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 125 毫秒
1.
The kinetics of formation and dissociation of [V(H2O)5NCS]2+ have been studied, as a function of excess metal-ion concentration, temperature, and pressure, by the stopped-flow technique. The thermodynamic stability of the complex was also determined spectrophotometrically. The kinetic and equilibrium data were submitted to a combined analysis. The rate constants and activation parameters for the formation (f) and dissociation (r) of the complex are: k/M ?1 · S?1 = 126.4, k/s?1 = 0.82; ΔH /kJ · mol?1 = 49.1, ΔH/kJ · mol?1 = 60.6; ΔS/ J·K?1·mol?1= ?39.8, ΔSJ·K?1·mol?1 = ?43.4; ΔV/cm3·mol?1 = ?9.4, and ΔV/cm3 · mol?1 =?17.9. The equilibrium constant for the formation of the monoisothiocynato complex is K298/M ?1 = 152.9, and the enthalpy and entropy of reaction are ΔH0/kJ · mol?1 = ? 11.4 and ΔS0/J. K?1mol?1 = +3.6. The reaction volume is ΔV0/cm3· mol?1 = +8.5. The activation parameters for the complex-formation step are similar to those for the water exchange on [V(H2O)6]3+ obtained previously by NMR techniques. The activation volumes for the two processes are consistent with an associative interchange, Ia, mechanism.  相似文献   

2.
The title cation ( = Ni2L) is formed in a variety of reactions (Schemes 1 and 2) in systems containing Ni2+ and (2-thiolatoethyl)-diphenylphosphine (= L?) in the absence of coordinating anions at Ni2+/L? ratios > 0.5 in apolar or moderately polar media. Solid [Ni2L3]CIO4 and [Ni2L3]BPh4 have been isolated. Job's plots confirm the Ni2L- stoichiometry in solution. 31P-NMR data are consistent with ≥ 97% Ni2L (vs. ? 3% of hypothetical Ni3L) at equilibrium and support the suggested configuration (Fig. 2). The equilibrium between NiL2 + NiL2Br2 and Ni2L + Br? varies with the solvent composition in CH23Cl2/EtOH mixtures. The rate of formation of Ni2L2Br2 from Ni2L and bromide (in high excess) in CH2Cl2 is first-order in [Ni2L]tot but depends on the ratio [Bu4NBr]tot/[Ni2L3 · ClO4]tot, even at a high excess of bromide. This is interpreted by efficient competition in ion-aggregate formation between the small perchlorate concentration introduced as the counterion of Ni2L, and the large excess of bromide.  相似文献   

3.
The cluster [Pt3(m?-CO)3(PCy3)3] can be protonated with HBF4 · OEt2 to form the cluster [Pt3(m?-CO)3(PCy3)3](m?3-H) +BF ( 2 ). This unstable compound was isolated and characterised by NMR and IR spectroscopy.  相似文献   

4.
The replacement of Cl? by ethylenediamine (en) in PdCl has been followed spectrophotometrically at 25°C and μ = 1 (NaClO4); it proceeds in two steps leading to Pd(en)Cl2 and Pd(en), respectively. The observed rate constants are discussed in terms of the mechanism proposed by Reinhardt [1] for the successive ammination reactions of PdCl.  相似文献   

5.
The kinetics of O2-uptake of five-coordinated Co2+/tren complexes (tren = 2,2′, 2″-tris(2-aminoethyl)amine) have been studied extensively. The kinetics of formation of (tren)Co(O2, OH)Co(tren)3+ exhibits two steps. The rate law of O2-addition, the first step, was of the form: rate = (k[H+] + kKa)/([H+] + Ka) [Co(tren)2+][O2]. Second-order rate constants k = 220 ± 19 M ?1s?1 and k = 1.8 ± .035 · 103M ?1s?1 agreed well from O2-uptake and (stopped-flow) spectrophotometric measurements. The protonation constant of the hydroxo complex obtained by equlibrium measurements (spectrophotometric and by pH-titration) in anaerobic conditions (pKa = 10.03) agreed well with that derived from kinetic data (p Ka = 9.93); k and k are about a factor 100 smaller than those for the pseudooctahedral Co(trien) (H2O). This and the fact that several other Co(II) complexes with five-coordinated geometry do not exhibit oxygen affinity led to the proposal that the oxygenation mechanism for Co2+/tren complexes involves fast preequilibria between Co(tren) (H2O)2+ and Co(tren) (H2O) and only the latter is assumed to be reactive. The enhanced rate at high pH is explained by rate determining H2O-exchange in the O2-addition step and the ability of coordinated OH? to labilize the neighbouring H2O. This mechanism is furthermore supported by the formation of one kinetically preferred isomer of the peroxo-bridged dicobalt(III) complex (O2 cis to the tertiary N-atom) and the large negative activation entropy (?30 eu). The second step is the intramolecular bridging reaction: is independent of [Co(tren)2+] and [O2] but exhibits a pH-dependence of the form k3 = k3[H + ]/(Ka + [H+]); k?3 ( = 5 · 10?5 s?1) was determined independently and from the two rate constants the equilibrium constant was calculated as ≈ 105. The ligand combination as in Co(tren)2+ was shown to provide an excellent balance to form a reversible oxygen carrier; possible reasons for this are discussed.  相似文献   

6.
The bidentate diphosphine ligand, 3,3′-oxybis[(dipenylphosphino)methylbenzene] ( 1 ) forms monomeric, trans-square-planar complexes MX2( 1 ) (M = Ni, Pd, Pt; X = Cl?, Br? I?, and, in part, N, NCS?, CN?, NO) as well as Pt(H)Cl( 1 ), Pt(H)Br( 1 ), and RhCl(CO)( 1 ). Polymeric species have been observed with substitutionally inert metal centres: trans-[PtCl2( 1 )]2 and cis-[PtCl2( 1 )]n (mean value of n ≈ 4–5) 31P-NMR, and selected IR and UV/VIS parameters are reported. Ligand 1 shows a marked preference for trans-spanning and monomeric chelate formation, despite its various degrees of freedom of internal rotation in the lignad backbone. The readily available ligand 1 as well as analogues with other donor atoms, therefore, appear useful in most potential applications of trans-spanning chelate ligands. The crystal structure of AgCl( 1 )·0.5 (CH3)2C?O·0.39 C6H12 (space group C2/c,a = 21.02 Å, b = 14.57 Å, c = 24.79 Å, β = 99.77°, V = 7531.4 Å3, Z = 8) confirms the presence of three-coordinate Ag( I ), with a coordination intermediate between a trigonal-planar and a T-shaped geometry (P-Ag-P = 145.61(8)°).  相似文献   

7.
Using a new mathematical treatment, the nature and stability constants of the simple and mixed complex-species of copper(II) with hydroxyde and ammonia as ligands have been determined. The solubility curves of CuO in heterogeneous equilibrium have been identified in function of pH only and in function of pH and pNH3tot at 25° and unit ionic strength (NaClO4). The predominent species in the relatively dilute system limited by the ionic strength are [Cu2+], [Cu(OH)2], [Cu(OH)], [Cu(OH)], [Cu(NH3)], [Cu(NH3)], [Cu(NH3)], [Cu(NH3) (OH)+], [Cu(NH3)3(OH)+] and [Cu(NH3)2(OH)2].  相似文献   

8.
In aqueous acetonitrile (AN), Cu (I) forms the complexes Cu(AN)L+ and CuL with a series of substituted imidazoles (L). Stability constants logK of Cu(AN)+ + L ? Cu(AN)L+ and logβ2 were near 5 and 12, resp., log units for all ligands. The rate of autoxidation is described by ?d[O2]/dt=[CuL]2[O2](ka/(1+kb[CuL]) + (kc[L]+kd)/([CuL] + ke[Cu])), implying competition between one- or two-electron reduction of O2. The value of kc decreases from 5500M ?2S ?1 for unsubstituted imidazole to about 40M ?2S ?1 for 2-methylimidazole or 1,2-dimethyl-imidazole and essentially zero for the corresponding 2-ethyl-derivatives. On the other hand, ka and kb are much less influenced by the nature of the ligands, all values being near 5 · 104M ?2S ?1 and 103M ?1, respectively, for the complexes with the last four bases. Thus rather subtle sterical changes may strongly influence the relative importance of different pathways in the reduction of dioxygen by cuprous complexes.  相似文献   

9.
In the context of our studies on ruthenium(II) complexes containing polyazaheterocyclic ligands, we have determined the rate constants of quenching by molecular oxygen (kq) of the metal-to-ligand charge-transfer-excited state of a series of homoleptic [RuL3] complexes (where L stands for 2,2′-bipyridine (bpy), 1,10-phenanthroline (phen), 2,2′-bipyrazine (bpz), 4,7-diphenyl-1,10-phenanthroline (dip), diphenyl-1,10-phenanthroline-4,7-disulfonate (dpds), and 1, 10-phenanthroline-5-octadecanamide (poda)) in H2O and in MeOH. These compounds are singlet-oxygen (O2(1Δg)) sensitizers, and quantum yields of singlet-oxygen production (ΨΔ) in both solvents are also reported. Values of kq and ΨΔ depend on the nature of the ligand L and on the solvent, ΨΔ values showing a large range of variation (0.2 to 1.0). In MeOH, the only pathway for quenching of the excited [RuL3] complexes by molecular oxygen is energy transfer: the fraction of quenched excited states yielding singlet oxygen (?) is unity for all compounds in the series investigated. Changing from MeOH to H2O has several remarkable effects: higher kq and lower ΨΔ values are observed; ? drops to ca. 0.5 except for [Ru(bpz)3]2+. In fact, [Ru(bpz)3]2+ is by far the weakest reductant in the series and behaves differently from the other complexes, with lowest kq and ΨΔ values and a ? equal to 1 in both solvents. Results are interpreted on the basis of the role played by charge-transfer interactions between the sensitizer excited state and molecular oxygen in the quenching mechanism. RuII Complexes based on the 4,7-diphenyl-1, 10-phenanthroline (dip) ligand are very efficient and stable singlet-oxygen sensitizers with ΨΔ values close to unity in air-saturated MeOH.  相似文献   

10.
Recrystallization of Cs3P7 from liquid NH3 yields the triammoniate Cs3P7·3 NH3, which loses the weakly bound NH3 of crystallization below 253 K. A low-temperature crystal-structure analysis shows that Cs3P7· NH3 consists of a framework of heptaphosphanortricyclane anions P and Cs+ cations with NH3 molecules completing the coordination of the cations. The framework is built from Cs3P7 layers connected by only few Cs…?P interactions, the interlayer gap being filled by a two-dimensional network of NH3. The Cs7P7 part of the structure completes a family of alkali-metal-polyphosphide substructures which range from [RbP7]2? or [CsP11]2? chains over [Cs2Pn]? layers (n = 7, 11) to now [Cs3P7] frameworks.  相似文献   

11.
Protonation and Cu(II) complexation equilibria of L -phenyhilaninamide, N2-methyl-L-phenylalaninamide, N2, N2-dimethyl-L-phenylalaninamide, L -valinamide, and L -prolinamide have been studied by potentiometry in aqueous solution. The formation constants of the species observed, CuL2+, CuL, CuLH, CuL2H and CuL2H?2, are discussed in relation to the structures of the ligands. Possible structures of bisamidato complexes are proposed on the ground of VIS and CD spectra. Since Cu(II) complexes of the present ligands (pH range 6–8) perform chiral resolution of dansyl- and unmodified amino acids in HPLC (reversed phase), it is relevant for the investigation of the resolution mechanism to know which are the species potentially involved in the recognition process.  相似文献   

12.
Polynuclear Cobalt Complexes. III. On meso-[(en)2Co(OH)2Co(en)2] (NO3)4 The structure of the title compound has been determined by X-ray analysis and refined to R = 0,060. The crystals are monoclinic, space group very probably C2/m, with cell dimensions a = 16,113(3), b = 10,946(2), c = 7,193(2) Å, β = 113,22(2)° and Z = 2. The configuration of the cation is ΔΛ IR. and UV./VIS. spectroscopic data are given.  相似文献   

13.
Synthesis, redox, photophysical, and photochemical properties of Ru(NN) complexes NN = 2-((2′-pyridyl)thiazole (pyth), 2-(2′-pyrazyl)thiazole (pzth), 2,2′-bithiazole (bth), 5-(2′-pyridyl)-1,2,4-thiadiazole (pytda), 2-(2′-pyridyl)imidazole (pyim), 1-methyl-2-(2′-pyridyl)imidazole (Mepyim), and 2-(2′-pyridyl)oxazole (pyox)) are described. Oxidation potentials for the Ru3+/2+ couples in MeCN varied from about 0.80 V to 1.60 V vs. NHE. Three reduction waves were observed in all the cases except for Ru(pyim) and Ru(Mepyim) complexes and asigned to the one-electron reduction of each bidentate ligand. Absorption spectra contained bands in the UV (280–325 nm) and VIS (437–481 nm) regions which have been assigned to ligand-centered π-π* and metal-to-ligand charge-transfer dπ-π* transitions, respectively. Emission spectra at 77 K were determined for all the complexes presenting maxima in the 580–650-nm region, with vibrational progression in some of them. Only pyth, pzth, bth, and pytda tris-chelates showed luminescence at room temperature in aqueous solution, with quantum yields ranging from 0.0013 to 0.0095 and excited-state lifetimes from 55 to 390 ns, as determined from pulsed laser techniques. Their E0–0 spetroscopic energies have been estimated from emission wavelength maxima at 77 K which, in turn, have allowed calculation of excited-state redox potentials. A plot of E0–0 vs. ΔE1/2, where ΔE1/2 = E1/2(3+/2+) ? E1/2(2+/+), was linear with a slope of ca. 1.1 and a correlation coefficient of 0.999, demonstrating an identical nature of the orbital involved in spectroscopic and electrochemical processes. Photochemical properties of Ru(NN) complexes have been tested using methyl viologen (MV2+) in Ar-purged aqueous solution at pH 5. Stern-Volmer treatment has led to the determination of bimolecular quenching constants (0.5 to 2 × 109m?1·s?1) which parallel electron-transfer free-energy changes. Homogeneous back-reaction of primarily produced MV and Ru(NN) has been measured resulting to be slightly higher than diffusion control and independent of ligand nature. Rate constants for the scavenging of Ru(NN) by added edta have been also determined (1.7 to 8.2 × 108M?1 · S?1). Under such conditions, net production of MV is attained with quantum yields varying from 0.003 to 0.038 (single-shot laser results).  相似文献   

14.
The temperature dependence of the emission lifetime of the series of complexes Ru(bpy)n(4,4′-dpb) (bpy = 2,2′bipyridine, 4,4′-dpb = 4,4′-diphenyl-2,2′-bipyridine) has been studied in propionitrile/butyronitrile (4:5 v/v) solutions in the range 90–293 K. The obtained photophysical parameters show that the energy separation between the metal-to-ligand charge tranfer (3MLCT) emitting level and the photoreactive metal-centered (3MC) level changes across the series (ΔE = 3960, 4100, 4300, and 4700 cm?1 for Ru(bpy)), Ru(bpy)2(4,4′-dpb)2+, Ru(bpy)(4,4′-dpb), and Ru(4,4′-dpb), respectively, where ΔE is the energy separation between the minimum of the 3MLCT potential curve and 3MLCT – 3MC crossing point. Comparison between spectral and electrochemical data indicated that the changes in ΔE are due to stabilization of the MLCT levels in complexes containing 4,4′-dpb with respect to Ru(bpy)2+3. The photochemical data for the same complexes (as I? salts) have been obtained in CH2Cl2 in the presence of 0.01M Cl? upon irradiation at 462 nm. The complexes containing 4,4′-dpb are more photostable than Ru(bpy). Comparison between the data for thermal population of the 3MC photoreactive state and those for photochemistry indicated that the overall photochemical process is governed by (i) a thermal redistribution between the emitting and photoreactive excited states, and (ii) mechanistic factors, likely related to the size of the detaching ligand.  相似文献   

15.
Crystal Structure and Data from Vibrational Spectra of cis-Na2[Pd(SO3)2en] · 4 H2O The compound cis-Na2[Pd(SO3)2en] · 4 H2O (en = 1,2-diaminoethane) crystallizes in the orthorhombic space group Pnma with a = 623.7(2), = 1070.9(10), c = 1989.5(30) pm and Z = 4. In the [Pd(SO3)2en]2? anions the trans-influence of the sulfite ligands manifests itself in long Pd? N bonds with short Pd? S distances. A set of Na+ ions is present in face-sharing octahedra Na(OH2)6+, forming rods [Na(OH2)6/2]+ parallel to [100]. A second set of Na+ ions is surrounded by two H2O molecules and four O atoms from SO3 ligands of two anions to form likewise octahedra with face-sharing, yielding rods [Na(OH2)2/2{(OSO2)2Pd en}2/2]? parallel to [100]. Comparatively low v(Pd? N)-frequencies reveal the trans-influence of the sulfite ligands also in the vibrational spectra.  相似文献   

16.
The copper-catalyzed oxidation of ascorbic acid (AscH2) has been studied with a Clark electrode in aqueous MeCN. CuI or CuII may be equally used as the source of metal ion, without influence on the rate law. At sufficiently high [MeCN], the rate of the overall reaction is essentially given by the rate of CuI autoxidation: the reaction is of first order with respect to [Cu] and [O2] and shows an inverse-square dependence on [MeCN] as observed for the autoxidation of Cu. The pH dependence is complicated by the combination of the intrinsic pH effect on autoxidation with an additional term in the rate law which is directly proportional to [AscH?]. The latter term is explained by direct oxidation of the organic substrate by the primary dioxygen adduct of CuI, CuO. For [MeCN] < 0.7M , a gradual and pH-dependent transformation of this rate law and deviation from the first-order dependence on [O2] is indicated.  相似文献   

17.
The title compound and its potassium analog have been prepared from corresponding aqueous solutions of 99TcO at pH ≈? 2 with SO2 as a reducing agent. An X-ray structure determination of the Na-salt showed Tc coordinated to the tetradentate N(CH2COO) ligand (NTA). Two Tc-NTA moieties are joined via two bridging O-atoms into a four-membered Tc2O2 ring. The observed diamagnetism, a strong absorption band at 19 950 cm?1, and a short Tc-Tc distance of 2.363 Å are typical for the Tc2O2-fragment with its strong metal-metal interaction. The structural trans-influence at Tc and the network of H-bonds are consistent with Tc in oxidation state IV.  相似文献   

18.
Crystal structures and electrical properties of radical-cation salts of the chiral organic donor TMET (S,S,S,S,-bis-(dimethylethylenedithio)tetrathiafulvalene) are described. Two structural types, 2:1 with octahedral anions Pf, AsF, SbF, I (incommensurate), and 3:2 with tetrahedral anions BF?4, CIO?4, ReO?4 are observed. Resistivity measurements between 2 and 298 K indicate that the 3:2 types are organic metals, while the other compounds are semiconductors. (TMET)3(CIO4)2 is metallic down to about 120 K at ambient pressure and remains metallic down to 2 K at 8 kbar.  相似文献   

19.
Recent work on the spontaneous (= acid-independent) cleavage of the mono-ol cation, i.e. in Cl?/ClO and NO/ClO mixed-electrolyte media has established (by analysis of anion-competition experiments) the existence of reactive ion pairs of the mono-ol cation with Cl? and NO. Their existence must be allowed for in the analysis of the rate data for the acid-induced cleavage (pH 0–1) of the mono-ol cation in these mixed-electrolyte media. Thus, previous data for acidic Cl?/ClO media have been re-interpreted in this work, and new data for NO/ClO media have been analyzed in the same sense. This analysis removes an apparent discrepancy in the orders of magnitude of ion aggregate stability constants between the mono-ol and similar binuclear cations.  相似文献   

20.
The stability constants of the Ni2+ and Co2+ complexes with 1,5-diazacyclooctane-N,N′-diacetic acid (H2DACODA) have been determined potentiometrically in 0.5M KNO3 at 25°. Only M(DACODA) and M(DACODA)OH? were observed. In addition the formation and dissociation kinetics of the pentacoordinate complexes M(DACODA) has been studied in aqueous solution using a stopped-flow technique. Formation follows the rate law vf = kf [M2+] [HDACODA?]/[H+], which can be interpreted as a bimolecular process either between M2+ and DACODA2? (k) or between MOH+ and HDACODA? (k). The second order rate constants k are much higher than those expected from water exchange and can only be explained by a strong internal conjugate base effect. In the limiting case, however, this is equivalent to the second possible explanation, which assumes MOH+ and HDACODA? as reactive species. The dissociation rate is given by vd = (kML + k [H+]) · [M(DACODA)].  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