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.
Reaction of the dimeric rhodium complex [{(η5-C5Me5)Rh(μ-Cl)Cl}2] with an excess of 1-(4-cyanophenyl)-imidazole in dichloromethane afforded neutral mononuclear complex [(η5-C5Me5)RhCl2(CPI)] (CPI = 1-(4-cyanophenyl)-imidazole) 1. The complex 1 reacted with EPh3 (E = P, As, Sb) and N-N donor bases 2,2′-bipyridine and 1,10-phenanthroline in different solvents to give substitution products wherein, nature of the product was governed by polarity of the solvents employed in the reaction. Resulting complexes have been characterized by elemental analyses, spectral (FAB-MS, IR, 1H,13C, 31P NMR, UV-Vis, Emission) and electrochemical studies. Coordination of CPI through imidazole nitrogen and the presence of pendant nitrile group have been supported by spectral studies.  相似文献   

2.
The kinetic and thermodynamic characteristics of solid-phase ligand substitution reactions were determined for chromium(III) and cobalt(III) amine complexes with B10H 10 2– and C2B9H 12 in the outer sphere. The kinetic equation of the topochemical process was found to be of the formf()=(1–)2/3, corresponding to a reaction proceeeding on the interface between the phases (shrinking sphere). Two types of solid-phase ligand substitution reactions were found: endothermic and exothermic reactions taking place either through thermal activation of the metal-ligand bond (in substitution reactions by the anion B10H 10 2– ), or through the acid-base proton-exchange reaction between the anion C2B9H 12 entering the coordination sphere and the amine leaving it; in this case the process proceeds without mass loss. It could be demonstrated that the reactions occurring in crystalline complex salts cannot proceed by purely dissociative or associative mechanisms; depending on the structure of the crystal lattice, mutually adapted dissociative or associative mechanisms are feasible. Reactions proceeding by the first mechanism haveE a=300–500 kJ/mol and logA=30–50; the values for reactions proceeding by the second mechanism areE a= =180–250 kJ/mol and logA=15–25.
Zusammenfassung Die kinetischen und thermodynamischen Kennwerte der Ligandensubstitutionsreaktionen in fester Phase bei Chrom(III)- und Kobalt(III)-Aminkomplexen mit B10H 10 2– und C2B9H12/– als Anionen in der externen Sphäre wurden untersucht. Die kinetische Gleichung des topochemischen Prozesses hat die Formf()=(1–)2/3, was einer an der Grenzfläche zwischen zwei Phasen verlaufenden Reaktion entspricht. Zwei Typen von Ligandsubstitutionsreaktionen in fester Phase wurden gefunden: endotherme und exotherme Reaktionen verlaufen entweder infolge thermischer Aktivierung der Metall-Ligand-Bindung (in Substitutionsreaktionen des Anions B10H 10 2– ) oder durch Protonenaustausch zwischen dem in die Koordinationssphäre eintretende Anion C2B9H 12 und dem daraus austretenden Amin; in diesem Falle verläuft der Prozeß ohne Massenverlust. Es konnte gezeigt werden, daß die in kristallinen Komplexsalzen vor sich gehenden Reaktionen nicht nach einem rein dissoziativen oder assoziativen Mechanismus verlaufen können; abhängig von der Struktur des Kristallgitters ist ein wechselseitig angepasster dissoziativer oder assoziativer Mechanismus wahrscheinlich. Die nach dem ersten Mechanismus verlaufenden Reaktionen weisen fürE a Werte von 300 500 kJ/mol und für logA von 30–50 auf, während bei nach dem zweiten Mechanismus verlaufenden Reaktionen die entsprechenden Werte zwischen 180 und 250 kJ/mol bzw. 15 und 25 liegen.

() () B10H 10 2- C2B9 H 12/- . , f()=(1-)2/3, ( ). : - , - ( B10H 10 2- ), - 29 12 - ; ., . . , , a =300–500 / logA=30–50, –ina=180–250 / log=15–25.


The authors are grateful to F. G. Szabó (ETALON Factory, Baku) for help in the application of the derivatograph for quantitative measurements and for development of a method to record thermoanalytical curves for measuring enthalpies of reactions.  相似文献   

3.
The thermal properties and gas Chromatographie behaviour of manganese(II) and manganese(III) trifluoroacetylacetonates (TFA) were investigated by using the ligand vapour technique. The two chelates, Mn(TFA); and Mn(TFA)3, can be quantitatively eluted on a mixed-liquid phase (1.9% OV-17 ÷ 0.1% PEG-20M) at column temperatures above 210°C and 130–150°C, respectively; Mn(TFA)3 is completely converted to Mn(TFA)2 by thermal dissociation at column temperatures above 180°C and completely eluted as Mn(TFA)2 above 210°C. The chelates can be determined separately within errors of about 1% after a preliminary extraction.  相似文献   

4.
The source of protons associated with the ligand loss channel of HX((n - 1)+) from [Pt(II)(dien)X](n+) (X = Cl, Br and I for n = 1 and X = NC(5)H(5) for n = 2) in the gas phase was investigated by deuterium-labelling studies. The results of these studies indicate that these protons originate from both the amino groups and the carbon backbone of the dien ligand. In some instances (e.g. X = Br and I), the protons lost from the carbon backbone can be even more abundant than the protons lost from the amino groups. The gas-phase substitution reactions of coordinatively saturated [Pt(II)(L(3))L(a)](2+) complexes (L(3) = tpy or dien) were also examined using ion-molecule reactions. The outcome of the ion-molecule reactions depends on both the ancillary ligand (L(3)) as well as the leaving group (L(a)). [Pt(II)(tpy)L(a)](2+) complexes undergo substitution reactions, with a faster rate when L(a) is a good leaving group, while the [Pt(II)(dien)L(a)](2+) complex undergoes a proton transfer reaction.  相似文献   

5.
The effect of several imidazolium-based ionic liquids on the mechanism of a classical ligand substitution reaction of [Pt(terpyridine)Cl] (+) with thiourea was investigated. A detailed kinetic study as a function of the nucleophile concentration and temperature was undertaken under pseudo-first-order conditions using stopped-flow techniques. Polarity measurements were performed for the employed ionic liquids on the basis of solvatochromic effects, and they show similarities with conventional polar solvents. Density-functional theory calculations (RB3LYP/LANL2DZp) were employed to predict the ion-pair stabilization energy between the ionic components of the ionic liquids and/or between the anions of the ionic liquids and the cationic Pt (II) complex. These data illustrate how the anions of the ionic liquids can affect the investigated substitution reaction. In general, the substitution mechanism in ionic liquids was found to have an associative character similar to that in conventional solvents. The observed deviations reflect the influence of the ionic liquid on the interaction between the anionic component of the liquid and the positively charged complex.  相似文献   

6.
Rate constants for ligand substitution by bromide, thiocyanate and azide ions at iron(III) ion in dimethyl sulfoxide as solvent, determined by stopped-flow spectrophotometry, are similar and are consistent with a dissociative mechanism. In addition, azide ion gives a second much slower, reaction, attributed to formation of a binuclear complex. Results of similar measurements with thiocyanate ion in acetonitrile were more complicated, attributed to a marked influence of residual water on the reactivity of iron (III) ion.  相似文献   

7.
Summary Substitution reactions ofcis-Mo(CO)4(py)2, Mo(CO)5-(4-Mepy) and Mo(CO)5(py) with biacetylbis(phenylimine), bipyridine and 1,10-phenanthroline in toluene and 1,2-dichloroethane were studied as a function of temperature and pressure. The volumes of activation are +4 cm3 mol–1 and almost zero for the tetra- and penta-carbonyl systems, respectively, and independent of entering ligand and solvent. Furthermore, these values closely parallel the trend in S#. The results are discussed in terms of earlier suggested mechanisms and relevant data recently reported in the literature.  相似文献   

8.
Treatment of either RuHCl(CO)(PPh3)3 or MPhCl(CO)(PPh3)2 with HSiMeCl2 produces the five-coordinate dichloro(methyl)silyl complexes, M(SiMeCl2)Cl(CO)(PPh3)2 (1a, M = Ru; 1b, M = Os). 1a and 1b react readily with hydroxide ions and with ethanol to give M(SiMe[OH]2)Cl(CO)(PPh3)2 (2a, M = Ru; 2b, M = Os) and M(SiMe[OEt]2)Cl(CO)(PPh3)2 (3a, M = Ru; 3b, M = Os), respectively. 3b adds CO to form the six-coordinate complex, Os(SiMe[OEt]2)Cl(CO)2(PPh3)2 (4b) and crystal structure determinations of 3b and 4b reveal very different Os-Si distances in the five-coordinate complex (2.3196(11) Å) and in the six-coordinate complex (2.4901(8) Å). Reaction between 1a and 1b and 8-aminoquinoline results in displacement of a triphenylphosphine ligand and formation of the six-coordinate chelate complexes M(SiMeCl2)Cl(CO)(PPh3)(κ2(N,N)-NC9H6NH2-8) (5a, M = Ru; 5b, M = Os), respectively. Crystal structure determination of 5a reveals that the amino function of the chelating 8-aminoquinoline ligand is located adjacent to the reactive Si-Cl bonds of the dichloro(methyl)silyl ligand but no reaction between these functions is observed. However, 5a and 5b react readily with ethanol to give ultimately M(SiMe[OEt]2)Cl(CO)(PPh3)(κ2(N,N-NC9H6NH2-8) (6a, M = Ru; 6b, M = Os). In the case of ruthenium only, the intermediate ethanolysis product Ru(SiMeCl[OEt])Cl(CO)(PPh3)(κ2(N,N-NC9H6NH2-8) (6c) was also isolated. The crystal structure of 6c was determined. Reaction between 1b and excess 2-aminopyridine results in condensation between the Si-Cl bonds and the N-H bonds with formation of a novel tridentate “NSiN” ligand in the complex Os(κ3(Si,N,N)-SiMe[NH(2-C5H4N)]2)Cl(CO)(PPh3) (7b). Crystal structure determination of 7b shows that the “NSiN” ligand coordinates to osmium with a “facial” arrangement and with chloride trans to the silyl ligand.  相似文献   

9.
The kinetics and mechanism of successive substitution of chelate ligands in Cu(II) bissalicyladehydate have been studied by ESR method.
- (II).
  相似文献   

10.
11.
A mixed-valence Mn(III)-Mn(II)-Mn(III) trinuclear complex of stoichiometry MnIIIMnIIMnIII(Hsaladhp)2(Sal)4.2CH3CN (1), where H3saladhp is a tridentate Schiff-base ligand, has been structurally characterized with X-ray crystallography. The Mn(III)Mn(II)Mn(III) angles are strictly 180 degrees as required by crystallographic inversion symmetry. The complex is valence-trapped with two terminal Mn(III) ions in a distorted square pyramidal geometry. The Mn(III)...Mn(II) separation is 3.495 A. The trinuclear complex shows small antiferromagnetic exchange J coupling. The magnetic parameters obtained from the fitting procedure in the temperature range 10-300 K are J1 = -5.7 cm-1, g = 2.02, zJ = -0.19 cm-1, and R = 0.004. The EPR spectrum was obtained at 4 K in CHCl3 and in tetrahydrofuran glasses. The low-field EPR signal is a superposition of two signals, one centered around g = 3.6 and the other, for which hyperfine structure is observed, centered around g = 4.1 indicating an S = 3/2 state. In addition, there is a 19-line signal at g = 2.0. The multiline signal compares well with that observed for the S2 or S0* states of the oxygen-evolving complex. 1H NMR data reveal that the trinuclear compound keeps its integrity into the CHCl3 solution. Crystal data for complex 1: [C54H52N4O18Mn3], M = 1209.82, triclinic, space group P1, a = 10.367(6) A, b = 11.369(6) A, c = 13.967(8) A; alpha = 112.56(1) degree, beta = 93.42(2) degrees, gamma = 115.43(1) degree, Z = 1.  相似文献   

12.
13.
Hemophore HasA is a 19 kDa iron(III) hemoprotein that participates in the shuttling of heme to a specific membrane receptor. In HasA, heme iron has an original coordination environment with a His/Tyr pair as axial ligands. Recently developed two-dimensional protonless (13)C-detected experiments provide the sequence-specific assignment of all but three protein residues in the close proximity of the paramagnetic center, thus overcoming limitations due to the short relaxation times induced by the presence of the iron(III) center. Mono-dimensional (13)C and (15)N experiments tailored for the detection of paramagnetic signals allow the identification of resonances of the axial ligands. These experiments are used to characterize the conformational features and the electronic structure of the heme iron(III) environment. The good complementarity among (1)H-, (13)C-, and (15)N-detected experiments is highlighted. A thermal high-spin/low-spin equilibrium is observed and is related to a modulation of the strength of the coordination bond between the iron and the Tyr74 axial ligand. The key role of a neighboring residue, His82, for the stability of the axial coordination and its involvement in the heme delivery to the receptor is discussed.  相似文献   

14.
15.
The role of the solvent in reactions involving ions is considered in relation to the structure of liquids. The rate constants and activation parameters for ligand substitutions at divalent transition metal cations in various solvents are compared with those for solvent exchanges. The differences are related to structural properties of the solvents, represented by their heats of evaporation and fluidities, and interpreted with the aid of a model developed from that of Frank and Wen. Water is not a typical solvent.This paper was presented at the symposium, The Physical Chemistry of Aqueous Systems, held at the University of Pittsburgh, Pittsburgh, Pennsylvania, June 12–14, 1972, in honor of the 70th birthday of Professor H. S. Frank.  相似文献   

16.
The average magnetic susceptibility (1.2-100 K) and magnetisation (100–15000 Oe at 4.2 K) of two perchlorato manganese(III) porphyrins establish them to be high-spin, in contrast to the “anomalous” behaviour of analogous iron(III) porphyrins. An explanation of the origin of the zero-field splitting in high-spin manganese(III) porphyrins is presented.  相似文献   

17.
The variation of ligand para substituents on pyridyl donor groups of tridentate amine copper(I) complexes was carried out in order to probe electronic effects on the equilibrium between mu-eta2:eta2-(side-on)-peroxo [Cu(II)2(O2(2-))]2+ and bis(mu-oxo) [Cu(III)2(O(2-))2] species formed upon reaction with O2. [Cu(I)(R-PYAN)(MeCN)n]B(C6F5)4 (R-PYAN = N-[2-(4-R-pyridin-2-yl)-ethyl]-N,N',N'-trimethyl-propane-1,3-diamine, R = NMe2, OMe, H, and Cl) (1R) vary over a narrow range in their Cu(II)/Cu(I) redox potentials (E(1/2) vs Fe(cp)2(+/0) = -0.40 V for 1(NMe2), -0.38 V for 1(OMe), -0.33 V for 1H, and -0.32 V for 1Cl) and in C-O stretching frequencies of their carbonyl adducts, 1R-CO: nu(C-O) = 2080, 2086, 2088, and 2090 cm(-1) for R = NMe2, OMe, H, and Cl, respectively. However, within this range of electronic properties for 1R, dioxygen reactivity is significantly affected. The reaction of 1Cl or 1H with O2 at -78 degrees C in CH2Cl2 gives UV-vis and resonance Raman spectra indicative of a mu-eta2:eta2-(side-on)-peroxo dicopper(II) adduct (2R). Compound 1(OMe) reacts with O2, yielding equilibrium mixtures of side-on peroxo (2(OMe)) and bis(mu-oxo) (3(OMe)) species. Oxygenation of 1(NMe2) leads to the sole generation of the bis(mu-oxo) dicopper(III) complex (3(NMe2)). A solvent effect was also observed; in acetone or THF, increased ratios of bis(mu-oxo) relative to side-on peroxo complex are observed. Thus, the equilibrium between a dicopper side-on peroxo and bis(mu-oxo) species can be tuned by ligand design-specifically, more electron donating ligands favor the formation of the latter isomer, and the peroxo/bis(mu-oxo) equilibrium can be shifted from one extreme to the other within the same ligand system. Observations concerning the reactivity of the dioxygen adducts 2H and 3(NMe2) toward external substrates are also presented.  相似文献   

18.
A new binucleating ligand incorporating four oxime groups, butane-2,3-dione O-[4-aminooxy-2,3-bis-(2-hydroxyimino-1-methyl-propylideneaminooxymethyl)-but-2-enyl]-dioxime, (H4mto), has been synthesized and its dinuclear cobalt(III), copper(II), and homo- and hetero-tetranuclear copper(II)–manganese(II) complexes have been prepared and characterized by 1H- and 13C-n.m.r., i.r., magnetic moments and mass spectral studies. Elemental analyses, stoichiometric and spectroscopic data indicate that the metal ions in the complexes are coordinated to the oxime nitrogen atoms (C=N) and the data support the proposed structure for H4mto and its complexes. Moreover, dinuclear cobalt(III) and copper(II) complexes of H4mto have a 2:1 metal:ligand ratio.  相似文献   

19.
Manganese silicalite-2 was synthesized at high pH using the molecular cluster Mn 12O 12(O 2CCH 3) 16 as a Mn source. The silicalite-2 (ZSM-11) materials were synthesized using 3,5-dimethyl- N, N-diethylpiperdinium hydroxide as a structure-directing agent to produce phase-pure ZSM-11 materials. No precipitation of manganese hydroxide was observed, and synthesis resulted in the incorporation of up to 2.5 mol % Mn into the silicalite-2 with direct substitution into the framework verified by the linear relationship between the unit cell volume and loading. The Mn is reduced to Mn (II) during hydrothermal synthesis and incorporated into the silicalite-2 framework during calcination at 500 degrees C. Further calcination at 750 degrees C does not affect the crystallinity but oxidizes essentially all of the Mn (II) to Mn (III) in the framework. The large difference in oxidation temperatures between the II and III oxidation states provides a means of producing relatively pure manganese(II) and manganese(III) silicalite-2 materials for applications such as catalysis.  相似文献   

20.
The kinetics of piperidine replacement by pyridine at the pentacyano(piperidine)-ferrate(II) complex ion was studied under pseudo-first order conditions in binary aqueous mixtures of methanol, t-butanol, p-dioxane, and glycerol, from a mole fraction of co-solvent from 0 to about 0.15. The observed variations can be explained considering the degree of hydrophobic interaction between released ligand and water molecules which changes according to the structure-forming or struucture-breaking effect of added co-solvent on water.  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