.2 V (RHE) in order to prevent any possible interference of other reactions such as H2 and O2 evolution. It has been shown that rhodium electrode with a layer of surface compounds formed anodically at E<<1.23 V (RHE) behaves like a reversible metal‐oxide electrode within the range of pH values from ca. 1.0 to ca. 8.0. It has been presumed that the stationary potential of such electrode is determined by the equilibrium of the following electrochemical reaction: Rh+3H2O??Rh(OH)3+3H++3e?. The pH‐dependence of the reversible potential of Eequation/tex2gif-inf-6.gif electrode has been found to be: Eequation/tex2gif-inf-8.gif=Ei=0st=0.69?0.059 pH, V. In acid solutions (pH<2.0) rhodium hydroxide dissolves into the electrolyte, therefore, to reach equilibrium, the solution must be saturated with Rh(OH)3. This has been achieved by adding Rh3+ ions in the form of Rh2(SO4)3. The solubility product of Rh(OH)3, estimated from the experimental Eequation/tex2gif-inf-16.gif?pH dependence obtained, is ca. 1.0×10?48, which is close to the value given in literature.  相似文献   

7.
The reaction of rhodium(I) carbonyl chloride, [Rh(CO)2Cl]2, with dichromate, cerium(IV) sulfate, hexachloroplatinic acid or p-benzoquinone in aqueous hydrochloric acid proceeds by consumption of 4 equivalents of oxidizing agent per mole or rhodium(I) in accordance with the equation RhI(CO)2  4e + H2O → RhIII(CO) + 2H+ + CO2A “cyclic” oxidation mechanism is suggested.  相似文献   

8.
Summary The ammoniation ofcis-[Rh(en)2Cl2] · (ClO4) in liquid NH3 was studied at constant ionic medium of 0.20 m perchlorate in the 0 to 35° range. The complex reacts in two distinct steps to givecis-[Rh(en)2(NH3)2] · (ClO4)3, with the intermediate formation ofcis-[Rh(en)2(NH3)Cl] · (ClO4)2. Both steps follow a conjugate-base mechanism. Activation parameters were obtained for the acid-base preequilibrium and the rate-determining step. The entropies of activation for the rate-determining step are 0 and –42 JK–1mol–1 for the first and second ammoniations respectively. These values are considerably lower than those found for the cobalt(III) analogues. The entropy changes for the acid-base equilibria are –84 and –36 JK–1mol–1 respectively, which is less negative than those values found for the cobalt(III) analogues. Trans-[Rh(en)2I2] · (ClO4) ammoniates totrans-[Rh(en)2(NH3)I] · (ClO4)2. The contribution of spontaneous ammoniation to the overall reaction oftrans-[Rh(en)2I2] · (ClO4) is negligible, so the uniqueness oftrans-[Co(en)2Cl2] · (ClO4) among cobalt(III) complexes in this respect is not reproduced for thetrans-dihalotetraamine structure in rhodium(III) complexes. A comparison of cobalt(III) and rhodium(III) amines with respect to activation parameters and the influence of formal charge of the metal complex on reactivity indicates a more associative type of activation for rhodium(III).  相似文献   

9.
The anionic rhodium carbonyl clusters [Rh7(CO)16]3− and [Rh14(CO)25]4− can be easily prepared by a new simple and high yield one-pot synthesis starting from RhCl3·nH2O dissolved in ethylene glycol and involving two steps: (i) treatment of RhCl3·nH2O under 1 atm of CO at 50 °C to give [Rh(CO)2Cl2]; (ii) addition of a base (CH3CO2Na or Na2CO3) followed by reductive carbonylation under 1 atm of CO at an adequate temperature (50 °C for [Rh7(CO)16]3−; 150 °C for [Rh14(CO)25]4−). These new syntheses are more convenient than those previously reported, especially since such clusters are not accessible via silica surface-mediated reactions. This different behavior is due to the particular stabilization on the silica surface and under 1 atm of CO of an anionic carbonyl cluster, called A, which does not allow the formation of a higher nuclearity carbonyl cluster, called B, which was shown to be the key-intermediate in the synthesis of [Rh14(CO)25]4− working in ethylene glycol solution. Although it was not possible to isolate crystals of A and B suitable for X-ray structural determination, a combination of cyclovoltammetry, one of the few examples so far available of the use of this technique for anionic rhodium carbonyl clusters, infrared spectroscopy and elemental analyses suggest that A and B are probably the never reported [Rh7(CO)14] and [Rh15(CO)28]3− clusters, respectively. In particular the tentative formulation of the two clusters was carried out by a non-conventional method based on the existence of a linear correlation between carbonyl frequencies of the main band and the [(charge/Rh atoms)/CO number] ratio.  相似文献   

10.
Molecular hydrogen (H2) is considered one of the most promising fuels to decarbonize the industrial and transportation sectors, and its photocatalytic production from molecular catalysts is a research field that is still abounding. The search for new molecular catalysts for H2 production with simple and easily synthesized ligands is still ongoing, and the terpyridine ligand with its particular electronic and coordination properties, is a good candidate to design new catalysts meeting these requirements. Herein, we have isolated the new mono-terpyridyl rhodium complex, [RhIII(tpy)(CH3CN)Cl2](CF3SO3) (Rh-tpy), and shown that it can act as a catalyst for the light-induced proton reduction into H2 in water in the presence of the [Ru(bpy)3]Cl2 (Ru) photosensitizer and ascorbate as sacrificial electron donor. Under photocatalytic conditions, in acetate buffer at pH 4.5 with 0.1 M of ascorbate and 530 μM of Ru, the Rh-tpy catalyst produces H2 with turnover number versus catalyst (TONCat*) of 300 at a Rh concentration of 10 μM, and up to 1000 at a concentration of 1 μM. The photocatalytic performance of Ru/Rh-tpy/HA/H2A has been also compared with that obtained with the bis-dimethyl-bipyridyl complex [RhIII(dmbpy)2Cl2]+ (Rh2) as a catalyst in the same experimental conditions. The investigation of the electrochemical properties of Rh-tpy in DMF solvent reveals that the two-electrons reduced state of the complex, the square-planar [RhI(tpy)Cl] (RhI-tpy), is quantitatively electrogenerated by bulk electrolysis. This complex is stable for hours under an inert atmosphere owing to the π-acceptor property of the terpyridine ligand that stabilizes the low oxidation states of the rhodium, making this catalyst less prone to degrade during photocatalysis. The π-acceptor property of terpyridine also confers to the Rh-tpy catalyst a moderately negative reduction potential (Epc(RhIII/RhI) = −0.83 V vs. SCE in DMF), making possible its reduction by the reduced state of Ru, [RuII(bpy)(bpy•−)]+ (Ru) (E1/2(RuII/Ru) = −1.50 V vs. SCE) generated by a reductive quenching of the Ru excited state (*Ru) by ascorbate during photocatalysis. A Stern–Volmer plot and transient absorption spectroscopy confirmed that the first step of the photocatalytic process is the reductive quenching of *Ru by ascorbate. The resulting reduced Ru species (Ru) were then able to activate the RhIII-tpy H2-evolving catalyst by reduction generating RhI-tpy, which can react with a proton on a sub-nanosecond time scale to form a RhIII(H)-tpy hydride, the key intermediate for H2 evolution.  相似文献   

11.
Multinuclear NMR data (13C, 31P, 13C–{31P}, 13C–{103Rh} and 31P–{103Rh}) for a series of mono- and di-substituted derivatives of Rh6(CO)16 containing neutral two electron donor ligands [Rh6(CO)15L, (L=NCMe, py, cyclooctene, PPh3, P(OPh)3,1/2(μ2,η1:η1-dppe)); Rh6(CO)14(LL), (LL=cis-CH2=CMe-CMe=CH2, dppm, dppe, (P(OPh)3)2)] are reported; these data show that the solid state structure is maintained in solution. Detailed assignments of the 13CO NMR spectra of Rh6(CO)15(PPh3) and Rh6(CO)14(dppm) clusters have been made on the basis 13C–{103Rh} double resonance measurements and the specific stereochemical features of the observed long range couplings in these clusters have been studied. The stereochemical dependence of 3J(P–C) for terminal carbonyl ligands is discussed and the values of 3J(P–C) are found to be mainly dependent on the bond angles in the P–Rh–Rh–C fragment; these data enable the fine structure of the complex multiplets in the 13C–{1H} and 31P–{1H} NMR spectra of Rh6(CO)14 (dppm) to be simulated. Variable temperature 13C–{1H} NMR measurements on Rh6(CO)15(PPh3) reveal the carbonyl ligands in this complex to be fluxional. The fluxional process involves exchange of all the CO ligands except the two terminal CO's associated with the rhodium trans to the substituted rhodium and can be explained by a simple oscillation of the PPh3 on the substituted rhodium atom aided by concomitant exchange of the unique terminal CO on this rhodium with adjacent μ3-CO's.  相似文献   

12.
The dimeric rhodium(II) complexes [Rh2(leu)4(H2O)2]- (ClO4)4 and [Rh2(pro)4(H2O)2](ClO4)4 have been prepared and characterized by elemental analyses, i.r., u.v.–vis. and 1H-n.m.r. spectroscopy. The amino acid molecules are coordinated as bridging ligands via their carboxylato groups. Cyclic voltammetry in DMF has shown that the complexes undergo a quasi-reversible reduction to yield dimers containing a Rh 2 3+ core. Oxidation processes within the 0–1.5V range were not observed.  相似文献   

13.
Binuclear Rh(II) compounds [Rh2(μ-OOCCH3)2(dbbpy)2(H2O)2](CH3COO)2 (1) (dbbpy = 4,4′-di-tert-butyl-2,2′-bipyridine), [Rh2(μ-OOCCH3)2(dbbpy)2(H2O)2](BF4)2·H2O·CH3CN (2), [Rh2(CH3COO)2(C18H24N2)2(CH3CN)2](BF4)2·4CH3CN (3) and {[Rh2(μ-OOCCH3)2(dbbpy)2][BF4]}n (4) have been synthesized and characterized with spectroscopic methods. Structure of complex 3 has been determined using X-ray crystallography. Rhodium atoms in compound 3 have distorted octahedral coordination with O and N atoms in equatorial positions and Rh atom and CH3CN molecule in axial coordination sites. Reduction of rhodium(II) compounds with aqueous 2-propanol leads to the formation of polymetallic compound {[Rh2(μ-OOCCH3)2(dbbpy)2][BF4]}n (4) containing [Rh2]3+ core. Compound 4 shows strong antiferromagnetic properties, μ = 0.18–1.73 M.B. in the range 1.8–300 K, J = −597 cm−1. Electrochemistry of compounds 3 and 4 in CH3CN has been investigated. Compound 4 exhibits a poorly reversible oxidation system at E1/2 = −0.92 V (ΔEp = 0.19 V) and in solution in DMF is slowly oxidized to 3 even in total absence of oxygen. Complex 3 is irreversibly oxidized to Rh(III) compound at Epa = 1.48 V and irreversibly reduced at Epc = −1.02 V to lead to the unstable polynuclear complex 4 in CH3CN.  相似文献   

14.
Cathodic electrodeposition of ZnSe is studied in aqueous acidic selenite (Se(IV)O2) baths of free Zn2+ ions and alkaline selenosulfite (Se0SO32–) baths of zinc complex ions. The synthesis of polycrystalline, cubic ZnSe by Se(IV)-diffusion-controlled electrodeposition from acidic solutions is optimized in terms of selenite concentration, while the effect of Ti, Ni, anodized Ti and CdSe substrates on the obtained layers growth and structural properties is investigated. Furthermore, the production of stoichiometric though amorphous ZnSe is demonstrated using selenosulfite solutions. The results are valuated by means of X-ray diffraction, scanning electron microscopy and reflectance spectroscopy techniques.  相似文献   

15.
Reduction of [Rh2(-OAc)2(bpy)2(H2O)2](OAc)2 and [Rh2Cl2(-OAc)2(bpy)2] · 3H2O complexes with ethanol and [Cr2(OAc)4(H2O)2] has been investigated using e.p.r. and u.v.–vis. spectra. The results indicate that stable complexes containing the [Rh2 3+] entity are not formed. The X-ray structure of [Rh2Cl2(-OAc)2(bpy)2] · 3H2O has been determined. Coordination around the Rh atom is in the form of a distorted octahedron. The complex shows an almost ideal eclipsed conformation. The equatorial coordination sites are occupied by bridging carboxylato ligands and 2,2-bipyridine and axial positions by the Cl ligand and the rhodium atom. The Rh–Rh distance is 2.574 Å.  相似文献   

16.
It was shown that the monomeric rhodium sulfate complexes [Rh(H2O)4(SO4)]+, trans-[Rh(H2O)2(SO4)2]?, cis-[Rh(H2O)2(SO4)2]?, and [Rh(SO4)3]3? were not predominant forms in aqueous solutions. The 103Rh NMR chemical shifts of the complexes were assigned, and the conditions for their formation in solutions, concentration parameters, and acidity at which the fraction of the monomers was maximal were determined. The constants of formation of the complexes and ion pair (IP) were estimated: K IP = 8 ± 3.5, K 1 ≈ 8, K 2trans ≈ 1, K 2cis ≈ 1, and K 3 ≈ 2.  相似文献   

17.
Ethanolic solutions of RhIII chloride exposed to γ-radiation under CO atmosphere are shown to be totally reduced into RhI complexes (Rh2(CO)4Cl2 and Rh(CO)2Cl-2) within a few hours with a radiolytic reduction yield of about 6.0 elementary reductions/100 eV (6.2·10-7 mol·J-1). The chloride ions freed in the medium inhibit further reduction through Rh(CO)2Cl-2 formation. On addition of copper metal under the same conditions, RhIII is transformed into Rh6(CO)16 with a conversion yield 50%. This cluster is formed via Rh2(CO)4Cl2 although Rh(CO)2Cl-2 is also present under these conditions. Rh6(CO)16 cluster is also formed under radiolysis by direct reduction of Rh2(CO)4Cl2, but metallic rhodium and other reduced products are obtained at the same time.  相似文献   

18.
Overall kinetic and chronopotentiometric studies were performed during Al anodising in H2SO4, 0–5% w/v, bath solutions pure and saturated by Al2(SO4)3. Peculiarities in film growth mechanism and nanostructure in these cases appeared, like significant differences of porosity and its dependence on film thickness, different critical current density above which pitting appears, salt deposition on pitted surface regions in saturated bath, etc. The different conditions inside pores are responsible for this behaviour like almost depletion of H+ during a long initial transient stage in the first case, supersaturation and formation of Al2(SO4)3 nanoparticle micelles on pore surface in the second case, etc. Differences in film growth mechanism also appeared between these and alike baths at higher acidity. Anodising in low acidity saturated baths shows superiority for growing low porosity films at specific conditions. New technologies may be suggested to produce optimal films of desired structure.  相似文献   

19.
Complex formation in a system Rh(III)-H2SO4-H2O was studied by the 103Rh and 17O NMR spectroscopy at room temperature. The formation of two interrelated systems of mononuclear and polynuclear complexes was established in the above solutions. The predominant species in the first system is a labile ionic pair {Rh(H2O) 6 3+ SO 4 2? }+, while in the second system, two inert binuclear complexes [Rh2(μ-SO4)2(H2O)8]2+ and [Rh2(μ-SO4)(μ-OH)(H2O)8]3+ prevail.  相似文献   

20.
On X-ray photoelectron spectra of the Au-Rh/TiO2 catalysts the position of Au4f peak was practically unaffected by the presence of rhodium, the peak position of Rh3d, however, shifted to lower binding energy with the increase of gold content of the catalysts. Rh enrichment in the outer layers of the bimetallic crystallites was experienced. The bands due to Au0-CO, Rh0-CO and (Rh0)2-CO were observed on the IR spectra of bimetallic samples, no signs for Rh+-(CO)2 were detected on these catalysts. The results were interpreted by electron donation from titania through gold to rhodium and by the higher particle size of bimetallic crystallites.  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