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.
The binding of water to the dimethylphosphate anion (DMP?) was calculated using the PCILO method. We found binding energies of 25.95 kcal·mol?1 in the O1-P-O3 plane bridging the anionic oxygen atoms and 19.3 kcal·mol?1 for the one-site association of a water molecule to an anionic oxygen atom of DMP?. In this range one water molecule added to DMP? in the O1 … O3 bridged configuration has a significantly higher binding energy to DMP? than water molecules added to other binding sites. The total binding energy of 5 water molecules to DMP? is 92 kcal·mol?1, a quantity which is about 10% less than the sum of the binding energies of the corresponding monohydrates.  相似文献   

2.
An internal NMR monitor for the study of lanthanide ion (Ln3+) binding to phospholipid bilayer membranes has been developed. The dimethylphosphate anion, DMP?, forms labile complexes with Ln3+ in aqueous solution and in solutions also containing bilayer dispersions. The hyperfine shift in the DMP? resonance induced by Pr3+ ions has been used to determine the overall thermodynamic formation constants for the Pr(DMP)2+ and Pr(DMP)2+ complexes: 81 (M?1) and 349 (M?2) at 52°C; the limiting hyperfine shift (31P) at 52°C is 91.5 ppm downfield. These parameters, applied to the observed DMP? hyperfine shift in the presence of the membrane, establish both the free Pr3+ concentration and the amount of Pr3+ bound to the phospholipid surface. Extensive data for the binding of Pr3+ to the outer surfaces of sonicated vesicles yield a limiting hyperfine shift per Pr3+ of 181.6 ppm downfield for the dipalmitoylphosphatidylcholine 31P resonance at 52°C, clearly demonstrating that the binding stoichiometry is two DPPCs per Pr3+. A Hill analysis indicates that the binding data are more anti-cooperative than a realistic Langmuir isotherm, yet more cooperative than a Stern isotherm incorporating electrostatic considerations at the Debye-Hückel level. Fittings to specific models lead to a cooperative model in which tense (T) sites, with low affinity for Pr3+, present in the absence of metal ions, quickly give way to relaxed (R) sites (two DPPCs per site), with much higher affinity for Pr3+, as the amount of Pr3+ bound to the surface increases. The intrinsic equilibrium constants for the binding of Pr3+ to DPPC vesicles are 2 M?1 and 3 000 M?1 for the T and R sites, respectively, at 52°C. The distribution coefficient between these sites ([R]/[T]) in the absence of Ln3+ is 0.14 at 52°C. We picture the binding site conversion as a head-group conformational change involving mostly the choline moiety. Sketchy results for binding on the inside vesicle surface indicate that the overall affinity for Pr3+ is significantly greater and suggest that the site stoichiometry may be different.  相似文献   

3.
Using the semiempirical potential functions, conformational energies of the model compounds DMP?, d(pCp), d(pGp), and d(pCpGpCp) are calculated, and the B → Z transition is discussed along the pseudorotational path of the sugar ring. For dimethylmonophosphate anion, DMP?, the energy contour map is presented and the stabilities of the phosphodiester conformations discussed. For the sugar ring without the base attached, the minimum energies for each sugar-puckering form are calculated along the pseudorotational path. The energy barrier of the interconversion between the C(3′)-endo form and the C(2′)-endo form is calculated to be about 2.0 kcal/mol. From the conformational energy calculations of the interconversions of mononucleoside diphosphates, d(pCp) and d(pGp), between the C(2′)-endo conformer and the C(3′)-endo conformer, the purine sugar segment is known to be more convertible than the pyrimidine sugar segment. The results also support the finding that the pseudorotational transition occurred with the O(1′)-endo form more easily than with the O(1′)-exo form. Based on the results of conformational studies of DMP?, d(pCp), and d(pGp), a topological transition of the handedness of the model compound, d(pCpGpCp), is studied. The left-handed Z-form is found to be less stable by about 8.5 kcal/mol than is the right-handed B-form. The energy barrier of the Z → B transition is calculated to be about 17.4 kcal/mol. The contributions of the electrostatic and nonbonded energies to the energy barrier are discussed in connection with the conformation changes of the model compound, d(pCpGpCp).  相似文献   

4.
S P Fodor  P A Starr  T G Spiro 《Biopolymers》1985,24(8):1493-1500
Raman spectra have been recorded for poly(dG-dT) · poly(dA-dC) and poly(dA-dT) · poly(dA-dT) in low salt and at high concentrations of CsF. Poly(dG-dT) · poly(dA-dC) shows no change in the 682-cm?1 guanine mode, demonstrating the absence of the Z-structure at high salt. The 790-cm?1 phosphodiester symmetric stretch, however, shifts up 5 cm?1 in 4.3M CsF, suggesting a slight conformational change, associated with ion binding or hydration changes. Poly(dA-dT) · poly(dA-dT) shows an additional broad band at 816 cm?1, attributed to the phosphodiester modes associated with the C3′-endo deoxyribose units in the alternating B-structure. In this case, both the 841- and the 816-cm?1 asymmetric phosphodiester stretches, associated with the C2′- and C3′-endo units, shift down on addition of CsF in a sequential manner. Correlation of this sequence with that previously observed for the two 31P-nmr resonances, establishes that the phosphodiester stretching frequencies depend on the conformation of the 5′-sugar, and not on the 3′-sugar.  相似文献   

5.
The magnetic shielding constant of the 31P nucleus of the dimethylphosphate anion is calculated by an ab initio method for different values of the torsion angles about the PO ester bond and different orientations of the methyl groups. The results obtained tend to show that both types of conformational parameters contribute to the value of σ31P. The largest shielding is obtained when the methyl groups are staggered with respect to the PO bond; the smallest, for the eclipsed arrangement. Measurements carried out on the 16 deoxyribodinucleoside monophosphates show that in the majority of cases σ31P is shifted toward lower field for the dimers having large values of 3JPH3′ and 3JPH5′(5″). The theoretical results are discussed in relation to experimental data for polynucleotides and nucleic acids.  相似文献   

6.
The conformations accessible to the internucleotide phosphodiester group in deoxydinucleoside monophosphates, deoxydinucleoside triphosphates, and deoxypolynucleotides have been explored in detail by potential energy calculations. The two most predominant conformations for the nucleotide moiety (3E and 2E) and their possible combinations (3E?3E, 3E?2E, 2E?2E, 2E?3E) have been employed, similar to our earlier studies on polyribonucleotides. The internucleotide P-O bond torsions are very sensitive to the sugar pucker (3E and 2E) and sugar type (ribose and 2′-deoxyribose) on the 3′-residue of dinucleoside phosphates. The preferred phosphodiester conformations found for the deoxydinucleoside monophosphates and triphosphates, in general, follow the same pattern as those obtained for ribose sugars when the sugar on the 3′-side of the molecule has the 3E sugar-ring conformation. The internucleotide P-O bonds show a greater degree of conformational freedom when the 3′-sugar has the 2E pucker. The double gauche g?g? conformation for the phosphodiester, which leads to the overlap of the adjacent bases, is shown to be one of the energetically most favored conformations for all the sequence of sugar puckers. It is found that the 2E?2E sequence of sugar puckers shows a greater energetic preference for the stacked helical conformation (g?g?) than the (3E?3E) and the mixed sugar-pucker combinations. This effect becomes more pronounced in going from a dinucleoside monophosphate to a dinucleoside triphosphate suggesting that the 2′-deoxy sugars favor the 2E sugar pucker in di-, oligo-, and polydeoxyribonucleotide structures. In addition to g?g?, the conformations g+g?, tg?, g?t, tg+, and g+t are also found to be possible for the phosphodiester in a polydeoxyribonucleotide and their populations depend to some extent on the sugar-pucker sequence. It is shown that the short-range intramolecular interactions involving the sugar and the phosphate groups dictate to a large extent the backbone conformations of nucleic acids and polynucleotides.  相似文献   

7.
The effect of magnesium ions on the parameters of the DNA helix-coil transition has been studied for the concentration range 10?6–10?1M at the ionic strengths of 10?3M Na+. Special attention has been given to the region of low ion concentrations and to the effect of polyvalent metallic impurities present in DNA. It has been shown that binding with Mg++ increases the DNA stability, the effect being observed mainly in the concentration range 10?6–10?4M. At[Mg++]>10?2M the thermal stability of DNA starts to decrease. The melting range extends to concentrations ~10?5M and then decreases to 7–8°C at the ion content of 10?3M. Asymmetry of the melting curves is observed at low ionic strengths ([Na+] = 10?3M) and [Mg++] ? 10?5M. The results, analyzed in terms of the statistical thermodynamic theory of double-stranded homopolymers melting in the presence of ligands, suggest that the effects observed might be due to the ion redistribution from denatured to native DNA. An experimental DNA–Mg++ phase diagram has been obtained which is in good agreement with the theory. It has been shown that thermal denaturation of the system may be an efficient method for determining the ion-binding constants for both native and denatured DNA.  相似文献   

8.
We examined three sonicated, specific-seqiemce polydeoxynucleotides in solution over a wide range of concentrations of several salts by 13P-nmr spectroscopy, and we found that the alternating copolymer poly(dAdT)·poly(dAdT) exhibits a dinucleotide repeat unit in all five salts and at all concentrations studied, as indicated by the presence of a doubled in its 31P-nmr spectra. The two components of the doublet show selective shift effects. The upfield component is assigned to dApdT in the gauche?-gauche? conformation and shifts upfield in all four monovalent salts used, relative to a single-stranded oligonucleotide control. The downfield component is assigned to dTpdA in the trans-gauche? conformation and shifts downfield with increasing CsF concentration but remains essentially constant in LiCl, NaCl, and CsCl. These changes indicate a fast noncooperative transition for poly(dAdT)·poly-(dAdT) from a presumed right-handed dinucleotide-repeat B-form to another conformation with a dinucleotide-repeat structure, via a continuum of structures that may differ in the extent of the winding of the double helix. Ethanol causes the upfield component to collapse into the other component, indicating conversion to a structure with a mononucleotide repeat unit and a trans-gauche? conformation. Up to 1M Mg2+ appears to have no significant effect on the phosphodiester conformations of poly(dAdT)·poly(dAdT). By contrast, poly-(dGdC)·poly(dGdC) gives a slow cooperative transition from what is considered to be a right-handed regular B-form to a left-handed Z-form on increasing MgCl2 and NaCl concentrations, although we observed no changes in chemical shifts below the transition points. The homopolymer poly(dA)·poly(dT) exhibits no unusual shift effects or transitions upon the addition of salts when compared to the oligonucleotide control and is considered to be a regular B-form with a gauche?-gauche? phosphodiester backbone conformation. These differences emphasize the distinct secondary structures of DNAs of different sequences and their selective responses to changes in solution conditions.  相似文献   

9.
A series of model compounds containing 3‐amino‐1H‐pyrazole‐5‐carboxylic acid residue with N‐terminal amide/urethane and C‐terminal amide/hydrazide/ester groups were investigated by using NMR, Fourier transform infrared, and single‐crystal X‐ray diffraction methods, additionally supported by theoretical calculations. The studies demonstrate that the most preferred is the extended conformation with torsion angles ? and ψ close to ±180°. The studied 1H‐pyrazole with N‐terminal amide/urethane and C‐terminal amide/hydrazide groups solely adopts this energetically favored conformation confirming rigidity of that structural motif. However, when the C‐terminal ester group is present, the second conformation with torsion angles ? and ψ close to ±180° and 0°, respectively, is accessible. The conformational equilibrium is observed in NMR and Fourier transform infrared studies in solution in polar environment as well as in the crystal structures of other related compounds. The observed conformational preferences are clearly related to the presence of intramolecular interactions formed within the studied residue. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Molecular structures for parallel DNA and RNA double helices with Hoogsteen pairing are proposed for the first time. The DNA helices have sugars in the C2′-endo region and the phosphodiester conformations are (trans, gauche?), and the RNA helices have sugars in the C3′-endo region and the phosphodiester conformations are (gauche?, gauche?). A pseudorotational symmetry relates the two parallel strands of DNA helices and a screw symmetry relates the two strands of RNA helices, which have an associated tilt of the The conformational space of parallel helices with Hoogsteen base pairing, unlike the Watson-Crick duplex, is highly restricted due to the unique positioning of the symmetry axis in the former case. The features of the parallel double helix with Hoogsteen pairing are compared with the Watson-Crick duplex and the corresponding triple helix. © 1994 John Wiley & Sons, Inc.  相似文献   

11.
Nitrosamines are carcinogenic and mutagenic only after metabolic activation via endoplasmic reticulum bound mixed function oxidase enzyme systems. Rencently a new photochemical process has been discovered by which nitrosamines are converted into unknown mutagenic compounds by irradiation with long wavelength UV light (> 335 nm) in the presence of phosphate ion at neutral pH. The mutagenic activity is detected by Ames Salmonella Typhimurium strain TA100 in the absence of rat liver microsomes. We have shown that mutagen production with nitrosomorpholine is inhibited in the presence of light by various spin trapping agents (N-t-butyl-phenylnitrone, etc.). Concurrent with this inhibition a stable free radical signal has been detected whose kinetics of formation is similar to the time course of mutagen formation during irradiation in the absence of spin trap. The free radical signal is formed only when phosphate or similar ions are present in the reaction mixture. Monomethylphosphate and dimethylphosphate can substitute for phosphate ion but with small ESR signals and mutagen formation. Trimethylphosphate gives a weak, time independent ESR signal and does not cause mutagen formation. The ESR splitting constants (aN and aH) for signals generated with each of the different phosphate species show differences which suggest that these ions may be components of some intermediate free radical species that is involved in stable mutagen formation. Arsenate ion inhibits mutagen formation in the presence of phosphate but is able in the absence of phosphate to form a ESR signal similar to that observed with phosphate ion.  相似文献   

12.
The effect of ribose(O2′)-methylation on the stability of (O2′)-methylated polyribonucleotide helices has been studied by conformational energy calculations. The preferred orientation of the methyl group is found to further stabilize the helical phosphodiester conformation (g?,g?) due to the enhanced short-range interactions arising between the methyl groups and the adjacent ribose moieties. The experimentally observed increase in melting temperature of (O2′)-alkylated polyribonucleotides is thus attributable to the enhanced stability of the helical backbone conformation.  相似文献   

13.
The Raman spectrum of polycytidylic acid was investigated in the pH range of 6.6–4.1. The thermal melting temperatures and the nature of the thermal melting profiles change in this range as monitored by the three Raman band envelopes, which include the 780-, 805-cm?1 bands, the 1190-, 1285-cm?1 bands, and the 1527-cm?1 band. By coupling these data with the theory of Raman scattering intensity and quantitative pH profiles for cytidine, it is shown that the band envelopes studied exhibit specific, yet different information regarding the thermal melting process. The band envelopes at 1170–1310 and 1527 cm?1, which are a sensitive function of both the extent of protonation and base stacking (hypochromic), reveal Tm values which agree with values derived from uv melting profiles. The 760–830-cm?1 envelope, which is not directly sensitive to cytosine residue protonation, but includes information associated with base stacking (the 780-cm?1 band) and the nature of the phosphodiester backbone (the frequency-dependent 805-cm?1 component), exhibits Tm values which deviate from the values obtained from the other bands. The observed differences are pH-dependent and correlate well with the extent of deprotonation that takes place in the denaturation process. Details of the spectrum of neutral and protonated poly(C) from pH 7 to 4.1 are discussed and related to the nature of the thermal denaturation process.  相似文献   

14.
Dirk Stigter 《Biopolymers》1998,46(7):503-516
We have studied electrostatic properties of DNA with a discrete charge model consisting of a cylindrical dielectric core with a radius of 8 Å and a dielectric constant Di = 4, surrounded by two helical strings of phosphate point charges at 10 Å from the axis, immersed in an aqueous medium with dielectric constant Dw = 78.54. Eliminating the dielectric core makes potentials in the phosphate surface less negative by about 0.5 kT/e. Salt effects are evaluated for the model without a dielectric core, using the shielded Coulomb potential. Smearing the phosphate charges increases their potential by about 2.5 kT/e, due mostly to the self-potential of the smeared charge. Potentials in the center of the minor and major grooves vary less than 0.02 kT/e along their helical path. The potential in the center of the minor groove is from 1.0 to 1.7 kT/e, more negative than in the center of the major groove, depending on dielectric core and salt concentration. So multivalent cations and also larger cationic ligands, such as some antibiotics, are likely to adsorb in the minor groove, in agreement with earlier computations by A. and B. Pullman. Dielectric effects on the surface potential and the local potential variations are found to be relatively small. Bending of DNA is studied by placing a multivalent cation, MZ+, in the center of the minor or major groove, curving DNA around it for a certain length, and calculating the free energy difference between the bent and the straight configuration. Boltzmann averaged bending angles, 〈β〉, are found to be maximal in 0.03M monovalent salt, for a length of about 50 or 25 Å of curved DNA when an MZ+ ion is adsorbed in the minor or the major groove, respectively. When the dielectric constant of water is used throughout the calculation, we find maximal bends of 〈β〉 = 11° for M2+ and 〈β〉 = 16° for M3+ in the minor groove, 〈β〉 = 13° for M3+ in the major groove. The absence of bends in DNA adsorbed to mica in the presence of Mg salts supports the role of Mg2+ in “ion bridging” between DNA and mica. The treatment of the effective dielectric constant between two points outside a dielectric cylinder in water is appended. © 1998 John Wiley & Sons, Inc. Biopoly 46: 503–516, 1998  相似文献   

15.
A new program, TALOS-N, is introduced for predicting protein backbone torsion angles from NMR chemical shifts. The program relies far more extensively on the use of trained artificial neural networks than its predecessor, TALOS+. Validation on an independent set of proteins indicates that backbone torsion angles can be predicted for a larger, ≥90 % fraction of the residues, with an error rate smaller than ca 3.5 %, using an acceptance criterion that is nearly two-fold tighter than that used previously, and a root mean square difference between predicted and crystallographically observed (?, ψ) torsion angles of ca 12º. TALOS-N also reports sidechain χ1 rotameric states for about 50 % of the residues, and a consistency with reference structures of 89 %. The program includes a neural network trained to identify secondary structure from residue sequence and chemical shifts.  相似文献   

16.
Polyamino acids which are soluble and helical in acetic acid and dichloroacetic acid (DCA) have been observed to undergo a helix to random chain transition upon the addition of lithium salts of strong acids. The transition can be reversed by diluting the salt. Apparently only lithium cations are able to bring about the polycarbobenzoxy-L -lysine (PCBL) transition in acetic acid, whereas the anions display a varying degree of effectiveness; ClO4? > Br? > TSA? > Cl? > NO3?. The lithium salts of carboxylate anions such as OAc? and TFA? do not cause polymer unwinding in acetic acid. Neither do the acids, TSA, HCl, TFA, or DCA induce the transformation in acetic acid. Poly-L -alanine (PLA) in DCA unfolds as LiBr is added, but does not unfold in the presence of 0.5M (CH3)4NBr, 0.25M CsBr, or 0.32M HCl. These results are explained on the basis of a direct interaction of the lithium salt with the polymer amide groups to form an ion-pair complex. The extent to which the union of the ion pair can dissociate from the complex in the low dielectric constant, environment determines the degree of unfolding of the polymer. The anion dissociation equilibrium presumably therefore would lie in the same order as given above. Acids such as HCl and TSA are considered to substantially protonate and ion-pair with the polymer, but do not readily dissociate the anion partner from the complex, and therefore do not produce an unstable positively charged helical structure.  相似文献   

17.
H Broch  D Vasilescu 《Biopolymers》1979,18(4):909-930
This paper reports a systematic PCILO study of the conformation of the nucleic acid backbone. The authors principally studied the ω′ and ω phosphodiester torsion angles of the disugar triphosphate model as a simultaneous function of (1) the sugar nature, ribose or deoxyribose, (2) the different combinations of the sugar ring puckers C(2′)-endo-C(2′)-endo, C(3′)-endo-C(3′)-endo, C(3′)-endo-C(2′)-endo, and C(2′)-endo-C(3′)-endo, and (3) the different conformations around the ψ(C4′–C5′) exocyclic bond. The dependence of the (ω′,ω) conformational energy maps upon these different factors, is discussed. The results are in very good agreement with the observed structures of ribonucleic (RNA10, RNA11, A′-RNA12, tRNAPhe) and deoxyribonucleic acids (D-DNA, C-DNA 9.3, B-DNA 10, A-DNA 11). Thus the validity of this model, the disugar triphosphate unit, is ensured. The main conclusions that can be drawn from this systematic study are the following:
  • 1 The torsion around P-05′ (angle ω) is, as a general rule, more flexible than the torsion around P-03′ (angle ω′).
  • 2 There is no notable difference between the ribose–triphosphate units and the deoxyribose–triphosphate units for the C(3′)-endo–C(3′)-endo and C(3′)-endo–C(2′)-endo sugar puckers.
  • 3 The deoxyribose–triphosphate units with C(2′)-endo–C(2′)-endo and C(2′)-endo–C(3′)-endo sugar puckers show much more ω′ flexibility than the ribose–triphosphate units with the same sugar puckers and cis position for the 2′hydroxyl group.
  • 4 The preferred values of ω′ are independent of the sugar nature (ribose or deoxyribose) and of ψ values; they are correlated with the sugar pucker of the first sugar-phosphate unit:
    • C(3′)-endo-C(3′)-endo and C(3′)-endo-C(2′)-endo puckers ? ω′ ? 240° (g? region)
    • C(2′)-endo-C(2′)-endo and C(2′)-endo-C(3′)-endo puckers ? ω′ 180° (t region)
  • 5 The preferred values of ω are independent of the nature and the puckering of the sugars; they are correlated with the rotational state of the torsion angle ψ(C4′–C5′): ψ ? 60° (gg) ? ω ? 300° (g?), ψ ? 180° (gt) or 300° (tg) ? ω ? 60° (g+)
  相似文献   

18.
Introducing experimental values as restraints into molecular dynamics (MD) simulation to bias the values of particular molecular properties, such as nuclear Overhauser effect intensities or distances, dipolar couplings, 3 J-coupling constants, chemical shifts or crystallographic structure factors, towards experimental values is a widely used structure refinement method. Because multiple torsion angle values ϕ correspond to the same 3 J-coupling constant and high-energy barriers are separating those, restraining 3 J-coupling constants remains difficult. A method to adaptively enforce restraints using a local elevation (LE) potential energy function is presented and applied to 3 J-coupling constant restraining in an MD simulation of hen egg-white lysozyme (HEWL). The method succesfully enhances sampling of the restrained torsion angles until the 37 experimental 3 J-coupling constant values are reached, thereby also improving the agreement with the 1,630 experimental NOE atom–atom distance upper bounds. Afterwards the torsional angles ϕ are kept restrained by the built-up local-elevation potential energies.  相似文献   

19.
Real and imaganiry parts of complex dielectric constant of dilute solutions of DNA in 10?3M NaCl with molecular weight ranging from 0.4 × 106 to 4 × 106 were measured at frequencies from 0.2 Hz to 30 kHz. Dielectric increments Δε were obtained from Cole-Cole plots and relaxation times τD from the loss maximum frequency. The τD of all samples agrees well with twice of the maximum viscoelastic relexation time in the Zimm theory, indicating that the low-frequency dielectric relaxiation should be ascribed to be the rotation of DNA. The rms dipole moment, which was obtained from Δε, agree well with that calculated from the counterion fluctuation theory. The dielectric increment was found to be greatly depressed in MgCl2, which is resonably interpreted in terms of a strong binding of Mg++ ions with DNA.  相似文献   

20.
The anomeric effect of 2-substituted 1,4-dioxane derivatives was calculated and compared with the values for substituted cyclohexane. The bond lengths, bond angles, torsion angles, and relative energies of axial and equatorial conformers of 2-substituted 1,4-dioxanes were calculated by the second-order Møller–Plesset (MP2), density functional theory (DFT/B3LYP), and Hartree–Fock (HF) methods using 6-31G basis set. The energy differences between the axial and equatorial conformers, endo and exo-anomeric effects, repulsive non-bond and H-bonding interactions were investigated. A linear free energy relationship (LFER) between calculated (MP2/6-31G) anomeric effect and inductive substituent constants (σI) was obtained for 2-substituted-1,4-dioxanes (slope = 6.19 and r2 = 0.967). The calculated energy differences indicate lower equatorial orientation for 2-substituted-1,4-dioxanes compared to the 2-substituted-tetrahydropyrans. The contribution of resonance, hyperconjugation, inductive, steric, hydrogen bonding, electrostatic interaction, and level of theory influences the anomeric effect.  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