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.
The role of DT-diaphorase (DTD, EC 1.6.99.2) in the bioreductive activation of mitomycin C was examined using purified rat hepatic DTD. The formation of adducts with reduced glutathione (GSH), binding of [3H]mitomycin C to DNA, and mitomycin C-induced DNA interstrand cross-linking were used as indicators of bioactivation. Mitomycin C was metabolized by DTD in a pH-dependent manner with increasing amounts of metabolism observed as the pH was decreased from 7.8 to 5.8. The major metabolite observed during DTD-mediated reduction of mitomycin C was 2,7-diaminomitosene. GSH adduct formation, binding of [3H]mitomycin C and mitomycin C-induced DNA interstrand cross-linking were observed during DTD-mediated metabolism. In agreement with the pH dependence of metabolism, increased bioactivation was observed at lower pH values. Temporal studies and experiments using authentic material showed that 2,7-diaminomitosene could be further metabolized by DTD resulting in the formation of mitosene adducts with GSH. DNA cross-linking during either chemical (sodium borohydride) or enzymatic (DTD) mediated reduction of mitomycin C could be observed at pH 7.4, but it increased as the pH was decreased to 5.8, showing the critical role of pH in the cross-linking process. These data provide unequivocal evidence that the obligate two-electron reductase DTD can bioactivate mitomycin C to reactive species which can form adducts with GSH and DNA and induce DNA cross-linking. The use of mitomycin C may be a viable approach to the therapy of tumors high in DTD activity, particularly when combined with strategies to lower tumor pH.  相似文献   

2.
The cytotoxic action of the antitumor antibiotic mitomycin C occurs primarily at the level of DNA. Using highly sensitive fluorescence assays which depend on the enhancement of ethidium fluorescence only when it intercalates duplex regions of DNA, three aspects of mitomycin C action on DNA have been studied: (a) cross-linking events, (b) alkylation without necessarily cross-linking, and (c) strand breakage. Cross-linking of DNA is determined by the return of fluorescence after a heat denaturation step at alkaline pH's. Under these conditions denatured DNA gives no fluorescence. The cross-linking was independently confirmed by S1-endonuclease (EC 3.1.4.-) digestion. At relatively high concentrations of mitomycin the suppression of ethidium fluorescence enhancement was shown not to be due to depurination but rather to alkylation, as a result of losses in potential intercalation sites. A linear relationship exists between binding ratio for mitomycin and loss of fluorescence. The proportional decrease in fluorescence with pH strongly suggests that the alkylation is due to the aziridine moiety of the antibiotic under these conditions. A parallel increase in the rate and overall efficiency of covalent cross-linking of DNA with lower pH suggests that the cross-linking event, to which the primary cytotoxic action has been linked, occurs sequentially with alkylation by aziridine and then by carbamate. Mitomycin C, reduced chemically, was shown to induce single strand cleavage as well as monoaklylation and covalent cross-linking in PM2 covalently closed circular DNA. The inhibition of this cleavage by superoxide dismutase (EC 1.15.1.1) and catalase (EC 1.11.1.6), and by free radical scavengers suggests that the degradation of DNA observed to accompany the cytotoxic action of mitomycin C is largely due to the free radical O2. In contrast to the behavior of the antibiotic streptonigrin, mitomycin C does not inactivate the protective enzymes superoxide dismutase or catalase. Lastly, mitomycin C is able to cross-link DNA in the absence of reduction at pH 4. This is consistent with the postulated cross-linking mechansims.  相似文献   

3.
We have assayed the cross-linking of oligonucleotides containing repeated mitomycin-reactive CpG sites in order to assess the factors that enhance activation of the carbamoyl function at C10, yielding efficient mitomycin cross-linking. Drugs studied include mitomycin C (MC), N-methylmitomycin A (NMA), and the aziridinomitosene of NMA (MS). Drugs were reduced both by catalytic hydrogenation and by diothionite. We find that cross-linking by fully reduced NMA can be increased severalfold by addition of either excess dithionite reductant or the oxidant FeCl3. Enhancement by FeCl3 is not seen with MC or MS, but excess dithionite increases cross-linking by all three compounds. We explain the action of Fe3+ by postulating production of the semiquinone of the monoadduct of mitomycin reacted at the C1-position; according to this mechanism, departure of the carbamate from C10 is more efficient for the semiquinone than for the hydroquinone. However, our results imply that the hydroquinone can also function as a cross-linking agent. Excess dithionite, beyond that required for stoichiometric reduction, activates the carbamate 2-3-fold for cross-linking. We find that the fully reduced leucoaziridinomitosene is highly unstable in solution, yet it produces efficient cross-liking. Hence, this compound is highly reactive in DNA alkylation and a good candidate for the role of primary alkylating agent.  相似文献   

4.
K Ueda  T Komano 《Nucleic acids research》1984,12(17):6673-6683
Mitomycin C reduced with sodium borohydride induced the DNA damage at deoxyguanosines preferentially in dinucleotide sequence G-T. The DNA damage produced strand breaks when subsequently heated. The DNA damage scarcely occurred when the end-labeled DNA was preincubated with ethidium bromide or actinomycin D before the addition of mitomycin C and the reducing agent. Fully reduced mitomycin C did not induce the DNA damage. The mitomycin C-inducing DNA damage seems to require the intercalation of the partially reduced mitomycin C of short life time, probably semiquinone radical, between DNA base pairs. The inhibitory effects of sodium chloride and radical scavengers suggested that the requirement of the covalent bond formation of mitomycin C to DNA and the involvement of oxygen radicals in the DNA damage. 7-N-(p-hydroxyphenyl)mitomycin C, which is reported to show a higher antitumor activity and a lower toxicity than mitomycin C, was readily reduced with dithiothreitol and induced the sequence-specific DNA damage, whereas mitomycin C was not.  相似文献   

5.
The kinetic relation between the photoinactivation and photooxidation of mitomycin C in the presence of riboflavin was investigated. The photoinactivation was tested for lambda-phage induction in Escherichia coli K-12 (lambda) cells and colony formation of E. coli Bs-1 cells. Mitomycin C lost its phage-inducing and antibiotic activities when the antibiotic was irradiated in vitro with visible light in the presence of riboflavin. The loss of phage-inducing activity followed a Stern-Volmer type equation with respect to the dose of irradiation, and the inactivation constant was evaluated to be 0.96X10(-4)m2/J. The initial rate of decay of mitomycin C by photooxidation in the presence of riboflavin obeyed first order kinetics, and its cross section was estimated to be 2.51X10(-6)m/J independent of the intensity of incident light. The cross section for photooxidation was found to be proportional to the inactivation constant. These results suggest that the photoinactivation of mitomycin C is caused by its photooxidation. In order to rationalize this conclusion, a mechanism of photooxidation was proposed and reactions in vivo of the photoproduct were discussed in relation to the inactivation.  相似文献   

6.
The anti-cancer drug mitomycin C is metabolically activated to bind and cross-link DNA. The cross-linking contributes significantly to the cytotoxicity. The complex chemical structure of mitomycin C allows its metabolism by several known (cytosolic NAD(P)H:quinone oxidoreductase and microsomal NADPH:cytochrome P450 reductase) and unknown enzymes. The identification of new enzymes/proteins that metabolize mitomycin C and like drugs is an area of significant research interest since these studies have direct implications in drug development and clinical usage. In the present studies, we have investigated a role of cytosolic glucose regulatory protein GRP58 in mitomycin C-induced DNA cross-linking and cytotoxicity. The control and GRP58 siRNA were transfected in human colon carcinoma HCT116 cells in culture. The transfection of GRP58 siRNA but not control siRNA significantly inhibited GRP58 in human colon carcinoma HCT116 cells. The inhibition of GRP58 led to decrease in mitomycin C-induced DNA cross-linking and cytotoxicity. These results establish a role of GRP58 in mitomycin C-induced DNA cross-linking and cytotoxicity. Site-directed mutagenesis of cysteines to serines in thioredoxin domains of GRP58 and cross-linking assays revealed that both N- and C-terminal thioredoxin domains are required for GRP58-mediated mitomycin C-induced DNA cross-linking. These results suggest that GRP58 might be an important target enzyme for further studies on mitomycin C and similar drug therapy.  相似文献   

7.
Mammalian NADPH-ferredoxin reductase (EC 1.18.1.2) functions in the mitochondrial electron transport chain for cytochrome P-450-dependent steroid hydroxylation. Significant homology of three-dimensional structure exists in the surroundings of FAD between NADPH-ferredoxin reductase and NADH-cytochrome b5 reductase. The latter is involved in the bioreduction of mitomycin C (MC), a prototype antitumor agent. In this study, we assessed the capacity of NADPH-ferredoxin reductase to activate MC. Mitomycin C increased the NADPH oxidase activity of NADPH-ferredoxin reductase. In the absence of ferredoxin, the Km value of NADPH-ferredoxin reductase for MC was 73.5 +/- 2.3 microM. While in the presence of 500 nM ferredoxin, a Lineweaver-Burk plot exhibited a biphasic curve. NADPH-ferredoxin reductase-mediated reduction of MC resulted in the formation of an alkylated complex of 4-(p-nitrobenzyl) pyridine and an increase in plasmide DNA single-strand breaks under hypoxic conditions. With the addition of 500 nM ferredoxin, the amount of the alkylated complex of 4-(p-nitrobenzyl) pyridine and the plasmide DNA single-strand breaks increased by 40% and 37%, respectively. However, neither alkylated complex of 4-(p-nitrobenzyl) pyridine nor DNA strand breaks was observed in the presence of SOD and catalase under aerobic conditions. These findings demonstrate that NADPH-ferredoxin reductase is capable of catalyzing the bioactivation of mitomycin C under hypoxic conditions in vitro.  相似文献   

8.
The mitomycins are a group of antitumor antibiotics that covalently bind to DNA upon reductive activation. Mitomycin A (1b; MA) is more toxic than its clinically useful mitomycin C (1a; MC). The greater toxicity of mitomycin A has been previously attributed to its higher reduction potential. In this report, the DNA alkylation products of reductively activated MA were isolated and characterized by conversion to the known 7-amino mitosene-deoxyguanosine adducts. The three major adducts formed were identified as a monoadduct, N2-(2"beta-amino-7"-methoxymitosen-1"alpha-yl)- 2'-deoxyguanosine (5), a decarbamoyl monoadduct, N2-(2"beta-amino-10"-decarbamoyl-7"-methoxymitosen-1"alpha-y l)-2'- deoxyguanosine (6), and a bisadduct, N2-(2"beta-amino-10"-deoxyguanosin-N2-yl-7-methoxymitosen-1" alpha- yl)-2'-deoxyguanosine (7). Under all reductive activation conditions employed, MA selectively alkylated the 2-amino group of guanine in DNA, like MC. In addition, both MA and MC alkylated DNA and cross-linked oligonucleotides to a similar extent. However, variations in the reductive activation conditions (H2/PtO2, Na2S2O4, or enzymatic) affected the distribution of the three major MA adducts in a different manner than the distribution of MC adducts was affected. A mechanism is proposed wherein the 7-methoxy substituent of MA allows initial indiscriminate activation of either of the drugs' two electrophilic sites. While oxygen inhibited cross-linking by MC, similar aerobic conditions exhibited little influence on the cross-linking ability of MA. Hence, the greater toxicity of MA may be influenced by increased and nonselective activation and cross-link formation in both aerobic and anaerobic cells. This effect is a direct consequence of the higher redox potential of MA as compared to MC.  相似文献   

9.
Sequence specificity of heat-labile sites in DNA induced by mitomycin C   总被引:4,自引:0,他引:4  
K Ueda  J Morita  T Komano 《Biochemistry》1984,23(8):1634-1640
The sequence specificity of the mitomycin C-DNA interaction was directly determined by using DNA sequencing techniques and by using 3'- or 5'-end-labeled DNA fragments of defined sequence as substrates. Mitomycin C reduced with sodium borohydride induced heat-labile sites in DNA preferentially at specific sequences. The heat-labile sites were induced most preferentially at the dinucleotide sequence G-T ( especially Pu G-T), which was determined by scanning autoradiograms with a microdensitometer after gel electrophoresis. DNA was cleaved at the 3' side of deoxyguanosines and of some deoxyadenosines by heat treatment. Oligonucleotides produced by heat treatment after reaction with reduced mitomycin C contained phosphoryl groups at the 5' termini. The 3' termini seemed not to have simple structures, judging from their electrophoretic mobilities. Oxygen radicals such as singlet oxygen and hydroxyl radical were possibly involved in the induction of heat-labile sites.  相似文献   

10.
The in vivo efficacy and ultrastructural effects of mitomycin C were determined against alveolar hydatid disease in experimentally infected animals and compared to mebendazole treatment. Mitomycin C inhibited the mean cyst mass of treated versus control animals by 84.1% which was statistically significant at the alpha = 0.01 level. Mebendazole given daily inhibited the mean cyst mass by 80.1%, while mebendazole administration on the same treatment schedule as that used for mitomycin C inhibited the mean cyst mass by 70.4%. Ultrastructurally, mitomycin C was not observed to affect the tegumental microtriches (microvilli) or the microtubular system. However, an increase in the number and accumulation of round to oval electrondense vesicles was observed within the subtegument. These inclusion bodies became vacuolated, subsequently degenerated, and formed myelin-like figures. Mitomycin C, like mebendazole, was only cystistatic in its effects on the cyst stage of Echinococcus multilocularis as evidenced by the growth of treated cyst material following inoculation into helminth-free animals.  相似文献   

11.
Mitomycin C (MC) and its structural analogs porfiromycin (PM), BMY-25282 and BL-6783 are toxic to EMT6 cells under aerobic and hypoxic conditions. The mitomycin antibiotics are hypothesized to exert cytotoxicity under hypoxic conditions by cross-linking DNA following reductive activation, while aerobic cytotoxicity may involve DNA cross-linking by these agents and/or damage due to the generation of oxygen radicals. Previous findings (Pritsos and Sartorelli, 1986) indicated that the rank order of cytotoxicity for a series of mitomycins was the same as the rank order for the rate of oxygen consumption induced by these agents. As an additional approach to explore the role of oxygen radicals in the aerobic cytotoxicity of the four agents studied, EMT6 cells were treated with the mitomycins in the presence of the superoxide dismutase inhibitor diethyldithiocarbamate (DETC). DETC, which decreased superoxide dismutase activity in EMT6 cells, increased the cytotoxicity of BMY-25282 and BL-6783 by half an order of magnitude, but did not affect the toxicity of PM or MC to these cells. DNA cross-links, a proposed cytotoxic lesion induced by BMY-25282, however, were not detectably increased in EMT6 cells exposed to this agent in the presence of DETC in spite of the large increase in cytotoxicity under these treatment conditions. No single strand breaks were detected in cells exposed to either BMY-25282 or BMY-25282 plus DETC. The findings support the concept that oxygen radicals may have a role in the aerobic cytotoxicity of some of the mitomycin antibiotics, and that the lesions responsible for cytotoxicity produced by oxygen radicals may not reside entirely at the level of DNA.  相似文献   

12.
D M Peterson  J Fisher 《Biochemistry》1986,25(14):4077-4084
Mitomycin c in the presence of NADPH and brewers' yeast NADPH: (acceptor) oxidoreductase (Old Yellow enzyme, EC 1.6.99.1) is transformed, at pH 8.0 and with anaerobicity, to two major mitosene products (the cis- and trans-1-hydroxy-2,7-diaminomitosenes; respective yields of 45 and 30%). These arise by covalent trapping by solvent of a quinone methide intermediate, obtained by rearrangement of the mitomycin c hydroquinone. At lower pH (6.5), the major product of this reaction is 2,7-diaminomitosene, which arises by covalent trapping of the quinone methide by H+. In the former instance the quinone methide acts as an electrophile and in the latter as a nucleophile. A detailed kinetic analysis indicates that the role of the NADPH and Old Yellow enzyme is to initiate an autocatalytic reaction, propagated by mitomycin c reduction by the mitosene hydroquinones (arising by the electrophilic pathway). The evidence supporting this conclusion is the formation of oxidized mitosene products, under the rigorously anaerobic reaction circumstance, the nonstoichiometric participation of NADPH, a dependence of the velocity on the total mitomycin c concentration, the pH dependence of the reaction, and the accurate simulation of the overall kinetic course with a mathematical model of the autocatalytic pathway. These observations indicate that the autocatalytic pathway may be dominant during in vitro mitomycin c anaerobic reductive activation by other reducing agents and that (as with anthracycline reductive activation) oxidation of the mitosene hydroquinone obtained from nucleophile addition to the quinone methide may be a necessary event for the formation of stable covalent adducts in vivo.  相似文献   

13.
Oligodeoxyribonucleotides cross-linked by reductively activated mitomycin C (MC) were prepared and purified for the first time. The cross-linked products were structurally characterized by nucleoside and MC-nucleoside adduct analysis. Optimal conditions were established for the cross-linking reaction, resulting in high yields, typically in the 20-50% range. Nuclease digests of the cross-linked oligonucleotides yielded the same bifunctional MC-deoxyguanosine adduct as that previously isolated from DNA exposed to MC in vitro and in vivo [Tomasz et al. (1987) Science 235, 1204]. The cross-linked oligonucleotides displayed broad thermal melting profiles, greatly increased Tm, and complex circular dichroism spectra. Phosphodiester linkages at the cross-link were resistant to spleen exonuclease, nuclease P1, and TaqI and ClaI restriction endonucleases; snake venom diesterase action was uninhibited. The cross-links are stable to heat at neutral pH but are removed by treatment in hot piperidine or by the reducing agents Na2S2O4 and dithiothreitol. Mechanisms are proposed for these reactions. These studies define optimal methods for introducing mitomycin cross-links into DNA fragments at a specific site, providing a versatile tool to study the effects of the MC cross-links on DNA structure and function.  相似文献   

14.
Reassignment of the guanine-binding mode of reduced mitomycin C   总被引:1,自引:0,他引:1  
Mitomycin C (1) is a clinically used antitumor antibiotic that binds covalently to deoxyribonucleic acid under reductive or acidic catalysis. We have determined the structures of the adducts resulting from attack of reductively activated 1 on the dinucleoside phosphate d(GpC) to be N2-(2' beta, 7'-diaminomitosen-1'alpha-yl)-2'-deoxyguanosine (2) and its 1' beta-isomer (3). This represents a revision of the previously reported structures for these adducts in that the mitomycin residue is linked to the N2- rather than O6-position of 2'-deoxyguanosine. This revision is the result of applying to the mitomycin case a newly developed general method that leads to unambiguous assignment of the linkage position in complex alkylated guanosines. The method as described here takes advantage of the resolution enhancement gained by calculation of the second derivatives of absorbance Fourier transform infrared spectra. In addition, we present 1H NMR data that corroborate the assigned structures of 2 and 3 and that should serve as a useful reference for future investigations into the binding of mitomycin C to DNA. The convenient synthesis of adducts 2 and 3 from deoxyguanosine and mitomycin C reported here should facilitate such investigations as well. Furthermore, we demonstrate a useful acetylation procedure for adducts and metabolites of mitomycin C that furnishes spectroscopically superior chemical derivatives (e.g., triacetates 4 and 5, derived from acetylation of adducts 2 and 3).  相似文献   

15.
Summary Mitomycin C, a known inhibitor of DNA synthesis, was injected into white prepupae ofPhormia regina, Adults which developed from these prepupae showed alterations of the bristle pattern, loss of whole bristle organs, and the formation of bristles without sockets or sockets without bristle shafts. Dose-dependence was found for all modifications. For the abdominal microchaetae, the period of maximum sensitivity to the drug began at 16 h after puparium formation, that is well after all of the macrochaetae and most of the microchaetae of the thorax and the head had grown insensitive. Bristle forming trichogen and tormogen cells developed high degrees of polyteny with distinctly banded chromosomes. Photometric determination of the amount of Feulgen-DNA per nucleus led to estimations of DNA classes ranging from 256C to 2048 C. DNA contents of nuclei from Mitomycin C treated animals were significantly lower during the actual growth of the bristle apparatus, but reached approximately the same level as the controls prior to the time of emergence. Cytological investigations proved that doses of Mitomycin C which yielded bristle organs either without sockets or without shafts do not affect the differential division of the bristle mother cell. Polytene chromosomes damaged by Mitomycin C displayed a diffuse and irregular banding pattern. Possible modes of action of Mitomycin C on replicating polytene chromosomes are discussed.  相似文献   

16.
Different variants of the comet assay were used to study the genotoxic and cytotoxic properties of the following eight compounds: chloral hydrate, colchicine, hydroquinone, DL-menthol, mitomycin C, sodium iodoacetate, thimerosal and valinomycin. Colchicine, mitomycin C, sodium iodoacetate and thimerosal induced genotoxic effects. The other compounds were found to be inactive. The compounds were tested in the standard comet assay as well as in the all cell comet assay (recovery of floating cells after treatment), designed in our laboratory for adherently-growing cells. This latter procedure proved to be more adequate for the assessment of the cytotoxicity for some of the compounds tested (hydroquinone, DL-menthol, thimerosal, valinomycin). Colchicine was positive in the standard comet assay (3h treatment) and in the all cell comet assay (24h treatment). Sodium iodoacetate and thimerosal were positive in the standard and/or the all cell comet assay. Chloral hydrate, hydroquinone, sodium iodoacetate, mitomycin C and thimerosal were also tested in the modified comet assay using lysed cells. Mitomycin C and thimerosal showed effects in this assay, whereas sodium iodoacetate was inactive. This indicates that it does not induce direct DNA damage. Compounds that are known or suspected to form DNA-DNA cross-links or DNA-protein cross-links (chloral hydrate, hydroquinone, mitomycin C and thimerosal) were checked for their ability to reduce ethyl methanesulfonate (EMS)-induced DNA damage. This mode of action could be demonstrated for mitomycin C only.  相似文献   

17.
The effect of mitomycin C on the accumulation of specific mRNAs was studied in asynchronously growing Swiss 3T3 cells, as well as in synchronously growing serum stimulated ts13 cells (a temperature-sensitive mutant from the BHK cell line). It was observed that the steady-state level of p53 RNA experienced some increase in 3T3 cells treated for 24 h with the drug. In addition, mitomycin when applied to serum stimulated ts13 cells increased the level of p2F1 RNA. Mitomycin diminished the level of core histone H3 RNA, a finding consistent with the inhibitory action of this compound on DNA replication.  相似文献   

18.
Mitomycin C, a DNA cross-linking agent, was administered for a week intraperitoneally to a normal 9-week-old rat in which hepatocyte proliferative activity had almost ceased. In the rat treated with mitomycin C, the number of polyploid cells with the DNA amounts more than 4C was not increased in comparison with that in the control rat without any treatment. However, the partial hepatectomy in the rat pretreated with mitomycin C provoked prominent hepatocyte polyploidization much greater in degree than that found in the hepatectomized rat without mitomycin C treatment. These results seem to indicate that the existence of cross-linking DNA damages is a latent factor necessary for the induction of abortive mitosis of a cell after completion of DNA synthesis, which results in the production of a mononuclear polyploid cell in one-step higher DNA class. Cross-linking DNA damages and DNA synthesis are the essential factors for the manifestation of polyploidization.  相似文献   

19.
Addition of different concentrations of sodium arsenite to the fermentation medium vised for the production of mitomycin antibiotics byStreptomyces caespitosus hindered the biosynthesis of mitomycins and led to the accumulation of 2-oxoglutarate, pyruvate and acetone. Mitomycin C isolated and purified using thin-layer chromatography in low concentration of about 0.1 μg/ml did not affect the RNA, DNA and protein biosynthesis of the growingBacillus subtilis, while at 10 μg/ml mitomycin C markedly affected RUA, DNA and protein biosynthesis.  相似文献   

20.
DNA sequence specificity of mitomycin cross-linking   总被引:2,自引:0,他引:2  
Using a gel electrophoresis assay, we show that the target DNA sequence cross-linked by N-methylmitomycin A, its aziridinomitosene, and mitomycin C is CpG, in strong preference over GpC. The yield per CpG site increases as the number of successive CpG sequences increases. Molecular modeling reveals no systematic difference between the energies of mitomycin cross-links at CpG in comparison with GpC. However, the distance between guanine amino groups in CpG sequences is nearly the same as the distance in the cross-linked adduct, whereas the amino group separation at GpC sites is substantially larger in the starting DNA than in the adduct. We suggest that the favorable placement of the second reaction center in CpG greatly accelerates the second step in the cross-linking reaction. As shown by a competition assay, mitomycins bind A-T and G-C sequences noncovalently equally well, even though the only sequence that yields appreciable cross-linking is CpG. N-Methylmitomycin A and its aziridinomitosene are found to be better cross-linking agents than is mitomycin C.  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