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.
Several toluene monooxygenase-producing organisms were tested for their ability to oxidize linear alkenes and chloroalkenes three to eight carbons long. Each of the wild-type organisms degraded all of the alkenes that were tested. Epoxides were produced during the oxidation of butene, butadiene, and pentene but not hexene or octadiene. A strain of Escherichia coli expressing the cloned toluene-4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 was able to oxidize butene, butadiene, pentene, and hexene but not octadiene, producing epoxides from all of the substrates that were oxidized. A T4MO-deficient variant of P. mendocina KR1 oxidized alkenes that were five to eight carbons long, but no epoxides were detected, suggesting the presence of multiple alkene-degrading enzymes in this organism. The alkene oxidation rates varied widely (ranging from 0. 01 to 0.33 micromol of substrate/min/mg of cell protein) and were specific for each organism-substrate pair. The enantiomeric purity of the epoxide products also varied widely, ranging from 54 to >90% of a single epoxide enantiomer. In the absence of more preferred substrates, such as toluene or alkenes, the epoxides underwent further toluene monooxygenase-catalyzed transformations, forming products that were not identified.  相似文献   

2.
Summary Newly isolated and already available strains of alkene-utilizing bacteria were able to oxidize ethene, propene or 1-butene to the respective 1,2-epoxides. Resting-cell suspensions of organisms isolated on propene and butene, when grown on these substrates converted ethene quantitatively to epoxyethane. Some, but not all ethene-utilizing strains accumulated 1,2-epoxypropane or 1,2-epoxybutane when propene or butene was supplied, although not quantitatively because the epoxides produced were partially further metabolized. Suitable epoxide producers which eventually may be employed as biocatalysts in a biotechnological process were used for immobilization in calcium alginate and K-carrageenan; after immobilization, 60%–100% activity for epoxide production was retained.  相似文献   

3.
We assayed the tolerance to solvents of three toluene-degrading Pseudomonas putida strains and Pseudomonas mendocina KR1 in liquid and soil systems. P. putida DOT-T1 tolerated concentrations of heptane, propylbenzene, octanol, and toluene of at least 10% (vol/vol), while P. putida F1 and EEZ15 grew well in the presence of 1% (vol/vol) propylbenzene or 10% (vol/vol) heptane, but not in the presence of similar concentrations of octanol or toluene. P. mendocina KR1 grew only in the presence of heptane. All three P. putida strains were able to become established in a fluvisol soil from the Granada, Spain, area, whereas P. mendocina KR1 did not survive in this soil. The tolerance to organic solvents of all three P. putida strains was therefore assayed in soil. The addition to soil of 10% (vol/wt) heptane or 10% (vol/wt) propylbenzene did not affect the survival of the three P. putida strains. However, the addition of 10% (vol/wt) toluene led to an immediate decrease of several log units in the number of CFU per gram of soil for all of the strains, although P. putida F1 and DOT-T1 subsequently recovered. This recovery was influenced by the humidity of the soil and the incubation temperature. P. putida DOT-T1 recovered from the shock faster than P. putida F1; this allowed the former strain to become established at higher densities in polluted sites into which both strains had been introduced.  相似文献   

4.
N-Nitrosodimethylamine (NDMA) is a potent carcinogen and an emerging contaminant in groundwater and drinking water. The metabolism of NDMA in mammalian cells has been widely studied, but little information is available concerning the microbial transformation of this compound. The objective of this study was to elucidate the pathway(s) of NDMA biotransformation by Pseudomonas mendocina KR1, a strain that possesses toluene-4-monooxygenase (T4MO). P. mendocina KR1 was observed to initially oxidize NDMA to N-nitrodimethylamine (NTDMA), a novel metabolite. The use of 18O2 and H218O revealed that the oxygen added to NDMA to produce NTDMA was derived from atmospheric O2. Experiments performed with a pseudomonad expressing cloned T4MO confirmed that T4MO catalyzes this initial reaction. The NTDMA produced by P. mendocina KR1 did not accumulate, but rather it was metabolized further to produce N-nitromethylamine (88 to 94% recovery) and a trace amount of formaldehyde (HCHO). Small quantities of methanol (CH3OH) were also detected when the strain was incubated with NDMA but not during incubation with either NTDMA or HCHO. The formation of methanol is hypothesized to occur via a second, minor pathway mediated by an initial α-hydroxylation of the nitrosamine. Strain KR1 did not grow on NDMA or mineralize significant quantities of the compound to carbon dioxide, suggesting that the degradation process is cometabolic.  相似文献   

5.
Eleven strains of alkene-utilizing bacteria belonging to the genera Mycobacterium, Nocardia, and Xanthobacter were tested for their ability to grow with C1 to C6 alkanes, C2 to C6 alkenes, alkadienes, and monoterpenes furnished individually as sole sources of carbon and energy in a mineral salts medium. A limited number of alkenes and alkanes supported growth of the bacteria; some bacteria were unable to grow on any of the saturated hydrocarbons tested. Monoterpenes were frequently used as carbon and energy sources by alkene-utilizing bacteria belonging to the genera Mycobacterium and Nocardia. Washed cell suspensions of alkene-grown bacteria attacked the whole range of alkenes tested, whereas only three strains were able to oxidize alkanes as well. The alkenes tested were oxidized either to water and carbon dioxide or to epoxyalkanes. Few epoxides accumulated in stoichiometric amounts from the corresponding alkenes, because most epoxides formed were further converted to other compounds like alkanediols.  相似文献   

6.
The bioremediation of polluted groundwater and toxic waste sites requires that bacteria come into close physical contact with pollutants. This can be accomplished by chemotaxis. Five motile strains of bacteria that use five different pathways to degrade toluene were tested for their ability to detect and swim towards this pollutant. Three of the five strains (Pseudomonas putida F1, Ralstonia pickettii PKO1, and Burkholderia cepacia G4) were attracted to toluene. In each case, the response was dependent on induction by growth with toluene. Pseudomonas mendocina KR1 and P. putida PaW15 did not show a convincing response. The chemotactic responses of P. putida F1 to a variety of toxic aromatic hydrocarbons and chlorinated aliphatic compounds were examined. Compounds that are growth substrates for P. putida F1, including benzene and ethylbenzene, were chemoattractants. P. putida F1 was also attracted to trichloroethylene (TCE), which is not a growth substrate but is dechlorinated and detoxified by P. putida F1. Mutant strains of P. putida F1 that do not oxidize toluene were attracted to toluene, indicating that toluene itself and not a metabolite was the compound detected. The two-component response regulator pair TodS and TodT, which control expression of the toluene degradation genes in P. putida F1, were required for the response. This demonstration that soil bacteria can sense and swim towards the toxic compounds toluene, benzene, TCE, and related chemicals suggests that the introduction of chemotactic bacteria into selected polluted sites may accelerate bioremediation processes.  相似文献   

7.
Sixteen new cultures of propane-utilizing bacteria were isolated from lake water from Warinanco Park, Linden, N.J. and from lake and soil samples from Bayway Refinery, Linden, N.J. In addition, 19 known cultures obtained from culture collections were also found to be able to grow on propane as the sole carbon and energy source. In addition to their ability to oxidize n-alkanes, resting-cell suspensions of both new cultures and known cultures grown on propane oxidize short-chain alkenes to their corresponding 1,2-epoxides. Among the substrate alkenes, propylene was oxidized at the highest rate. In contrast to the case with methylotrophic bacteria, the product epoxides are further metabolized. Propane and other gaseous n-alkanes inhibit the epoxidation of propylene. The optimum conditions for in vivo epoxidation are described. Results from inhibition studies indicate that a propane monooxygenase system catalyzes both the epoxidation and hydroxylation reactions. Experiments with cell-free extracts show that both hydroxylation and epoxidation activities are located in the soluble fraction obtained after 80,000 × g centrifugation.  相似文献   

8.
Pure cultures of methylotrophs and methanotrophs are known to oxidize methyl bromide (MeBr); however, their ability to oxidize tropospheric concentrations (parts per trillion by volume [pptv]) has not been tested. Methylotrophs and methanotrophs were able to consume MeBr provided at levels that mimicked the tropospheric mixing ratio of MeBr (12 pptv) at equilibrium with surface waters (≈2 pM). Kinetic investigations using picomolar concentrations of MeBr in a continuously stirred tank reactor (CSTR) were performed using strain IMB-1 and Leisingeria methylohalidivorans strain MB2T — terrestrial and marine methylotrophs capable of halorespiration. First-order uptake of MeBr with no indication of threshold was observed for both strains. Strain MB2T displayed saturation kinetics in batch experiments using micromolar MeBr concentrations, with an apparent Ks of 2.4 μM MeBr and a Vmax of 1.6 nmol h−1 (106 cells)−1. Apparent first-order degradation rate constants measured with the CSTR were consistent with kinetic parameters determined in batch experiments, which used 35- to 1 × 107-fold-higher MeBr concentrations. Ruegeria algicola (a phylogenetic relative of strain MB2T), the common heterotrophs Escherichia coli and Bacillus pumilus, and a toluene oxidizer, Pseudomonas mendocina KR1, were also tested. These bacteria showed no significant consumption of 12 pptv MeBr; thus, the ability to consume ambient mixing ratios of MeBr was limited to C1 compound-oxidizing bacteria in this study. Aerobic C1 bacteria may provide model organisms for the biological oxidation of tropospheric MeBr in soils and waters.  相似文献   

9.
Fatty epoxides are used as plasticizers and plastic stabilizers and are intermediates for the production of other chemical substances. The currently used industrial procedure for fatty epoxide synthesis requires a strong acid catalyst which can cause oxirane ring opening and side product formation. To find a replacement for the acid catalyst, we have been conducting research on a peroxygenase enzyme from oat (Avena sativa) seeds and have devised a method for immobilization of this enzyme using a hydrophobic membrane support. In this study, fatty acids and fatty methyl esters commonly encountered in commercial vegetable oils were tested as substrates for immobilized peroxygenase, and the epoxide products were characterized. The epoxidation time course of linoleic acid showed two distinct phases with nearly complete conversion to monoepoxide before diepoxide was produced. The diepoxide formed from linolenic acid was found to be 9,10-15,16-diepoxy-12-octadecenoic acid, and only a trace of triepoxide was obtained. Additionally it was discovered that acyclic alkenes with internal double bonds, a cyclic alkene, and an alkene with an aromatic substituent were substrates of peroxygenase. However, alkenes with terminal unsaturation were unreactive. With every substrate examined, oat seed peroxygenase exhibited specificity for epoxidation, producing no other products, and oxirane ring opening did not occur.  相似文献   

10.
Wild type, mutant, and recombinant bacterial strains capable of oxidizing aromatic hydrocarbons were screened for their ability to oxidize anisole (methoxybenzene) and phenetole (ethoxybenzene). Toluene-induced cells ofPseudomonas putida F39/D transformed anisole to a compound tentatively identified ascis-1,2-dihydroxy-3-methoxyclohexa-3,5-diene (anisole-2,3-dihydrodiol), 2-methoxyphenol, catechol, and trace amounts of phenol while phenetole was converted primarily tocis-1,2-dihydroxy-3-ethoxycyclohexa-3,5-diene (phenetole-2,3-dihydrodiol) and 2-ethoxyphenol. Induced cells ofPseudomonas sp. NCIB 9816/11 andBeijerinckia sp. B8/36 transformed anisole to phenol, and phenetole to phenol and ethenyloxybenzene. Toluene-induced cells ofP. putida BG1 converted anisole to phenol but did not oxidize phenetole. In contrast, toluene-induced cells ofP. mendocina KR1, which oxidize toluene via monooxygenation at thepara position, transformed anisole to 4-methoxyphenol, and phenetole to 2-, 3- and 4-ethoxyphenol. The involvement of toluene and naphthalene dioxygenases in the reactions catalyzed by strains F39/D and NCIB 9816/11, respectively, was confirmed with recombinantE. coli strains expressing the cloned dioxygenase genes. The results show that the oxygenases from differentPseudomonas strains oxidize anisole and phenetole to different hydroxylated products.  相似文献   

11.
Wild-type toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 oxidizes toluene to p-cresol (96%) and oxidizes benzene sequentially to phenol, to catechol, and to 1,2,3-trihydroxybenzene. In this study T4MO was found to oxidize o-cresol to 3-methylcatechol (91%) and methylhydroquinone (9%), to oxidize m-cresol and p-cresol to 4-methylcatechol (100%), and to oxidize o-methoxyphenol to 4-methoxyresorcinol (87%), 3-methoxycatechol (11%), and methoxyhydroquinone (2%). Apparent Vmax values of 6.6 ± 0.9 to 10.7 ± 0.1 nmol/min/ mg of protein were obtained for o-, m-, and p-cresol oxidation by wild-type T4MO, which are comparable to the toluene oxidation rate (15.1 ± 0.8 nmol/min/mg of protein). After these new reactions were discovered, saturation mutagenesis was performed near the diiron catalytic center at positions I100, G103, and A107 of the alpha subunit of the hydroxylase (TmoA) based on directed evolution of the related toluene o-monooxygenase of Burkholderia cepacia G4 (K. A. Canada, S. Iwashita, H. Shim, and T. K. Wood, J. Bacteriol. 184:344-349, 2002) and a previously reported T4MO G103L regiospecific mutant (K. H. Mitchell, J. M. Studts, and B. G. Fox, Biochemistry 41:3176-3188, 2002). By using o-cresol and o-methoxyphenol as model substrates, regiospecific mutants of T4MO were created; for example, TmoA variant G103A/A107S produced 3-methylcatechol (98%) from o-cresol twofold faster and produced 3-methoxycatechol (82%) from 1 mM o-methoxyphenol seven times faster than the wild-type T4MO (1.5 ± 0.2 versus 0.21 ± 0.01 nmol/min/mg of protein). Variant I100L produced 3-methoxycatechol from o-methoxyphenol four times faster than wild-type T4MO, and G103S/A107T produced methylhydroquinone (92%) from o-cresol fourfold faster than wild-type T4MO and there was 10 times more in terms of the percentage of the product. Variant G103S produced 40-fold more methoxyhydroquinone from o-methoxyphenol than the wild-type enzyme produced (80 versus 2%) and produced methylhydroquinone (80%) from o-cresol. Hence, the regiospecific oxidation of o-methoxyphenol and o-cresol was changed for significant synthesis of 3-methoxycatechol, methoxyhydroquinone, 3-methylcatechol, and methylhydroquinone. The enzyme variants also demonstrated altered monohydroxylation regiospecificity for toluene; for example, G103S/A107G formed 82% o-cresol, so saturation mutagenesis converted T4MO into an ortho-hydroxylating enzyme. Furthermore, G103S/A107T formed 100% p-cresol from toluene; hence, a better para-hydroxylating enzyme than wild-type T4MO was formed. Structure homology modeling suggested that hydrogen bonding interactions of the hydroxyl groups of altered residues S103, S107, and T107 influence the regiospecificity of the oxygenase reaction.  相似文献   

12.
The enzymatic synthesis of α,β-halohydrins from gaseous alkenes is described. The enzymatic reaction required an alkene, a halide ion, dilute hydrogen peroxide, and a haloperoxidase enzyme. A wide range of gaseous alkenes were suitable for this reaction, including those containing isolated, conjugated, and cumulative carbon-carbon double bonds. Chlorohydrins, bromohydrins, and iodohydrins could be formed. The combining of this enzymatic synthesis with a previously described enzymatic synthesis of epoxides from α,β-halohydrins provides an alternate pathway, other than the well-known enzymatic direct epoxidation pathway, from alkene to an epoxide.  相似文献   

13.
Enantiopure sulfoxides are valuable asymmetric starting materials and are important chiral auxiliaries in organic synthesis. Toluene monooxygenases (TMOs) have been shown previously to catalyze regioselective hydroxylation of substituted benzenes and phenols. Here we show that TMOs are also capable of performing enantioselective oxidation reactions of aromatic sulfides. Mutagenesis of position V106 in the α-hydroxylase subunit of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 and the analogous position I100 in toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 improved both rate and enantioselectivity. Variant TomA3 V106M of TOM oxidized methyl phenyl sulfide to the corresponding sulfoxide at a rate of 3.0 nmol/min/mg protein compared with 1.6 for the wild-type enzyme, and the enantiomeric excess (pro-S) increased from 51% for the wild type to 88% for this mutant. Similarly, T4MO variant TmoA I100G increased the wild-type oxidation rate by 1.7-fold, and the enantiomeric excess rose from 86% to 98% (pro-S). Both wild-type enzymes showed lower activity with methyl para-tolyl sulfide as a substrate, but the improvement in the activity and enantioselectivity of the mutants was more dramatic. For example, T4MO variant TmoA I100G oxidized methyl para-tolyl sulfide 11 times faster than the wild type did and changed the selectivity from 41% pro-R to 77% pro-S. A correlation between regioselectivity and enantioselectivity was shown for TMOs studied in this work. Using in silico homology modeling, it is shown that residue I100 in T4MO aids in steering the substrate into the active site at the end of the long entrance channel. It is further hypothesized that the main function of V106 in TOM is the proper positioning or docking of the substrate with respect to the diiron atoms. The results from this work suggest that when the substrate is not aligned correctly in the active site, the oxidation rate is decreased and enantioselectivity is impaired, resulting in products with both chiral configurations.  相似文献   

14.
The solvent-tolerant strain Pseudomonas putida DOT-T1E has been engineered for biotransformation of toluene into 4-hydroxybenzoate (4-HBA). P. putida DOT-T1E transforms toluene into 3-methylcatechol in a reaction catalyzed by toluene dioxygenase. The todC1C2 genes encode the α and β subunits of the multicomponent enzyme toluene dioxygenase, which catalyzes the first step in the Tod pathway of toluene catabolism. A DOT-T1EΔtodC mutant strain was constructed by homologous recombination and was shown to be unable to use toluene as a sole carbon source. The P. putida pobA gene, whose product is responsible for the hydroxylation of 4-HBA into 3,4-hydroxybenzoate, was cloned by complementation of a Pseudomonas mendocina pobA1 pobA2 double mutant. This pobA gene was knocked out in vitro and used to generate a double mutant, DOT-T1EΔtodCpobA, that was unable to use either toluene or 4-HBA as a carbon source. The tmo and pcu genes from P. mendocina KR1, which catalyze the transformation of toluene into 4-HBA through a combination of the toluene 4-monoxygenase pathway and oxidation of p-cresol into the hydroxylated carboxylic acid, were subcloned in mini-Tn5Tc and stably recruited in the chromosome of DOT-T1EΔtodCpobA. Expression of the tmo and pcu genes took place in a DOT-T1E background due to cross-activation of the tmo promoter by the two-component signal transduction system TodST. Several independent isolates that accumulated 4-HBA in the supernatant from toluene were analyzed. Differences were observed in these clones in the time required for detection of 4-HBA and in the amount of this compound accumulated in the supernatant. The fastest and most noticeable accumulation of 4-HBA (12 mM) was found with a clone designated DOT-T1E-24.  相似文献   

15.
Summary Methanotrophic bacteria have been shown to oxidize gaseous alkenes to the corresponding epoxides utilizing an NADH2-dependent methane monooxygenase. A cell paste of methane-grown methylotrophs was coated on porous glass beads. The production of propylene oxide from propylene was performed in a gas-solid bioreactor to ensure continuous production and removal of product epoxide from the microenvironment of the biocatalyst. The amount of propylene oxide produced before cofactor regeneration was between 120–145 moles/20 mg cells in about 10 h depending on the microbial strains used. The conversion rate for propylene was 2.7%. Regeneration of cofactor NADH2 was performed in the bioreactor with the vapor of a cosubstrate, methanol.  相似文献   

16.
Epoxide hydratase was solubilized from human liver microsomal fractions and purified to an extent where the specific activity was 40-fold greater than that of the liver homogenate. Combination of homogenate and purified preparation showed that the increase in activity was not due to the removal of an inhibitor. Monosubstituted oxiranes with a lipophilic substituent larger than an ethyl group (isopropyl, t-butyl, n-hexyl, phenyl) readily interacted as substrates or inhibitors with this purified human epoxide hydratase, whereas those with a small substituent (methyl, ethyl, vinyl) were inactive, probably reflecting greater affinity of the former epoxides owing to lipophilic binding sites near the active site of the enzyme. In a series of oxiranes having a lipophilic substituent of sufficient size (styrene oxides), monosubstituted as well as 1,1- and cis-1,2-disubstituted oxiranes readily served as substrates or inhibitors of the enzyme, but not the trans-1,2-disubstituted, tri- or tetra-substituted oxiranes. trans-Substitution at the oxirane ring apparently prevents access of the oxirane ring to the active site by steric hindrance. Epoxide hydratase was also solubilized from microsomal fractions of rat and guinea-pig liver and purified by the same procedure. Structural requirements for effective interaction of substrates, inhibitors and activators were qualitatively identical for epoxide hydratase from the three sources. However, several quantitative differences were observed. Thus human hepatic epoxide hydratase seems to be very similar to, although not identical with, the enzyme from guinea pig or rat. Studies with epoxide hydratase from the latter two species therefore appear to be significant with respect to man. In addition, knowledge of structural requirements for epoxides to serve as substrates for human epoxide hydratase may prove useful for drug design. Compounds which need aromatic or olefinic moieties for their desired effect would not be expected to lead to accumulation of epoxides if their structure was such as to allow for a metabolically produced epoxide to be rapidly consumed by epoxide hydratase.  相似文献   

17.
Unspecific peroxygenases (EC 1.11.2.1) represent a group of secreted heme-thiolate proteins that are capable of catalyzing the mono-oxygenation of diverse organic compounds, using only H2O2 as a co-substrate. Here we show that the peroxygenase secreted by the fungus Agrocybe aegerita catalyzed the oxidation of 20 different alkenes. Five branched alkenes, among them 2,3-dimethyl-2-butene and cis-2-butene, as well as propene and butadiene were epoxidized with complete regioselectivity. Longer linear alkenes with a terminal double bond (e.g. 1-octene) and cyclic alkenes (e.g. cyclohexene) were converted into the corresponding epoxides and allylic hydroxylation products; oxidation of the cyclic monoterpene limonene yielded three oxygenation products (two epoxides and an alcohol). In the case of 1-alkenes, the conversion occurred with moderate stereoselectivity, in which the preponderance for the (S)-enantiomer reached up to 72% ee for the epoxide product. The apparent Michaelis–Menten constant (Km) for the epoxidation of the model substrate 2-methyl-2-butene was 5 mM, the turnover number (kcat) 1.3 × 103 s?1 and the calculated catalytic efficiency, kcat/Km, was 2.5 × 105 M?1 s?1. As epoxides represent chemical building blocks of high relevance, new enzymatic epoxidation pathways are of interest to complement existing chemical and biotechnological approaches. Stable and versatile peroxygenases as that of A. aegerita may form a promising biocatalytic platform for the development of such enzyme-based syntheses.  相似文献   

18.
Aerobic bacterial growth on aromatic hydrocarbons typically requires oxygenase enzymes, which are known to fortuitously oxidize nongrowth substrates. In this study, we found that oxidation of diethyl ether by toluene 2-monooxygenase supported more rapid growth of Burkholderia cepacia G4/PR1 than did the aromatic substrates n-propylbenzene and o-xylene. The wild-type Burkholderia cepacia G4 failed to grow on diethyl ether. Purified toluene 2-monooxygenase protein components oxidized diethyl ether stoichiometrically to ethanol and acetaldehyde. Butyl methyl ether, diethyl sulfide, and 2-chloroethyl ethyl ether were oxidized by B. cepacia G4/PR1.  相似文献   

19.
We developed a biocatalyst by cloning the styrene monooxygenase genes (styA and styB) from Pseudomonas fluorescens ST responsible for the oxidation of styrene to its corresponding epoxide. Recombinant Escherichia coli was able to oxidize different aryl vinyl and aryl ethenyl compounds to their corresponding optically pure epoxides. The results of bioconversions indicate the broad substrate preference of styrene monooxygenase and its potential for the production of several fine chemicals.  相似文献   

20.
The use of peptide mass fingerprinting with matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS) was demonstrated to identify and phenotypically characterize toluene-degrading bacteria via biomarkers of degradation and taxonomical classification. Pseudomonas putida F1, P. mendocina KR1, and Burkholderia sp. JS150 were grown on toluene, extracted, electrophoretically separated, and analyzed by MALDI-TOF MS. Catabolic enzymes were identified and results substantiated using tandem MS.  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