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.
An ascomycetous yeast strain isolated from activated sludge could decolorize Reactive Black 5 azo dye at 200 mg l?1 up to 90 % within 12–18 h under agitated condition. Yeast decolorization ability was investigated at different RB5 concentrations and, at higher dye concentration, 500 mg l?1, the decolorization was found to be 98 % after 36 h incubation time. Extensive decolorization (95–99 %) was obtained in presence of five other azo dyes, Reactive Orange 16, Reactive Red 198, Direct Blue 71, Direct Yellow 12, and Direct Black 22, by isolated yeast. HPLC analysis, UV–vis spectra and colorless biomass obtained after complete decolorization showed that the decolorization occured through a biodegradation mechanism. Decolorization was occurred during the exponential growth phase which is associated to primary metabolism. Laccase production by the yeast cells was not detected. The isolated yeast was characterized according to phenotypical and molecular procedures and was closely related (99 % identity) to Issatchenkia orientalis.  相似文献   

2.
The objectives of this study were to investigate: (1) the capacity of Enterococcus faecalis on the decolorization of the azo dyes Acid Red 27 and Reactive Red 2; and (2) the growth characteristics of E. faecalis on those dyes. E. faecalis was able to decolorize Acid Red 27 and Reactive Red 2 effectively. High decolorization efficiency (95–100%) was achieved within 3 h of incubation for Acid Red 27, and 12 h for Reactive Red 2, at room temperature, neutral pH, static and non-aerated condition. Growth characteristics of E. faecalis on azo dyes, which were indicated by cell growth rate, biomass production, and growth yield, was worse than the control. E. faecalis grew better on Acid Red 27 rather than Reactive Red 2.  相似文献   

3.
Summary Kodam et al. reported a 100% decolorization of the sulfonated azo dyes Reactive Red 2, Reactive Red 141, Reactive Orange 4, Reactive Orange 7 and Reactive Violet 5 by an unidentified bacterium, KMK 48. High effectiveness was attained within 36 h of incubation at room temperature and neutral pH. Optimum decolorization took place strictly under aerobic conditions, which is contrary to other well-documented reports. Thus, this microorganism seems to be potentially effective for bioremediation of textile-dyeing industry effluents.  相似文献   

4.
Sphingomonas sp strain 1CX was isolated from a wastewater treatment plant and is capable of aerobically degrading a suite of azo dyes, using them as a sole source of carbon and nitrogen. All azo dyes known to be decolorized by strain 1CX (Orange II, Acid Orange 8, Acid Orange 10, Acid Red 4, and Acid Red 88) have in their structure either 1-amino-2-naphthol or 2-amino-1-naphthol. In addition, an analysis of the structures of the dyes degraded suggests that there are certain positions and types of substituents on the azo dye which determine if degradation will occur. Growth and dye decolorization occurs only aerobically and does not occur under fermentative or denitrification conditions. The mechanism by which 1CX decolorizes azo dyes appears to be through reductive cleavage of the azo bond. In the case of Orange II, the initial degradation products were sulfanilic acid and 1-amino-2-naphthol. Sulfanilic acid, however, was not used by 1CX as a growth substrate. The addition of glucose or inorganic nitrogen inhibited growth and decoloration of azo dyes by 1CX. Attempts to grow the organism on chemically defined media containing several different amino acids and sugars as sources of nitrogen and carbon were not successful. Phylogenetic analysis of Sphingomonas sp strain 1CX shows it to be related to, but distinct from, other azo dye-decolorizing Sphingomonas spp strains isolated previously from the same wastewater treatment facility. Received 19 May 1999/ Accepted in revised form 11 August 1999  相似文献   

5.
Wastewater effluents from the textile and other dye-stuff industries contain significant amounts of synthetic dyes that require treatment to prevent groundwater contamination. In research aimed at biotechnology for treatment of azo dyes, this study examined 288 strains of azo-dye degrading bacteria to identify efficient strains and determine incubation times required for decolorization. Initial enrichment cultures were carried out using a mixture of four structurally different dyes (Acid Red 88, Reactive Black 5, Direct Red 81, and Disperse Orange 3) as the sole source of C and N to isolate the bacteria from soil, activated sludge, and natural asphalt. Six strains were selected for further study based on their prolific growth and ability to rapidly decolorize the dyes individually or in mixtures. Treatment times required by the most efficient strain, AS96 (Shewanella putrefaciens) were as short as 4 h for complete decolorization of 100 mg l−1 of AR-88 and DR-81 dyes under static conditions, and 6 and 8 h, respectively, for complete decolorization of RB-5 and DO-3. To our knowledge, these bacterial strains are the most efficient azo-dye degrading bacteria that have been described and may have practical application for biological treatment of dye-polluted wastewater streams.  相似文献   

6.
Anthraquinone-2-sulfonate was immobilized on ceramsites (AQS-ceramsites) using a novel adsorption/covalence coupling method and their effects on the anaerobic bio-decolorization rates of azo dyes by salt-tolerant AQS-reducing (STAR) community were investigated. The results showed that AQS-ceramsites mediated specific bio-decolorization rates of four azo dyes Acid Yellow 36, Reactive Red 2, Acid Red 27 and Acid Orange 7 increase 2.3–6.4 fold than those lacking ceramsites in the presence of 50 g/L NaCl. Moreover, repeated experiments with AQS-ceramsites showed that the decolorization efficiencies of azo dyes could remain over 98% of their original value. These results indicated that AQS-ceramsites functioning as redox mediators exhibited good catalytic activity and stability under saline conditions. The dynamics of the STAR community structure revealed by PCR-DGGE also showed that the presence of AQS-ceramsites made STAR bacteria keeping predominant in the catalytic system. Therefore, it can be concluded that this novel solid redox mediator is potentially useful for the treatment of saline dye wastewater.  相似文献   

7.
Shan X J  Jin X S  Zhou Z P  Dai F Q 《农业工程》2012,32(5):240-243
Anthropogenic activities and environmental changes have had a significant effect on the fishery ecosystem, biological characteristics, and population dynamics of marine fishes. Overfishing threatens the sustainability of many populations. We evaluated changes in the biological characteristics, distribution, and abundance of Cleisthenes herzensteini using bottom trawl survey data collected from 1985 to 2010 in the central and southern Yellow Sea. The dominant body length of C. herzensteini during spring was 80–160 mm in 1986, 60–160 mm in 1998, and 41–80 mm and 111–170 mm in 2010. During summer, the dominant body length was 80–180 mm and 130–169 mm in 2000 and 2007, respectively. During autumn, the dominant body length was 60–160 mm, 100–180 mm, and 90–149 mm in 1985, 2000, and 2009, respectively. During winter, the dominant body length was 80–200 mm, 120–220 mm, and 100–200 mm in 1985, 1999, and 2010, respectively. The dominant body length decreased gradually from 1985 to 2010 (excluding spring, 2010), illustrating the “miniaturization” of the C. herzensteini population. Growth was significantly different between male and female individuals, with male individuals forming a “smaller-size type”. The sex ratio of C. herzensteini was relatively stable during spring and summer, but significantly different during autumn and winter. The diet of C. herzensteini also changed significantly from 1985 to 2010. During 1985–1986, the diet consisted primarily of Crangon affinis, Eualus sinensis and Gammaridae species. C. affinis, Engraulis japonicus, and Ammodytes personatus were dominant during 1998–2000, whereas C. affinis was the dominant prey species during 2009–2010. Thus, there was a clear decrease in dietary diversity, with a shift to benthos shrimp, particularly C. affinis, which accounted for 82.58% of the total diet (by weight) in 2010. The gastric vacuous rate also decreased in every season and the gonad developmental stage changed with each season. The distribution of C. herzensteini shifted northward and offshore and became more concentrated. The average catch per haul of C. herzensteini decreased in spring and autumn. The average catch per haul ranged from 1.44 kg h-1 to 0.14 kg h-1 in spring and the percentage by weight ranged from 6.53% to 1.28%. The average catch per haul ranged from 3.03 kg h-1 to 0.26 kg h-1 in autumn and the percentage by weight ranged from 8.00% to 0.60%. The average catch per haul increased significantly during summer, ranging from 0.18 kg h-1 to 0.58 kg h-1, with a percentage by weight of 0.03–0.80%. The average catch per haul was relatively stable in winter (around 1.00 kg h-1), but the percentage by weight gradually increased during 1985–2010. Taken together, our results suggested that the population structure, diet composition, and distribution of C. herzensteini had been altered during the last three decades. To address this, it is essential to initiate measures to conserve the C. herzensteini resource.  相似文献   

8.
The nonspecific ability of anaerobic sludge bacteria obtained from cattle dung slurry was investigated for 17 different dyes in a batch assay system using sealed serum vials. Experiments using Reactive Violet 5 (RV 5) showed that sludge bacteria could effectively decolorize solutions having dye concentrations up to 1000 mg l−1 with a decolorization efficiency of above 75% during 48 h of incubation. Headspace gas composition of anaerobic batch systems for varying dye concentration revealed that lower concentrations of RV 5 (upto 500 mg l−1) were found to be stimulatory to the methanogenic activity of sludge bacteria. However at higher dye concentrations, the headspace gas composition was found to be similar to batch assay controls without dye, indicating that dye at higher concentrations was inhibitory to methanogenic bacteria of sludge. The optimum inoculum and incubation temperature for maximum decolorization of RV 5 was found to be 9.0 g l−1(in terms of total solids) and 37°C, respectively. Of sixteen other dyes tested, nine (Reactive Black 5, Reactive Blue 31, Reactive Blue 28, Reactive Red HE8B, Reactive Yellow, Reactive Golden Yellow, Mordant Orange, Novatic Olive R S/D & Navilan Yellow GL) were decolorized with more than 88% efficiency; three (Orange II, Navy Blue HER & Novatic Blue BC S/D) were decolorized with about 50–65% efficiency, whereas other three dyes (Procion Orange H2R, Procion Brilliant Blue HGR & Novatic Blue BC S/D) were decolorized with less than 40% efficiency. Though Ranocid Fast Blue was decolorized with about 92.5% efficiency, this was merely due to sorption, whereas the other dyes were decolorized due to biotransformation.  相似文献   

9.
Twenty-two azo dyes were used to study the influence of substituents on azo dye biodegradability and to explore the possibility of enhancing the biodegradabilities of azo dyes without affecting their properties as dyes by changing their chemical structures. Streptomyces spp. and Phanerochaete chrysosporium were used in the study. None of the actinomycetes (Streptomyces rochei A10, Streptomyces chromofuscus A11, Streptomyces diastaticus A12, S. diastaticus A13, and S. rochei A14) degraded the commercially available Acid Yellow 9. Decolorization of monosulfonated mono azo dye derivatives of azobenzene by the Streptomyces spp. was observed with five azo dyes having the common structural pattern of a hydroxy group in the para position relative to the azo linkage and at least one methoxy and/or one alkyl group in an ortho position relative to the hydroxy group. The fungus P. chrysosporium attacked Acid Yellow 9 to some extent and extensively decolorized several azo dyes. A different pattern was seen for three mono azo dye derivatives of naphthol. Streptomyces spp. decolorized Orange I but not Acid Orange 12 or Orange II. P. chrysosporium, though able to transform these three azo dyes, decolorized Acid Orange 12 and Orange II more effectively than Orange I. A correlation was observed between the rate of decolorization of dyes by Streptomyces spp. and the rate of oxidative decolorization of dyes by a commercial preparation of horseradish peroxidase type II, extracellular peroxidase preparations of S. chromofuscus A11, or Mn(II) peroxidase from P. chrysosporium. Ligninase of P. chrysosporium showed a dye specificity different from that of the other oxidative enzymes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The vertical biomass allocation patterns of roots grown under standardised conditions were determined for species representing the major New Zealand indigenous grass genera Chionochloa and Festuca. Ten ramets, each of 2–3 tillers from garden collections of each species were grown in irrigated vertical sand columns in a glasshouse, and harvested after 168 days. Chionochloa teretifolia, Chionochloa macra, and Chionochloa crassiusucula, characteristic of alpine environments failed to produce new roots and died. However, most of the Chionochloa taxa (Chionochloa beddiei, Chionochloa pallens, Chionochloa rigida ssp. rigida, Chionochloa rubra ssp. cuprea, Chionochloa vireta), developed extensive new roots that reached the base of the one metre sand column. Roots of Chionochloa cheesemanii and Chionochloa conspicua reached 80–90 cm depth. Two Festuca taxa (Festuca actae, Festuca luciarum) had roots to 1 m depth, and roots of Festuca coxii, Festuca matthewsii ssp. latifundii, Festuca matthewsii ssp. matthewsii, Festuca multinodis, and Festuca novae-zelandiae grew to 70–90 cm depth. The edaphic specialists (Festuca deflexa, Chionochloa spiralis, Chionochloa defracta) were all shallow rooting.Species of Festuca maintained at least 40% of the root mass in the upper 10 cm of the column and most of the Chionochloa taxa had less than 40% of root mass in the upper zone. Genotype level variation in root mass less than 10 cm deep was greater in Chionochloa than in Festuca, and least in the edaphic specialist grasses.  相似文献   

11.
Twenty-two azo dyes were used to study the influence of substituents on azo dye biodegradability and to explore the possibility of enhancing the biodegradabilities of azo dyes without affecting their properties as dyes by changing their chemical structures. Streptomyces spp. and Phanerochaete chrysosporium were used in the study. None of the actinomycetes (Streptomyces rochei A10, Streptomyces chromofuscus A11, Streptomyces diastaticus A12, S. diastaticus A13, and S. rochei A14) degraded the commercially available Acid Yellow 9. Decolorization of monosulfonated mono azo dye derivatives of azobenzene by the Streptomyces spp. was observed with five azo dyes having the common structural pattern of a hydroxy group in the para position relative to the azo linkage and at least one methoxy and/or one alkyl group in an ortho position relative to the hydroxy group. The fungus P. chrysosporium attacked Acid Yellow 9 to some extent and extensively decolorized several azo dyes. A different pattern was seen for three mono azo dye derivatives of naphthol. Streptomyces spp. decolorized Orange I but not Acid Orange 12 or Orange II. P. chrysosporium, though able to transform these three azo dyes, decolorized Acid Orange 12 and Orange II more effectively than Orange I. A correlation was observed between the rate of decolorization of dyes by Streptomyces spp. and the rate of oxidative decolorization of dyes by a commercial preparation of horseradish peroxidase type II, extracellular peroxidase preparations of S. chromofuscus A11, or Mn(II) peroxidase from P. chrysosporium. Ligninase of P. chrysosporium showed a dye specificity different from that of the other oxidative enzymes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The production of ligninolytic enzymes by the fungus Schizophyllum sp. F17 using a cost-effective medium comprised of agro-industrial residues in solid-state fermentation (SSF) was optimized. The maximum activities of the enzymes manganese peroxidase (MnP), laccase (Lac), and lignin peroxidases (LiP) were 1,200, 586, and 109 U/L, respectively, on day 5 of SSF. In vitro decolorization of three structurally different azo dyes by the extracellular enzymes was monitored to determine its decolorization capability. The results indicated that crude MnP, but not LiP and Lac, played a crucial role in the decolorization of azo dyes. After optimization of the dye decolorization system with crude MnP, the decolorization rates of Orange IV and Orange G, at an initial dye concentration of 50 mg/L, were enhanced to 76 and 57%, respectively, after 20 min of reaction at pH 4 and 35°C. However, only 8% decolorization of Congo red was observed. This enzymatic reaction system revealed a rapid decolorization of azo dyes with a low MnP activity of 24 U/L. Thus, this study could be the basis for the production and application of MnP on a larger scale using a low-cost substrate.  相似文献   

13.
Dye decolorization capacity of two white-rot fungi, Irpex lacteus and Phanerochaete chrysosporium, was compared in N-limited liquid cultures. The agitated cultures showed lower ability to decolorize azo dyes Reactive Orange 16 and Naphthol Blue Black than static cultures. Similar effect was also observed with other structurally different synthetic dyes. The effect of surfactants on the decolorization process is discussed. A significant increase in the Reactive Orange 16 decolorization by the agitated I. lacteus cultures was observed after adding 0.1% Tween 80, following a higher Mn-dependent peroxidase production. The in vitro dye decolorization using the purified enzyme proved its decolorization ability.  相似文献   

14.
The biological decolorization of the textile azo dye Reactive Red 2 was investigated using a mixed, mesophilic methanogenic culture, which was developed with mixed liquor obtained from a mesophilic, municipal anaerobic digester and enriched by feeding a mixture of dextrin/peptone as well as media containing salts, trace metals and vitamins. Batch decolorization assays were conducted with the unacclimated methanogenic culture and dye decolorization kinetics were determined as a function of initial dye, biomass, and carbon source concentrations. Dye decolorization was inhibited at initial dye concentrations higher than 100 mg l-1 and decolorization kinetics were described based on the Haldane model. The effect of long-term culture exposure to the reactive dye on decolorization kinetics, culture acclimation, as well as possible dye mineralization was tested using two reactors fed weekly for two years with an initial dye concentration of 300 mg l-1 and a mixture of dextrin/peptone. The maximum dye decolorization rate after a 2-year acclimation at an initial dye concentration of 300 mg l-1 was more than 10-fold higher as compared to that obtained with the unacclimated culture. Aniline and the o-aminohydroxynaphthalene derivative resulting from the reductive azo bond cleavage of the dye were detected, but further transformation(s) leading to dye mineralization were not observed. Reactive Red 2 did not serve as the carbon and energy source for the mixed culture, and dye decolorization was sustained by the continuous addition of dextrin and peptone. Thus, biological decolorization of reactive azo dyes is feasible under conditions of low redox potential created and maintained in overall methanogenic systems, but supply of a biodegradable carbon source is necessary.  相似文献   

15.
The present study deals with the decolorization of Acid Black 210 by a bioluminescent bacterium, Vibrio harveyi TEMS1, isolated from coastal seawater of Izmir Bay, Turkey. Maximum rate of decolorization of Acid Black 210 was observed when Luria Bertani medium was used. Decolorization of Acid Black 210 was 38.9% and 93.9% at 24 h under shaking and static conditions, respectively. The optimum dye-decolorizing activity of the culture was obtained at 100 ppm initial dye concentration and incubation temperature of 20°C. Vibrio harveyi TEMS1 was also tested for its ability to decolorize four azo dyes (Acid Black 24, Acid Blue 7, Acid Green 20, Acid Yellow 36) in addition to Acid Black 210.  相似文献   

16.
Successful decolorization of azo dyes (Orange II, Amido Black 10, Reactive Black 5, and Reactive Red 120) and industrial textile dye influents and effluents with sulfate-reducing bacteria from within a biosulfidogenic reactor was achieved with decolorizations ranging from 96% to 49% over 144 h. Concomitant with the decrease in absorbance of the dye in the visible region (480-620 nm) was an increase in the absorbance at 280 nm, over 48 h, suggesting an increase in concentration of single aromatic amines. With an extended period of time there was a subsequent decrease in the absorbance at 280 nm indicating that the aromatic amines had been degraded. The anthraquinone dye, Reactive Blue 2, remained unchanged after 144 h of incubation in the biosulfidogenic reactor and was only rapidly decolored at 192 h, implying that certain factors are induced in the reactor to break down this non-azo dye. The fastest decolorization/degradation rates and highest hydrogenase enzyme production were observed with Orange II, while the slowest decolorization/degradation rate and least enzyme production were with Reactive Blue 2, suggesting that these processes are controlled, to a certain degree, by an enzymatic mechanism. With sulfate-reducing bacteria that had been cultured on a lactate medium, there was complete decolorization of both authentic dyes and industrial influents and effluents as monitored by the decrease of absorbance in the visible region (480-620 nm). There was, however, very little breakdown of the single aromatic compounds as the absorbance at 280 nm remained fairly significant. This supports the suggestion that, within the biosulfidogenic reactor, there are factors other than the identified hydrogenases that are responsible for degradation of the aromatic compounds.  相似文献   

17.
Presence of huge amount of salts in the wastewater of textile dyeing industry is one of the major limiting factors in the development of an effective biotreatment system for the removal of azo dyes from textile effluents. Bacterial spp. capable of thriving under high salt conditions could be employed for the treatment of saline dye-contaminated textile wastewaters. The present study was aimed at isolating the most efficient bacterial strains capable of decolorizing azo dyes under high saline conditions. Fifty-eight bacterial strains were isolated from seawater, seawater sediment, and saline soil, using mineral salt medium enriched with 100?mg?l?1 Reactive Black-5 azo dye and 50?g NaCl l?1 salt concentration. Bacterial strains KS23 (Psychrobacter alimentarius) and KS26 (Staphylococcus equorum) isolated from seawater sediment were able to decolorize three reactive dyes including Reactive Black 5, Reactive Golden Ovifix, and Reactive Blue BRS very efficiently in liquid medium over a wide range of salt concentration (0–100?g NaCl l?1). Time required for complete decolorization of 100?mg dye l?1 varied with the type of dye and salt concentration. In general, there was an inverse linear relationship between the velocity of the decolorization reaction (V) and salt concentration. This study suggested that bacteria isolated from saline conditions such as seawater sediment could be used in designing a bioreactor for the treatment of textile effluent containing high concentration of salts.  相似文献   

18.
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   

19.
【目的】在无营养条件下,利用白腐真菌绒毛栓孔菌(Trametes pubescens)菌丝体对染料进行脱色可减少试验成本,提高染料处理的实用性。【方法】将该菌株液体培养的菌丝体在无营养条件下对染料进行脱色,并对其中脱色效果较好的偶氮染料刚果红的脱色过程进行分析。在此过程中,测定了该菌株分泌的胞外胞内酶活力,优化影响因子如初始pH值、温度、染料浓度和盐度,同时利用气相色谱-质谱联用技术分析无营养条件下偶氮染料刚果红的降解产物。植物毒性试验测定刚果红经绒毛栓孔菌菌丝体脱色前后的毒性变化。【结果】菌丝体对偶氮染料刚果红有较好的脱色效果,在初始pH值为2.0,温度为30°C,染料浓度为80 mg/L,盐度为2.5%(质量体积比)时,150 r/min转速下培养7 d后脱色率可达80.52%。在此过程中,菌丝体可被连续使用2次,且其所分泌的酶系可降解染料。此外,通过气相色谱-质谱联用分析得到刚果红的降解产物为萘胺、联苯胺和叠氮萘。植物毒性试验显示在无营养条件下的绒毛栓孔菌菌丝体对染料有明显的脱毒作用。【结论】研究发现绒毛栓孔菌菌丝体在无营养条件下的偶氮染料废水处理中具有广阔的应用前景。  相似文献   

20.
The nonviable biomass of Aspergillus niger, Aspergillus japonica, Rhizopus nigricans, Rhizopus arrhizus, and Saccharomyces cerevisiae were screened for biosorption of textile dyes. The selected anionic reactive dyes were C.I. Reactive Black 8, C.I. Reactive Brown 9, C.I. Reactive Green 19, C.I. Reactive Blue 38, and C.I. Reactive Blue 3. Experiments were conducted at initial dye concentration of 50, 100, 150 and 200mg/L. The effect of initial dye concentration, dose of biosorbent loading, temperature, and pH on adsorption kinetics was studied. S. cerevisiae and R. nigricans were good biosorbents at initial dye concentration of 50mg/L, 1g% (w/v) biomass loading and 29+/-1 degrees C. R. nigricans adsorbed 90-96% dye in 15min, at 20 degrees C and pH 6.0. The data showed an optimal fit to the Langmuir and Freundlich isotherms. The maximum uptake capacity (Q(o)) for the selected dyes was in the range 112-204mg/g biomass.  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