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.
Like pH, salt concentration can have a dramatic effect on enzymatic catalysis. Here, a general equation is derived for the quantitative analysis of salt-rate profiles: k(cat)/K(M) = (k(cat)/K(M))(MAX)/[1+([Na+]/K[Na+])(n')], where (k(cat)/K(M))(MAX) is the physical limit of k(cat)/K(M), K(Na+) is the salt concentration at which k(cat)/K(M) = (k(cat)/K(M))(MAX)/2, and -n' is the slope of the linear region in a plot of log(k(cat)/K(M)) versus log [Na+]. The value of n' is of special utility, as it reflects the contribution of Coulombic interactions to the uniform binding of the bound states. This equation was used to analyze salt effects on catalysis by ribonuclease A (RNase A), which is a cationic enzyme that catalyzes the cleavage of an anionic substrate, RNA, with k(cat)/K(M) values that can exceed 10(9) M(-1) s(-1). Lys7, Arg10, and Lys66 comprise enzymic subsites that are remote from the active site. Replacing Lys7, Arg10, and Lys66 with alanine decreases the charge on the enzyme as well as the value of n'. Likewise, decreasing the number of phosphoryl groups in the substrate decreases the value of n'. Replacing Lys41, a key active-site residue, with arginine creates a catalyst that is limited by the chemical conversion of substrate to product. This change increases the value of n', as expected for a catalyst that is more sensitive to changes in the binding of the chemical transition state. Hence, the quantitative analysis of salt-rate profiles can provide valuable insight into the role of Coulombic interactions in enzymatic catalysis.  相似文献   

2.
Deuterium kinetic solvent isotope effects for the human alpha-thrombin-catalyzed hydrolysis of (1) substrates with selected P(1)-P(3) sites, Z-Pro-Arg-7-amido-4-methylcoumarin (7-AMC), N-t-Boc-Val-Pro-Arg-7-AMC, Bz-Phe-Val-Arg-4-nitroanilide (pNA), and H-D-Phe-L-Pip-Arg-pNA, are (DOD)k(cat) = (2.8-3.3) +/- 0.1 and (DOD)(k(cat)/K(m)) = (0.8-2.1) +/- 0.1 and (2) internally fluorescence-quenched substrates (a) (AB)Val-Phe-Pro-Arg-Ser-Phe-Arg-Leu-Lys(DNP)-Asp-OH, an optimal sequence, and (b) (AB)Val-Ser-Pro-Arg-Ser-Phe-Gln-Lys(DNP)-Asp-OH, recognition sequence for factor VIII, are (DOD)k(cat) = 2.2 +/- 0.2 and (DOD)(k(cat)/K(m)) = (0.8-0.9) +/- 0.1, at the pL (L = H, D) maximum, 8.4-9.0, and (25.0-26.0) +/- 0.1 degrees C. The most plausible models fitting the partial isotope effect (proton inventory) data have been selected on the basis of lowest values of the reduced chi squared and consistency of fractionation factors at all substrate concentrations, assuming rate-determining acylation. The data for Z-Pro-Arg-7-AMC are consistent with a single-proton bridge at the transition state phi(TS) = 0.39 +/- 0.05 and components for solvent reorganization phi(S) = 0.8 +/- 0.1 and phi(S) = 1.22 for k(cat) and k(cat)/K(m), respectively. The data for tripeptide amides fit bowl-shaped curves; an example is N-t-Boc-Val-Pro-Arg-7-AMC: phi(TS)(1) = phi(TS)(2) = 0.57 +/- 0.01 and phi(S) = 1 for k(cat) and 1.6 +/- 0.1 for k(cat)/K(m). Proton inventories for the nonapeptide (2b) are linear. The data for k(cat) for H-D-Phe-L-Pip-Arg-pNA and the decapeptide (2a) are most consistent with two identical fractionation factors for catalytic proton bridging, phi(TS)(1) = phi(TS)(2) = 0.68 +/- 0.02 and a large inverse component (phi(S) = 3.1 +/- 0.5) for the latter, indicative of substantial solvent reorganization upon leaving group departure. Proton inventory curves for k(cat)/K(m) for nearly all substrates are dome-shaped with an inverse isotope effect component (phi(S) = 1.2-2.4) originating from solvent reorganization during association of thrombin with substrate. These large contributions from medium effects are in full accord with the conformational adjustments required for the fulfillment of the dual, hemostatic and thrombolytic, functions of thrombin.  相似文献   

3.
Beta-D-Xylosidase/alpha-L-arabinofuranosidase from Selenomonas ruminantium is the most active enzyme known for catalyzing hydrolysis of 1,4-beta-D: -xylooligosaccharides to D-xylose. Temperature dependence for hydrolysis of 4-nitrophenyl-beta-D-xylopyranoside (4NPX), 4-nitrophenyl-alpha-L-arabinofuranoside (4NPA), and 1,4-beta-D-xylobiose (X2) was determined on and off (k (non)) the enzyme at pH 5.3, which lies in the pH-independent region for k (cat) and k (non). Rate enhancements (k (cat)/k (non)) for 4NPX, 4NPA, and X2 are 4.3 x 10(11), 2.4 x 10(9), and 3.7 x 10(12), respectively, at 25 degrees C and increase with decreasing temperature. Relative parameters k (cat) (4NPX)/k (cat) (4NPA), k (cat) (4NPX)/k (cat) (X2), and (k (cat)/K (m))(4NPX)/(k (cat)/K (m))(X2) increase and (k (cat)/K (m))(4NPX)/(k (cat)/K (m))(4NPA), (1/K (m))(4NPX)/(1/K (m))(4NPA), and (1/K (m))(4NPX)/(1/K (m))(X2) decrease with increasing temperature.  相似文献   

4.
The gene encoding a glycoside hydrolase family 43 beta-xylosidase (GbtXyl43A) from the thermophilic bacterium Geobacillus thermoleovorans strain IT-08 was synthesized and cloned with a C-terminal His-tag into a pET29b expression vector. The recombinant gene product termed GbtXyl43A was expressed in Escherichia coli and purified to apparent homogeneity. Michaelis-Menten kinetic parameters were obtained for the artificial substrates p-nitrophenyl-beta-D: -xylopyranose (4NPX) and p-nitrophenyl-alpha-L: -arabinofuranose (4NPA), and it was found that the ratio k (cat)/K (m) 4NPA/k (cat)/K (m) 4NPX was approximately 7, indicting greater catalytic efficiency for 4NP hydrolysis from the arabinofuranose aglycon moiety. Substrate inhibition was observed for the substrates 4-methylumbelliferyl xylopyranoside (muX) and the arabinofuranoside cogener (muA), and the ratio k (cat)/K (m) muA/k (cat)/K (m) muX was approximately 5. The enzyme was competitively inhibited by monosaccharides, with an arabinose K (i) of 6.8 +/- 0.62 mM and xylose K (i) of 76 +/- 8.5 mM. The pH maxima was 5.0, and the enzyme was not thermally stable above 54 degrees C, with a t (1/2) of 35 min at 57.5 degrees C. GbtXyl43A showed a broad substrate specificity for hydrolysis of xylooligosaccharides up to the highest degree of polymerization tested (xylopentaose), and also released xylose from birch and beechwood arabinoxylan.  相似文献   

5.
Chondroitin AC lyase from Flavobacterium heparinum degrades chondroitin sulfate glycosaminoglycans via an elimination mechanism resulting in disaccharides or oligosaccharides with delta4,5-unsaturated uronic acid residues at their nonreducing end. Mechanistic details concerning the ordering of the bond-breaking and -forming steps of this enzymatic reaction are nonexistent, mainly due to the inhomogeneous nature of the polymeric substrates. The creation of a new class of synthetic substrates for this enzyme has allowed the measurement of defined and reproducible k(cat) and K(m) values and has expanded the range of mechanistic studies that can be performed. The primary deuterium kinetic isotope effect upon k(cat)/K(m) for the abstraction of the proton alpha to the carboxylic acid was measured to be 1.67 +/- 0.07, showing that deprotonation occurs in a rate-limiting step. Using substrates with leaving groups of differing reactivity, a flat linear free energy relationship was produced, indicating that the C4-O4 bond is not broken in a rate-determining step. Taken together, these results strongly suggest a stepwise mechanism. Consistent with this was the measurement of a secondary deuterium kinetic isotope effect upon k(cat)/K(m) of 1.01 +/- 0.03 on a 4-[(2)H]-substrate, indicating that no sp2 character is developed at C4 during the rate-limiting step, thereby ruling out a concerted syn-elimination.  相似文献   

6.
Incubation of farnesyl diphosphate (1) with Penicillium roqueforti aristolochene synthase yielded (+)-aristolochene (4), accompanied by minor quantities of the proposed intermediate (S)-(-)germacrene A (2) and the side-product (-)-valencene (5) in a 94:4:2 ratio. By contrast, the closely related aristolochene synthase from Aspergillus terreus cyclized farnesyl diphosphate only to (+)-aristolochene (4). Site-directed mutagenesis of amino acid residues in two highly conserved Mg(2+)-binding domains led in most cases to reductions in both k(cat) and k(cat)/K(m) as well as increases in the proportion of (S)-(-)germacrene A (2), with the E252Q mutant of the P. roqueforti aristolochene synthase producing only (-)-2. The P. roqueforti D115N, N244L, and S248A/E252D mutants were inactive, as was the A. terreus mutant E227Q. The P. roqueforti mutant Y92F displayed a 100-fold reduction in k(cat) that was offset by a 50-fold decrease in K(m), resulting in a relatively minor 2-fold decrease in catalytic efficiency, k(cat)/K(m). The finding that Y92F produced (+)-aristolochene (4) as 81% of the product, accompanied by 7% 5 and 12% 2, rules out Tyr-92 as the active site Lewis acid that is responsible for protonation of the germacrene A intermediate in the formation of aristolochene (4).  相似文献   

7.
We have compared endo- and exo-type protease reactions and characterized the enzymatic reaction mechanisms by determining all kinetic parameters (k(on), k(off), k(cat), K(d) = k(off)/k(on), and K(m) = (k(off) + k(cat))/k(on)) by following the mass change of the formation and the decay of the enzyme-substrate (ES) complex (k(on) and k(off)), and the formation of the product (k(cat)) on a 27 MHz quartz-crystal microbalance in aqueous solutions. The K(m) value was nearly equal to the K(d) value for the endo-type protease (subtilisin and alpha-chymotrypsin); however, in the case of exo-type protease (carboxypeptidase P), the K(m) value was quite different from the K(d) value, due to k(cat) > k(off).  相似文献   

8.
The penM and pntM genes from the pentalenolactone biosynthetic gene clusters of Streptomyces exfoliatus UC5319 and Streptomyces arenae TU?469 were predicted to encode orthologous cytochrome P450s, CYP161C3 and CYP161C2, responsible for the final step in the biosynthesis of the sesquiterpenoid antibiotic pentalenolactone (1). Synthetic genes optimized for expression in Escherichia coli were used to obtain recombinant PenM and PntM, each carrying an N-terminal His(6)-tag. Both proteins showed typical reduced-CO UV maxima at 450 nm, and each bound the predicted substrate, pentalenolactone F (4), with K(D) values of 153 ± 14 and 126 ± 11 μM for PenM and PntM, respectively, as determined by UV shift titrations. PenM and PntM both catalyzed the oxidative rearrangement of 4 to 1 when incubated in the presence of NADPH, spinach ferredoxin, ferredoxin reductase, and O(2). The steady-state kinetic parameters were k(cat) = 10.5 ± 1.7 min(-1) and K(m) = 340 ± 100 μM 4 for PenM and k(cat) = 8.8 ± 0.9 min(-1) and K(m) = 430 ± 100 μM 4 for PntM. The in vivo function of both gene products was confirmed by the finding that the corresponding deletion mutants S. exfoliatus/ΔpenM ZD22 and S. arenae/ΔpntM ZD23 no longer produced pentalenolactone but accumulated the precursor pentalenolactone F. Complementation of each deletion mutant with either penM or pntM restored production of antibiotic 1. Pentalenolactone was also produced by an engineered strain of Streptomyces avermitilis that had been complemented with pntE, pntD, and either pntM or penM, as well as the S. avermitilis electron-transport genes for ferredoxin and ferrodoxin reductase, fdxD and fprD.  相似文献   

9.
Incubation of farnesyl diphosphate (1) with the W308F or W308F/H309F mutants of pentalenene synthase, an enzyme from Streptomyces UC5319, yielded pentalenene (2), accompanied by varying proportions of (+)-germacrene A (7) with relatively minor changes in k(cat) and k(cat)/K(m). By contrast, single H309 mutants gave rise to both (+)-germacrene A (7) and protoilludene (8) in addition to pentalenene (2). Mutation to glutamate of each of the three aspartate residues in the Mg(2+)-binding aspartate-rich domain, (80)DDLFD, resulted in reduction in the k(cat)/K(m) for farnesyl diphosphate and formation of varying proportions of pentalenene and (+)-germacrene A (7). Formation of (+)-germacrene A (7) by the various pentalenene synthase mutants is the result of a derailment of the natural anti-Markovnikov cyclization reaction, and not simply the consequence of trapping of a normally cryptic, carbocationic intermediate. Both the N219A and N219L mutants of pentalenene synthase were completely inactive, while the corresponding N219D mutant had a k(cat)/K(m) which was 3300-fold lower than that of the wild-type synthase, and produced a mixture of pentalenene (2) (91%) and the aberrant cyclization product beta-caryophyllene (9) (9%). Finally, the F77Y mutant had a k(cat)/K(m) which was reduced by 20-fold compared to that of the wild-type synthase.  相似文献   

10.
The gene encoding aspartate ammonia lyase (aspB) from Bacillus sp. YM55-1 has been cloned and overexpressed, and the recombinant enzyme containing a C-terminal His(6) tag has been purified to homogeneity and subjected to kinetic characterization. Kinetic studies have shown that the His(6) tag does not affect AspB activity. The enzyme processes L-aspartic acid, but not D-aspartic acid, with a K(m) of approximately 15 mM and a k(cat) of approximately 40 s(-1). By using this recombinant enzyme in the reverse reaction, a set of four N-substituted aspartic acids were prepared by the Michael addition of hydroxylamine, hydrazine, methoxylamine, and methylamine to fumarate. Both hydroxylamine and hydrazine were found to be excellent substrates for AspB. The k(cat) values are comparable to those observed for the AspB-catalyzed addition of ammonia to fumarate ( approximately 90 s(-1)), whereas the K(m) values are only slightly higher. The products of the enzyme-catalyzed addition of hydrazine, methoxylamine, and methylamine to fumarate were isolated and characterized by NMR spectroscopy and HPLC analysis, which revealed that AspB catalyzes all the additions with excellent enantioselectivity (>97 % ee). Its broad nucleophile specificity and high catalytic activity make AspB an attractive enzyme for the enantioselective synthesis of N-substituted aspartic acids, which are interesting building blocks for peptide and pharmaceutical synthesis as well as for peptidomimetics.  相似文献   

11.
Two enzymes of unknown function from the amidohydrolase superfamily were discovered to catalyze the deamination of N-6-methyladenine to hypoxanthine and methyl amine. The methylation of adenine in bacterial DNA is a common modification for the protection of host DNA against restriction endonucleases. The enzyme from Bacillus halodurans, Bh0637, catalyzes the deamination of N-6-methyladenine with a k(cat) of 185 s(-1) and a k(cat)/K(m) of 2.5 × 10(6) M(-1) s(-1). Bh0637 catalyzes the deamination of N-6-methyladenine 2 orders of magnitude faster than adenine. A comparative model of Bh0637 was computed using the three-dimensional structure of Atu4426 (PDB code: 3NQB) as a structural template and computational docking was used to rationalize the preferential utilization of N-6-methyladenine over adenine. This is the first identification of an N-6-methyladenine deaminase (6-MAD).  相似文献   

12.
A compartmentalized tyramide labeling system (CoaTi) employing flow cytometry for sorting of yeast cells was developed as ultrahigh-throughput screening for Glucose oxidase (GOx) from Aspergillus niger. CoaTi combines in vitro compartmentalization technology with the CARD reporter system which uses fluorescein tyramide labels for detection of peroxidase activity. Physical connection between cells and fluorescein tyramide radicals was achieved by compartmentalization of yeast cells inside microdroplets of single water-in-oil emulsions. After reaction cells were recovered from single emulsions and sorted by flow cytometry, an error prone PCR mutant library of Glucose oxidase (GOx) containing 10(7) cells and ~10(5) of different GOx variants was screened. Mutagenic conditions of GOx mutant library were selected to generate <1 % of active GOx population in order to explore influence of high mutation frequency on GOx activity. GOx variant Mut12 that contains 5 mutations (N2Y, K13E, T30V, I94V, K152R) showed a 1.2 times decreased K(m) (22.0 vs 18.1 mM) and a 2.7 fold increased k(cat) (150 s(-1) vs 54.8 s(-1)) compared to wt GOx. Compared to the employed parent B11 GOx (16 mM, 80 s(-1)) it has a slightly increased K(m) and 1.8 times increased k(cat).  相似文献   

13.
The isochorismate-pyruvate lyase from Pseudomonas aeruginosa (PchB) catalyzes two pericyclic reactions, demonstrating the eponymous activity and also chorismate mutase activity. The thermodynamic parameters for these enzyme-catalyzed activities, as well as the uncatalyzed isochorismate decomposition, are reported from temperature dependence of k(cat) and k(uncat) data. The entropic effects do not contribute to enzyme catalysis as expected from previously reported chorismate mutase data. Indeed, an entropic penalty for the enzyme-catalyzed mutase reaction (ΔS(++) = -12.1 ± 0.6 cal/(mol K)) is comparable to that of the previously reported uncatalyzed reaction, whereas that of the enzyme-catalyzed lyase reaction (ΔS(++) = -24.3 ± 0.2 cal/(mol K)) is larger than that of the uncatalyzed lyase reaction (-15.77 ± 0.02 cal/(mol K)) documented here. With the assumption that chemistry is rate-limiting, we propose that a reactive substrate conformation is formed upon loop closure of the active site and that ordering of the loop contributes to the entropic penalty for converting the enzyme substrate complex to the transition state.  相似文献   

14.
Catalytic effects of galactose oxidase on the oxidation of beta-D-galactose-carrying lipids with an oligo-ethylene glycol spacer (number of ethylene glycol units (n)=1, 2, 3, 6, 9, 13, and 20) were examined. The affinity of galactose oxidase for the galactose residue in the amphiphile (estimated by the inverse of the Michaelis constant, K(m)) was much higher than those for free D-galactose and small beta-D-galactopyranosides, and dependent on the length of the ethylene glycol spacer. That is, both below and above the critical micellar concentration, the 1/K(m) values decreased with an increase in the n value. The effectiveness of the enzyme, which can be estimated by the k(cat)/K(m) value, showed the same tendency as the 1/K(m) value. These results could be attributed to the role of the nonpolar environment around the galactose residue in the binding by the enzyme. A significant enhancement of the enzymatic oxidation of galactose residue on the liposome surface was also observed.  相似文献   

15.
Several mechanisms have been considered as principal factors in enhancing the catalytic reaction velocity of enzymes: approximation, covalent catalysis, general acid-based catalysis, and strain. Among them, the strain on the substrate and/or the enzyme is often found to be brought about on association of the substrate and the enzyme. If this strain is released in the transition state, it contributes to enhancing the k(cat) value, although it does not change the k(cat)/K(m) value. In aspartate aminotransferase, however, we found by analysis of the Schiff base pK(a) values that the unliganded enzyme carries a strain in the protonated Schiff base formed between the coenzyme pyridoxal phosphate and a lysine residue. This bond is cleaved in most of the reaction intermediates, including the transition state. As a result, the activation energy between the free enzyme plus substrate and the transition state is decreased by 16 kJ/mol, equal to the value of the strain energy. The net effect of this strain is enhancement (10(3)-fold) of the catalytic efficiency in terms of k(cat)/K(m), the more important indicator of the catalytic efficiency at low concentration of the substrate.  相似文献   

16.
Hemoproteins are known to react with the strong nitrating and oxidizing agent peroxynitrite according to different mechanisms. In this article, we show that the iron(iii) forms of the sperm whale myoglobin (sw Mb) mutants H64A, H64D, H64L, F43W/H64L, and H64Y/H93G catalyze the isomerization of peroxynitrite to nitrate. The two most efficient catalysts are H64A (k(cat) = (5.8 +/- 0.1) x 10(6) M(-1) s(-1), at pH 7.5 and 20 degrees C) and H64D metMb (k(cat) = (4.8 +/- 0.1) x 10(6) M(-1) s(-1), at pH 7.5 and 20 degrees C). The pH dependence of the values of k(cat) shows that HOONO is the species which reacts with the heme. In the presence of physiologically relevant concentrations of CO(2) (1.2 mM), the decay of peroxynitrite is accelerated by these metMb mutants via the concurring reaction of HOONO with their iron(iii) centers. Studies in the presence of free added tyrosine show that the metMb mutants prevent peroxynitrite-mediated nitration. The efficiency of the different sw metMb mutants correlates with the value of k(cat). Finally, we show that sw WT-metMb is nitrated to a larger extent than horse heart metMb, a result that suggests that the additional Tyr151 is a site of preferential nitration. Again, the extent of nitration of the tyrosine residues of the metMb mutants correlates with the values of k(cat).  相似文献   

17.

The hydrolysis of bis-(nitrophenyl) phosphate (BNPP) as model for secondary phosphate esters is analyzed at six different concentrations of Eu(III) ions and four temperatures between 303 and 343 K. Eyring plots yield activation enthalpy parameters which with, e.g. [Eu 3+ ] between 0.40 and 10.0 mM drop from 130 to 74 kJ mol m 1 , respectively, with a relatively small drop in the opposing entropy contributions. The observed saturation profiles allow for the first time to evaluate the influence of j H and j S separately on the Michaelis-Menten values K M and k cat , showing that the catalytic metal-ion effects are largely due to changes in k cat and dominated by favorable j H changes. Preliminary studies of the solvent influence show a surprising difference between water mixtures with either ethanol or DMSO. With, e.g. 40% ethanol, one observes a doubling of the rate constant, with 40% DMSO an almost 10-fold rate decrease. In both cases, a linear correlation with the solvent polarity parameter E T is found.  相似文献   

18.
Electron transfer between Fe(CN)(6)(3-) and Fe(CN)(6)(4-) in homogeneous aqueous solution with K(+) as the counterion normally proceeds almost exclusively by a K(+)-catalyzed pathway, but this can be suppressed, and the direct Fe(CN)(6)(3)(-)-Fe(CN)(6)(4-) electron transfer path exposed, by complexing the K(+) with crypt-2.2.2 or 18-crown-6. Fe((13)CN)(6)(4-)-NMR line broadening measurements using either crypt-2.2.2 or (with extrapolation to zero uncomplexed [K(+)]) 18-crown-6 gave consistent values for the rate constant and activation volume (k(0) = (2.4 +/- 0.1) x 10(2) L mol(-1) s(-1) and Delta V(0) = -11.3 +/- 0.3 cm(3) mol(-1), respectively, at 25 degrees C and ionic strength I = 0.2 mol L(-1)) for the uncatalyzed electron transfer path. These values conform well to predictions based on Marcus theory. When [K(+)] was controlled with 18-crown-6, the observed rate constant k(ex) was a linear function of uncomplexed [K(+)], giving k(K) = (4.3 +/- 0.1) x 10(4) L(2) mol(-2) s(-1) at 25 degrees C and I = 0.26 mol L(-1) for the K(+)-catalyzed pathway. When no complexing agent was present, k(ex) was roughly proportional to [K(+)](total), but the corresponding rate constant k(K)' (=k(ex)/[K(+)](total)) was about 60% larger than k(K), evidently because ion pairing by hydrated K(+) lowered the anion-anion repulsions. Ionic strength as such had only a small effect on k(0), k(K), and k(K)'. The rate constants commonly cited in the literature for the Fe(CN)(6)(3-/4-) self-exchange reaction are in fact k(K)'[K(+)](total) values for typical experimental [K(+)](total) levels.  相似文献   

19.
Beta-D-xylosidase from the ruminal anaerobic bacterium, Selenomonas ruminantium (SXA), catalyzes hydrolysis of beta-1,4-xylooligosacharides and has potential utility in saccharification processes. The enzyme, heterologously produced in Escherichia coli and purified to homogeneity, has an isoelectric point of approx 4.4, an intact N terminus, and a Stokes radius that defines a homotetramer. SXA denatures between pH 4.0 and 4.3 at 25 degrees C and between 50 and 60 degrees C at pH 5.3. Following heat or acid treatment, partially inactivated SXA exhibits lower k(cat) values, but similar K(m) values as untreated SXA. D-glucose and D-xylose protect SXA from inactivation at high temperature and low pH.  相似文献   

20.
The complex [Cu2(L-66)]2+ (L-66 = a,a'-bis?bis[2-(1'-methyl-2'-benzimidazolyl)ethyl]amino?-m-xylene) undergoes fully reversible oxygenation at low temperature in acetone. The optical [lambda(max) = 362 (epsilon 15000), 455 (epsilon 2000), and 550 nm (epsilon 900M(-1)cm(-1))] and resonance Raman features (760 cm(-1), shifted to 719cm(-1)(-1) with 18O2) of the dioxygen adduct [Cu2(L-66)(O2)]2+ indicate that it is a mu-eta2:eta2-peroxodicopper(II) complex. The kinetics of dioxygen binding, studied at - 78 degrees C, gave the rate constant k1 = 1.1M(-1) 5(-1) for adduct formation, and k(-1) =7.8 x 10(-5)s(-1), for dioxygen release from the Cu2O2 complex. From these values, the O2 binding constant K= 1.4 x 10(4)M(-1) at -78 degrees C could be determined. The [Cu2(L-66)(O2)]2+ complex performs the regiospecific ortho-hydroxylation of 4-carbomethoxyphenolate to the corresponding catecholate and the oxidation of 3,5-di-tert-butylcatechol to the quinone at -60 degrees C. Therefore, [Cu2(L-66)]2+ is the first synthetic complex to form a stable dioxygen adduct and exhibit true tyrosinase-like activity on exogenous phenolic compounds.  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