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 39 毫秒
1.
《Inorganica chimica acta》1988,149(2):259-264
The bis(N-alkylsalicylaldiminato)nickel(II) complexes Ni(R-sal)2 with R = CH(CH2OH)CH(OH)Ph (I), R = CH(CH3)CH(OH)Ph (II) and R = CH2CH2Ph (III; Ph = phenyl) were prepared and characterized. In the solid state I and II are paramagnetic (μ = 3.2 and 3.3 BM at 20 °C, respectively), whereas III is diamagnetic. It follows from the UV-Vis spectra that in acetone solution I is six-coordinate octahedral and III is four-coordinate planar, the spectrum of II showing characteristics of both modes of coordination. Vis spectrophotometry and stopped-flow spectrophotometry were applied to study the kinetics of ligand substitution in I–III by H2salen (= N,N′-disalicylidene-ethylenediamine) in the solvent acetone at different temperatures. The kinetics follow a second-order rate law, rate = k[H2-salen] [complex]. At 20 °C the sequence of rate constants is k(III):k(II):k(I) = 11 850:40.6:1. The activation parameters are ΔH(I) = 112, ΔH(II) = 40.7, ΔH(III) = 35.7 kJ mol−1 and ΔS(I) = 92, ΔS(II) = −103, ΔS(III) = −89 J K−1 mol−1. The enormous difference in rate between complexes I, II and III, which is less pronounced in methanol, is attributed to the existence of a fast equilibrium planar ⇌ octahedral, which is established in the case of I and II by intramolecular octahedral coordination through the hydroxyl groups present in the organic group R. An A-mechanism is suggested to control the substitution in the sense that the entering ligand attacks the four-coordinate planar complex, the octahedral complex being kinetically inert.  相似文献   

2.
Nickel(II) complexes with the compartmental Schiff bases derived from 2,6-diformyl-4-chlorophenol and 1,5-diamino-3-thiapentane (H2L1) or 3,3′-diamino-N-methyl-dipropylamine (H2L2) were synthesized, and the crystal structures of [Ni(L1)- (py)2] and [Ni(L2)(dmf)]·H20 were determined by X-ray crystallography.Ni(L1)(py)2 is monoclinic, space group C2/c, with a= 18.457(6), b = 11.116(7), c= 16.098(6) Å, and β = 115.79(5)°; Dc = 1.49 g cm−3 for Z = 4. The structure was refined to the final R of 6.9%. The molecule has C2 symmetry. The nickel atom is six-coordinated octahedral. Selected bond lengths are: NiO 2.04(1) Å, NiN (L1) 2.08(1) Å, NiN(py) 2.17(1) Å.[Ni(L2)(dmf)]·H2O is monoclinic, space group P21/n, with a = 17.329(6), b = 13.322(7), c = 12.476(7) Å and β = 95.43(5)°; Dc = 1.45 g cm−3 for Z = 4. The structure was refined to the final R of 5.1%. The nickel atom is bonded in the octahedral geometry to the bianionic pentadentate ligand L2 and to one molecule of dimethylformamide. Selected bond lengths are: NiO (charged) 2.063(3) Å (mean value), NiO (neutral) 2.120(3) Å, NiN (planar) 2.050(3) Å (mean value), NiN (tetrahedral) 2.177(3) Å.  相似文献   

3.
The kinetics of malonate replacement in bis- (malonato)oxovanadate(IV), [VO(mal)2H2O]2−(hereafter water molecule will be omitted), by oxalate has been studied by the stopped-flow method. The reaction was found to consist of two consecutive steps (k1 and k2: first-order rate constants) passing through a mixed ligand complex, [VO(mal)(ox)]2−. The rates for each step depended linearly on the concentrations of free oxalate species, Hox and ox2−. The second-order rate constants for the replacement by ox2− were much larger in the k1 step than in the k2 step and the activation parameters were determined as follows: ΔH= 43.5 ± 5.6 kJ mol−1, ΔS±-53 ± 19 J K−1 mol−1 and ΔH≠= 43.6 ± 0.5 kJ mol−1, δS≠ = -62 ± 2 J K−l mol−1 for the k1 and k2 steps, respectively. The volume of activation was determined to be -0.65 ± 0.75 cm3 mol−1 at 20.2 °C by the high-pressure stopped-flow method for the apparent rate constants.  相似文献   

4.
《Inorganica chimica acta》1986,121(2):175-183
Chloride anation of trans-Pt(CN)4ClOH2 has been studied with and without Pt(CN)42− present at 25.0°C by use of stopped-flow and conventional spectrophotometry and a 1.00 M perchlorate medium. The rate law in the absence of Pt(CN)42− is Rate=(p1 + p2 [H+] ) [Cl]2 [complex]/(1 + q [Cl]) with p1=(3.0 ± 0.1) × 10−5 M−2s−1, p2=(3.6 ± 0.1) × 10−5 M−3 s−1 and q=(0.62 ± 0.02) M−1. It is compatible with a chloride assistance via an intermediate of the type Cl-Cl-Pt(CN)4···OH22−, in which the reactivity of the aqua ligand is enhanced due to a partial reduction of the platinum. This mechanism of halide assistance is in principle the same as the modified reductive elimination oxidative addition (REOA) mechanism proposed by Poë, in which the intermediate is not split into free halogen, platinum(II) and water, and in which electron transfer not necessarily involves complete reduction to platinum(II). To avoid confusion with complete reductive eliminations, reactions without split of the intermediates are here termed halide-assisted reactions. The pH-dependence indicates acid catalysis via a protonated intermediate ClClPt(CN)4···OH3.The Pt(CN)42−accelerated path has the rate law Rate=
[Cl-] [Pt(CN)42−] [complex] where k=(39.9±0.5) M−2 s−1 and Ka=(4.0±0.2)10−2 M is the protolysis constant of trans-Pt(CN)4ClOH2−.Reaction between PtCl5OH2 and chloride is accelerated by Pt(CN)42− and gives PtCl62− as the reaction product. The rate law is Rate=k [Cl] [Pt(CN)42−] [PtCl5OH2] with k=(5.6 ± 0.2)10−3 M−2 s−1 at 35.0°C and for a 1.50 M perchlorate acid medium. The reaction takes place without central ion exchange. Alternative mechanisms with two consecutive central ion exchanges can be excluded. The role of Pt(CN)42− in this reaction is very similar to that of the assisting halide in the halide assisted anations. [p ]Reaction between trans-Pt(CN)4ClOH2 and PtCl42− gives Pt(CN)42− and PtCl5OH2 as products and has the rate law Rate=k[PtCl42−] [trans-Pt(CN)4ClOH2] with k=(3.32 ± 0.02) M−1 s−1 at 25 °C for a 1.00 M perchloric acid medium. The formation of an aqua complex as the primary reaction product and the rate independent of [Cl] shows that formation of a bridged intermediate of the type Pt(II)Cl4ClPt(IV)(CN)4OH23− is formed in the initial reaction step, not five-coordinated PtCl53−.  相似文献   

5.
The preparation of the planar yellow [Ni([8]aneN2)2](ClO4)2 is described. The complex dissociates in basic solution, with rate = kOH[NiL][OH?] (L = 1,5-diazacyclo-octane). At 25 °C, kOH = 4.5 x 10?2 M?1 s?1 and the corresponding activation parameters are ΔH = 69.2 kJ mol?1 and ΔS298 = ?38.6 J K?1 mol?1. Acid catalysed dissociation in quite slow even in strongly acidic solutions. The kinetic data in this case can be fitted to the expression Kobs = ko + KH[H+], where ko relates to a solvolytic pathway and kH to the acid catalysed pathway. At 60 °C, Ko = 2 x 10?5 s?1 and kH is 2 x 10?5 M?1 s?1. Possible mechanisms for these reactions are considered.The Ni(II)/Ni(III) redox couple for NiLn+ is irreversible on Pt using MeCN as solvent.  相似文献   

6.
The crystal and molecular structure of nitrosyltris-(trimethylphosphine)nickel(O) hexafluorophosphate, {Ni(NO)(PMe3)3}PF6, has been determined from three dimensional single crystal X-ray analysis. The compound crystallizes in the orthorhombic space group Pnma with Z = 4 and a unit cell of dimensions: a = 16.253(3), b = 10.536(1) and c = 12.228(2) Å. The structure was solved by conventional heavy atom techniques and refined by least-squares methods to R1 = 0.036 and R2 = 0.048 respectively for 1085. independent reflections. The coordination geometry around the nickel is a slightly distorted tetrahedron with an average PNiP angle of 105.63° and PNiN angle 113.03°. The nickel nitrosyl group is slightly bent with an NiNO angle of 175.4(5)°. The bending occurs in the ClPlNiNO plane toward Pl. The structure is compared with other tetrahedral {MNO}10 phosphine complexes and the MNO bonding is discussed.  相似文献   

7.
In CD3CN solutions the kinetic parameters characterising rotation about the CNMe2 and CNH2 bonds in [UO2(1,1-DMU)5]2+ (1,1-DMU = 1,1- dimethylurea) were determined as: k(265 K) = 39.1 ± 0.4 and 2960 ± 60 s?1, ΔH3 = 49.1 ± 0.76 and 61.1 ±0.5 kJ mol?1, ΔS2 = ?28.3 ± 2.7 and 53.1 ± 2.2 J K?1 mol?1 respectively from 1H NMR studies. Resonances arising from the three isomeric 1,3-DMU (= 1,3-dimethylurea) ligands were observed for [UO2(1,3-DMU)5]2+ in CD3CN solution and the kinetic parameters characterising their isomerisations were also determined. The three isomers of 1,3-DMU have not previously been detected in solution and it appears that coordination of 1,3-DMU to UO22+ increases the barrier to rotation about the carbon nitrogen bond, as is also shown to be the case for 1,1-DMU.  相似文献   

8.
《Inorganica chimica acta》1987,128(2):169-173
The axial adduct formation of the iron(II) complex of 2,3,9,10-tetraphenyl-l,4,8,11-tetraaza-1,3,8,10-cyclotetradecatetraene (L) with imidazole in dimethyl sulfoxide has been investigated spectrophotometrically at various temperatures and pressures. In the presence of a large excess of imidazole the reaction with the two phases has been observed. The first faster reaction is the formation of the monoimidazole complex of FeL2+, and the second slower reaction corresponds to the formation of the bisimidazole complex. Activation parameters are as follows: for the first step with k1 (25.0°C) = (6.8 ±0.2)×105 mol−1 kg s−1, ΔH31 = 47.5 ± 4.9 kJ mol−1, ΔS31 = 26±16 J K−1 mol−1, and ΔV31 (30.0°C) = 27.2±1.5 cm3 mol−1; for the second step with k2 (25.0°C) = 26.8±0.8 mol−1 kg s−1, ΔH32 = 91.6± 0.8 kJ mol−1, ΔS32 = 90±3 J K−1 mol−1, and ΔV32 (35.0°C) = 21.8±0.9 cm3 mol−1. The large positive activation volumes strongly indicate a dissociative character of the activation process.  相似文献   

9.
The rates of deuterium exchange reactions of malondialdehyde (MDA) and deuterated malondialdehyde (MDAd) have been studied as a function of acidity and the content of dimethyl sulfoxide (DMSO) in binary mixtures with D2O . MDA incorporates deuterium from D2O solutions in a first-order reaction with a rate constant (kobs) that depends on the acid concentration. From this dependence, a catalytic constant, kcat, can be derived (kcatMDA = 2.25 × 105M?s?1). Similar kinetic behavior was found for MDAd in H2O solutions, and in this case, kcatMDA = 1.56 × 105M?1s?1. Results from reactions of MDA and MDAd in identical H2OD2O mixtures show that primary and secondary isotope effects are small (kH/kD = 1.13) and that solvent isotope effects cause most of the differences found between reactions in D2O and H2O. Reactions in binary DMSOd6D2O mixtures show a six-fold rate increase as the proportion of DMSOd6 increases from 50% to 90%. These results also illustrate the relatively high reactivity of MDA at pH values well above its pKa and the importance of medium composition on its reaction rate.  相似文献   

10.
《Inorganica chimica acta》1987,133(2):295-300
The compound K4[Mo2(SO4)4]Br·4H2O has been made and its crystal structure determined. Space group P4/mnc; unit cell dimensions, a = 11.903(2), c = 8.021(1) Å, V = 1136(1) Å3. The compound is isomorphous with the analogous chloride whose structure has been reported. The MoMo and MoBr distances are 2.169(2) and 2.926(1) Å, respectively and the [Mo2(SO4)4] 3− ions reside on crystallographic special positions with 4/m symmetry. The Raman spectra of both the bromo and chloro compounds have been measured and the MoMo stretching frequency is 370 ± 1.5 cm−1 in each, for the compounds containing the natural isotopic distribution of molybdenum. The chloro compound has been prepared containing the pure isotope 92Mo as well, and the Raman spectra recorded. The v(MoMo) band is shifted by 6.8 ± 0.5 cm−1. The compound K4[Mo2(SO4)4]·2H2O has also been prepared with Mo at natural abundance and with the pure isotope 100Mo, whereby a shift of 8.5 ± 0.5 cm−1 was found. These and other results will be discussed with regard to the similarity of the Raman spectra of the Mo2(S04)43− and M02(S04)44− species.  相似文献   

11.
The solvent kinetic isotope effects (SKIE) on the yeast α-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 °C. With p-nitrophenyl-d-glucopyranoside (pNPG), the dependence of kcat/Km on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, DOD(kcat/Km), of 1.9 (±0.3). The two pKas characterizing the pH profile were increased in D2O. The shift in pKa2 of 0.6 units is typical of acids of comparable acidity (pKa=6.5), but the increase in pKa1 (=5.7) of 0.1 unit in going from H2O to D2O is unusually small. The initial velocities show substrate inhibition (Kis/Km~200) with a small solvent isotope effect on the inhibition constant [DODKis=1.1 (±0.2)]. The solvent equilibrium isotope effects on the Kis for the competitive inhibitors d-glucose and α-methyl d-glucoside are somewhat higher [DODKi=1.5 (±0.1)]. Methyl glucoside is much less reactive than pNPG, with kcat 230 times lower and kcat/Km 5×104 times lower. The solvent isotope effect on kcat for this substrate [=1.11 (±0. 02)] is lower than that for pNPG [=1.67 (±0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

12.
《Inorganica chimica acta》1988,148(2):233-240
The complexes CodptX3 and [Codpt(H2O)X2]ClO4 (X = Cl, Br; dpt = dipropylenetriamine = NH(CH2CH2CH2NH2)2) have been prepared and characterized. Rate constants (s−1) for aqueous solution at 25 °C and μ = 0.5 M (NaClO4), for the acid-independent sequential ractions.
have been measured spectrophotometrically. For X = Cl: k1 ⋍ 2 × 10−2, k2 = 1.7 × 10−4 and k3 = 4.8 × 10−6, and for X = Br: k1 ⋍ 2 × 10−2, k2 = 5.25 × 10−4 and k3 = 2.5 × 10−5 The primary equation was found to be acid independent, while the secondary and tertiary aquations were acid-inhibited reactions. For the second step, the rate of the reaction was given by the rate equation
where Ct is the complex concentration in the aqua-and hydroxodihalo species, k2 is the rate constant for the acid-dependent pathway and Ka is the equilibrium constant between the hydroxo and aqua complex ions. The activation parameters were evaluated, for X = Cl: ΔH2 = 106.3 ± 0.4 kJ mol−1 and ΔS2 = 40.2 ± 1.7 J K−1 mol, and for X = Br: ΔH2 = 91.6 ± 0.4 kJ mol−1 and ΔS2 = 0.4 ± 1.7 J K−1 mol−1. The results are discussed and detailed comparisons of the reactivities of these complexes with other haloaminecobalt(III) species are presented.  相似文献   

13.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

14.
The synthesis and characterisation of a series of dinuclear and polynuclear coordination compounds with 4-allyl-1,2,4-triazole are described. Dinuclear compounds were obtained for Mn(II) and Fe(II) with composition [M2(Altrz)5(NCS)4], and for Co(II) and Ni(II) with composition [M2(Altrz)4(H2O)(NCS)4](H2O)2. The crystal structure of [Co2(Altrz)4(H2O)(NCS)4](H2O)2 was solved at room temperature. It crystallizes in the monoclinic space group P21/n. The lattice constants are a = 18.033(3) Å, b = 13.611(2) Å, c = 15.619(3) Å, β = 92.04(2)° Z = 4. One cobalt ion has an octahedrally arranged donor set of ligands consisting of three vicinal nitrogens of 1,2-bridging triazoles (CoN = 2.14–2.15 Å), one terminal triazole nitrogen (CoN = 2.12 Å) and two N-bonded NCS anions (CON = 2.08 Å). The other Co(II) ion has the same geometry, but the terminal triazole ligand is replaced by H2O (CoO = 2.15 Å). The crystal structure is stabilised by hydrogen bonding through H2O molecules, S-atoms of the NCS anions and the lone-pair electron of the monodentate triazole. The magnetic exchange in the Mn, Co and Ni compounds is antiferromagnetic with J-values of ?0.4 cm?1, ?10.9 cm?1 and ?8.7 cm?1 respectively. The Co compound was interpreted in terms of an Ising model. For [Zn2(Altrz)5(NCS)2]∞[Zn(NCS)4], [Cu2(Altrz)3(NCS)4]∞ and [Cd2(Altrz)3(NCS)4]∞ chain structures are proposed. In the Cu compound thiocyanates appear to be present, bridging via the nitrogen atom, as deduced from the IR spectrum.  相似文献   

15.
Synthetic [125I]-Tyr23, Phe2, Nle4-adrenocorticotropin (ACTH)-(1–38) ([125I]-ACTH analog) with full biological potency and near theoretical specific radioactivity (1800 ± 75 Ci/mmol) was used to investigate ACTH receptors on isolated rat adipocytes derived from 42-day-old rats. Binding to adipocytes was studied in the presence of 1% bovine serum albumin (BSA) as well as 4% BSA. The interaction of the [125I]-ACTH analog with adipocytes was highly specific, rapid, saturable, and reversible. Scatchard analysis of the binding data obtained in medium containing 1% BSA revealed a single class of binding sites with an apparent KD = 170 ± 11.9 pM. Competition experiments with unlabeled ACTH also yielded a comparable value for the apparent KD (143 ± 16.5 pm). The number of receptors per adipocyte was quite low (521–841/cell). The stimulation of lipolysis by ACTH was closely correlated with the binding, the apparent Km being 145–177 pm. At a concentration of 4% BSA in the incubation medium, the binding curve was shifted significantly to the right (apparent KD = 446 ± 77 pM) and the binding capacity was also significantly enhanced (1663 ± 208/cell) without any change in the apparent Km for glycerol release (187 ± 7.1 pm).  相似文献   

16.
The outer sphere reductions of Co(NH3)5B3+ by Fe(CN)5A3− have been studied. The observed pseudo first order rate constants (Co complex in excess) obey the dependence kobs=Kosket[Co]/(1 +Kos[Co]), as expected for outer sphere electron transfer reactions. Values of the fundamental electron transfer rate constants ket have been determined, along with the equilibrium constant Kos for a range of reactions in which A and B are pyridyl ligands of different sizes. The first order electron transfer rate constants vary in a manner that is consistcnt with adiabatic electron transfer. The outer sphere ion pairing equilibrium constants Kos have been calculated: Kos=8.6 ± 0.1 × 102 M−1 when A and B=pyridine; Kos=1.07 ± 0.09 × 103 M−1 where A=pyridine, B=1-phenyl-3-(4-pyridyl)propane; Kos=1.86 ± 0.11 × 103 M−1 when A=4,4′-bipyridine, B=pyridine; Kos=1.27 ± 0.08 × 103 M−1 when A=4,4′-bipyridine, B=4-phenylpyridine. Distances of closest approach between the metal centers in the reactive ion pairs are compared, and it is concluded that there is a common mechanism, in which the ammonia side of the cobalt complex approaches the cyano side of the iron complex in each reactive ion pair.The distance of closest approach between the two metal centers (a) was calculated from the experimental values for the ion pairing equilibrium constant Kos at 25 °C: 5.2 Å when A=4,4′-bipyridine, B=pyridine; 5.4 Å when A=4,4′-bipyridine, B=4-phenylpyridine; 5.5 Å when A=pyridine, B=1-phenyl-3-(4-pyridyl)propane; 5.7 Å when A=B=pyridine. These relatively short metal-metal distances, when compared to the X-ray structure of the compound [Co(NH3)5(4-phenylpyridine)]2[S2O6]3· 4H2O, do not support an ion pair orientation in which the two substituted pyridine ligands A and B are oriented toward each other. [P21/c,a=7.399(3), b=22.355(10), c=13.776(4) Å, β=92.02(3)°, R=0.070.] The crystallographic results show that if the two pseudo-octahedral coordination spheres are oriented in the reactive ion pair so that an ammonia face of the cobalt complex is at hydrogen bonding distance from a cyano face on the iron complex, the metal-metal distance is 5.3 Å, a distance which is in agreement with the kinetic results.  相似文献   

17.
The kinetics and mechanism of a linear trihydroxamic acid siderophore (deferriferrioxamine B, H4DFB+) ligand exchange with Al(H2O)63+ to form mono(deferriferrioxamine B)aluminum(III) (Al(H2O)4H3DFB)3+ have been investigated at 25 °C over the [H+] range 0.001−1.0 M and I = 2.0 M (HClO4/NaClO4) by 27Al NMR. Kinetic results are consistent with Al(H2O)4(H3DFB)3+ formation and dissociation proceeding through a parallel path mechanistic scheme involving Al(H2O)63+(k2/k−1) and Al(H2O)5(OH)2+(k2/k−2) where k1 = 0.13 M−1 s−1, k−1 = 8.7 × 10−3 M−1 s−1, k2 = 2.7 × 103 M−1 s−1, and k−2 = 9.6 × 10−4 s−1. Relative complex formation rates at Al(H2O)63+ and Al(H2O)5OH2+, and comparison with kinetic data for a series of synthetic hydroxamic acids, suggest that an interchange mechanism is operative. These results are also discussed in relation to kinetic data for the corresponding iron(III)-deferriferrioxamine B system.  相似文献   

18.
《Inorganica chimica acta》1986,115(2):153-161
In the reaction of the tetradentate ligand 3,3′-(1,4- butanediyldiamino) bis (3-methyl-2-butanone)-dioxime (BnAO) with nickel(II) and copper(II), the monomeric [Ni(BnAO-H)]I·H2O and a mixed monomer/dimer salt [Cu(BnAO-H)H2O]2[(Cu(BnAO-H))2](ClO4)4, respectively, are formed, and all complexes have an intramolecular hydrogen bond between cis oxime groups. The OHO bonds give the characteristic infrared absorptions as well as the downfield proton-NMR signal (Ni complex). [Ni(BnAO-H)]I·H2O crystallizes in space group P21/a with a=13.511(2), b=10.599(2), c=14.096(2) Å, β=97.52°, Z=4 and Dc=1.623 g/cm3. The structure was solved by Patterson and Fourier methods and refined by full-matrix least-squares techniques to a final R of 0.021 for 2124 reflections with I 2σ(I). The nickel(II) atom in the complex has slightly distorted square planar geometry with an intramolecular O···O contact of 2.417(7) Å. The copper(II) complex crystallizes in space group P21/c with a =13.425(2), b=21.446(3), c=14.349(4) Å, β= 104.4(5)°, Z=8 (monomers) and Dc=1.485 g/cm3. The final R value for this complex was 0.053 for 3033 reflections with I 2σ(I). This structure contains a monomeric [Cu(BnAO-H)H2O]+ ion and a dimeric [(Cu(BnAO-H))2]2+ ion, having intramolecular O···O hydrogen bonds of 2.421(5) and 2.531(5) Å, respectively. The copper(II) ions have square-pyramidal coordination with the axial positions occupied by an oxygen of the water of hydration in the monomer and by an oxime oxygen atom in the dimer. A center of symmetry relates the two halves of the dimer. The copper atom in each case is out of the plane of the four nitrogen atoms toward the axial site. The copper(II) complex is unusual in that the crystal contains both a monomer and a dimer.  相似文献   

19.
《Inorganica chimica acta》1988,149(1):151-154
The extraction equilibrium of the hydronium-uranium(VI)-dicyclohexano-24-crown-8 complex was carried out in the crown ether1,2-dichloroethaneHCl aqueous solution system at different temperatures. The extraction complex has the overall composition (L)2·(H3O+·χH2O)2·UO2Cl42− (L = dicyclohexano-24-crown-8). The values of the extraction equilibrium constants (Kex) increase steadily with a decrease in temperature: 13.5 (298 K), 7.96 (301 K), 4.20 (303 K) and 2.07 (305 K). A plot of log Kex against 1/T shows a straight line. The value of the enthalpy change, ΔH°, was calculated from the slope and equals −212 kJ mol−1. The value of the entropy change, ΔS°, was calculated from ΔH° and Kex and equals −690 J K−1 mol−1, whereas ΔG° = −6.45 kJ mol−1. Comparing these thermodynamic parameters with those of the dicyclohexano-18-crown-6 isomer A [1] (ΔS° = −314 J K−1 mol−1, ΔH° = −101 kJ mol−1 and ΔG° = −8.37 kJ mol−1), it can be seen that ΔH° and ΔS° are more negative for the former than for the latter, and both are enthalpy-stabilized complexes. The molecular structure of the complex has the feature that there are two H5O2+ ions in it, in contrast to the H3O+ ions in the dicyclohexano-18-crown-6 isomer A complex [1]. Each of the H5O2+ ions is held in the crown ether cavity by four hydrogen bonds. The H5O2+ ion has a central bond. The uranium atom forms UO2Cl42− as a counterion away from the crown ether. The formation of this complex is in good agreement with more negative entropy change and less negative free energy change, as mentioned above.  相似文献   

20.
《Inorganica chimica acta》1988,150(1):81-100
The (NH3)5CoOC(NH2)23+ ion is consumed in water according to the rate law k(obs.) = k1 + k2[OH], where k1 = 4.0 × 10−5 s−1 and k2 = 14.2 M−1 s−1 (0–0.1 M [OH];μ = 1.1 M, NaClO4, 25 °C). A hitherto unrecognized intramolecular O- to N- linkage isomerization reaction has been detected. In strongly acid solution only aquation to (NH3)5CoOH23+ is observed, but in 0.1–1.0 M [OH], 7% of the directly formed products is the urea-N complex (NH3)5CoNHCONH22+ which has been isolated. In the neutral pH region a much greater proportion (25%) of the products is the urea-N species. These results are interpreted in terms of an urea-O to urea-N linkage isomerization reaction competing with hydrolysis for both spontaneous (k1) and base-catalyzed (k2) pathways; the rearrangement is not observed in strongly acidic solution (pH ⩽ 1) because the protonated N-bonded isomer (pKa ≈ 3) is unstable with respect to the O-bonded form. The appearance of the isomerization pathway as the pH is raised in the 0–6 region is commensurate with a rate increase which cannot be attributed to a contribution from the base catalysis term k2[OH]. It is argued that this observation establishes, for the spontaneous pathway, that hydrolysis and linkage isomerization are separate reaction pathways — there is no common intermediate. The product distribution and rate data lead to the complete rate law, k(obs.) = k1 + k2[OH] = (ks + kON) + (kOH + kON) [OH] for the reactions of the O-bonded isomers, where ks, kOH are the specific rates for hydrolysis, and kON, kON are the specific rates for O- to N-linkage isomerization, by spontaneous and base-catalyzed pathways respectively; kON = 1.3 × 10−5 s−1 and kON = 1.1 M−1 s−1 (μ = 1.0 M, NaClO4, 25 °C). The O- to N- linkage isomerization has been observed also for complexes of N-methylurea, N,N-dimethylurea and N-phenylurea, but not for the N,N′-dimethylurea species. There is an approximately statistical relationship among the data for −NH2 capture (versus H2O), while −NHR and −NR2 do not compete with water as nucleophiles for Co(III) in either the spontaneous or base-catalyzed hydrolysis processes. For each urea-O complex, O- to N-isomerization is a more significant parallel reaction in the spontaneous as opposed to the base-catalyzed pathway. This is interpreted as being indicative of more associative character in the spontaneous route to products, a conclusion supported by other evidence. Some activation parameter data have been recorded and the effect of the N-substitution on the rates of solvolysis (H2O, Me2SO) is discussed. The urea-N complexes have been isolated as their deprotonated forms, [(NH3)5CoNHCONRR′](ClO4)2·xH2O (R,R′ = H, CH3). They are kinetically inert in neutral to basic solution but in acid they protonate (H2O, pKa 2–3; μ = 1.0 M, 25 °C) and then isomerize rapidly back to their O-bonded forms. Some solvolysis accompanies this N- to O-rearrangement in H2O and Me2SO. Specific rates and activation parameters are reported. The kinetic data follow a rate law of the form kNO(obs.) = (k + kNO)[H+]/(Ka + [H+]) and the active species in the reaction is the protonated form; k, kNO are the specific rates for hydrolysis and isomerization, respectively. Proton NMR data establish that the site of protonation (in Me2SO) is the cobalt-bound nitrogen atom. For the unsubstituted urea species (NH3)5CoNH2CONH23+, diastereotopic exo-NH2 protons arising from restricted rotation about the CN bond are observed. The relevance to the mechanism of the linkage isomerization process is considered. 13C and 1H NMR and electronic absorption spectral data are presented, and distinctions between linkage isomers and the solution structures (electronic and conformational) are discussed. The urea-N/urea-O complex equilibrium is governed by the relation KNO(obs.) = KNO[H+]/[H+](Ka), where KNO is the equilibrium constant = [(NH35Co(urea-O)3+]/[(NH3)5Co(urea-N)3+]. Values for KNO(=kNO/kON = 260 and pKa ≈ 3 for the NH2CONH2 system are consistent with the stability of the N-isomer in feebly acidic to basic solution (e.g. pH 6, KNO(obs.) = 2.6 × 10−2) and instability in acid solution (e.g. pH 1, KNO(obs.) = 240). The equilibrium data for this and other urea complexes of (NH3)5Co(III) are contrasted with the result for the analogous Rh(III)NH2CONH2 system KNO ≈ 1).  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