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.
p-Hydroxyphenylpyruvate dioxygenase (HPPD) is a key enzyme in tyrosine catabolism and is the molecular target site of β-triketone pharmacophores used to treat hypertyrosinemia in humans. In plants, HPPD is involved in the biosynthesis of prenyl quinones and tocopherols, and is the target site of β-triketone herbicides. The β-triketone-rich essential oil of manuka (Leptospermum scoparium), and its components leptospermone, grandiflorone and flavesone were tested for their activity in whole-plant bioassays and for their potency against HPPD. The achlorophyllous phenotype of developing plants exposed to manuka oil or its purified β-triketone components was similar to that of plants exposed to the synthetic HPPD inhibitor sulcotrione. The triketone-rich fraction and leptospermone were approximatively 10 times more active than that of the crude manuka oil, with I50 values of 1.45, 0.96 and 11.5 μg mL−1, respectively. The effect of these samples on carotenoid levels was similar. Unlike their synthetic counterpart, steady-state O2 consumption experiments revealed that the natural triketones were competitive reversible inhibitors of HPPD. Dose-response curves against the enzyme activity of HPPD provided apparent I50 values 15.0, 4.02, 3.14, 0.22 μg mL−1 for manuka oil, triketone-rich fraction, leptospermone and grandiflorone, respectively. Flavesone was not active. Structure-activity relationships indicate that the size and lipophilicity of the side-chain affected the potency of the compounds. Computational analysis of the catalytic domain of HPPD indicates that a lipophilic domain proximate from the Fe2+ favors the binding of ligands with lipophilic moieties.  相似文献   

2.
It was recently shown that the structure of the fluorophore attached to the acyl chain of phosphatidylcholine analogs determines their mechanism of transport across the plasma membrane of yeast cells (Elvington et al., J. Biol Chem. 280:40957, 2005). In order to gain further insight into the physical properties of these fluorescent phosphatidylcholine (PC) analogs, the rate and mechanism of their intervesicular transport was determined. The rate of spontaneous exchange was measured for PC analogs containing either NBD (7-nitrobenz-2-oxa-1,3-diazol-4-yl), Bodipy FL (4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene), Bodipy 530 (4,4-difluoro-5,7-diphenyl-4-bora-3a,4a-diaza-s-indacene), or Bodipy 581 (4,4-difluoro-5-(4-phenyl-1,3-butadienyl)-4-bora-3a,4a-diaza-s-indacene) attached to a five or six carbon acyl chain in the sn-2 position. The rate of transfer between phospholipid vesicles was measured by monitoring the increase in fluorescence as the analogs transferred from donor vesicles containing self-quenching concentrations to unlabeled acceptor vesicles. Kinetic analysis indicated that the transfer of each analog occurred by diffusion through the water phase as opposed to transfer during vesicle collisions. The vesicle-to-monomer dissociation rate constants differed by over four orders of magnitude: NBD-PC (kdis = 0.115 s− 1; t1/2 = 6.03 s); Bodipy FL-PC (kdis = 5.2 × 10− 4; t1/2 = 22.2 min); Bodipy 530-PC (kdis = 1.52 × 10− 5; t1/2 = 12.6 h); and Bodipy 581-PC (kdis = 5.9 × 10− 6; t1/2 = 32.6 h). The large differences in spontaneous rates of transfer through the water measured for these four fluorescent PC analogs reflect their hydrophobicity and may account for their recognition by different mechanisms of transport across the plasma membrane of yeast.  相似文献   

3.
Combined pH-metric, UV-Vis, 1H NMR and EPR spectral investigations on the complex formation of M(II) ions (M=Co, Ni, Cu and Zn) with N-(2-benzimidazolyl)methyliminodiacetic acid (H2bzimida, hereafter H2L) in aqueous solution at a fixed ionic strength, I=10−1 mol dm−3, at 25 ± 1 °C indicate the formation of M(L), M(H−1L) and M2(H−1L)+ complexes. Proton-ligand and metal-ligand constants and the complex formation equilibria have been elucidated. Solid complexes, [M(L)(H2O)2] · nH2O (n=1 for M = Co and Zn, n=2 for M = Ni) and {Cu (μ-L) · 4H2O}n, have been isolated and characterized by elemental analysis, spectral, conductance and magnetic measurements and thermal studies. Structures of [Ni(L)(H2O)2] · 2H2O and {Cu(μ-L) · 4H2O}n have been determined by single crystal X-ray diffraction. The nickel(II) complex exists in a distorted octahedral environment in which the metal ion is coordinated by the two carboxylate O atoms, the amino-N atom of the iminodiacetate moiety and the pyridine type N-atom of the benzimidazole moiety. Two aqua O atoms function as fifth and sixth donor atoms. The copper(II) complex is made up of interpenetrating polymeric chains of antiferromagnetically coupled Cu(II) ions linked by carboxylato bridges in syn-anti (apical-equatorial) bonding mode and stabilized via interchain hydrogen bonds and π-π stacking interactions.  相似文献   

4.
The N-terminal segment of the Semliki Forest virus polyprotein is an intramolecular serine protease that cleaves itself off after the invariant Trp267 from a viral polyprotein and generates the mature capsid protein. After this autoproteolytic cleavage, the free carboxylic group of Trp267 interacts with the catalytic triad (His145, Asp167 and Ser219) and inactivates the enzyme. We have deleted the last 1-7 C-terminal residues of the mature capsid protease to investigate whether removal of Trp267 regenerates enzymatic activity. Although the C-terminally truncated polypeptides do not adopt a defined three-dimensional structure and show biophysical properties observed in natively unfolded proteins, they efficiently catalyse the hydrolysis of aromatic amino acid esters, with higher catalytic efficiency for tryptophan compared to tyrosine esters and kcat/KM values up to 5 × 105 s−1 M−1. The enzymatic mechanism of these deletion variants is typical of serine proteases. The pH enzyme activity profile shows a pKa1 = 6.9, and the Ser219Ala substitution destroys the enzymatic activity. In addition, the fast release of the first product of the enzymatic reaction is followed by a steady-state second phase, indicative of formation and breakdown of a covalent acyl-enzyme intermediate. The rates of acylation and deacylation are k2 = 4.4±0.6 s−1 and k3 = 1.6±0.5 s−1, respectively, for a tyrosine derivative ester substrate, and the amplitude of the burst phase indicates that 95% of the enzyme molecules are active. In summary, our data provide further evidence for the potential catalytic activity of natively unfolded proteins, and provide the basis for engineering of alphavirus capsid proteins towards hydrolytic enzymes with novel specificities.  相似文献   

5.
In this study, we have structurally characterized the amidase of a nitrile-degrading bacterium, Rhodococcus sp. N-771 (RhAmidase). RhAmidase belongs to amidase signature (AS) family, a group of amidase families, and is responsible for the degradation of amides produced from nitriles by nitrile hydratase. Recombinant RhAmidase exists as a dimer of about 107 kDa. RhAmidase can hydrolyze acetamide, propionamide, acrylamide and benzamide with kcat/Km values of 1.14 ± 0.23 mM− 1s− 1, 4.54 ± 0.09 mM− 1s− 1, 0.087 ± 0.02 mM− 1s− 1 and 153.5 ± 7.1 mM− 1s− 1, respectively. The crystal structures of RhAmidase and its inactive mutant complex with benzamide (S195A/benzamide) were determined at resolutions of 2.17 Å and 2.32 Å, respectively. RhAmidase has three domains: an N-terminal α-helical domain, a small domain and a large domain. The N-terminal α-helical domain is not found in other AS family enzymes. This domain is involved in the formation of the dimer structure and, together with the small domain, forms a narrow substrate-binding tunnel. The large domain showed high structural similarities to those of other AS family enzymes. The Ser-cis Ser-Lys catalytic triad is located in the large domain. But the substrate-binding pocket of RhAmidase is relatively narrow, due to the presence of the helix α13 in the small domain. The hydrophobic residues from the small domain are involved in recognizing the substrate. The small domain likely participates in substrate recognition and is related to the difference of substrate specificities among the AS family amidases.  相似文献   

6.
Mixed ligand complexes: [Co(L)(bipy)] · 3H2O (1), [Ni(L)(phen)] · H2O (2), [Cu(L)(phen)] · 3H2O (3) and [Zn(L)(bipy)] · 3H2O (4), where L2− = two -COOH deprotonated dianion of N-(2-benzimidazolyl)methyliminodiacetic acid (H2bzimida, hereafter, H2L), bipy = 2,2′ bipyridine and phen = 1,10-phenanthroline have been isolated and characterized by elemental analysis, spectral and magnetic measurements and thermal studies. Single crystal X-ray diffraction studies show octahedral geometry for 1, 2 and 4 and square pyramidal geometry for 3. Equilibrium studies in aqueous solution (ionic strength I = 10−1 mol dm−3 (NaNO3), at 25 ± 1 °C) using different molar proportions of M(II):H2L:B, where M = Co, Ni, Cu and Zn and B = phen, bipy and en (ethylene diamine), however, provides evidence of formation of mononuclear and binuclear binary and mixed ligand complexes: M(L), M(H−1L), M(B)2+, M(L)(B), M(H−1L)(B), M2(H−1L)(OH), (B)M(H−1L)M(B)+, where H−1L3− represents two -COOH and the benzimidazole N1-H deprotonated quadridentate (O, N, O, N), or, quinquedentate (O, N, O, N, N) function of the coordinated ligand H2L. Binuclear mixed ligand complex formation equilibria: M(L)(B) + M(B)2+ ? (B)M(H−1L)M(B)+ + H+ is favoured with higher π-acidity of the B ligands. For Co(II), Ni(II) and Cu(II), these equilibria are accompanied by blue shift of the electronic absorption maxima of M(II) ions, as a negatively charged bridging benzimidazolate moiety provides stronger ligand field than a neutral one. Solution stability of the mixed ligand complexes are in the expected order: Co(II) < Ni(II) < Cu(II) > Zn(II). The Δ log KM values are less negetive than their statistical values, indicating favoured formation of the mixed ligand complexes over the binary ones.  相似文献   

7.
Human arylamine N-acetyltransferase 1 (NAT1) is a xenobiotic-metabolizing enzyme that biotransforms aromatic amine chemicals. We show here that biologically-relevant concentrations of inorganic (Hg2+) and organic (CH3Hg+) mercury inhibit the biotransformation functions of NAT1. Both compounds react irreversibly with the active-site cysteine of NAT1 (half-maximal inhibitory concentration (IC50) = 250 nM and kinact = 1.4 × 104 M−1 s−1 for Hg2+ and IC50 = 1.4 μM and kinact = 2 × 102 M−1 s−1 for CH3Hg+). Exposure of lung epithelial cells led to the inhibition of cellular NAT1 (IC50 = 3 and 20 μM for Hg2+ and CH3Hg+, respectively). Our data suggest that exposure to mercury may affect the biotransformation of aromatic amines by NAT1.  相似文献   

8.
Ferric heme proteins bind weakly basic ligands and the binding affinity is often pH dependent due to protonation of the ligand as well as the protein. In an effort to find a small, neutral ligand without significant acid/base properties to probe ligand binding reactions in ferric heme proteins we were led to consider the organonitriles. Although organonitriles are known to bind to transition metals, we have been unable to find any prior studies of nitrile binding to heme proteins. In this communication we report on the equilibrium and kinetic properties of acrylonitrile binding to cytochrome c peroxidase (CcP) as well as the oxidation of acrylonitrile by CcP compound I. Acrylonitrile binding to CcP is independent of pH between pH 4 and 8. The association and dissociation rate constants are 0.32 ± 0.16 M−1 s−1 and 0.34 ± 0.15 s−1, respectively, and the independently measured equilibrium dissociation constant for the complex is 1.1 ± 0.2 M. We have demonstrated for the first time that acrylonitrile can bind to a ferric heme protein. The binding mechanism appears to be a simple, one-step association of the ligand with the heme iron. We have also demonstrated that CcP can catalyze the oxidation of acrylonitrile, most likely to 2-cyanoethylene oxide in a “peroxygenase”-type reaction, with rates that are similar to rat liver microsomal cytochrome P450-catalyzed oxidation of acrylonitrile in the monooxygenase reaction. CcP compound I oxidizes acrylonitrile with a maximum turnover number of 0.61 min−1 at pH 6.0.  相似文献   

9.
Unlike superoxide dismutases (SODs), superoxide reductases (SORs) eliminate superoxide anion (O2•−) not through its dismutation, but via reduction to hydrogen peroxide (H2O2) in the presence of an electron donor. The microaerobic protist Giardia intestinalis, responsible for a common intestinal disease in humans, though lacking SOD and other canonical reactive oxygen species-detoxifying systems, is among the very few eukaryotes encoding a SOR yet identified. In this study, the recombinant SOR from Giardia (SORGi) was purified and characterized by pulse radiolysis and stopped-flow spectrophotometry. The protein, isolated in the reduced state, after oxidation by superoxide or hexachloroiridate(IV), yields a resting species (Tfinal) with Fe3+ ligated to glutamate or hydroxide depending on pH (apparent pKa = 8.7). Although showing negligible SOD activity, reduced SORGi reacts with O2•− with a pH-independent second-order rate constant k1 = 1.0 × 109 M− 1 s− 1 and yields the ferric-(hydro)peroxo intermediate T1; this in turn rapidly decays to the Tfinal state with pH-dependent rates, without populating other detectable intermediates. Immunoblotting assays show that SORGi is expressed in the disease-causing trophozoite of Giardia. We propose that the superoxide-scavenging activity of SOR in Giardia may promote the survival of this air-sensitive parasite in the fairly aerobic proximal human small intestine during infection.  相似文献   

10.
Catalase-peroxidases (KatGs) are unique bifunctional heme peroxidases that exhibit peroxidase and substantial catalase activities. Nevertheless, the reaction pathway of hydrogen peroxide dismutation, including the electronic structure of the redox intermediate that actually oxidizes H2O2, is not clearly defined. Several mutant proteins with diminished overall catalase but wild-type-like peroxidase activity have been described in the last years. However, understanding of decrease in overall catalatic activity needs discrimination between reduction and oxidation reactions of hydrogen peroxide. Here, by using sequential-mixing stopped-flow spectroscopy, we have investigated the kinetics of the transition of KatG compound I (produced by peroxoacetic acid) to its ferric state by trapping the latter as cyanide complex. Apparent bimolecular rate constants (pH 6.5, 20 °C) for wild-type KatG and the variants Trp122Phe (lacks KatG-typical distal adduct), Asp152Ser (controls substrate access to the heme cavity) and Glu253Gln (channel entrance) are reported to be 1.2 × 104 M− 1 s− 1, 30 M− 1 s− 1, 3.4 × 103 M− 1 s− 1, and 8.6 × 103 M− 1 s− 1, respectively. These findings are discussed with respect to steady-state kinetic data and proposed reaction mechanism(s) for KatG. Assets and drawbacks of the presented method are discussed.  相似文献   

11.
The folding mechanism and stability of dimeric formate dehydrogenase from Candida methylica was analysed by exposure to denaturing agents and to heat. Equilibrium denaturation data yielded a dissociation constant of about 10−13 M for assembly of the protein from unfolded chains and the kinetics of refolding and unfolding revealed that the overall process comprises two steps. In the first step a marginally stable folded monomeric state is formed at a rate (k1) of about 2 × 10−3 s−1 (by deduction k−1 is about10−4 s−1) and assembles into the active dimeric state with a bimolecular rate constant (k2) of about 2 × 104 M−1 s−1. The rate of dissociation of the dimeric state in physiological conditions is extremely slow (k−2 ∼ 3 × 10−7 s−1).  相似文献   

12.
Microorganisms living in arsenic-rich geothermal environments act on arsenic with different biochemical strategies, but the molecular mechanisms responsible for the resistance to the harmful effects of the metalloid have only partially been examined. In this study, we investigated the mechanisms of arsenic resistance in the thermophilic bacterium Thermus thermophilus HB27. This strain, originally isolated from a Japanese hot spring, exhibited tolerance to concentrations of arsenate and arsenite up to 20 mM and 15 mM, respectively; it owns in its genome a putative chromosomal arsenate reductase (TtarsC) gene encoding a protein homologous to the one well characterized from the plasmid pI258 of the Gram + bacterium Staphylococcus aureus. Differently from the majority of microorganisms, TtarsC is part of an operon including genes not related to arsenic resistance; qRT-PCR showed that its expression was four-fold increased when arsenate was added to the growth medium. The gene cloning and expression in Escherichia coli, followed by purification of the recombinant protein, proved that TtArsC was indeed a thioredoxin-coupled arsenate reductase with a kcat/KM value of 1.2 × 104 M− 1 s− 1. It also exhibited weak phosphatase activity with a kcat/KM value of 2.7 × 10− 4 M− 1 s− 1. The catalytic role of the first cysteine (Cys7) was ascertained by site-directed mutagenesis. These results identify TtArsC as an important component in the arsenic resistance in T. thermophilus giving the first structural–functional characterization of a thermophilic arsenate reductase.  相似文献   

13.
Understanding the structure, folding, and interaction of membrane proteins requires experimental tools to quantify the association of transmembrane (TM) helices. Here, we introduce isothermal titration calorimetry (ITC) to measure integrin αIIbβ3 TM complex affinity, to study the consequences of helix–helix preorientation in lipid bilayers, and to examine protein-induced lipid reorganization. Phospholipid bicelles served as membrane mimics. The association of αIIbβ3 proceeded with a free energy change of − 4.61 ± 0.04 kcal/mol at bicelle conditions where the sampling of random helix–helix orientations leads to complex formation. At bicelle conditions that approach a true bilayer structure in effect, an entropy saving of > 1 kcal/mol was obtained from helix–helix preorientation. The magnitudes of enthalpy and entropy changes increased distinctly with bicelle dimensions, indicating long-range changes in bicelle lipid properties upon αIIbβ3 TM association. NMR spectroscopy confirmed ITC affinity measurements and revealed αIIbβ3 association and dissociation rates of 4500 ± 100 s− 1 and 2.1 ± 0.1 s− 1, respectively. Thus, ITC is able to provide comprehensive insight into the interaction of membrane proteins.  相似文献   

14.
15.
The rate of conversion of 1 to N-(2-methoxyphenyl)phthalimide (2) within [HCl] range 5.0 × 10−3-1.0 M at 1.0 M ionic strength (by NaCl) reveals the presence of both uncatalyzed and specific acid-catalyzed kinetic terms in the rate law. Intramolecular carboxamide group-assisted cleavage of amide bond of 1 reveals rate enhancement of much larger than 106-fold compared to the expected rate of analogous intermolecular reaction.  相似文献   

16.
Using an unsymmetrical N-tridentate (2-pyridyl)alkylamine ligand methyl[2-(2-pyridyl)ethyl](2-pyridylmethyl)amine (mpepma), a new tricopper(II) complex [{(mpepma)Cu(OClO3)}33-OCO2)][ClO4] (1), having triply bridging μ3-carbonate group, with three symmetry related triangular array of copper(II) centers has been synthesized. In 1, each copper ion is five-coordinate (almost perfect square pyramidal) by three nitrogens from the ligand (two pyridyl and an amine) and an oxygen atom of the carbonate group in the equatorial plane and by an oxygen atom of the perchlorate ion in the apical position. Magnetic susceptibility data (2-300 K) were fitted using the Hamiltonian: , which corresponds to a triangular array of spins, corrected with a zJ′ intertrimer interaction or zero-field splitting within the ground state, to fit low temperature data. It is concluded that the copper(II) ions in 1 are ferromagnetically coupled. The best parameters obtained are g = 2.11, J = 18 ± 1 cm−1, zJ′ = −0.36 ± 0.02 cm−1 or D = 0.50 ± 0.05 cm−1.  相似文献   

17.
(E)-2-(2-(2-hydroxyphenyl)hydrazono)-1-phenylbutane-1,3-dione (H2L) was synthesized by azocoupling of diazonium salt of 2-hydroxyaniline with 1-phenylbutane-1,3-dione and characterized by IR, 1H and 13C NMR spectroscopies and X-ray diffraction analysis. In solution, H2L exists as a mixture of the enol-azo and hydrazone tautomeric forms and a decrease of temperature and of solvent polarity shifts the tautomeric balance to the hydrazone form. In the solid state, H2L crystallizes from ethanol-water in the monohydrate hydrazone form, as shown by X-ray analysis. The dissociation constants of H2L (pK1 = 5.98 ± 0.04, pK2 = 9.72 ± 0.03) and the stability constants of its copper(II) complex (log β1 = 11.01 ± 0.07, log β2 = 20.19 ± 0.08) were determined by the potentiometric method in aqueous-ethanol solution. The copper(II) complex [Cu2(μ-L)2]n was isolated in the solid state and found by X-rays to be a coordination polymer of a binuclear core with a distorted square pyramidal metal coordination geometry.  相似文献   

18.
The joint substitution of three active-site residues in Escherichia colil-aspartate aminotransferase increases the ratio of l-cysteine sulfinate desulfinase to transaminase activity 105-fold. This change in reaction specificity results from combining a tyrosine-shift double mutation (Y214Q/R280Y) with a non-conservative substitution of a substrate-binding residue (I33Q). Tyr214 hydrogen bonds with O3 of the cofactor and is close to Arg374 which binds the α-carboxylate group of the substrate; Arg280 interacts with the distal carboxylate group of the substrate; and Ile33 is part of the hydrophobic patch near the entrance to the active site, presumably participating in the domain closure essential for the transamination reaction. In the triple-mutant enzyme, kcat′ for desulfination of l-cysteine sulfinate increased to 0.5 s− 1 (from 0.05 s− 1 in wild-type enzyme), whereas kcat′ for transamination of the same substrate was reduced from 510 s− 1 to 0.05 s− 1. Similarly, kcat′ for β-decarboxylation of l-aspartate increased from < 0.0001 s− 1 to 0.07 s− 1, whereas kcat′ for transamination was reduced from 530 s− 1 to 0.13 s− 1. l-Aspartate aminotransferase had thus been converted into an l-cysteine sulfinate desulfinase that catalyzes transamination and l-aspartate β-decarboxylation as side reactions. The X-ray structures of the engineered l-cysteine sulfinate desulfinase in its pyridoxal-5′-phosphate and pyridoxamine-5′-phosphate form or liganded with a covalent coenzyme-substrate adduct identified the subtle structural changes that suffice for generating desulfinase activity and concomitantly abolishing transaminase activity toward dicarboxylic amino acids. Apparently, the triple mutation impairs the domain closure thus favoring reprotonation of alternative acceptor sites in coenzyme-substrate intermediates by bulk water.  相似文献   

19.
Absorbance difference spectroscopy and redox titrations have been applied to investigate the properties of photosystem I from the chlorophyll d containing cyanobacterium Acaryochloris marina. At room temperature, the (P740+ − P740) and (FA/B − FA/B) absorbance difference spectra were recorded in the range between 300 and 1000 nm while at cryogenic temperatures, (P740+A1 − P740A1) and (3P740 − P740) absorbance difference spectra have been measured. Spectroscopic and kinetic evidence is presented that the cofactors involved in the electron transfer from the reduced secondary electron acceptor, phylloquinone (A1), to the terminal electron acceptor and their structural arrangement are virtually identical to those of chlorophyll a containing photosystem I. The oxidation potential of the primary electron donor P740 of photosystem I has been reinvestigated. We find a midpoint potential of 450 ± 10 mV in photosystem I-enriched membrane fractions as well as in thylakoids which is very similar to that found for P700 in chlorophyll a dominated organisms. In addition, the extinction difference coefficient for the oxidation of the primary donor has been determined and a value of 45,000 ± 4000 M− 1 cm− 1 at 740 nm was obtained. Based on this value the ratio of P740 to chlorophyll is calculated to be 1:~ 200 chlorophyll d in thylakoid membranes. The consequences of our findings for the energetics in photosystem I of A. marina are discussed as well as the pigment stoichiometry and spectral characteristics of P740.  相似文献   

20.
GOX is the most widely used enzyme for the development of electrochemical glucose biosensors and biofuel cell in physiological conditions. The present work describes the production of a recombinant glucose oxidase from Penicillium amagasakiense (yGOXpenag) displaying a more efficient glucose catalysis (kcat/KM(glucose) = 93 μM−1 s−1) than the native GOX from Aspergillus niger (nGOXaspng), which is the most industrially used (kcat/KM(glucose) = 27 μM−1 s−1). Expression in Pichia pastoris allowed easy production and purification of the recombinant active enzyme, without overglycosylation. Its biotechnological interest was further evaluated by measuring kinetics of ferrocinium-methanol (FMox) reduction, which is commonly used for electron transfer to the electrode surface. Despite their homologies in sequence and structure, pH-dependant FMox reduction was different between the two enzymes. At physiological pH and temperature, we observed that electron transfer to the redox mediator is also more efficient for yGOXpenag than for nGOXaspng(kcat/KM(FMox) = 27 μM−1 s−1 and 17 μM−1 s−1 respectively). In our model system, the catalytic current observed in the presence of blood glucose concentration (5 mM) was two times higher with yGOXpenag than with nGOXaspng. All our results indicated that yGOXpenag is a better candidate for industrial development of efficient bioelectrochemical devices used in physiological conditions.  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