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 31 毫秒
1.
Decolorization of several dyes (Red HE-8B, Malachite Green, Navy Blue HE-2R, Magenta, Crystal Violet) and an industrial effluent with growing cells ofPhanerochœte chrysosporium in shake and static culture was demonstrated. All the dyes and the industrial effluent were decolorized to some extent with varying percentages of decolorization (20–100%). The rate of decolorization was very rapid with Red HE-8B, an industrial dye. Decolorization rates for all the dyes in static condition were found to be less than the shake culture and also dependent on biomass concentration.  相似文献   

2.
The ability of a Brazilian strain ofPleurotus pulmonarius to decolorize structurally different synthetic dyes (including azo, triphenylmethane, heterocyclic and polymeric dyes) was investigated in solid and submerged cultures. Both were able to decolorize completely or partially 8 of 10 dyes (Amido Black, Congo Red, Trypan Blue, Methyl Green, Remazol Brilliant Blue R, Methyl Violet, Ethyl Violet, Brilliant Cresyl Blue). No decolorization of Methylene Blue and Poly R 478 was observed. Of the four phenol-oxidizing enzymes tested in culture filtrates (lignin peroxidase, manganese peroxidase, aryl alcohol oxidase, laccase),P. pulmonarius produced only laccase. Both laccase activity and dye decolorization were related to glucose and ammonium starvation or to induction by ferulic acid. The decolorizationin vivo was tested using three dyes — Remazol Brilliant Blue R, Trypan Blue and Methyl Green. All of them were completely decolorized by crude extracellular extracts. Decolorization and laccase activity were equally affected by pH and temperature. Laccase can thus be considered to be the major enzyme involved in the ability ofP. pulmonarius to decolorize industrial dyes.  相似文献   

3.
Aspergillus sojae B-10 was immobilized and used to treat model dye compounds. The model wastewater, containing 10 ppm of azo dyes such as Amaranth, Sudan III, and Congo Red, was treated with cells attached to a rotating disc contactor (RDC). Amaranth was decolorized more easily than were Sudan III and Congo Red. Decolorization of Amaranth began within a day, and the dye was completely decolorized within 5 days of incubation. Both Sudan III and Congo Red were almost completely decolorized after 5 days of incubation. Semicontinuous decolorization of azo by reusing attached mycelia resulted in almost complete decolorization in 20 days. This experiment indicated that decolorization was successfully conducted by removing azo dyes withAspergillus sojae B-10.  相似文献   

4.
The little studied white rot fungus Ischnoderma resinosum was tested for its ability to decolorize seven different synthetic dyes. The strain efficiently decolorized Orange G, Amaranth, Remazol Brilliant Blue R, Cu-phthalocyanin and Poly R-478 on agar plates and in liquid culture at a relatively high concentration of 2–4 and 0.5–1 g l−1, respectively. Malachite Green and Crystal Violet were decolorized to a lower extent up to the concentration of 0.1 g l−1. Decolorization capacity of I. resinosum was higher than that in Phanerochaete chrysosporium, Pleurotus ostreatus or Trametes versicolor. In contrast with these thoroughly examined fungi, I. resinosum was able to degrade a wide spectrum of chemically and structurally different synthetic dyes. I. resinosum also efficiently decolorized dye mixtures. In liquid culture, Orange G and Remazol Brilliant Blue R were decolorized most rapidly; the process was not affected by different nitrogen content in the media. Shaken cultivation strongly inhibited the decolorization of Orange G.  相似文献   

5.
Several dyes, representing different structural classes, inhibit [3H]flunitrazepam binding to brain specific receptors in the rat with 50% inhibition in the 1 to 100 M range. Crystal Violet and Methyl Violet 2B inhibited more potently in the forebrain than in the cerebellum. Congo Red yielded a Hill number near 2.3, probably reflecting positive cooperativity between interacting binding sites in benzodiazepine receptor complexes. Toluidine Blue 0 was the most potent of the dyes tested (IC50=1 M in cerebellum) and inhibited more potently in cerebellum than in forebrain.  相似文献   

6.
Red HE7B (RHE7B, 100 mg l−1), a sulfonated azo dye, was decolorized at static condition by Pseudomonas desmolyticum NCIM 2112 in 72 h with 71% reduction in chemical oxygen demand (COD). Extracellular lignin peroxidase (LiP) has played a crucial role in breakdown of the dye by asymmetric cleavage and reductases in the initial 24 h incubation to break azo bonds of some dye molecules. Dye also induced the activity of aminopyrine N-demethylase, one of the enzymes of mixed function oxidase system. Decolorization and degradation were analyzed by using UV–vis and high-pressure liquid chromatography (HPLC). The Fourier transform infrared spectroscopy (FTIR) analysis revealed that P. desmolyticum preferred C–N and SO bonds to break down the RHE7B. GC–MS identification of 8-amino-naphthalene-1,3,6,7-tetraol and 2-hydroxyl-6-oxalyl-benzoic acid as final metabolites supports the degradation of RHE7B by desulfonation before and after ring cleavage. Aerobic degradation of amines and reduced phytotoxicity increased the applicability of this microorganism for dye removal.

Scientific relevance of the paper

This is the first report on degradation of Red HE7B by oxidative enzymes and on further degradation by desulfonation before and after ring cleavage.  相似文献   

7.
A broad-spectrum dye-decolorizing bacterium, strain DN322, was isolated from activated sludge of a textile printing wastewater treatment plant. The strain was characterized and identified as a member of Aeromonas hydrophila based on Gram staining, morphology characters, biochemical tests, and nearly complete sequence analysis of 16S rRNA gene and the gyrase subunit beta gene (gyrB). Strain DN322 decolorized a variety of synthetic dyes, including triphenylmethane, azo, and anthraquinone dyes. For color removal, the most suitable pH and temperature were pH 5.0–10.0 and 25–37°C, respectively. Triphenylmethane dye, e.g., Crystal Violet, Basic Fuchsin, Brilliant Green, and Malachite Green (50 mg l−1) were decolorized more than 90% within 10 h under aerobic culture condition and Crystal Violet could be used as sole carbon source and energy source for cell growth. The color removal of triphenylmethane dyes was due to a soluble cytosolic enzyme, and the enzyme was an NADH/NADPH-dependent oxygenase; For azo and anthraquinone dyes, e.g., Acid Amaranth, Great Red GR, Reactive Red KE-3B, and Reactive Brilliant Blue K-GR (50 mg l−1) could be decolorized more than 85% within 36 h under anoxic condition. This strain may be useful for bioremediation applications.  相似文献   

8.
Biodegradation of Methyl red by Galactomyces geotrichum MTCC 1360   总被引:1,自引:0,他引:1  
Galactomyces geotrichum MTCC 1360 can decolorize triphenylmethane, azo and reactive high exhaust textile dyes. At shaking condition this strain showed 100% decolorization of a toxic azo dye Methyl red (100 m gl−1) within 1 h in deionized water at 30 °C. The degradation of Methyl red was possible through a broad pH (3–12) and temperature (5–50 °C) range. Glucose and mycelium concentration had increased the decolorization rate, but the addition of 1 gl−1 molasses in deionized water made decolorization possible in only 10 min. Induction in the NADH–dichloro phenol indophenol (NADH–DCIP) reductase, Malachite green reductase, laccase and lignin peroxidase (Lip) activities were observed in the cells obtained after complete decolorization, showing that there is direct involvement in the degradation of Methyl red. The absence of N-N′-dimethyl-p-phenylenediamine (DMPD) in 5 °C, 2-aminobenzoic acid (ABA) in 50 °C and both the compounds in 30 °C sample have shown the differences in the metabolic fate of Methyl red at different temperatures. The untreated dye at 300 mg l−1 concentration showed 88% germination inhibition in Sorghum bicolor, whereas it was 72% in Triticum aestivum. There was no germination inhibition for both the plants by Methyl red metabolites at 300 mg l−1 concentration.

The scientific relevance of the paper

The azo dye Methyl red (100 mg l−1) was decolorized by G. geotrichum MTCC 1360 within 1 h at shaking condition in deionized water. This organism could decolorize Methyl red at wide pH and temperature ranges. Decolorization time was reduced to 10 min by the addition of molasses to deionized water. There was induction in laccase and Lip, NADH–DCIP reductase and Malachite green reductase activities. The metabolic fate of Methyl red changes with temperature which can be evidenced by the formation of 2-ABA at 5 °C, N-N′-DMPD at 50 °C and both the compounds were absent at 30 °C. Phytotoxicity showed that metabolites of dye had induced shoot and root length of both the tested plants.  相似文献   

9.
Decolourisation of reactive dyes Drimarene Blue X3LR and Remazol Brilliant Blue R by white rot fungi Funalia trogii was studied under static conditions. The effect of various conditions such as mycelial age, initial dye and glucose concentrations on decolourisation were also investigated. Decolourisation activity of F. trogii was compared with Phanerochaete chrysosporium known as test microorganism. It was found that 7-day-old cultures were more effective than 5-day-old cultures of F. trogii for decolourisation of these dyes. Decolourisations by F. trogii of both dyes were increased with glucose concentration decreasing. In contrast, decolourisations by P. chrysosporium were decreased. F. trogii decolourised 92–98% of both dyes within 4–10 h. However, P. chrysosporium partially decolourised (11–20%) these dyes during 10days incubation period under the same conditions.  相似文献   

10.
Decolorization of textile reactive azo dyes by a strain of bacteria (SL186) isolated from a contaminated site was investigated. SL186 was identified as Clostridium bifermentans by phenotypic characterization and 16S rDNA sequence comparison. Under anaerobic conditions, SL186 had decolorized the dyes Reactive Red 3B-A, Reactive Black 5, and Reactive Yellow 3G-P by over 90% after 36 h post-inoculation. The bacterium retained decolorizing activity over a wide range of pH values (6–12), with peak activity at pH 10. Additionally, SL186 decolorized a relatively high concentration of Reactive Red 3B-A dye (1,000 ppm) by over 80% and raw industrial effluent effectively. The addition of glucose increased the decolorization rate a little. Spectrophotometric analyses of the reactive dyes showed no distinct peak indicating aromatic amines. However, a new peak was detected between 300 and 450 nm from the decolorized raw industrial effluent. These results suggest that C. bifermentans SL186 is a suitable bacterium for the biological processing of dye-containing wastewater.  相似文献   

11.
Zhang SJ  Yang M  Yang QX  Zhang Y  Xin BP  Pan F 《Biotechnology letters》2003,25(17):1479-1482
Three reactive dyes were rapidly adsorbed by the mycelium pellets of Penicillium oxalicum. Dye removal of Reactive Blue 19 was up to 60% in 10 min and 91% in 80 min. Dye adsorption isotherms fitted Langmuir model well and the maximum adsorption capacities at 20 °C were calculated to be 160 mg g–1 for Reactive Blue 19, 122 mg g–1 for Reactive Red 241 and 137 mg g–1 for Reactive Yellow 145, respectively. The pellets exhibited a high dye adsorption capacity (80–180 mg g–1) for all of the 3 dyes over a wide pH range (pH 2–10), and the maximum adsorption was obtained at pH 2. The adsorption capacity was mildly increased by increasing salinity.  相似文献   

12.
Of seven fungal strains tested for their ability to decolourise three structurally diverse synthetic dyes, Phanerochaete sordida, Bjerkandera sp. BOS55, Phlebia radiata, and Phanerochaete chrysosporium had average values of maximum decolourisation rates higher than 0.2 [Absorbance] d–1. All seven fungi produced manganese peroxidase (MnP) but laccase activity was detected only in Phlebia radiata. No lignin peroxidase (LiP) activity was observed.  相似文献   

13.
The white rot fungus, Fomes lividus, was isolated from the logs of Shorea robusta in the Western Ghats region of Tamil Nadu, India. The fungus was tested for decolorization of azo dyes such as orange G (50 M) congo red (50 M) amido black 10B (25 M) and also for colour removal from dye industry effluents. The results revealed that the fungus could remove only 30.8% of orange G in the synthetic solution, whereas congo red and amido black 10B were removed by 74.0 and 98.9% respectively. A dye industry effluent was treated by the fungus in batch and continuous mode. In batch mode treatment, a maximum decolorization of 84.4% was achieved on day 4, and in continuous mode a maximum decolorization of 37.5% was obtained on day 5. The colour removal by the basidiomycete fungus might be due to adsorption of the dyes to the mycelial surface and metabolic breakdown. These results suggested that the batch mode treatment of Fomes lividus is one of the most efficient ways for colour removal in dye industry effluents.  相似文献   

14.
The cyanobacterium Spirulina platensis was used to verify the possibility of employing microalgal biomass to reduce the contents of nitrate and phosphate in wastewaters. Batch tests were carried out in 0.5 dm3 Erlenmeyer flasks under conditions of light limitation (40 mol quanta m–2 s–1) at a starting biomass level of 0.50 g/dm3 and varying temperature in the range 23–40°C. In this way, the best temperature for the growth of this microalga (30°C) was determined and the related thermodynamic parameters were estimated. All removed nitrate was used for biomass growth (biotic removal), whereas phosphate appeared to be removed mainly by chemical precipitation (abiotic removal). The best results in terms of specific and volumetric growth rates ( =0.044 day–1, Q x =33.2 mg dm–3 day–1) as well as volumetric rate and final yield of nitrogen removal ( =3.26 mg dm–3 day–1, =0.739) were obtained at 30°C, whereas phosphorus was more effectively removed at a lower temperature. In order to simulate full-scale studies, batch tests of nitrate and phosphate removal were also performed in 5.0 dm3 vessels (mini-ponds) at the optimum temperature (30°C) but increasing the photon fluence rate to 80 mol quanta m–2 s–1 and varying the initial biomass concentration from 0.25 to 0.86 g/dm3. These additional tests demonstrated that an increase in the inoculum level up to 0.75 g/dm3 enhanced both NO3 and PO4 3– removal, confirming a strict dependence of these processes on biomass activity. In addition, the larger surface area of the ponds and the higher light intensity improved removal yields and kinetics compared to the flasks, particularly concerning phosphorus removal ( =0.032–0.050 day–1, Q x =34.7–42.4 mg dm–3 day–1, =3.24–4.06 mg dm–3 day–1, =0.750–0.879, =0.312–0.623 mg dm–3 day–1, and =0.224–0.440).  相似文献   

15.
Summary Three retinal-containing pigments have been detected inHalobacterium halobium membranes: bacteriorhodopsin (bR), halorhodopsin (hR), and slow-cycling rhodopsin (sR). The first two hyperpolarize the cell membrane by electrogenic transport of H+ and Cl, respectively. The third pigment, sR, may be a photosensory receptor since mutants lacking bR and hR retain their retinal-dependent phototaxis responses. We monitored light-induced changes in fluorescence of several voltage-sensitive dyes in cells and membrane vesicles. Red light-induced potential changes generated by bR and hR were similar to signals described previously. Signals generated by hR could be identified using four criteria: wavelength dependence, Cl dependence, shunting by valinomycin and K+, and the absence of these signals in hR-deficient mutants. The absence (detection limit 0.5 mV) of hyperpolarization signals in bRhRsR+ vesicles and cells shows that sR photochemical reactions are nonelectrogenic. Two signals independent of bR and hR were measured: blue light caused a decrease and red light an increase in dye fluorescence. Both signals appear to derive from sR on the basis of their retinal-dependence and action spectra. In a retinal-deficient mutant strain (Flx3R), both sR signals appeared after addition of all-trans retinal. In this strain retinal also restores phototaxis sensitivity within the same time scale. The retinal concentration dependence for all four parameters monitored—the attractant (red) and repellent (blue) phototaxis, and the red light and blue light-induced fluorescence signals—is the same. This correlation is consistent with the hypothesis that both attractant and repellent responses are mediated by sR, as suggested by Bogomolni and Spudich (Proc. Natl. Acad. Sci. USA.79:6250–6254 (1982)).  相似文献   

16.
Summary Green plants were repeatedly regenerated from suspension-derived protoplasts of Kentucky blue grass (Poa pratensis L.) cv. Geronimo. One suspension was capable of donating competent protoplasts during long-term culture i. e. 10–16 months after its establishment. The plating efficiency of the protoplasts from three different suspension lines varied from 0.004% in the lowest up to 1.5% in the highest responding line, using agarose-bedding in nutrition medium devoid of nurse or feeder cultures. Green plants germinated from polyembryos, which developed from 0.4–2.7% of the protoplast-derived microcolonies. A total of 127 plants were successfully transferred to soil.Abbreviations ABA abscisic acid - BAP 6-benzylaminopurine - 2,4-D 2,4-dichlorophenoxyacetic acid - FW fresh weight - PE plating efficiency  相似文献   

17.
Botryosphaeran, a (13;16)-β-d-glucan produced by Botryosphaeria rhodina MAMB-05, was found to be present in a triple helix conformation from helix–coil transition studies using Congo Red. The triple helix conformation was disrupted at increasing alkali concentrations. Conformational changes were also observed using phenanthrene as a fluorescent probe, and the fluorescence intensity decreased 80% in the presence of dimethyl sulfoxide. The results confirmed the triple helix conformation of botryosphaeran, an important property manifesting biological response modifying activity.  相似文献   

18.
George  T.S.  Gregory  P.J.  Robinson  J.S.  Buresh  R.J.  Jama  B. 《Plant and Soil》2002,246(1):53-63
A field experiment in western Kenya assessed whether the agroforestry species Tithonia diversifolia (Hemsley) A. Gray, Tephrosia vogelii Hook f., Crotalaria grahamiana Wight & Arn. and Sesbania sesban (L) Merill. had access to forms of soil P unavailable to maize, and the consequences of this for sustainable management of biomass transfer. The species were grown in rows at high planting density to ensure the soil under rows was thoroughly permeated by roots. Soil samples taken from beneath rows were compared to controls, which included a bulk soil monolith enclosed by iron sheets within the tithonia plot, continuous maize, and bare fallow plots. Three separate plant biomass samples and soil samples were taken at 6-month intervals, over a period of 18 months. The agroforestry species produced mainly leaf biomass in the first 6 months but stem growth dominated thereafter. Consequently, litterfall was greatest early in the experiment (0–6 months) and declined with continued growth. Soil pH increased by up to 1 unit (from pH 4.85) and available P increased by up to 38% (1 g P g–1) in agroforestry plots where biomass was conserved on the field. In contrast, in plots where biomass was removed, P availability decreased by up to 15%. Coincident with the declines in litterfall, pH decreased by up to 0.26 pH units, plant available P decreased by between 0.27 and 0.72 g g–1 and Po concentration decreased by between 8 and 35 g g–1 in the agroforestry plots. Declines in Po were related to phosphatase activity (R2=0.65, P<0.05), which was greater under agroforestry species (0.40–0.50 nmol MUB s–1 g–1) than maize (0.28 nmol MUB s–1 g–1) or the bare fallow (0.25 nmol MUB s–1 g–1). Management of tithonia for biomass transfer, decreased available soil P by 0.70 g g–1 and Po by 22.82 g g–1. In this study, tithonia acquired Po that was unavailable to maize. However, it is apparent that continuous cutting and removal of biomass would lead to rapid depletion of P stored in organic forms.  相似文献   

19.
A 6.3 kb DNA fragment containing genes responsible for azo-dye decolorization was cloned and expressed in Escherichia coli. The resulting recombinant strain E. coli CY1 decolorized 200 mg azo dye (C.I. Reactive Red 22) l–1 at 28 °C at 8.2 mg g cell–1 h–1, while the host (E. coli DH5) had no color-removal activity. Addition of 0.5 mM isopropyl--d-thiogalacto-pyranoside (IPTG) increased the decolorization rate 3.4-fold. The dependence of the decolorization rate on initial dye concentration essentially followed Monod-type kinetics and the maximal rate occurred with the dye at 600 mg l–1. The decolorization rate of E. coli CY1 was optimal at 40 °C and pH 11. Aeration (increased dissolved O2 level) strongly inhibited the decolorization, but decolorization occurred effectively under static incubation conditions (no agitation was employed). The CY1 strain also exhibited excellent stability during repeated-batch operations.  相似文献   

20.
Decolorization of azo dyes by Rhodobacter sphaeroides   总被引:5,自引:0,他引:5  
Song ZY  Zhou JT  Wang J  Yan B  Du CH 《Biotechnology letters》2003,25(21):1815-1818
Rhodobacter sphaeroides AS1.1737 decolorized more than 90% of several azo dyes (200 mg dyes l–1) in 24 h. The optimal culture conditions were: anaerobic illumination (1990 lx), peptone as carbon source, temperature 35–40 °C and pH 7–8. Intracellular crude enzyme from this strain had azoreductase activity, optimized temperature as 45–50 °C, and decolorization kinetics which were consistent with a ping-pong mechanism.  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