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 46 毫秒
1.
The influence of stem lacunar structure on the potential of diffusion and mass flow to meet estimated root O2 demands was evaluated and compared in four submersed aquatic plant species. Internodal lacunae formed large continuous gas canals which were constricted at the nodes by thin, perforated diaphragms. Gas transport studies showed that nodes had little effect on diffusion, but significantly reduced mass flow. Measured diffusive resistances approximated those predicted by Fick's first law, ranged from 203 to 5107 × 108 s m−4 and increased as lacunar area decreased in Potamogeton praelongus, two Myriophyllum species and Elodea canadensis. Our analysis suggested that diffusion could satisfy estimated root O2 demands given the development of relatively steep O2 gradients (0.15–0.35 mol O2 mor−1 per 0.5 m stem) between shoots and roots. Plants with high resistances (e.g. > 750 × 108 s m−4) and long lacunar pathlengths may be unable, even during active photosynthesis, to support the O2 demands of a large root system by diffusion alone. Measured nodal resistances to mass flow approximated those predicted by Hagen-Poiseuille law and ranged from 46 to 2029 × 108 Pa s m−3. Our analysis suggested that these resistances were quite low and that relatively small pressure differentials (< 150 Pa per 0.5 m stem) could drive mass flow at rates which would support root O2 demands. Possible mechanisms whereby plant architecture may serve to maintain these pressure differentials are proposed.  相似文献   

2.
The performance of a laboratory-scale sewage treatment system composed of an up-flow anaerobic sludge blanket (UASB) reactor and a moving bed biofilm reactor (MBBR) at a temperature of (22–35 °C) was evaluated. The entire treatment system was operated at different hydraulic retention times (HRT’s) of 13.3, 10 and 5.0 h. An overall reduction of 80–86% for CODtotal; 51–73% for CODcolloidal and 20–55% for CODsoluble was found at a total HRT of 5–10 h, respectively. By prolonging the HRT to 13.3 h, the removal efficiencies of CODtotal, CODcolloidal and CODsoluble increased up to 92, 89 and 80%, respectively. However, the removal efficiency of CODsuspended in the combined system remained unaffected when increasing the total HRT from 5 to 10 h and from 10 to 13.3 h. This indicates that, the removal of CODsuspended was independent on the imposed HRT. Ammonia-nitrogen removal in MBBR treating UASB reactor effluent was significantly influenced by organic loading rate (OLR). 62% of ammonia was eliminated at OLR of 4.6 g COD m−2 day−1. The removal efficiency was decreased by a value of 34 and 43% at a higher OLR’s of 7.4 and 17.8 g COD m−2 day−1, respectively. The mean overall residual counts of faecal coliform in the final effluent were 8.9 × 104 MPN per 100 ml at a HRT of 13.3 h, 4.9 × 105 MPN per 100 ml at a HRT of 10 h and 9.4 × 105 MPN per 100 ml at a HRT of 5.0 h, corresponding to overall log10 reduction of 2.3, 1.4 and 0.7, respectively. The discharged sludge from UASB–MBBR exerts an excellent settling property. Moreover, the mean value of the net sludge yield was only 6% in UASB reactor and 7% in the MBBR of the total influent COD at a total HRT of 13.3 h. Accordingly, the use of the combined UASB–MBBR system for sewage treatment is recommended at a total HRT of 13.3 h.  相似文献   

3.
We have used two different approaches to determine hydrodynamic parameters for mucins secreted by guinea-pig tracheal epithelial cells in primary culture. Cells were cultured under conditions that promote mucous cell differentiation. Secreted mucins were isolated as the excluded fraction from a Sepharose CL-4B gel filtration column run under strongly dissociating conditions. Biochemical analysis confirmed the identity of the high molecular weight material as mucins. Analytical ultracentrifugation was used to study the physical properties of the purified mucins. The weight average molecular mass (M w ) for three different preparations ranged from 3.3×106 to 4.7×106 g/mol (corresponding to an average structure of 1 – 2 subunits), and the sedimentation coefficient from 25.5 to 35 S. Diffusion coefficients ranging from 4.5×10–8 to 6.4×10–8 cm2/s were calculated using the Svedberg equation. A polydispersity index (M z /M w ) of ∼1.4 was obtained. Diffusivity values were also determined by image analysis of mucin granule exocytosis captured by videomicroscopy. The time course of hydration and dissolution of mucin was measured and a relationship is presented which models both phases, each with first order kinetics, in terms of a maximum radius and rate constants for hydration and dissolution. A median diffusivity value of 8.05×10–8 cm2/s (inter-quartile range = 1.11×10–7 to 6.08×10–8 cm2/sec) was determined for the hydration phase. For the dissolution phase, a median diffusivity value of 6.98×10–9 cm2/s (inter-quartile range = 1.47×10–8 to 3.25×10–9 cm2/sec) was determined. These values were compared with the macromolecular diffusion coefficients (D 20,w ) obtained by analytical ultracentrifugation. When differences in temperature and viscosity were taken into account, the resulting D 37,g was within the range of diffusivity values for dissolution. Our findings show that the physicochemical properties of mucins secreted by cultured guinea-pig tracheal epithelial cells are similar to those of mucins of the single or double subunit type purified from respiratory mucus or sputum. These data also suggest that measurement of the diffusivity of dissolution may be a useful means to estimate the diffusion coefficient of mucins in mucus gel at the time of exocytosis from a secretory cell. Received: 10 March 1998 / Accepted: 27 March 1998  相似文献   

4.
Summary The effect of low concentrations of hydrogen peroxide (H2O2) (5 × 10−7−9.5 × 10−7 M) on cell growth and antibody production was investigated with murine hybridoma cells (Mark 3 and anti-hPL) in culture. Cell growth, measured by flow cytometry with morphological parameters, was significantly stimulated by H2O2 (8 × 10−7 M) but H2O2 concentration of 7 × 10−6 M and above increased cell death. H2O2 stimulation of antibody production was nonsignificant. The metabolism of cells treated with 8 × 10−7 or 1 × 10−5 M H2O2 was similar to that of the control in terms of glucose and glutamine consumption, lactate and ammonia production, and amino acid concentrations in the medium. The concentrations of lactate dehydrogenase, a marker of cell death, in test and control cells were similar. However, concentrations of intracellular free radicals measured by flow cytometry with dihydrorhodamine 123 (DHR 123) and dichlorofluorescein diacetate (DCFH-DA) as fluorochromes were different. The reactive oxygen species content of cells in 8 × 10−7 M H2O2 was similar to that of the controls, but there was a sudden, marked production of superoxide anions (detected with DHR 123) and H2O2 or peroxides (detected with DCFH-DA) by cells incubated with 1 × 10−5 M H2O2 which increased with increasing H2O2 until cell death.  相似文献   

5.
The effects of a novel nonpeptide NK1 tachy-kinin receptor antagonist, SR 140333, on the functional consequences of NK1 receptor activation in a human astrocytoma cell line, U373MG, were investigated. Radioligand binding conducted with 125l-Bolton-Hunter substance P revealed a competitive inhibition by SR 140333 and its R enantiomer SR 140603 with Ki values of 0.74 and 7.40 nM, respectively. The NK1-selective agonist, [Sar9,Met(O2)11]-substance P, stimulated the formation of inositol phosphates with an EC50 of 3.8 × 10?9M. SR 140333 blocked the stimulatory effect of this agonist (10?7M) with an IC50 of 1.6 × 10?9M,whereas the effect of another NK1 agonist, septide (EC50= 1.5 × 10?8M)was antagonized with an IC50 of 2.1 × 10?10M.Enhancement of [3H]taurine release by [Sar9,Met(O2)11]-substance P (EC50= 7.4 × 10?9M) was also inhibited by SR 140333 with an IC50 of 1.8 × 10?9 M. SR 140603 was 10-fold less potent than SR 140333 in inhibiting inositol monophosphate formation and [3H]taurine release. The calcium mobilization induced by [Sar9,Met(O2)11]-substance P (10?8M) was totally prevented by 10?8MSR 140333. Patchclamp experiments showed that SR 140333 depressed the outward current evoked by 5 × 10?8M [Sar9, Met(O2)11]-substance P with an IC50 of 1.3 × 10?9M. The expression of c-fos was stimulated by [Sar9,Met(O2)11]-substance P with an EC50 of 2.5 × 10?10M, an effect that was also inhibited by SR 140333 with an IC50 of 1.1 × 10?9M. The present results illustrate the sequential events of the response elicited by NK1 agonists, which were antagonized by SR 140333, demonstrating its powerful NK1 antagonist activity on a functional basis.  相似文献   

6.
Direct electron transfer of hemoglobin modified with quantum dots (QDs) (CdS) has been performed at a normal graphite electrode. The response current is linearly dependent on the scan rate, indicating the direct electrochemistry of hemoglobin in that case is a surface-controlled electrode process. UV–vis spectra suggest that the conformation of hemoglobin modified with CdS is little different from that of hemoglobin alone, and the conformation changes reversibly in the pH range 3.0–10.0. The hemoglobin in a QD film can retain its bioactivity and the modified electrode can work as a hydrogen peroxide biosensor because of its peroxidase-like activity. This biosensor shows an excellent response to the reduction of H2O2 without the aid of an electron mediator. The catalytic current shows a linear dependence on the concentration of H2O2 in the range 5 × 10−7–3 × 10−4 M with a detection limit of 6 × 10−8 M. The response shows Michaelis–Menten behavior at higher H2O2 concentrations and the apparent Michaelis–Menten constant is estimated to be 112 μM.  相似文献   

7.
Polyamine oxidase from Penicillium chrysogenum oxidized spermine rapidly and spermidine slightly at pH 7.5. The apparent Km values for spermine and spermidine were calculated to be 2.25 × 10?5 m and 9.54 × 10?6 m, respectively. The relative maximum velocities for spermine and spermidine were 3.37 × 10?3 m (H2O2) per min per mg of protein and 2.08 × 10?4 m (H2O2) per min per mg of protein, respectively. Spermine oxidation of the enzyme was competitively inhibited by spermidine and putrescine. The apparent Ki values by spermidine and putrescine were calculated to be 3.00 × 10?5 m and 1.80 × 10?8 m, respectively. On the other hand, polyamine oxidase from Aspergillus terreus rapidly oxidized both spermidine and spermine at pH 6.5. The apparent Km values for spermidine and spermine were 1.20 × 10?8 m and 5.37 × 10?7 m, respectively. The relative maximum velocities for spermidine and spermine were 1.55 × 10?2 m (H2O2) per min per mg of protein and 6.20 × 10?3 m (H2O2) per min per mg of protein, respectively.

Differential determination of spermine and spermidine was carried out using the two enzymes. The initial rate was assayed with Penicillium enzyme and the end point was measured afte addition of Aspergillus enzyme. Small amounts of polyamines (25 to 200 nmol of spermine and 25 to 250 nmol of spermidine) were assayed by solving two simultaneous equations obtained from the rate assay method and the end point assay method. The calculated values were in close agreement with those obtained by an amino-acid analyzer.  相似文献   

8.
The quenching of singlet molecular oxygen (1O2) by the flavylium cation form of six widespread anthocyanin derivatives: cyanidin 3-glucoside (CG), cyanidin 3-rutinoside (CR), cyanidin 3-galactoside (CGL), malvidin (M), malvidin 3-glucoside (MG) and malvidin 3,5-diglucoside (MDG) was studied in 1% HCl methanol solution by time-resolved phosphorescence detection (TRPD) of 1O2 and photostationary actinometry using perinaphthenone and methylene blue as sensitizers, respectively. The average value of the total (k0) and chemical (kc) quenching rate constants were ~ 4×108 m?1 s?1 and 3×106 m?1 s?1, respectively, indicating the good performance of the studied anthocyanins as catalytic quenchers of 1O2. The quenching efficiency was larger for malvidin than for cyanidin derivatives, probably by the extra electron-donating methoxy group in ring B of the malvidin derivatives; and it was also dependent on the number and type of glycosylated substitution. As observed for other phenolic-like derivatives, the quenching of 1O2 by anthocyanins was mediated by a charge-transfer mechanism, which was modulated by the total number of –OR substituents that increases the electron-donating ability of these compounds.  相似文献   

9.
 DNA binding by trans-[(H2O)(Pyr)(NH3)4RuII]2+ (Pyr=py, 3-phpy, 4-phpy, 3-bnpy, 4-bnpy) is highly selective for G7 with K G=1.1×104 to 2.8×104, with the more hydrophobic Pyr ligands exhibiting slightly higher binding. A strong dependence on ionic strength indicates that ion-pairing with DNA occurs prior to binding. At μ=0.05, d[RuII-DNA]/dt=k[RuII][DNA], where k=0.17–0.21 M–1 s–1 with the various Pyr ligands. The air oxidation of [(py)(NH3)4RuII] n -DNA to [(py)(NH3)4RuIII] n -DNA at pH 6 occurs with a pseudo-first-order rate constant of k obs=5.6×10–4 s–1 at μ=0.1, T=25  °C. Strand cleavage of plasmid DNA appears to occur by both Fenton/Haber-Weiss chemistry and by base-catalyzed routes, some of which are independent of oxygen. Base-catalyzed cleavage is more efficient than O2 activation at neutral pH and involves the disproportionation of covalently bound RuIII and, in the presence of O2, Ru-facilitated autoxidation to 8-oxoguanine. Disproportionation of [py(NH3)4RuIII] n -DNA occurs according to the rate law: d[RuII–GDNA]/dt=k 0[RuIII–GDNA]+k 1[RuIII–GDNA][OH], where k 0=5.4×10–4 s–1 and k 1=8.8 M–1 s–1 at 25  °C, μ=0.1. The appearance of [(Gua)(py)(NH3)4RuIII] under argon, which occurs according to the rate law: d[RuIII–G]/dt=k 0[RuIII–GDNA]+k 1[OH][RuIII–GDNA] (k 0=5.74×10–5 s–1, k 1=1.93×10–2 M–1 s–1 at T=25  °C, μ=0.1), is consistent with lysis of the N-glycosidic bond by RuIV-induced general acid hydrolysis. In air, the ratio of [Ru-8-OG]/[Ru-G] and their net rates of appearance are 1.7 at pH 11, 25  °C. Small amounts of phosphate glycolate indicate a minor oxidative pathway involving C4′ of the sugar. In air, a dynamic steady-state system arises in which reduction of RuIV produces additional RuII. Received: 11 November 1998 / Accepted: 3 March 1999  相似文献   

10.
The ventilatory equivalent for CO2 defines ventilatory efficiency largely independent of metabolism. An impairment of ventilatory efficiency may be caused by an increase in either anatomical or physiological dead space, the latter being the most important mechanism in the hyperpnoea of heart failure, pulmonary embolism, pulmonary hypertension and the former in restrictive lung disease. However, normal values for ventilatory efficiency have not yet been established. We investigated 101 (56 men) healthy volunteers, aged 16–75 years, measuring ventilation and gas exchange at rest (n = 64) and on exercise (modified Naughton protocol, n = 101). Age and sex dependent normal values for ventilatory efficiency at rest defined as the ratio ventilation:carbon dioxide output ( E:CO2), exercise ventilatory efficiency during exercise, defined as the slope of the linear relationship between ventilation and carbon dioxide output ( E vs CO2 slope), oxygen uptake at the anaerobic threshold and at maximum (O2AT,O2max, respectively) and breathing reserve were established. Ventilatory efficiency at rest was largely independent of age, but was smaller in the men than in the women [ E:CO2 50.5 (SD 8.8) vs 57.6 (SD 12.6) P<0.05]. Ventilatory efficiency during exercise declined significantly with age and was smaller in the men than in the women (men: ( E vs CO2 slope = 0.13 × age + 19.9; women: E vs CO2 slope = 0.12 × age + 24.4). The O2AT and O2max were 23 (SD 5) and 39 (SD 7) ml O2 · kg · min−1 in the men and 18 (SD 4) and 32 (SD 7) in the women, respectively, and declined significantly with age. The O2AT was reached at 58 (SD 9)% O2max. Breathing reserve at the end of exercise was 41% and was independent of sex and age. It was concluded from this study that ventilatory efficiency as well as peak oxygen uptake are age and sex dependent in adults. Accepted: 11 June 1997  相似文献   

11.
Diversity and activity of aquatic fungi under low oxygen conditions   总被引:1,自引:0,他引:1  
1. The objective was to test whether a decrease in oxygen concentration in streams affects the diversity and activity of aquatic hyphomycetes and consequently leaf litter decomposition. 2. Senescent leaves of Alnus glutinosa were immersed for 7 days in a reference stream, for fungal colonization, and then incubated for 18 days in microcosms at five oxygen concentrations (4%, 26%, 54%, 76% and 94% saturation). Leaf decomposition (as loss of leaf toughness), fungal diversity, reproduction (as spore production) and biomass (ergosterol content) were determined. 3. Leaf toughness decreased by 70% in leaves exposed to the highest O2 concentration, whereas the decrease was substantially less (from 25% to 45%) in treatments with lower O2. Fungal biomass decreased from 99 to 12 mg fungi g−1 ash‐free dry mass on exposure to 94% and 4% O2 respectively. Sporulation was strongly inhibited by reduction of dissolved O2 in water (3.1 × 104 versus 1.3 × 103 spores per microcosms) for 94% and 4% saturation respectively. 4. A total of 20 species of aquatic hyphomycetes were identified on leaves exposed to 94% O2, whereas only 12 species were found in the treatment with 4% O2 saturation. Multidimensional scaling revealed that fungal assemblages exposed to 4% O2 were separated from all the others. Articulospora tetracladia, Cylindrocarpon sp. and Flagellospora curta were the dominant species in microcosms with 4% O2, while Flagellospora curvula and Anguillospora filiformis were dominant at higher O2 concentrations. 5. Overall results suggest that the functional role of aquatic hyphomycetes as decomposers of leaf litter is limited when the concentration of dissolved oxygen in streams is low.  相似文献   

12.
 Dithionite has been found to reduce directly (without mediators) the Escherichia coli R2 subunit of ribonucleotide reductase. With dithionite (∼10 mM) in large excess, the reaction at 25  °C is complete in ∼10 h. Preparations of E. coli R2 have an FeIII 2 (met-R2) component in this work at ∼40% levels, alongside the fully active enzyme FeIII 2 . . . Tyr*, which has a tyrosyl radical at Tyr-122. In the pH range studied (7–8) the kinetics are biphasic. Rate laws for both phases give [S2O4 2–] and not [S2O4 2–]1/2 dependencies, and saturation kinetics are observed for the first time in R2 studies. No dependence on pH was detected. The kinetics (25  °C) of the first phase are reproduced in separate experiments using only met-R2, with association of S2O4 2– to met-R2, K=330 M–1, occurring prior to electron transfer, k et=4.8×10–4 s–1, I=0.100 M (NaCl). The second phase assigned to the reaction of FeIII 2 . . . Tyr* with S2O4 2– gives K=800 M–1 and k et=5.6×10–5 s–1. Bearing in mind the substantially smaller reduction potential for FeIII 2 compared to Tyr*, this is a quite remarkable finding, with implications similar to those already reported for the reaction of R2 with hydrazine, but with additional information provided by the saturation kinetics. The similarity in rates for the two phases (∼fourfold difference) suggests that reduction of FeIII 2 is occurring in both cases, and since S2O4 2– is involved a two-equivalent change is proposed with the formation of FeII 2 . . . Tyr* in the case of active R2. As a sequel to the second phase, intramolecular reduction of the strongly oxidising Tyr* by the FeII 2 is rapid, and further decay of FeIIFeIII is also fast. There is no stable mouse met-R2 form, and the single-phase reaction with dithionite gives saturation kinetics with K=208 M–1 and k et=1.7±10–3 s–1. Mechanistic implications, including the applicability of a pathway for electron transfer via FeA, are considered. Received: 25 February 1998 / Received: 20 August 1998  相似文献   

13.
Based on the inhibition effect of methimazole (MMI) on the reaction of luminol–H2O2 catalyzed by gold nanoparticles, a novel chemiluminescence (CL) method was developed for the determination of MMI. Under the optimum conditions, the relative CL intensity was linearly related to MMI concentration in the range from 5.0 × 10?8 to 5.0 × 10?5 mol L?1. The detection limit was 1.6 × 10?8 mol L?1 (S/N = 3), and the RSD for 6.0 × 10?6 mol L?1 MMI was 4.83 (n = 11). This method has high sensitivity, wide linear range, inexpensive instrumentation and has been applied to detect MMI in pharmaceutical tablets and pig serum samples. Furthermore, a possible reaction mechanism is discussed. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

14.
Kinetics of 1-hydroxypyrene (1-HP) oxidation catalyzed with recombinant Coprinus cinereus (rCiP) and horseradish (HRP) peroxidases was investigated with a special emphasis for developing a nanomolar hydrogen peroxide (H2O2) detection system. At pH 8.0 the bimolecular constants of 1-HP oxidation with the ferryl compounds of rCiP and HRP were equal to (1.0 ± 0.3) × 108 M−1 s−1 and (0.6 ± 0.2) × 108 M−1 s−1, respectively. High bimolecular constants and fluorescence quantum yield of 1-HP (0.66) permitted detection as low as 21 nM of H2O2. To optimize the detection system 1-HP oxidation was modeled at steady-state conditions in the range pH 5.0 to pH 8.0. The 1-HP based detection system was compared with the Amplex Red system. The peroxidase-catalyzed 1-HP oxidation system was used for determination of ozone in the air.  相似文献   

15.
Ming L  Xi X  Liu J 《Biotechnology letters》2006,28(17):1341-1345
A platinized carbon paste prepared via electrodeposition had a preferential electrocatalytic action toward H2O2. Therefore, we have developed a new amperometric glucose biosensor based on the immobilization of glucose oxidase on to the electrochemically platinized carbon paste. The proposed biosensor is free of potential interferences due to its cathodic detection of glucose at the potential of 0.0 V (vs. Ag/AgCl). It also shows acceptable analytical performance in terms of linearity (6 × 10−5 to 1.2 × 10−2 M, r = 0.998), detection limit (2 × 10−5 M), response time (20–30 s), reproducibility (RSD = 4.4%), and storage life (t 0.80 = 45 days). All these advantages of the biosensor raise potential possibilities for its medical or other biotechnical applications.  相似文献   

16.
Genome size has been studied for the first time in the Colchicum genus. Values obtained by flow cytometry were quite stable and specific to each taxon: C. autumnale L., 2C=5.89±0.22 pg, equivalent to 5.7×10 bp (2 n=4 x=36); C. alpinum DC., 2C=8.06±0.24 pg, equivalent to 7.8×10 bp (2 n=6 x=56); C. lusitanum Brot., 2C=10.7±0.67 pg, equivalent to 10.3×10 bp (2 n= approx. 10 x=90–92 and 94–96); C. multiflorum Brot., 2C=16.5±0.69 pg, equivalent to 15.9×10 bp, (2 n= approx. 16 x=140–148); C. corsicum Baker, 2C=21.3±0.99 pg, equivalent to 20.6×10 bp (2 n=22 x= approx. 198±2). These values are well below those published for Liliaceae stricto sensu. In Colchicum species from the western Mediterranean area, genome size was highly correlated with ploidy level ( R 2=0.99, P<0.001). This relationship is consistent with most previous results, with our new chromosome number count for C. corsicum and with our correction of published erroneous counts for C. lusitanum (2 n=102–108). The Calabrian population appeared to be distinct from all of the other plants of the C. alpinum group. Reproducible and accurate, cytometry appears to be a particularly appropriate method for studying this polyploid genus in Western Europe, taking into account that chromosome numbers are difficult to enumerate. It can guide future taxonomic research because it reveals similarities or differences between taxa within this difficult complex.  相似文献   

17.
We discuss the energetics of a cladoceran, Simocephalus vetulus at different temperatures (8.0 ± 1.0, 15.0 ± 1.0, 21.0 ± 1.0 and 28.0 ± 1.0 °C) and food (Chlamydomonas sp.) concentrations (25 × 103, 50 × 103, 75 × 103 and 100 × 103 cells ml−1). Increase in temperature accelerated ingestion and, to some extent, oxygen consumption. The study revealed a high reproduction efficiency in S. vetulus. Net growth efficiency (ECI) was higher (13.17–41.18%) in pre-adults than in adults (2.71–8.40%). The assimilated energy (A) increased with increasing food concentration at all temperatures. Assimilation efficiency (AD) decreased with increasing food concentrations. The energy used for growth (P) was nearly constant at all food levels because the egested energy increased and assimilation efficiency decreased as food concentration increased.  相似文献   

18.
Abstract: Cell and tissue concentrations of NO2? and NO3? are important indicators of nitric oxide synthase activity and crucial in the regulation of many metabolic functions, as well as in nonenzymatic nitric oxide release. We adapted the capillary electrophoresis technique to quantify NO2? and NO3? levels in single identified buccal neurons and ganglia in the opisthobranch mollusc Pleurobranchaea californica, a model system for the study of the chemistry of neuron function. Neurons were injected into a 75-µm separation capillary and the NO2? and NO3? were separated electrophoretically from other anions and detected by direct ultraviolet absorbance. The limits of detection for NO2? and NO3? were <200 fmol (<4 µM in the neurons under study). The NO2? and NO3? levels in individual neurons varied from 2 mM (NO2?) and 12 mM (NO3?) in neurons histochemically positive for NADPH-diaphorase activity down to undetectable levels in many NADPH-diaphorase-negative cells. These results affirm the correspondence of histochemical NADPH-diaphorase activity and nitric oxide synthase in molluscan neurons. NO2? was not detected in whole ganglion homogenates or in hemolymph, whereas hemolymph NO3? averaged 1.8 ± 0.2 × 10?3M. Hemolymph NO3? in Pleurobranchaea was appreciably higher than values measured for the freshwater pulmonate Lymnaea stagnalis (3.2 ± 0.2 × 10?5M) and for another opisthobranch, Aplysia californica (3.6 ± 0.7 × 10?4M). Capillary electrophoresis methods provide utility and convenience for monitoring NO2?/NO3? levels in single cells and small amounts of tissue.  相似文献   

19.
 Kinetics of the steady-state oxidation of n–alkylferrocenes (alkyl = H, Me, Et, Bu and C5H11) by H2O2 to form the corresponding ferricenium cations catalyzed by horseradish peroxidase has been studied in micellar systems of Triton X-100, CTAB, and SDS, mostly at pH 6.0 and 25  °C. The rate of oxidation of ferrocenes with longer alkyl radicals is too slow to be measured. The reaction obeying the [RFc]:[H2O2] = 2 : 1 stoichiometry is strictly first-order in both HRP and RFc in a wide concentration range. The corresponding observed second-order rate constants k, which refer to the interaction of the peroxidase compound II (HRP-II) with RFc, decrease with the elongation of the alkyl substituent R, and this in turn is accompanied by an increase in the formal redox potentials E°′ in the same medium. Increasing the surfactant concentration lowers the rate constants k, the effect being due to the nonproductive binding of RFc to micelles rather than to enzyme inactivation. The micellar effects are accounted for in terms of the Berezin pseudo-phase model of micellar catalysis applied to the interaction of enzyme with organometallic substrates. The oxidation was found to occur primarily in the aqueous pseudo-phase and the calculated intrinsic second-order rate constants k w are (1.9 ± 0.5)×105, (2.7 ± 0.1)×104, and (5.9 ± 0.6)×103 M–1 s–1 for HFc, EtFc, and n–BuFc, respectively. The data obtained were used for estimating the self-exchange rate constants for the HRP-II/HRP couple in terms of the Marcus formalism. Received: 15 July 1996 / Accepted: 15 November 1996  相似文献   

20.
A highly selective and simple chemiluminescence (CL) method for determination of penicillin G potassium (PGK) was developed. In the proposed method, CL was elicited from PGK upon its oxidation with H2O2. The light emission was enhanced in the presence of N‐cetyl‐N,N,N‐trimethylammonium bromide (CTMAB). An experimental design, central composite design (CCD), was used to realize the optimized variables, including pH, surfactant (CTMAB) and H2O2 concentrations. Under optimum condition, the calibration graph was linear in the range 3.3 × 10?3–3.3 × 10?1 mmol/L, with a detection limit of 8.8 × 10?4 mmol/L for PGK. The precision was calculated by analysing samples containing 1.6 × 10?1 mmol/L PGK (n = 5) and the relative standard deviation (RSD) was 1.40%. The utility of this method was demonstrated by determining PGK in pharmaceutical formulations for injection. The proposed method was validated by a reference method. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号