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.
In the photosynthetic reaction center from Rhodobacter sphaeroides, the primary (Q(A)) and secondary (Q(B)) electron acceptors are both ubiquinone-10, but with very different properties and functions. To investigate the protein environment that imparts these functional differences, we have applied X-band HYSCORE, a 2D pulsed EPR technique, to characterize the exchangeable protons around the semiquinone (SQ) in the Q(A) and Q(B) sites, using samples of (15)N-labeled reaction centers, with the native high spin Fe(2+) exchanged for diamagnetic Zn(2+), prepared in (1)H(2)O and (2)H(2)O solvent. The powder HYSCORE method is first validated against the orientation-selected Q-band ENDOR study of the Q(A) SQ by Flores et al. (Biophys. J.2007, 92, 671-682), with good agreement for two exchangeable protons with anisotropic hyperfine tensor components, T, both in the range 4.6-5.4 MHz. HYSCORE was then applied to the Q(B) SQ where we found proton lines corresponding to T ≈ 5.2, 3.7 MHz and T ≈ 1.9 MHz. Density functional-based quantum mechanics/molecular mechanics (QM/MM) calculations, employing a model of the Q(B) site, were used to assign the observed couplings to specific hydrogen bonding interactions with the Q(B) SQ. These calculations allow us to assign the T = 5.2 MHz proton to the His-L190 N(δ)H···O(4) (carbonyl) hydrogen bonding interaction. The T = 3.7 MHz spectral feature most likely results from hydrogen bonding interactions of O1 (carbonyl) with both Gly-L225 peptide NH and Ser-L223 hydroxyl OH, which possess calculated couplings very close to this value. The smaller 1.9 MHz coupling is assigned to a weakly bound peptide NH proton of Ile-L224. The calculations performed with this structural model of the Q(B) site show less asymmetric distribution of unpaired spin density over the SQ than seen for the Q(A) site, consistent with available experimental data for (13)C and (17)O carbonyl hyperfine couplings. The implications of these interactions for Q(B) function and comparisons with the Q(A) site are discussed.  相似文献   

2.
Steered molecular dynamics simulations of protein-ligand interactions   总被引:1,自引:0,他引:1  
Molecular recognition and specific protein-ligandinteractions are central to many biochemical processes,such as enzyme catalysis, assembly of organelles, en-ergy transduction, signaling, diverse control functions,and replication, expression and storage of the geneticmaterial[1]. Moreover, protein-ligand interactions pro-vide the mechanism of many drug therapies and un-derstanding of such interactions is thus significant forrational drug design[1,2]. For the experimental studiesof protein-ligan…  相似文献   

3.
Polychlorinated biphenyls (PCBs) comprise a group of persistent organic pollutants that differ significantly in their physicochemical properties, their persistence, and their biological activities. They can be metabolized via hydroxylated and dihydroxylated metabolites to PCB quinone intermediates. We have recently demonstrated that both dihydroxy PCBs and PCB quinones can form semiquinone radicals (SQ(*-)) in vitro. These semiquinone radicals are reactive intermediates that have been implicated in the toxicity of lower chlorinated PCB congeners. Here we describe the synthesis of selected PCB metabolites with differing degrees of chlorination on the oxygenated phenyl ring, e.g., 4,4'-dichloro-biphenyl-2,5-diol, 3,6,4'-trichloro-biphenyl-2,5-diol, 3,4,6,-trichloro-biphenyl-2,5-diol, and their corresponding quinones. In addition, two chlorinated o-hydroquinones were prepared, 6-chloro-biphenyl-3,4-diol and 6,4'-dichloro-biphenyl-3,4-diol. These PCB (hydro-)quinones readily react with oxygen or via comproportionation to yield the corresponding semiquinone free radicals, as detected by electron paramagnetic resonance spectroscopy (EPR alias ESR). The greater the number of chlorines on the (hydro-)quinone (oxygenated) ring, the higher the steady-state level of the resulting semiquinone radical at near neutral pH.  相似文献   

4.
The competition between the P(+)Q(A)(-) --> PQ(A) charge recombination (P, bacteriochlorophyll pair acting as primary photochemical electron donor) and the electron transfer to the secondary quinone acceptor Q(A)(-)Q(B) --> Q(A)Q(B)(-) (Q(A) and Q(B), primary and secondary electron accepting quinones) was investigated in chromatophores of Rb. capsulatus, varying the temperature down to -65 degrees C. The analysis of the flash-induced pattern for the formation of P(+)Q(A)Q(B)(-) shows that the diminished yield, when lowering the temperature, is not due to a homogeneous slowing of the rate constant k(AB) of the Q(A)(-)Q(B) --> Q(A)Q(B)(-) electron transfer but to a distribution of conformations that modulate the electron transfer rate over more than 3 orders of magnitude. This distribution appears "frozen", as no dynamic redistribution was observed over time ranges > 10 s (below -25 degrees C). The kinetic pattern was analyzed to estimate the shape of the distribution of k(AB), showing a bell-shaped band on the high rate side and a fraction of "blocked" reaction centers (RCs) with very slow k(AB). When the temperature is lowered, the high rate band moves to slower rate regions and the fraction of blocked RCs increases at the expense of the high rate band. The RCs that recombine from the P(+)Q(A)Q(B)(-) state appear temporarily converted to a state with rapid k(AB), indicating that the stabilized state described by Kleinfeld et al. (Biochemistry 1984, 23, 5780-5786) is still accessible at -60 degrees C.  相似文献   

5.
The unbinding process of E2020 ((R,S)-1-benzyl-4-[(5,6-dimethoxy-1-indanon)-2-yl]-methylpiperidine) leaving from the long active site gorge of Torpedo californica acetylcholinesterase (TcAChE) was studied by using steered molecular dynamics (SMD) simulations on a nanosecond scale with different velocities, and unbinding force profiles were obtained. Different from the unbinding of other AChE inhibitors, such as Huperzine A that undergoes the greatest barrier located at the bottleneck of the gorge, the major resistance preventing E2020 from leaving the gorge is from the peripheral anionic site where E2020 interacts intensively with several aromatic residues (e.g., Tyr70, Tyr121, and Trp279) through its benzene ring and forms a strong direct hydrogen bond and a water bridge with Ser286 via its O24. These interactions cause the largest rupture force, approximately 550 pN. It was found that the rotatable bonds of the piperidine ring to the benzene ring and dimethoxyindanone facilitate E2020 to pass the bottleneck through continuous conformation change by rotating those bonds to avoid serious conflict with Tyr121 and Phe330. The aromatic residues lining the gorge wall are the major components contributing to hydrophobic interactions between E2020 and TcAChE. Remarkably, these aromatic residues, acting in three groups as "sender" and "receiver", compose a "conveyer belt" for E2020 entering and leaving the TcAChE gorge.  相似文献   

6.
The selective (15)N isotope labeling was used for the identification of the nitrogen involved in a hydrogen bond formation with the semiquinone in the high-affinity Q(H) site in the cytochrome bo(3) ubiquinol oxidase. This nitrogen produces dominating contribution to X-Band (14)N ESEEM spectra. The 2D ESEEM (HYSCORE) experiments with the Q(H) site SQ in the series of selectively (15)N labeled bo(3) oxidase proteins have directly identified the N(epsilon) of R71 as an H-bond donor. In addition, selective (15)N labeling has allowed us for the first time to determine weak hyperfine couplings with the side-chain nitrogens from all residues around the SQ. Those are reflecting a distribution of the unpaired spin density over the protein in the SQ state of the quinone processing site.  相似文献   

7.
The entering and leaving processes of Huperzine A (HupA) binding with the long active-site gorge of Torpedo californica acetylcholinesterase (TcAChE) have been investigated by using steered molecular dynamics simulations. The analysis of the force required along the pathway shows that it is easier for HupA to bind to the active site of AChE than to disassociate from it, which for the first time interprets at the atomic level the previous experimental result that unbinding process of HupA is much slower than its binding process to AChE. The direct hydrogen bonds, water bridges, and hydrophobic interactions were analyzed during two steered molecular dynamics (SMD) simulations. Break of the direct hydrogen bond needs a great pulling force. The steric hindrance of bottleneck might be the most important factor to produce the maximal rupture force for HupA to leave the binding site but it has a little effect on the binding process of HupA with AChE. Residue Asp72 forms a lot of water bridges with HupA leaving and entering the AChE binding gorge, acting as a clamp to take out HupA from or put HupA into the active site. The flip of the peptide bond between Gly117 and Gly118 has been detected during both the conventional MD and SMD simulations. The simulation results indicate that this flip phenomenon could be an intrinsic property of AChE and the Gly117-Gly118 peptide bond in both HupA bound and unbound AChE structures tends to adopt the native enzyme structure. At last, in a vacuum the rupture force is increased up to 1500 pN while in water solution the greatest rupture force is about 800 pN, which means water molecules in the binding gorge act as lubricant to facilitate HupA entering or leaving the binding gorge.  相似文献   

8.
The hydride affinities of 80 various p- and o-quinones in DMSO solution were predicted by using B3LYP/6-311++G (2df,p)//B3LYP/6-31+G* and MP2/6-311++G**//B3LYP/6-31+G* methods, combined with the PCM cluster continuum model for the first time. The results show that the hydride affinity scale of the 80 quinones in DMSO ranges from -47.4 kcal/mol for 9,10-anthraquinone to -124.5 kcal/mol for 3,4,5,6-tetracyano-1,2-quinone. Such a long scale of the hydride affinities (-47.4 to -124.5 kcal/mol) indicates that the 80 quinones can form a large and useful library of organic oxidants, which can provide various organic hydride acceptors that the hydride affinities are known for chemists to choose in organic syntheses. By examining the effect of substituent on the hydride affinities of quinones, it is found that the hydride affinities of quinones in DMSO are linearly dependent on the sum of the Hammett substituent parameters sigma: DeltaGH-(Q) approximately -16.0Sigmasigmai - 70.5 (kcal/mol) for p-quinones and DeltaGH-(Q) approximately -16.2Sigmasigmai - 81.5 (kcal/mol) for o-quinones only if the substituents have no large electrostatic inductive effect and large ortho-effect. Study of the effect of the aromatic properties of quinone on the hydride affinities showed that the larger the aromatic system of quinone is, the smaller the hydride affinity of the quinone is, and the decrease of the hydride affinities is linearly to take place with the increase of the number of benzene rings in the molecule of quinones, from which the hydride affinities of aromatic quinones with multiple benzene rings can be predicted. By comparing the hydride affinities of p-quinones and the corresponding o-quinones, it is found that the hydride affinities of o-quinones are generally larger than those of the corresponding p-quinones by ca. 11 kcal/mol. Analyzing the effect of solvent on the hydride affinities of quinones showed that the effects of solvent (DMSO) on the hydride affinities of quinones are mainly dependent on the electrostatic interaction of the charged hydroquinone anions (QH-) with solvent (DMSO). All the information disclosed in this work should provide some valuable clues to chemists to choose suitable quinones or hydroquinones as efficient hydride acceptors or donors in organic syntheses and to predict the thermodynamics of hydride exchange between quinones and hydroquinones in DMSO solution.  相似文献   

9.
Abstract— Flash photolysis techniques have been used to study the effect of 8α-substitution on flavin triplet state formation and decay and on the properties of neutral and anionic serniquinones. Compared with riboflavin, the N(1) and N(3) isomers of 8α-histidylriboflavin show a lower triplet yield (?10%) and a faster rate of decay (? 4-Cfold). Acetylation of the histidyl a-amino groups and of the flavin ribityl side chain results in a 2-fold increase in triplet yield and a 2-fold slower rate of decay. The yield of neutral 8α-substituted flavin semiquinones upon flash photolysis in the presence of EDTA was approximately 50% that given by riboflavin. These substituted flavin neutral semiquinones dismutated at a rate 2–3 times slower than the corresponding unsubstituted form, although the anionic semiquinones dismutated at approximately the same rate. In the presence of oxygen, the kinetics of semiquinone decay changed from second order to pseudo-first order upon raising the pH, thus showing anionic semiquinone oxidation as seen previously with unsnbstituted flavins. The pK values for the ionization of the neutral 8α-substituted Aavin semiquinones are 1–1.5 units lower than the unsubstituted form. The anionic 8α-substituted flavin semiquinones react with oxygen at a rate 2–10 times more slowly than does the riboflavin form. Such alterations in properties probably reflect the electron-withdrawing effect of the 8α-substituents on the flavin ring system.  相似文献   

10.
Chlorine adsorption on small neutral, anionic, and cationic silver clusters Ag(n) (n=2-7) has been studied using the PW91PW91 density functional method. It was found that the adsorption of chlorine on the lowest-energy bare clusters does not always produce the lowest-energy complexes. In addition, the binding of chlorine can greatly change the geometries of the silver clusters in some cases. Among various possible adsorption sites, bridge site is energetically preferred for the neutral Ag(n) while top site is energetically more preferred for the anionic Ag(n) with n< or =6. For cationic clusters, adsorptions on bridge and face sites have similar binding energies, which are much larger than those on top sites. Natural bond orbital analyses show that irrespective of charge state, electrons always transfer from silver atoms to adsorbate and silver acts like alkali metals in the interaction with chlorine atom. Significant odd-even alternation patterns in the properties of the complexes have been observed: Even-electron clusters often have higher ionization energies, lower electron affinities, and higher dissociation energies than their odd-electron neighbors. It was also found that chlorine atoms bind more strongly with odd-electron bare clusters than with even-electron bare clusters. These patterns reveal that even-electron clusters are more stable than odd-electron clusters.  相似文献   

11.
12.
用ESR研究了菲醌(PQ)、四氟对寒醌(TClQ)、2,3-二氯-5,6-二氰苯醌(DDQ)和苯醌(BQ)与三乙胺、(Et3bN)的电子转移反应过程。实验结果表明,醌上取代基吸电子能力越强,越易与Et3N反应,但所形成半醌负离子自由基的稳定性,则并未有此规律,而是由自由基终止机理所决定。由实验得到了DDQ与Et3N的表变曲线。本文讨论了DDQ与Et3N反应的机理,并得其反应的微分方程,用实验拟合曲线确定速率常数。然后,对该方程求解,与实验曲线比较初步确定了该反应的历程。  相似文献   

13.
The neutral oxime reactivator RS194B with a seven-membered ring has shown better efficacy towards the tabun-inhibited AChE than that of RS69N with a six-membered ring and RS41A with a five-membered ring. The difference in the efficacy of these reactivators has remained unexplored. We report here the origin of the difference of efficacy of these reactivators based on the conformational analysis, quantum chemical calculations and steered molecular dynamics (SMD) simulations. The conformational analysis using B3LYP/6-31G(d) level of theory revealed that RS41A and RS194B are more stable in gauche conformation due to the gauche effect (–N–C–C–N– bonds) whereas RS69N prefers anti-conformation. The SMD simulations show that RS194B retains in more stable gauche conformation inside the active gorge of AChE during different time intervals that experiences more hydrogen bonding, hydrophobic interactions with the catalytic anionic site (CAS) residues and weaker interactions with the peripheral anionic site (PAS) residues compared to RS41A and RS69N. In an effort to design an even superior reactivator, RS194B-S has been chosen with a subtle change in the geometry of RS194B by replacing the carbonyl oxygen with the sulfur atom. The newly designed reactivator RS194B-S can also be a promising candidate to reactivate tabun-inhibited AChE.  相似文献   

14.
Escherichia coli alkaline phosphatase (AP) can hydrolyze a variety of chemically diverse phosphate monoesters while making contacts solely to the transferred phosphoryl group and its incoming and outgoing atoms. Strong interactions between AP and the transferred phosphoryl group are not present in the ground state despite the apparent similarity of the phosphoryl group in the ground and transition states. Such modest ground-state affinity is required to curtail substrate saturation and product inhibition and to allow efficient catalysis. To investigate how AP achieves limited affinity for its ground state, we first compared binding affinities of several related AP ligands. This comparison revealed a paradox: AP has a much stronger affinity for inorganic phosphate (P(i)) than for related compounds that are similar to P(i) geometrically and in overall charge but lack a transferable proton. We postulated that the P(i) proton could play an important role via transfer to the nearby anion, the active site serine nucleophile (Ser102), resulting in the attenuation of electrostatic repulsion between bound P(i) and the Ser102 oxyanion and the binding of P(i) in its trianionic form adjacent to a now neutral Ser residue. To test this model, isotope-edited Fourier transform infrared (FTIR) spectroscopy was used to investigate the ionic structure of AP-bound P(i). The FTIR results indicate that the P(i) trianion is bound and, in conjunction with previous studies of pH-dependent P(i) binding and other results, suggest that P(i) dianion transfers its proton to the Ser102 anion of AP. This internal proton-transfer results in stronger P(i) binding presumably because the additional negative charge on the trianionic P(i) allows stronger electrostatic interactions within the AP active site and because the electrostatic repulsion between bound P(i) and anionic Ser102 is eliminated when the transferred P(i) proton neutralizes Ser102. Indeed, when Ser102 is neutralized the P(i) trianion binds AP with a calculated K(d) of ≤290 fM. These results suggest that electrostatic repulsion between Ser102 and negatively charged phosphate ester substrates contributes to catalysis by the preferential destabilization of the reaction's E·S ground state.  相似文献   

15.
Zhang T  Anslyn EV 《Organic letters》2006,8(8):1649-1652
[structure: see text] A cadmium-centered tris-boronic acid receptor was synthesized, and its binding properties toward various anionic sugars were determined. This receptor shows high affinity for different anionic sugars, especially gluconic acid, which has an association constant near approximately 10(7) M(-)(1) at neutral pH. Further, using an indicator displacement assay, a color change of pyrocatechol violet was observed upon addition of anionic sugars. This colorimetric test was used as a facile screening technique to qualitatively analyze guest affinities.  相似文献   

16.
 To explore the interactions between ubiquinones and oxygen in living organisms, the thermodynamics of a series of electron and hydrogen transfer reactions between semiquinone radicals, as well as their corresponding protonated forms, and oxygen, singlet or triplet, were studied using the hybrid Hartree–Fock–density functional theory method Becke's three parameter hybrid method with the Lee, Yang, and Parr correlation functional. Effects of the solvent and of the isoprenyl tail on the electron and hydrogen transfer reactions were also investigated. It is found that semiquinone radicals (semiquinone anion radicals or protonated semiquinone radicals) cannot react with triplet oxygen to form the superoxide anion radical O2 . In contrast, neutral quinones can scavenge O2 efficiently. In the gas phase, only protonated semiquinone radicals can react spontaneously with singlet oxygen to produce peroxyl radical (HO2). However, both semiquinone anion radicals and protonated semiquinone radicals can react with singlet oxygen to produce harmful oxygen radicals (O2 a l l b u l l and HO2, respectively) in aqueous and protein environments. The free-energy changes of the corresponding reactions obtained for different ubiquinone systems are very similar. It clearly shows that the isoprenyl tail does not influence the electron and hydrogen transfer reactions between semiquinone radicals and oxygen significantly. Results of electron affinities, vertical ionization potentials, and proton affinities also show that the isoprenyl tail has no substantial effect on the electronic properties of ubiquinones. Received: 3 July 2000 / Accepted: 6 September 2000 / Published online: 21 December 2000  相似文献   

17.
CO adsorption on small neutral, anionic, and cationic silver clusters Ag(n) (n = 1-7) has been studied with use of the PW91PW91 density functional theory (DFT) method. The adsorption of CO on-top site, among various possible sites, is energetically preferred irrespective of the charge state of the silver cluster. The cationic silver clusters generally have a greater tendency to adsorb CO than the anionic and neutral silver ones, except for n = 3 and 4, and the binding energies reach a local minimum at n = 5. The binding energies on the neutral clusters, instead, reach a local maximum at n = 3, which is about 0.87 eV, probably large enough to be captured in the experiments. Binding of CO to the silver clusters is generally weaker than that to the copper and gold counterparts at the same size and charge state. This is due to the weaker orbital interaction between silver and CO, which is caused by the larger atomic radius of the silver atom. In contrast, Au atoms with a larger nuclear charge but a similar atomic radius to silver owing to the lanthanide contraction are able to have a stronger interaction with CO.  相似文献   

18.
The pleckstrin homology (PH) domain of general receptor for phosphoionositides 1 (GRP1-PHD) binds specifically to phosphatidylinositol (3,4,5)-triphosphate (PIP3), and acts as a second messenger. Using an extensive array of molecular dynamics (MD) simulations employing highly mobile membrane mimetic (HMMM) model as well as complementary full membrane simulations, we capture differentiable binding and dynamics of GRP1-PHD in the presence of membranes containing PC, PS, and PIP3 lipids in varying compositions. While GRP1-PHD forms only transient interactions with pure PC membranes, incorporation of anionic lipids resulted in stable membrane-bound configurations. We report the first observation of two distinct PIP3 binding modes on GRP1-PHD, involving PIP3 interactions at a “canonical” and at an “alternate” site, suggesting the possibility of simultaneous binding of multiple anionic lipids. The full membrane simulations confirmed the stability of the membrane bound pose of GRP1-PHD as captured from our HMMM membrane binding simulations. By performing additional steered membrane unbinding simulations and calculating nonequilibrium work associated with the process, as well as metadynamics simulations, on the protein bound to full membranes, allowing for more quantitative examination of the binding strength of the GRP1-PHD to the membrane, we demonstrate that along with the bound PIP3, surrounding anionic PS lipids increase the energetic cost of unbinding of GRP1-PHD from the canonical mode, causing them to dissociate more slowly than the alternate mode. Our results demonstrate that concurrent binding of multiple anionic lipids by GRP1-PHD contributes to its membrane affinity, which in turn control its signaling activity. © 2019 Wiley Periodicals, Inc.  相似文献   

19.
The recently developed MM/GBSA_IE method is applied to computing hot and warm spots in p53/PMI-MDM2/MDMX protein–protein interaction systems. Comparison of the calculated hot (>2 kcal/mol) and warm spots (>1 kcal/mol) in P53 and PMI proteins interacting with MDM2 and MDMX shows a good quantitative agreement with the available experimental data. Further, our calculation predicted hot spots in MDM2 and MDMX proteins in their interactions with P53 and PMI and they help elucidate the interaction mechanism underlying this important PPI system. In agreement with the experimental result, the present calculation shows that PMI has more hot and warm spots and binds stronger to MDM2/MDMX. The analysis of these hot and warm spots helps elucidate the fundamental difference in binding between P53 and PMI to the MDM2/MDMX systems. Specifically, for p53/PMI-MDM2 systems, p53 and PMI use essentially the same residues (L54, I61, Y67, Q72, V93, H96, and I99) of MDM2 for binding. However, PMI enhanced interactions with residues L54, Y67, and Q72 of MDM2. For the p53/PMI-MDMX system, p53 and PMI use similar residues (M53, I60, Y66, Q71, V92, and Y99) of MDMX for binding. However, PMI exploited three extra residues (M61, K93, and L98) of MDMX for enhanced binding. In addition, PMI enhanced interaction with four residues (M53, Y66, Q71, and Y99) of MDMX. These results gave quantitative explanation on why the binding affinities of PMI-MDM2/MDMX interactions are stronger than that of p53-MDM2/MDMX although their binding modes are similar. © 2018 Wiley Periodicals, Inc.  相似文献   

20.
Compounds of the form [(pap)M(Q(2-))] (pap = phenylazopyridine; Q = 3,5-di-tert-butyl-benzoquinone, M = Pd, 1a and 1b, M = Pt, 2a and 2b; Q = 4-tert-butyl-benzoquinone, M = Pd, 3a and 3b; M = Pt, 4a and 4b) were synthesized in a one-pot reaction. The geometrical isomers, which are possible because of the built in asymmetry of these ligands, have been separated by using different temperatures and variable solubility. Structural characterization of 1b shows that the metal centers are in a square planar environment, the pap ligand is in the unreduced neutral state and the quinones are in the doubly reduced, Q(2-) catecholate form. Cyclic voltammetric measurements on the complexes display two one-electron oxidations and two one-electron reductions. EPR and vis-NIR spectra of the one-electron oxidized forms of the complexes indicate that the first oxidation takes place on the Q(2-) ligands to produce a metal bound semiquinone (Q˙(-)) radical. Reduction takes place on the pap ligand, generating metal bound pap˙(-) as seen from the (14)N (I = 1) coupling in their EPR spectrum. All the complexes in their [(pap)M(Q(2-))] neutral forms show strong absorptions in the NIR region which are largely LLCT (ligand to ligand charge transfer) in origin. These NIR bands can be tuned over a wide energy range by varying the metal center as well as the Q ligand. In addition, the intensity of NIR bands can be switched on and off by a simple electron transfer at relatively low potentials. DFT studies were used to corroborate these findings.  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