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 46 毫秒
1.
N,N-Dimethylation of the H-Dmt-Tic-NH-CH(R)-R′ series of compounds produced no significant affect on the high δ-opioid receptor affinity (Ki=0.035–0.454 nM), but dramatically decreased that for the μ-opioid receptor. The effect of N-methylation was independent of the length of the linker (R); however, the bioactivities were affected by the chemical composition of the third aromatic group (R′): phenyl (Ph) (5′–8′) elicited a greater reduction in μ-affinity (40–70-fold) compared to analogues containing 1H-benzimidazole-2-yl (Bid) (9-fold). The major consequences of N,N-dimethylation on in vitro bioactivity were: (i) a loss of δ-agonism coupled with the appearance of potent δ antagonism (4′–7′) (pA2=8.14–9.47), while 1 exhibited only a 160-fold decreased δ agonism (1′) and the δ antagonism of 8 enhanced >10-fold (pA2=10.62, 8′); and (ii) a consistent loss of μ-affinity resulted in enhanced δ-opioid receptor selectivity. With the exception of compound 1′, the change in the hydrophobic environment at the N-terminus and formation of a tertiary amine by N,N-dimethylation in analogues of the Dmt-Tic pharmacophore produced potent δ-selective antagonists.  相似文献   

2.
3.
Two novel, weakly antiferromagnetically coupled, tetranuclear copper(II) complexes [Cu4(PAP)22-1,1-N3)22-1,3-N3)22-CH3OH)2(N3)4 (1) (PAP = 1,4-bis-(2′-pyridylamino)phthalazine) and [Cu4(PAP3Me)22-1,1-N3)22-1,3-N3)2(H2O)2(NO2)2]- (NO3)2 (2) (PAP3Me = 1,4-bis-(3′-methyl-2′-pyridyl)aminophthalazine) contain a unique structural with two μ2-1,1-azide intramolecular bridges, and two μ2-1,3-azide intermolecular bridges linking pairs of copper(II) centers. Four terminal azide groups complete the five-coordinate structures in 1, while two terminal waters and two nitrates complete the coordination spheres in 2. The dinuclear complexes [Cu2(PPD)(μ2-1,1-N3)(N3)2(CF3SO3)]CH3OH) (3) and [Cu2(PPD)(μ2-1,1-N3)(N3)2(H2O)(ClO4)] (4) (PPD = 3,6-bis-(1′-pyrazolyl)pyridazine) contain pairs of copper centers with intramolecular μ2-1,1-azid and pyridazine bridges, and exhibit strong antiferromagnetic coupling. A one-dimensional chain structure in 3 occurs through intermolecular μ2-1,1-azide bridging interactions. Intramolecular Cu-N3-Cu bridge angles in 1 and 2 are small (107.9 and 109.4°, respectively), but very large in 3 and 4 (122.5 and 123.2°, respectively), in keeping with the magnetic properties. 2 crystallizes in the monoclinic system, space group C2/c with a = 26.71(1), b = 13.51(3), c = 16.84(1) Å, β = 117.35(3)° and R = 0.070, Rw = 0.050. 3 crystallizes in the monoclinic system, space group P21/c with a = 8.42(1), b = 20.808(9), c = 12.615(4) Å, β = 102.95(5)° and R = 0.045, Rw = 0.039. 4crystallizes in the triclinic system, space group P1, with a = 10.253(3), b = 12.338(5), c = 8.072(4) Å, = 100.65(4), β = 101.93(3), γ = 87.82(3)° and R = 0.038, Rw = 0.036 . The magnetic properties of 1 and 2 indicate the presence of weak net antiferromagnetic exchange, as indicated by the presence of a low temperature maximum in χm (80 K (1), 65 K (2)), but the data do not fit the Bleaney-Bowers equation unless the exchange integral is treated as a temperature dependent term. A similar situation has been observed for other related compounds, and various approaches to the problem will be discussed. Magnetically 3 and 4 are well described by the Bleaney-Bowers equation, exhibiting very strong antiferromagnetic exchange (− 2J = 768(24) cm−1 (3); − 2J = 829(11) cm−1 (4)).  相似文献   

4.
The first examples of binary palladium(II) derivatives of unsaturated carboxylic acids are reported. It was found that the interaction of Pd3(μ-OAc)6 with the ,β-unsaturated 1-methylcrotonic (tiglic) and crotonic acids leads to the corresponding carboxylates of composition Pd3[μ-O2CC(R′) = CHMe]6, where R′ = Me (1) or H (2). The new compounds have been characterized by elemental analysis, solid and solution IR, 1H and 13C NMR, and ESI mass spectrometry. The crystal structure of 1 has been determined. This molecule displays a central Pd3 cyclic core with Pd–Pd distances of 3.093–3.171 Å. Each Pd–Pd bond is bridged by a pair of carboxylate ligands, one above and the other below the Pd3 plane, providing a square planar coordination for each Pd atom in an approximate D3h overall symmetry arrangement. Solution spectroscopic data show that the bridging η112 interaction of the carboxylates of 1 and 2 is readily displaced, with a change of the ligand to the terminal (η1) coordination mode.  相似文献   

5.
5e-tert-Butyl-2e-[4-(substituted-ethynyl)phenyl]-1,3-dithianes with selected functional groups (R) on the ethynyl moiety are potent blockers of the GABA-gated chloride channel measured as inhibitor concentration (IC50) for 4-n-[3H]propyl-1-(4-ethynylphenyl)-2, 6,7-trioxabicyclo[2.2.2]octanebinding to bovine brain membranes. The terminal R substituents were introduced by coupling 5e-tert-butyl-2e-(4-iodophenyl)-1,3-dithiane with HC ≡ CR or 5e-tert-butyl-2e-(4-ethynylphenyl)-1,3-dithiane with XR. The potency of the parent compound (R=H) with an IC50 of 21 μM is equaled or exceeded by up to 7-fold (i.e. IC50 = 3–21 μM) by several carboxylic acids [R = (CH2)nCO2H (n = 0–3), (CH2nOCH2CO2H (n = 1–3) and CH2SCH2 CO2H] and their esters and two phosphonic acids (CH2CH2PO3H2 and CH2OCH2PO3H2) but not their esters. These carboxyl and phosphonic acids (and their salts) include the most potent water-soluble chloride channel blockers known. Conversion to the monosulfones increases activity of the R = H and CH2OH analogs by 1.2- to 3-fold but decreases that of the R = CH2CH2CO2R′ (R′ = H or CH3) derivatives by 3- to 13-fold. Quantitative structure-activity analyses for 44 2-[4-(substituted-ethynyl)phenyl]-dithianes suggests that the principal feature of the R substituent for high activity is its polarizable volume modeled as molecular refractivity, i.e. this substituent is not a well-defined pharmacophore and undergoes a structurally non-specific interaction with the receptor. These observations lay the background for preparing candidate affinity probes.  相似文献   

6.
Reaction of [Au(η2-Ar){CH2C(O)R}Cl] (Ar=C6H4N=N- Ph-2, R=Me, C6H2(OMe)3-3′,4′,5′; Ar=C6H3(N=NC6H4Me- 4′)-2, Me-5, R=Me) with PPh3 and NaClO4·H2O (1:2:1) at room temperature, leads to reductive elimination giving [Au(PPh3)2]ClO4 and the corresponding carbon-carbon coupling product ArCH2C(O)R. A similar process takes place when complexes [Au(η2-Ar){CH2C(O)R}(PPh3)Cl] are refluxed in tetrahydrofuran, through elimination of [Au(PPh3)Cl].  相似文献   

7.
Novel thyrotropin-releasing hormone (TRH, pGlu-His-Pro-NH2) analogs, made by solid phase, were derived from the general scaffold pGlu-(D/L)Agl(X)-Pro-NH2 where Agl = aminoglycine. Analogs ranged from X being a proton to an acylating agent derived from substituted (aromatic heterocyclic rings) formic or acetic acids or an aminotriazolyl moiety (3′-amino-1H-1′,2′,4′-triazolyl) built on N of aminoglycine or Nβ of ,β-diaminoproprionic acid (Dpr). X was expected to mimic the electronic and structural characteristics of the imidazole ring of histidine. Analogs were purified by HPLC, characterized by mass spectrometry and isolated as either diastereoisomeric mixtures or pure isomers. Analogs, tested for their binding affinity to mouse pituitary TRH receptors, have apparent equilibrium inhibitory constants >1 μM.  相似文献   

8.
The cationic monoalkylated derivatives of the well-known metalloligand [Pt2(μ-S)2(PPh3)4], viz. [Pt2(μ-S)(μ-SR)(PPh3)4]+ (R = n-Bu, CH2Ph) are themselves able to act as metalloligands towards the Ph3PAu+ and R′Hg+ (R′ = Ph or ferrocenyl) fragments, by reaction with Ph3PAuCl or R′HgCl, respectively. The resulting dicationic products [Pt2(μ-SR)(μ-SAuPPh3)(PPh3)4]2+ and [Pt2(μ-SR)(μ-SHgR′)(PPh3)4]2+ are readily isolated as their hexafluorophosphate salts, and have been fully characterised by spectroscopic techniques and an X-ray structure determination on [Pt2(μ-SR)(μ-SHgFc)(PPh3)4](PF6)2.  相似文献   

9.
All-E-(3R,6′R)-3-hydroxy-3′,4′-didehydro-β,γ-carotene (anhydrolutein I) and all-E-(3R,6′R)-3-hydroxy-2′,3′-didehydro-β,ε-carotene (2′,3′-anhydrolutein II) have been isolated and characterized from extracts of human plasma using semipreparative high-performance liquid chromatography (HPLC) on a C18 reversed-phase column. The identification of anhydroluteins was accomplished by comparison of the UV-Vis absorption and mass spectral data as well as HPLC-UV-Vis-mass spectrometry (MS) spiking experiments using fully characterized synthetic compounds. Partial synthesis of anhydroluteins from the reaction of lutein with 2% H2SO4 in acetone, in addition to anhydrolutein I (54%) and 2′,3′-anhydrolutein II (19%), also gave (3′R)-3′-hydroxy-3,4-dehydro-β-carotene (3′,4′-anhydrolutein III, 19%). While anhydrolutein I has been shown to be usually accompanied by minute quantities of 2′,3′-anhydrolutein II (ca. 7–10%) in human plasma, 3′,4′-anhydrolutein III has not been detected. The presence of anhydrolutein I and II in human plasma is postulated to be due to acid catalyzed dehydration of the dietary lutein as it passes through the stomach. These anhydroluteins have also been prepared by conversion of lutein diacetate to the corresponding anhydrolutein acetates followed by alkaline hydrolysis. However, under identical acidic conditions, loss of acetic acid from lutein diacetate proceeded at a much slower rate than dehydration of lutein. The structures of the synthetic anhydroluteins, including their absolute configuration at C(3) and C(6′) have been unambiguously established by 1H NMR and in part by 13C NMR, and circular dichroism.  相似文献   

10.
The reaction of benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-isopropylidene-β-lactoside with 1,11-ditosyloxy-3,6,9-trioxaundecane gave benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-isopropylidene-3,2′-O--(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (2, 47%). Acid hydrolysis of 2 and condensation of the product with 1,14-ditosyloxy-3,6,9,12-tetra-oxatetradecane afforded benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-(3,6,9,12-tetraoxa-tetradecane-1,14-diyl)-3,2′-O-(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (29%). Similarly, the reaction of benzyl 2,6,2′,4′,6′-penta-O-benzyl-β-lactoside with Ts[OCH2CH2]4OTs gave benzyl 2,6,2′,4′,6′-penta-O-benzyl-3,3′-O-(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (78%). 1H-N.m.r. spectroscopy has been used to study the formation of host-guest complexes with some of these macrocyclic compounds and benzyl ammonium thiocyanate.  相似文献   

11.
The reaction between [(η6-p-cymene)Ru(H2O)3]X2 and 4,7-phenanthroline (phen) leads to the formation of the rectangular tetranuclear complexes [(η6-p-cymene)4Ru4(μ-4,7-phen-N4,N7)2(μ-OH)4]X4 (X = NO3, 1a; SO3CF3, 1b) which have been structurally characterised by X-ray crystallography. 1H NMR spectroscopic studies suggest the presence of a partially dissociated dinuclear species of type [(η6-p-cymene)2Ru2(μ-4,7-phen-N4,N7)(solv)4]4+ in equilibrium with the tetranuclear cyclic species found in the solid state. The temperature effect for this equilibrium was studied by variable temperature 1H NMR experiments in D2O and MeOD. The results reveal that the proportion of the tetranuclear species increases with the polarity of the solvent which favour stacking interactions between the phenanthroline moieties. In addition, the reactivity of the tetranuclear species towards the nucleosides guanosine (Guo), cytidine (Cyt), 2′-deoxythymidine (Thy) and 2′-deoxyadenosine (dAdo) has been monitored by 1H NMR as a potential model for the interaction of the 1 species with the probable DNA target. The results reveal that the 1 systems are able to bind the nucleobases endocyclic nitrogen atoms of Guo Cyt, and dAdo.  相似文献   

12.
The reaction of N,N,N′,N′-tetrakis(2-pyridylmethyl)ethylenediamine (tpen) with VCl3 in CH3CN yields Cl3V(tpen)VCl3 which was hydrolyzed in water in the presence of oxygen affording [V2O2(μ-OH)2(tpen)]I2·2H2O, the crystal structure of which has been determined. Asyn-{OV(μ-OH)2VO}2+ core has been identified where the V(IV) centers are antiferromagnetically coupled (J = −150 cm−1;g = 1.80).  相似文献   

13.
The dinuclear Pt---Si complex {(Ph3P)Pt{μ-η2-H---SiH(IMP)]}2 (trans-1a–cis-1b=3:1; IMP=2-isopropyl-6-methylphenyl) reacted with basic phosphines such as 1,2-bis(diphenylphosphino)ethane (dppe) and dimethylphenylphosphine (PMe2Ph) to afford different dinuclear Pt---Si complexes with loss of H2, {(P)2Pt[μ-SiH(IMP)]}2 [P=dppe, trans-2a (major), cis-2b (trace); PMe2Ph, 3 (trans only)]. Complexes 2 and 3 were characterized by multinuclear NMR spectroscopy and X-ray crystallography (2a). In contrast, the reaction of 1a,b with the sterically demanding tricyclohexylphosphine (PCy3) afforded {(Cy3P)Pt{μ-η2-H---SiH(IMP)]}2 (trans-4a–cis-4b 2:1) analogous to 1a,b where the central Pt2Si2(μ-H)2 core remains intact but the PPh3 ligands have been replaced by PCy3. Complexes 4a and 4b was characterized by multinuclear NMR and IR spectroscopies.  相似文献   

14.
Tungsten phosphoranylideneketene complexes of the type Tp′(CO)(p-OC6H4R)W(η2-(C,C)---O=CC---PR′2Ph) (R=NO2, R′=Me (6a); R=NO2, R′=Ph (6b); R=CN, R′=Me (7a); R=CN, R′=Ph (7b); R=Cl, R′=Ph (8b)) have been synthesized from phosphonium carbyne precursors in a reaction that reflects coupling of carbonyl and carbyne ligands. In addition to these products, aryloxycarbyne complexes Tp′(CO)2WCO(p-C6H4NO2) (9a), Tp′(CO)2WCO(p-C6H4CN) (9b), and Tp′(CO)2WCO(p-C6H4Cl) (9c)) have been prepared via substitution of the phosphonium carbyne phosphine with an aryloxide nucleophile. The product ratio of substitution at the carbyne carbon to carbonyl–carbyne coupling can be tuned by variation of the aryloxide para-substituent. Aryloxy carbyne complexes are the favored products with stronger nucleophiles, while weaker nucleophiles result in a mixture of aryloxy carbyne complexes and η2-ketenyl coupled complexes. Formation of η2-ketenyl complexes is favored for the least nucleophilic aryloxides. Ketenyl complexes 6a and 6b were methylated at the ketenyl oxygen to form cationic alkyne complexes [Tp′(CO)(p-OC6H4NO2)W(η2-(C,C)---CH3OCCPR2Ph)][OTf] (R=Me (10a), R=Ph (10b)). The structures of η2-ketenyl complexes 6a and 7b and the structure of cationic alkyne complex 10a were determined by X-ray crystallography.  相似文献   

15.
Monobridged-dinuclear platinum(II) complexes, where the bridging ligand is 4,4′-dipyrazolylmethane, have been prepared for use as potential anticancer agents. The complexes synthesized include [{cis-PtCl2(NH3)}2(μ-dpzm)], [{trans-PtCl2(Me2SO)}2(μ-dpzm)] and [{cis-PtCl2(Me2SO)}2(μ-dpzm)]. The characterization of these complexes is based on microanalytical, IR and 1H NMR data.  相似文献   

16.
The crystal and the molecular structure of 4,1′,6′-trichloro-4,1′,6′-trideoxy-galacto-sucrose (TGS) was determined by X-ray analysis at 294 K. Crystals of TGS are orthorhombic, space group P212121, with a = 7.318(3), b = 12.027(4), c = 18.136(5) Å, V = 1596(1) Å3, Z = 4; Dx = 1.655 g.cm-3, λ(MoK) = 0.71073 Å, μ(MoK) = 5.44 cm-1, F(000) = 816. The X-ray intensities of 2649 reflections with I 2.5σ(I) were measured with Zr-filtered MoK-radiation. The structure was solved by the Patterson procedure and refined by full-matrix least-squares to a final R-value of 0.0298. Large conformational differences between TGS and sucrose were observed, particularly in the conformation of the glycosidic linkage. These differences originate from chlorine substitution, which affects intramolecular hydrogen bonding and sweet-taste glucophores.  相似文献   

17.
The aim of the current study was to characterize the effects of chemical ischemia and reperfusion at the transductional level in the brain. Protein kinase C isoforms (, β1, β2, γ, δ and ) total levels and their distribution in the particulate and cytosolic compartments were investigated in superfused rat cerebral cortex slices: (i) under control conditions; (ii) immediately after a 5-min treatment with 10 mM NaN3, combined with 2 mM 2-deoxyglucose (chemical ischemia); (iii) 1 h after chemical ischemia (reperfusion). In control samples, all the PKC isoforms were detected; immediately after chemical ischemia, PKC β1, δ and isoforms total levels (cytosol + particulate) were increased by 2.9, 2.7 and 9.9 times, respectively, while isoform was slightly reduced and γ isoform was no longer detectable. After reperfusion, the changes displayed by , β1, γ, δ and were maintained and even potentiated, moreover, an increase in β2 (by 41 ± 12%) total levels became significant. Chemical ischemia-induced a significant translocation to the particulate compartment of PKC isoform, which following reperfusion was found only in the cytosol. PKC β1 and δ isoforms particulate levels were significantly higher both in ischemic and in reperfused samples than in the controls. Conversely, following reperfusion, PKC β2 and isoforms displayed a reduction in their particulate to total level ratios. The intracellular calcium chelator, 1,2-bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid, 1 mM, but not the N-methyl-d-asparate receptor antagonist, MK-801, 1 μM, prevented the translocation of β1 isoform observed during ischemia. Both drugs were effective in counteracting reperfusion-induced changes in β2 and isoforms, suggesting the involvement of glutamate-induced calcium overload. These findings demonstrate that: (i) PKC isoforms participate differently in neurotoxicity/neuroprotection events; (ii) the changes observed following chemical ischemia are pharmacologically modulable; (iii) the protocol of in vitro chemical ischemia is suitable for drug screening.  相似文献   

18.
In the present paper, the modulation of the basolateral membrane (BLM) Na+-ATPase activity of inner cortex from pig kidney by angiotensin II (Ang II) and angiotensin-(1–7) (Ang-(1–7)) was evaluated. Ang II and Ang-(1–7) inhibit the Na+-ATPase activity in a dose-dependent manner (from 10−11 to 10−5 M), with maximal effect obtained at 10−7 M for both peptides. Pharmacological evidences demonstrate that the inhibitory effects of Ang II and Ang-(1–7) are mediated by AT2 receptor: The effect of both polypeptides is completely reversed by 10−8 M PD 123319, a selective AT2 receptor antagonist, but is not affected by either (10−12–10−5 M) losartan or (10−10–10−7 M) A779, selective antagonists for AT1 and AT(1–7) receptors, respectively. The following results suggest that a PTX-insensitive, cholera toxin (CTX)-sensitive G protein/adenosine 3′,5′-cyclic monophosphate (cAMP)/PKA pathway is involved in this process: (1) the inhibitory effect of both peptides is completely reversed by 10−9 M guanosine 5′-O-(2-thiodiphosphate) (GDPβS; an inhibitor of the G protein activity), and mimicked by 10−10 M guanosine 5′-O-(3-thiotriphosphate) (GTPγS; an activator of the G protein activity); (2) the effects of both peptides are mimicked by CTX but are not affected by PTX; (3) Western blot analysis reveals the presence of the Gs protein in the isolated basolateral membrane fraction; (4) (10−10–10−6 M) cAMP has a similar and non-additive effect to Ang II and Ang-(1–7); (5) PKA inhibitory peptide abolishes the effects of Ang II and Ang-(1–7); and (6) both angiotensins stimulate PKA activity.  相似文献   

19.
Acid dissociation constants of 2,3-diphytanyl-sn-glycero-1-phosphoryl-sn-3′-glycero-1′-methylphosphate (PGP-Me), the major phospholipid in extreme halophiles (Halobacteriaceae), and of the demethylated 2,3-diphytanyl-sn-glycero-1-phosphoryl-sn-3′-glycero-1′-phosphate (PGP) and its deoxy analog 2,3-diphytanyl-sn-glycero-1-phosphoryl-1′-(1′,3′-propanediol-3′-phosphate) (dPGP), were calculated by an original mathematical procedure from potentiometric titration data in methanol/water (1:1, v/v) and found to be as follows: for PGP-Me (and presumably PGP), pK1=3.00 and pK2=3.61; for PGP, pK3=11.12; and for dPGP, pK1=2.72, pK2=4.09, and pK3=8.43. High-resolution 31P NMR spectra of intact PGP-Me in benzene/methanol or in aqueous dispersion showed two resonances corresponding to the two P-OH groups, each of which exhibited a chemical shift change in the pH range 2.0–4.5, corresponding to acid dissociation constants pK1=pK2=3.2; no further ionization (pK3) was detected at higher pH values in the range 5–12. The present results show that PGP-Me titrates as a dibasic acid in the pH range 2–8, but above pH 8, it titrates as a tribasic acid, presumably PGP, formed by hydrolysis of the methyl group during the titration with KOH. Calculation of the concentrations of the ionic molecular species of PGP-Me, PGP and dPGP as a function of pH showed that the dianionic species predominate in the pH range 5–9, covering the optimal pH for growth of Halobacteriaceae. The results are consistent with the concept that the hydroxyl group of the central glycerol moiety in PGP-Me and PGP is involved in the formation of an intramolecular hydrogen-bonded cyclic structure of the polar headgroup, which imparts greater stability to the dianionic form of PGP-Me and PGP in the pH range 5–9 and facilitates lateral proton conduction by a process of diffusion along the membrane surface of halobacterial cells.  相似文献   

20.
Complexes RuCl3(PPh3)L2 (L = MeIm (1a, Im (1b)) and [RuCl2(PPh3)2(bipy)]Cl·4H2O (2) have been synthesized via the ruthenium(III) precursor RuCl3(PPh3)2 (DMA), and characterized, including an X-ray structural analysis for 1a (MeIm = N-methylimidazole, Im = imidazole, bipy = 2,2′-bipyridyl, and DMA = N, N′-dimethylacetamide). Crystals of 1a are monoclinic, space group P21/n, A = 10.5491(5), B = 20.4934(9), C = 12.8285(4) Å, β = 90.166(4)°, Z = 4. The structure, which reveals a mer configuration for the chlorides, and cis-methylimidazoles, was solved by conventional heavy atom methods and was refined by full-matrix least-square procedures to R = 0.041 and Rw = 0.042 for 3328 reflections with I 3σ(I). From the RuCl2(PPh3)3 precursor, the ruthenium(II) complexes RuCl2(PPh3)2L2 and [RuCl(PPh3)L4]Cl have been made (L = Im or MeIm), while [RuCl(dppb)Im3]Cl has been made from [RuCl2(dppb)]2(μ-dppb) (dppb = Ph2P(CH2)4PPh2).  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