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.
Abstract— The activities of Na-K ATPase and acetylcholinesterase in the rat brain cortex were measured at different postnatal ages as a function of temperature. It was found that compared to acetylcholinesterase, Na-K ATPase is more strongly affected by the rise in temperature and that this response is further enhanced with age. Arrhenius plots of the data were prepared and the apparent energies of activation were computed for each plot. It was observed that all plots were biphasic except that for Na-K ATPase of the immature (5-day-old) brain which showed no transition temperature, with an apparent energy of activation of 15.5 kcal/mol. The enzyme from the mature brain (25-day-old) showed an average transition temperature of 22.6°C, with average apparent energies of activation of 15.3 and 27.2 kcal/mol above and below the transition temperature respectively. The cortex of 1-day-old rat showed no Na-K ATPase activity. Arrhenius plots of acetylcholinesterase studied at ages 1, 5 and 25 days postnatally all showed transition temperatures which increased from an average of 16.1°C for 1-day-old to 17 and 21.5°C for 5- and 25-day-old animals respectively. The average apparent energies of activation for acetylcholinesterase below the transition temperature changed from 8.3 kcal/mole at day 1 to 8.7 and 7.2 kcal/mol at days 5 and 25, while above the transition temperature they were 4.3, 5.2 and 4.1 at days 1, 5 and 25 respectively. The results are discussed in terms of the differences and changes in the interactions of Na-K ATPase and acetylcholinesterase with membrane lipids during the postnatal phase of brain development.  相似文献   

2.
Various parameters of protein synthesis were analyzed in rabbit reticulocytes exposed to various temperatures for up to five hours. Between 10 degrees C and 40 degrees C total protein synthesis exhibited two different apparent activation energies (36 kcal/mole, 10-24 degrees C; 22 kcal/mole, 24-40 degrees C), as did protein elongation and release (35 kcal/mole, 10-25 degrees C; 12 kcal/mole, 25-40 degrees C). However, the level of polysomes remained essentially unchanged between 0 degrees C and 42 degrees C which implies that the activation energy for polypeptide initiation is quite similar to that for elongation and is also biphasic. This situation is different from that in cultured mouse L-cells where the polysome level is dependent on temperatures. Nevertheless, reticulocytes and L-cells appear to be similar in their temperature dependence of initiation and in their rate of elongation (5-6 amino acids/second at 36 degrees C.  相似文献   

3.
Chinese hamster cells in suspension were exposed to 20 kHz ultrasound (US) at 54 W/cm2 and various temperatures between 2 and 44 °C. Activation energies were 2.6 and 24 kcal/mole below and above 35 °C, respectively. Procaine, a local anaesthetic drug known to increase membrane fluidity, enhanced cellular inactivation by US above 41 °C, increasing the activation energy to 62 kcal/mole. The inactivation of the bacterium Salmonella typhimurium by US was also dependent on the exposure temperature, with an activation energy of 2.9 kcal/mole between 2 and 44 °C. These data are most simply explained by the hypothesis that membranes are a major target for cellular inactivation by US and that the fluidity of the membranes is important in this respect.  相似文献   

4.
The energy metabolism of an anaerobic obligate thermophile, Clostridium thermocellum, has been examined as a function of incubation temperature using 31P NMR spectroscopy. Specifically investigated were the generation and availability of ATP as a function of temperature, activation energies for key processes in energy metabolism including formation of a pH gradient across the cell membrane, transport of key nutrients, and initial steps in glycolysis, and the existence of a membrane phase transition in the intact organism. Cells generate ATP via glycolysis at all temperatures examined; hence, limitation of the energy supply is not directly responsible for the lack of growth of this organism at low temperatures. Estimations of activation energies show a distinct hierarchy in the ATP-utilizing reactions examined. Conservation of ATP hydrolysis energy as delta pH has the lowest activation energy (less than or equal to 4 kcal/mol), two transport processes exhibit 10 kcal/mol activation energies, and early phosphorylation steps in glycolysis have significantly higher activation energies (approximately 25 kcal/mol). Neither the membrane-bound ATPase responsible for formation of the pH gradient nor the permease involved in phosphate transport shows evidence of a change in behavior around the phase transition temperature determined for extracted lipids of C. thermocellum. Line widths of inorganic phosphate do show a break in behavior around 35-40 degrees C. Possible explanations for this behavior are discussed.  相似文献   

5.
The temperature dependence and effects of sodium and potassium chloride on purified preparations of sarcolemmal Ca2+-activated ATPase were investigated. It was shown that within the concentration range of 0,1--1,0 M both salts have the same effect on the enzyme activity. A low ionic strength and concentration of the salts of 0,1 M the temperature maximum was 45 degrees and the shapes of temperature curves were the same. The Arrhenius plots showed a break at 16--19 degrees. The apparent activation energies were 27,3 kcal/mole below and 17,1 kcal/mole above the break point. At high ionic strength (0,5 M) the temperature maximum was observed at 40 degrees and the apparent activation energies decreased down to 18,0 kcal/mole below and 11,5 kcal/mole above the break point.  相似文献   

6.
Acidic fibroblast growth factor (aFGF) is unstable at physiological temperatures in the absence of polyanions such as heparin. Therefore, the effect of temperature on the kinetics of refolding of aFGF has been examined in the presence and absence of several polyanions. The protein folds into its native state at temperatures up to 30 degrees C without polyanions with an activation energy of approximately 14 kcal/mol, but does not acquire native structure above this temperature. When heparin, inositol hexasulfate, or sulfate ion are present, aFGF refolds below 30 degrees C with a slightly reduced activation energy (10-11 kcal/mol). In addition, the protein now also renatures between 30 and 50 degrees C with activation energies of 1-2 (heparin), 16 (inositol hexasulfate), and 7 (sulfate) kcal/mol. Trace heavy metals appear to inhibit the refolding process, but a molecular chaperone (bovine 70-kDa heat shock cognate protein) and a peptidylprolyl isomerase (the FK506-binding protein) have no effect. It is concluded that the rate of refolding of aFGF at physiological temperatures is probably controlled by the interaction of a native-like state of the protein with an unknown polyanionic species.  相似文献   

7.
Effect of reduced temperatures on protein synthesis in mouse L cells.   总被引:1,自引:0,他引:1  
N Craig 《Cell》1975,4(4):329-335
The rate of incorporation of leucine into protein, the rate of polypeptide elongation and termination, and the relative quantity and size of polysomes were analyzed in mouse L cells grown in suspension culture at various temperatures between 0 degrees C and 36 degrees C. Between 10 degrees C and 36 degrees C protein synthesis exhibited two different apparent activation energies (39 kcal/mole, 10-25 degrees C; 14 kcal/mole, 25-36 degrees C), whereas elongation and termination had only one (16 kcal/mole). Below 36 degrees C, the polysome level and size decreased, reaching a minimum of 30% of the control 36 degrees C values at 10 degrees C; below 10 degrees C the level increased again back to control values at 0 degrees C. The polysome decline was time dependent, requiring about 5 hr to reach the equilibrium value. This decline is completely reversible within 60 min, even in the presence of 4 mug/ml of actinomycin D, and even after 15 hr of incubation at the lower temperature. The results suggest that polypeptide initiation is rate limiting, particularly below 25 degrees C; whereas above this temperature, elongation or perhaps some other process may be limiting. These results are quite different from those obtained for E. coli and rabbit reticulocyte protein synthesis.  相似文献   

8.
Various parameters of protein synthesis were analyzed in rabbit reticulocytes exposed to various temperatures for up to five hours. Between 10°C and 40°C total protein synthesis exhibited two different apparent activation energies (36 kcal/mole, 10–24°C; 22 kcal/mole, 24–40°C), as did protein elongation and release (35 kcal/mole, 10–25°C; 12 kcal/mole, 25–40°C). However, the level of polysomes remained essentially unchanged between 0°C and 42°C which implies that the activation energy for polypeptide initiation is quite similar to that for elongation and is also biphasic. This situation is different from that in cultured mouse L-cells where the polysome level is dependent on temperatures. Nevertheless, reticulocytes and L-cells appear to be similar in their temperature dependence of initiation and in their rate of elongation (5–6 amino acids/second at 36°C).  相似文献   

9.
Temperature Dependence of Vasopressin Action on the Toad Bladder   总被引:6,自引:4,他引:2  
Toad bladders were challenged with vasopressin at one temperature, fixed on the mucosa with 1% glutaraldehyde, and then subjected to an osmotic gradient at another temperature. Thus, the temperature dependence of vasopressin action on membrane permeability was distinguished from the temperature dependence of osmotic water flux. As the temperature was raised from 20° to 38°C, there was a substantial increase in the velocity of vasopressin action, but osmotic flux was hardly affected. In this range of temperature the apparent energy of activation for net water movement across the bladder amounted to only 1.2 kcal/mole, a value well below the activation energy for bulk water viscosity. It is suggested that osmotic water flux takes place through narrow, nonpolar channels in the membrane. When the temperature was raised from 4° to 20°C, both vasopressin action as well as osmotic water flux were markedly enhanced. Activation energies for net water movement were now 8.5 kcal/mole (4°–9°C) and 4.1 kcal/mole (9°–20°C), indicating that the components of the aqueous channel undergo conformational changes as the temperature is lowered from 20°C. At 43°C bladder reactivity to vasopressin was lost, and irreversible changes in selective permeability were observed. The apparent energy of activation for net water movement across the denatured membrane was 6.6 kcal/mole. Approximately 1 µosmol of NaCl was exchanged for 1 µl of H2O across the denatured membrane.  相似文献   

10.
The temperature dependence of membrane interactions between PR8 influenza virus and virus receptor (GD1a)-containing liposomes was studied. For quantitation, the octadecylrhodamine B chloride (R18) membrane marker was incorporated into liposomes at quenched concentrations. Upon interaction with target membranes, the marker gets diluted, and dequenching can be measured in a fluorescence spectrophotometer. Rate constants were calculated from the dequenching curves under low pH conditions, which allow for fusion, and at neutral pH, where no specific fusion occurs. Activation energies were determined from Arrhenius plots. The results were compared with the temperature dependence of other viral activities like infectivity, hemolysis, and fusion with erythrocytes. For the slow reaction at pH 7.4, where only non-specific lipid transfer takes place, the activation energy was about 24 kcal/mole between 15 degrees C and 45 degrees C. For the fast, hemagglutinin (HA)-specific fusion reaction (pH 5.3), a very low activation energy (approximately 7 kcal/mole) was found between 25 degrees C and 37 degrees C, whereas below 25 degrees C it was much higher (approximately 34 kcal/mole). The temperature range with low activation energy coincides with the one for optimal infectivity, hemolysis, and fusion with erythrocytes. Furthermore, it is the same range in which the conformational change of HA takes place, which in the absence of a partner membrane leads to an irreversible inactivation of the fusion protein.  相似文献   

11.
An attempt has been made to assess the validity of applying the frictional and viscous coefficients of bulk water to the movement of water and solutes through the urinary bladder of the toad. The temperature dependence of diffusion of THO, C14-urea, C14-thiourea, and net water transfer across the bladder was determined in the presence and absence of vasopressin. The activation energy for diffusion of THO was 9.8 kcal per mole in the absence of vasopressin and 4.1 kcal per mole with the hormone present. Activation energies simultaneously determined following vasopressin for diffusion and net transfers of water were similar, and in the same range as known activation energies for diffusion and viscous flow in water. Urea had activation energies for diffusion of 4.1 and 3.9 kcal per mole in the absence and presence of vasopressin, respectively. Thiourea had a high activation energy for diffusion of 6.3 kcal per mole, which was unchanged, 6.6 kcal per mole, following hormone. These findings suggest that in its rate-limiting permeability barrier, water is present in a structured state, offering a high resistance to penetration by water. Vasopressin enlarges the aqueous channels so that the core of water they contain possesses the physical properties of ordinary bulk water. Urea penetrates the tissue via these aqueous channels while thiourea is limited by some other permeability barrier.  相似文献   

12.
The effect of temperature on the permeability of nonelectrolytes across liposomal membranes above and below their transition temperature has been studied by using an osmotic method. Below their transition temperature, liposomes are osmotically insensitive structures but, on addition of gramicidin A, the water permeability so increased that the permeability of solutes could be studied. The measured activation energies for permeation of a variety of nonelectrolytes has been found to increase when a) there is an increase in the capability of the solutes to form hydrogen bonds in water, b) the cholesterol concentration in the membranes increases and c) the membranes pass from a liquid-crystalline to a solid-crystalline state. The change in the activation energy for permeation per hydrogen bond is about 1.8 kcal/mole for all the different liposome systems investigated; the only solute tested that deviated from this correlation was urea, whose activation energy for permeation across a gramicidin-containing system was much lower than expected from its hydrogen-bonding capacity. This finding suggests that urea is permeating across the gramicidin pore. Although the literature contains only incomplete data relating the activation energies for permeation of nonelectrolytes across biological membranes to their hydrogen-bonding capacity, the available evidence suggests that there is a similar correlation to that found in liposomes. Thus, the average increase in the activation energy per hydrogen bond for permeation across ox red cell membranes (Jacobs, Glassman & Parpart, J. Cell. Comp. Physiol. 7:197, 1935) is 2.2 plus or minus 0.4 kcal/mole, a value that is similar to that obtained in liposomes. However, the activation energies for water and urea are - in such a system - very much lower than expected, suggesting that they, too, are permeating by some parallel route such as an aqueous pore.  相似文献   

13.
The effect of temperature on the velocity of rhodamine phalloidin-labelled F-actin moving in vitro on rabbit skeletal myosin has been studied. Translating actin filaments were visualized by epi-fluorescence in an inverted microscope, equipped with temperature control (+/- 0.2 K) of the stage and objective. Images were recorded in real time at magnifications of 400x or 160x by an intensified CCD camera on video tape. Motion of individual filaments was tracked by hand and velocities determined using frame times recorded simultaneously on the video tape. Velocity changed from 12 microns per second at 42 degrees C to 11 nm per second at 3 degrees C. The Arrhenius plot is non-linear, with the data following a cubic regression curve with no evident breaks or jumps. Data taken over the temperature range from single preparations followed the same curve for both heating and cooling; this indicates reversibility and absence of hysteresis. A hyperbolic model that smoothly translates with temperature between two asymptotic activation energies fits the data above 7 degrees C: these energies are 50(+/- 5) kJ per mole (Q10 = 1.9) at high temperatures and 289(+/- 29) kJ per mole (Q10 = 76.5) at low temperature with a transition temperature of 15.4(+/- 0.6) degrees C. These values are compared with other measurements made in vitro, in solution studies and on muscle fibres. An Arrhenius activation energy of 50 kJ per mole and a transition temperature of 15 degrees C are consistent with previous determinations but 289 kJ per mole is significantly greater than has been seen at low temperatures in other systems. This may indicate a different rate-limiting step in the kinetics of skeletal myosin driving actin filaments in vitro below 15 degrees C. Current determinations of the myosin "step-size" assume that the actin velocity is determined by the rate of ATP hydrolysis; the data confirm similar activation energies above 20 degrees C but they show that the temperature dependencies and activation energies are different at lower temperatures, implying uncoupling of the two processes.  相似文献   

14.
Two colchicine analogs with modifications only in the C ring are better inhibitors than colchicine of cell growth and tubulin polymerization. Radiolabeled thiocolchicine (with a thiomethyl instead of a methoxy group at position C-10) and N-acetylcolchinol O-methyl ether (NCME) (with a methoxy-substituted benzenoid instead of the methoxy-substituted tropone C ring) were prepared for comparison with colchicine. Scatchard analysis indicated a single binding site with KD values of 1.0-2.3 microM. Thiocolchicine was bound 2-4 times as rapidly as colchicine, but the activation energies of the reactions were nearly identical (18 kcal/mol for colchicine, 20 kcal/mol for thiocolchicine). NCME bound to tubulin in a biphasic reaction. The faster phase was 60 times as fast as colchicine binding at 37 degrees C, and a substantial reaction occurred at 0 degrees C. The rate of the faster phase of NCME binding changed relatively little as a function of temperature, so the activation energy was only 7.0 kcal/mol. Dissociation reactions were also evaluated, and at 37 degrees C the half-lives of the tubulin-drug complexes were 11 min for NCME, 24 h for thiocolchicine, and 27 h for colchicine. Relative dissociation rates as a function of temperature varied little among the drug complexes. Activation energies for the dissociation reactions were 30 kcal/mol for thiocolchicine, 27 kcal/mol for NCME, and 24 kcal/mol for colchicine. Comparison of the activation energies of association and dissociation yielded free energies for the binding reactions of -20 kcal/mol for NCME, -10 kcal/mol for thiocolchicine, and -6 kcal/mol for colchicine. The greater effectiveness of NCME and thiocolchicine as compared with colchicine in biological assays probably derives from their more rapid binding to tubulin and the lower free energies of their binding reactions.  相似文献   

15.
Data concerning the temperature dependence of ouabain-sensitive (Na+ + K+)-activated ATPase have enabled estimates of the apparent activation energies of this process to be obtained. Arrhenius plots show a point of inflection at about 20 °; at higher temperatures the activation energy is about 13.5 kcal/mole while below this temperature the value increases to 28.5 kcal/mole. Storage at −5 ° or reduction in total cation concentration without alteration of the Na+:K+ ratio causes no significant change in these values, although the specific activity is markedly reduced. Reduction in the sodium concentration alone, however, increases the apparent activation energy at lower temperatures. These results support the hypothesis that two independent processes are involved in ATP hydrolysis, one operating above the critical temperature and one operating below this temperature. Storage, or reduction in the concentrations of both sodium and potassium ions, appears to reduce the number of functional ATPase units, without significantly altering the properties of those which can still hydrolyze ATP. Reduction in the sodium concentration alone, however, may also cause some inhibition of all units. This is more marked at lower temperatures, and may arise from competition by potassium for sodium-binding sites.  相似文献   

16.
The endocytosis of diferric transferrin and accumulation of its iron by freshly isolated rabbit reticulocytes was studied using 59Fe-125I-transferrin. Internalized transferrin was distinguished from surface-bound transferrin by its resistance to release during treatment with Pronase at 4 degrees C. Endocytosis of diferric transferrin occurs at the same rate as exocytosis of apotransferrin, the rate constants being 0.08 min-1 at 22 degrees C, 0.19 min-1 at 30 degrees C, and 0.45 min-1 at 37 degrees C. At 37 degrees C, the maximum rate of transferrin endocytosis by reticulocytes is approximately 500 molecules/cell/s. The recycling time for transferrin bound to its receptor is about 3 min at this temperature. Neither transferrin nor its receptor is degraded during the intracellular passage. When a steady state has been reached between endocytosis and exocytosis of the ligand, about 90% of the total cell-bound transferrin is internal. Endocytosis of transferrin was found to be negligible below 10 degrees C. From 10 to 39 degrees C, the effect of temperature on the rate of endocytosis is biphasic, the rate increasing sharply above 26 degrees C. Over the temperature range 12-26 degrees C, the apparent activation energy for transferrin endocytosis is 33.0 +/- 2.7 kcal/mol, whereas from 26-39 degrees C the activation energy is considerably lower, at 12.3 +/- 1.6 kcal/mol. Reticulocytes accumulate iron atoms from diferric transferrin at twice the rate at which transferrin molecules are internalized, implying that iron enters the cell while still bound to transferrin. The activation energies for iron accumulation from transferrin are similar to those of endocytosis of transferrin. This study provides further evidence that transferrin-iron enters the cell by receptor-mediated endocytosis and that iron release occurs within the cell.  相似文献   

17.
The (Na+ +K+)-activated, Mg2+-dependent ATPase from rabbit kidney outer medulla was prepared in a partially inactivated, soluble form depleted of endogenous phospholipids, using deoxycholate. This preparation was reactivated 10 to 50-fold by sonicated liposomes of phosphatidylserine, but not by non-sonicated phosphatidylserine liposomes or sonicated phosphatidylcholine liposomes. The reconstituted enzyme resembled native membrane preparations of (Na+ +K+)-ATPase in its pH optimum being around 7.0, showing optimal activity at Mg2+:ATP mol ratios of approximately 1 and a Km value for ATP of 0.4 mM. Arrhenius plots of this reactivated activity at a constant pH of 7.0 and an Mg2+: ATP mol ratio of 1:1 showed a discontinuity (sharp change of slope) at 17 degrees C, with activation energy (Ea) values of 13-15 kcal/mol above this temperature and 30-35 kcal below it. A further discontinuity was also found at 8.0 degrees C and the Ea below this was very high (greater than 100 kcal/mol). Increased Mg2+ concentrations at Mg2+:ATP ratios in excess of 1:1 inhibited the (Na+ +K+)-ATPase activity and also abolished the discontinuities in the Arrhenius plots. The addition of cholesterol to phosphatidylserine at a 1:1 mol ratio partially inhibited (Na+ +K+)-ATPase reactivation. Arrhenius plots under these conditions showed a single discontinuity at 20 degrees C and Ea values of 22 and 68 kcal/mol above and below this temperature respectively. The ouabain-insensitive Mg2+-ATPase normally showed a linear Arrhenius plot with an Ea of 8 kcal/mol. The cholesterol-phosphatidylserine mixed liposomes stimulated the Mg2+-ATPase activity, which now also showed a discontinuity at 20 degrees C with, however, an increased value of 14 kcal/mol above this temperature and 6 kcal/mol below. Kinetic studies showed that cholesterol had no significant effect on the Km values for ATP. Since both cholesterol and Mg2+ are known to alter the effects of temperature on the fluidity of phospholipids, the above results are discussed in this context.  相似文献   

18.
The influence of low temperature on soybean (Glycine max [L.] Merr. cv. Wells) energy transduction via mitochondrial respiration and dehydrogenases was investigated in this study during imbibition and germination. Mitochondria were isolated from embryonic axes of seeds treated at 10 and 23 C (control) by submergence in H2O for 6 hours and maintenance for an additional 42 hours in a moist environment. Arrhenius plots of initial respiration rates revealed that those from cold-treated axes had respiratory control (RC) ratios of near 1.0 above an inflection in the plot at 8 C. Arrhenius plots of control axes mitochondrial respiration showed RC ratios of 2.8 above and 5.0 below an inflection temperature of 12.5 C. Energies of activation for mitochondrial respiration between 20 and 30 C for the cold and control treatments were 7.8 and 15.6 kcal/mole, respectively. These data indicate possible differences in mitochondrial membranes, degree of mitochondrial integrity, and mitochondrial enzyme complement between the two treatments.  相似文献   

19.
The effects of acute and long-term changes in temperature upon catalytic and calcium regulatory function of red (slow oxidative) and white (fast glycolytic) muscle from striped bass (Morone saxatilis) were determined. Acclimation to 5 degrees C or 25 degrees C had no significant effect on catalytic function (ATPase activity) or regulatory sensitivity (Ca++-activation) of myofibrils from either muscle type. Substantial differences between red and white muscle were found in the intrinsic thermal sensitivity of maximally-activated Mg++-Ca++ myofibrillar ATPase. Arrhenius plots of myofibrillar ATPase from white muscle show one significant breakpoint at 29 degrees C, with activation energies (Ea) of 2.3 and 23.4 kcal mole-1 at temperatures above and below this transition, respectively. Arrhenius plots of myofibrillar ATPase from red muscle show two transitions occurring at 22 and 9 degrees C, with Ea of 7.6 kcal mole-1 above 22 degrees C and 18.3 kcal mole-1 between 9 and 22 degrees C. Activation energies for myofibrils from red muscle increase substantially to approximately 107.3 kcal mole-1 below the 9 degrees C breakpoint. Differences in the intrinsic thermal sensitivity of red and white muscle catalytic function are apparently due to interaction of actomyosins and calcium regulatory proteins which are specific to each muscle type. The results suggest that capacity for sustained swimming in striped bass, which is powered exclusively by red muscle, will be severely impaired at cold temperature unless compensations occur above the level of contractile proteins.  相似文献   

20.
This study was an attempt to observe the effects of temperature on adsorption and one-step growth of the virus N-1 infecting the nitrogen-fixing cyanobacterium Nostoc muscorum. Adsorption rate was found to be maximum at 40° C whereas no adsorption occurred at 10° C. The Q 10 value was about 2.03 and the energy of activation, Ea was 16.3 kcal/ mole for the adsorption process. The development cycle of the virus was temperature sensitive. With increase in temperature, a gradual increase in inhibition of virus yield i.e. 8.33% at 30° C, 35.3% at 35° C and complete inhibition at 40° C was observed. Out of 7 h latent period, the early 4 h were temperature sensitive and heat treatment had a reversible inhibitory effect on virus development. The temperature treatment did not affect the rise period but burst-size was reduced.List of Abbreviattions PFU plaque-forming units - IM input multiplicity  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