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 isometric tension of single fibers isolated from glycerinated rabbit psoas muscle was measured at various temperatures using Mg-ITP as a substrate. The tension developed in Mg-ITP decreased linearly as the temperature was reduced from 24 degrees C to 4 degrees C. Myosin formed the myosin--product complex predominantly via ATP hydrolysis at the burst site during Mg-ATP hydrolysis, irrespective of temperature, and the tension developed in Mg-ATP decreased linearly as the temperature decreased (Yoshida and Tawada (1976) J. Biochem. 80, 861). During Mg-ITP hydrolysis, myosin forms the myosin*-product complex predominantly at the burst site above 20 degrees C, while myosin forms the myosin*-substrate complex below 8 degrees C (Hozumi (1976) Eur. J. Biochem. 63, 241). However, the temperature dependence of tension development in Mg-ITP is linear, as with Mg-ATP, as mentioned above. This temperature dependence is not compatible with some muscle models which assume the formation of the myosin*-product complex by cross-bridges prior to combination with actin during contraction.  相似文献   

2.
The effects of temperature on Mg-ITPase activity of heavy meromyosin and myosin subfragment 1 were measured in 0.1 M KC1. The initial burst of Pi liberation was one mol per mol of heavy meromyosin or two mol of myosin subfragment 1, i.e. one mol per two mol of myosin active sites, at 20 degrees C. However, it was almost zero mol below 8degrees C. Effects of KC1 concentration and pH on ITPase activity of heavy meromyosin at 20 degrees C were different from those below 8 degrees C, suggesting that the rate-limiting step in the Mg-ITP hydrolysis of myosin depends on temperature. The effect of temperature on the actin activation of heavy meromyosin Mg-ITPase was analyzed by measuring the temperature dependence of double-reciprocal plots of ITPase activity against actin concentration. The extent of actin activation was larger at low temperture. The results presented in this paper might be explained by assuming the existence of two kinds of active sites on a myosin molecule.  相似文献   

3.
Small-angle neutron scattering has been used to study structural features of lamellar bilayer membranes of dimyristoylphosphatidylcholine (DMPC) and DMPC mixed with various amount of cholesterol. The studies were recorded at a fixed hydration level of 17% 2H2O, i.e. just below saturation. Bragg reflections gives information on the ripple structure and on the bilayer periodicity. The crystalline Lc phase, which was stabilized after long time storage at low temperature, exhibits major small angle scattering when cholesterol is mixed into the membrane. The intermediate P beta' gel-phase, which is characteristic by the rippled structure, is dramatically stabilized by the introduction of cholesterol. The ripple structure depends significantly both on the cholesterol content and on the temperature. At high temperatures, T greater than 15 degrees C, the inverse ripple periodicity varies basically linearly with cholesterol content, and approach zero (i.e. periodicity goes to infinite) at 20 mol% cholesterol, approximately. At lower temperatures the correlation is more complex. The data indicate additional phase boundaries below 2 mol% and at approx. 8 mol%. Secondary rippled structures are observed in the low temperature L beta'-phase for cholesterol content below approx. 8 mol%. The data gives detailed insight into the phosphatidylcholine cholesterol phase diagram, which is discussed on the basis of a simple model in which the cholesterol complexes are fixed to the defect stripes of the rippled structure.  相似文献   

4.
The pH-activity curve of heavy meromyosin ATPase [EC 3.6.1.3] was measured at various temperatures. The pH-activity curve at higher temperatures showed a maximum at low pH and a minimum at pH 7 to 8 as has been already reported. At lower temperatures it was sigmoidal in shape, similar to a simple dissociation curve of pKa 6 to 7. The pH-activity curve at intermediate temperatures appeared to be inbetween the two extreme shapes. These changes in pH-activity curve with temperature were found to be common in the presence of divalent cations such as Mg2+, Mn2+, and Ca2+. The ATPase mechanism may be identical in the presence of any divalent cation, and the rate determining step revealing the steady state rate alters by changing the temperature. The transition temperatures estimated at pH 8 were 10 degrees, 8 degrees, and about 5 degrees in the presence of MnCl2, CaCl2, and MgCl2, respectively. The difference in the temperature coefficients above and below the transition temperature was most distinct in the presence of MnCl2, and vague in the presence of CaCl2. A similar change of pH-activity curve with temperature was found with heavy meromyosin ITPase in the presence of MgCl2.  相似文献   

5.
The conformations of the transitory intermediates of the myosin ATPase occurring during the hydrolytic cycle, enzyme without ligand, enzyme-substrate complex and two different forms of enzyme-product complex, have been characterized in terms of numbers and classes of reactive thiol groups based on incorporation of radioactively labeled alkylation reagent. The techniques employed allowed this to be done under steady-state conditions in the presence of high ligand concentrations on intact myosin from rabbit fast skeletal muscles at low ionic strength where the protein is in the gel state as it is in muscle. The binding of a divalent cation (Mg2+ or Ca2+) nucleotide complex exposes thiol-1 as well as thiol-2 groups. The long-lived ATPase intermediate occurring at temperatures above 10 degrees C adopts the same conformation with Mg2+ and Ca2+ ions. This intermediate does not protect the thiol-1 and thiol-2 groups but exposes a number of thiol-3 groups which seem to be located distant from the active site. The conformation of the intermediate prevailing in the presence of ATP changes with lowering temperature below 10 degrees C and is identical with that found in the presence of ADP at 0 degree C indicating a change in the rate-limiting step of the hydrolytic cycle. In the absence of divalent cations no such temperature-dependent change in conformation was observed. Evaluation of the activation entropies shows that the structure of the long-lived intermediate occurring above 10 degrees C in the presence of Mg2+ ions goes through a transformation from low to high order at around 20 degrees C. In the case of the monovalent-cation-stimulated ATPase a constant activation energy of around 70 kJ/mol, typical of many enzyme reactions, was found over the entire temperature range from 0--35 degrees C.  相似文献   

6.
Patterns of ear temperatures were examined in 2- to 45-day-old calves. At variable heat loads, peripheral temperature levels varied with the heat load with little fluctuations. When 2 to 3 days old, high ear temperatures were maintained for 10-14 hr/day at 10-15 degrees C ambient temperatures in 5 of 8 calves. In near-natural conditions, fluctuating ear temperatures were the rule in the vasodilated, intermediate and vasoconstricted states; fluctuations were symmetrical in the two ears in the vasodilated and intermediate states only. The possible occurrence of cold-induced vasodilation at 10-20 degrees C ambient temperature is discussed.  相似文献   

7.
8.
A psychrotrolerant acetate-oxidizing sulfate-reducing bacterium (strain akvb(T)) was isolated from sediment from the northern part of The North Sea with annual temperature fluctuations between 8 and 14 degrees C. Of the various substrates tested, strain akvb(T) grew exclusively by the oxidation of acetate coupled to the reduction of sulfate. The cells were motile, thick rods with round ends and grew in dense aggregates. Strain akvb(T) grew at temperatures ranging from -3.6 to 26.3 degrees C. Optimal growth was observed at 20 degrees C. The highest cell specific sulfate reduction rate of 6.2 fmol cell(-1) d(-1) determined by the (35)SO(2-)(40) method was measured at 26 degrees C. The temperature range of short-term sulfate reduction rates exceeded the temperature range of growth by 5 degrees C. The Arrhenius relationship for the temperature dependence of growth and sulfate reduction was linear, with two distinct slopes below the optimum temperatures of both processes. The critical temperature was 6.4 degrees C. The highest growth yield (4.3-4.5 g dry weight mol(-1) acetate) was determined at temperatures between 5 and 15 degrees C. The cellular fatty acid composition was determined with cultures grown at 4 and 20 degrees C, respectively. The relative proportion of cellular unsaturated fatty acids (e.g. 16:1omega7c) was higher in cells grown at 4 degrees C than in cells grown at 20 degrees C. The physiological responses to temperature changes showed that strain akvb(T) was well adapted to the temperature regime of the environment from which it was isolated. Phylogenetic analysis showed that strain akvb(T) is closest related to Desulfobacter hydrogenophilus, with a 16S rRNA gene sequence similarity of 98.6%. DNA-DNA-hybridization showed a similarity of 32% between D. hydrogenophilus and strain akvb(T). Based on phenotypic and DNA-based characteristics we propose that strain akvb(T) is a member of a new species, Desulfobacter psychrotolerans sp. nov.  相似文献   

9.
A high molecular weight 'cryogel' was obtained as insoluble complexes by cold incubation at near-freezing temperatures from heparinized plasma of patients with rheumatoid arthritis. After the cryogel was solubilized at 37 degrees C, 1:1 complex of fibrinogen and fibronectin was purified at room temperature by affinity chromatography on a gelatin-Sepharose 4B. Hydrodynamic properties of the complex were investigated as a function of temperature and NaCl concentration using a dynamic light scattering. The diffusion coefficients of the complex at 20 degrees C decreased with increasing of NaCl concentration as free fibronectin. The complex appears to be a more compact form at low ionic concentration, which is associated with conformational changes of fibronectin. The diffusion coefficient of the complex at 20 degrees C in 0.05 M TrisHCl(pII7.4) containing 0.5 M NaCl was estimated as 8.5 x 10(-8) cm2s-1. The complex did not dissociate over the temperature range from 20 to 37 degrees C. The diffusion coefficients of the complex decreased significantly at 12 degrees C and 40 degrees C. The thermal denaturation of fibrinogen molecule in the complex was observed at 40 degrees C. The CONTIN analysis of the light scattering data showed that the complex associated to form higher aggregates at 15 degrees C, but not at near-freezing temperature. The equilibrium between the complex and higher aggregates appeared reversible.  相似文献   

10.
An anaerobic ammonium oxidation (anammox) process for ammonia-rich wastewater treatment has not been reported at temperatures below 15 degrees C. This study used a gel carrier with entrapped anammox bacteria to obtain a stable nitrogen removal performance at low temperatures. In a continuous feeding test, a high nitrogen conversion rate (6.2 kg N m(-3) day(-1)) was confirmed at 32 degrees C. Nitrogen removal activity decreased gradually with decreasing operation temperature; however, it still occurred at 6 degrees C. Nitrogen conversion rates at 22 and 6.3 degrees C were 2.8 and 0.36 kg N m(-3) day(-1), respectively. Moreover, the stability of anammox activity below 20 degrees C was confirmed for more than 130 days. In batch experiments, anammox gel carriers were characterized with respect to temperature. The optimum temperature for anammox bacteria was found to be 37 degrees C. Furthermore, it was clear that the temperature dependence changed at about 28 degrees C. The apparent activation energy in the temperature range from 22 to 28 degrees C was calculated as 93 kJ mol(-1), and that in the range from 28 to 37 degrees C was 33 kJ mol(-1). This value agrees with the result of a continuous feeding test (94 kJ mol(-1), between 6 and 22 degrees C). The nitrogen removal performance demonstrated at the low temperatures used in this study will open the door for the application of anammox processes to many types of industrial wastewater treatment.  相似文献   

11.
Endocytosis of 125I-mannose-bovine serum albumin (BSA) and exocytosis of 125I-mannose-poly-D-lysine by rabbit alveolar macrophages were examined as a function of temperature. A plot for total ligand uptake (cell-associated ligand plus degraded ligand) versus time shows a single inflection point at 20 degrees C. Ligand degradation does not occur below 20 degrees C. Internalization of surface-bound 125I-mannose-BSA is negligible below 10 degrees C. The rate constant for internalization increases dramatically above 20 degrees C: 0.02 min-1 at 20 degrees C, 0.05 min-1 at 25 degrees C, 0.13 min-1 at 30 degrees C, and 0.29 min-1 at 35 degrees C. 125I-Mannose-N-acetyl-poly-D-lysine preloaded in lysosomes is exocytosed in a temperature and time-dependent fashion. Even at lower temperatures (2-10 degrees C), secretion of 125I-mannose-N-acetyl-poly-D-lysine was detected, indicating that movement of lysosomal content to plasma membrane and beyond cannot be suppressed at these temperatures. Thus, the temperature dependence of exocytosis of an 125I-labeled ligand is quite different from that of endocytosis, suggesting that the two processes are controlled by different mechanisms. Stimulation of secretion of preloaded 125I-mannose-N-acetyl-poly-D-lysine by mannose-BSA was more pronounced at lower temperatures with a sharp inflection point at 10 degrees C. These findings suggest that endosomes containing newly internalized mannose-BSA interact with the exocytosis pathway and enhance secretion of 125I-mannose-N-acetyl-poly-D-lysine from lysosomes.  相似文献   

12.
1. The reactions of cytochrome omicron in intact cells of aerobically grown Escherichia coli with O2 and CO have been studied at low temperature. 2. Flash photolysis of CO-liganded cells in the presence of O2 and at temperatures between -79 and -102 degrees C results in the oxidation of kinetically heterogeneous beta-type cytochromes (including cytochrome omicron), but not of cytochrome d. 3. The reaction of reduced cytochrome omicron with O2 involves O2 binding to give intermediate(s) with spectral characteristics similar to that of the reduced oxidase-CO complex. Observation in the alpha-region suggests that unexplained ligand dissociation accompanies the initial O2 binding. 4. At temperatures below -98 degrees C, an 'end point' in the reaction is reached; further reaction and oxidation of cytochrome omicron occurs on raising the temperature. 5. There is a linear relationship between the rate of formation of the oxygen compound and the O2 concentration up to 0.5 mM. The second-order constant for its formation (k+1) is 0.91 M-1.S-1 at -101 degrees C. The reaction is not readily reversible, the value of k-1 being 1.4 X 10(-5) S-1 and the kd 1.5 X 10(-5) M. 6. The energy of activation for this reaction at low temperatures is 29.9kJ (7.1 kcal)/mol. 7. The reaction with O2 is distinguished from that with CO by the markedly lower velocity and high photolytic reversibility of the latter. 8. Comparisons are drawn between the intermediate(s) in the O2 reaction of cytochrome omicron in E. coli and those identified in other bacteria and in the reaction of cytochrome aa3 with O2.  相似文献   

13.
The orientational order and rotational dynamics of 1-palmitoyl-2-[[2-[4-(6-phenyl-trans-1,3,5- hexatrienyl)phenyl]ethyl] carbon yl]-3-sn-phosphatidylcholine (DPH-PC) in dilinoleoylphosphatidylethanolamine (DLPE) and 1-palmitoyl-2-oleoylphosphatidylcholine (POPC) binary lipid mixtures were investigated. A previous study (Biochim. Biophys. Acta 731 (1983) 177) indicated that the empirical phase diagram of POPC/DLPE can roughly be divided into three zones. They are the lamellar (15% PC and higher), intermediate (5-15% PC) and inverted hexagonal (0-5% PC) phases. As the lipids changed from the lamellar to intermediate phase, the order parameter increased at all temperatures (1-50 degrees C). On the contrary, the rotational diffusion decreased at high temperatures (20-50 degrees C) but increased at low temperatures (1-10 degrees C). These results indicate that the intermediate phase is in a stressed state at high temperatures but in a highly mobile amorphous state at low temperatures. As the lipid progressed from the intermediate toward hexagonal phase, the order parameter decreased abruptly at all temperatures. The ratio of order parameter in the intermediate phase to that in the hexagonal phase was calculated. This ratio was found to increase linearly with temperature, indicating that a distinct change in the packing symmetry of lipids occurred as temperature increased. From the intermediate to hexagonal phase, the rotational diffusion increased slightly at high temperatures but declined abruptly at low temperatures. These results further agreed with the stressed and amorphous natures of the intermediate phases as described above.  相似文献   

14.
Tension responses to ramp stretches of 1-3% Lo (fiber length) in amplitude were examined in resting muscle fibers of the rat at temperatures ranging from 10 degrees C to 36 degrees C. Experiments were done using bundles of approximately 10 intact fibers isolated from the extensor digitorum longus (a fast muscle) and the soleus (a slow muscle). At low temperatures (below approximately 20 degrees C), the tension response consisted of an initial rise to a peak during the ramp followed by a complex tension decay to a plateau level; the tension decay occurred at approximately constant sarcomere length. The tension decay after a standard stretch at approximately 3-4.Lo/s contained a fast, an intermediate, and a (small amplitude) slow component, which at 10 degrees C (sarcomere length approximately 2.5 microns) were approximately 2000.s-1, approximately 150.s-1, and approximately 25.s-1 for fast fibers and approximately 2000.s-1, approximately 70.s-1 and approximately 8.s-1 for slow fibers, respectively. The fast component may represent the decay of interfilamentary viscous resistance, and the intermediate component may be due to viscoelasticity in the gap (titin, connectin) filament. The two- to threefold fast-slow muscle difference in the rate of passive tension relaxation (in the intermediate and the slow components) compares with previously reported differences in the speed of their active contractions; this suggests that "passive viscoelasticity" is appropriately matched to contraction speed in different muscle fiber types. At approximately 35 degrees C, the fast and intermediate components of tension relaxation were followed by a delayed tension rise at approximately 10.s-1 (fast fibers) and 2.5.s-1 (slow fibers); the delayed tension rise was accompanied by sarcomere shortening. BDM (5-10 mM) reduced the active twitch and tetanic tension responses and the delayed tension rise at 35 degrees C; the results indicate stretch sensitive activation in mammalian sarcomeres at physiological temperatures.  相似文献   

15.
H Ellens  J Bentz  F C Szoka 《Biochemistry》1986,25(14):4141-4147
The initial kinetics of fusion and leakage of liposomes composed of N-methylated dioleoylphosphatidylethanolamine (DOPE-Me) have been correlated with the phase behavior of this lipid. Gagné et al. [Gagné, J., Stamatatos, L., Diacovo, T., Hui, S. W., Yeagle, P., & Silvius, J. (1985) Biochemistry 24, 4400-4408] have shown that this lipid is lamellar (L alpha) below 20 degrees C, is hexagonal (HII) above 70 degrees C, and shows isotropic 31P NMR resonances at intermediate temperatures. This isotropic state is also characterized by complex morphological structures. We have prepared DOPE-Me liposomes at pH 9.5 and monitored the temperature dependence of the mixing of aqueous contents, leakage, and changes in light scattering upon reduction of the pH to 4.5. At and below 20 degrees C, where the lipid is in the L alpha phase, there is very little aggregation or destabilization of the liposomes. Between 30 and 60 degrees C, i.e., where the lipid is in the isotropic state, the initial rates of liposome fusion (mixing of aqueous contents) and leakage increase. At temperatures approaching that where the hexagonal HII phase transition occurs, the initial rates and extents of fusion decrease, whereas leakage is enhanced. Similar results were found for dioleoylphosphatidylethanolamine/dioleoylphosphatidylcholine (2:1) liposomes. These results clearly establish a common mechanism between the appearance of the isotropic state (between the L alpha and HII phases) and the promotion of liposome fusion. We propose a simple model to explain both the observed behavior of phosphatidylethanolamine-containing membranes with respect to liposome fusion and/or lysis and the beginning of the L alpha-HII phase transition.  相似文献   

16.
BHK 21 cells can adhere to a protein-coated plastic dish in the presence of Ca2+ at temperatures above 12 degrees C. However, they cannot adhere below 8 degrees C. The ESR spectrum of cells spin-labeled with a stearic acid label indicated that the membrane fluidity changed characteristically at 10 degrees C, 20 degrees C, and 30 degrees C. The critical temperature for cell adhesion coincided well with one of the characteristic temperatures for the membrane fluidity change. In the case of adhesion in the presence of Mg2+, no such correlation was observed.  相似文献   

17.
The alpine cockroach Celatoblatta quinquemaculata is common at altitudes of around 1500 m on the Rock and Pillar range of Central Otago, New Zealand where it experiences freezing conditions in the winter. The cockroach is freeze tolerant, but only to c. -9 degrees C. The cause of death at temperatures below this is unknown but likely to be due to osmotic damage to cells (shrinkage). This study compared the effect of different ice nucleation temperatures (-2 and -4 degrees C) on the viability of three types of cockroach tissue (midgut, Malpighian tubules and fat body cells) and cooling to three different temperatures (-5, -8, -12 degrees C). Two types of observations were made (i) cryomicroscope observations of ice formation and cell shrinkage (ii) cell integrity (viability) using vital stains. Cell viability decreased with lower treatment temperatures but ice nucleation temperature had no significant effect. Cryomicroscope observations showed that ice spread through tissue faster at -4 than -2 degrees C and that intracellular freezing only occurred when nucleated at -4 degrees C. From temperature records during cooling, it was observed that when freezing occurred, latent heat immediately increased the insect's body temperature close to its melting point (c. -0.3 degrees C). This "rebound" temperature was independent of nucleation temperature. Some tissues were more vulnerable to damage than others. As the gut is thought to be the site of freezing, it is significant that this tissue was the most robust. The ecological importance of the effect of nucleation temperature on survival of whole animals under field conditions is discussed.  相似文献   

18.
Sublingual and oesophageal temperatures were compared at various air temperatures in 16 subjects. In warm air (25-44 degrees C) sublingual temperatures stabilized within plus or minus 0-45 degrees C of oesophageal temperatures, but in air at room temperature (18-24 degrees C) they were sometimes as much as 1-1 degrees C below and in cold air (5-10 degrees C) as much as 4-4 degrees C below oesophageal readings. The sublingual-oesophageal temperature difference in cold air was greatly reduced by keeping the face warm, but it was not reduced in two patients breathing through tracheostomies and thereby eliminating cold air flow from the nose and pharynx. Parotid saliva temperature was low and saliva flow high during exposure, and cold saliva seemed to be mainly responsible for the erratic depression of sublingual temperature in the cold. These results indicate hazards in the casual use of sublingual temperatures, and indicate that external heat may have to be supplied to enable them to give reliable clinical assessments of body temperature.  相似文献   

19.
The efficiency of methylation of tRNA by a thermostable tRNA(guanosine-2')-methyltransferase (Gm-methylase) was examined at various temperatures using several species of tRNA isolated from Escherichia coli, yeast and Bacillus subtilis, each possessing different thermal properties. The optimal temperature for the methylation reaction was ca. 20 degrees C lower than the melting temperature of the tRNA in each case. Arrhenius plots of the methylation reactions with various tRNAs gave straight lines below the optimal temperatures in all cases, with similar activation energies of between 10 and 14 kcal/mol. Above the optimal temperatures, the methyl acceptor activity decreased as the incubation temperature was raised to 80 degrees C, at which point the methylase was still active. A correlation was observed between the remaining methyl acceptor activity and the hyperchromicity of tRNA. These results suggest that Gm-methylase recognizes the tertiary structure of tRNA, and it is not the substrate tRNA but the enzyme which is activated by heat.  相似文献   

20.
The acclimation of C(4) photosynthesis to low temperature was studied in the montane grass Muhlenbergia montana in order to evaluate inherent limitations in the C(4) photosynthetic pathway following chilling. Plants were grown in growth cabinets at 26 degrees C days, but at night temperatures of either 16 degrees C (the control treatment), 4 degrees C for at least 28 nights (the cold-acclimated treatment), or 1 night (the cold-stress treatment). Below a measurement temperature of 25 degrees C, little difference in the thermal response of the net CO(2) assimilation rate (A) was observed between the control and cold-acclimated treatment. By contrast, above 30 degrees C, A in the cold-acclimated treatment was 10% greater than in the control treatment. The temperature responses of Rubisco activity and net CO(2) assimilation rate were similar below 22 degrees C, indicating high metabolic control of Rubisco over the rate of photosynthesis at cool temperatures. Analysis of the response of A to intercellular CO(2) level further supported a major limiting role for Rubisco below 20 degrees C. As temperature declined, the CO(2) saturated plateau of A exhibited large reductions, while the initial slope of the CO(2) response was little affected. This type of response is consistent with a Rubisco limitation, rather than limitations in PEP carboxylase capacity. Stomatal limitations at low temperature were not apparent because photosynthesis was CO(2) saturated below 23 degrees C at air levels of CO(2). In contrast to the response of photosynthesis to temperature and CO(2) in plants acclimated for 4 weeks to low night temperature, plants exposed to 4 degrees C for one night showed substantial reduction in photosynthetic capacity at temperatures above 20 degrees C. Because these reductions were at both high and low CO(2), enzymes associated with the C(4) carbon cycle were implicated as the major mechanisms for the chilling inhibition. These results demonstrate that C(4) plants from climates with low temperature during the growing season can fully acclimate to cold stress given sufficient time. This acclimation appears to involve reversal of injury to the C(4) cycle following initial exposure to low temperature. By contrast, carbon gain at low temperatures generally appears to be constrained by the carboxylation capacity of Rubisco, regardless of acclimation time. The inability to overcome the Rubisco limitation at low temperature may be an inherent limitation restricting C(4) photosynthetic performance in cooler climates.  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