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 74 毫秒
1.
Studies on Oxide Catalysts. XLii. Redox Behaviour of Nickel in Zeolites NiNa? Y. 4. Influence of Composition on the Reducibility of Nickel in Zeolites NiNa? Y By chemical analysis (reaction with K2Cr2O7) and ESCA investigations we determined the degree of reduction in reduced samples NiNa-Y as function of the mole ratio SiO2/Al2O3 (module), of the Ni2+ degree of exchange and the kind of the second cations. (NH4+, Ca2+, Co2+, and Nd3+) in the temperature region of 620–770 K. The degree of nickel reduction increases with increasing module, decreasing degree of exchange and decreasing number of Brönsted acidic centres. This behaviour is caused by the influence of the interaction between cations Ni2+ and zeolite lattice on the reduction equilibrium.  相似文献   

2.
Studies on Oxide Catalysts. XXXIV. Redoxbehaviour of Nickel in Zeolite NiNaY. 1. Reducibility and Reoxidizability of Nickel in Zeolites NiNaY The properties of metallic nickel in reduced (470–870 K) and reoxidized (470, 670 K) samples were studied by chemical analysis (reaction with K2Cr2O7) and spectroscopic methods (FMR, IR after CO adsorption, UV/VIS). The reduction of Ni2+ cations from oxidic clusters proceeds in an onestep reaction. Contrary to this, isolated Ni2+ cations are reduced stepwise to Ni+ cations and subsequently to metallic nickel. The reduction degree depends in characteristic manner on the reduction temperature. Metallic nickel which was reduced at temperatures < 620 K, can be completely reoxidized at 470 K. Higher temperatures result in metallic aggregations which are not completely reoxidized even at 670 K.  相似文献   

3.
Characterization and Catalytic Activity of Ni2+ -X and -Y Zeolites. II. Reducibility of Ni2+ by Low Olefines and the Dimerization Activity of the Ni2+ -Zeolites The reducibility of Ni2+ in X and Y zeolites by hydrogen, but-1-ene, propene, and ethene is compared. The degree of reduction was determined after isothermal reduction and reoxidation by the TPR method. At 673 K on X zeolites the reducibility decreases in the order: H2 > but-1-ene, propene > ethene. On Y zeolites an inversion takes place: but-1-ene, propene > H2, ethene. The mechanism of reduction by olefins should be determined by an intermediate splitting off of a hydride ion as a reducing species. Such a mechanism explains the higher degree of reduction in the more acid Y zeolites. Assuming low valent nickel as an active center in ethen dimerization the induction period results from the reduction of Ni2+ ions.  相似文献   

4.
Penicillamine Complexes of Nickel, Chromium, and Molybdenum — Structural Particularity and Biological/Medical Relevance The compounds Tl2[NiII(H2O)6][NiII(D-pen)(L-pen)]2[NiII(SCN)2(H2O)4] 1 , Tl[NiII(D-pen)2H] · H2O 2 , Tl[CrIII(D-pen)2] 3 , and Na2[MoO4(pen)2] · 3 CH3OH · 3 H2O 4 have been prepared by the reaction of nickel nitrate (for 1 ), nickel acetate (for 2 ), potassium chromate (for 3 ), and sodium molybdate (for 4 ) with D- and D, L-penicillamine, respectively. They were characterized by single-crystal X-ray structure analysis and other physical methods. Whereas penicillamine acts as a bidentate (N, S)-ligand in 1 and 2 , CrIII (in 3 ), and MoV (in 4 ) are coordinated to the three ligand atoms N, O, and S. The presence of three different types of NiII-complexes a cationic, a neutral, and an anionic one in 1 is remarkable. For crystal data see Inhaltsübersicht.  相似文献   

5.
By using the node‐and‐spacer approach in suitable solvents, four new heterotrimetallic 1D chain‐like compounds (that is, containing 3d–3d′–4f metal ions), {[Ni(L)Ln(NO3)2(H2O)Fe(Tp*)(CN)3] ? 2 CH3CN ? CH3OH}n (H2L=N,N′‐bis(3‐methoxysalicylidene)‐1,3‐diaminopropane, Tp*=hydridotris(3,5‐dimethylpyrazol‐1‐yl)borate; Ln=Gd ( 1 ), Dy ( 2 ), Tb ( 3 ), Nd ( 4 )), have been synthesized and structurally characterized. All of these compounds are made up of a neutral cyanide‐ and phenolate‐bridged heterotrimetallic chain, with a {? Fe? C?N? Ni(? O? Ln)? N?C? }n repeat unit. Within these chains, each [(Tp*)Fe(CN)3]? entity binds to the NiII ion of the [Ni(L)Ln(NO3)2(H2O)]+ motif through two of its three cyanide groups in a cis mode, whereas each [Ni(L)Ln(NO3)2(H2O)]+ unit is linked to two [(Tp*)Fe(CN)3]? ions through the NiII ion in a trans mode. In the [Ni(L)Ln(NO3)2(H2O)]+ unit, the NiII and LnIII ions are bridged to one other through two phenolic oxygen atoms of the ligand (L). Compounds 1 – 4 are rare examples of 1D cyanide‐ and phenolate‐bridged 3d–3d′–4f helical chain compounds. As expected, strong ferromagnetic interactions are observed between neighboring FeIII and NiII ions through a cyanide bridge and between neighboring NiII and LnIII (except for NdIII) ions through two phenolate bridges. Further magnetic studies show that all of these compounds exhibit single‐chain magnetic behavior. Compound 2 exhibits the highest effective energy barrier (58.2 K) for the reversal of magnetization in 3d/4d/5d–4f heterotrimetallic single‐chain magnets.  相似文献   

6.
Zusammenfassung Der Komplex des Ni2+ mit o-Methylbenzamidoxim (oMB), [Ni(oMB)2], wurde mit H2O2 in Anwesenheit von Cu2+ oder unter Luftdurchblasen in alkalischer Lösung mit O2 zu [Ni(oMB)4] oxydiert. Die Ligandenzahl und die Bildungskonstante sindn=4 und lgK=7,33. Die Vierwertigkeit des Nickels wurde analytisch nachgewiesen. Die Oxydation verändert dasoMB nicht, wie aus dem IR-Spektrum ersichtlich ist. Die Geschwindigkeitskonstante istk=36,3 l/mol–1 sec–1 und entspricht einer bimolekularen Reaktion.
Complexes of quadrivalent Ni with o-methylbenzamide oxime
The complex of Ni2+ with o-methylbenzamide oxime [Ni(oMB)2] was oxidized to [Ni(oMB)4] with H2O2 in the presence of Cu2+ or with O2 by air-blowing in alkaline solution. The ligand number and the formation constant aren=4 and lgK=7.33, resp. It was proved analytically that nickel is 4-valent. The IR-spectra showed that the oxidation does not attackoMB. The speed constant isk=36.3 l/mole–1 sec–1, corresponding to a bimolecular reaction.
  相似文献   

7.
Zusammenfassung Laugefeuchte Gemische aus metallischem Nickel und dessen Hydroxiden bzw. Oxidhydraten verschiedener Wertigkeitsstufen können in Proben von 50–200 mg bezüglich Zusammensetzung und effektiver Wertigkeit nach folgendem Schema analysiert werden:Reduktion der Probe mit alkalischer Arsenat(III)-Lösung, Umsatz des entstandenen Arsenats(V) mit KJ zu photometrisch bestimmbarem elementarem Jod, Extraktion der im Rückstand enthaltenen Nickelverbindungen mit verd. H2SO4, Auflösung des metallischen Nickels in H2SO4 + H2O2 und photometrische Bestimmung der Ni2+-Lösungen als ÄDTA-Komplex. Die effektive Wertigkeit der Nickelverbindungen kann auf ±0,02 Wertigkeitseinheiten genau ermittelt werden. Die Ergebnisse stimmen gut mit elektrochemisch erhaltenen Werten überein.Beachtet werden muß ein starker Phototropieeffekt der zu photometrierenden Jodlösung. Ein 380facher Überschuß an As(III) stört die As(V)-Bestimmung nicht.
Analysis of mixtures of metallic nickel and oxide-hydrates of bivalent and higher-valency nickel
Mixtures consisting of metallic nickel and the hydroxides or oxide-hydrates in various valency stages moistened with alkali in samples of 50 to 200 mg can be analysed by the following scheme: Reduction of the sample with alkaline As(III) solution, reaction of As(V) with KI setting free elementary, photometrically determinable iodine, extraction of the Ni-compounds present in the residue with dilute H2SO4, dissolution of the metallic M in H2SO4 + H2O2, and photometric determination of the Ni2+ solutions as EDTA complex.In this manner the effective valency of the Ni-compounds can be determined within ± 0.02 valency units, The results agree well with those obtained electrochemically.The iodine solution from which the extinction has to be measured photometrically shows a strong phototropic effect.An excess of 380 parts of As(III) does not interfere with the As(V) determination.


Auszugsweise vorgetragen auf der Hauptversammlung der Gesellschaft Deutscher Chemiker, 13. bis 18. September 1971 in Karlsruhe [6].

Dem Vorstand der VARTA AGdanke ich für die freundliche Genehmigung zur Publikation. Besonderer Dank gebührt Fräulein H. Slomp, die alle Arbeiten durchgeführt und durch ihre Fähigkeit zu beobachten und zu kombinieren wesentlich zur Lösung des Problems beigetragen hat.  相似文献   

8.
Irradiations of Ni/TiO2 catalyst by UV in hydrogen at 77 K produced not only Ni+ ions on the catalyst surface, but also Ni3+ and Ti3+ species in bulk or near the interface between nickel and titania. These photo-generated species were detected and characterized by low temperature electron paramagnetic resonance (EPR) spectroscopy. Relative spin concentrations of the photogenerated paramagnetic species (Nin+ and Ti3+) varied with the nickel content in titania. A high nickel content in the sample resulted in a high peak intensity ratio of Nin+ to Ti3+. It was found that the photoinduced self-redox reaction of Ni2+ ions to form Ni+ and Ni3+ ions has a priority over the photoreduction of Ti4+ to Ti3+ ions. The characteristic EPR spectrum of the Ni3+ (3d7) ions with g1 = 2.268, g2 = 2.237, and g3 = 2.045 indicates that the Ni3+ ions are most likely located in the substitutional sites of TiO2, possibly near the surface rutile phase. The Ni+ species (3d9) with g4 = 2.130 and g1 = 2.063 are on the surface of TiO2. Both Ni+ and Ni3+ ions are quite stable in hydrogen. The Ni3+ ions seem to be responsible for anchoring the nickel ions onto titania and stablizing the Ni+ species on the surface. The Ni+ ions are thus free from oxygen poisoning and still show a high activity toward olefin oligomerization.  相似文献   

9.
Reactions that proceed within mixed ethylene–methanol cluster ions were studied using an electron impact time-of-flight mass spectrometer. The ion abundance ratio, [(C2H4)n(CH3OH)mH+]/[(C2H4)n(CH3OH)m+], shows a propensity to increase as the ethylene/methanol mixing ratio increases, indicating that the proton is preferentially bound to a methanol molecule in the heterocluster ions. The results from isotope-labelling experiments indicate that the effective formation of a protonated heterocluster is responsible for ethylene molecules in the clusters. The observed (C2H4)n(CH3OH)m+ and (C2H4)n(CH3OH)m–1CH3O+ ions are interpreted as a consequence of the ion–neutral complex and intracluster ion–molecule reaction, respectively. Experimental evidence for the stable configurations of heterocluster species is found from the distinct abundance distributions of these ions and also from the observation of fragment peaks in the mass spectra. Investigations on the relative cluster ion distribution under various conditions suggest that (C2H4)n(CH3OH)mH+ ions with n + m ≤ 3 have particularly stable structures. The result is understood on the basis of ion–molecule condensation reactions, leading to the formation of fragment ions, $ {\rm CH}_2=\!=\mathop {\rm O}\limits^ + {\rm CH}_3 $ and (CH3OH)H3O+, and the effective stabilization by a polar molecule. The reaction energies of proposed mechanisms are presented for (C2H4)n(CH3OH)mH+(n + m ≤ 3) using semi-empirical molecular orbital calculations.  相似文献   

10.
Nickel ferrite nanospheres were successfully synthesized by a reverse emulsion-assisted hydrothermal method. The reverse emulsion was composed of water, cetyltrimethyl ammonium bromide, polyoxyethylene(10)nonyl phenyl ether, iso-amyl alcohol and hexane. During the hydrothermal process, β-FeO(OH) and Ni0.75Fe0.25(CO3)0.125(OH)2·0.38H2O (INCHH) nanorods formed first and then transformed into nickel spinel ferrite nanospheres. The phase transformation mechanism is proposed based on the results of X-ray powder diffraction, transmission electron microscopy and energy-dispersive X-ray spectroscopy, etc. Nickel ferrite may form at the end of the INCHH nanorods or from the solution accompanied by the dissolution of β-FeO(OH) and INCHH nanorods. The X-ray photoelectron spectroscopy analysis shows that a few Fe3+ ions have been reduced to Fe2+ ions during the formation of nickel ferrite. The maximum magnetization of the nickel ferrite nanospheres obtained after hydrothermal reaction for 30 h is 55.01 emu/g, which is close to that of bulk NiFe2O4.  相似文献   

11.
A porous metal–organic framework (MOF), [Ni2(dobdc)(H2O)2]?6 H2O (Ni2(dobdc) or Ni‐MOF‐74; dobdc4?=2,5‐dioxido‐1,4‐benzenedicarboxylate) with hexagonal channels was synthesized using a microwave‐assisted solvothermal reaction. Soaking Ni2(dobdc) in sulfuric acid solutions at different pH values afforded new proton‐conducting frameworks, H+@Ni2(dobdc). At pH 1.8, the acidified MOF shows proton conductivity of 2.2×10?2 S cm?1 at 80 °C and 95 % relative humidity (RH), approaching the highest values reported for MOFs. Proton conduction occurs via the Grotthuss mechanism with a significantly low activation energy as compared to other proton‐conducting MOFs. Protonated water clusters within the pores of H+@Ni2(dobdc) play an important role in the conduction process.  相似文献   

12.
Ni/SiO2 materials with identical composition (SiO2/Ni = 1.0) have been synthesized by precipitation of Ni(NO3)2 · 6H2O solution with Na2CO3 solution on the silica gel, obtained at three different pH values. The present investigation was undertaken in an endeavor to study the effects of the silica gel support type and the reduction temperature on the formation and dispersion of the metallic nickel phase in the reduced Ni/SiO2 precursors of the vegetable oil hydrogenation catalyst. The physicochemical characterization of the unreduced and reduced precursors has been accomplished appropriately by powder X-ray diffraction, infrared spectroscopy, temperature programmed reduction and H2-chemisorption techniques. It can be stated that the texture peculiarities of the silica gels used as supports influence on the crystalline state and distribution of the deposited Ni-containing phases during the preparation of the precursors, on the reduction temperature of the investigated solids as well as on the bulk size and surface dispersion of the arising metallic nickel particles. It was shown that two types of Ni2+-species are formed during the synthesis procedure, namely basic nickel carbonate-like and Ni-phyllosilicate with different extent of presence, location and strength of interaction. The different location of these species is supposed to result in various strength of Ni-O and Ni-O-Si interaction, thus determining the overall reducibility of the precursors. It was specified that the Ni2+-species are strongly bonded to the surface of the silica gel obtained at neutral pH value and weakly bonded to the surface of those prepared in acidic and alkaline conditions. It was established that the precursor, derivates from the silica gel obtained at alkaline conditions, demonstrates both significant reduction of the Ni2+ ions at 430°C and finely dispersed metallic nickel particles on its surface. High dispersion of the metallic nickel might be the crucial reason for achieving of high activity in the vegetable oil hydrogenation.  相似文献   

13.
Three inorganic–organic composite sandwich-type phosphotungstates [Ni(tepa)(H2O)]4H2[Ni4(H2O)2(α-B-PW9O34)2]·8H2O (1), (enH2)3[Ni2(H2O)10][Ni4(H2O)2(α-B-PW9O34)2]·en·8H2O (2) and (enH2)10[Mn4(H2O)2(α-B-PW9O34)2]2·20H2O (3) (tepa=tetraethylenepentamine and en=ethylenediamine) have been synthesized by the hydrothermal reaction of the trivacant Keggin polyoxoanion [α-A-PW9O34]9− with Ni2+ or Mn2+ ions in the presence of tepa or en and structurally characterized by IR spectra, elemental analysis, thermogravimetric analysis and variable temperature magnetic susceptibility. X-ray crystallographic analyses indicate that they all contain the classical tetra-M sandwiched polyoxoanions [M4(H2O)2(α-B-PW9O34)2]10− (M=Ni2+ or Mn2+) and nickel-organoamine cations or organoamine cations work as the charge balance ions. The tetra-M clusters in 1, 2 and 3 exhibit the familiar structural type of a β-junction at the sites of metal incorporation. The study of magnetic property of 1 is indicative of a typical ferromagnetic coupling between Ni2+ cations.  相似文献   

14.
Studies on Nickel Oxide Mixed Catalysts. I. Structural Properties of NiO/SiO2 Catalysts Structural properties of NiO/SiO2 catalysts prepared by precipitation-deposition and impregnation have been investigated. A nickel-layer-silicate like structure is formed only in the precipitated catalysts showed infrared spectroscopic measurements. After thermal treatment at 723 K by means of magnetic measurements and reflectance spectroscopy besides Ni2+ ions in octahedral environment tetrahedral coordinated Ni2+ions were found. The part of tetrahedral coordinated Ni2+ ions is independent of the NiO-content up to 40 mole % NiO. Nickel oxide is formed above a content of 40 mole %. In the case of the impregnated catalysts nickel oxide cluster are formed on the surface after annealing at 723 K.  相似文献   

15.
The formation of nickel citrate complexes was studied at ionic strength values of 0.1 and 0.3 mol/l (Et4NCl) and 298.15 K by potentiometric titration. The NiCit?, NiHCit, and NiH2Cit+ complexes were formed in a Ni2+ ion-citric acid (H3Cit) system. The thermodynamic formation constants of the nickel(II) citrate complexes were calculated in an aqueous solution at \(I = 0:\log \beta _{NiCit^ - }^0 \) = 6.86 ± 0.12 (Ni2+ + Cit3? ai NiCit?), logK 1 0 = 4.18 ± 0.10 (Ni2+ + HCit2? ai NiHCit), and logK 2 0 = 2.24 ± 0.11 (Ni2+ + H2Cit? ai NiH2Cit+). The spectral properties of the Ni2+-H3Cit system were studied by spectrometry. The conditions of calorimetric determination of the thermal effects of formation of the nickel citrate complexes in an aqueous solution were optimized on the basis of the calculated stability constants of the Ni(II) complexes with H3Cit.  相似文献   

16.
A series of heterometallic 3d–Gd3+ complexes based on a lanthanide metalloligand, [M(H2O)6][Gd(oda)3] ? 3 H2O [M=Cr3+ ( 1‐Cr )] (H2oda=2,2′‐oxydiacetic acid), [M(H2O)6][MGd(oda)3]2 ? 3 H2O [M=Mn2+ ( 2‐Mn ), Fe2+ ( 2‐Fe ) and Co2+ ( 2‐Co )], and [M3Gd2(oda)6(H2O)6] ? 12 H2O [M=Ni2+ ( 3‐Ni ), Cu2+ ( 3‐Cu ), and Zn2+ ( 3‐Zn )], are reported. Magnetic and heat‐capacity studies revealed a significant impact on the magnetocaloric effect depending on the anisotropy of the 3d transition metal ions, as confirmed by comparison of the observed maximum values of ?ΔSm between complexes 2‐Co and 1‐Cr . In these two complexes, the 3d metal ions have the same spin (S=3/2 for Co2+ and Cr3+ ions), and the theoretical calculation suggested a larger ?ΔSm value for 2‐Co (47.8 J K?1 kg?1) than 1‐Cr (37.5 J K?1 kg?1); however, the significant anisotropy of Co2+ ions in 2‐Co , which can result in smaller effective spins, gives a smaller value of ?ΔSm for 2‐Co (32.2 J K?1 kg?1) than for 1‐Cr (35.4 J K?1 kg?1) at ΔH=9 T.  相似文献   

17.
The kinetics of stripping of Ni2+ from a Ni‐BTMPPA complex, dissolved in a kerosene solution of BTMPPA (H2A2, Cyanex 272), by acidic sulfate‐acetato solution, was studied using the single (falling) drop technique and flux (F) method of data treatment. The empirical flux equation at 303 K is Fb (kmol/m2s) = 10?4.35 [Ni2+] (1+10?3.42 [H+]?1)?1 ([H2A2](o)0.5+2.50 [H2A2](o))?1 (1+6[SO42?]) (1+3.20 [Ac?]). Activation energy (Ea), entropy change in activation (ΔS±), and enthalpy change in activation (ΔH±) were measured under different experimental conditions. Based on the empirical flux equation, Ea and ΔS±, the mechanism of Ni2+ stripping is provided. In a low [H+] region, the stripping reaction steps appear as [NiA+] → Ni2+ + A? and [Ni(HA2)2](int) → [NiHA2]+(int) + HA2(int)? in lower and higher concentration regions of free BTMPPA, respectively, provided [SO42?] and [Ac?] are kept low. However, at higher [H+] concentrations, the stripping is under diffusion control. With increasing [SO42?] and [Ac?], the enhancement of the rate is attributed to the attack of the Ni(II) complex by SO42? or HSO4? and Ac? to form NiSO4 or NiHSO4+ and NiAc+ complexes. Negative ΔS± values indicate that the rate‐determining stripping reaction steps occur via an substitution nucleophilic, bimolecular (SN2) mechanism.  相似文献   

18.
The gas phase reactions of metal ions (Al+, Cu+) with amine molecules [CH3NH2=MA, (CH3)2NH=DMA] were investigated using a laser ablation‐molecular beam method. The directly associated product complex ions,Al+‐MA and Al+‐DMA, and the dehydrogenation product ions, Cu+(CH2NH) and Cu+(C2H5N), as well as hydrated ion Cu+(NC2H5·H2O), have been obtained and recorded from the reactions of the metal ions and organic amine molecules, and density functional theory (B3LYP) calculations have been performed to reveal the optimized geometry, energetics, and reaction mechanism of the title reactions with basis set 6‐311+G(d,p) adopted.  相似文献   

19.
Two nickel complexes supported by tridentate NS2 ligands, [Ni2(κN,S,S,S′‐NPh{CH2(MeC6H2R′)S}2)2] ( 1 ; R′=3,5‐(CF3)2C6H3) and [Ni2(κN,S,S,S′‐NiBu{CH2C6H4S}2)2] ( 2 ), were prepared as bioinspired models of the active site of [NiFe] hydrogenases. The solid‐state structure of 1 reveals that the [Ni2(μ‐ArS)2] core is bent, with the planes of the nickel centers at a hinge angle of 81.3(5)°, whereas 2 shows a coplanar arrangement between both nickel(II) ions in the dimeric structure. Complex 1 electrocatalyzes proton reduction from CF3COOH at ?1.93 (overpotential of 1.04 V, with icat/ip≈21.8) and ?1.47 V (overpotential of 580 mV, with icat/ip≈5.9) versus the ferrocene/ferrocenium redox couple. The electrochemical behavior of 1 relative to that of 2 may be related to the bent [Ni2(μ‐ArS)2] core, which allows proximity of the two Ni???Ni centers at 2.730(8) Å; thus possibly favoring H+ reduction. In contrast, the planar [Ni2(μ‐ArS)2] core of 2 results in a Ni???Ni distance of 3.364(4) Å and is unstable in the presence of acid.  相似文献   

20.
Charaterization and Catalytic Activity of Ni2+ Exchanged X and Y Zeolites. I. TPR Studies on NiNaX and NiNaY Zeolites . The structure of TPR spectra of NiNaX and NaNiY zeolites variously exchanged is determined by the location of the cations. In case of X zeolites a peak appears with a maximum at 750–800 K (reduction on SII and SI, positions) and for higher exchange degrees an additional one at about 1000 K (reduction on SI positions). Three ranges of reduction may be separated in case of Y zeolites (reduction on SII, SI′, and SI). With increasing Si/Al ratios the maximum of the hightemperature peak is shifted to higher temperatures. The reduction at temperatures up to 800 K resulted in higher reduction degrees for X reolites while the overall reduction up to high temperatures led to higher reduction degrees for Y zeolites. The kinetic analysis by means of two different methods yielded the following activation energies: (85 ± 10) or (86 ± 2) kJ/mole, respectively, for the low-temperature peak, and (223 ± 12) or (214 ± 2) kJ/mole, respectively, for the high-temperature peak.  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