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.
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   

2.
The kinetics of the electron transfer between reduced spinach [2Fe-2S]-ferredoxin and the optically active complexes [Co((R,R)- or (S,S)-alamp)py]+ ( I ), [Co((R,R)- or (S,S)-promp)H2O]+ ( IIa ), and [Co((R,R)- or (S,S)-promp)py]+ ( IIb ) have been investigated. The reactions are stereoselective, and for I and IIa , the Stereoselectivity strongly depends on temperature due to large differences in the activation enthalpy between enantiomeric reagents. Isokinetic behaviour is observed between enantiomers, the ΔΔH values being largely compensated by the ΔΔS values. The compensation behaviour is explained by the combination of stereochemical interactions and desolvation processes on formation of the precursor complex or the transition state.  相似文献   

3.
Oxygentation of aqueous solutions of CoIII in presence of stoichiometric amounts of N-(2-aminoethyl)ethane-1,2-diamine (dien) and 1,3-diaminopropan-2-ol (dapo) produces μ-peroxocobalt(III) dimers. Acid cleavage (HCI) yields mer-exo(H)-, mer-endo (H)-, unsym-fac-exo(OH)-, and unsym-fac-endo(OH)-[CoCl(dien)(dapo)]2+ ( A–D )(X = Cl), resp. and unsym-fac-[Co-(dien)(dapo-N,N′,O)]3+ ( G ). Isomer seperation was achieved by fractional crystallization as ZnCl and ClO salts and by ion-exchange chromatography. The corresponding bromo, azido, nitrito-O, nitro-N, thiocyanato, hydroxo, and aqua complexes were also synthesized. Optically resolved samples were prepared for chiral compounds, and the complexes were structurally characterized by X-ray analyses ($ \mathop {\it \Lambda} \limits^ \to $(?)436(CD) -A (X = N3)), ($ \mathop {\it \Delta} \limits^ \to $(?)436(CD) -B ). (X = N3), $ \mathop {\it \Delta} \limits^ \to $ (+)436(CD) -B by their chiroptical properties, and by 13C-NMR spectroscopy supported by 1H-NMR, IR, CD, and UV/VIS spectroscopy. $ \mathop {\it \Lambda} \limits^ \to $(?)436(CD)-mer-exo(H)-[Co(N3)(dien)(dapo)](hydrogen di-O-benzoyl-L-tartrate)2.4 H2O crystallizes in the orthorhombic space group P212121, a = 7.676(1) Å, b = 19.457(1) Å, c = 34.702(2) Å. $ \mathop {\it \Lambda} \limits^ \to $(?)436(CD)-mer-endo(H)-[Co(N3)(dien)(dapo)] (hydrogen di-O-benzoyl-L-tartrate)2.2.75 H2O crystallizes in the triclinic space group P1, a = 8.062(3) Å b = 10.296(1) Å, c = 15.056(2) Å, alpha = 80.55(1)°, β = 85.18(2)°, γ = 89.10(2)°. $ \mathop {\it \Delta} \limits^ \to $(+)436(CD)-mer-endo(H)-[Co(N3)(dien)(dapo)](hydrogen di-O-benzoyl-L-tartrate)2. 5.75 H2O crystallizes in the triclinic space group P1, a = 7.742(1) Å, b = 10.014(1) Å, c = 18.045(2) Å, α = 99.57(1)°, β = 92.87(1)°, γ = 102.56(1)°. The absolute configurations of the three cations were determined unambiguously. Interconversions of the various isomers and derivatives and structural, configurational, and spectroscopic aspects are discussed in detail.  相似文献   

4.
D. C. polarography and cyclic voltammetry were used for investigating the reduction processes of the tris(1,10-phenanthroline)cobalt(II) and bis(2,9-dimethyl-1, 10-phenanthroline)-cobalt(II) perchlorates in 0.1 M solutions of tetraethylammonium perchlorate in acetonitrile. The first complex gave a four-step reduction wave; the first two steps were found to be diffusion controlled and reversible reductions from Co(phen)+ to Co(phen)3+ to Co(phen) to Co(phen;) occured. The second complex gave a six-step reduction wave; the first three steps were found to be diffusion controlled and were to be considered as successive reversible reductions from Co(2, 9dm-phen)+ to Co(2, 9dmphen), from Co(2, 9dmphen) to Co(2, 9dmphen)2 and from Co(2, 9dmphen)2 to Co(2, 9dmphen).  相似文献   

5.
Rb2Co3(H2O)2[B4P6O24(OH)2]: A Borophosphate with ‐Tetrahedral Anionic Partial Structure and Trimers of Octahedra (Co O12(H2O)2) Rb2Co3(H2O)2[B4P6O24(OH)2] is formed under mild hydrothermal conditions (T = 165 °C) from mixtures of RbOH(aq), CoCl2, H3BO3, and H3PO4 (molar ratio 1 : 1 : 1 : 2). The crystal structure (orthorhombic system) was solved by X‐ray single crystal methods (space group Pbca, No. 61; R‐values (all data): R1 = 0.0699, wR2 = 0.0878): a = 950.1(1) pm, b = 1227.2(2) pm, c = 2007.4(2) pm; Z = 4. The anionic partial structure consists of tetrahedral [B4P6O24(OH)28–] layers, which contain three‐ and nine‐membered rings. CoII is octahedrally coordinated by oxygen and oxygen and H2O ligands, respectively (coordination octahedra CoO6 and CoO4(H2O)2). Three adjacent coordination octahedra are condensed via common edges to form trimeric units (CoO12(H2O)2). The oxidation state +2 of cobalt was confirmed by magnetic measurements. The octahedral trimers are quasi‐isolated. No long‐range magnetic ordering occurs down to 2 K. Rb+ is disordered over three crystallographically independent sites within channels of the structure running parallel [010]; the coordination sphere of Rb+ is formed by nine oxygen species of the tetrahedral layers, one OH group and one H2O molecule.  相似文献   

6.
Reactions of oxygenated cobalt(II) complexes. XII. A binuclear μ-peroxodicobalt(III) complex with a macrocyclic bridging ring
  • 1 XI: siehe [1].
  • Singly bridged [(tren) (NH3) CoO2(NH3) (tren)]4+ reacts with excess tren by replacement of NH3 in cis-position to the peroxo group and formation of a new type of doubly bridged μ-peroxo complex. An X-ray structure determination of [(tren)-Co(O2, tren)Co(tren)] (ClO4)4 · 2 H2O showed that the additional tren forms a macrocyclic bridging ring. The conformation of the CoOOCo group is transoid with a dihedral angle of 20°. The crystals are monoclinic with space group P21/c. The lattice constants are a = 9,798, b = 26,385, c = 16,385 Å, β = 110,2° with four formula units in the cell. The final R value is 0,124. ClO anions are disordered. The reactions of [(tren)Co(O2, tren)Co(tren)]4+ in aqueous solution are compared with those of [(tren) (NH3) CoO2Co (NH3tren)]4+. In acidic solution the new complex mainly decomposes to CoII and O2. In alcaline medium the bridging tren is replaced by an OH bridge, forming the well characterized doubly bridged [(tren)-Co(O2, OH)Co(tren)]3+. Differing from the singly bridged bis (ammino) complex, the reactions of which show no pH dependency at all, the decomposition of the tren bridged complex is H+-catalyzed. The kinetic data have been interpreted as (i) preceding fast protonation step which is followed by a conformational change of the bridging ring, (ii) acid hydrolysis of a Co-μ-tren bond and (iii) fast cleavage of the Co-OO bond which is labilized by coordinated H2O.  相似文献   

    7.
    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   

    8.
    Bimetallic and trimetallic compounds containing unsupported bonds of subgroup 4 metals (M = Ti, Zr, Hf) and Co were prepared by hydride elimination (A) from RM derivatives (R1 = PhCH2; RN; R2 = Me, Et)) and by salt elimination (B) from RMX (X = Cl, Br; R.1 = PhCH2, RN and R3O; R3= i-Pr, n-Bu)) by reaction with HCo(CO)4 and Na[Co(CO)4], respectively. Compounds RMCo(CO)4 with R1 = PhCH2, RM[Co(CO)4]2 R.1 = PhCH2, were prepared both by methods A and B, while (R3O)4-n Ti[Co(Co)4]n (n = 1, 2) compounds were obtained by reaction B. Several tertiary phosphine and phosphite derivatives of the former two types were obtained by substitution of a carbonyl group by PR ligand or by A type reaction of HCo(CO)3(PR with RM compounds.  相似文献   

    9.
    The transverse relaxation rate of H2O in Al(H2O) has been measured as a function of temperature (255 to 417 K) and pressure (up to 220 MPa) using the 17O-NMR line-broadening technique, in the presence of Mn(II) as a relaxation agent. At high temperatures the relaxation rate is governed by chemical exchange with bulk H2O, whereas at low temperatures quadrupolar relaxation is prevailing. Low-temperature fast-injection 17O-NMR was used to extend the accessible kinetic domain. The samples studied contained Al3+ (0.5 m), Mn2+ (0.2–0.5 m), H + (0.2–3.1 m) and 17O-enriched (20–40%) H2O. Non-coordinating perchlorate was used as counter ion. The following H2O exchange parameters were obtained: k = (1.29 ± 0.04) s−1, ΔH* = (84.7 ± 0.3) kJ mol−1, ΔS* = +(41.6 ± 0.9) J K −1 mol−1, and ΔV = +(5.7 ± 0.2) cm3 mol−1, indicating a dissociative interchange, Id, mechanism. These results of H2O exchange on Al(H2O) are discussed together with the available complex-formation rate data and permit also the assignment of Id mechanisms to these latter reactions.  相似文献   

    10.
    Kinetik und mechanismus der oxydation von diaqua(nitrilotriacetato)-cobaltat(II) durch peroxodisulfat in wäßrigsauren lösungen Die Kinetik der Oxydation von CoII-NTA durch Peroxodisulfat (S2O) in saurem Medium photometrisch untersucht. Die Stöchiometrie der Reaktion ist: 2CoII-NTA? + S2O → 2CoIII?NTA + 2SO. Im pH-Bereich 4,2—5,4 folgt die Reaktion dem Geschwindigkeitsgesetz [H+] und [S2O] sind die Wasserstoff- bzw. Peroxodisulfationen-Konzentration, KH ist die Dissoziationskonstante des CoII(NTA)H und kH ist die Geschwindigkeitskonstante für den Elektronenübergang. Die Aktivierungsparameter werden mitgeteilt und der mögliche Mechanismus für den Elektronenübergang wird diskutiert.  相似文献   

    11.
    The sulfato bridged MnII phen complex [Mn(phen)(H2O)2SO4] ( 4 ) consists of [Mn(H2O)2(phen)(SO4)2/2] chains, which are generated from Mn atoms in distorted octahedral sites interlinked by bidentate sulfato bridges. Interdigitation of phen ligands leads to the formation of [Mn(H2O)2(phen)(SO4)2/2] double‐chains stabilized by significant π—π stacking interactions. The double‐chains are held together via interchain hydrogen bonds formed between water and sulfato O atoms. Crystal data: orthorhombic, P212121 (no. 19), a = 6.674(1), b = 10.359(1), c = 19.913(3)Å, V = 1376.7(3)Å3, Z = 4, R = 0.0439 and wR2 = 0.0935 for 1830 out of 2286 reflections with Fo2 > 2σ(Fo2).  相似文献   

    12.
    Ab initio calculations are performed with 6–31G basis set to study the geometry and binding of the H3O, H5O, H7O, and H9O complexes. The H3O complex is also investigated with the 6–31 G* basis set and MP 2 (Moller–Plesset perturbation theory of second order).  相似文献   

    13.
    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   

    14.
    Kinetic data for aqueous‐phase reactions of sulfate anion radicals (SO) with perfluorocarboxylates (CnF2n+1C(O)O?) are needed to evaluate removal and transformation processes of CnF2n+1C(O)O? species in the environment, but rate constants for the reactions of SO with CnF2n+1C(O)O? (kn) have been reported only for short‐chain CnF2n+1C(O)O? species (n = 1–3). Since CnF2n+1C(O)O? reacted with SO to form CmF2m+1C(O)O? (m < n), we determined relative rates kn?1/kn for the reactions of SO with CnF2n+1C(O)O? (n = 4–7), along with conversion ratios for conversion of CnF2n+1C(O)O? into Cn?1F2n?1C(O)O?n) and into Cn?2F2n?3C(O)O?n) at 298 K. SO was photolytically generated from S2O by use of sunlamps (λ ≈ 310 nm). Even if kn and kn?1 change with increasing ionic strength, kn?1/kn can be determined when kn?1/kn and αn remain almost constant during the reaction. The values of kn/k1 for n = 4–7 were nearly equal, and their average was 0.82 ± 0.04 (2σ). Conversion ratios of αn and βn were mostly independent of n for n = 4–7, and their averages were 0.77 ± 0.07 (2σ) and 0.13 ± 0.08, respectively. Branching ratios of reactions of a possible intermediate (CnF2n+1O?), reaction of CnF2n+1O? with H2O, and fission of the C? C bond of CnF2n+1O?, seemed to determine αn and βn. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 735–747, 2009  相似文献   

    15.
    An MP4(full,SDTQ)/6-311++G(d,p)//MP2(full)/6-311++G(d,p) ab initio study was performed of the reactions of formyl and isoformyl cations with H2O and NH3, which play an important role in flame and interstellar chemistries. Two different confluent channels were located leading to CO+H3O+/NH. The first one corresponds to the approach of the neutral molecule to the carbon atom of the cations. The second one leads to the direct proton transfer from the cations to the neutrals. At 900 K the separate products CO+H3O+/NH are the most stable species along the Gibbs energy profiles for the processes. For the reaction with H2O the reaction channel leading to HC(OH) (protonated formic acid) is disfavored with respect to the two CO+H3O+ channels in agreement with the experimental evidence that H3O+ is the major ion observed in hydrocarbon flames. According to our calculations, NH+H2O are considerably more stable in Gibbs energy than NH3+H3O+;NH will predominate in the reaction zone when ammonia is added to CH4+Ar diffusion flame, as experimentally observed. At 100 K the most stable structures are the intermediate complexes CO…HOH/HNH. Particularly the CO…HOH complex has a lifetime large enough to be detected and, therefore, could play a certain role in interstellar chemistry. ©1999 John Wiley & Sons, Inc. J Comput Chem 20: 1432–1443, 1999  相似文献   

    16.
    Azide anation and racemization of optically pure mer-exo(H)- and mer-endo(H)-[Co(OH)(dien)(dapo)]2+ ( A and B (X = OH), resp.; dien = N-(2-aminoethyl)ethane-1,2-diamine; dapo = 1,3-diaminopropan-2-ol) involve the same symmetrical pentacoordinate intermediate as the base hydrolyses of the corresponding mer-exo(H)- and mer-endo(H)-[CoX(dien)(dapo)]2+ species A and B , respectively, where X = Cl, Br, or N3. The kinetic parameters of the anation process are fully compatible with the independently measured competition ratio. The rate data reveal that substitution of OH? is unexpectedly fast, viz. it is not consistent with the usual sequence Br? > Cl? > H2O > N > OH?. This behavior is interpreted on the basis of an internal conjugate base mechanism which involves an amino-hydroxo/aminato-aqua tautomerism, viz. the reaction is actually an OH? -catalyzed substitution of [CoH2O(dien)(dapo)]3+ where deprotonation occurs effectively at the secondary-amine site NH of dien.  相似文献   

    17.
    Introduction A series of lanthanide sulfide complexes have beenlargely used for ceramics and thin film materials1 andthese complexes could be prepared from the precursorswhich are the compounds containing lanthanide-sulfurbonds.2-4 For instance, the compounds synthesized with[(alkyl)2dtc]-, phen?H2O and lanthanide salts were usedas the volatile precursors for preparing lanthanide sul-fide, its friction properties in lubricant was investigatedin literature 5 and the preparation and propertie…  相似文献   

    18.
    Rates of solvolysis of ions [Co(3Rpy)4Cl2]+ with R = Me and Et have been measured over a range of temperatures for a series of water-rich water + methanol mixtures to investigate the effect of changes in solvent structure on the solvolysis of complexes presenting a largely hydrophobic surface to the solvent. The variation of the enthalpies and entropies of activation with solvent composition has been determined. A free energy cycle relating the free energy of activation in water to that in water + methanol is applied using free energies of transfer of individual ionic species from water into water + methanol. Data for the free energy of transfer of chloride ions ΔG(Cl?) from both the spectrophotometric solvent sorting method and the TATB method for separating ΔG(salt) into ΔG(i) for individual ions are used: irrespective of the source of ΔG(Cl?), in general, ?ΔG(Co(Rpy)4Cl2+) > ?ΔG(Co(Rpy)4Cl2+), where Rpy = py, 4Mepy, 4Etpy, 3Etpy, and 3Mepy, showing that changes in solvent structure in water-rich water + methanol mixtures generally stabilize the cation in the transition state more than the cation in the initial state for this type of complex ion. A similar result is found when the free energy cycle is applied to the solvolysis of the dichloro (2,2′,2″-triaminotriethylamine)cobalt(III) ion. The introduction of a Me or Et group on the pyridine ring in [Co(Rpy)4Cl2]+ has little influence on the difference {ΔG(Co(Rpy)4Cl2+)?ΔG(Co(Rpy)4Cl2+)} in water + methanol with the mol fraction of methanol < 0.20.  相似文献   

    19.
    Experiments are described in which the kinetics of cathodic hydrogen evolution from the unhydrated H3O+ ion in pure CF3SO H3O+ are compared with those from an aqueous solution of CF3SO3H where the proton is mainly in a fully hydrated state as H9O. From the acid hydrate, which exists mainly as the ionic compound CF3SOH3O+, rates of H2 evolution at Ni, Pt, and Hg electrodes, measured at a given overpotential or expressed as exchange current densities, are between about 3.5 and 20 times slower than those from the same electrolyte in dilute (1.0M) aqueous solution. Allowing for the concentration differences in these two types of system and double-layer effects, the rate constants are between about 9.4 and 216 times smaller for the reaction from H3O+ than from H9O at the above electrodes. The evaluation of apparent heats of activation for H2 evolution from the two types of proton sources allows ratios of real frequency factors to be calculated for discharge from H3O+ and H9O. These data have a bearing on the theoretical conclusions regarding proton discharge mechanisms and show that frequency factor effects can be as important as activation energy differences in determining the rates of proton discharge from different proton sources. The results are discussed in terms of current ideas about electron and proton transfer in electrochemical reactions, the state of hydration of H+, and the role of discharge from paired CF3SO and H3O+ ions. In particular, the molecular mechanics of discharge of the proton from the molecular ion H3O+ can be different from that from the fully hydrated H+ ion where many more HO- vibrational and librational modes can be involved in the process of activation of the H9O entity.  相似文献   

    20.
    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   

    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