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 possible contributions of the mechanochemical triggering effect to the enzymatic activation of the carbon-cobalt bond of coenzyme B12 (5'-deoxyadenosylcobalamin, AdoCbl) for homolytic cleavage have been studied by molecular modeling and semiempirical molecular orbital calculations. Classically, this effect has envisioned enzymatic compression of the axial Co-N bond in the ground state to cause upward folding of the corrin ring and subsequent sterically induced distortion of the Co-C bond leading to its destabilization. The models of this process show that in both methylcobalamin (CH3Cbl) and AdoCbl, compression of the axial Co-N bond does engender upward folding of the corrin ring, and that the extent of such upward folding is smaller in an analog in which the normal 5,6-dimethylbenzimidazole axial ligand is replaced by the sterically smaller ligand, imidazole (CH3(lm)Cbl and Ado(lm)Cbl). Furthermore, in AdoCbl, this upward folding of the corrin is accompanied by increases in the carbon-cobalt bond length and in the Co-C-C bond angle (which are also less pronounced in Ado(Im)Cbl), and which indicate that the Co-C bond is indeed destabilized by this mechanism. However, these effects on the Co-C bond are small, and destabilization of this bond by this mechanism is unlikely to contribute more than ca. 3 kcal mol(-1) towards the enzymatic catalysis of Co-C bond homolysis, far short of the observed ca. 14 kcal mol(-1). A second version of mechanochemical triggering, in which compression of the axial Co-N bond in the transition state for Co-C bond homolysis stabilizes the transition state by increased Co-N orbital overlap, has also been investigated. Stretching the Co-C bond to simulate the approach to the transition state was found to result in an upward folding of the corrin ring, a slight decrease in the axial Co-N bond length, a slight displacement of the metal atom from the plane of the equatorial nitrogens towards the "lower" axial ligand, and a decrease in strain energy amounting to about 8 kcal mol(-1) for both AdoCbl and Ado(Im)Cbl. In such modeled transition states, compression of the axial Co-N bond to just below 2.0 A (the distance subsequently found to provide maximal stabilization of the transition state by increased orbital overlap) required about 4 kcal mol(-1) for AdoCbl, and about 2.5 kcal mol(-1) for Ado(Im)Cbl. ZINDO/1 calculations on slightly simplified structures showed that maximal electronic stabilization of the transition state by about 10 kcal mol(-1) occurred at an axial Co-N bond distance of 1.96 A for both AdoCbl and Ado(Im)Cbl. The net result is that this type of transition state mechanochemical triggering can provide 14 kcal mol(-1) of transition state stabilization for AdoCbl, and about 15.5 kcal mol(-1) for the Ado(Im)Cbl, enough to completely explain the observed enzymatic catalysis. These results are discussed in the light of current knowledge about class I AdoCbl-dependent enzymes, in which the coenzyme is bound in its "base-off" conformation, with the lower axial ligand position occupied by the imidazole moiety of an active site histidine residue, and the class II enzymes, in which AdoCbl binds to the enzyme in its "base-on" conformation, and the pendent 5,6-dimethylbenzimidazole base remains coordinated to the metal during Co-C bond activation.  相似文献   

2.
1H NMR line broadening is found to be an effective complimentary method to chemical trapping for determining the rates and activation parameters for organo-metal bond homolysis events that produce freely diffusing radicals. Application of this method is illustrated by measurement of bond homolysis activation parameters for a series of organo-cobalt porphyrin complexes ((TPP)Co-C(CH3)2CN (ΔH = 19.5±0.9 kcal mol−1, ΔS = 12±3 cal°K−1 mol−1), (TMP)Co-C(CH3)2CN (ΔH = 20±1 kcal mol−1S = 13±2 cal°K−1 mol−1), (TAP)Co-C(CH3)2CO2CH3H = 18.2±0.5 kcal mol−1, ΔS = 12±2 cal °K−1 mol−1), (TAP)Co-CH(CH3)C6H5H = 22.5±0.5, ΔS = 17±2 cal °K−1 mol−1)). The line broadening method is particularly useful in determining activation parameters for dissociation of weakly bonded organometallics where the rate of homolysis can exceed the range measurable by conventional chemical trapping methods.  相似文献   

3.
Rates of stepwise anation of cis-Cr(ox)2(H2O2) with SCN/N3, Cr(acac)2(H2O)2+ with SCN and Cr(atda)(H2O)2 with SCN have been investigated in weakly acidic aqueous solutions. Rate constants, kI and kII for the two steps in each system, are composite as kx = kx0+kxX[X] (x = I, II; X = SCN, N3). These rate constants have been evaluated also as the corresponding ΔH and ΔS values. The results obtained and the plausible Id mechanism seem to suggest Cr---OOC bond dissociation (hence a strongly negative ΔS) generating the transition state in each system with outer-sphere association forming the precursor complex in the X dependent paths.  相似文献   

4.
Aqueous solutions of adenosylcobalamin (AdoCbl) were thermolyzed with excess beta-mercaptoethanol under anaerobic conditions. The product studies reveal that approximately 90% Co-C bond homolysis occurs, to yield Co(II)cobalamin, 5'-deoxyadenosine, and the disulfide product from the combination of two HOCH2CH2S* radicals, 2,2'-dithiodiethanol; there is also approximately 10% Co-C bond heterolysis, yielding Co(III)cobalamin, adenine, and 2,3-dihydroxy-4-pentenal. The kinetic studies show there is a first-order dependence on AdoCbl and zero-order dependence on thiol under the higher [RSH] experimental conditions used, consistent with the rate-determining step at high [RSH] being the generation of Ado*. The kinetic results require that, in enzyme-free AdoCbl solution, adenosyl radical (Ado*) is formed as a discrete intermediate which then abstracts H* from the added thiol. The activation parameters for Co-C bond homolysis in the presence of thiol trap are the same within experimental error as the activation parameters for Co-C bond homolysis without trap, standard delta H(obs) = 29(2) kcal mol(-1) and standard delta S(obs) = -1(5) e.u. The results, in comparison to the rate of Co-C bond homolysis in ribonucleoside triphosphate reductase (RTPR), reveal that RTPR accelerates Co-C bond cleavage in AdoCbl by approximately 10(10+/-1). The recent literature evidence bearing on the exact mechanism of RTPR enzymic cleavage of the Co-C bond of AdoCbl is briefly discussed, notably the fact that this mechanism is presently controversial, but does involve at least coupled (and possibly concerted) Co-C cleavage, -S-H cleavage, and C-H (Ado-H) formation steps.  相似文献   

5.
The kinetics of the anation reactions of [M(RNH2)5H2O]3+ (M = Rh, R = H, Me, Et, Pr; M = Cr, R = H, Me, Pr) with several ligands (H3PO4/H2PO4, H3PO3/H2PO3, CF3COO, Br, Cl, SCN) have been studied at different temperatures and acidities at I = 1.0 M (LiClO4. Results obtained for the anation rate constants and thermal activation parameters are compared with the previously published data for R = H, in order to establish the effects of the amine substituents in the reaction mechanism proposed for the substitution reactions of these complexes. The results obtained are interpreted on the basis of a mechanism where the bond formation process is more important in the substitution on M = Cr complexes than in that of the M = Rh complexes, as already pointed out for the published ΔΛ values for the water exchange on these systems. A simple Langford-Gray classification becomes inadequate to describe these situations where the increase of the steric demand of the amine substituents shifta the Ia-Id classification to the Id side, although no dramatic changes in the reaction mechanism are found. It is concluded that a More O'Ferall ‘continuous’ type of approach to the mechanism classification of the substitution reactions is much more useful in this case.  相似文献   

6.
Subtle changes in the M?ssbauer parameters are observed while going from methyl- to ethyl- to adenosylcobalamin, and also when the "base" is detached from the cobalt. The observation of these changes demonstrates that the Co-C bond, among others, remains intact after the Auger event, accompanying the electron-capture decay of the cobalt-57. The differences between ethylcobalamin and the other two organocobalamins in the magnitude of the quadrupole splittings have been interpreted on the basis of the sigma-donating tendency of the organic moiety and the Co-C bond length. The latter is presumably determined by the steric hindrance offered to the group in approaching the cobalt atom. The ethyl- and adenosylcobalamins in their "base-off" form exhibit a larger quadrupole splitting than the corresponding "base-on" form. In the "bas-off" form, the cobalt atom is perhaps raised above the mean plane of the four equatorial nitrogen atoms of the corrin ring, which may result in the diminution of the delocalization of the 3dpi electron density. The higher population of dpi orbitals and the enhanced metallic character of the dz2, resulting from shrinkage of the Co-C bond length, enhances the magnitude of the quadrupole splitting.  相似文献   

7.
Li X  Shen L  Zhang D  Qi H  Gao Q  Ma F  Zhang C 《Biosensors & bioelectronics》2008,23(11):1624-1630
A simple and highly sensitive electrochemical impedance spectroscopy (EIS) biosensor based on a thrombin-binding aptamer as molecular recognition element was developed for the determination of thrombin. The signal enhancement was achieved by using gold nanoparticles (GNPs), which was electrodeposited onto a glassy carbon electrode (GCE), as a platform for the immobilization of the thiolated aptamer. In the measurement of thrombin, the change in interfacial electron transfer resistance of the biosensor using a redox couple of [Fe(CN)6]3−/4− as the probe was monitored. The increase of the electron transfer resistance of the biosensor is linear with the concentration of thrombin in the range from 0.12 nM to 30 nM. The association and dissociation rate constants of the immobilized aptamer–thrombin complex were 6.7 × 103 M−1 s−1 and 1.0 × 10−4 s−1, respectively. The association and dissociation constants of three different immobilized aptamers binding with thrombin were measured and the difference of the dissociation constants obtained was discussed. This work demonstrates that GNPs electrodeposited on GCE used as a platform for the immobilization of the thiolated aptamer can improve the sensitivity of an EIS biosensor for the determination of protein. This work also demonstrates that EIS method is an efficient method for the determination of association and dissociation constants on GNPs modified GCE.  相似文献   

8.
Due to contradictions in the literature we have redetermined the acid-base properties of riboflavin (=RiFl; vitamin B2), i.e. 7,8-dimethyl-10-ribityl-isoalloxazine, and of flavin mononucleotide (FMN2−), also known as riboflavin 5′-phosphate, via potentiometric pH titrations (I = 0.1 M, NaNO3; 25 °C). In contrast to various claims, the isoalloxazine ring cannot be protonated at pH > 1, a result in agreement with an early study (pKa = −0.2; L. Michaelis, M.P. Schubert and C.V. Smythe, J. Biol. Chem., 116 (1936) 587–607); deprotonation of the ring system occurs in both compounds with pKa 10. The pKa value of 0.7 determined for the deprotonation of H2(FMN) must be attributed to the release of the first proton from the fully protonated phosphate group; its second proton is released with pKa = 6.18 in agreement with the acidity constants of various other monoprotonated monophosphate esters. The stability constants of the 1:1 complexes formed between Mg2+, Ca2+, Sr2+, Ba2+, Mn2+, Co2+, Ni2+, Cu2+, Zn2+ or Cd2+ (---M2+) and FMN2− were determined by potentiometric pH titrations in aqueous solution (I = 0.1 M, NaNO3; 25 °C). The log stability constants of all these M(FMN) complexes are about 0.2 log units higher than expected from the basicity of the phosphate group. This slight stability increase cannot be attributed to the formation of a seven-membered chelate involving the ribit-hydroxy group at C-4′ as the stability constants for the M2+ 1:1 complexes of glycerol 1-phosphate (G1P2−) demonstrate: G1P2− contains the same structural unit which would also allow in this case the formation of the mentioned seven-membered chelate; however, the stability of the M(G1P) complexes is solely determined by the basicity of the phosphate group. Hence, in agreement with earlier conclusions (J. Bidwell, J. Thomas and J. Stuehr, J. Am. Chem. Soc., 108 (1986) 820–825) regarding Ni(FMN) one must conclude that the slight stability increase of the M(FMN) complexes has to be attributed to the isoalloxazine ring. The equality of the stability increase of the complexes for all the mentioned ten metal ions precludes its attribution to an interaction with an N site and makes a specific interaction with an O site also somewhat unlikely. In addition, carbonyl oxygens appear as not very favorable for the formation of macrochelates by a further interaction with already phosphate-coordinated metal ions. Therefore, we propose that the slight but significant stability increase originates from M(FMN) species (with a formation degree of about 30%) in which the hydrophobic flavin residue is close to the metal ion, thereby lowering the ‘effective’ dielectric constant in the microenvironment of the metal ion and thus indirectly promoting the −PO32−/M2+ interaction.  相似文献   

9.
Carbonylation of the anionic iridium(III) methyl complex, [MeIr(CO)2I3] (1) is an important step in the new iridium-based process for acetic acid manufacture. A model study of the migratory insertion reactions of 1 with P-donor ligands is reported. Complex 1 reacts with phosphites to give neutral acetyl complexes, [Ir(COMe)(CO)I2L2] (L = P(OPh)3 (2), P(OMe)3 (3)). Complex 2 has been isolated and fully characterised from the reaction of Ph4As[MeIr(CO)2I3] with AgBF4 and P(OPh)3; comparison of spectroscopic properties suggests an analogous formulation for 3. IR and 31P NMR spectroscopy indicate initial formation of unstable isomers of 2 which isomerise to the thermodynamic product with trans phosphite ligands. Kinetic measurements for the reactions of 1 with phosphites in CH2Cl2 show first order dependence on [1], only when the reactions are carried out in the presence of excess iodide. The rates exhibit a saturation dependence on [L] and are inhibited by iodide. The reactions are accelerated by addition of alcohols (e.g. 18× enhancement for L = P (OMe)3 in 1:3 MeOH-CH2Cl2). A reaction mechanism is proposed which involves substitution of an iodide ligand by phosphite, prior to migratory CO insertion. The observed rate constants fit well to a rate law derived from this mechanism. Analysis of the kinetic data shows that k1, the rate constant for iodide dissociation, is independent of L, but is increased by a factor of 18 on adding 25% MeOH to CH2Cl2. Activation parameters for the k1 step are ΔH = 71 (±3) kJ mol, ΔS = −81 (±9) J mol−1 K−1 in CH2Cl2 and ΔH = 60(±4) kJ mol−1, ΔS = −93(± 12) J mol−1 K−1 in 1:3 MeOH-CH2Cl2. Solvent assistance of the iodide dissociation step gives the observed rate enhancement in protic solvents. The mechanism is similar to that proposed for the carbonylation of 1.  相似文献   

10.
P.Muir Wood 《BBA》1974,357(3):370-379
The rate of electron transfer between reduced cytochrome ƒ and plastocyanin (both purified from parsley) has been measured as k = 3.6 · 107 M−1 · s−1, at 298 °K and pH 7.0, with activation parameters ΔH = 44 kJ · mole−1 and ΔS = +46 J · mole−1 · °K−1. Replacement of cytochrome ƒ with red algal cytochrome c-553, Pseudomonas cytochrome c-551 and mammalian cytochrome c gave rates at least 30 times slower: k = 5 · 105, 7.5 · 105 and 1.0 · 106 M−1 · s−1, respectively.

Similar measurements made with azurin instead of plastocyanin gave k = 6 · 106 and approx. 2 · 107 M−1 · s−1 for reaction of reduced azurin with cytochrome ƒ and algal cytochrome respectively.

Rate constants of 115 and 80 M−1 · s−1 were found for reduction of plastocyanin by ascorbate and hydroquinone at 298 °K and pH 7.0. The rate constants for the oxidation of plastocyanin, cytochrome ƒ, Pseudomonas cytochrome c-551 and red algal cytochrome c-553 by ferricyanide were found to be between 3 · 104 and 8 · 104 M−1 · s−1.

The results are discussed in relation to photosynthetic electron transport.  相似文献   


11.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

12.
Rotational barriers about the M-S bonds of 16-electron bent metallocene monothiolates (η5-C5H5)2Zr(Cl) (SR) (R = −CH3, −CH2CH3, −CH(CH3)2, −C(CH3)3) (1a–d) have been measured by dynamic 1H NMR methods: 32, 33, 35 and 26 kJ mol−1, respectively. The ground-state orientation about the Zr-S bonds of 1 that maximizes Spπ → Mdπ bonding (Cl-Zr-S-R ≈ 90°) also maximizes CpR steric interaction, whereas the rotational transition-state orientation (Cl-Zr-S-R ≈ 0°) is one that minimizes Spπ → Mdπ bonding and maximizes ClR steric interaction. Deviation from a ground-state orientation that is ideal for Spπ → Mdπ bonding might be expected as the size of the R group and CpR steric interaction increases. Thus, the aberrant trend for the R = −C(CH3)3 derivative could be attributed to a ground-state steric effect where the sterically demanding −C(CH3)3 group forces an unfavorable (misdirected) orientation for Mdπ-Spπ bonding, but a favorable orientation with respect to CpR and ClR steric interactions. However, the solid-state structures of (η5-C5H5)2Zr(SR)2 (R = −CH3, −CH2CH3, −CH(CH3)2, −C(CH3)3) (2a–d) exhibit regular variation of their metric parameters as evidenced by their Zr-S-C bond angles of 108, 109, 113, and 124° and S-Zr-S′ bond angles of 97, 99, 100 and 106°, respectively. Neither the S′-Zr-S-R torsion angles nor the dihedral angles that describe the relationship between the S/Zr/S′ and Cp(centroid)/Zr/Cp′ (centroid) planes (both indicators of the relative orientation of the Zr dπ acceptor orbital and the thiolate S pπ donor orbital) reflect the steric demand of the R group. Thus, the size of the R group imposes a measured effect on the geometry of 2 and the tert-butyl group is not extraordinary. Although the enthalpic and entropic effects could not be deconvoluted for rotation about the Zr-S bond of 1 in the present study, literature precedents suggest that both enthalpic and entropic effects may play a role in determining the irregular trend that is observed.  相似文献   

13.
1. Rate constants for reduction of paraquat ion (1,1′-dimethyl-4,4′-bipyridy-lium, PQ2+) to paraquat radical (PQ+·) by eaq and CO2· have been measured by pulse radiolysis. Reduction by eaq is diffusion controlled (k = 8.4·1010 M−1·s−1) and reduction by CO2· is also very fast k = 1.5·1010 M−1·s−1).

2. The reaction of paraquat radical with oxygen has been analysed to give rate constants of 7.7·108 M−1·s−1 and 6.5·108 M−1·s−1 for the reactions of paraquat radical with O2 and O2·, respectively. The similarity in these rate constants is in marked contrast to the difference in redox potentials of O2 and O2· (− 0.59 V and + 1.12 V, respectively).

3. These rate constants, together with that for the self-reaction of O2·, have been used to calculate the steady-state concentration of O2· under conditions thought to apply at the site of reduction of paraquat in the plant cell. On the basis of these calculations the decay of O2· appears to be governed almost entirely by its self-reaction, and the concentration 5 μm away from the thylakoid is still 90% of that at the thylakoid itself. Thus, O2· persists long enough to diffuse as far as the chloroplast envelope and tonoplast, which are the first structures to be damaged by paraquat treatment. O2· is therefore sufficiently long-lived to be a candidate for the phytotoxic product formed by paraquat in plants.  相似文献   


14.
The kinetics and equilibria of complex formation by Ga(III) with NCS in aqueous solution have been measured over a range of acidities and temperatures, the contributing paths to the reaction resolved, and their rate constants and activation parameters determined. The hydrolysis equilibria required to carry out this resolution of kinetic behaviour have also been measured.

Unlike the other reported complexation reactions of Ga(III) in aqueous solution, the separate reaction pathways can be assigned with no ambiguity. At 25 °C and ionic strength 0.5 M, the observed forward rate constant for the complex formation is described by {k1 + k2K1h/[H+] + k3K1hK2h/[H+]2} M−1 s−1. For these conditions, the first and second successive hydrolysis constants of Ga(H2O)63+ are given by pK1h = 3.69 ± 0.01 and pK2h = 3.74 ± 0.04. The rate constants corresponding to the reactions of the species Ga(H2O)63+, Ga(H2O)5(OH)2+ and Ga(H2O)4(OH)2+ with NCS are k1 = 57 ± 4 M−1 −1, k2 = (1.08 ± 0.01) × 105 M−1 s−1 and k3 = 3 × 106 M−1 s−1 respectively. The complexation equilibrium quotient [GaNCS2+]/([Ga3+][NCS]) has been independently determined by spectrophotometric titration to be 20.8 ± 0.3 M−1 at 25 °C and ionic strength 0.5 M.

These kinetic results lead to an interpretation of the data, and a reinterpretation of other data for aquo-Ga(III) complex formation kinetics from the literature which support the assignment of a dissociative interchange mechanism for these reactions rather than the associative activation mode sometimes proposed.  相似文献   


15.
Rate constants determined by the stopped-flow method for four protein-protein reactions at 25°C, pH's in the range 5.8–7.5. I = 0.10 M (NaCI), are as follows: cytochrome c(II) with plastocyanin, PCu(II). 1.5 × 106 M−1 sec−1, pH 7.6; high-potential iron-sulfur protein (Hipip) with PCu(II), 3.7 × 105 M−1sec−1. pH 5.8; cytochrome c(II) with azurin, ACu(ll). 6.4 × 103 M−1sec−1, pH 6.1; Hipip with ACu(II), 2.2 × 105 M−1sec−1, pH 5.8. Activation parameters have been determined for all four reactions; they indicate higher enthalpy requirements and less negative entropy requirements for the PCu(II) as opposed to ACu(II) reactions. Equilibrium constants K for association prior to electron transfer are < 150 M−1 for the cytochrome c(II) reduction of PCu(II) (estimated charges 8 + and 9-,respectively), and < 300 M−1 for the other reactions, indicating no favorable interactions. Rate constants have been analyzed in terms of the simple Marcus theory, which has previously given an excellent fit to thirteen protein-protein reactions considered by Wherland and Pecht. No similar correlation exists in the present studies, and calculated rate constants differ by orders of magnitude from experimentally determined values.  相似文献   

16.
[Fe(TIM)(CH3CN)2](PF6)2 (1) (TIM = 2,3,9,10-tetramethyl-1,4,8,11-tetraazacyclodeca-1,3,8,10-tetraene) forms a complex with NO reversibly in CH3CN (53±1% converted to the NO complex) or 60% CH3OH/40% CH3CN (81±1% conversion). Quantitative NO complexation occurs in H2O or CH3OH solvents. The EPR spectrum of [Fe(TIM)(solvent)NO]2+ in frozen 60/40 CH3OH/CH3CN at 77 K shows a three line feature at g=2.01, 1.99 and 1.97 of an S=1/2FeNO7 ground state. The middle line exhibits a three-line N-shf coupling of 24 G indicating a six-coordinate complex with either CH3OH or CH3CN as a ligand trans to NO. In H2O [Fe(TIM)(H2O)2]2+ undergoes a slow decomposition, liberating 2,3-butanedione, as detected by 1H NMR in D2O, unless a π-acceptor axial ligand, L=CO, CH3CN or NO is present. An equilibrium of 1 in water containing CH3CN forms [Fe(TIM)(CH3CN)(H2O)]2+ which has a formation constant KCH3CN=320 M−1. In water KNOKCH3CN since NO completely displaces CH3CN. [Fe(TIM)(CH3CN)2]2+ binds either CO or NO in CH3CN with KNO/KCO=0.46, sigificantly lower than the ratio for [FeII(hemes)] of 1100 in various media. A steric influence due to bumping of β-CH2 protons of the TIM macrocycle with a bent S=1/2 nitrosyl as opposed to much lessened steric factors for the linear Fe---CO unit is proposed to explain the lower KNO/KCO ratio for the [Fe(TIM)(CH3CN)]2+ adducts of NO or CO. Estimates for formation constants with [Fe(TIM)]2+ in CH3CN of KNO=80.1 M−1 and KCO=173 M are much lower than to hemoglobin (where KNO=2.5×1010 M−1 and KCO=2.3×107) due to a reversal of steric factors and stronger π-backdonation from [FeII(heme)] than from [FeII(TIM)(CH3CN)]2+.  相似文献   

17.
The reactions of complex (C5Me5)Ir(Cl) (CO) (Me) (1a) with cyclohexylisocyanide and phosphines (L=CyNC, PHPh2, PMePh2, PMe2Ph) give the products of alkyl migratory insertion (C5Me5Ir(Cl) (COMe) (L), in toluence or tetrahydrofuran at 323 K or higher temperature. The phenyl analogue (C5Me5)Ir(Cl)(CO)(Ph) or the iodide complexes (C5Me5)Ir(I) (CO) (R) (R=Me, Ph_are not reactive under the same conditions. The reaction of (C5Me5)Ir(Cl)(CO)(Me) with PMePh2 and PMe2Ph in acetonitrile yields the chloride substitution product [(C5Me5)Ir(CO)(L)(Me)]+Cl. Kinetic measurements for the reactions of (C5Me5)Ir(Cl)(CO)(Me) in toluene are first order in the iridium complex and exhibit a saturation dependence on the incoming donors L. Analysis of the data suggests a two-step process involving (i) rapid formation of a molecular complex [(C5Me5)Ir(Cl)(CO)(Me), (L)], in which the structure of 1a is unperturbed within the limits of spectroscopic analysis, and (ii) rate determining methyl migration. The reaction parameters are K for the pre-equilibrium step (K = 1.5 (CyNC), 7.3 (PHPh2), 7.1 (PMePh2) dm3 mol−1 at 323 K) and k2 for the slow carbon---carbon bond formation (k2 (105) = 6.9 (CyNC), 1.2 (PHPh2), 1.0 (PMePh2) s−1 at 323 K). The activation parameters for the methyl migration step in the reaction with PMePh2 obtained between 308 and 338 K, are ΔH = 106±16 kJ mol−1 and ΔS = − 14±5 J K−1 mol−1. The reaction of 1a with PMePh2 proceeds at similar rates in tetrahydrofuran (K = 3.7 dm3 mol−1, k2 (105) = 1.2 s−1, 323 K). The crystal structure of (C5Me5)Ir(Cl)(COMe) (PMe2Ph) has been determined by X-ray diffraction. C20H29ClOPIr: Mr = 544.1, monoclinic, P21/n, A = 8.084 (2), B = 9.030(2), C = 28.715 (3) Å, β = 91.41 (3)°, Z = 4, Dc = 1.71 g cm−3, V = 2095.5 Å3, room temperatyre, Mo K, γ = 0.71069, μ = 65.55 cm−1, F(000) = 1044, R = 0.037 for 2453 independent observed reflections. The complex shows a deformed tetrahedral coordination assuming the η5-C5Me5 molecular fragment as a single coordination site. The iridium-chlorine bond is staggered with respect to two adjacent C(ring)-methyl bonds, while the Ir---P and the Ir---COMe bonds are eclipsed with respect to C(ring)-methyl bonds.  相似文献   

18.
The square-planar bis chelate complexes Ni(R-sal)2 (= bis(N-alkyl)salicylaldiminato)nickel(II)) with R = (CH2)2Ph (I; Ph = phenyl), (CH2)3Ph (II), (CH2)4Ph (III) and (CH2)2(4-hydroxyphenyl) (IV) were prepared and characterized. ComplexesII and III meet the steric requirements for intramolecular aromatic ring stacking. Stopped-flow spectrophotometry was used to study the kinetics of ligand substitution in complexesI–IV by H2salen (=N,N′-disalicylidene-ethylenediamine) in acetone. For the substitution of the two bidentate ligands in Ni(R-sal)2 only one step is kinetically observed which follows a second-order rate law, rate =k[H2salen] [Ni(R-sal)2], with k = 43.4 (I), 64.0 (II), 87.0 (III) and 49.5 (IV) M−1 s−1 at 298 K. It is found, therefore, that the size of k does not change significantly upon lengthening of the alkane chain in Ni(Ph(CH2)nsal)2 from n = 2 to 4 and that there is no kinetic evidence for intramolecular stacking interactions. The equilibrium constants and thermodynamic parameters for the formation of the bis adductsIII·(py)2 and III·(MeOH)2 in acetone are reported.  相似文献   

19.
Single crystal X-ray diffraction studies of trans-[(Ph3P)2Pd(Ph)X] (X = F (1), Cl (2), Br (3), and I (4) were carried out. The four structures split in two isostructural and isomorphous groups, namely orthorhombic for 1 and 2 (space group Pbca, Z = 8) and triclinic for 3 and 4 (space group P-1, Z = 2). According to the Pd---C bond length, the trans influence of X within these pairs follows the trend Cl>F and 1>Br. However, the trans influence of Cl is slightly stronger than that of Br. Both structural and 13C NMR studies revealed that electron-donating effects of (Ph3P)2PdX increase along the series X=I− for the Pd centre in [(Ph3P)2Pd(Ph)] were studied by 31P NMR in rigorously anhydrous CH2Cl2 solutions, and equilibrium constants and ΔG values were obtained for all possible combinations. The sequence F > Cl > Br > I is characteristic of halide preference for the Pd complexes. Dissolving 1 and PPN Cl in dry CH2Cl2 resulted in the release of ‘naked’ F which fluorinated the solvent smoothly to give a mixture of CH2ClF and CH2F2 in high yield. When chloroform was used instead of CH2Cl2, dichlorocarbene was generated slowly, forming the corresponding cyclopropane in the presence of styrene. All observations were rationalized successfully in terms of the filled/filled effect and push/pull interactions.  相似文献   

20.
The complex Pt(bph) (CO)2 crystallizes in the space group Cmcm with a = 18.647(6), B = 9.566(2) and C = 6.4060(5) Å. The geometry of the molecule is slightly distorted from square planar with a Pt---C(CO) bond distance of 1.98(2) Å and a Pt---C(bph) bond distance of 2.04(2) Å. The Pt(bph)(CO)2 complex serves as a precursor for the preparation of a wide variety of Pt(bph)X2 complexes, where X = monodentate ligands such as acetonitrile, pyridine, etc., and X2 = bidentate ligands such as bypyridine, 1,10-phenanthroline, etc. In the solid state, the complex exhibits a green color, but when ground with an alkali metal salt turns deep blue to purple. In CH2Cl2, the color disappears but optical transitions are observed at 271 nm (2.7 × 104 M−1 cm−1), 303 nm (1.1 × 104 M−1 cm−1) and 330 nm (5.5 × 103 M−1 cm−1). The complex is a weak emitter exhibiting a structured spectrum in CH2Cl2 at r.t. with maxima located at 562 and 594 nm and an emission lifetime of 3.1 μs when excited at 337 nm.  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