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.
A laccase from the culture filtrate of white rot fungus Daedalea flavida MTCC-145 has been purified and characterized. The method involved concentration of the culture filtrate by ultrafiltration and an anion exchange chromatography on diethylaminoethyl (DEAE) cellulose. The sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) and native polyacrylamide gel electrophoresis (native PAGE) both gave single protein bands indicating that the enzyme preparation was pure. The molecular mass of the enzyme determined from SDS-PAGE analysis was 75.0 kDa. Purification fold was 21.5 while recovery of the enzyme activity was 11.52%. Using 2,6-dimethoxyphenol, diammonium salt of 2,2'-[azino-bis-(3-ethylbenzthiazoline-6-sulfonic acid)] and 3,5-dimethoxy-4-hydroxybenzaldehyde azine as substrates, the Km, kcat, and k cat/K m values of the laccase were found to be 440 µM, 6.45 s–1, 1.47 × 104 M–1 s–1; 366 µM, 6.45 s–1, 1.76 × 104 M–1 s–1; and 226 µM, 6.45 s–1, 2.85 × 104 M–1 s–1, respectively. The pH and temperature optima were 4.5 and 50°C, respectively. The enzyme was most stable at pH 5.0 when exposed for 1 h. The purified laccase has yellow color and shows no absorption band around 610 nm characteristic of blue laccases. The enzyme transforms toluene and substituted toluenes to corresponding benzaldehyde and substituted benzaldehydes in the absence of mediator molecules with higher catalytic efficiency as compared to other known laccases.  相似文献   

2.
Acetic acid bacteria Gluconobacter oxydans subsp. industrius RKM V-1280 were immobilized into a synthetic matrix based on polyvinyl alcohol modified with N-vinylpyrrolidone and used as biocatalysts for the development of bioanodes for microbial fuel cells. The immobilization method did not significantly affect bacterial substrate specificity. Bioanodes based on immobilized bacteria functioned stably for 7 days. The maximum voltage (fuel cell signal) was reached when 100–130 μM of an electron transport mediator, 2,6-dichlorophenolindophenol, was added into the anode compartment. The fuel cell signals reached a maximum at a glucose concentration higher than 6 mM. The power output of the laboratory model of a fuel cell based on the developed bioanode reached 7 mW/m2 with the use of fermentation industry wastes as fuel.  相似文献   

3.
A laccase from the culture filtrate of Phellinus linteus MTCC-1175 has been purified to homogeneity. The method involved concentration of the culture filtrate by ammonium sulphate precipitation and an anion exchange chromatography on DEAE-cellulose. The SDS-PAGE and native-PAGE gave single protein band indicating that the enzyme preparation was pure. The molecular mass of the enzyme determined from SDS-PAGE analysis was 70 kDa. Using 2.6-dimethoxyphenol, 2.2′[azino-bis-(3-ethylbonzthiazoline-6-sulphonic acid) diammonium salt] (ABTS) and 4-hydroxy-3,5-dimethoxybenzaldehyde azine as the substrates, the K m, k cat and k cat/K m values of the laccase were found to be 160 μM, 6.85 s?1, 4.28 × 104 M?1 s?1, 42 μM, 6.85 s?1, 16.3 × 104 M?1 s?1 and 92 μM, 6.85 s?1, 7.44 × 104 M?1 s?1, respectively. The pH and the temperature optima of the P. linteus MTCC-1175 laccase were 5.0 and 45°C, respectively. The activation energy for thermal denaturation of the enzyme was 38.20 kJ/mole/K. The enzyme was the most stable at pH 5.0 after 1 h reaction. In the presence of ABTS as the mediator, the enzyme transformed toluene, 3-nitrotoluene and 4-chlorotoluene to benzaldehyde, 3-nitrobenzaldehyde and 4-chlorobenzaldehyde, respectively.  相似文献   

4.
Although there have been many studies on bacterial removal of soluble azo dyes, much less information is available for biological treatment of water-insoluble azo dyes. The few bacterial species capable of removing Sudan dye generally require a long time to remove low concentrations of insoluble dye particles. The present work examined the efficient removal of Sudan I by Shewanella oneidensis MR-1 in the presence of redox mediator. It was found that the microbially reduced anthraquinone-2,6-disulfonate (AQDS) could abiotically reduce Sudan I, indicating the feasibility of microbially-mediated reduction. The addition of 100 μM AQDS and other different quinone compounds led to 4.3–54.7 % increase in removal efficiencies in 22 h. However, adding 5-hydroxy-1,4-naphthoquinone into the system inhibited Sudan I removal. The presence of 10, 50 and 100 μM AQDS stimulated the removal efficiency in 10 h from 26.4 to 42.8, 54.9 and 64.0 %, respectively. The presence of 300 μM AQDS resulted in an eightfold increase in initial removal rate from 0.19 to 1.52 mg h?1 g?1 cell biomass. A linear relationship was observed between the initial removal rates and AQDS concentrations (0–100 μM). Comparison of Michaelis–Menten kinetic constants revealed the advantage of AQDS-mediated removal over direct reduction. Different species of humic acid could also stimulate the removal of Sudan I. Scanning electronic microscopy analysis confirmed the accelerated removal performance in the presence of AQDS. These results provide a potential method for the efficient removal of insoluble Sudan dye.  相似文献   

5.
6.
Abstract

Myeloperoxidase is very susceptible to reducing radicals because the reduction potential of the ferric/ferrous redox couple is much higher compared with other peroxidases. Semiquinone radicals are known to reduce heme proteins. Therefore, the kinetics and spectra of the reactions of p-hydroquinone, 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone with compounds I and II were investi-gated using both sequential-mixing stopped-flow techniques and conventional spectrophotometric measurements. At pH 7 and 15°C the rate constants for compound I reacting with p-hydroquinone, 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone were determined to be 5.6±0.4×107 M-1s-1, 1.3±0.1×106 M-1s-1 and 3.1±0.3×106 M-1s-1, respectively. The corresponding reaction rates for compound II reduction were calculated to be 4.5±0.3×106 M-1s-1, 1.9±0.1×105 M-1s-1 and 4.5±0.2×104 M-1s-1, respectively. Semiquinone radicals, produced by compounds I and II in the classical peroxidation cycle, promote compound III (oxymyeloperoxidase) formation. We could monitor formation of ferrous myeloperoxidase as well as its direct transition to compound III by addition of molecular oxygen. Formation of ferrous myeloperoxidase is shown to depend strongly on the reduction potential of the corresponding redox couple benzoquinone/semiquinone. With 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone as substrate, myeloperoxidase is extremely quickly trapped as compound III. These MPO-typical features could have potential in designing specific drugs which inhibit the production of hypochlorous acid and consequently attenuate inflammatory tissue damage.  相似文献   

7.
《Journal of bryology》2013,35(1):151-158
Abstract

The concentration of chlorophyll a, b, and total chlorophyll have been monitored on a seasonal basis in Brachythecium rutabulum. Total chlorophyll increases during summer full canopy conditions from 1.70 mg chl g?1 on 8 May to 11.1 mg chl g?1 on 11 October. Photosynthetic-illumination curves show that during this period light saturation declines from 200 μmol m?2s?1 to 30 μmol m?2s?1 by 6 July, and light compensation falls dramatically from 65 μmol m?2s?1 to 4 μmol m?2s?1. The data also appear to support the conclusion that there is concurrently an increase in the density of photosynthetic units by the end of September.  相似文献   

8.
Photosynthetic pigments, C, N, and P tissue composition, and photosynthetic rate were measured from April to October in the brown alga Phyllariopsis purpurascens (C. Agardh) Henry et South (Laminariales, Phaeophyta) growing at a 30-m depth in the Strait of Gibraltar. Ir-radiance reaching the population ranged from 13.5 to 27.5 mol.m-2.mo-1. The available light for this species, expressed as a percentage of the irradiance above the water, was 1.8%. Dissolved inorganic nitrogen forms, NO3-and NH4+, were constant from April to October, whereas phosphate was depleted in August. Chlorophyll a decreased from 520.0 ± 165.0 to 199.6 ± 159.9 μg.g-1 dry weight; in contrast, chlorophyll c and carotenoids did not change until September but increased threefold in October. C:N and N:P ratios changed in the same way and in the same range. They were constant until July but increased from 15–17 up to 42 (C:N) and from 14 to 40 (N:P) in October, suggesting a severe P limitation of growth of this species. The dark respiration rate and the light compensation point were constant from April to October (0.5 ± 0.1 μmol O2. m-2.s-1 and 6.5 ± 0.2 μmol.m-2. s-1, respectively), whereas the maximum rate of apparent photosynthesis, light onset saturation parameter, and half saturation constant for light were maximum in April to May (3.7 μmol O2. m-2.s-1and 40 and 41.5 μmol.m-2. s-1, respectively) and October (3.6 μmol O2. m-2.s-1 and 50 and 53.7 μmol.m-2. s-1, respectively). They were minimum in August (1.2 μmol O2.m-2.s-1 and 11.3 and 12 μmol.m-2.s-1, respectively). These minimum figures yielded a negative carbon budget in August and 0 in September, whereas it was positive the rest of the year. Photosynthetic efficiency, estimated by the ratio between maximum apparent photosynthesis and light half saturation constant, showed a strong agreement with productivity measured by means of an independent method. These results indicate that lamina expansion in this species is controlled by photosynthetic efficiency.  相似文献   

9.
It is demonstrated that cyanobacteria (both azotrophic and non‐azotrophic) contain heme b oxidoreductases that can convert chlorite to chloride and molecular oxygen (incorrectly denominated chlorite ‘dismutase’, Cld). Beside the water‐splitting manganese complex of photosystem II, this metalloenzyme is the second known enzyme that catalyses the formation of a covalent oxygen–oxygen bond. All cyanobacterial Clds have a truncated N‐terminus and are dimeric (i.e. clade 2) proteins. As model protein, Cld from Cyanothece sp. PCC7425 (CCld) was recombinantly produced in Escherichia coli and shown to efficiently degrade chlorite with an activity optimum at pH 5.0 [kcat 1144 ± 23.8 s?1, KM 162 ± 10.0 μM, catalytic efficiency (7.1 ± 0.6) × 106 M?1 s?1]. The resting ferric high‐spin axially symmetric heme enzyme has a standard reduction potential of the Fe(III)/Fe(II) couple of ?126 ± 1.9 mV at pH 7.0. Cyanide mediates the formation of a low‐spin complex with kon = (1.6 ± 0.1) × 105 M?1 s?1 and koff = 1.4 ± 2.9 s?1 (KD ~ 8.6 μM). Both, thermal and chemical unfolding follows a non‐two‐state unfolding pathway with the first transition being related to the release of the prosthetic group. The obtained data are discussed with respect to known structure–function relationships of Clds. We ask for the physiological substrate and putative function of these O2‐producing proteins in (nitrogen‐fixing) cyanobacteria.  相似文献   

10.
Candida utilis was grown on a pineapple cannery effluent in a chemostat at dilution rates ranging between 0.05 and 0.65 h–1 to establish optimal conditions for biomass production and chemical oxygen demand (COD) reduction. Sucrose, fructose and glucose were the main sugars in the effluent. Maximum value for cell yield coefficient and productivity were (0.686, gx/gs) and (2.96, gx/l/h) at a dilution rate of 0.425 and 0.475 h–1, respectively, while maximum COD reduction (98%) was attained at a dilution rate of 0.1 h–1. The maintenance coefficient attained a value of (0.093, gs/gx/h). An increase in dilution rate produced a higher protein content of the biomass.  相似文献   

11.
Rates of net photosynthesis and respiration were determined for Pithophora oedogonia (Mont.) Wittr. acclimatized to 56 combinations of light (7–1200 μE m?2 s?1) and temperature (5–35°C). Conditions for maximum net photosynthesis were estimated to be 26°C and 970 μE m?2 s?1. The rate of net photosyntheses varied considerably with temperature, with the maximum measured value (9.67 mg O2 h?1 g dry wt.?1) occurring at 25°C. Respiration rate increased with temperature and the light received just prior to measurement. The maximum respiration rate (7.05 mg O2 g?1 h?1) occurred at 30°C and 1200 μE m?2 s?1. Exposure of Pithophora to light levels of 600 or 1200 μE m?2 s?1 prior to determination of the respiration rate resulted in significantly elevated levels of oxygen consumption at temperatures ≥ 15°C. The relationship between light, temperature and photosynthesis and respiration were summarized as three-dimensional response surfaces.  相似文献   

12.
Recovery from 60 min of photoinhibitory treatment at photosynthetic photon flux densities of 500, 1400 and 2200 μMmol m?2 s? was followed in cells of the green alga Chlamydomonas reinhardtii grown at 125 μMmol m?2 s?1. These light treatments represent photoregulation, moderate photoinhibition and strong photoinhibition, respectively. Treatment in photoregulatory light resulted in an increased maximal rate of oxygen evolution (Pmax) and an increased quantum yield (Φ), but a 15% decrease in Fv/FM. Treatment at moderately photoinhibitory light resulted in a 30% decrease in Fv/FM and an approximately equal decrease in Φ. Recovery in dim light restored Fv/FM within 15 and 45 min after high light treatment at 500 and 1400 μMmol m?2 s?1, respectively. Convexity (Θ), a measure of the extent of co-limitation between PS II turnover and whole-chain electron transport, and Φ approached, but did not reach the control level during recovery after exposure to 1400 μMmol m?2 s?1, whereas Pmax increased above the control. Treatment at 2200 μMmol m?2 s?1 resulted in a strong reduction of the modeled parameters Φ, Θ and Pmax. Subsequent recovery was initially rapid but the rate decreased, and a complete recovery was not reached within 120 min. Based on the results, it is hypothesized that exposure to high light results in two phenomena. The first, expressed at all three light intensities, involves redistribution within the different aspects of PS II heterogeneity rather than a photoinhibitory destruction of PS II reaction centers. The second, most strongly expressed at 2200 μmol m?2 s?1, is a physical damage to PS II shown as an almost total loss of PS IIα and PS II QB-reducing centers. Thus recovery displayed two phase, the first was rapid and the only visible phase in algae exposed to 500 and 1400 μmol m?2 s?1. The second phase was slow and visible only in the later part of recovery in cells exposed to 2200 μmol m?2 s?1.  相似文献   

13.
Biogeochemical, isotope geochemical and microbiological investigation of Lake Svetloe (White Sea basin), a meromictic freshwater was carried out in April 2014, when ice thickness was ~0.5 m, and the ice‐covered water column contained oxygen to 23 m depth. Below, the anoxic water column contained ferrous iron (up to 240 μμM), manganese (60 μM), sulfide (up to 2 μM) and dissolved methane (960 μM). The highest abundance of microbial cells revealed by epifluorescence microscopy was found in the chemocline (redox zone) at 23–24.5 m. Oxygenic photosynthesis exhibited two peaks: the major one (0.43 μmol C L?1 day?1) below the ice and the minor one in the chemocline zone, where cyanobacteria related to Synechococcus rubescens were detected. The maximum of anoxygenic photosynthesis (0.69 μmol C L?1 day?1) at the oxic/anoxic interface, for which green sulfur bacteria Chlorobium phaeoclathratiforme were probably responsible, exceeded the value for oxygenic photosynthesis. Bacterial sulfate reduction peaked (1.5 μmol S L?1 day?1) below the chemocline zone. The rates of methane oxidation were as high as 1.8 μmol CH4 L?1 day?1 at the oxi/anoxic interface and much lower in the oxic zone. Small phycoerythrin‐containing Synechococcus‐related cyanobacteria were probably involved in accumulation of metal oxides in the redox zone.  相似文献   

14.
In order to separate the net effect of growth at elevated [CO2] on stomatal conductance (gs) into direct and acclimatory responses, mid‐day values of gs were measured for plants grown in field plots in open‐topped chambers at the current ambient [CO2], which averaged 350 μmol mol?1 in the daytime, and at ambient + 350 μmol mol?1[CO2] for winter wheat, winter barley, potato and sorghum. The acclimatory response was determined by comparing gs measured at 700 μmol mol?1[CO2] for plants grown at the two [CO2]. The direct effect of increasing [CO2] from 350 to 700 μmol mol?1 was determined for plants grown at the lower concentration. Photosynthetic rates were measured concurrently with gs. For all species, growth at the higher [CO2] significantly reduced gs measured at 700 μmol mol?1[CO2]. The reduction in gs caused by growth at the higher [CO2] was larger for all species on days with low leaf to air water vapour pressure difference for a given temperature, which coincided with highest conductances and also the smallest direct effects of increased [CO2] on conductance. For barley, there was no other evidence for stomatal acclimation, despite consistent down‐regulation of photosynthetic rate in plants grown at the higher [CO2]. In wheat and potato, in addition to the vapour pressure difference interaction, the magnitude of stomatal acclimation varied directly in proportion to the magnitude of down‐regulation of photosynthetic rate through the season. In sorghum, gs consistently exhibited acclimation, but there was no down‐regulation of photosynthetic rate. In none of the species except barley was the direct effect the larger component of the net reduction in gs when averaged over measurement dates. The net effect of growth at elevated [CO2] on mid‐day gs resulted from unique combinations of direct and acclimatory responses in the various species.  相似文献   

15.
The selective, sensitive method of analysis of ascorbic acid by high performance liquid chromatography with electrochemical detection (HPLC/EC) has been used to determine the ascorbic acid content of cell extracts from yeasts grown in glucose-free medium, 0.3 M D-glucose, and 0.112 M L-galactono-1,4-lactone. Saccharomyces cerevisiae (strain G-25 and its tetraploid) and a commercial baker's yeast contained less than 2 μg ascorbic acid g?1 wet wt. of cells when grown for 22 h in glucose-free medium. In 0.3 M D-glucose, only the commercial baker's yeast gave a slight increase (2–50 μg g?1 wet wt. in 22 h). In 0.112 M L-galactono-1,4-lactone, all three strains produced ascorbic acid (372–587 μg g?1 wet wt. in 22 h). Lypomyces starkeyi, a species previously reported to contain a significant amount of ascorbic acid (Heick et al., Can. J. Biochem., 47 (1972) 752), was essentially devoid of ascorbic acid under all three conditions of incubation although it did contain an HPLC/EC reactive peak (RT = 0.87 relative to ascorbic acid) that was readily oxidized by charcoal in the presence of oxygen. The identity of this new compound remains to be determined.  相似文献   

16.
The aim of current study is to identify the kinetic characteristics and elucidate the possible transformation pathways of the interaction between the redox mediator (anthraquinone-2,6-disulfonate, AQDS) and goethite during the process of microbial goethite reduction by Shewanella putrefaciens, a dissimilatory iron reduction bacterium (DIRB). Speciations of both AQDS and microbially reduced ferrous iron are used to characterize the interaction process among S. putrefaciens, AQDS and goethite. Due to the complexities of the natural environment, two pre-incubation reaction systems of the “DIRB–goethite” and the “DIRB–AQDS” are introduced to investigate the dynamics of goethite reduction and redox transformation of AQDS. Results show that the characteristics of the microbial goethite reduction and the kinetic transformation between two species of the redox mediator are different in two pre-incubation reaction systems. Both abiotic and enzymatic reactions and their coupling regulate the kinetic process for “redox mediatoriron” interaction in the presence of DIRB. This study will help to understand the characteristics and mechanism of microbial reduction of the Fe(III) oxide and transformation of redox mediator.  相似文献   

17.
The role of aldose reductase (ALR2) in diabetes mellitus is well-established. Our interest in finding ALR2 inhibitors led us to explore the inhibitory potential of new thiosemicarbazones. In this study, we have synthesized adamantyl-thiosemicarbazones and screened them as aldehyde reductase (ALR1) and aldose reductase (ALR2) inhibitors. The compounds bearing phenyl 3a, 2-methylphenyl 3g and 2,6-dimethylphenyl 3m have been identified as most potent ALR2 inhibitors with IC50 values of 3.99 ± 0.38, 3.55 ± 0.26 and 1.37 ± 0.92 µM, respectively, compared with sorbinil (IC50 = 3.14 ± 0.02 μM). The compounds 3a, 3g, and 3m also inhibit ALR1 with IC50 value of 7.75 ± 0.28, 7.26 ± 0.39 and 7.04 ± 2.23 µM, respectively. Molecular docking was also performed for putative binding of potent inhibitors with target enzyme ALR2. The most potent 2,6-dimethylphenyl bearing thiosemicarbazone 3m (IC50 = 1.37 ± 0.92 µM for ALR2) and other two compound 3a and 3g could potentially lead for the development of new therapeutic agents.  相似文献   

18.
Diurnal and seasonal trends in net photosynthetic rate (P N), stomatal conductance (g), transpiration rate (E), vapour pressure deficit, temperature, photosynthetic photon flux density, and water use efficiency (WUE) were compared in a two-year-old Dalbergia sissoo and Hardwickia binata plantation. Mean daily maximum P N in D. sissoo ranged from 21.40±2.60 μmol m−2 s−1 in rainy season I to 13.21±2.64 μmol m−2 s−1 in summer whereas in H. binata it was 20.04±1.20 μmol m−2 s−1 in summer and 13.64±0.16 μmol m−2 s−1 in winter. There was a linear relationship between daily maximum P N and g s in D. sissoo but there was no strong linear relationship between P N and g s in H. binata. In D. sissoo, the reduction in g s led to a reduction in both P N and E enabling the maintenance of WUE during dry season thereby managing unfavourable environmental conditions efficiently whereas in H. binata, an increase in g s causes an increase of P N and E with a significant moderate WUE.  相似文献   

19.
Molecular and microscopic studies were performed to identify Chroococcidiopsis sp., an endolithic cyanobacterium, isolated from gypsum rocks of Atacama Desert (Chile). It was adapted to grow in mineral liquid medium, with 9 mM nitrate, bubbled with CO2-enriched air (2.5 % v/v), and continuously illuminated with a white light of 70 μmol photons m–2 s–1. The obtained biomass (productivity of 0.21 g L–1 d–1) had a C/N ratio of 6.67, and it contained carbohydrates (45.40 % of dry weight), proteins (36.72 %), lipids (5.60 %) nucleic acids (3.90 %) and ashes (8.28 %). The lipid fraction was particularly rich in palmitic (29.86 % of total fatty acids), linoleic (18.20 %), palmitoleic (12.75 %), linolenic (10.92 %), stearic (9.64 %) and capric acid (6.29 %). Chroococcidiopsis sp. accumulated phycobiliproteins in a light-dependent process and produced 204 mg g–1, under incident light of 10 μmol photons·m–2·s–1, with a relative abundance of 40.9 % for phycocyanin, 23.3 % for phycoerythrin, and 35.8 % for allophycocyanin. The biomass from this cyanobacterium can be a good source of these pigments, especially APC (maximum of 95 mg g dw−1), which are of interest for pharmacological, cosmetic, and food industries.  相似文献   

20.
The aim of this study was to elucidate the kinetic constraints during the redox biotransformation of the azo dye, Reactive Red 2 (RR2), and carbon tetrachloride (CT) mediated by soluble humic acids (HAs) and immobilized humic acids (HAi), as well as by the quinoid model compounds, anthraquinone-2,6-disulfonate (AQDS) and 1,2-naphthoquinone-4-sulfonate (NQS). The microbial reduction of both HAs and HAi by anaerobic granular sludge (AGS) was the rate-limiting step during decolorization of RR2 since the reduction of RR2 by reduced HAi proceeded at more than three orders of magnitute faster than the electron-transferring rate observed during the microbial reduction of HAi by AGS. Similarly, the reduction of RR2 by reduced AQDS proceeded 1.6- and 1.9-fold faster than the microbial reduction of AQDS by AGS when this redox mediator (RM) was supplied in soluble and immobilized form, respectively. In contrast, the reduction of NQS by AGS occurred 1.6- and 19.2-fold faster than the chemical reduction of RR2 by reduced NQS when this RM was supplied in soluble and immobilized form, respectively. The microbial reduction of HAs and HAi by a humus-reducing consortium proceeded 1,400- and 790-fold faster than the transfer of electrons from reduced HAs and HAi, respectively, to achieve the reductive dechlorination of CT to chloroform. Overall, the present study provides elucidation on the rate-limiting steps involved in the redox biotransformation of priority pollutants mediated by both HAs and HAi and offers technical suggestions to overcome the kinetic restrictions identified in the redox reactions evaluated.  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