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.
The effect of Dipyridamole (10(-6)-10(-3) M) on the photomobilized electron transport in the system of quinone acceptors Q(A)-Q(B) of isolated photosynthetic reaction centers of Rhodobacter sphaeroides and on its temporary stabilization on Q(B) was studied. Depending on the type of the detergent present in the reaction center (lauryl dimethylamine oxide, Triton X-100, sodium dodecyl sulfate, and sodium cholate), dipyridamole could increase the time of the electron transfer to Q(B). The dipyridamole effect on the efficiency of the electron stabilization on Q(B) for reaction centers with different detergents was revealed in slowing down the process of dark reduction of photoactive bacteriochlorophyll from Q(B) at initial concentrations of added dipyridamole (10(-6)-10(-5) M) with following acceleration of the process at the dipyridamole concentrations of 10(-4)-10(-3) M. The pH lowering from 6.8-7.0 to 5.9-6.0 increased the dipyridamole effect. The possibility of the dipyridamole effect on the structural-dynamic state of the reaction center complex, including its hydrogen bond system, which influences the studied parameters of functional activity, is suggested.  相似文献   

2.
Dipyridamole, 2,6-bis(diethanolamino)-4,8-dipiperidinopyrimido(5, 4-d)pyrimidine, is employed in clinical practice as a vasodilator. It can also inhibit a specific membrane protein (glycoprotein P) which pumps anticancer drugs out of tumor cells. Dipyridamole (10-4 M) markedly slows down the kinetics of the electrogenic phase of the photoelectric response in Rhodobacter sphaeroides chromatophores. This phase is due to proton transfer from the external medium to the secondary quinone acceptor in the reaction center. In purple membranes of bacterium Halobacterium salinarium containing bacteriorhodopsin dipyridamole (in its charged state) significantly slowed the kinetics of proton transfer from the primary donor, Asp-96 (in membranes from bacteria of wild type), or from the external medium (in D96N mutant) to the Schiff base. It is suggested that dipyridamole can influence the structural-dynamic state of membrane proteins including modification of the structure of their hydrogen bonds involved in proton-transport processes.  相似文献   

3.
1. In order to check our hypothesis of the electrostatic nature of the acceleration of methanesulfonylation of acetylcholinesterase (acetylcholine hydrolase, EC 3.1.1.7) with cationic accelerators, equations were solved for methane-sulfonylation with two accelerators and the reaction was studied in the presence of some single accelerators, including the sodium cation, and in the presence of two acclerators simultaneously. 2. The second-order rate constants for methanesulfonylation of the complexes between the enzyme and accelerators decamethonium, tetraethylammonium and tetramethylammonium are 90, 88 and 17 1 - mol-1 - s-1, respectively, which corresponds to a maximal acceleration of 29, 28 and 5.5 times, respectively. The dissociation constants for the binding of these accelerators to the enzyme, obtained from our acceleration experiments, are 3.7 - 10(-6), 3.2 - 10(-4) and 1.4 - 10(-3) M, respectively. These values are in good agreement with the dissociation constants of these ligands as inhibitors of acetylcholinesterase. It is interesting to note that the sodium cation also accelerates the methane-sulfonylation up to around three times, the corresponding second-order rate constant and the dissociation constant being 10 1 - mol-1 - s-1 and 1.3 M, respectively. 3. All tested cations compete in the acceleration with each other; they seem to accelerate the reaction in the same way and from the same site, the catalytic anionic site. 4. These findings confirm the hypothesis of the electrostatic nature of acceleration.  相似文献   

4.
Effect of modulators on protein kinase A (PKA) activity, promastigote growth and their ability to infect peritoneal macrophages was monitored. PKA inhibitors reduced [Protein Kinase Inhibitor (PKI) - 56%; H89 - 54.5%] kemptide phosphorylation by Leishmania major promastigote lysates, while activators increased phosphorylation (8-CPT-cAMP - 88%; Sp-cAMPS-AM - 152%). Activation was specifically inhibited by PKI. Phosphodiesterase inhibitors also increased kemptide phosphorylation (dipyridamole - 171%; rolipram - 106%; and 3-isobutyl-1-methyl-xanthine - 154%). Parasite proliferation was significantly retarded (200 nM H89; 100 μM myristoylated-PKI) or completely inhibited (500 nM H89) by culturing with PKA inhibitors. Incubation with dipyridamole or Sp-cAMPS-AM also inhibited proliferation. Brief treatment (2 h) with either H89, myristoylated-PKI, dipyridamole or Sp-cAMPS-AM reduced initial macrophage infection at days 1 and 2 (>40%) and on day 3 (>78% only for 100 μM myr-PKI). Characterization of leishmanial cAMP mediated signal transduction pathways will serve as the basis for the new drug design.  相似文献   

5.
Papaverine and dipyridamole induce the interphase death of thymocytes rapidly growing four hours later and reaching its maximum by the seventh-eighth hour of the cell incubation. To induce death of thymocytes no constant presence of these preparations in the incubation medium is needed, a definite (for each of preparations) time of the contact with cells being enough. The interphase death of thymocytes induced by papaverine and dipyridamole is preceded by acceleration of the release of adenine nucleotide catabolism products from cells mainly as hypoxanthine and inosine, respectively. These both processes are induced by papaverine for a shorter period of its incubation with cells than by dipyridamole and the joined use of these substances intensifies the above processes. The analysis of the data obtained indicates that thymocytes under the effect of papaverine die rather from the exhaustion of the adenine nucleotide pool, than from a decrease in the adenylate charge of cells. Exogenous adenosine essentially removes the toxic effect of papaverine but not of dipyridamole. Addition of adenine and inosine to thymocytes does not affect their survival rate in the presence of the preparations under study.  相似文献   

6.
The scrape-loading/dye transfer technique was applied on the bovine aortic endothelial cell line GM-7373 to analyze the effects of the antithrombolytic drug dipyridamole on gap junction coupling in endothelial cells. We found that a cell treatment for 24 h with dipyridamole in therapeutically relevant concentrations (1–100 μM) increased gap junction coupling in a dose dependent manner. Similar to dipyridamole, forskolin as well as 8-Br-cAMP increased the gap junction coupling, while dibutyryl-cGMP (db-cGMP) did not affect the gap junction coupling of the GM-7373 endothelial cells. In parallel, a pharmacological inhibition of protein kinase A (PKA) with N-[2-(p-bromocinnamylamino)ethyl]-5-isoquinolinesulfonamide dihydrochloride (H-89), antagonised the action of dipyridamole on gap junction coupling. We propose that the observed dipyridamole induced increase in gap junction coupling in endothelial cells is related to a cAMP-PKA dependent phosphorylation pathway. The report shows that gap junction coupling in endothelial cells is a suitable therapeutic target for treatment of cardiovascular diseases.  相似文献   

7.
The rat aortic smooth muscle cell line A-10 was used to investigate the effect of dipyridamole on the gap junction coupling of smooth muscle cells. The scrape loading/dye transfer (SL/DT) technique revealed that dipyridamole concentrations between 5 μM and 100 μM significantly increased gap junction coupling. The adenosine receptor antagonist MRS 1754, as well as the PKA inhibitors Rp-cAMPS and H-89 were able to inhibit the dipyridamole-related increase in coupling, while forskolin and Br-cAMP also induced an enhancement of the gap junction coupling. Regarding the time-dependent behaviour of dipyridamole, a short-term effect characterised by an oscillatory reaction was observed for application times of less than 5 h, while applications times of at least 6 h resulted in a long-term effect, characterised by a constant increase of gap junction coupling to its maximum levels. This increase was not altered by prolonged presence of dipyridamole. In parallel, a short application of dipyridamole for at least 15 min was found to be sufficient to evoke the long-term effect measured 6 h after drug washout. We propose that in both the short-term and long-term effect, cAMP-related pathways are activated. The short-term phase could be related to an oscillatory cAMP effect, which might directly affect connexin trafficking, assembly and/or gap junction gating. The long-term effect is most likely related to the new expression and synthesis of connexins. With previous data from a bovine aortic endothelial cell line, the present results show that gap junction coupling of vascular cells is a target for dipyridamole.  相似文献   

8.
A mutant murine T-cell line which overproduces purines and excretes massive quantities of inosine into the culture medium has served as a cell culture model for overproduction hyperuricemia (Ullman, B., Wormsted, M. A., Cohen, M. B., and Martin, D. W., Jr. (1982) Proc. Natl. Acad. Sci. U. S. A. 79, 5127-5131). Incubation of these cells with micromolar concentrations of dipyridamole, a potent inhibitor of nucleoside transport, prevents the excretion of inosine and depresses the rate of purine biosynthesis to that of wild type cells. These concentrations of dipyridamole have no effect on cellular growth rate or on the intracellular nucleoside triphosphate or phosphoribosylpyrophosphate pools. We suggest that dipyridamole might also be useful in ameliorating purine overproduction associated with hyperuricemia and gout.  相似文献   

9.
Polymorphonuclear neutrophils (PMN) are thought to play a role in reperfusion injury and ischemia. These effects are partly mediated by toxic oxygen species (superoxide anion, hydrogen peroxide and hydroxyl radical) acting at the level of the endothelium. It was demonstrated recently that the superoxide anion reacts with nitric oxide (NO) and that interaction leads to the generation of highly toxic peroxynitrite. Several drugs were tested so far in order to affect PMN function. It was demonstrated that dipyridamole (2,6-bis-diethanolamino-4,8-dipiperidinopyrimido-(5,4-d)-pyrimidine) can influence neutrophil function by inhibiting adenosine uptake. However, this action can not fully explain all of the observed effects of dipyridamole action on PMN metabolism. The aim of our study was to evaluate the influence of dipyridamole on nitric oxide production by activated polymorphonuclear neutrophils. Incubation of PMNs with hydroxylamine (HA) and phorbol myristate acetate (PMA) generated nitrite (36.4+/-4.2 nmol/h 2x10(6) PMN), dipyridamole at 100 micromol/l, 50 micromol/l and 10 micromol/l caused a considerable drop in nitrite production (11.8+/-1.8, 19.7+/-2.7 and 27.4+/-3.2 nmol/h, respectively). Neither adenosine nor the adenosine analogue could mimic the dipyridamole effect. Moreover theophylline, an adenosine inhibitor could not reverse the dipirydamole action on PMN metabolism. We also found that dipyridamole inhibited hydrogen peroxide release from neutrophils. Catalase that scavenges hydrogen peroxide also largely abolished nitric oxide release from PMN. It is evident that dipyridamole inhibits hydroxylamine-augmented nitric oxide production by activated polymorphonuclear neutrophils through an adenosine-independent mechanism.  相似文献   

10.
The inhibition of inorganic anion transport by dipyridamole (2,6-bis(diethanolamino)-4,8-dipiperidinopyrimido[5,4-d] pyrimidine) takes place only in the presence of Cl-, other halides, nitrate or bicarbonate. At any given dipyridamole concentration, the anion flux relative to the flux in the absence of dipyridamole follows the equation: Jrel = (1 + alpha 2[Cl-])/(1 + alpha 4[Cl-]) where alpha 2 and alpha 4 are independent of [Cl-] but dependent on dipyridamole concentration. At high [Cl-] the flux approaches alpha 2/alpha 4, which decreases with increasing dipyridamole concentration. Even when both [Cl-] and dipyridamole concentration assume large values, a small residual flux remains. The equation can be deduced on the assumption that Cl- binding allosterically increases the affinity for dipyridamole binding to band 3 and that the bound dipyridamole produces a non-competitive inhibition of sulfate transport. The mass-law constants for the binding of Cl- and dipyridamole to their respective-binding sites are about 24 mM and 1.5 microM, respectively (pH 6.9, 26 degrees C). Dipyridamole binding leads to a displacement of 4,4'-dibenzoylstilbene-2,2'-disulfonate (DBDS) from the stilbenedisulfonate binding site of band 3. The effect can be predicted quantitatively on the assumption that the Cl- -promoted dipyridamole binding leads to a competitive replacement of the stilbenedisulfonates. For the calculations, the same mass-law constants for binding of Cl- and dipyridamole can be used that were derived from the kinetic studies on Cl- -promoted anion transport inhibition. The newly described Cl- binding site is highly selective with respect to Cl- and other monovalent anion species. There is little competition with SO4(2-), indicating that Cl- binding involves other than purely electrostative forces. The affinity of the binding site to Cl- does not change over the pH range 6.0-7.5. Dipyridamole binds only in its deprotonated state. Binding of the deprotonated dipyridamole is pH-independent over the same range as Cl- binding.  相似文献   

11.
Rate constants for the hydrolysis of L-proline methyl ester to form proline and methanol in D(2)O buffered at neutral pD and 25 degrees C and the deuterium enrichment of the proline product determined by electrospray ionization mass spectrometry are reported. The data give k(DO) = 5.3 +/- 0.5 M(-1) s(-1) as the second-order rate constant for carbon deprotonation of N-protonated proline methyl ester by deuterioxide ion in D(2)O at 25 degrees C and I = 1.0 (KCl). The data provide good estimates of carbon acidities of pK(a) = 21 for N-protonated proline methyl ester and pK(a) = 29 for proline zwitterion in water and of the second-order rate constant k(HO) = 4.5 x 10(-5) M(-1) s(-1) for carbon deprotonation of proline zwitterion by hydroxide ion at 25 degrees C. There is no detectable acceleration of the deprotonation of N-protonated proline methyl ester by the Br?nsted base 3-quinuclidinone in water, and it is not clear that such Br?nsted catalysis would make a significant contribution to the rate acceleration for deprotonation of bound proline at proline racemase. A comparison of the first-order rate constants k(HO)[HO(-)] = 4.5 x 10(-11) s(-1) for deprotonation of free proline zwitterion in water at pH 8 and k(cat) = 2600 s(-1) for deprotonation of proline bound to the active site of proline racemase at pH 8 shows that the enzymatic rate acceleration for proline racemase is ca. 10(13)-fold. This corresponds to a 19 kcal/mol stabilization of the transition state for deprotonation of the enzyme-bound carbon acid substrate by interaction with the protein catalyst. It is suggested that (1) much of the rate acceleration of the enzymatic over the nonenzymatic reaction in water may result from transfer of the substrate proline zwitterion from the polar solvent water to a nonpolar enzyme active site and (2) the use of thiol anions rather than oxygen anions as Br?nsted bases at this putative nonpolar enzyme active site may be favored, because of the smaller energetic price for desolvation of thiol anions than for desolvation of the more strongly solvated oxygen anions.  相似文献   

12.
The increase in temperature leads to a decrease in pKa of the group responsible for the activation of CaATP2- hydrolysis by myosin in the alkaline zone of pH. At 20-25 degrees the pKa value is about 9. The value of ionization heat (deltaHi) calculated from pKa temperature dependence is 7.6+/-+/-0.8 kcal/mol. These values are approximated to the values known for phenol hydroxyl of tyrosine. It has been demonstrated that the acceleration of CaATP2- hydrolysis at alkaline values of pH is accompanied by an increase in the Arrhenius energy of activation (Ea), determined from the temperature dependence of the maximal reaction rate (V). The increase of Ea at alkaline values of pH is apparent and is due to an increase in the concentration of a deprotonized form of the enzyme, having a higher activity. A comparison of activation parameters of the reaction at alkaline and neutral values of pH permits to conclude that the acceleration of CaATP2- hydrolysis at alkaline values of pH is due to the acceleration of the limiting step of the reaction. It has also been found that at alkaline values of pH the power of myosin binding with ADP, a competitive inhibitor and the reaction product, is decreased. It is assumed that the acceleration of ATP hydrolysis at alkaline values of pH is due to accelerated dissociation of the reaction products from the active centre of the enzyme, as a result of ionization of a functional group of myosin, probably of the tyrosine residue.  相似文献   

13.
Whole cells and cell-free preparations of the methylotrophic bacteria, Pseudomonas sp. AM 1 and Achromobacter parvulus, can oxidize formate at tis concentration in the reaction medium up to 1 M. The respiration of whole cells is registered at a concentration of formate greater than 10(-2) M, while that of cell-free extracts at a formate concentration greater than 5 X 10(-5) M. This seems to be due to the presence of a permeability barrier in cells for formate. The oxidation of reduced TMPD and exogenous cytochrome c by the membrane preparations of the two bacteria is inhibited by formate and cyanide; Ki50% = 2.5 X 10(-2) and 10(-6) M, respectively. The oxidation of NADH by the membrane preparations of the bacteria is not inhibited by 1 M formate and 5 X 10(-4) M cyanide but is inhibited by formaldehyde with Ki50% = 3 X 10(-2) M. Formaldehyde has no effect on the oxidation of reduced TMPD and cytochrome c at concentrations greater than 2 X 10(-1) M. These data indicate that respiration of the studied methylotrophic bacteria in the presence of high formate concentrations should be attributed in the presence of a branched electron transport chain in them; one branch of the chain is resistant to formate and cyanide, but is sensitive to formaldehyde.  相似文献   

14.
Human equilibrative nucleoside transporters (hENT) 1 and 2 differ in that hENT1 is inhibited by nanomolar concentrations of dipyridamole and dilazep, whereas hENT2 is 2 and 3 orders of magnitude less sensitive, respectively. When a yeast expression plasmid containing the hENT1 cDNA was randomly mutated and screened by phenotypic complementation in Saccharomyces cerevisiae to identify mutants with reduced sensitivity to dilazep, clones with a point mutation that converted Met33 to Ile (hENT1-M33I) were obtained. Characterization of the mutant protein in S. cerevisiae and Xenopus laevis oocytes revealed that the mutant had less than one-tenth the sensitivity to dilazep and dipyridamole than wild type hENT1, with no change in nitrobenzylmercaptopurine ribonucleoside (NBMPR) sensitivity or apparent uridine affinity. To determine whether the reciprocal mutation in hENT2 (Ile33 to Met) also altered sensitivity to dilazep and dipyridamole, hENT2-I33M was created by site-directed mutagenesis. Although the resulting mutant (hENT2-I33M) displayed >10-fold higher dilazep and dipyridamole sensitivity and >8-fold higher uridine affinity compared with wild type hENT2, it retained insensitivity to NBMPR. These data established that mutation of residue 33 (Met versus Ile) of hENT1 and hENT2 altered the dilazep and dipyridamole sensitivities in both proteins, suggesting that a common region of inhibitor interaction has been identified.  相似文献   

15.
In isolated rat adipocytes, basal as well as insulin-stimulated 3-O-methylglucose transport was inhibited nearly completely (maximal inhibition: 95%) by the nucleoside transport inhibitors dipyridamole (IC50 = 5 microM), nitrobenzylthioguanosine (20 microM), nitrobenzylthioinosine (35 microM) and papaverine (130 microM). Transport kinetics in the presence of 10 microM dipyridamole revealed a significant increase in the transport Km value of 3-O-methylglucose (3.45 +/- 0.6 vs 2.36 +/- 0.29 mM in the controls) as well as a decrease in the Vmax value (4.84 +/- 0.95 vs 9.03 +/- 1.19 pmol/s per microliter lipid in the controls). Half-maximally inhibiting concentrations of dipyridamole were one order of magnitude higher than those inhibiting nucleoside (thymidine) uptake (0.48 microM). The inhibitory effect of dipyridamole (5 microM) reached its maximum within 30 s. The agent failed to affect insulin's half-maximally stimulating concentration (0.075 nM) indicating that it did not interfere with the mechanism by which insulin stimulates glucose transport. Further, dipyridamole fully suppressed the glucose-inhibitable cytochalasin B binding (IC50 = 1.65 +/- 0.05 microM). The data indicate that nucleoside transport inhibitors reduce glucose transport by a direct interaction with the transporter or a closely related protein. It is suggested that glucose and nucleoside transporters share structural, and possibly functional, features.  相似文献   

16.
The effect of the formation of mixed disulphides--protein-glutathione--on the proteolysis rate was studied using soluble fractions of proteins from different rat tissues as substrates. It was shown that the binding of oxidized glutathione to proteins increases the proteolysis rate under the effect of trypsin and chymotrypsin. When the concentration of oxidized glutathione is 5.10(-4) M a 1.2-1.4-fold increase in the proteolysis rate is registered and when the concentration is 5.10(-3) M a 1.4-1.8 fold increase is observed.  相似文献   

17.
Using pulse radiolysis and laser flash photolysis, we have investigated the reactions of the deleterious species, e(-)(aq), HO&z.rad;, O(2)(*)(-) and O(2)((1)Delta(g)) with 10 water-soluble cyclopropyl-fused C(60) derivatives including a mono-adduct dendro[60]fullerene (d) and C(60) derivatives based on C(60)[C(COOH)(2)](n=2-6), some of which are known to be neuroprotective in vivo. The rate constants for reactions of e(-)(aq) and HO&z.rad; lie in the range 0.5-3.3 x 10(10) M(-1) s(-1). The d and bis-adduct monoanion radicals display sharp absorption peaks around 1000 nm (epsilon = 7 000-11 500 M(-1) cm(-1)); the anions of the tris-, tetra-, and penta-adduct derivatives have broader, weaker absorptions. The monohydroxylated radicals have their most intense absorption maxima around 390-440 nm (epsilon = 1000-3000 M(-1) cm(-1)). The anion and hydroxylated radical absorption spectra display a blue-shift as the number of addends increases. The radical anions react with oxygen (k approximately 10(7)-10(9) M(-1) s(-1)). The reaction of O(2)(*)(-) with the C(60) derivatives does not occur via an electron transfer. The rate constants for singlet oxygen reaction with the dendrofullerene and eee-derivative in D(2)O at pH 7.4 are k approximately 7 x 10(7) and approximately 2 x 10(7) M(-1) s(-1) respectively, in contrast to approximately 1.2 x 10(5) M(-1) s(-1) for the reaction with C(60) in C(6)D(6). The large acceleration of the rates for electron reduction and singlet oxygen reactions in water is due to a solvophobic process.  相似文献   

18.
The inactivation of prostaglandin E2 (PGE2) was decreased in the pulmonary circulation of isolated rat lungs, when either dipyridamole or sulfinpyrazone was infused into the pulmonary artery at the concentration of 20 μM. After pulmonary injection of 7.1 nmoles of 14C-PGE2 the amount of 15-oxo-metabolites of PGE2 in the effluent was 3.91 ± 0.19 nmoles from control lungs and 2.05 ± 0.19 nmoles (2P < 0.001) in that from 20 μM dipyridamole treated lungs. The corresponding values for control and 20 μM sulfinpyrazone lungs were 4.11 ± 0.25 and 3.03 ± 0.14 nmoles (2P < 0.01), respectively. The amounts of unmetabolized PGE2 were correspondingly increased in the effluents from dipyridamole and sulfinpyrazone (20 μM) lungs. Neither dipyridamole nor sulfinpyrazone had at concentration of 2 μM any significant effect on the amount of 15-oxo-metabolites in the effluent, although the amount of unmetabolized PGE2 was slightly increased in 2 μM sulfinpyrazone experiments.  相似文献   

19.
The effects of dipyridamole on the metabolism of adenosine added to dog whole blood were studied in vitro at 37°C. The half-lives for 8.8μM and 100μM adenosine were 3.42 and 6.89 min, respectively. Dipyridamole, in concentrations of 10?7 to 10?4M increased the half-life of 8.8μM adenosine 2 to 5-fold. The disappearance rate of adenosine in the presence of 10?4 dipyridamole was similar to the disappearance rate of adenosine in plasma. Inosine formation was enhanced by dipyridamole. Control blood samples not receiving exogenous adenosine showed an increase in endogenous adenosine and inosine during 30 min of incubation. Endogenous production of nucleosides was unaffected by dipyridamole. Analysis showed that endogenous adenosine approached a steady-state concentration of 1μM. The results indicate that the half-life for adenosine in dog whole blood is significantly greater than previously reported and that dipyridamole is effective in inhibiting adenosine disappearance in concentrations as low as 10?7M. The implications of endogenous production of adenosine are discussed.  相似文献   

20.
Kinetic studies of the reduction of Pseudomonas aeruginosa ferricytochrome c551 by Fe(EDTA)2- have been made. The reaction was found to follow a second-order rate law: k 4.2 x 10(3) M(-1) s(-1) [25 degrees, micro0.1 M, pH 7.0 (phosphate)]; deltaH+/+ 3.2 kcal/ mol; AS+/+ -30 cal/mol-deg. The electrostatics-corrected self-exchange rate constant (k11 corr) calculated for cytochrome c551 based on the Fe(EDTA)2- cross reaction is 2 M(-1) s(-1), as compared to a value of 6 M(-1) s(-1) for horse heart cytochrome c. The close correspondence of the two k11 corr values is taken as an indication that the two proteins employ very similar electron transfer mechanisms in their reactions with Fe(EDTA)(2-). It is proposed that this mechanism involves reagent contact, but little protein conformational change, at the partially exposed heme edge.  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