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.
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   

2.
Rate and equilibrium constants at 25 °C, pH ∼ 1, and ionic strength 0.10 for hydrolysis of the two non-equivalent chlorides of dichloro[S-methyl-l-cysteine(N,S)]platinum(II) isomers, denoted [PtCl2(SmecysH)], and the resultant chloro-aqua species have been determined by NMR, potentiometric, and spectrophotometric methods. Though hydrolysis constants, Kh, for the two chlorides are similar (pKh = 4-5), the rate of hydrolysis of the chloride trans to coordinated S, kh = 3.4 × 10−3 s−1, is 2-3 orders of magnitude faster than the kh for the other chloride, 2.3 × 10−6 s−1, and for the cancer drug cisplatin, cis-[PtCl2(NH3)2], 5.2 × 10−5 s−1. Relative rates of hydrolysis determined under three different experimental conditions (pH ∼ 1 in 0.10 M HNO3, high pH in 0.10 M NaOH, and at low pH with Ag+ assistance) are consistent: the Cl trans to S is 100-1000 times more labile than the Cl cis to S. Potentiometric and NMR methods were also used to estimate pKa values of all aqua species, which are comparable to values reported for corresponding aqua species derived from cisplatin.  相似文献   

3.
Recombinant β-galactosidase from Lactobacillus plantarum WCFS1, homologously over-expressed in L. plantarum, was purified to apparent homogeneity using p-aminobenzyl 1-thio-β-d-galactopyranoside affinity chromatography and subsequently characterized. The enzyme is a heterodimer of the LacLM-family type, consisting of a small subunit of 35 kDa and a large subunit of 72 kDa. The optimum pH for hydrolysis of its preferred substrates o-nitrophenyl-β-d-galactopyranoside (oNPG) and lactose is 7.5 and 7.0, and optimum temperature for these reactions is 55 and 60 °C, respectively. The enzyme is most stable in the pH range of 6.5-8.0. The Km, kcat and kcat/Km values for oNPG and lactose are 0.9 mM, 92 s−1, 130 mM−1 s−1 and 29 mM, 98 s−1, 3.3 mM−1 s−1, respectively. The L. plantarum β-galactosidase possesses a high transgalactosylation activity and was used for the synthesis of prebiotic galacto-oligosaccharides (GOS). The resulting GOS mixture was analyzed in detail, and major components were identified by using high performance anion exchange chromatography with pulsed amperometric detection (HPAEC-PAD) as well as capillary electrophoresis. The maximal GOS yield was 41% (w/w) of total sugars at 85% lactose conversion (600 mM initial lactose concentration). The enzyme showed a strong preference for the formation of β-(1→6) linkages in its transgalactosylation mode, while β-(1→3)-linked products were formed to a lesser extent, comprising ∼80% and 9%, respectively, of the newly formed glycosidic linkages in the oligosaccharide mixture at maximum GOS formation. The main individual products formed were β-d-Galp-(1→6)-d-Lac, accounting for 34% of total GOS, and β-d-Galp-(1→6)-d-Glc, making up 29% of total GOS.  相似文献   

4.
Two new β-carbonic anhydrases (CAs, EC 4.2.1.1) from the bacterial pathogen Salmonella enterica serovar Typhimurium, stCA 1 and stCA 2, were characterized kinetically. The two enzymes possess appreciable activity as catalysts for the hydration of CO2 to bicarbonate, with kcat of 0.79 × 106 s−1 and 1.0 × 106 s−1, and kcat/Km of 5.2 × 107 M−1 s−1 and of 8.3 × 107 M−1 s−1, respectively. A large number of simple/complex inorganic anions as well as other small molecules (sulfamide, sulfamic acid, phenylboronic acid, phenylarsonic acid, dialkyldithiocarbamates) showed interesting inhibitory properties towards the two new enzymes, with several low micromolar inhibitors discovered. As many strains of S. enterica show extensive resistance to classical antibiotics, inhibition of the β-CAs investigated here may be useful for developing lead compounds for novel types of antibacterials.  相似文献   

5.
ADP-ribosyl cyclase and NAD+ glycohydrolase (CD38, E.C.3.2.2.5) efficiently catalyze the exchange of the nicotinamidyl moiety of NAD+, nicotinamide adenine dinucleotide phosphate (NADP+) or nicotinamide mononucleotide (NMN+) with an alternative base. 4′-Pyridinyl drugs (amrinone, milrinone, dismerinone and pinacidil) were efficient alternative substrates (kcat/KM = 0.9-10 μM−1 s−1) in the exchange reaction with ADP-ribosyl cyclase. When CD38 was used as a catalyst the kcat/KM values for the exchange reaction were reduced two or more orders of magnitude (0.015-0.15 μM−1 s−1). The products of this reaction were novel dinucleotides. The values of the equilibrium constants for dinucleotide formation were determined for several drugs. These enzymes also efficiently catalyze the formation of novel mononucleotides in an exchange reaction with NMN+, kcat/KM = 0.05-0.4 μM−1 s−1. The kcat/KM values for the exchange reaction with NMN+ were generally similar (0.04-0.12 μM−1 s−1) with CD38 and ADP-ribosyl cyclase as catalysts. Several novel heterocyclic alternative substrates were identified as 2-isoquinolines, 1,6-naphthyridines and tricyclic bases. The kcat/KM values for the exchange reaction with these substrates varied over five orders of magnitude and approached the limit of diffusion with 1,6-naphthyridines. The exchange reaction could be used to synthesize novel mononucleotides or to identify novel reversible inhibitors of CD38.  相似文献   

6.
The kinetics of Ca2+-dependent conformational changes of human cardiac troponin (cTn) were studied on isolated cTn and within the sarcomeric environment of myofibrils. Human cTnC was selectively labeled on cysteine 84 with N-((2-(iodoacetoxy)ethyl)-N-methyl)amino-7-nitrobenz-2-oxa-1,3-diazole and reconstituted with cTnI and cTnT to the cTn complex, which was incorporated into guinea pig cardiac myofibrils. These exchanged myofibrils, or the isolated cTn, were rapidly mixed in a stopped-flow apparatus with different [Ca2+] or the Ca2+-buffer 1,2-Bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid to determine the kinetics of the switch-on or switch-off, respectively, of cTn. Activation of myofibrils with high [Ca2+] (pCa 4.6) induced a biphasic fluorescence increase with rate constants of >2000 s−1 and ∼330 s−1, respectively. At low [Ca2+] (pCa 6.6), the slower rate was reduced to ∼25 s−1, but was still ∼50-fold higher than the rate constant of Ca2+-induced myofibrillar force development measured in a mechanical setup. Decreasing [Ca2+] from pCa 5.0-7.9 induced a fluorescence decay with a rate constant of 39 s−1, which was approximately fivefold faster than force relaxation. Modeling the data indicates two sequentially coupled conformational changes of cTnC in myofibrils: 1), rapid Ca2+-binding (kB ≈ 120 μM−1 s−1) and dissociation (kD ≈ 550 s−1); and 2), slower switch-on (kon = 390s−1) and switch-off (koff = 36s−1) kinetics. At high [Ca2+], ∼90% of cTnC is switched on. Both switch-on and switch-off kinetics of incorporated cTn were around fourfold faster than those of isolated cTn. In conclusion, the switch kinetics of cTn are sensitively changed by its structural integration in the sarcomere and directly rate-limit neither cardiac myofibrillar contraction nor relaxation.  相似文献   

7.
Efficient electron transfer from reductase domain to oxygenase domain in nitric oxide synthase (NOS) is dependent on the binding of calmodulin (CaM). Rate constants for the binding of CaM to NOS target peptides was only determined previously by surface plasmon resonance (SPR) (Biochemistry 35, 8742-8747, 1996) suggesting that the binding of CaM to NOSs is slow and does not support the fast electron transfer in NOSs measured in previous and this studies. To resolve this contradiction, the binding rates of holo Alexa 350 labeled T34C/T110W CaM (Alexa-CaM) to target peptides from three NOS isozymes were determined using fluorescence stopped-flow. All three target peptides exhibited fast kon constants at 4.5 °C: 6.6 × 108 M− 1 s− 1 for nNOS726-749, 2.9 × 108 M− 1 s− 1 for eNOS492-511 and 6.1 × 108 M− 1 s− 1 for iNOS507-531, 3-4 orders of magnitude faster than those determined previously by SPR. Dissociation rates of NOS target peptides from Alexa-CaM/peptide complexes were measured by Ca2+ chelation with ETDA: 3.7 s− 1 for nNOS726-749, 4.5 s− 1 for eNOS492-511, and 0.063 s− 1 for iNOS507-531. Our data suggest that the binding of CaM to NOS is fast and kinetically competent for efficient electron transfer and is unlikely rate-limiting in NOS catalysis. Only iNOS507-531 was able to bind apo Alexa-CaM, but in a very different conformation from its binding to holo Alexa-CaM.  相似文献   

8.
Previous work demonstrated that a mixture of NH4Cl and KNO3 as nitrogen source was beneficial to fed-batch Arthrospira (Spirulina) platensis cultivation, in terms of either lower costs or higher cell concentration. On the basis of those results, this study focused on the use of a cheaper nitrogen source mixture, namely (NH4)2SO4 plus NaNO3, varying the ammonium feeding time (T = 7-15 days), either controlling the pH by CO2 addition or not. A. platensis was cultivated in mini-tanks at 30 °C, 156 μmol photons m−2 s−1, and starting cell concentration of 400 mg L−1, on a modified Schlösser medium. T = 13 days under pH control were selected as optimum conditions, ensuring the best results in terms of biomass production (maximum cell concentration of 2911 mg L−1, cell productivity of 179 mg L−1 d−1 and specific growth rate of 0.77 d−1) and satisfactory protein and lipid contents (around 30% each).  相似文献   

9.
Two extracellular chitinases (designated as Chi-56 and Chi-64) produced by Massilia timonae were purified by ion-exchange chromatography, ammonium sulfate precipitation, and gel-filtration chromatography. The molecular mass of Chi-56 was 56 kDa as determined by both SDS-PAGE and gel-filtration chromatography. On the other hand, Chi-64 showed a molecular mass of 64 kDa by SDS-PAGE and 28 kDa by gel-filtration chromatography suggesting that its properties may be different from those of Chi-56. The optimum temperature, optimum pH, pI, Km, and Vmax of Chi-56 were 55 °C, pH 5.0, pH 8.5, 1.1 mg mL−1, and 0.59 μmol μg−1 h−1, respectively. For Chi-64, these values were 60 °C, pH 5.0, pH 8.5, 1.3 mg mL−1, and 1.36 μmol μg−1 h−1, respectively. Both enzymes were stimulated by Mn2+ and inhibited by Hg2+, and neither showed exochitinase activity. The N-terminal sequences of Chi-56 and Chi-64 were determined to be Q-T-P-T-Y-T-A-T-L and Q-A-D-F-P-A-P-A-E, respectively.  相似文献   

10.
The reactions of a dioxotetraamine Cu(II) complex [Cu(H−2L)] (L is 6-(9-fluorenyl)-1,4,8,11-tetraazaandencane-5,7-dione)with O2 − were investigated by electrochemistry, UV-Vis spectrophotometry and pulse radiolysis, respectively. In DMSO solution, [CuII(H−2L)] was oxidized into [CuIII(H−2L)]+ by O2 −, a consecutive reaction was observed with [CuIII(H−2L)(O2 2−)] − as intermediates (k1=1.71×103 M−1 s−1, k2=1.2×10−2 s−1). The mechanism of O2 − dismutation catalyzed by the complex involved alternate oxidation and reduction of Cu(II) by O2 − and the kcat is 6.07 × 107 M−1 s−1 (pH 7.4).  相似文献   

11.
Kinetic studies of X exchange on [AuX4] square-planar complexes (where X=Cl and CN) were performed at acidic pH in the case of chloride system and as a function of pH for the cyanide one. Chloride NMR study (330-365 K) gives a second-order rate law on [AuCl4] with the kinetic parameters: (k2Au,Cl)298=0.56±0.03 s−1 mol−1 kg; ΔH2‡ Au,Cl=65.1±1 kJ mol−1; ΔS2‡ Au,Cl=−31.3±3 J mol−1 K−1 and ΔV2 Au,Cl=−14±2 cm3 mol−1. The variable pressure data clearly indicate the operation of an Ia or A mechanism for this exchange pathway. The proton exchange on HCN was determined by 13C NMR as a function of pH and the rate constant of the three reaction pathways involving H2O, OH and CN were determined: k0HCN,H=113±17 s−1, k1HCN,H=(2.9±0.7)×109 s−1 mol−1 kg and k2HCN,H=(0.6±0.2)×106 s−1 mol−1 kg at 298.1 K. The rate law of the cyanide exchange on [Au(CN)4] was found to be second order with the following kinetic parameters: (k2Au,CN)298=6240±85 s−1 mol−1 kg, ΔH2 Au,CN=40.0±0.8 kJ mol−1, ΔS2 Au,CN=−37.8±3 J mol−1 K−1 and ΔV2 Au,CN=+2±1 cm3 mol−1. The rate constant observed varies about nine orders of magnitude depending on the pH and HCN does not act as a nucleophile. The observed rate constant of X exchange on [AuX4] are two or three orders of magnitude faster than the Pt(II) analogue.  相似文献   

12.
A cDNA clone (GenBank Accession No. AY835398) encoding a sesquiterpene synthase, (E)-β-farnesene synthase, has been isolated from Artemisia annua L. It contains a 1746-bp open reading frame coding for 574 amino acids (66.9 kDa) with a calculated pI = 5.03. The deduced amino acid sequence is 30-50% identical with sequences of other sesquiterpene synthases from angiosperms. The recombinant enzyme, produced in Escherichia coli, catalyzed the formation of a single product, β-farnesene, from farnesyl diphosphate. The pH optimum for the recombinant enzyme is around 6.5 and the Km- and kcat-values for farnesyl diphosphate, is 2.1 μM and 9.5 × 10−3 s−1, respectively resulting in the efficiency 4.5 × 10−3 M−1 s−1. The enzyme exhibits substantial activity in the presence of Mg2+, Mn2+ or Co2+ but essentially no activity when Zn2+, Ni2+ or Cu2+ is used as cofactor. The concentration required for maximum activity are estimated to 5 mM, 0.5 mM and <10 μM for Mg2+, Co2+ or Mn2+, respectively. Geranyl diphosphate is not a substrate for the recombinant enzyme.  相似文献   

13.
Two 15N-labelled cis-Pt(II) diamine complexes with dimethylamine (15N-dma) and isopropylamine (15N-ipa) ligands have been prepared and characterised. [1H,15N] HSQC NMR spectroscopy is used to obtain the rate and equilibrium constants for the aquation of cis-[PtCl2(15N-dma)2] at 298 K in 0.1 M NaClO4 and to determine the pKa values of cis-[PtCl(H2O)(15N-dma)2]+ (6.37) and cis-[Pt(H2O)2(15N-dma)2]2+ (pKa1 = 5.17, pKa2 = 6.47). The rate constants for the first and second aquation steps (k1 = (2.12 ± 0.01) × 10−5 s−1, k2 = (8.7 ± 0.7) × 10−6 s−1) and anation steps (k−1 = (6.7 ± 0.8) × 10−3 M−1 s−1, k−2 = 0.043 ± 0.004 M−1 s−1) are very similar to those reported for cisplatin under similar conditions, and a minor difference is that slow formation of the hydroxo-bridged dimer is observed. Aquation studies of cis-[PtCl2(15N-ipa)2] were precluded by the close proximity of the NH proton signal to the 1H2O resonance.  相似文献   

14.
The synthesis and characterisation of cis- and trans-[Co(tmen)2(NCCH3)2](ClO4)3 are described. Solvolysis rates have been measured by both 1H NMR spectroscopy and UV-Vis spectrophotometry in dimethyl sulfoxide at 298.2 K. The cis isomer undergoes solvolysis by consecutive first-order reactions, k1=5.61 × 10−4 and k2=5.35 × 10−4 s−1, each with steric retention. The measured solvolysis rate (single step reaction) for the trans isomer is k=1.54 × 10−5 s−1. The solvent exchange rates have been measured by 1H NMR spectroscopy in CD3CN at 298.2 K: kex(cis)=kct + kcc=2.0 × 10−5 and kex(trans)=ktc + ktt=4.56 × 10−6 s−1. From these data, the measured cis-trans isomerisation rate (1.71 × 10−6 s−1) and equilibrium position in CH3CN (17% trans), the steric course for substitution in the exchange processes has been determined: trans reactant - 69% trans product; cis reactant - 99% cis product. Aquation rates for cis- and trans-[Co(tmen)2(NCCH3)2](ClO4)3 have also been determined spectrophotometrically and by NMR; kcis=1.3 × 10−4 and ktrans=2.7 × 10−5 s−1. In both cases the steric course for the primary aquation step is indeterminate because the subsequent steps are faster. Where data are available, the [Co(tmen)2X2]n+ complexes are found to be consistently much more reactive than their [Co(en)2X2]n+ analogues.  相似文献   

15.
Two new MnIII complexes Na[Mn(5-SO3-salpnOH)(H2O)] ⋅ 5H2O (1) and Na[Mn(5-SO3-salpn)(MeOH)] ⋅ 4H2O (2) (5-SO3-salpnOH = 1,3-bis(5-sulphonatosalicylidenamino)propan-2-ol, 5-SO3-salpn = 1,3-bis(5-sulphonatosalicylidenamino)propane) have been prepared and characterized. Electrospray ionization-mass spectrometry, UV-visible and 1H NMR spectroscopic studies showed that the two complexes exist in solution as monoanions [Mn(5-SO3-salpn(OH))(solvent)2], with the ligand bound to MnIII through the two phenolato-O and two imino-N atoms located in the equatorial plane. The E1/2 of the MnIII/MnII couple (−47.11 (1) and −77.80 mV (2) vs. Ag/AgCl) allows these complexes to efficiently catalyze the dismutation of , with catalytic rate constants 2.4 × 106 (1) and 3.6 × 106 (2) M−1 s−1, and IC50 values of 1.14 (1) and 0.77 (2) μM, obtained through the nitro blue tetrazolium photoreduction inhibition superoxide dismutase assay, in aqueous solution of pH 7.8. The two complexes are also able to disproportionate up to 250 equivalents of H2O2 in aqueous solution of pH 8.0, with initial turnover rates of 178 (1) and 25.2 (2) mM H2O2 min−1 mM−1 catalyst−1. Their dual superoxide dismutase/catalase activity renders these compounds particularly attractive as catalytic antioxidants.  相似文献   

16.
The high light sensitivity of three submerged aquatic freshwater plant species, Egeria densa, Elodea nuttallii and Myriophyllum heterophyllum, which have been cultivated at a photosynthetically active radiation (PAR, 400-700 nm) of 70 μmol photons m−2 s−1, was studied by means of chlorophyll fluorescence and pigment analyses. Exposure of plants to 100, 300, 600 and 1000 μmol photons m−2 s−1 PAR for up to 360 min induced a strong reduction of the Fv/Fm ratio, indicating a pronounced inactivation of PSII even at the lowest PAR applied. These changes were accompanied by a reduction of the chlorophyll content to about 60-70% of control values at the highest PAR. Rapidly inducible photoprotective mechanisms were not affected, as derived from the rapid generation of pH-dependent energy dissipation under these conditions. At PAR higher than 100 μmol photons m−2 s−1, however, the primary quinone acceptor of photosystem II, QA, was reduced to about 80% and the effective quantum yield of photosystem II, ΦPSII, dropped to values of about 10%, indicating a high reduction state of the photosynthetic electron transport chain. These data support the notion that the three aquatic macrophytes have a very low capacity for the acclimation to higher light intensities.  相似文献   

17.
The response of rapid light–response curves (RLCs) of variable fluorescence to changes in short- and long-term photoacclimation status was studied in an estuarine benthic diatom. The diatom Nitzschia palea was grown under low- (LL, 20 μmol m−2 s−1) and high-light (HL, 400 μmol m−2 s−1) conditions, with the purpose of characterising the effects of long-term photoacclimation on (i) steady-state light–response curves (LC) of relative electron transport rate, rETR, (ii) the response of RLCs to changes in ambient irradiance (E, the irradiance to which the sample is acclimated to immediately before the RLCs), (iii) the relationship of RLCs to LC parameters and non-photochemical quenching (NPQ). Photoacclimation to LL and HL conditions induced distinct light–response patterns of rETR and NPQ. Higher growth light resulted in rETR vs. E curves with lower initial slopes (α, 0.591 μmol−1 m2 s vs. 0.661 μmol−1 m2 s, for HL and LL, respectively) and markedly higher maximum rates (rETRm, 95.9 vs. 29.3), reached under higher E levels (higher light-saturation coefficient, Ek: 162.4 μmol m−2 s−1 vs. 44.3 μmol m−2 s−1). Acclimation to HL induced bi-phasic NPQ vs. E curves, with minimum values reached under low E levels (15–25 μmol m−2 s−1) and not on dark-acclimated samples. The response of RLCs to changes in ambient irradiance varied with the long-term photoacclimation status of the samples. The initial slope, αRLC, decreased monotonically with E in LL cultures, from 0.68 to 0.25 μmol−1 m2 s, while varied bi-phasically in HL-acclimated samples. Typically, αRLC of HL cultures increased under low E, reaching a maximum of 0.61 μmol−1 m2 s under 25–55 μmol m−2 s−1, and decreased gradually under higher E levels to 0.25 μmol−1 m2 s. RLC maximum rETR, rETRm,RLC, and saturation coefficient Ek,RLC, increased with E following a saturation-like pattern, with the HL cultures presenting markedly higher values for all the E range (maximum rETRm,RLC values were 108.6 and 33.4 for HL and LL cultures, respectively). An inverse relationship was consistently found between αRLC and NPQ, both on LL and HL cultures, causing strong correlations (P < 0.001 in all cases) between NPQ and the high light-induced decrease of αRLC, ΔαRLC. RLCs were confirmed to also provide information on the long-term photoacclimation status, as significant correlations (P < 0.001 both for HL and LL cultures) were verified between Ek and an index based on RLC parameters, Êk, both for LL and HL cultures. These results reinforce the usefulness of RLCs as a tool for inferring on the short- and long-term photoacclimation status of samples with different long-term light histories, through the estimation of LC parameters and the monitoring of NPQ levels.  相似文献   

18.
Chen S  Hu Q  Hu M  Luo J  Weng Q  Lai K 《Bioresource technology》2011,102(17):8110-8116
Fungal strain HU, isolated from activated sludge and identified as a member of the genus Cladosporium based on morphology and sequencing of 28S rRNA, was shown to degrade 90% of fenvalerate, fenpropathrin, β-cypermethrin, deltamethrin, bifenthrin, and permethrin (100 mg L−1) within 5 days. Fenvalerate was utilized as sole carbon and energy source and co-metabolized in the presence of sucrose. Degradation of fenvalerate occurred at pH 5-10 at 18-38 °C. The fungus first hydrolyzed the carboxylester linkage to produce α-hydroxy-3-phenoxy-benzeneacetonitrile and 3-phenoxybenzaldehyde, and subsequently degraded these two compounds with a qmax, Ks and Ki of 1.73 d−1, 99.20 mg L−1 and 449.75 mg L−1, respectively. Degradation followed first-order kinetics. These results show that the fungal strain may possess potential to be used in bioremediation of pyrethroid-contaminated environments.  相似文献   

19.
In biological systems, enzymes often use metal ions, especially Mg2+, to catalyze phosphodiesterolysis, and model aqueous studies represent an important avenue of examining the contributions of these ions to catalysis. We have examined Mg2+ and Ca2+ catalyzed hydrolysis of the model phosphodiester thymidine-5′-p-nitrophenyl phosphate (T5PNP). At 25 °C, we find that, despite their different Lewis acidities, these ions have similar catalytic ability with second-order rate constants for attack of T5PNP by hydroxide (kOH) of 4.1 × 10−4 M−1s−1 and 3.7 × 10−4 M−1s−1 in the presence of 0.30 M Mg2+ and Ca2+, respectively, compared to 8.3 × 10−7 M−1s−1 in the absence of divalent metal ion. Examining the dependence of kOH on [M2+] at 50 °C indicates different kinetic mechanisms with Mg2+ utilizing a single ion mechanism and Ca2+ operating by parallel single and double ion mechanisms. Association of the metal ion(s) occurs prior to nucleophilic attack by hydroxide. Comparing the kOH values reveals a single Mg2+ catalyzes the reaction by 1800-fold whereas a single Ca2+ ion catalyzes the reaction by only 90-fold. The second Ca2+ provides an additional 10-fold catalysis, significantly reducing the catalytic disparity between Mg2+ and Ca2+.  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