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.
To gain new insights into the function of photosystem II (PSII) herbicides DCMU (a urea herbicide) and bromoxynil (a phenolic herbicide), we have studied their effects in a better understood system, the bacterial photosynthetic reaction center of the terbutryn-resistant mutant T4 of Blastochloris (Bl.) viridis. This mutant is uniquely sensitive to these herbicides. We have used redox potentiometry and time-resolved absorption spectroscopy in the nanosecond and microsecond time scale. At room temperature the P(+)(*)Q(A)(-)(*) charge recombination in the presence of bromoxynil was faster than in the presence of DCMU. Two phases of P(+)(*)Q(A)(-)(*) recombination were observed. In accordance with the literature, the two phases were attributed to two different populations of reaction centers. Although the herbicides did induce small differences in the activation barriers of the charge recombination reactions, these did not explain the large herbicide-induced differences in the kinetics at ambient temperature. Instead, these were attributed to a change in the relative amplitude of the phases, with the fast:slow ratio being approximately 3:1 with bromoxynil and approximately 1:2 with DCMU at 300 K. Redox titrations of Q(A) were performed with and without herbicides at pH 6.5. The E(m) was shifted by approximately -75 mV by bromoxynil and by approximately +55 mV by DCMU. As the titrations were done over a time range that is assumed to be much longer than that for the transition between the two different populations, the potentials measured are considered to be a weighted average of two potentials for Q(A). The influence of the herbicides can thus be considered to be on the equilibrium of the two reaction center forms. This may also be the case in photosystem II.  相似文献   

2.
Crafts , A. S., and S. Yamaguchi . (U. California, Davis.) Absorption of herbicides by roots. Amer. Jour. Bot. 47(4): 248—255. Illus. 1960.—Many herbicides are used through soil. When 2,4-D*2 was applied to culture-solution barley, bean, cotton, and Zebrina plants, there was evidence that the herbicide is held at high concentration by the roots. Very little of the labeled compound moved into the tops of barley, Zebrina and bean; a fair quantity was found in cotton foliage. Barley seedlings allowed to absorb 2,4-D*, ATA*, MH*, urea*, monuron*, dalapon*, simazin*, P32 and IAA* showed interesting differences. All chemicals were highly sorbed by roots, 2,4-D* was moved to tops in least amount, urea was next; the other seven were moved in larger quantities. ATA*, MH*, IAA*, P32, and dalapon* seemed readily phloem mobile; monuron* and simazin* seemed limited to xylem movement. Results are discussed in relation to the mechanism of root uptake.  相似文献   

3.
The redox potential of Q(A) in photosystem II (PSII) is known to be lower by approximately 100 mV in the presence of phenolic herbicides compared with the presence of DCMU-type herbicides. In this study, the structural basis underlying the herbicide effects on the Q(A) redox potential was studied using Fourier transform infrared (FTIR) spectroscopy. Light-induced Q(A)(-)/Q(A) FTIR difference spectra of Mn-depleted PSII membranes in the presence of DCMU, atrazine, terbutryn, and bromacil showed a strong CO stretching peak of Q(A)(-) at 1,479 cm(-1), while binding of phenolic herbicides, bromoxynil and ioxynil, induced a small but clear downshift by approximately 1 cm(-1). The CO peak positions and the small frequency difference were reproduced in the S(2)Q(A)(-)/S(1)Q(A) spectra of oxygen-evolving PSII membranes with DCMU and bromoxynil. The relationship of the CO frequency with herbicide species correlated well with that of the peak temperatures of thermoluminescence due to S(2)Q(A)(-) recombination. Density functional theory calculations of model hydrogen-bonded complexes of plastoquinone radical anion showed that the small shift of the CO frequency is consistent with a change in the hydrogen-bond structure most likely as a change in its strength. The Q(A)(-)/Q(A) spectra in the presence of bromoxynil, and ioxynil, which bear a nitrile group in the phenolic ring, also showed CN stretching bands around 2,210 cm(-1). Comparison with the CN frequencies of bromoxynil in solutions suggested that the phenolic herbicides take a phenotate anion form in the Q(B) pocket. It was proposed that interaction of the phenolic C-O(-) with D1-His215 changes the strength of the hydrogen bond between the CO of Q(A) with D2-His214 via the iron-histidine bridge, causing the decrease in the Q(A) redox potential.  相似文献   

4.
Effect of two photosynthetic inhibitor herbicides, atrazine (both purified and formulated) and [3-(3,4-dichlorophenyl)-1,1-dimethyl urea] (DCMU), on the growth, macromolecular contents, heterocyst frequency, photosynthetic O2 evolution and dark O2 uptake of wild type and multiple herbicide resistant (MHR) strain of diazotrophic cyanobacterium A. variabilis was studied. Cyanobacterial strains showed gradual inhibition in growth with increasing dosage of herbicides. Both wild type and MHR strain tolerated < 6.0 mg L(-1) of atrazine (purified), < 2.0 mg L(-1) of atrazine (formulated) and < 0.4 mg L(-1) of DCMU indicating similar level of herbicide tolerance. Atrazine (pure) (8.0 mg L(-1)) and 4.0 mg L(-1) of atrazine (formulated) were growth inhibitory concentrations (lethal) for both wild type and MHR strain indicating formulated atrazine was more toxic than the purified form. Comparatively lower concentrations of DCMU were found to be lethal for wild type and MHR strain, respectively. Thus, between the two herbicides tested DCMU was more growth toxic than atrazine. At sublethal dosages of herbicides, photosynthetic O2 evolution showed highest inhibition followed by chlorophyll a, phycobhiliproteins and heterocyst differentiation as compared to carotenoid, protein and respiratory O2 uptake.  相似文献   

5.
Some herbicides act by binding to the exchangeable quinone site in the photosystem II (PSII) reaction centre, thus blocking electron transfer. In this article, it is hypothesized that the plant is killed by light-induced oxidative stress initiated by damage caused by formation of singlet oxygen in the reaction centre itself. This occurs when light-induced charge pairs in herbicide-inhibited PSII decay by a charge recombination route involving the formation of a chlorophyll triplet state that is able to activate oxygen. The binding of phenolic herbicides favours this pathway, thus increasing the efficiency of photodamage in this class of herbicides.  相似文献   

6.
Many herbicides of different chemical structure inhibit photosynthetic electron flow by interrupting the photosyn‐thetic electron flow by interrupting the photosynthetic electron transport chain between the primary acceptor (QA) and the secondary acceptor (QB) of photosystem 2 (PS2). Thermoluminescence (TL) originates from PS2, and the bands of the glow curve can be related to the charge recombination between positively charged donors and negatively charged acceptors. The glow curve of TL is strongly influenced by addition of PS2 herbicides. The herbicide treatment shifts the peak position and activation energy of the TL band related to QA, suggesting that herbicide binding affects the midpoint redox potential not only of Q B but also that of QA. On the basis of the band shift the herbicides of various chemical structures can be classified into different “thermodynamical” groups which relfect the differences in the binding properties of these herbicides. As a new approach TL seems to be a useful technique in studying the mechanism and site of action of herbicides that inhibit electron transport of PS2.  相似文献   

7.
The folding through the thylakoid membrane of the D-1 herbicide binding polypeptide and of the homologous D-2 subunit of photosystem II is predicted from comparison of amino acid sequences and hydropathy index plots with the folding of the subunits L and M of a bacterial photosystem. As the functional amino acids involved in Q and Fe binding in the bacterial photosystem of R. viridis, as indicated by the X-ray structure, are conserved in the homologous D-1 and D-2 subunits of photosystem II, a detailed topology of the binding niche of QB and of herbicides on photosystem II is proposed. The model is supported by the observed amino acid changes in herbicide tolerant plants and algae. These changes are all in the binding domain on the matrix side of the D-1 polypeptide, and turn out to be of functional significance in the QB binding.New inhibitors of QB function are described. Their chemical structure, i.e. pyridones, quinolones, chromones and benzodiones, contains the features of the phenolic type herbicides. Their essential elements, -charges at particular atoms, QSAR and steric requirements for optimal inhibitory potency are discussed and compared with the classical herbicides of the urea/triazine type.  相似文献   

8.
Wim F.J. Vermaas  Charles J. Arntzen   《BBA》1983,725(3):483-491
We have analyzed the binding of synthetic quinones and herbicides which inhibit electron transport at the acceptor side of Photosystem II (PS II) of the photosynthetic electron-transport chain in thylakoid membranes. These data show that quinones and PS II-directed herbicides compete for binding to a common binding environment within a PS II region which functions as the Q / PQ oxidoreductase. We observed that (1) synthetic quinones cause a parallel inhibition of electron transport and [14C]herbicide displacement, and (2) herbicide binding is affected both by the fully oxidized and fully reduced form of a quinone. Quinone function and inhibitor binding were also investigated in thylakoids isolated from triazine-resistant weed biotypes. We conclude the following. (1) The affinity of the secondary accepting quinone, B, is decreased in resistant thylakoids. (2) The observation that the equilibrium concentration of reduced Q after transferring one electron to the acceptor side of PS II is increased in resistant as compared to susceptible chloroplasts may be explained both by a decrease in the affinity of PQ for the herbicide / quinone binding environment, and by a decrease of the midpont redox potential of the B / B couple. (3) The binding environment regulating quinone and herbicide affinity may be divided roughly into two domains; we suggest that the domain regulating quinone head-group binding is little changed in resistant membranes, whereas the domain-regulating quinone side-group binding (and atrazine) is altered. This results in increased inhibitory activity of tetrachloro-p-benzoquinone and phenolic herbicides, which are hypothesized to utilize the quinone head-group domain. The two domains appear to be spatially overlapping because efficient atrazine displacement by tetrachloro-p-benzoquinone is observed.  相似文献   

9.
Certain phenolic compounds represent a distinct class of Photosystem (PS) II Q(B) site inhibitors. In this paper, we report a detailed study of the effects of 2,4,6-trinitrophenol (TNP) and other phenolic inhibitors, bromoxynil and dinoseb, on PS II energetics. In intact PS II, phenolic inhibitors bound to only 90-95% of Q(B) sites even at saturating concentrations. The remaining PS II reaction centers (5-10%) showed modified Q(A) to Q(B) electron transfer but were sensitive to urea/triazine inhibitors. The binding of phenolic inhibitors was 30- to 300-fold slower than the urea/triazine class of Q(B) site inhibitors, DCMU and atrazine. In the sensitive centers, the S(2)Q(A)(-) state was 10-fold less stable in the presence of phenolic inhibitors than the urea/triazine herbicides. In addition, the binding affinity of phenolic herbicides was decreased 10-fold in the S(2)Q(A)(-) state than the S(1)Q(A) state. However, removal of the oxygen-evolving complex (OEC) and associated extrinsic polypeptides by hydroxylamine (HA) washing abolished the slow binding kinetics as well as the destabilizing effects on the charge-separated state. The S(2)-multiline electron paramagnetic resonance (EPR) signal and the 'split' EPR signal, originating from the S(2)Y(Z) state showed no significant changes upon binding of phenolic inhibitors at the Q(B) site. We thus propose a working model where Q(A) redox potential is lowered by short-range conformational changes induced by phenolic inhibitor binding at the Q(B) niche. Long-range effects of HA-washing eliminate this interaction, possibly by allowing more flexibility in the Q(B) site.  相似文献   

10.
Black-grass (Alopecurus myosuroides) is a major weed of wheat in Europe, with several populations having acquired resistance to multiple herbicides of differing modes of action. As compared with herbicide-susceptible black-grass, populations showing herbicide cross-resistance contained greatly elevated levels of a specific type I glutathione transferase (GST), termed AmGST2, but similar levels of a type III GST termed AmGST1. Following cloning and expression of the respective cDNAs, AmGST2 differed from AmGST1 in showing limited activity in detoxifying herbicides but high activities as a glutathione peroxidase (GPOX) capable of reducing organic hydroperoxides. In contrast to AmGST2, other GPOXs were not enhanced in the herbicide-resistant populations. Treatment with a range of herbicides used to control grass weeds in wheat resulted in increased levels of hydroperoxides in herbicide-susceptible populations but not in herbicide-resistant plants, consistent with AmGST2 functioning to prevent oxidative injury caused as a primary or secondary effect of herbicide action. Increased AmGST2 expression in black-grass was associated with partial tolerance to the peroxidizing herbicide paraquat. The selective enhancement of AmGST2 expression resulted from a constitutively high expression of the respective gene, which was activated in herbicide-susceptible black-grass in response to herbicide safeners, dehydration and chemical treatments imposing oxidative stress. Our results provide strong evidence that GSTs can contribute to resistance to multiple herbicides by playing a role in oxidative stress tolerance in addition to detoxifying herbicides by catalysing their conjugation with glutathione.  相似文献   

11.
Delftia acidovorans MC1 was continuously cultivated under nutristat conditions with elevated concentrations of the herbicides (RS)-2-(2,4-dichlorophenoxy)propionate [(RS)-2,4-DP] and 2,4-dichlorophenoxyacetate (2,4-D). The presence of 1-5 mM of either of these compounds did not essentially inhibit growth. Moreover, substrate consumption was not essentially affected at pH values of 7.0-9.0 selected by reason of alkaline in situ conditions found e.g. on contaminated building rubble but was decreased at pH 9.3. The adenylate energy charge declined to some degree as the herbicide concentration rose, the extent of this increasing as the pH rose. This was caused by an increase in the concentration of ADP and in particular AMP, in contrast to the fairly constant ATP level of around 4 nmol/mg dry mass with (RS)-2,4-DP and 2 nmol/mg with 2,4-D. Comparison of the individual growth parameters with theoretical data taking into account maintenance coefficients of 0.48 mmol (RS)-2,4-DP/g*h and 0.6 mmol 2,4-D/g*h revealed that the culture followed purely kinetic rules. This excludes the necessity of using substrate to a significant extent to satisfy extra efforts in energy for homeostasic work under these accentuated conditions.  相似文献   

12.
R Busi  M M Vila-Aiub  S B Powles 《Heredity》2011,106(5):817-824
The dynamics of herbicide resistance evolution in plants are influenced by many factors, especially the biochemical and genetic basis of resistance. Herbicide resistance can be endowed by enhanced rates of herbicide metabolism because of the activity of cytochrome P450 enzymes, although in weedy plants the genetic control of cytochrome P450-endowed herbicide resistance is poorly understood. In this study we have examined the genetic control of P450 metabolism-based herbicide resistance in a well-characterized Lolium rigidum biotype. The phenotypic resistance segregation in herbicide resistant and susceptible parents, F1, F2 and backcross (BC) families was analyzed as plant survival following treatment with the chemically unrelated herbicides diclofop-methyl or chlorsulfuron. Dominance and nuclear gene inheritance was observed in F1 families when treated at the recommended field doses of both herbicides. The segregation values of P450 herbicide resistance phenotypic traits observed in F2 and BC families was consistent with resistance endowed by two additive genes in most cases. In obligate out-crossing species such as L. rigidum, herbicide selection can easily result in accumulation of resistance genes within individuals.  相似文献   

13.
A strain of bacteria able to detoxicate the herbicide diuron in pure culture was isolated from sites contaminated with different urea herbicides. Diuron was used as a sole source of carbon and energy by this isolate which is a Gram-negative, aerobic, rod-shaped bacterium with a single polar flagellum, and grows at 40 degrees C. The strain has been identified as Pseudomonas sp.  相似文献   

14.
The herbicides trifluralin (alpha,alpha,alpha-trifluoro-2,6-dinitro-N, N-dipropyl-p-toluidine) and diallate (S-[2,3-dichloroallyl] diisopropylthiocarbamate) inhibit electron transport, ATP synthesis, and cytochrome f reduction by isolated spinach (Spinacia oleracea L.) chloroplasts. Both compounds inhibit noncyclic electron transport from H(2)O to ferricyanide more than 90% in coupled chloroplasts at concentrations less than 50 mum. Neither herbicide inhibits electron transport in assays utilizing only photosystem I activity, and the photosystem II reaction elicited by addition of oxidized p-phenylenediamine or 2,5-dimethylquinone is only partially inhibited by herbicide concentrations which block electron flow from H(2)O to ferricyanide. Inhibition of ATP synthesis parallels inhibition of electron flow in all noncyclic assay systems, and cyclic ATP synthesis catalyzed by either diaminodurene or phenazine metho-sulfate is susceptible to inhibition by both herbicides. These results indicate that trifluralin and diallate both inhibit electron flow in isolated chloroplasts at a point in the electron transport chain between the two photosystems.  相似文献   

15.
Various indicators of pesticide environmental risk have been proposed, and one of the most widely known and used is the environmental impact quotient (EIQ). The EIQ has been criticized by others in the past, but it continues to be used regularly in the weed science literature. The EIQ is typically considered an improvement over simply comparing the amount of herbicides applied by weight. Herbicides are treated differently compared to other pesticide groups when calculating the EIQ, and therefore, it is important to understand how different risk factors affect the EIQ for herbicides. The purpose of this work was to evaluate the suitability of the EIQ as an environmental indicator for herbicides. Simulation analysis was conducted to quantify relative sensitivity of the EIQ to changes in risk factors, and actual herbicide EIQ values were used to quantify the impact of herbicide application rate on the EIQ Field Use Rating. Herbicide use rate was highly correlated with the EIQ Field Use Rating (Spearman’s rho >0.96, P-value <0.001) for two herbicide datasets. Two important risk factors for herbicides, leaching and surface runoff potential, are included in the EIQ calculation but explain less than 1% of total variation in the EIQ. Plant surface half-life was the risk factor with the greatest relative influence on herbicide EIQ, explaining 26 to 28% of the total variation in EIQ for actual and simulated EIQ values, respectively. For herbicides, the plant surface half-life risk factor is assigned values without any supporting quantitative data, and can result in EIQ estimates that are contrary to quantitative risk estimates for some herbicides. In its current form, the EIQ is a poor measure of herbicide environmental impact.  相似文献   

16.
Binding of herbicides to photosystem II inhibits the electron transfer from Q(A) to Q(B) due to competition of herbicides with plastoquinone bound at the Q(B) site. We investigated herbicide binding to monomeric and dimeric photosystem II core complexes (PSIIcc) isolated from Thermosynechococcus elongatus by a combination of different methods (isothermal titration and differential scanning calorimetry, CD spectroscopy and measurements of the oxygen evolution) yielding binding constants, enthalpies and stoichiometries for various herbicides as well as information regarding stabilization/destabilization of the complex. Herbicide binding to detergent-solubilized PSIIcc can be described by a model of single independent binding sites present on this important membrane protein. Interestingly, binding stoichiometries herbicide:PSIIcc are lower than 1:1 and vary depending on the herbicide under study. Strong binding herbicides such as terbutryn stabilize PSIIcc in thermal unfolding experiments and endothermically binding herbicides like ioxynil probably cause large structural changes accompanied with the binding process as shown by differential scanning calorimetry experiments of the unfolding reaction of PSIIcc monomer in the presence of ioxynil. In addition we studied the occupancy of the Q(B) sites with plastoquinone (PQ9) by measuring flash induced fluorescence relaxation yielding a possible explanation for the deviations of herbicide binding from a 1:1 herbicide/binding site model.  相似文献   

17.
De-esterification is an important degradation or detoxification mechanism of sulfonylurea herbicide in microbes and plants. However, the biochemical and molecular mechanisms of sulfonylurea herbicide de-esterification are still unknown. In this study, a novel esterase gene, sulE, responsible for sulfonylurea herbicide de-esterification, was cloned from Hansschlegelia zhihuaiae S113. The gene contained an open reading frame of 1,194 bp, and a putative signal peptide at the N terminal was identified with a predicted cleavage site between Ala37 and Glu38, resulting in a 361-residue mature protein. SulE minus the signal peptide was synthesized in Escherichia coli BL21 and purified to homogeneity. SulE catalyzed the de-esterification of a variety of sulfonylurea herbicides that gave rise to the corresponding herbicidally inactive parent acid and exhibited the highest catalytic efficiency toward thifensulfuron-methyl. SulE was a dimer without the requirement of a cofactor. The activity of the enzyme was completely inhibited by Ag(+), Cd(2+), Zn(2+), methamidophos, and sodium dodecyl sulfate. A sulE-disrupted mutant strain, ΔsulE, was constructed by insertion mutation. ΔsulE lost the de-esterification ability and was more sensitive to the herbicides than the wild type of strain S113, suggesting that sulE played a vital role in the sulfonylurea herbicide resistance of the strain. The transfer of sulE into Saccharomyces cerevisiae BY4741 conferred on it the ability to de-esterify sulfonylurea herbicides and increased its resistance to the herbicides. This study has provided an excellent candidate for the mechanistic study of sulfonylurea herbicide metabolism and detoxification through de-esterification, construction of sulfonylurea herbicide-resistant transgenic crops, and bioremediation of sulfonylurea herbicide-contaminated environments.  相似文献   

18.
Summary

Four aquatic hyphomycetes and one terrestrial fungus were examined for their responses to the phenoxy herbicides (±)-MCPP and 2,4-D as both single and binary preparations with respect to hyphal extension, sporulation and respiration. Hyphal extension of all species was unaffected at concentrations less than 100 mg l?1. At higher concentrations there was a variable inhibitory response to the herbicides but no clear pattern was observed between the five fungi. The binary herbicide mixture had a weak synergistic effect on inhibition of growth rate. For the four aquatic hyphomycetes sporulation was reduced at several herbicide concentrations, but there was no consistent reduction over the experimental period. Flagellospora curvula and Clavariopsis aquatica showed increased sporulation at 100 and 1000 mg l?1 for only some herbicide combinations. The respiration rates of the fungi varied with species and herbicide concentration and ranged between stimulation at 100 mg l?1 to inhibition at 4000 mg l?1. The results indicate that the five fungi are not likely to be severely effected by the phenoxy herbicides at concentrations normally occurring in the field. The possible effects of these herbicides on nutrient cycling are briefly discussed.  相似文献   

19.
K. Zimmermann  M. Heck  J. Frank  J. Kern  A. Zouni 《BBA》2006,1757(2):106-114
Binding of herbicides to photosystem II inhibits the electron transfer from QA to QB due to competition of herbicides with plastoquinone bound at the QB site. We investigated herbicide binding to monomeric and dimeric photosystem II core complexes (PSIIcc) isolated from Thermosynechococcus elongatus by a combination of different methods (isothermal titration and differential scanning calorimetry, CD spectroscopy and measurements of the oxygen evolution) yielding binding constants, enthalpies and stoichiometries for various herbicides as well as information regarding stabilization/destabilization of the complex. Herbicide binding to detergent-solubilized PSIIcc can be described by a model of single independent binding sites present on this important membrane protein. Interestingly, binding stoichiometries herbicide:PSIIcc are lower than 1:1 and vary depending on the herbicide under study. Strong binding herbicides such as terbutryn stabilize PSIIcc in thermal unfolding experiments and endothermically binding herbicides like ioxynil probably cause large structural changes accompanied with the binding process as shown by differential scanning calorimetry experiments of the unfolding reaction of PSIIcc monomer in the presence of ioxynil. In addition we studied the occupancy of the QB sites with plastoquinone (PQ9) by measuring flash induced fluorescence relaxation yielding a possible explanation for the deviations of herbicide binding from a 1:1 herbicide/binding site model.  相似文献   

20.
WLR1, a biotype of Lolium rigidum Gaud. that had been treated with the sulfonylurea herbicide chlorsulfuron in 7 consecutive years, was found to be resistant to both the wheat-selective and the nonselective sulfonylurea and imidazolinone herbicides. Biotype SLR31, which became cross-resistant to chlorsulfuron following treatment with the aryloxyphenoxypropionate herbicide diclofop-methyl, was resistant to the wheat-selective, but not the nonselective, sulfonylurea and imidazolinone herbicides. The concentrations of herbicide required to reduce in vitro acetolactate synthase (ALs) activity 50% with respect to control assays minus herbicide for biotype WLR1 was greater than those for susceptible biotype VLR1 by a factor of >30, >30, 7,4, and 2 for the herbicides chlorsulfuron, sulfometuron-methyl, imazapyr, imazathapyr, and imazamethabenz, respectively. ALS activity from biotype SLR31 responded in a similar manner to that of the susceptible biotype VLR1. The resistant biotypes metabolized chlorsulfuron more rapidly than the susceptible biotype. Metabolism of 50% of [phenyl-U-14C]chlorsulfuron in the culms of two-leaf seedlings required 3.7 h in biotype SLR31, 5.1 h in biotype WLR1, and 7.1 h in biotype VLR1. In all biotypes the metabolism of chlorsulfuron in the culms was more rapid than that in the leaf lamina. Resistance to ALS inhibitors in L. rigidum may involve at least two mechanisms, increased metabolism of the herbicide and/or a herbicide-insensitive ALS.  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