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.
Porcine submaxillary mucin (PSM) is a glycoprotein composed of a protein core and frequent, short oligosaccharide side chains. We report static and dynamic light scattering experiments and intrinsic viscosities for PSM in aqueous solvent systems. In 0.1M NaCl solution, the data suggest PSM exists as large, internally branched, highly hydrated, polydisperse aggregates that slowly dissociate to give a stable species of weight-average molecular weight (Mw) 7.4 × 106. In 6M GdnHCl solution, the noncovalent bonds between PSM molecules are broken, giving a highly elongated molecule of Mw = 2.0 × 106. The irreversible nature of this dissociation suggests that the forces that stabilize the native aggregates of PSM in 0.1M NaCl are specific in nature. On reduction of PSM with mercaptoethanol, the polydispersity decreases and Mw also decreases to 9 × 105. A discrete change is observed in the solution properties of PSM in 0.1M NaCl at a concentration of 2mg/mL, manifested by a sudden decrease in the translational diffusion coefficient, an increase in viscosity number, and a decrease in slope of the osmotic compressibility. We tentatively propose that a weak and reversible secondary association process occurs at this concentration, although a purely hydrodynamic interaction cannot be ruled out.  相似文献   

2.
Conformation of mucous glycoproteins in aqueous solvents   总被引:5,自引:0,他引:5  
Light-scattering techniques have been used to measure the z-average radius of gyration Rg z-average translational diffusion coefficient Dt and weight–average molecular weight Mw of porcine submaxillary mucin (PSM) in solution. PSM isolated at low shear in the presence of protease inhibitors has a Mw about twice as large as a sample prepared without these precautions. The former sample has a Mw of 17 × 106 in 0.1M NaCl, which decreases to 8 × 106 in 6M guanidine hydrochloride (GdnHCl) and then to 2 × 106 on addition of 0.1M mercaptoethanol to the 6M GdnHCl solution. The Rg or D values obtained for PSM in this work superimpose with those of other authors for different mucin glycoproteins, leading to linear log–log relationships to the molecular weight of the protein core. Comparison of these results with those in the literature for denatured proteins suggest that mucins are linear random coils in which the protein core is stiffened by the presence of the oligosaccharide side chains. The length of the oligosaccharides and the nature of the solvent have little effect on the extension of the protein core. This suggests that the stiffness of the protein core is maintained by steric repulsion of the residues at the beginning of the oligosaccharide chains.  相似文献   

3.
Laser light-scattering has been used to investigate the size of native proteoglycan aggregates (PGA-aA1) from day-8 chick limb-bud chondrocyte cultures isolated under associative extraction and purification conditions in 0.4M guanidinium chloride (GdnHCl) solution. Dynamic light-scattering measurements yielded a hydrodynamic radius, Rs, of 244 ± 10 nm for PGA-aA1 in 0.4M GdnHCl, and a weight-average molecular weight (M w) of 150 ± 50 × 106 was obtained from a Zimm plot. Disaggregation in 4.0M GdnHCl aqueous solution yielded proteoglycan subunits (PGS) with Rs = 39 ± 2 nm, M w = 1.6 ± 0.3 × 106, which reassembled in 0.4M GdnHCl to form “reconstituted native” aggregates (PGA-raA1) with Rs = 121 ± 6 nm, M w = 17 ± 3 × 106. A second specimen of PGA-aA1 had Rs = 192 ± 10 nm, M w = 100 ± 10 × 106. The latter value was estimated from an empirical relationship between M w and Rs. After dissociation, this specimen reassembled to form PGA-raA1 with Rs = 85 ± 5 nm, M w = 12 ± 1 × 106. These data are compared with those for a specimen of reconstituted aggregate (PGA-A1) that had been extracted under dissociative conditions and then reaggregated by dialysis to 0.4M GdnHCl aqueous solution, for which Rs = 138 ± 9 nm, M w = 45 ± 8 × 106. From these values, we have calculated the weight-average number of subunits per aggregate Nw: 111 for PGA-aA1 and 12 for raA1 (70 and 7 for the second PGA-aA1 and PGA-raA1 specimen, respectively) as compared to 32 for PGA-A1. The numbers of subunits per aggregate were also determined from electron micrographs of spread specimens. The latter results show the same trends as those obtained by light scattering, but lead in each case to lower numbers of subunits per aggregate. These data demonstrate conclusively that PGA samples exhibit a higher degree of aggregation in solution than visualized in typical electron microscopy (EM) preparations, probably due to disaggregation during EM specimen preparation. Since Nw determined both by light scattering (LS) and by EM are larger for native versus reconstituted aggregate samples, our data point to a more compact aggregation of subunits along the hyaluronic acid (HA) chains in the former.  相似文献   

4.
Diffusion of bovine serum albumin in a neutral polymer solution   总被引:3,自引:0,他引:3  
G D Phillies 《Biopolymers》1985,24(2):379-386
The diffusion coefficient D of bovine serum albumin through various solutions (pH 7.0, 0.5M NaCl) of polythylene oxide (Mw ~ 1 × 105, 3 × 105) was studied with quasielastic light scattering. In solutions of the 1 × 105 polymer solution at polymer concentrations above 0.5 g/L, D is considerably greater than would have been expected from the viscosity of water:polymer mixtures, the deviations being larger at low protein concentration that at high protein concentration. With either polymer, D falls with increasing protein concentration.  相似文献   

5.
Histones are associated with DNA to form nucleosome essential for chromatin structure and major nuclear processes like gene regulation and expression. Histones consist of H1, H2A, H2B and H3, H4 type proteins. In the present study, combined histones from calf thymus were complexed with ct DNA and their binding affinities were measured fluorimetrically. All the five histones were resolved on SDS page and their binding with DNA was visualized. The values of biding affinities varied with pH and salt concentration. Highest affinity (4.0?×?105 M?1) was recorded at pH 6.5 in 50 mM phosphate buffer and 1.5?×?104 M?1 in 2 M NaCl at pH 7.0. The CD spectra support the highest binding affinity with maximum conformational changes at pH 7.0. The time-resolved fluorescence data recorded two life times for histone tyrosine residues at 300 nm emission in phosphate buffer pH 6.5. These life times did not show much change upon binding with DNA in buffer as well as in 2 M NaCl. The isothermal calorimetric studies yielded thermodynamic parameters ΔG, ΔH and ΔS as ?1.6?×?105 cal/mol, ?1.13?×?103 cal/mol and ?3.80 cal/mol/deg, respectively, evidencing a spontaneous exothermic reaction. The dominant binding forces in building the nucleosome are electrostatic interactions.  相似文献   

6.
A simple and sensitive specrophotometric method combined with solid-phase extraction (SPE) for the simultaneous determination of sodium linear-dodecylbenzenesulfonate (DBS) and sodium dodecyl sulfate (SDS) is described. The C2 (ethyl group bonded silicagel) cartridge could be repeatedly used more than 500 times for SPE, and it enabled the anionic surfactants to be concentrated by 50-fold. The calibration graph for DBS was linear in the range from 1.6×10?8 M to 5.0×10?7 M and for SDS from 2.0×10?9 M to 3.0×10?7 M. The relative standard deviation (n=5) for 5.0×10?7 M DBS was 3.1% and for 2.5×10?7 M SDS was 1.7%. The proposed method was applied to the simultaneous determination of DBS and SDS in river-water samples.  相似文献   

7.
By combining gel permeation chromatography (GPC) and light-scattering spectroscopy, including photon correlation and angular distribution of absolute scattered intensity, we were able to characterize immunologically active Haemophilus influenzae type b polysaccharide (HIB Ps) bovine serum albumin (BSA) conjugates in terms of equivalent hydrodynamic radius rh ~ (6.2 ± 0.6) × 102 Å, apparent radius of gyration rg ~ (5.4 ± 0.3) × 102 Å, apparent molecular weight Mw ~ (3.5 ± 0.4) × 106 g/mol, and a second virial coefficient A2 ~ (1.9 ± 0.3) × 10?4 cm3 mol/g2. We could study the effects of each of the processes in the conjugate formation according to the following procedure: BSA (dialysis, modification, fractionation) + HIB Ps → HIB Ps/BSA conjugate (conjugate formation, fractionation). Narrow distributions of HIB Ps BSA conjugate formation can be achieved using fractionated BSA.  相似文献   

8.
Abstract

Myeloperoxidase is very susceptible to reducing radicals because the reduction potential of the ferric/ferrous redox couple is much higher compared with other peroxidases. Semiquinone radicals are known to reduce heme proteins. Therefore, the kinetics and spectra of the reactions of p-hydroquinone, 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone with compounds I and II were investi-gated using both sequential-mixing stopped-flow techniques and conventional spectrophotometric measurements. At pH 7 and 15°C the rate constants for compound I reacting with p-hydroquinone, 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone were determined to be 5.6±0.4×107 M-1s-1, 1.3±0.1×106 M-1s-1 and 3.1±0.3×106 M-1s-1, respectively. The corresponding reaction rates for compound II reduction were calculated to be 4.5±0.3×106 M-1s-1, 1.9±0.1×105 M-1s-1 and 4.5±0.2×104 M-1s-1, respectively. Semiquinone radicals, produced by compounds I and II in the classical peroxidation cycle, promote compound III (oxymyeloperoxidase) formation. We could monitor formation of ferrous myeloperoxidase as well as its direct transition to compound III by addition of molecular oxygen. Formation of ferrous myeloperoxidase is shown to depend strongly on the reduction potential of the corresponding redox couple benzoquinone/semiquinone. With 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone as substrate, myeloperoxidase is extremely quickly trapped as compound III. These MPO-typical features could have potential in designing specific drugs which inhibit the production of hypochlorous acid and consequently attenuate inflammatory tissue damage.  相似文献   

9.
10.
A water soluble acidic heteropolysaccharide named WAF was isolated from Auricularia auricula‐judae by extracting with 0.9% NaCl solution. By using gas chromatography, gas chromatography‐mass spectrometry, and NMR, its chemical structure was determined to be composed of a backbone of α‐(1→3)‐linked D ‐mannopyranose residues with pendant side groups of β‐D ‐xylose, β‐D ‐glucose, or β‐D ‐glucuronic acid at position O6 or O2. Six fractions prepared from WAF with a weight‐average molecular mass (Mw) between 5.9 × 104 and 64.7 × 104 g/mol were characterized with laser light scattering and viscometry in 0.1M NaCl at 25°C. The dependence of intrinsic viscosity ([η]) and radius of gyration (Rg) on Mw for this polysaccharide were found to be [η] = 1.79 × 10?3Mw0.96 cm3 g?1 and Rg = 6.99 × 10?2 Mw0.54 nm. The molar mass per unit contour length (ML) and the persistence length (Lp) were estimated to be 1124 nm?1 and 11 nm, respectively. The WAF exhibited a semirigid character typical of linear polysaccharides. Molecular modeling was then used to predict the ordered and disordered states of WAF; the simulated ML and Lp were however much smaller than the experimental values. Taken altogether, the results suggested that WAF formed a duplex in solution. © 2010 Wiley Periodicals, Inc. Biopolymers 95: 217–227, 2011.  相似文献   

11.
Abstract: We identified and characterized 125I-endothelin-1 (125I-ET-1) binding sites in tumor capillaries isolated from human glioblastomas, using the quantitative receptor autoradiographic technique with pellet sections. Quantification was done using the computerized radioluminographic imaging plate system. High-affinity ET receptors were localized in capillaries from glioblastomas and the surrounding brain tissues (KD = 4.7 ± 1.0 × 10?10 and 1.6 ± 0.3 × 10?10M, respectively; Bmax = 161 ± 38 and 140 ± 37 fmol/mg, respectively; mean ± SEM, n = 5). BQ-123, a selective antagonist for the ETA receptor, potently competed for 125I-ET-1 binding to sections of the microvessels with IC50 values of 5.1 ± 0.3 and 5.1 ± 1.5 nM, and 10?6M BQ-123 displaced 84 and 58% of ET binding to capillaries from tumors and brains, respectively. In addition, competition curves obtained in the presence of increasing concentrations of ET-3 showed two components (IC50 = 5.7 ± 2.5 × 10?10 and 1.4 ± 0.2 × 10?6M for tumor microvessels, 1.8 ± 0.6 × 10?10 and 1.1 ± 0.3 × 10?6M for brain microvessels, respectively). Our results indicate that (a) the method we used is simple and highly sensitive for detecting and characterizing various receptors in tumor capillaries, especially in the case of a sparse specimen, and (b) capillaries in glioblastomas express specific high-affinity ET binding sites, candidates for biologically active ET receptors, which predominantly belong to the ETA subtype.  相似文献   

12.
The elongation growth of the Avena first internode segments was studied in the presence of one or several of the following growth substances: indoleacetic acid (IAA), 6-fur-furylamino purine (FAP, kinetin), 6-benzylamino purine (BAP), gibberellin A3 (GA3) and A4+7 (GA4+7), and abscisic acid (ABA). The cytokinins at concentrations of 10?7 to 10?6M stimulated growth with 4 to 6 per cent but this effect was not statistically significant. Concentrations higher than 5 × 10?6M inhibited growth. FAP and BAP (from 10?8M to 10?6M) had no significant interaction with any other growth substance used. The two-factor interactions of IAA × ABA, IAA × GA3, and GA3× ABA, as well as the three-factor interaction IAA × ABA × GA3 were significant. However, the IAA × ABA interaction was significant only when high concentration (10?6M) of ABA was used. The growth inhibition produced by 10?7 and 10?6M ABA was overcome by about equimolar concentrations of IAA. The stimulation of growth by GA3 and GA4+7 (10?9 to 10?7M) was prevented by simultaneous application of ABA, and it was reduced significantly by application of IAA (10?7 to 10?8M). GA3 at 10?8M combined with different concentrations of IAA gave slightly higher elongation than IAA alone but the observed values were significantly lower than expected assuming independent additive action.  相似文献   

13.
Summary Sodium chloride-tolerant plantlets of Dendrocalamus strictus were regenerated successfully from NaCl-tolerant embryogenic callus via somatic embryogenesis. The selection of embryogenic callus tolerant to 100 mM NaCl was made by exposing the callus to increasing (0–200 mM) concentrations of NaCl in Murashige and Skoog medium having 3% (w/v) sucrose, 0.8% (w/v) agar, 3.0 mg l−1 (13.6 μM) 2,4-dichlorophenoxyacetic acid (2,4-D), and 0.5mg l−1 (2.3μM) kinetin (callus initiation medium). The tolerance of the selected embryogenic callus to 100 mM NaCl was stable through three successive transfers on NaCl-free callus initiation medium. The tolerant embryogenic callus had high levels of Na+, sugar, free amino acids, and proline but a slight decline was recorded in K+ level. The stable 100 mM NaCl-tolerant embryogenic callus differentiated somatic embryos on maintenance medium [MS medium +3% sucrose +0.8% agar +2.0 mg l−1 (9.0 μM) 2,4-D+0.5 mg l−1 (2.3 μM) kinetin] supplemented with different (0–200 mM) concentrations of NaCl. About 39% of mature somatic embryos tolerant to 100 mM NaCl germinated and converted into plantlets in germination medium [half-strength MS+2% sucrose+0.02 mg l−1 (0.1 μM) α-naphthaleneacetic acid +0.1 mg l−1 (0.49 μM) indole-3-butyric acid] containing 100 mM NaCl. Of these plantlets about 31% established well on transplantation into a garden soil and sand (1:1) mixture containing 0.2% (w/w) NaCl.  相似文献   

14.
Human tracheobronchial mucin isolated from cystic fibrosis patients (CF HTBM) was purified using a combination of gel filtration and density gradient centrifugation. The resulting mucin was fractionated to reduce polydispersity and to facilitate studies of the molecular weight dependence of mucin viscoelasticity in concentrated solution. The viscoelastic properties of CF HTBM were examined in distilled water, 0.1M salt solutions and chaotropic solvents. In controlled strain experiments (strain ≥ 5%) with increasing mucin concentration, a crossover from sol to gel behavior is observed. The gel strength, as measured by the magnitude of the storage modulus at comparable mucin concentrations, is greatest for distilled water, intermediate for 0.1M NaCl, and lowest far 6M GdnHCl. In distilled water, high molecular weight mucin undergoes a sol-gel transition at ~ 12 mg/mL, and shows evidence of a plateau modulus at higher concentrations. The storage and loss moduli of concentrated high molecular weight fractions in 6M GdnHCl exhibit a power law dependence on frequency typical of weak gels near the sol–gel transition at 20 mg/mL. Similar rheology is observed in 0.1M NaCl and 0.091M NaCl/3 mM CaCl2, but with evidence for additional weak associations at low frequency. The power law exponent in these systems is 0.70 ± 0.02, in good agreement with prediction for networks formed by a percolation mechanism. Low molecular weight fractions in these solvents exhibit a fluid-like viscoelastic response. However, low molecular weight mucin in distilled water shows a strain-dependent increase in elasticity at low frequency indicative of weak intermolecular associations. Comparison of the rheological behavior of CF HTBM with our earlier studies of ovine submaxillary mucin lends support to the idea that carbohydrate side-chain interactions are important in the gelation mechanism of mucins. © 1995 John Wiley & Sons, Inc.  相似文献   

15.
The thermodynamics of ethidium ion binding to the double strands formed by the ribooligonucleotides rCA5G + rCU5G and the analogous deoxyribo-oligonucleotides dCA5G + dCT5G were determined by monitoring the absorbance versus temperature at 260 and 283 nm at several concentrations of oligonucleotides and ethidium bromide. A maximum of three ethidium ions bind to the oligonucleotides, which is consistent with intercalation and nearest-neighbor exclusion. For the ribo-oligonucleotide the binding mechanism is complex. Either two sites (assumed to be the intercalation sites at the two ends of the oligonucleotide) bind more strongly by a factor of 140 than the third site, or all sites are identical, but there is strong anticooperativity on binding (cooperativity parameter, 0.1). In sharp contrast, the binding to the same sequence (with thymine substituted for uracil) in the deoxyribo-oligonucleotide showed all sites equivalent and no cooperativity. For the ribo-oligonucleotides the enthalpy for ethidium binding is ?14 kcal/mol. The equilibrium constants at 25°C depend on the model; either K = 6 × 105M?1 for the two strong sites (4 × 103M?1 for the weak site) or K = 2.5 × 105M?1 for the intrinsic constant of the anticooperative model. For the equivalent deoxyribo-oligonucleotide the enthalpy of binding is -9 kcal/mol and the equilibrium constant at 25°C is a factor of 10 smaller (K = 2.5 × 104M?1).  相似文献   

16.
A phosphodiesterase I (EC 3.1.4.1; PDE-I) was purified from Walterinnesia aegyptia venom by preparative native polyacrylamide gel electrophoresis (PAGE). A single protein band was observed in analytical native PAGE and sodium dodecyl sulfate (SDS)-PAGE. PDE-I was a single-chain glycoprotein with an estimated molecular mass of 158 kD (SDS-PAGE). The enzyme was free of 5′-nucleotidase and alkaline phosphatase activities. The optimum pH and temperature were 9.0 and 60°C, respectively. The energy of activation (Ea) was 96.4, the Vmax and Km were 1.14 µM/min/mg and 1.9 × 10?3 M, respectively, and the Kcat and Ksp were 7 s?1 and 60 M ?1 min?1 respectively. Cysteine was a noncompetitive inhibitor, with Ki = 6.2 × 10?3 M and an IC50 of 2.6 mM, whereas adenosine diphosphate was a competitive inhibitor, with Ki = 0.8 × 10?3 M and an IC50 of 8.3 mM. Glutathione, o-phenanthroline, zinc, and ethylenediamine tetraacetic acid (EDTA) inhibited PDE-I activity whereas Mg2+ slightly potentiated the activity. PDE-I hydrolyzed thymidine-5′-monophosphate p-nitrophenyl ester most readily, whereas cyclic 3′-5′-AMP was least susceptible to hydrolysis. PDE-I was not lethal to mice at a dose of 4.0 mg/kg, ip, but had an anticoagulant effect on human plasma. These findings indicate that W. aegyptia PDE-I shares various characteristics with this enzyme from other snake venoms.  相似文献   

17.
Picrotoxin, 1 × 10?5M to 1.6 × 10?3M, had little or no effect on the amplitude of intracellularly recorded excitatory junctional potentials (EJPs) at extracellular calcium concentrations [Ca2+]0 ranging from 0.5 to 15 mM. The slope of the log EJP vs. log[Ca2+]0 relationship was approximately 1 with or without picrotoxin. The reduction of EJP amplitude resulting from the addition of 5 × 10?5M GABA was largely reversed by 10?5M picrotoxin.  相似文献   

18.
AISI-1020 carbon steel coupons were fixed onto a water circulation loop in order to study the effect of varying NaCl concentrations on formation of biofilms by natural populations of microorganisms. Overall, we observed a reduction in the number of bacteria attached to the metal surfaces as NaCl levels increased. At 12.85 and 80 g/l NaCl, the respective bacterial counts were: 1.7×109 CFU/cm2 and 7.5×102 CFU/cm2 for aerobic species; 1.3×104 CFU/cm2 and 2.1×10 CFU/cm2 for anaerobic species; and 1.8×103 CFU/cm2 and 4.6×10 CFU/cm2 for sulfate-reducing species. However, the opposite trend was observed for the numbers of iron-reducing bacteria: 4.1×106 CFU/cm2 at 12.85 g/l NaCl and 7.5 108 CFU/cm2 at 80 g/l NaCl, respectively. Fungal counts remained constant throughout the experimental period. The salt concentration at which the maximum corrosion rate was observed was 35 g/l. In view of the marked loss of metal mass recorded at this salinity, AISI-1020 carbon steel proved to belong to the group of alloys less resistant to corrosion. Journal of Industrial Microbiology & Biotechnology (2000) 25, 45–48. Received 07 December 1999/ Accepted in revised form 25 April 2000  相似文献   

19.
A study was made of the time course and kinetics of [3H]GABA uptake by dispersed cell cultures of postnatal rat cerebellum with and without neuronal cells. The properties of GABA neurons were calculated from the biochemical difference between the two types of cultures. It was found that for any given concentration of [3H]GABA, or any time up to 20 min, GABA neurons in cultures 21 days in vitro had an average velocity of uptake several orders of magnitude greater than that of nonneuronal cells. In addition, the apparent Km values for GABA neurons for high and low affinity uptake were 0.33 × 10−6 M and 41.8 × 10−4 M, respectively. For nonneuronal cells, the apparent Km for high affinity uptake was 0.29 × 10−6 M. The apparent Vmax values for GABA neurons for high and low affinity uptake were 28.7 × 10−6 mol/g DNA/min and 151.5 mmol/g DNA/min, respectively. For nonneuronal cells, the apparent Vmax for high affinity uptake was 0.06 × 10−6 mol/g DNA/min. No low affinity uptake system for nonneuronal cells could be detected after correcting the data for binding and diffusion. By substituting the apparent kinetic constants in the Michaelis-Menten equation, it was determined that for GABA concentrations of 5 × 10−9 M to 1 mM or higher over 99% of the GABA should be accumulated by GABA neurons, given equal access of all cells to the label. In addition, high affinity uptake of [3H]GABA by GABA neurons was completely blocked by treatment with 0.2 mM ouabain, whereas that by nonneuronal cells was only slightly decreased. Most (75–85%) of the [3H]GABA (4.4 × 10−6 M) uptake by both GABA neurons and nonneuronal cells was sodium and temperature dependent.  相似文献   

20.
Callus cultures of carnation, Dianthus caryophyllus L. ev. G. J. Sim, were grown on a synthetic medium of half strength Murashige and Skoog salts, 3 % sucrose, 100 mg/l of myo-inositol, 0.5 mg/l each of thiamin, HCl, pyridoxin, HCl and nicotinic acid and 10 g/l agar. Optimal concentrations of growth regulators were observed to be 3 × 10?6M indoleacetic acid (JAA) combined with 3 × 10?6M benzylaminopurin (BAP) or 10?6M 2,4-dichlorophenoxy acetic acid (2,4-D) alone. IAA + BAP caused a 100 fold increase in fresh weight over 4 weeks at 25°C. Addition of casein hydrolysate increased growth further. Cell suspension cultures worked best in media containing 2,4-D in which they had a doubling time of about 2 days. Filtered suspensions were successfully plated on agar in petri dishes, but division was never observed in single cells. The cultures initiated roots at higher concentrations of IAA or NAA, but all attempts to induce formation of shoots or em-bryoids gave negative results.  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