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 15 毫秒
1.
This article tries to rationalize the shortcomings of various model compounds and discusses requirements that a low-molecular compound must fulfill in order to become a potentially competitive catalyst for nitrogenases. For fundamental reasons, such a synthetic catalyst cannot be a precise structural duplicate of the active centers of nitrogenase. Results obtained with iron-sulfur carbonyl and diazene complexes further indicate that (1) the coupling and chronology of proton and electron transfer steps, (2) invariance of iron-sulfur distances within a wide range of electron density changes at the iron centers, and (3) Brönsted basic thiolate donors favoring the protonation of metal-sulfur cores and the formation of N–H···S bridges may be essential in order to reduce N2 via N2H2 and N2H4 to NH3 under mild conditions.  相似文献   

2.
A modified form of the Debye-Marcus equation relating electron transfer rate constants to charges on proteins and distances of electron transfer has been applied to the reaction of chemically modified cytochrome f, in which positively charged amino groups are replaced with negatively charged carboxyl groups. The rate of electron transfer from reduced cytochrome f to ferricyanide decreased with increasing ionic strength when the native and singly substituted cytochrome f were used, although a sharp decrease was observed in the former case. When doubly or more than triply substituted cytochrome f was used, the rate of electron transfer was almost constant or increased with increasing ionic strength, respectively. The kinetic-ionic strength effects on this reaction can be well explained by the Debye-Marcus equation in which the charge and radius of the protein are treated as variable parameters. The results show the importance of local positive charges of about 2.0 on native cytochrome f and effective radius of about 11 A of cytochrome f for the electron transfer to ferricyanide. Since the net charge on the native cytochrome f is negative and the calculated radius of the protein is 22.8 A, the above results indicate that positive charges on the electron transfer site control the electrostatic interactions in this reaction. Previously reported data which had been analyzed by using the total net charge and full radius of the protein, were also well explained by the local charge and effective radius of the protein.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Kinetic and thermodynamic studies involving the application of different high-pressure techniques, are very useful in gaining mechanistic information on the basis of volume changes that occur during inorganic and bioinorganic electron transfer reactions. The most fundamental type of electron transfer reaction concerns self-exchange reactions, for which the overall reaction volume is zero, and activation volumes can be measured and discussed. In the case of non-symmetrical electron transfer reactions, intra- and intermolecular processes can be studied and volume profiles can be constructed. Precursor complex formation can in some cases be recognized kinetically in such systems. Typical values of activation and reaction volumes are reviewed for various reversible and irreversible electron transfer reactions. Mechanistic conclusions reached on the basis of these parameters are presented. Volume profiles for electron transfer reactions enable a simplistic presentation of the reaction mechanism on the basis of intrinsic and solvational volume changes along the reaction coordinate.  相似文献   

4.
Use of rigorous equilibration kinetics to evaluate rate constants for the Fe(CN)6 4- reduction of horse-heart cytochrome c in the oxidized form, cyt c (III), has shown that limiting kinetics do not apply with concentrations of Fe(CN)6 4- (the reactant in excess) in the range 2-10 x 10(-4) M, I = 0.10 M (NaCl). The reaction conforms to a first-order rate law in each reactant, and at 25 degrees C, pH 7.2 (Tris), it is concluded that K for association prior to electron transfer is less than 200 M-1. From previous studies at 25 degrees C, ph 7.0 (10(-1) M phosphate), I = 0.242 M (NaCl), a value K = 2.4 x 10(3) M-1 has been reported. Had such a value applied, some or all of the redox inactive complexes Mo(CN)8 4-, Co(CN)6 3-, Cr(CN)6 3-, Zr(C2O4)4 4- present in amounts 5-20 x 10(-4) M would have been expected to associate at the same site and partially block the redox process. No effect on rats was observed. With the reductants Fe(CN)5(4-NH2-py)3- and Fe(CN)5(imid)3-, reactions proceeded to greater than 90% completion and rate laws were again first order in each reactant. Rate constants (M-1 sec-1) at 25 degrees C, pH 7.2 (Tris), I = 0.10 M (NaCl), are Fe(CN)6 4- (3.5 x 10(4)), Fe(CN)5(4-NH2py)3- (6.7 x 10(5), and Fe(CN)5(imid)3- (4.2 x 10(5). Related reactions in which cyt c(II) is oxidized are also first order in each reactant, Fe(CN)6 3- (9.1 x 10(6)), Fe(CN)5(NCS)3- (1.3 x 10(6)), Fe(CN)5(4-NH2py)2- (3.8 x 10(6) at pH 9.4), and Fe(CN)5(NH3)2- (2.75 x 10(6) at ph 8). Redox inactive Co(CN)6 3- (1.0 x 10(-3) M) has no effect on the reaction of Fe(CN)6 3- which suggests that a recent interpretation for the Fe(CN)6 3- oxidation of cyt c(II), I = 0.07 M, may also require reappraisal.  相似文献   

5.
The kinetics of methemoglobin reduction by Fe(EDTA)2? have been studied and found to follow a second order rate law with k = 29.0 M?1 s?1 [25°C, μ = 0.2 M, pH 7.0 (phosphate)], ΔH3 = 5.5 ± 0.7 kcal/mol, and ΔS2= ?33 ± 2 e.u.. The electrostatics-corrected self-exchange rate constant (k11corr) for hemoglobin based on the Fe(EDTA)2? cross-reaction is 2.79×10?3M?1 s?1. This rate constant is compared with others reported for a water-soluble iron porphyrin and calculated from published data for the reactions of myoglobin and hemoglobin with Fe(EDTA)2? and Fe(CDTA)2?/?. The k11corr values for these systems range over ten orders of magnitude with heme ? myoglobin > hemoglobin.  相似文献   

6.
Rapid reactions occur between [OsVI(tpy)(Cl)2(N)]X (X = PF6, Cl, tpy = 2,2′:6′,2″-terpyridine) and aryl or alkyl phosphi nes (PPh3, PPh2Me, PPhMe2, PMe3 and PEt3) in CH2Cl2 or CH3CN to give [OsIV(tpy)(Cl)2(NPPh3)]+ and its analogs. The reaction between trans-[OsVI(tpy)(Cl)2(N)]+ and PPh3 in CH3CN occurs with a 1:1 stoichiometry and a rate law first order in both PPh3 and OsVI with k(CH3CN, 25°C) = 1.36 ± 0.08 × 104 M s−1. The products are best formulated as paramagnetic d4 phosphoraniminato complexes of OsIV based on a room temperature magnetic moment of 1.8 μB for trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6), contact shifted 1H NMR spectra and UV-Vis and near-IR spectra. In the crystal structures of trans-[OsIV(tpy)(Cl)2( NPPh3)](PF6)·CH3CN (monoclinic, P21/n with a = 13.384(5) Å, b = 15.222(7) Å, c = 17.717(6) Å, β = 103.10(3)°, V = 3516(2) Å3, Z = 4, Rw = 3.40, Rw = 3.50) and cis-[OsIV(tpy)(Cl)2(NPPh2Me)]-(PF6)·CH3CN (monoclinic, P21/c, with a = 10.6348(2) Å, b = 15.146(9) ÅA, c = 20.876(6) Å, β = 97.47(1)°, V = 3334(2) Å3, Z = 4, R = 4.00, Rw = 4.90), the long Os-N(P) bond lengths (2.093(5) and 2.061(6) Å), acute Os-N-P angles (132.4(3) and 132.2(4)°), and absence of a significant structural trans effect rule out significant Os-N multiple bonding. From cyclic voltammetric measurements, chemically reversible OsV/IV and OsIV/III couples occur for trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6) in CH3CN at +0.92 V (OsV/IV) and −0.27 V (OsIV/III) versus SSCE. Chemical or electrochemical reduction of trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6) gives isolable trans-OsIII(tpy)(Cl)2(NPPh3). One-electron oxidation to OsV followed by intermolecular disproportionation and PPh3 group transfer gives [OsVI(tpy)Cl2(N)]+, [OSIII(tpy)(Cl)2(CH3CN)]+ and [Ph3=N=PPh3]+ (PPN+). trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6) undergoes reaction with a second phosphine under reflux to give PPN+ derivatives and OsII(tpy)(Cl)2(CH3CN) in CH3CN or OsII(tpy)(Cl)2(PR3) in CH2Cl2. This demonstrates that the OsVI nitrido complex can undergo a net four-electron change by a combination of atom and group transfers.  相似文献   

7.
8.
9.
The reaction of nitrones with transition metal nitrile complexes of the type [MCl2(PhCN)2] and [MCl4(MeCN)2] leads to different products, depending on the metal. Reaction of [PdCl2(PhCN)2] with RCHN(Me)O (R=Ph, p-C6H4Me) afford Δ4-1,2,4-oxadiazoline complexes as products of [2 + 3] cycloaddition of the nitrone across the C≡N bond of the nitrile. The oxophilic transition metal compounds [TiCl4(MeCN)2] and [ZrCl4(MeCN)2] undergo rapid ligand exchange to form nitrone complexes from which the nitrone can be released without decomposition. The closely related compounds [MoCl4(MeCN)2] and [WCl4(MeCN)2] mediate the hydrolysis of nitrones to the corresponding aldehydes.  相似文献   

10.
In human heme peroxidases the prosthetic group is covalently attached to the protein via two ester linkages between conserved glutamate and aspartate residues and modified methyl groups on pyrrole rings A and C. Here, monomeric recombinant myeloperoxidase (MPO) and the variants D94V and D94N were produced in Chinese hamster ovary cell lines. Disruption of the Asp(94) to heme ester bond decreased the one-electron reduction potential E'(0) [Fe(III)/Fe(II)] from 1 to -55 mV at pH 7.0 and 25 degrees C, whereas the kinetics of binding of low spin ligands and of compound I formation was unaffected. By contrast, in both variants rates of compound I reduction by chloride and bromide (but not iodide and thiocyanate) were substantially decreased compared with the wild-type protein. Bimolecular rates of compound II (but not compound I) reduction by ascorbate and tyrosine were slightly diminished in D94V and D94N. The presented biochemical and biophysical data suggest that the Asp(94) to heme linkage is no precondition for the autocatalytic formation of the other two covalent links found in MPO. The findings are discussed with respect to the known active site structure of MPO and its complexes with ligands.  相似文献   

11.
A family of 12 different mixed ligand complexes of iron with cyanide and substituted 1,10-phenanthroline was prepared. The electron transfer properties of each reagent were systematically manipulated by varying the substituent(s) on the aromatic ring system and the stoichiometry of the two types of ligands in the complex. Values for the standard reduction potentials of each member of this family of electron transfer reagents were determined and spanned from 500 to 900 mV. The one-electron transfer reactions between each of these substitution-inert reagents and the high potential blue copper protein, rusticyanin, from Thiobacillus ferrooxidans were studied by stopped flow spectrophotometry under acidic conditions. For comparison with the protein results, the kinetics of electron transfer between each of these reagents and sulfatoiron were also investigated. The Marcus theory of electron transfer was successfully applied to this set of kinetic data to demonstrate that 10 of the 12 reagents had equal kinetic access to the redox center of the rusticyanin and utilized the same reaction pathway for electron transfer. The utility of these synthetic electron transfer reagents in characterizing the electron transfer properties of very high potential, redox-active metalloproteins is illustrated.  相似文献   

12.
13.
The reduction kinetics of the photo-oxidized primary electron donor P700 in photosystem I (PS I) complexes from cyanobacteria Synechocystis sp. PCC 6803 were analyzed within the kinetic model, which considers electron transfer (ET) reactions between P700, secondary quinone acceptor A1, iron-sulfur clusters and external electron donor and acceptors – methylviologen (MV), 2,3-dichloro-naphthoquinone (Cl2NQ) and oxygen. PS I complexes containing various quinones in the A1-binding site (phylloquinone PhQ, plastoquinone-9 PQ and Cl2NQ) as well as F X-core complexes, depleted of terminal iron–sulfur F A/F B clusters, were studied. The acceleration of charge recombination in F X-core complexes by PhQ/PQ substitution indicates that backward ET from the iron–sulfur clusters involves quinone in the A1-binding site. The kinetic parameters of ET reactions were obtained by global fitting of the P700 + reduction with the kinetic model. The free energy gap ΔG 0 between F X and F A/F B clusters was estimated as ?130 meV. The driving force of ET from A1 to F X was determined as ?50 and ?220 meV for PhQ in the A and B cofactor branches, respectively. For PQ in A1A-site, this reaction was found to be endergonic (ΔG 0?=?+75 meV). The interaction of PS I with external acceptors was quantitatively described in terms of Michaelis–Menten kinetics. The second-order rate constants of ET from F A/F B, F X and Cl2NQ in the A1-site of PS I to external acceptors were estimated. The side production of superoxide radical in the A1-site by oxygen reduction via the Mehler reaction might comprise ≥0.3% of the total electron flow in PS I.  相似文献   

14.
Multicopper oxidases (MCOs) are unique among copper proteins in that they contain at least one each of the three types of biologic copper sites, type 1, type 2, and the binuclear type 3. MCOs are descended from the family of small blue copper proteins (cupredoxins) that likely arose as a complement to the heme-iron-based cytochromes involved in electron transport; this event corresponded to the aerobiosis of the biosphere that resulted in the conversion of Fe(II) to Fe(III) as the predominant redox state of this essential metal and the solubilization of copper from Cu2S to Cu(H2O) n 2+. MCOs are encoded in genomes in all three kingdoms and play essential roles in the physiology of essentially all aerobes. With four redox-active copper centers, MCOs share with terminal copper-heme oxidases the ability to catalyze the four-electron reduction of O2 to two molecules of water. The electron transfers associated with this reaction are both outer and inner sphere in nature and their mechanisms have been fairly well established. A subset of MCO proteins exhibit specificity for Fe2+, Cu+, and/or Mn2+ as reducing substrates and have been designated as metallooxidases. These enzymes, in particular the ferroxidases found in all fungi and metazoans, play critical roles in the metal metabolism of the expressing organism.  相似文献   

15.
16.
The energy-converting redox enzymes perform productive reactions efficiently despite the involvement of high energy intermediates in their catalytic cycles. This is achieved by kinetic control: with forward reactions being faster than competing, energy-wasteful reactions. This requires appropriate cofactor spacing, driving forces and reorganizational energies. These features evolved in ancestral enzymes in a low O(2) environment. When O(2) appeared, energy-converting enzymes had to deal with its troublesome chemistry. Various protective mechanisms duly evolved that are not directly related to the enzymes' principal redox roles. These protective mechanisms involve fine-tuning of reduction potentials, switching of pathways and the use of short circuits, back-reactions and side-paths, all of which compromise efficiency. This energetic loss is worth it since it minimises damage from reactive derivatives of O(2) and thus gives the organism a better chance of survival. We examine photosynthetic reaction centres, bc(1) and b(6)f complexes from this view point. In particular, the evolution of the heterodimeric PSI from its homodimeric ancestors is explained as providing a protective back-reaction pathway. This "sacrifice-of-efficiency-for-protection" concept should be generally applicable to bioenergetic enzymes in aerobic environments.  相似文献   

17.
The coordination chemistry of thioether functionalized cyclodiphosphazane ligand, cis-{tBuNP(OCH2CH2SCH3)}2 (1) is described. The reactions of 1 with [Pd (COD)Cl2] in 1:1, 1:2 and 2:1 M ratios afforded cis-[PdCl2{tBuNP(OCH2CH2SCH3)}2] (2), cis-[{PdCl2}2{tBuNP(OCH2CH2SCH3)}2] (3) and trans-[PdCl2{(tBuNP(OCH2CH2SCH3))2}2] (4), respectively. Treatment of 1 with [Pd(PEt3)Cl2]2 or [PdCl(η3-C3H5)]2 in appropriate molar ratios produce the mono- and binuclear complexes [PdCl2(PEt3{tBuNP(OCH2CH2SCH3)}2] (5) and [{PdCl(η3-C3H5)}2{tBuNP(OCH2CH2SCH3)}2] (6) in good yield. The reaction of 1 with [{Ru(p-cymene)Cl2}2] afforded the mononuclear cationic complex, [{(p-cymene)RuCl{tBuNP(OCH2CH2SCH3)}2]Cl (7), whereas the reactions of [Rh(COD)Cl]2, [Pt(COD)Cl2] and [Au(SMe2)Cl] with 1 yielded the corresponding P-coordinated neutral complexes, [RhCl(COD){tBuNP(OCH2CH2SCH3)}2] (8)cis-[PtCl2{tBuNP(OCH2CH2SCH3)}2] (9), respectively. The binuclear palladium(II) complex 3 was found to be an effective catalyst for the Suzuki-Miyaura cross-coupling reactions.  相似文献   

18.
Polarized absorption microspectrophotometry has been used to detect catalysis and intermolecular electron transfer in single crystals of two multiprotein complexes: (1) the binary complex between Paracoccus denitrificans methylamine dehydrogenase, which contains tryptophan-tryptophylquinone (TTQ) as a cofactor, and its redox partner, the blue copper protein amicyanin; (2) the ternary complex between the same two proteins and cytochrome c-551i. Continuous wave electron paramagnetic resonance has been used to compare the state of copper in polycrystalline powders of the two systems. While catalysis and intermolecular electron transfer from reduced TTQ to copper are too fast to be accessible to our measurements, heme reduction occurs over a period of several minutes. The observed rate constant is about four orders of magnitude lower than in solution. The analysis of the temperature dependence of this apparent constant provides values for the parameters H(AB), related to electronic coupling between the two centers, and lambda, the reorganizational energy, that are compatible with electron transfer being the rate-determining step. From these parameters and the known distance between copper and heme, it is possible to calculate the parameter beta, which depends on the nature of the intervening medium, obtaining a value typical of electron transfer across a protein matrix. These findings suggest that the ternary complex in solution might achieve a higher efficiency than the rigid crystal structure thanks to an as yet unidentified role of protein dynamics.  相似文献   

19.
Novel calix[4]pyrrole bearing vic-dioxime ligand (LH2) of the general formula, R1R2C2N2O2H2 (where, R1 = C6H5- and R2 = C39H50N5-) has been synthesized by the reaction of anti-chlorophenylglyoxime with 3-aminophenylcalix[4]pyrrole at room temperature. The mononuclear Cu(II), Ni(II) and Co(II)complexes of this vic-dioxime ligand were prepared and their structures were confirmed by elemental analysis, FT-IR, TGA and magnetic susceptibility measurements; the HMBC, DEPT, 1H and 13C NMR spectra of the LH2 ligand were also reported. The electrochemical property of the complexes was investigated in DMSO by cyclic voltammetry at 200 mV s−1 scan rate. The cyclic voltammetric measurements clearly indicated that Co(LH)2·2H2O complex differs from the Ni(LH)2 and Cu(LH)2 complexes upon the exhibition of quasi-reversible one-electron transfer reduction process in the negative region instead of an irreversible process.  相似文献   

20.
The fluorescence quenching of pyranine by benzoquinone in acetonitrile medium was studied using steady‐state and time‐resolved fluorescence techniques. The quenching process was characterized by a Stern–Volmer plot, which displayed a linear aspect. From the linear plot, the bimolecular quenching rate constant was obtained. The obtained rate constants are within diffused controlled limits. The results show that benzoquinone can efficiently quench the fluorescence of pyranine with dynamic quenching rate constants in the order of 1010 M–1 s–1, suggesting that the pyranine can act as a good electron donor for photoinduced electron transfer in artificial photosynthesis and organic solar cells. In addition, the electron injection dynamics of a pyranine/titanium dioxide semiconductor film was also investigated and electron injection from the excited state pyranine into the conduction band of titanium dioxide is suggested. These preliminary results hold promise for the possibility of using pyranine in dye‐sensitized solar cells. Copyright © 2016 John Wiley & Sons, Ltd.  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