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.
Comparison of extraction properties of macrocyclic calix[4,6]arenethiaethers (CATE) with their acyclic analogs R2S (R = C6H13, C8H17) for the recovery of [Rh(NO2)3(H2O)3]0 rhodium form from nitric acid solutions was carried out. Rhodium recovery with CATE (0.05 M) in the absence of accelerating additives under optimal conditions exceeds 90% at 5–10-fold preconcentration and is only 1–3% for R2S (1 M). Extraction kinetics was studied and hypothesis on the mechanism of multiple acceleration of rhodium recovery was proposed for CATE extraction, the mechanism includes the formation of intermediate product of colloidal- chemical nature on account of the surface activity of the macrocycle and its reaction with rhodium accompanied by rhodium chelation to the sulfur atoms of neighboring fragments of the macrocycle. The obtained results are of interest for the development of methods for the isolation of fission rhodium from nitrate–nitrite nitric acid solutions.  相似文献   

2.
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   

3.
Complex rhodium(III) salts of the composition cis-M[Rh(NH3)2(NO2)4nH2O, where M = K+, Cs+, Ag+, (CH3)4N+, are synthesized and described. Their molecular and crystal structures are determined by single crystal X-ray diffraction.  相似文献   

4.
The state of ruthenium in conjugated phases upon extraction of trans-[Ru(15NO)(15NO2)4(OH)]2? complex with tri-n-octylphosphine oxide (TOPO) in the presence of Zn2+ and subsequent back extraction with H15NO3 and NH3(concd.) solutions was studied by 15N NMR. Binuclear complexes [Ru(NO)(NO2)5?n (μ-NO2) n?1(μ-OH)Zn(TOPO) n ] and [Ru(NO)(NO2)4?n (ONO)(μ-NO2) n?1(μ-OH)Zn(TOPO) n ], where n = 2, 3, are predominant forms in extract. Kinetic restrictions for ruthenium extraction with TOPO solution in hexane and its back extraction with aqueous solutions of nitric acid and ammonia are eliminated in the absence of direct coordination of extractant to ruthenium. fac-Dinitronitrosyl forms [Ru(NO)(H2O)3(NO2)2]+, [Ru(NO)(H2O)2(NO2)2(NO3)]0 (3 and 6 M HNO3) and [Ru(NO)(H2O)(NO2)2(NO3)2]? (6 M HNO3) prevail in nitric acid back extracts. Equilibrium constant at ambient temperature (0.05 ± 0.01) was assessed for the coordination of second nitrate ion to nitrosylruthenium dinitronitrato complex. Complex species [Ru(NO)(NO2)4(OH)]2? and [Ru(NO)(NO2)3(ONO)(OH)]2? prevail in ammonia back extract.  相似文献   

5.
The extraction of rhodium and iridium with 4-(non-5-yl)pyridine (NP) was investigated. The rate of rhodium extraction increases with increasing concentration of NP and chloride ions. Spectroscopic studies indicate that the extracted species is an ion pair, RhCl3?6 3HNP+. Under the conditions of optimum Rh extraction ([Cl?]=3.7 M, [NP]=0.3 M, [H]=0.08 M), iridium is also extracted by NP with similar efficiency in the form of IrCl3?6 3HNP+. The use of hypophosphorous acid to labilize rhodium results in a better extraction of rhodium without significantly changing the extraction of iridium. The efficiency and kinetics of the rhodium extraction improve with increasing chloride concentration. For [Cl?] ? 3.7 M, [H3PO2]=2.5 M, [NP]=0.3 M and Ph ≈ 1.6, 82% of rhodium is extracted in 4 min and 95% in 30 min.  相似文献   

6.
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   

7.
Recently described and fully characterized trinuclear rhodium‐hydride complexes [{Rh(PP*)H}32‐H)33‐H)][anion]2 have been investigated with respect to their formation and role under the conditions of asymmetric hydrogenation. Catalyst–substrate complexes with mac (methyl (Z)‐ N‐acetylaminocinnamate) ([Rh(tBu‐BisP*)(mac)]BF4, [Rh(Tangphos)(mac)]BF4, [Rh(Me‐BPE)(mac)]BF4, [Rh(DCPE)(mac)]BF4, [Rh(DCPB)(mac)]BF4), as well as rhodium‐hydride species, both mono‐([Rh(Tangphos)‐ H2(MeOH)2]BF4, [Rh(Me‐BPE)H2(MeOH)2]BF4), and dinuclear ([{Rh(DCPE)H}22‐H)3]BF4, [{Rh(DCPB)H}22‐H)3]BF4), are described. A plausible reaction sequence for the formation of the trinuclear rhodium‐hydride complexes is discussed. Evidence is provided that the presence of multinuclear rhodium‐hydride complexes should be taken into account when discussing the mechanism of rhodium‐promoted asymmetric hydrogenation.  相似文献   

8.
The structures of two salts [Co(NH3)6][Rh(NO2)6] (I) and [Co(NH3)6][(NO2)3Rh(μ-NO2)1+x (μ-OH)2−x Rh(NO2)3]·(2−x)(H2O), x = 0.17 (II) are solved. Single crystals of the salts are obtained by the counter diffusion method through the gel of aqueous solutions of [Co(NH3)6]Cl3 and Na3[Rh(NO2)6]. The structure of [Co(NH3)6][Rh(NO2)6] is consistent with the diffraction data for a polycrystalline sample of poorly soluble fine salt formed in the exchange reaction between aqueous solutions of [Co(NH3)6]Cl3 and Na3[Rh(NO2)6]. The structure of [Co(NH3)6][(NO2)3Rh(μ-NO2)1+x (μ-OH)2−x Rh(NO2)3]·(2−x)(H2O), x = 0.17 exhibits the stabilizing effect of a large cation in the formation of novel, unknown previously coordination ions: [(NO2)3Rh(μ-NO2)(μ-OH)2Rh(NO2)3]3− and [(NO2)3Rh(μ-NO2)2(μ-OH)Rh(NO2)3]3−.  相似文献   

9.
The separation and isolation of many of the platinum group metals (PGMs) is currently achieved commercially using solvent extraction processes. The extraction of rhodium is problematic however, as a variety of complexes of the form [RhCln(H2O)6-n](n−3)− are found in hydrochloric acid, making it difficult to design a reagent that can extract all the rhodium. In this work, the synergistic combination of a primary amine (2-ethylhexylamine, LA) with a primary amide (3,5,5-trimethylhexanamide, L1) is shown to extract over 85 % of rhodium from 4 M hydrochloric acid. Two rhodium complexes are shown to reside in the organic phase, the ion-pair [HLA]3[RhCl6] and the amide complex [HLA]2[RhCl5(L1)]; in the latter complex, the amide is tautomerized to its enol form and coordinated to the rhodium centre through the nitrogen atom. This insight highlights the need for ligands that target specific metal complexes in the aqueous phase and provides an efficient synergistic solution for the solvent extraction of rhodium.  相似文献   

10.
A detailed analysis of the 35Cl/37Cl isotope effects observed in the 19.11 MHz 103Rh NMR resonances of [RhCln(H2O)6−n]3−n complexes (n = 3–6) in acidic solution at 292.1 K, shows that the ‘fine structure’ of each 103Rh resonance can be understood in terms of the unique isotopologue and in certain instances the isotopomer distribution in each complex. These 35Cl/37Cl isotope effects in the 103Rh NMR resonance of the [Rh35/37Cl6]3− species manifest only as a result of the statistically expected 35Cl/37Cl isotopologues, whereas for the aquated species such as for example [Rh35/37Cl5(H2O)]2−, cis-[Rh35/37Cl4(H2O)2] as well as the mer-[Rh35/37Cl3(H2O)3] complexes, additional fine-structure due to the various possible isotopomers within each class of isotopologues, is visible. Of interest is the possibility of the direct identification of stereoisomers cis-[RhCl4(H2O)2], trans-[RhCl4(H2O)2], fac-[RhCl3(H2O)3] and mer-[RhCl3(H2O)3] based on the 103Rh NMR line shape, other than on the basis of their very similar δ(103Rh) chemical shift. The 103Rh NMR resonance structure thus serves as a novel and unique ‘NMR-fingerprint’ leading to the unambiguous assignment of [RhCln(H2O)6−n]3−n complexes (n = 3–6), without reliance on accurate δ(103Rh) chemical shifts.  相似文献   

11.
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   

12.
Solutions of platinum(IV) nitrate were studied by 195Pt, 15N, 14N, and 17O NMR and IR and Raman spectroscopy. It was found that in nitric acid, two interrelated systems of nitrate complexes, mono- and polynuclear ones, coexist. The complexes predominating in concentrated solutions are [Pt2(μ-OH)(μ-NO3)(NO3)2(H2O)6 ? x (OH) x ](4 ? x)+, [Pt4(μ-OH)3(μ-NO3)3(NO3)3(H2O)9 ? x (OH) x ](7 ? x)+, [Pt4(μ-OH)4(μ-NO3)2(NO3)4(H2O)8 ? x (OH) x ](6 ? x)+, and [Pt4(μ-OH)6(NO3)3(H2O)16 ? x (OH) x ](7 ? x)+.  相似文献   

13.
The manifold enhancement of the rhodium extraction efficiency from nitric acid solutions of triaquatrinitrorhodium is discovered with the use of two-component mixed extractants based on alkylaniline (AA), dihexyl sulfide (DHS), dihexyl sulfoxide (DHSO), tributyl phosphate (TBP), and oxime ACORGA P5100. A mixture of unimolar solutions of alkylanilinium nitrate and DHS is found to be the most efficient extractant: at 35°C, this mixture quantitatively extracts rhodium within 5 min from aqueous solutions containing 0.06 to 3 mol/L HNO3. The extraction kinetics are studied. The following two-stage extraction mechanism is substantiated: the equilibrium of formation of a colloidal-chemical intermediate involving [Rh(H2O)3(NO2)3], HNO3, and (BHNO3)p (an associated form of the alkylanilinium salt) and the reaction of the intermediate with DHS (the rate-controlling stage).  相似文献   

14.
Kinetic studies in homogeneously Rh(III)-catalyzed oxidation of reducing sugars, i.e. maltose and lactose, by N-bromoacetamide (NBA) in the presence of perchloric acid have been made at 40 °C using mercuric acetate as Br ion scavenger. The results obtained for the oxidation of both reducing sugars show first-order dependence of the reactions on NBA at its low concentrations, which shifts towards zero-order at its higher concentrations. First-order kinetics in [Rh(III)] and zero-order kinetics in [reducing sugar] were observed. Positive effect of [Cl] was observed in the oxidation of both maltose and lactose. Order of reaction was found to be one and half (1.5) throughout the variation of [H+] in the oxidation of both maltose and lactose. An increase in the rate of reaction with the decrease in [Hg(OAc)2] and [NHA] was observed for both the redox systems. The rate of oxidation is unaffected by the change in ionic strength (μ) of the medium. The main oxidation products of the reactions were identified as formic acid and arabinonic acid in the case of maltose and formic acid, arabinonic acid and lyxonic acid in the case of lactose. A common mechanism for the oxidation of both maltose and lactose, showing the formation of most reactive activated complex, [RhCl4(H3O)H2OBr]+, and an unreactive complex, [RhCl4(H2O)(H2OBrHg)]2+, has been proposed. Various activation parameters have also been calculated and on the basis of these parameters, a suitable explanation for the reaction mechanism has been given.  相似文献   

15.
The transition state for the oxidative addition reaction [Rh(acac)(P(OPh)3)2] + CH3I, as well as two simplified models viz. [Rh(acac)(P(OCH3)3)2] and [Rh(acac)(P(OH)3)2], are calculated with the density functional theory (DFT) at the PW91/TZP level of theory. The full experimental model, as well as the simplified model systems, gives a good account of the experimental Rh-ligand bond lengths of both the rhodium(I) and rhodium(III) β-diketonatobis(triphenylphosphite) complexes. The relative stability of the four possible rhodium(III) reaction products is the same for all the models, with trans-[Rh(acac)(P(OPh)3)2(CH3)(I)] (in agreement with experimental data) as the most stable reaction product. The best agreement between the theoretical and experimental activation parameters was obtained for the full experimental system.  相似文献   

16.
Cationic methyl complex of rhodium(III), cis-[Rh(Acac)(PPh3)2(CH3)(Py)][BPh4] (1) as a single isomer with Py in the trans to PPh3 position, is formed upon the reaction of cis-[Rh(Acac)(PPh3)2(CH3)(CH3CN)][BPh4] with pyridine in methylene chloride solution.Complex 1 was characterized by elemental analysis and by 31P{1H} and 1H NMR spectra.Cationic pentacoordinate acetyl complexes, trans-[Rh(Acac)(PPh3)2(COCH3)][BPh4] (2) and trans-[Rh(BA)(PPh3)2(COCH3)][BPh4] (3), are prepared by action of carbon monoxide on cis-[Rh(Acac)(PPh3)2(CH3)(CH3CN)][BPh4] and cis-[Rh(BA)(PPh3)2(CH3)(CH3CN)][BPh4], respectively, in methylene chloride solutions.Complexes 2 and 3 were characterized by elemental analysis and by IR, 31P{1H}, 13C{1H} and 1H NMR. According to NMR data, 2 and 3 in solution are non-fluxional trigonal bipyramids with β-diketonate and acetyl ligands in the equatorial plane and axial phosphines.In solutions, 2 and 3 gradually isomerize into octahedral methyl carbonyl complexes trans-[Rh(Acac)(PPh3)2(CO)(CH3)][BPh4] (4) and trans-[Rh(BA)(PPh3)2(CO)(CH3)][BPh4] (5), respectively.Complexes 4 and 5 were characterized by IR, 31P{1H}, 13C{1H} and 1H NMR, without isolation.Upon the action of PPh3 on cis-[Rh(Acac)(PPh3)2(CH3)(CH3CN)][BPh4] and cis-[Rh(BA)(PPh3)2(CH3)(CH3CN)] [BPh4], reductive elimination of the methyl ligand as a phosphonium salt, [CH3PPh3][BPh4], occurs to give square planar rhodium(I) complexes [Rh(Acac)(PPh3)2] and[Rh(BA)(PPh3)2], respectively. The reaction products were identified in the reaction mixtures by 31P{1H} and 1H NMR.  相似文献   

17.
The kinetics of the reaction [Rh(H2O)6]3+ + H3PO4 ? [Rh(H2O)5H2PO4]2+ + H3O+ has been studied by 31P NMR; E a = 142 ± 12 kJ/mol, logA = 17 ± 2. An empirical dependence of the 31P NMR chemical shift on the equilibrium pH value has been found. The acid dissociation constant of the coordinated ion has been evaluated: pK = 1.5. The 31P NMR chemical shifts of individual [Rh(H2O)5H2PO4]2+ and [Rh(H2O)5HPO4]+ complex ions are, respectively, 14.5 and 15.8 ppm at 323 K.  相似文献   

18.
The reaction of Rh(CO)2acac with triphenylantimony in the presence of cesium benzoate in tetraethylene glycol/dimethyl ether solution resulted in the selective formation of [Rh12Sb(CO)27]3- (66% yield) after 3 h of contact time under ≈400 atm of carbon monixide and hydrogen (CO/H2  1) at 140–160°C. The cluster has been isolated as the [Cs(18-Crown-6)2]+, [(CH3)4]+, [(C2H5)4N]+, (Ph3P)2N]+ and [PhCH2N(C2H5)3]+ salts. The [(C2H5)4N]3 [Rh12Sb(CO)27] complex has been characterized via a complete three-dimensional X-ray diffraction study. The complex crystallizes in the space group R3c with a  23.258(13) Å, c  22.811(4) Å, V  10 686 Å3 and p(calcd.)  2.334 g cm-3 for mol.wt. 2503.66 and Z  6. Diffraction data were collected with an Enraf-Nonius CAD 4 automated diffractometer using graphite-monochromatized Mo-Kα radiation. The structure was solved by direct methods and refined by difference-Fourier and least-squares techniques. All non-hydrogen atoms have been located and refined: final discrepancy indices are Rf  3.5% and Rwf  4.6% for 3011 reflections. The anion's structure consists of twelve rhodium atoms situated at the corners of a distorted icosahedron with contacts of 2.807(1), 2.861(1), 2.874(1), 2.999(1), 3.017(1) and 3.334(1) Å and rhodium—antimony contacts of 2.712(0) Å. Rhodium—rhodium bond distances of 2.807 and 3.017 Å are in the range usually found for these complexes although a distance of 3.334 Å may be longer than expected from bonding interactions. The sum of the covalent radii of antimony and rhodium, 2.80 Å, is intermediate between the two observed RhSb contacts. The anion cluster structure is that of distorted icosahedron. This polyhedron has previously been found in [B12H12]2- but not with transition metal clusters. A comparison between the structures of rhodium carbonyl clusters and boranes shows the occurrence of similar structural features. Applications of bonding theories based on the boranes, such as Wade's rules, to rhodium carbonyl clusters shows the extent in which these rules are obeyed.  相似文献   

19.
The mixed-metal vinylidene clusters HFe3Rh(CO)11(CCHR) (R = H, C6H5) have been synthesized via the reaction of [HFe3(CO)3CCHR][P(C6H5)4] with [RhCl(CO)2]2 in the presence of a thallium salt. The reaction initially gives the [Fe3Rh(CO)11]CCHR][P(C6H5)4] cluster which leads to the final products by protonation. Spectroscopic data indicate a μ42 mode of bonding for the vinylidene ligand. A structure with a Fe3Rh core in a butterfly configuration and in which the rhodium atom occupy a wing-tip site is proposed. The catalytic activity of HFe3Rh(CO)11(CCH(C6H5)) (80% yield) has been checked in hydroformylation and hydrogenation. In hydroformylation the cluster shows the same activity as Rh4(CO)12, whereas in hydrogenation the mixed-metal system shows specific activity; isomerization of 1-heptene to cis and trans 2-heptene takes place with no more than 14% heptane formation. The cluster is broken down during the catalysis, and some H3Fe3CO)93-CCH2(C6H5)) is formed. The latter cluster is not an active catalyst, and under the same conditions use of Rh4(CO)12 results mainly in hydrogenation of 1-heptene. These observations suggest that the active species is a mixed iron-rhodium system.  相似文献   

20.
The reactions of several mono- and poly-nuclear carbonyl metallates with nitrosonium ion have been studied. Besides simple substitution of a carbon monoxide with NO+ some reactions yielded products containing other nitrogeneous ligands. When [CoRu3(CO)13]? reacts with NO+, low yields of the new nitrido cluster CoRu3N(CO)12 are formed. Prior conversion of [CoRu3(CO)13]? to the new hydrido cluster [H2CoRu3(CO)12]? under hydrogen, followed by nitrosylation, forms the new imido cluster H2Ru3(NH)(CO)9 in very low yield. The reaction of [FeCO3(CO)12]? with NO+ also generates an imido cluster, FeCo2(NH)(CO)9, in 15% yield. This cluster has been characterized by X-ray crystallography and was found to be similar to the tricobalt alkylidyne clusters. (Triclinic crystal system, P1 space group, Z=2, a 6.787(1), b 8.016(1), c 13.881(2) Å, α 95.50(1), β 100.77(1), γ 107.93(1)°. Modifications of the nitrosylations using NO+ were studied. In particular, the addition of triethylamine or N-t-butylbenzaldimine allowed the use of NO+ in THF without solvent decomposition. With [CpMo(CO)3]? and [CpFe(CO)2]? the N-nitrosoiminium species appears to form transient alkylmetals which further react to give the dimers [CpMo(CO)3]2 and [CpFe(CO)2]2.  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