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.
Quinones are widely used as medicines or redox agents. The chemical properties are based on the reactions against an electron donor. 9,10-Phenanthraquinone (PQ), which is a quinone contaminated in airborne particulate matters, forms redox cycling, not Michael addition, with electron donors. Redox cycling of PQ contributes to its toxicity, following generation of reactive oxygen species (ROS). Detoxification of quinones is generally thought to be two-electron reduction forming hydroquinones. However, a hydroquinone of PQ, 9,10-dihydroxyphenanthrene (PQH(2)), has been never detected itself, because it is quite unstable. In this paper, we succeeded in detecting PQH(2) as its stable derivative, 9,10-diacetoxyphenanthrene (DAP). However, higher concentrations of PQ (>4 microM) form disproportionately with PQH(2), producing the 9,10-phenanthraquinone radical (PQ(-)) which is a one-electron reducing product of PQ. In cellular experiments using DAP as a precursor of PQH(2), it was shown that PQH(2) plays a critical role in the oxidative protein damage and cellular toxicity of PQ, showing that two-electron reduction of PQ can also initiate redox cycling to cause oxidative stress-dependent cytotoxicity.  相似文献   

2.
9,10-Phenanthraquinone (PQ), a major quinone contained in diesel exhaust particles and atmospheric PM(2.5), undergoes one-electron reduction by flavin enzymes such as NADPH-cytochrome P450 reductase, leading to production of reactive oxygen species in vitro. We have detected an ESR signal for superoxide (O(2)(-)) and hydroxyl radicals ((.)OH) by the spin trap method when PQ was mixed with P450 reductase, NADPH, and iron(III). When we examined the effects of PQ on A549 human pulmonary epithelial cells, PQ induced apoptosis with a LC(50) of approximately 7 microM. Formation of protein carbonyls was also detected in cells after treatment with PQ, suggesting that PQ induces oxidative damage. Iron chelators such as 1,10-phenanthroline (OP), desferrioxamine mesylate, and deferiprone respectively afforded protection against the toxic effects of PQ. Furthermore, treatment of A549 cells with 10-20 microM PQ for 12 h specifically down-regulated protein levels of Cu,Zn-superoxide dismutase (Cu,Zn-SOD) and heme oxygenase-1 (HO-1) by more than 50%. Pretreatment of cells with OP (10 microM) markedly reduced the down-regulation of Cu,Zn-SOD and HO-1 and protein carbonyl formation in response to PQ. The inhibitor of Cu,Zn-SOD, diethyldithiocarbamate, enhanced the toxic effects of 5 microM PQ. The present findings suggest that PQ causes iron-mediated oxidative damage that is exacerbated by the concomitant down-regulation of Cu,Zn-SOD.  相似文献   

3.
The effects of adrenochrome, a metabolite of epinephrine (adrenaline), and 9,10-phenanthrenequinone (PQ), a component of diesel exhaust particles, on the stereoselective reduction of 4-benzoylpyridine (4-BP) were examined in pig heart cytosol. PQ was a potent inhibitor for the 4-BP reduction, while adrenochrome was a poor inhibitor. A similar result was observed in the effects of adrenochrome and PQ on the reduction of all-trans retinal. Furthermore, although PQ mediated efficiently the formation of superoxide anion radical through its redox cycling in pig heart cytosol, adrenochrome had no ability to mediate the superoxide formation. These may be because the reactivity for adrenochrome, catalyzed by pig heart carbonyl reductase (PHCR), is much lower than that for PQ. The optimal pH for the reduction of PQ in pig heart cytosol was around 5.5. Dicumarol, a potent inhibitor of DT-diaphorase, had little effect on the time course of NADPH oxidation during the reduction of PQ. Therefore, it is concluded that PHCR plays a critical role in superoxide formation through redox cycling of PQ.  相似文献   

4.
Two-electron oxidation of 2,2,5,7,8-pentamethyl-6-chromanol (PH), a model compound of alpha-tocopherol, gives the unstable 8a-hydroxy-2,2,5,7,8-pentamethyl-6-chromanone (POH) which rearranges to form stable 2-(3-hydroxy-3-methylbutyl)-3,5,6-trimethyl-1,4-benzoquinone (PQ). POH and PQ are isomers which have the same oxidation state. The aim of this work was to compare the ease of reduction of POH and PQ at room temperature using a variety of biological and chemical reductants in a reductant:POH (PQ) ratio of 20:1 (or 16:1). Ascorbic acid completely reduced POH to PH in 20 min, but had no effect on PQ after 40 min. Sodium ascorbate did not reduce POH or PQ at all after 40 min. Sodium dithionite reduced POH to PH (85%) in 20 min, but reduced PQ to its hydroquinone, PQH2 (67%) in 40 min. Dithiothreitol produced a slight reduction of POH to PH (21%) but reduced PQ to PQH2 (69%). NADH/FAD reduced POH and PQ to PQH2 (73% and 42%, respectively) in 10 min. It was concluded that POH is easier to reduce than PQ and more likely to form PH as a product, particularly under conditions of mild acidity.  相似文献   

5.
The oxidation of the PQ-pool after illumination with 50 or 500 micromol quantam(-2)s(-1) was measured in isolated thylakoids as the increase in DeltaA(263), i.e., as the appearance of PQ. While it was not observed under anaerobic conditions, under aerobic conditions it was biphasic. The first faster phase constituted 26% or 44% of total reappearance of PQ, after weak or strong light respectively. The dependence on oxygen presence as well as the correlation with the rate of oxygen consumption led to conclusion that this phase represents the appearance of PQ from PQ(*-) produced in the course of PQH(2) oxidation by superoxide accumulated in the light within the membrane.  相似文献   

6.
Photoinactivation of photosystem II (PSII) and light-dependent degradation of the reaction center II (RCII) protein D1 have been investigated in Chlamydomonas reinhardtii mutants D6, AC208, and B4 deficient in cytochrome b6/f, plastocyanin, and photosystem I (PSI) activity, respectively. These mutants possess active PSII and reduce plastoquinone (PQ) but cannot oxidize plastoquinol (PQH2) via light-dependent reduction of NADP. In light-exposed cells a high ratio PQH2/PQ and a low turnover of PQ/PQH2 at the RCII-QB site are maintained. In all mutants photoinactivation of RCII was slower as compared to the wild-type (wt) cells, and D1 degradation was drastically decreased. The degradation of D1 was also lower in the wt cells under anaerobic conditions and presence of ascorbate, while raising the concentration of dissolved oxygen increased the degradation of the D1 protein in the AC208 mutant. Photoinactivation and light-dependent degradation of the D1 protein were drastically increased in the Scenedesmus obliquus LF-1 mutant cells altered in its PSII manganese binding and thus unable to reduce PQ using water as an electron donor. Diuron inhibited the light-dependent degradation of D1 protein in both the LF-1 mutant and wt cells. Based on these results we propose that availability of PQ at the QB site is required for (i) the photoinactivation process of the RCII acceptor side followed by inactivation of the donor side leading to the generation of harmful cation radicals (Z+, P680+, chlz+) which damage the D1 protein, and (ii) the accessibility of the cleavage site of the damaged D1 protein to proteolytic degradation.  相似文献   

7.
Here we use crystal structures to investigate and review channels and pathways for the transfer of substrates (water, plastoquinone (PQ)) and products (electrons, protons, oxygen, reduced PQ (PQH(2))) to, and from, the redox active catalytic sites of photosystem II (PSII). A putative oxygen channel has been identified which is about 21A in length, leading from the water splitting site to the lumen. This channel follows a path along the lumenal surface of CP43, passing across the interface of the large extrinsic loop which joins the fifth and sixth transmembrane helices of this chlorophyll binding protein. In so doing it seems to minimise interactions with the excited states of chlorophylls bound within the PSII complex, especially those that constitute the primary electron donor, P680. Two additional channels leading from the water splitting site, and also exiting at the lumen, were also identified. Their hydrophilic nature suggests that they probably facilitate the delivery of water to, and protons from, the catalytic site. Also discussed are unique features in the electron transfer pathway of PSII, as compared with those of purple photosynthetic bacteria, and structural implications of the PSII Q(B)-site in terms of PQ protonation and PQ/PQH(2) diffusion.  相似文献   

8.
9.
It has been found that the solubility of PQH2 and alpha-TQH2 in hexane increased with the increase in PQ and alpha-TQ concentrations, respectively, that is connected with the formation of quinhydrone-type charge-transfer complexes. Measurements of the solubility of both prenylquinones and their reduced forms in hexane and acetone, at -30 degrees C and room temperature, showed a much higher affinity of the quinol forms for acetone than for hexane. In the case of quinones, the difference in affinity was not significant. The possibility of charge-transfer complex formation by PQ and alpha-TQ in thylakoid membranes and the influence of such complexes on the diffusion of PQH2 and alpha-TQH2 molecules have been considered.  相似文献   

10.
It has been found that plastoquinone (PQ) and alpha-tocopherol quinone (alpha-TQ) can form quinhydrone-type charge-transfer complexes on PQH2 and alpha-TQH2, respectively, both in the crystalline state and in solutions of organic solvents. The charge-transfer spectra of PQ/PQH2 mixtures in hydrophobic solvents showed two bands: one at 349-358 nm, the other at 430-440 nm, one charge-transfer band at 351-355 nm occurring in water-miscible solvents. The intensity ratio of these two bands varied with changing PQ/PQH2 ratio. The charge-transfer spectra of alpha-TQ/alpha-TQH2 mixtures in all solvents investigated showed one peak at 361-367 nm and a broad shoulder within the range 400-540 nm, whose shape varied depending on the solvent used. In the infrared spectrum of PQ and alpha-TQ (1700-1600 cm(-1)) splitting of the carbonyl band occurred and was caused by the presence of two peak. In the spectra of quinhydrones the splitting disappeared, this being brought about by the appearance of a new peak at the position of splitting, which originated from the complexed quinones. The possibility of the formation of such complexes in thylakoid membranes is discussed.  相似文献   

11.
Paraquat (PQ) is a well-known pneumotoxicant that exerts its toxic effect by elevating intracellular levels of superoxide. In addition, production of pro-inflammatory cytokines has possibly been linked to PQ-induced inflammatory processes through reactive oxygen species (ROSs) and nitric oxide (NO). However, the role of NO in PQ-induced cell injury has been controversial. To explore this problem, we examined the effect of NO on A549 cells by exposing them to the exogenous NO donor NOC18 or to cytokines; tumor necrosis factor-alpha, interleukin-1 beta and interferon-gamma, as well as PQ. Although the exogenous NO donor on its own had no effect on the release of lactate dehydrogenase (LDH), remarkable release was observed when the cells were exposed to high concentrations of NOC18 and PQ. This cellular damage caused by 1 mM NOC18 plus 0.2 mM PQ was ascertained by phase contrast microscopy. On the other hand, NO derived from 25-50 microM NOC18 added into the medium improved the MTT reduction activity of mitochondria, suggesting a beneficial effect of NO on the cells. Incubation of A549 cells with cytokines increased in inducible NO synthase (iNOS) expression and nitrite accumulation, resulting in LDH release. PQ further potentiated this release. The increase in nitrite levels could be completely prevented by NOS inhibitors, while the leakage of LDH was not attenuated by the inhibition of NO production with them. On the other hand, ROS scavenging enzymes, superoxide dismutase and catalase, inhibited the leakage of LDH, whereas they had no effect on the increase in the nitrite level. These results indicate that superoxide, not NO, played a key role in the cellular damage caused by PQ/cytokines. Our in vitro models demonstrate that NO has both beneficial and deleterious actions, depending on the concentrations produced and model system used.  相似文献   

12.
Paraquat (PQ) is a well-known pneumotoxicant that exerts its toxic effect by elevating intracellular levels of superoxide. In addition, production of pro-inflammatory cytokines has possibly been linked to PQ-induced inflammatory processes through reactive oxygen species (ROSs) and nitric oxide (NO). However, the role of NO in PQ-induced cell injury has been controversial. To explore this problem, we examined the effect of NO on A549 cells by exposing them to the exogenous NO donor NOC18 or to cytokines; tumor necrosis factor-α, interleukin-1 β and interferon-γ, as well as PQ. Although the exogenous NO donor on its own had no effect on the release of lactate dehydrogenase (LDH), remarkable release was observed when the cells were exposed to high concentrations of NOC18 and PQ. This cellular damage caused by 1 mM NOC18 plus 0.2 mM PQ was ascertained by phase contrast microscopy. On the other hand, NO derived from 25–50 μM NOC18 added into the medium improved the MTT reduction activity of mitochondria, suggesting a beneficial effect of NO on the cells. Incubation of A549 cells with cytokines increased in inducible NO synthase (iNOS) expression and nitrite accumulation, resulting in LDH release. PQ further potentiated this release. The increase in nitrite levels could be completely prevented by NOS inhibitors, while the leakage of LDH was not attenuated by the inhibition of NO production with them. On the other hand, ROS scavenging enzymes, superoxide dismutase and catalase, inhibited the leakage of LDH, whereas they had no effect on the increase in the nitrite level. These results indicate that superoxide, not NO, played a key role in the cellular damage caused by PQ/cytokines. Our in vitro models demonstrate that NO has both beneficial and deleterious actions, depending on the concentrations produced and model system used.  相似文献   

13.
9,10-Phenanthrenequinone (9,10-PQ), a major quinone in diesel exhaust particles, induces apoptosis via the generation of reactive oxygen species (ROS) because of 9,10-PQ redox cycling. We have found that intratracheal infusion of 9,10-PQ facilitates the secretion of surfactant into rat alveolus. In the cultured rat lung, treatment with 9,10-PQ results in an increase in a lower-density surfactant by ROS generation through redox cycling of the quinone. The surfactant contains aldo-keto reductase (AKR) 1C15, which reduces 9,10-PQ and the enzyme level in the surfactant increases on treatment with 9,10-PQ suggesting an involvement of AKR1C15 in the redox cycling of the quinone. In six human cell types (A549, MKN45, Caco2, Hela, Molt4 and U937) only type II epithelial A549 cells secrete three human AKR1C subfamily members (AKR1C1, AKR1C2 and AKR1C3) with the surfactant into the medium; this secretion is highly increased by 9,10-PQ treatment. Using in vitro enzyme inhibition analysis, we have identified AKR1C3 as the most abundantly secreted AKR1C member. The AKR1C enzymes in the medium efficiently reduce 9,10-PQ and initiate its redox cycling accompanied by ROS production. The exposure of A549 cells to 9,10-PQ provokes viability loss, which is significantly protected by the addition of the AKR1C3 inhibitor and antioxidant enzyme and by the removal of the surfactants from the culture medium. Thus, the AKR1C enzymes secreted in pulmonary surfactants probably participate in the toxic mechanism triggered by 9,10-PQ.  相似文献   

14.
The ligninolytic fungus Phanerochaete chrysosporium oxidized phenanthrene and phenanthrene-9,10-quinone (PQ) at their C-9 and C-10 positions to give a ring-fission product, 2,2'-diphenic acid (DPA), which was identified in chromatographic and isotope dilution experiments. DPA formation from phenanthrene was somewhat greater in low-nitrogen (ligninolytic) cultures than in high-nitrogen (nonligninolytic) cultures and did not occur in uninoculated cultures. The oxidation of PQ to DPA involved both fungal and abiotic mechanisms, was unaffected by the level of nitrogen added, and was significantly faster than the cleavage of phenanthrene to DPA. Phenanthrene-trans-9,10-dihydrodiol, which was previously shown to be the principal phenanthrene metabolite in nonligninolytic P. chrysosporium cultures, was not formed in the ligninolytic cultures employed here. These results suggest that phenanthrene degradation by ligninolytic P. chrysosporium proceeds in order from phenanthrene----PQ----DPA, involves both ligninolytic and nonligninolytic enzymes, and is not initiated by a classical microsomal cytochrome P-450. The extracellular lignin peroxidases of P. chrysosporium were not able to oxidize phenanthrene in vitro and therefore are also unlikely to catalyze the first step of phenanthrene degradation in vivo. Both phenanthrene and PQ were mineralized to similar extents by the fungus, which supports the intermediacy of PQ in phenanthrene degradation, but both compounds were mineralized significantly less than the structurally related lignin peroxidase substrate pyrene was.  相似文献   

15.
16.
The formation of beta-D-glucopyranosides (glucuronides) by the UDP-glucuronosyltransferases (UGTs) is a significant metabolic pathway that facilitates the elimination of small hydrophobic molecules such as drugs, dietary constituents, steroids, and bile acids. We elucidate here that an anti-oxidative response leads to induction of UGT1A1 through the Nrf2-Keap1 pathway. When human HepG2 cells were treated with the prooxidants tert-butylhydroquinone and beta-naphthoflavone, cellular UGT1A1 glucuronidation activities were increased. The induction of UGT1A1 proceeded following the overexpression of Nrf2 and was blocked following overexpression of Keap1, demonstrating that Keap1 suppresses Nrf2 activation of the UGT1A1 gene. Loss of function analysis for Nrf2 conducted by small interfering RNA revealed that induction of UGT1A1 was not seen in Nrf2 knock-out cells. To examine the contribution of oxidants toward the regulation of human UGT1A1 in vivo, transgenic mice bearing the human UGT1 locus (Tg-UGT1) were treated with tert-butylhydroquinone. Human UGT1A1 was markedly increased in small and large intestines as well as in liver. Gene mapping experiments including transfections of UGT1A1 reporter gene constructs into HepG2 cells coupled with functional analysis of Nrf2 expression and binding to anti-oxidant-response elements (ARE) resulted in identification of an ARE in the phenobarbital-response enhancer module region of the UGT1A1 gene. The ARE flanks the recently identified Ah receptor xenobiotic-responsive element. The results suggest that Nrf2-Keap1-dependent UGT1A1 induction by prooxidants might represent a key adaptive response to cellular oxidative stress that defends against a variety of environmental insults, including electrophile attacks and chemical carcinogenesis.  相似文献   

17.
The aim of this study is to investigate the molecular mechanisms underlying delayed progressive pulmonary fibrosis, a characteristic of subacute paraquat (PQ) poisoning. Epithelial-mesenchymal transition (EMT) has been proposed as a cause of organ fibrosis, and transforming growth factor-β (TGF-β) is suggested to be a powerful mediator of EMT. We thus examined the possibility that EMT is involved in pulmonary fibrosis during PQ poisoning using A549 human alveolar epithelial cells in vitro. The cells were treated with various concentrations of PQ (0–500 μM) for 2–12 days. Short-term (2 days) high-dose (>100 μM) treatments with PQ induced cell death accompanied by the activation of caspase9 as well as a decrease in E-cadherin (an epithelial cell marker), suggesting apoptotic cell death with the features of anoikis (cell death due to the loss of cell-cell adhesion). In contrast, long-term (6–12 days) low-dose (30 μM) treatments with PQ resulted in a transformation into spindle-shaped mesenchymal-like cells with a decrease of E-cadherin as well as an increase of α-smooth muscle actin (α-SMA). The mesenchymal-like cells also secreted the extracellular matrix (ECM) protein fibronectin into the culture medium. The administration of a TGF-β1 receptor antagonist, SB431542, almost completely attenuated the mesenchymal transformation as well as fibronectin secretion, suggesting a crucial role of TGF-β1 in EMT-like cellular response and subsequent fibrogenesis. It is noteworthy that despite the suppression of EMT-fibrogenesis, apoptotic death was observed in cells treated with PQ+SB431542. EMT-like cellular response and subsequent fibrogenesis were also observed in normal human bronchial epithelial (NHBE) cells exposed to PQ in a TGF-β1-dependent manner. Taken together, our experimental model reflects well the etiology of PQ poisoning in human and shows the involvement of EMT-like cellular response in both fibrogenesis and resistance to cell death during subacute PQ poisoning of pulmonary epithelial cells.  相似文献   

18.
Tumor hypoxia is an important therapeutic target, and it can potentially be exploited by hypoxia-activated prodrugs. However, physiological hypoxia in normal tissues is a limitation. One solution would be to confine activation to severely (pathologically) hypoxic tissue, using hypoxia-activated prodrugs that provide a bystander effect through diffusion of the activated cytotoxin to adjacent regions at intermediate oxygen concentrations (associated with partial radioresistance). To evaluate this requirement, we identified five hypoxia-activated prodrugs with at least 10-fold higher potency against a cell line (A549-P540(puro)) overexpressing human cytochrome P450 reductase (P450R) relative to A549-Lo21 cells with 200-fold lower P450R activity. Bystander killing by these hypoxia-activated prodrugs was tested in anoxic multicellular layer co-cultures of these two cell lines. Cytotoxic potency against A549-Lo21 cells was unaffected by the presence of A549-P450(puro) cells for tirapazamine and RSU-1069 but increased more than 10-fold for the aziridinyldintrobenzamide CB 1954, more than 14-fold for the corresponding nitrogen mustard SN 23862, and 15-fold for its water-soluble analog SN 23816. The cytotoxic extracellular metabolites resulting from hypoxic nitroreduction of CB 1954 and SN 23862 by A549-P450(puro) cells were identified by LC/MS and bioassay methods. For SN 23862, these included the 2-amine metabolite, previously, identified as the bystander metabolite from aerobic activation by the E. coli nfsB nitroreductase, but also novel di-reduced metabolites. Cytotoxicity of SN 23862 to A549-P450(puro) cells was inhibited by lower concentrations of oxygen than for tirapazamine. The combination of selective activation under severe hypoxia with an efficient bystander effect identifies the dinitrobenzamide mustards for further development as hypoxia-activated prodrugs.  相似文献   

19.
Mitochondrial oxidative stress is a contributing factor in the etiology of numerous neuronal disorders. However, the precise mechanism(s) by which mitochondrial reactive oxygen species modify cellular targets to induce neurotoxicity remains unknown. In this study, we determined the role of mitochondrial aconitase (m-aconitase) in neurotoxicity by decreasing its expression. Incubation of the rat dopaminergic cell line, N27, with paraquat (PQ(2+) ) resulted in aconitase inactivation, increased hydrogen peroxide (H(2) O(2) ) and increased ferrous iron (Fe(2+) ) at times preceding cell death. To confirm the role of m-aconitase in dopaminergic cell death, we knocked down m-aconitase expression via RNA interference. Incubation of m-aconitase knockdown N27 cells with PQ(2+) resulted in decreased H(2) O(2) production, Fe(2+) accumulation, and cell death compared with cells expressing basal levels of m-aconitase. To determine the metabolic role of m-aconitase in mediating neuroprotection, we conducted a complete bioenergetic profile. m-Aconitase knockdown N27 cells showed a global decrease in metabolism (glycolysis and oxygen consumption rates) which blocked PQ(2+) -induced H(+) leak and respiratory capacity deficiency. These findings suggest that dopaminergic cells are protected from death by decreasing release of H(2) O(2) and Fe(2+) in addition to decreased cellular metabolism.  相似文献   

20.
Nitropyrene, the predominant nitropolycyclic hydrocarbon found in diesel exhaust, is a mutagenic and tumorigenic environmental pollutant that requires metabolic activation via nitroreduction and ring oxidation. In order to determine the role of ring oxidation in the mutagenicity of 1-nitropyrene, its oxidative metabolites, 1-nitropyrene 4,5-oxide and 1-nitropyrene 9,10-oxide, were synthesized and their mutation spectra were determined in the coding region of hprt gene of CHO cells by a PCR amplification of reverse-transcribed hprt mRNA, followed by a DNA sequence analysis. A comparison of the two metabolites for mutation frequencies showed that 1-nitropyrene 9,10-oxide was 2-times higher than 1-nitropyrene 4,5-oxide. The mutation spectrum for 1-nitropyrene 4,5-oxide was base substitutions (33/49), one base deletions (11/49) and exon deletions (5/49). In the case of 1-nitropyrene 9,10-oxide, base substitutions (27/50), one base deletions (15/50), and exon deletions (8/50) were observed. Base substitutions were distributed randomly throughout the hprt gene. The majority of the base substitutions in mutant from 1-nitropyrene 4,5-oxide treated cells were A-->G transition (15/33) and G-->A transition (8/33). The predominant base substitution, A-->G transition (11/27) and G-->A transition (8/27), were also observed in mutant from 1-nitropyrene 9,10-oxide treated cells. The mutation at the site of adenine and guanine was consistent with the previous results, where the sites of DNA adduct formed by these compounds were predominant at the sites of purines. A comparison of the mutational patterns between 1-nitropyrene 4,5-oxide and 1-nitropyrene 9,10-oxide showed that there were no significant differences in the overall mutational spectrum. These results indicate that each oxidative metabolite exhibits an equal contribution to the mutagenicity of 1-nitropyrene, and ring oxidation of 1-nitropyrene is an important metabolic pathway to the formation of significant lethal DNA lesions.  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