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 796 毫秒
1.
The ligands (L) bis (2-pyridyl) methane (BPM) and 6-methyl-bis (2-pyridyl)methane (MBPM) form the three complexes CuL2+, CuL, and Cu2L2H with Cu2+. Stability constants are log K1 = 6.23 ± 0.06, log K2 = 4.83 ± 0.01, and log K (Cu2L2H + 2H2+ ? 2 CuL2+) = ?10.99 ± 0.03 for BPM and 4.56 ± 0.02, 2.64 ± 0.02, and ?11.17 ± 0.03 for MBPM, respectively. In the presence of catalytic amounts of Cu2+, the ligands are oxygenated to the corresponding ketones at room temperature and neutral pH. With BPM and 2,4,6-trimethylpyridine (TMP) as the substrate and the buffer base, respectively, the kinetics of the oxygenation can be described by the rate law with k1 = (5.9 ± 0.2) · 10?13 mol l?1 s?1, k2 = (4.0 ± 0.6) · 10?4 mol?1 ls?1, k3 = (1.1 ± 0.1) · 10?12 mol l?1 s?1, and k4 = (9 ± 2) · 10?14 mol l?1 s?1.  相似文献   

2.
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   

3.
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   

4.
The kinetics of the bromate ion-iodide ion-L-ascorbic acid clock reaction was investigated as a function of temperature and pressure using stopped-flow techniques. Kinetic results were obtained for the uncatalyzed as well as for the Mo(VI) and V(V) catalyzed reactions. While molybdenum catalyzes the BrO-I? reaction, vanadium catalyzes the direct oxidation of ascorbic acid by bromate ion. The corresponding rate laws and kinetic parameters are as follows. Uncatalyzed reaction: r2 = k2[BrO] [I?][H+]2, k2 = 38.6 ± 2.0 dm9 mol?3 s?1, ΔH? = 41.3 ± 4.2 kJmol?1, ΔS? = ?75.9 ± 11.4 Jmol?1 K?1, ΔV? = ?14.2 ± 2.9 cm3 mol?1. Molybdenum-catalyzed reaction: r2 = k2[BrO] [I?] [H+]2 + kMo[BrO] [I?] [ H+]2[M0(VI)], kMo = (2.9 ± 0.3)106 dm12 mol?4 s?1, ΔH? = 27.2 ± 2.5 kJmol?1, ΔS? = ?30.1 ± 4.5 Jmol?1K?1, ΔV? = 14.2 ± 2.1 cm3 mol?1. Vanadium-catalyzed reaction: r1 = kV[BrO] [V(V)], kV = 9.1 ± 0.6 dm3 mol?1 s?1, ΔH? = 61.4 ± 5.4 kJmol?1, ΔS? = ?20.7 ± 3.1 Jmol?1K?1, ΔV? = 5.2 ± 1.5 cm3 mol?1. On the basis of the results, mechanistic details of the BrO-I? reaction and the catalytic oxidation of ascorbic acid by BrO are elaborated. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
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   

6.
The autoxidation of two cuprous complexes, Cu(NH3)2+ and Cu(imidazole)2+, has been studied by following the oxygen consumption with a coated oxygen sensor, and the formation of CuII by means of a stopped flow technique. The reaction was found to be of third order being proportional to the concentrations of CuI, oxygen, and free ligand. pH variation was without effect in the range studied. The rate constants are kIM = 5,5 ·103 12· Mol?2·s?1 for imidazole, and kNH3 = 1,6·104 12· Mol?2· s?1 for NH3 as ligand, resp. An apparent activation energy of less than 2 Kcal/mole has been found for the autoxidation of the cuprous imidazole complex. This can be explained by the assumption of a rapidly playing equilibrium proceeding the rate determining step.  相似文献   

7.
The complexation of 1-methyl-2-hydroxymethyl-imidazole (L) with Cu(I) and Cu(II) has been studied in aqueous acetonitrile (AN). Cu(I) forms three complexes, Cu(AN)L+, CuL2+, and Cu(AN)H?1L, with stability constants logK(Cu(AN)+ + L ? Cu(AN)L+) = 4.60 ± 0.02, logβ2 = 11.31 ± 0.04, and logK(Cu(AN)H?1L+H+ ? Cu(AN)L+) = 10.43 ± 0.08 in 0.15M AN. The main species for Cu(II) are CuL2+, CuH?1L+, CuH?1L2+, and CuH?2L2. The autoxidation of CuL2+ was followed with an oxygen sensor and spectrophotometrically. Competition between the formation of superoxide in a one-electron reduction of O2 and a path leading to H2O2 via binuclear (CuL2)2O was inferred from the rate law with ka = (2.31 ± 0.12) · 104M ?2S ?1, kb = (1.0 ± 0.2) · 103M ?1, kc = (2.85 ± 0.07) · 102M ?2S ?1, kd = 3.89 ± 0.14M ?1S ?1, ke = 0.112 ± 0.004, kf = (2.06 ± 0.24) · 10?10M S ?1, kg = (1.35 ± 0.07) · 10?7 S ?1, and kh = (6.8 ± 1.4) · 10?7M ?1 S ?1.  相似文献   

8.
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   

9.
Kinetics of the complex formation of chromium(III) with alanine in aqueous medium has been studied at 45, 50, and 55°C, pH 3.3–4.4, and μ = 1 M (KNO3). Under pseudo first-order conditions the observed rate constant (kobs) was found to follow the rate equation: Values of the rate parameters (kan, k, KIP, and K) were calculated. Activation parameters for anation rate constants, ΔH(kan) = 25 ± 1 kJ mol?1, ΔH(k) = 91 ± 3 kJ mol?1, and ΔS(kan) = ?244 ± 3 JK?1 mol?1, ΔS(k) = ?30 ± 10 JK?1 mol?1 are indicative of an (Ia) mechanism for kan and (Id) mechanism for k routes (‥substrate Cr(H2O) is involved in the k route whereas Cr(H2O)5OH2+ is involved in k′ route). Thermodynamic parameters for ion-pair formation constants are found to be ΔH°(KIP) = 12 ± 1 kJ mol?1, ΔH°(K) = ?13 ± 3 kJ mol?1 and ΔS°(KIP) = 47 ± 2 JK?1 mol?1, and ΔS°(K) = 20 ± 9 JK?1 mol?1.  相似文献   

10.
The kinetics of the acqueous-phase reactions of the free radicals ·OH, ·Cl, and SO· with the halogenated acetates, CH2FCOO?, CHF2COO?, CF3COO?, and with CH2ClCOO?, CHCl2COO?, CCl3COO? were investigated. Generally, the reactivity decreases with increasing halogen substitution and is in the order k(·OH) > k(SO·) > k(·Cl), but there is no general relation between the effect on reactivity of chlorine and fluorine substitution. © 1995 John Wiley & Sons, Inc.  相似文献   

11.
The complexation of Cu+ by the potentially tripod like ligand cis, cis-1, 3, 5 cyclohexanetriamine (chta) has been studied potentiometrically in aqueous acetonitrile (an). The expected tetracoordinated species Cu (chta) ? (an)+ was formed only at rather high pH with log K (Cu (an)+ + chta ? Cu (chta) · (an)+) = 6.94. Quite unexpectedly the most stable complex in neutral solution was the trimetric species Cu3 (chta) with log K (3 Cu+ + 2 chta ? Cu3 (chta)) = 31.75. In addition, the ternary complexes Cu (LH2) · (an)3+ and Cu (LH) · (an)2+ (L = chta) are formed at low pH. From model considerations, Cu3 (chta) must contain two ligand molecules with all amino groups in equatorial position, linked by three linearly coordinated Cu+-ions. Cu3 (chta)3+2 shows no measurable reactivity towards dioxygen. At pH values above 9, very rapid O2-uptake due to Cu (chta) · (an)+ is observed. In this reaction, Cu+-autoxidation is stoichiometrically coupled to ligand oxidation, followed by a much slower Cu-catalyzed secondary reaction of the primary oxidation product of chta. Hydrogen peroxide and likely also superoxide, are involved in the coupled Cu+/ligand oxidation.  相似文献   

12.
Ab initio calculations of potential energy, dipole moment, equilibrium OH distance, force constants, and anharmonic frequencies, and correlation between these quantities, are presented for a water molecule and an OH? ion in a uniform electric field of varying field strength. It is explained why a bound H2O molecule in nature always experiences a frequency downshift with respect to the free molecule, and a bound OH?1 ion, either a downshift or an upshift. The frequency-field variation is well accounted for by the expression ΔνOH α ?E‖ · (d μ/drOH + 1/2 · ?μ/?rOH). A frequency maximum occurs at the field strength where ?μ‖tot/?rOH ~ 0. Two cases can be discerned: (1) the frequency maximum falls at a positive field strength when dμ/drOH is positive (this is the situation for OH?), and (2) the maximum frequency falls at a negative field when dμ/drOH is negative (this occurs for water). In general, for an OH bond in a bonding situation where the intermolecular interactions are dominated by electrostatic forces, the nonlinearity of the frequency shift with respect to an applied field is governed by how close to the frequency maximum one is, i.e., by both dμ/drOH and ?μ/?rOH. Correlation curves between the external linear force constant, kext, and rOH,e are closely linear over the whole field range studied here, whereas the frequency vs. rOH,e and force constants vs. rOH,e correlation curves form two approximately linear, parallel branches, corresponding to “before” and “after” the maximum in the frequency vs. field curves. Each branch of the ν vs. rOH,e curves has a slope of ~ ? 16,000 cm?1/Å. © 1993 John Wiley & Sons, Inc.  相似文献   

13.
The difference in steric strain between the oxidized and the reduced forms of tetraaminecopper complexes is correlated with the corresponding reduction potentials. The experimentally determined data considered range from ?0.54 to ?0.04 V (vs. NHE) in aqueous solution and from ?0.35 to ?0.08 V (vs. NHE) in MeCN. The observed and/or computed geometries of the tetraaminecopper(II) complexes are distorted octahedral or square-pyramidal (4 + 2 or 4+1) with (distorted) square-planar CuN4 chromophores (CuII? N = 1.99–2.06 Å; Cu? O ≈ 2.5 Å; Cu? O ≈ 2.3 Å), those of the tetraaminecopper(I) complexes are (distorted) tetrahedral (four-coordinate; CuI? N = 2.12–2.26 Å; tetrahedral twist angle ?? = 30–90°). The reduction potentials of CuII/I couples with primary-amine ligands and those with macrocyclic secondary-amine ligands were correlated separately with the corresponding strain energies, leading to slopes of 70 and 61 kJ mol?1 V?1, with correlation coefficients of 0.89 and 0.91, respectively. The approximations of the model (entropy, solvation, electronic factors) and the limits of applicability are discussed in detail and in relation to other approaches to compute reduction potentials of transition-metal compounds.  相似文献   

14.
Ab initio calculations of potential energy, dipole moment, equilibrium OH distance, force constants, and anharmonic frequencies, and correlations between these quantities, are presented for a water molecule and an OH? ion in a uniform electric field of varying field strength. It is explained why a bound H2O molecule in nature always experiences a frequency downshift with respect to the free molecule, and a bound OH? ion either a downshift or an upshift. The frequency-field variation is well accounted for by the expression ΔνOH ∝ ?E·(dμ/drOH + 1/2 · ?μ/?rOH). A frequency maximum occurs at the field strength where ?μ/?rOH ~ 0. Two cases can be discerned: (1) the frequency maximum falls at a positive field strength when dμ/drOH is negative (this is the situation for OH?), and (2) the maximum frequency falls at a negative field when dμ/drOH is positive (this occurs for water). In general, for an OH bond in a bonding situation where the intermolecular interactions are dominated by electrostatic forces, the nonlinearity of the frequency shift with respect to an applied field is governed by how close to the frequency maximum one is, i.e., by both dμ/drOH and ?μ/?rOH. Correlation curves between the external linear force constant, kext, and rOH,e are closely linear over the whole field range studied here, whereas the frequency vs. rOH,e and force constants vs. rOH,e correlation curves form two approximately linear, parallel branches, corresponding to “before” and “after” the maximum in the frequency vs. field curves. Each branch of the v vs. rOH,e curves has a slope of ~ ?16,000 cm?1/Å. © 1993 John Wiley & Sons, Inc.  相似文献   

15.
Acid hydrolysis of the ester function in Δ-(?)5892-(RR)-[Co (trien) (glyOEt) Cl]2+ ((?)- 1 ) produces optically pure Δ-(?)589-(RR)-[Co (trien) (glyOH)Cl]2+ ((?)- 4 ). Hg2+ induced removal of chloride in (?)- 4 follows the rate law kobs = kHg [Hg2+] with kHg = (1.36 ± 0.03) × 10?2 M?1s?1, 25°, μ=1.0, and produces optically pure Δ-(?)5892-(RR)-[Co(trien) (glyO)]2+ ((?)- 2 ). Competition by NO occurs in this reaction ([NO], 1M, 3%) indicating a path whereby external nucleophiles (Y?NO, H2O) compete with the intramolecular carboxylate function for an intermediate of reduced coordination number. Rapid ring closure to 2 must ensue for Y ? H2O. Base hydrolysis of chloride in (±)- 1 produces (±)- 2 together with its diastereoisomer β2-(RS, SR)-[Co(trien) (glyO)]2+, ((±)- 3 ), in which one secondary amine function has an inverted configuration. 2 and 3 incorporate 18O-labelled solvent into the Co-O position of the coordinated carboxylate moiety ( 2: 9.0%; 3: 12.3%) indicating that at least part of the product arises via intramolecular hydrolysis in β2-hydroxo ethylglycinate intermediates (Fig. 4). Base hydrolysis of (?)- 4 follows the rate law Kobs = kOH[OH?] with kOH = (6.3 ± 0.6) × 10?4M?1 S?1, 25°, μ = 1.0 producing (?)- 2 (37-45%) and (?)- 3 (63-55%), the ratio being somewhat medium dependent. Competition by added N (1M) occurs using (±) - 4 forming β2-(RR, SS)-[Co (trien) (glyO)N3]+ (~2%) and β2-(RS, SR)-[Co (trien) (glyO)N3]+ (~ 13%). Mutarotation at the secondary nitrogen centre is shown to occur after the rate determining loss of Cl? in 1 and 4 and before the formation of 2 and 3 . It is concluded that this secondary nitrogen is the site of deprotonation in the reactive conjugate bases of 1 and 4 , and possible mechanisms for the mutarotation process are considered.  相似文献   

16.
Structural Interactions of Planar and Non‐planar Bis(1,2‐dithiosquarato)metalate Host Lattices with CuII Complexes – Structure and EPR Investigations 1,2‐Dithiosquaratometalates (M = Cu, Ni, Zn) are available by direct synthesis from metal salts with dipotassium‐1,2‐dithiosquarate. The structural influence of the planar and nonplanar host lattice systems (BzlEt3N)2[Cu/Ni(dtsq)2] and (BzlEt3N)2[Cu/Zn(dtsq)2] on the geometrical and electronic structure of the CuII guest complex [Cu(dtsq)2]2– is studied by EPR spectroscopy. The used host lattices (BzlEt3N)2[Ni(dtsq)2] (planar) and (BzlEt3N)2[Zn(dtsq)2] (tetrahedral) are characterized by X‐ray structure analysis. (BzlEt3N)2[Ni(dtsq)2] crystallizes in the triclinic unit cell P1 with a = 9.1021(8) Å, b = 9.4190(8) Å, c = 11.0119(10) Å, α = 92.8560(10)°, β = 95.375(2)°, γ = 104.5180(10)° and Z = 1. (BzlEt3N)2[Zn(dtsq)2] crystallizes in the monoclinic unit cell C2/c with a = 21.1299(14) Å, b = 16.6641(11) Å, c = 13.8324(9) Å, β = 123.9100(10)° and Z = 4. The g and A Cu tensors in the Cu/Ni system are nearly axial symmetric (g|| = 2.122, g = 2.028; A = –159.5 · 10–4 cm–1, A = –36.9 · 10–4 cm–1). The coordination geometry of the CuII guest complex in the tetrahedral Cu/Zn system is rather distorted, which is shown by the changed g and A Cu tensor parameters (g|| = 2.143, g = 2.042; A = –103.0 · 10–4 cm–1, A ≈ –5.0 · 10–4 cm–1). The spin density distribution is discussed using EHT molecular orbital calculations.  相似文献   

17.
An analysis of the former works devoted to the reactions of I(III) in acidic nonbuffered solutions gives new thermodynamic and kinetic information. At low iodide concentrations, the rate law of the reaction IO + I? + 2H+ ? IO2H + IOH is k+B [IO][I?][H+]2k?B [IO2H][IOH] with k+B = 4.5 × 103 M?3s?1 and k?B = 240 M?1s?1 at 25°C and zero ionic strength. The rate law of the reaction IO2H + I? + H+ ? 2IOH is k+C [IO2H][I?][H+] – k?C [IOH]2 with k+C = 1.9 × 1010 M?2s?1 and k?C = 25 M?1s?1. These values lead to a Gibbs free energy of IO2H formation of ?95 kJ mol?1. The pKa of iodous acid should be about 6, leading to a Gibbs free energy of IO formation of about ?61 kJ mol?1. Estimations of the four rate constants at 50°C give, respectively, 1.2 × 104 M?3s?1, 590 M?1s?1, 2 × 109 M?2s?1, and 20 M?1 s?1. Mechanisms of these reactions involving the protonation IO2H + H+ ? IO2H and an explanation of the decrease of the last two rate constants when the temperature increases, are proposed. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 647–652, 2008  相似文献   

18.
One unit of S(IV) (SO2 or SHO3?) is oxidized per 2 units of [NiIII(cyclam)] species to obtain sulfate. Kinetic analyses have been done by varying the acidities (0.013 ? [H+] ? 1.0 M) and halide concentrations (0.000 ? [X?] ? 0.012 M; X=Cl and Br) at constant ionic strength (μ = 1.0 M). The rate law that incorporates the [X?] and [H+] dependence is ?d[NiIII]T/dt=2k[NiIII]T[S(IV)]T where 2k={ka[H+] + kbK + kKX[H+] [X?] + kKXK[X?]} {[H+] + K}?1 {1 + KX[X?]}?1, here ka=87 ± 7 M?1 s?1, kb=(2.5 ± 0.5)×103 M?1 s?1 and pK = 1.8 ± 0.2. Rate constants ka and kb are attributed to the reactions of [NiIII(cyclam) (H2O)2]3+ with SO2 and SHO3?, respectively. Monohalo species apparent equilibrium constants KCl=(1600 ± 400) M?1 and KBr=(190 ± 20) M?1 and rate constants k=80 ± 8 M?1 s?1 and k = 140 ± 15 M?1 s?1 are ascribed to the protonated pathway, involving the [NiIII(cyclam) (H2O)X]2+ and SO2(aq) reaction pairs. The other two rate constants of k=(5 ± 1)×103 M?1 s?1 and k=(3.1 ± 0.5)×104 M?1 s?1, refer to the deprotonated pathway and are assigned to the [NiIII(cyclam) (H2O)X]2+ /SHO3? redox couple. A deuterium H2O / D2O isotope effect of 2.1–2.8 can be attributed partially to an equilibrium isotope effect at low acidity though a small kinetic isotope (2.5 ± 0.5) effect is evident for the dihydrogen sulfito pathway, ka. The kinetic isotope effect and the absence of sulfite radical scavenging effects are explained by a mechanism entailing migration of a hydride from sulfur to the NiIII center to produce a NiIII—H species, which rapidly comproportionates, and S(VI). © 1993 John Wiley & Sons, Inc.  相似文献   

19.
The oxygen-exchange reaction of V10O with bulk water has been followed by time-dependent 17O-NMR spectroscopy (buffered solutions, pH ~ 5.5, [V10]total ~ 0.17m, T = 298 K). It is shown that all seven structurally different sites of O-atoms are kinetically similar but, in contrast to earlier studies, not identical (6 h ? ‘t1/2’ ? 11 h). The kinetic similarity of the various structural sites implies the some (but not full) O scrambling is involved. Two possible mechanisms with a ‘half-bonded’ and an ‘open’ intermediate are discussed in detail to interpret the experimental results. A computer simulation of the exchange reaction based on these models is presented. It is shown that the ‘half-bonded-intermediate’ mechanism is consistent with the experimental data and the following parameters are calculated: formation of the intermediate: k1 = 5.8 · 10?3 s?1, k?1 = 6.7 · 10?2 s?1, [intermediate] ≈ 8%; all activated O-atoms exchange within the lifetime of the intermediate (τ ~ 15 s), and the calculated exchange rate of the intermediate (k2 ? 0.60 s?1) is consistent with earlier assumptions (k2 ≈ 0.5 s?1). It is shown that a simulation based on the ‘open-intermediate’ mechanism results in kinetic parameters which are not consistent with the kinetics of the formation of cyclic metavanadates ((VO)n, n = 4,5) from decavanadate, since the required formation rate is by a factor ~ 102 too fast, and the equilibrium concentration of metavanadates is by a factor of ~ 2 too large (under the conditions of the O-exchange experiments of decavanadate (T = 298 K, [V10]total ≈ 0.17m, pH ~ 5.55) the total amount of metavanadates present is ~ 8%, with [(VO)4]/[(VO)5] ~ 4:1; a qualitative analysis of the kinetics of the formation of metavanadates (vo kinetics; the exact mechanism of the back-reaction (at least second-order) is not known with certainty) leads to k1 ? 4·10?5 s?1). O exchange of decavanadates via equilibrated metavanadates would lead to full scrambling of the O sites and is not consistent with the observed differences in the exchange rates. From the qualitative kinetic parameters of the metavanadate formation kinetics, it can be concluded that any contribution of an ‘open’ or an ‘metavanadate’ mechanism is of the order of 1–2% at most.  相似文献   

20.
Crystal Structure of SrZn(OH)4 · H2O Colorless crystals of SrZn(OH)4 · H2O are obtained by electrochemical oxidation of Zn in a zinc/iron pair in an aqueous ammonia solution saturated with strontium hydroxide. The X-ray crystal structure determination was now successful including all hydrogen positions: P1 , Z = 2, a = 6.244(1) Å, b = 6.3000(8) Å, c = 7.701(1) Å, α = 90.59(1)°, β = 112.56(2)°, γ = 108.66(2)°, N(F ≥ 3σF) = 1967, N(Var.) = 84, R/Rw = 0.020/0.024. In SrZn(OH)4 · H2O Zn2+ is tetrahedrally coordinated by four OH? -ions while Sr2+ has 6 OH? and one H2O as neighbours. The polyhedra around Sr2+ are connected to chains which are linked three-dimensionally by isolated tetrahedra [Zn(OH)4]. Hydrogen bonds between H2O as donor and OH? are characterized by raman spectroscopy.  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