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 984 毫秒
1.
A mediator-free amperometric hydrogen peroxide biosensor was prepared by immobilizing horseradish peroxidase (HRP) enzyme on colloidal Au modified platinum (Pt) wire electrode, which was modified by poly 2,6-pyridinediamine (pPA). The modified process was characterized by electrochemical impedance spectroscopy (EIS), and the electrochemical characteristics of the biosensor were studied by cyclic voltammetry, linear sweep voltammetry and chronoamperometry. The biosensor displayed an excellent electrocatalytical response to reduction of H2O2 without the aid of an electron mediator, the linear range was 4.2 × 10−7–1.5 × 10−3 mol/L (r = 0.9977), with a detection limit of 1.4 × 10−7 mol/L. Moreover, the performance and factors influencing the resulted biosensor were studied in detail. The studied biosensor exhibited permselectivity, good stability and good fabrication reproducibility.  相似文献   

2.
The kinetics and thermodynamics of Ga(III) exchange between gallium mononitrilotriacetate and human serum transferrin as well as those of the interaction between gallium-loaded transferrin and the transferrin receptor 1 were investigated in neutral media. Gallium is exchanged between the chelate and the C-site of human serum apotransferrin in interaction with bicarbonate in about 50 s to yield an intermediate complex with an equilibrium constant K 1 = (3.9 ± 1.2) × 10−2, a direct second-order rate constant k 1 = 425 ± 50 M−1 s−1 and a reverse second-order rate constant k −1 = (1.1 ± 3) × 104 M−1 s−1. The intermediate complex loses a single proton with proton dissociation constant K 1a = 80 ± 40 nM to yield a first kinetic product. This product then undergoes a modification in its conformation which lasts about 500 s to produce a second kinetic intermediate, which in turn undergoes a final extremely slow (several hours) modification in its conformation to yield the gallium-saturated transferrin in its final state. The mechanism of gallium uptake differs from that of iron and does not involve the same transitions in conformation reported during iron uptake. The interaction of gallium-loaded transferrin with the transferrin receptor occurs in a single very fast kinetic step with a dissociation constant K d = 1.10 ± 0.12 μM and a second-order rate constant k d = (1.15 ± 0.3) × 1010 M−1 s−1. This mechanism is different from that observed with the ferric holotransferrin and suggests that the interaction between the receptor and gallium-loaded transferrin probably takes place on the helical domain of the receptor which is specific for the C-site of transferrin and HFE. The relevance of gallium incorporation by the transferrin receptor-mediated iron-acquisition pathway is discussed.  相似文献   

3.
While the naturally occurring reducing agents glutathione (GSH) and ascorbate (H2A) alone are ineffective at reducing iron(III) sequestered by the siderophore ferrioxamine B, the addition of an iron(II) chelator, sulfonated bathophenanthroline (BPDS), facilitates reduction by either reducing agent. A mechanism is described in which a ternary complex is formed between ferrioxamine B and BPDS in a rapidly established pre-equilibrium step, which is followed by rate limiting reduction of the ternary complex by glutathione or ascorbate. Spectral, thermodynamic, and kinetic evidence are given for ternary complex formation. Ascorbate was found to be slightly more efficient at reducing the ternary complex than glutathione (k4=2.1×10−3 M−1 s−1 and k4=6.3×10−4 M−1 s−1, respectively) at pH 7. Reduction is followed by a rapid ligand exchange step where iron is released from ferrioxamine B to form tris-(BPDS)iron(II). The implications of these results for siderophore mediated iron transport and release are discussed.  相似文献   

4.
Formaldehyde ferredoxin oxidoreductase from Pyrococcus furiosus is a homotetrameric protein with one tungstodipterin and one [4Fe–4S] cubane per 69-kDa subunit. The enzyme kinetics have been studied under steady-state conditions at 80 °C and pre-steady state conditions at 50 °C, in the latter case via monitoring of the relatively weak (ε ≈ 2 mM−1 cm−1) optical spectrum of the tungsten cofactor. The steady-state data are consistent with a substrate substituted-enzyme mechanism for three substrates (formaldehyde plus two ferredoxin molecules). The K M value for free formaldehyde (21 μM) with ferredoxin as an electron acceptor is approximately 3 times lower than the value measured when benzyl viologen is used as an acceptor. The K M of ferredoxin (14 μM) is an order of magnitude less than previously reported values. An explanation for this discrepancy may be the fact that high concentrations of substrate are inhibitory and denaturing to the enzyme. Pre-steady-state difference spectra reveal peak shifts and a lack of isosbestic points, an indication that several processes happen in the first seconds of the reaction. Two fast processes (k obs1 = 4.7 s−1, k obs2 = 1.9 s−1) are interpreted as oxidation of the substrate followed by rearrangement of the active site. Alternatively, these processes could be the entry/binding of the substrate followed by its oxidation. The release of the product and the electron shuffling over the tungsten and iron–sulfur center in the absence of an external electron acceptor are slower (k obs3 = 6.10 × 10−2 s−1, k obs4 = 2.18 × 10−2 s−1). On the basis of these results in combination with results from previous electron paramagnetic resonance studies an activation route plus catalytic redox cycle is proposed.  相似文献   

5.
The objective of this work was to determine (1) the effect of rotational speed (N) and lifters on the oxygen transfer coefficient (k L) of a mineral solution and (2) the effect of solids concentration of a slurry soil-mineral solution on k L, at a fixed value N (0.25 s−1); in both cases the treatment was carried out in an aerated rotating drum reactor (RDR) operated at atmospheric pressure. First, the k L for the mineral solution was in the range 6.38 × 10−4–7.69 × 10−4 m s−1, which was of the same order of magnitude as those calculated for closed rotating drums supplied with air flow. In general, k L of RDR implemented with lifters was superior or equal to that of RDR without lifters. For RDR implemented with lifters, k L increased with N in the range 6.65 × 10−4–10.51 × 10−4 m s−1, whereas k L of RDR without lifters first increased with N up to N = 0.102 s−1, and decreased beyond this point. Second, regarding soil slurry experiments, an abrupt fall of k L (ca. 50%) at low values of the solid concentration (C v) and an asymptotic pattern at high C v were observed at N = 0.25 s−1. These results suggest that mass transfer phenomena were commanded by the slurry properties and a semi-empirical equation of the form Sh = f(Re, Sc) seems to corroborate this finding.  相似文献   

6.
Kinetic analysis of binding of [3H][N-[2-[4-(2-[O-methyl-3H]methoxyphenyl)-1-piperazinyl]ethyl]-N-(2-pyridinyl)cyclohexane carboxamide ([3H]WAY100635) to 5-HT1A receptors in rat hippocampal membranes has revealed complex regulation mechanism for this radioligand. Saturation binding experiments revealed that [3H]WAY100635 binds to a single class of receptors with very high apparent affinity (K D = 87 ± 4 pM, B max = 15.1 ± 0.2 fmol/mg protein). The binding was almost irreversible, as the dissociation rate constant obtained k off = (7.8 ± 1.1) × 10−3 min−1, means that equilibrium with this radioligand cannot be achieved before 7.5 h incubation at 25°C. Systematic association kinetic studies of [3H]WAY100635 binding revealed sharp reaction acceleration at higher radioligand concentration, proposing mechanism of positive cooperativity. The affinities of antagonists determined from competition with [3H]WAY100635 did not coincide with their abilities to inhibit 5-HT-dependent activation of [35S]GTPγS binding probably due to the ligand’s kinetic peculiarities. Thus, [3H]WAY100635 appears to be an excellent tool for determining receptor binding sites, but its applicability in equilibrium studies is strongly limited.  相似文献   

7.
Hwang JW  Choi YB  Park S  Choi CY  Lee EY 《Biodegradation》2007,18(1):91-101
A two-stage reactor system was developed for the continuous degradation of gas-phase trichloroethylene (TCE). Methylosinus trichosporium OB3b was immobilized on activated carbon in a TCE degradation reactor, trickling biofilter (TBF). The TBF was coupled with a continuous stirred tank reactor (CSTR) to allow recirculation of microbial cells from/to the TBF for the reactivation of inactivated cells during TCE degradation. The mass transfer aspect of the TBF was analyzed, and mass transfer coefficient of 3.9 h−1 was estimated. The loss of soluble methane monooxygenase (sMMO) activity was modeled based on a material balance on the CSTR and TBF, and transformation capacity (T c) was determined to be 20.2 mol mg−1. Maximum TCE degradation rate of 525 mg 1−1 d−1 was obtained and reactor has been stably operated for more than 270 days.  相似文献   

8.
The effects of aliphatic hydrocarbons (n-hexadecane andn-dodecane) on the volumetric oxygen mass transfer coefficient (k L a) were studied in flat alveolar airlift reactor and continuous stirred tank reactors (CSTRs). In the flat alveolar airlift reactor, high aeration rates (>2 vvm) were required in order to obtain efficient organic-aqueous phase dispersion and reliablek L a measurements. Addition of 1% (v/v)n-hexadecane orn-dodecane increased thek l a 1.55-and 1.33-fold, respectively, compared to the control (superficial velocity: 25.8×10−3 m/s, sparger orifice diameter: 0.5 mm). Analysis of the gas-liquid interfacial areaa and the liquid film mass transfer coefficientk L suggests that the observedk L a increase was a function of the media's liquid film mass transfer. Addition of 1% (v/v)n-hexadecane orn-dodecane to analogous setups using CSTRs led to ak L a increase by a factor of 1.68 and 1.36, respectively (superficial velocity: 2.1×10−3 m/s, stirring rate: 250 rpm). These results propose that low-concentration addition of oxygen-vectors to aerobic microbial cultures has additional benefit relative to incubation in purely aqueous media.  相似文献   

9.
The bacteriostatic potency of the cerium-humic acid complex was evaluated by experimental measurement of this complex interaction with E. coli, Bacillus pyocyaneus, Staphylococcus aureus, Leuconostoc and Streptococcus faecalis, and by comparison bacteriostatic effects with the cerium-citrate complex. The experimental results indicated that the cerium-humic acid complex strongly inhibited growth of all five bacterial strains, and its diameter of bacteriostatic circles were more than 30 mm. The minimal bacteria-inhibiting concentration were 1×10−3, 2×10−3 and 1×10−2 mol/L for E. coli and Bacillus pyocyaneus, Staphylococcus aureus, and Leuconostoc and Streptococcus faecalis individually, and the measured minimal bactericidal concentrations were 2×10−3 and 1×10−2 mol/L for Bacillus pyocyaneus, E. coli, and Leuconostoc. To kill Staphylococcus aureus and Streptococcus faecalis, the concentration had to be more than 1×10−2 mol/L. On the contrary, we found that cerium-citrate complex did not inhibit the growth of the above five bacteria, but stimulated bacterial growth. The completely different bacteriostatic results of two cerium complexes may hint that the association and chemical properties of the two complexes were different.  相似文献   

10.
This article reports rate constants for thiol–thioester exchange (k ex), and for acid-mediated (k a), base-mediated (k b), and pH-independent (k w) hydrolysis of S-methyl thioacetate and S-phenyl 5-dimethylamino-5-oxo-thiopentanoate—model alkyl and aryl thioalkanoates, respectively—in water. Reactions such as thiol–thioester exchange or aminolysis could have generated molecular complexity on early Earth, but for thioesters to have played important roles in the origin of life, constructive reactions would have needed to compete effectively with hydrolysis under prebiotic conditions. Knowledge of the kinetics of competition between exchange and hydrolysis is also useful in the optimization of systems where exchange is used in applications such as self-assembly or reversible binding. For the alkyl thioester S-methyl thioacetate, which has been synthesized in simulated prebiotic hydrothermal vents, k a = 1.5 × 10−5 M−1 s−1, k b = 1.6 × 10−1 M−1 s−1, and k w = 3.6 × 10−8 s−1. At pH 7 and 23°C, the half-life for hydrolysis is 155 days. The second-order rate constant for thiol–thioester exchange between S-methyl thioacetate and 2-sulfonatoethanethiolate is k ex = 1.7 M−1 s−1. At pH 7 and 23°C, with [R″S(H)] = 1 mM, the half-life of the exchange reaction is 38 h. These results confirm that conditions (pH, temperature, pK a of the thiol) exist where prebiotically relevant thioesters can survive hydrolysis in water for long periods of time and rates of thiol–thioester exchange exceed those of hydrolysis by several orders of magnitude.  相似文献   

11.
Spent sulfidic caustic was applied to sulfur utilizing autotrophic denitrification as the simultaneous source of electron donor and alkalinity. The two experiment set-up of upflow anoxic hybrid growth reactor (UAHGR) and upflow anoxic suspended growth reactor (UASGR) was adopted and nitrate removals were similar in both reactors. Approximately 90% of the initial nitrate was denitrified at nitrate loading rate of 0.15∼0.40 kgNO3 /m3·d. The experimental stoichiometric ratio of sulfate production to nitrate removal was ranged from 1.5 to 2.1 mgSO4 2−/mgNO3 . During the operation period, denaturing gradient gel electrophoresis (DGGE) analysis of polymerase chain reaction (PCR)-amplified 16S rDNA fragments for the sludge sample of both reactors showed the change of microbial communities. Thiobacillus denitrificans-like microorganism occupied 28.5% (18 clones) of the 63 clones by cloning the PCR products from the sludge sample of UAHGR. Acidovorax avenae, which can reduce nitrate to nitrogen gas while oxidizing phenol (heterotrophic denitrifier), was also found in 7 clones (11.1%). Although an organic carbon source was not added to the medium, a microorganism (Kaistella koreensis) capable of oxidizing organic compounds was found in 7 clones (11.1%). Therefore, the microbial community of spent sulfidic caustic applied autotrophic denitrification process well corresponds to the substrate components of spent sulfidic caustic. Through the batch cultivation of microorganisms in UAHGR, the microbial kinetic coefficients of spent sulfidic caustic applied autotrophic denitrification were estimated to be μ max = 0.097 h−1, k d = 0.0021 h−1, K s = 200 mgNO3 /L, and Y = 0.31 mgMLVSS/mgNO3 .  相似文献   

12.
The aim of the study was to verify the hypothesis if copper could influence the activity of sodium-transporting systems in erythrocyte membrane that could be related to essential hypertension. The examined group of patients consisted of 15 men with hypertension. The control group was 11 healthy male volunteers. The Na+/H+ exchanger (NHE) activity in erythrocytes was determined according to Orlov et al. The activity of transporting systems (ATP-Na+/K+; co-Na+/K+/Cl; ex-Na+/Li+; free Na+ and K+ outflow [Na+, K+-outflow]) was determined according to Garay's method. The concentration of copper in plasma was assessed using atomic absorption spectrometry. The activity of ATP-Na+/K+ (μmol/L red blood cells [RBCs]/h) in hypertensive patients was 2231.5±657.6 vs 1750.5±291 in the control (p<0.05), the activity of co-Na+/K+/Cl (μmol/L RBCs/h) in hypertensives was 171.3±77.9 vs 150.7±53.9 in the control (NS). Na+-outflow (μmol/L RBCs/h) in hypertensives was 118.3±51.6 vs 113.3±24.4 in the control (NS). The K+-outflow (μmol/L RBCs/h) in hypertensives was 1361.7±545.4 vs 1035.6±188.3 in the control (NS). The activity of ex-Na+/Li+ (μmol/L RBCs/h) in hypertensive patients was 266.1±76.1 vs 204.1±71.6 in the control (p<0.05). NHE activity (mmol/L RBCs/h) in hypertensives was 9.7±2.96 vs 7.7±1.33 in the control (p<0.05). In hypertensive patients, negative correlation was found between the activity of Na+/K+/Cl co-transport and plasma copper concentration (R s=−0.579, p <0.05) and between the activity of ex-Na+/Li+ and plasma copper concentration (R s=−0.508, p<0.05). Plasma copper concentration significantly influences the activity of sodium transporting systems in erythrocyte membrane. Copper supplementation could be expected to provide therapeutic benefits for hypertensive patients.  相似文献   

13.
Both HA-CdS and HB-CdS (Hys-CdS, Hys represents HA, HB) complex systems were established according to the dynamics of heterogeneous electron-transfer process m = ES* /S+ - E < 0\mu = E_{S^* /S^ + } - E< 0 . In these systems, the electron transferring from1Hys* to conduction band of CdS is feasible. Determined from the fluorescence quenching, the apparent association constants (Kapp) between Hypocrellin A (HA), Hypocrellin B. (HB) and CdS sol. were about 940 (mol/L)−1, 934 (mol/L)−1, respectively. Fluorescence lifetime measurements gave the rate constant for the electron transfer process from1HA*,1HB* into conduction band of CdS semiconductor as 5.16 × 109 s−1, 5.10 × 109 s−1, respectively. TEMPO (2,2,6,6-tetramethy-1-piperdinyloxy), a stable nitroxide radical, was used in the kinetic study of the reduction reaction taking place on the surface of a CdS colloidal semiconductor, kinetics equation of the reaction was determined with the electron paramagnetic resonance (EPR) method, and the reaction order of TEMPO is zero. When Hys were added, the rate of EPR increased greatly. By comparing rate constants, the Hys-CdS systems were revealed to be about 350 times more efficient than CdS sol. alone in the photoreduction of TEMPO under visible light. It suggests that Hys can be used as efficient sensitizers of a colloidal semiconductor in the application of solar energy.  相似文献   

14.
By means of microcalorimetry, the effect of four copper(II) complexes on Tetrahymena growth was investigated. The extent and duration of the inhibitory effect on the metabolism, judged by the rate constant, k, and the half inhibition concentration, IC50, varied with the different complexes. The results showed that the half inhibition concentrations IC50 of CuCl2, (C9H6NO)2Cu and [Cu(phen)2]Cl2⋅6H2O were 9.9 × 10−4, 2.0 × 10−4, and 2.6 × 10−4 mol/L, respectively. The sequence of antibiotic activity of these three complexes was: (C9H6NO)2Cu > [Cu(phen)2]Cl2⋅6H2O > CuCl2. The growth rate constants of [Cu(phen)3]Cl2⋅6H2O did not change obviously with the increase of concentrations, but [Cu(phen)3]Cl2⋅6H2O also can prolong the time of Tetrahymena growth.  相似文献   

15.
Kim HT  Ko HJ  Kim N  Kim D  Lee D  Choi IG  Woo HC  Kim MD  Kim KH 《Biotechnology letters》2012,34(6):1087-1092
A gene, alg7D, from Saccharophagus degradans, coding for a putative alginate lyase belonging to the family of polysaccharide lyase-7, was overexpressed in Escherichia coli. The properties of the recombinant Alg7D were characterized. The enzyme endolytically depolymerized alginate by β-elimination into oligo-alginates with degrees of polymerization of 2–5. Its activity was maximal at 50°C and pH 7 and was slightly increased in the presence of Na+. The K M , V max , k cat , and k cat /K M values were: 3 mg ml−1, 6.2 U mg−1, 1.9 × 10−2 s−1, and 6.3 × 10−3 mg−1 ml s−1, respectively.  相似文献   

16.
A medicinal mushroom, Phellinus linteus, was successfully cultivated using a cheese-processing waste, whey, and the optimal bioconversion conditions for the maximum mycelial growth rate was also estimated through solid-state cultivation experiments. Response surface analysis with a face-centered design (center point replication = 5) was applied to statistically approximate the simultaneous effects of the three variables, i.e., substrate concentration (10–30 g lactose l−1), temperature (20–30°C), and pH (4–6), on the mycelial growth rate of P. linteus. The following is a partial cubic model where η is the mycelial growth rate (K r ) and x k is the corresponding variable term (k = substrate concentration, temperature, and pH in order): η = −23.8 + 8.67 × 10−2 x 1 + 1.48x 2 + 1.77x 3 + 8.00 × 10−4 x 1 x 2 + 7.25 × 10−2 x 1 x 3 + 5.13 × 10−2 x 2 x 3 −1.28 × 10−2 x 12 –3.18 × 10−2 x 22. −2.64 × 10−1 x 32 −3.28 × 10−3 x 1 x 2 x 3 + 4.68 × 10−4 x 12 x 2. The produced response surface model proved to be significant (r 2 > 0.99, P-value <0.0001, coefficient of variation <5%) to describe the explored space. Temperature was found to be the most significant factor of dominant effects on the mycelial growth rate, and other variables such as temperature2, pH, pH2, and (substrate concentration2 × temperature) also showed significant effects on the model output. The maximum mycelial growth rate was predicted to be 2.80 mm d−1 at 29.7 g lactose l−1, 26.2°C, and pH 5. Our results proved a good potential of whey to serve as an alternative growth medium for cultivating P. linteus mycelia. This may provide another potential for managing this nutrient-rich waste in a cost-effective way.  相似文献   

17.
The relationship between scale and body growth for emigrating Atlantic salmon, Salmo salar, smolts was previously not understood and therefore was examined in this study using mark-recapture techniques. The size of smolts at time of recapture was significantly greater than when marked (P = 0.0002). The growth in length of smolts emigrating 5 km over an average of 20 days was 7.7 ± 6.1 mm per day. Instantaneous somatic growth (G body) ranged from 7.0 × 10−4 to 5.1 × 10−3 (mean = 2.7 × 10−3 ± 1.3 × 10−3). The mean number of plus growth circuli present per scale was significantly greater for smolts when recaptured compared to when marked (P = 0.0014). The instantaneous growth rate of scales (G scale) ranged from 1.4 × 10−3 to 11.5 × 10−3 (mean = 6.6 × 10−3 ± 4.3 × 10−3). The relationship between body size and scale radius showed positive allometry rather than isometry. The relationship of G scale with G body showed positive allometry indicating that scales grew at a slightly faster rate than the body during the emigratory period.  相似文献   

18.
Extracellular secretion of lignin peroxidase from Pycnoporus sanguineus MTCC-137 in the liquid culture growth medium amended with lignin containing natural sources has been shown. The maximum secretion of lignin peroxidase has been found in the presence of saw dust. The enzyme has been purified to homogeneity from the culture filtrate of the fungus using ultrafiltration and anion exchange chromatography on DEAE-cellulose. The purified lignin peroxidase gave a single protein band in sodium dodecylsulphate polyacrylamide gel electrophoresis corresponding to the molecular mass 40 kDa. The K m, k cat and k cat/K m values of the enzyme using veratryl alcohol and H2O2 as the substrate were 61 M, 2.13 s−1, 3.5 × 104 M−1s−1 and 71 M, 2.13 s−1, 3.0 × 104 M−1 s−1 respectively at the optimum pH of 2.5. The temperature optimum of the enzyme was 25°C.  相似文献   

19.
The harmful effects of surfactants to the environment are well known. We were interested in investigating their potential toxicity in a pure culture of Acinetobacter junii, a phosphate (P)-accumulating bacterium. Results showed a high acute toxicity of sodium dodecyl sulfate (SDS) and hexadecyltrimethylammonium bromide (HDTMA) against A. junii. The estimated EC50 values of the HDTMA for the inhibition of CFUs in the pure culture of A. junii was 3.27 ± 1.12 × 10−7 mol L−1 and for the inhibition of the P-uptake rates 2.47 ± 0.51 × 10−6 mol L−1. For SDS, estimated EC50 values for the inhibition of CFUs in the pure culture of A. junii was 5.00 ± 2.95 × 10−6 mol L−1 and for the inhibition of the P-uptake rates 3.33 ± 0.96 × 10−4 mol L−1. The obtained EC50 values in the standardised yeast toxicity test using Saccharomyces cerevisiae were 3.03 ± 0.38 × 10−4 and 4.33 ± 0.32 × 10−5 mol L−1 for SDS and HDTMA, respectively. These results emphasized the need to control concentrations of surfactants entering the activated sludge system. The negative effects of these toxicants could greatly decrease populations of P-accumulating bacteria, as well as eukaryotic organisms, inhabiting activated sludge systems, which in turn could result in the decrease of the system efficiency.  相似文献   

20.
Gamma linolenic acid (GLA) degradation in Spirulina followed first-order reaction kinetics. At an accelerated temperature range of 45 to 55°C, the degradation rate constants (k r) of GLA obtained were 4.0 × 10−2 to 8.8 × 10−2 day−1. The energy of activation (E a) was 16.53 kcal mol−1, and the Q10 was 2.22. Based on 20% GLA degradation, the shelf life of sun-dried Spirulina at 30°C is 263 days or 8.6 months using the Arrhenius plot, and 258 days or 8.5 months using the Q 10 approach. Presented at the 6th Meeting of the Asia Pacific Society of Applied Phycology, Manila, Philippines.  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