首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Combining micellar electrokinetic capillary chromatography (MEKC) and nuclear magnetic resonance (NMR) experimentation, we shed light on the structural basis for the chirally selective solubilization of atropisomeric binaphthyl compounds by bile salt micelles comprised of cholate (NaC) or deoxycholate (NaDC). The model binaphthyl analyte R,S‐BNDHP exhibits chirally selective interactions with primary micellar aggregates of cholate and deoxycholate, as does the closely related analyte binaphthol (R,S‐BN). Chiral selectivity was localized, by NMR chemical shift analysis, to the proton at the C12 position of these bile acids. Correspondingly, MEKC results show that the 12α‐OH group of either NaC or NaDC is necessary for chirally selective resolution of these model binaphthyl analytes by bile micelles, and the S isomer is more highly retained by the micelles. With NMR, the chemical shift of 12β‐H was perturbed more strongly in the presence of S‐BNDHP than R‐BNDHP. Intermolecular NOEs demonstrate that R,S‐BNDHP and R,S‐BN interact with a similar hydrophobic planar pocket lined with the methyl groups of the bile salts, and are best explained by the existence of an antiparallel dimeric unit of bile salts. Finally, chemical shift data and intermolecular NOEs support different interactions of the enantiomers with the edges of dimeric bile units, indicating that R,S‐BNDHP enantiomers sample the same binding site preferentially from opposite edges of the dimeric bile unit. Chirality 28:525–533, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

2.
Bile micelles play an important role in oral absorption of low‐solubility compounds. Bile micelles can affect solubility, dissolution rate, and permeability. For the pH–solubility profile in bile micelles, the HendersonHasselbalch equation should be modified to take bile‐micelle partition into account. For the dissolution rate, in the NernstBrunner equation, the effective diffusion coefficient in bile‐micelle media should be used instead of the monomer diffusion coefficient. The diffusion coefficient of bile micelles is 8‐ to 18‐fold smaller than that of monomer molecules. For permeability, the effective diffusion coefficient in the unstirred water layer adjacent to the epithelial membrane, and the free fraction at the epithelial membrane surface should be taken into account. The importance of these aspects is demonstrated here using several in vivo and clinical oral‐absorption data of low‐solubility model compounds. Using the theoretical equations, the food effect on oral absorption is further discussed.  相似文献   

3.
In order to define an enantioselective nuclear magnetic resonance (NMR) method for the antiasthmatic drug montelukast, a series of nine easily available products were evaluated as NMR chiral solvating agents (CSAs): D‐dibenzoyltartaric acid, D‐ditoluoyltartaric acid, (+)‐camphorsulfonic acid, (S)‐BINOL, (S)‐3,3’‐diphenyl‐2,2’‐binaphthyl‐1,1’‐diol, (R)‐3,3'′‐di‐9‐anthracenyl‐1,1'′‐bi‐2‐naphthol, (R)‐3,3'′‐di‐9‐phenanthrenyl‐1,1'′‐bi‐2‐naphthol, Pirkle's alcohol, and (?)‐cinchonidine. It was proved that most of the studied agents constitute diastereomeric complexes with both drug enantiomers in CD2Cl2 or CDCl3 solutions, thus permitting the direct 1H NMR detection of the unwanted S‐enantiomer, even at levels of 0.75%. (?)‐Cinchonidine was found to be the more convenient CSA in terms of NMR enantiodiscrimination power and ease of experimental requirements. The final method was validated and applied to the fast monitoring of the optical purity of montelukast “in‐process” samples, circumventing the need for tedious and slower analytical procedures like enantioselective chromatography or capillary electrophoresis. In addition, a method for the enantiopurity control of the commercial drug (montelukast sodium salt) was also established using (S)‐BINOL as NMR CSA. Chirality 25: 780–786, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
Mexiletine, an effective class IB antiarrhythmic agent, and its analogs were resolved on three different crown ether‐based chiral stationary phases (CSPs), one (CSP 1 ) of which is based on (+)‐(18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid and the other two (CSP 2 and CSP 3 ) are based on (3,3’‐diphenyl‐1,1’‐binaphthyl)‐20‐crown‐6. Mexiletine was resolved with a resolution (RS) of greater than 1.00 on CSP 1 and CSP 3 containing residual silanol group‐protecting n‐octyl groups on the silica surface, but with a resolution (RS) of less than 1.00 on CSP 2 . The chromatographic behaviors for the resolution of mexiletine analogs containing a substituted phenyl group at the chiral center on the three CSPs were quite dependent on the phenoxy group of analytes. Namely, mexiletine analogs containing 2,6‐dimethylphenoxy, 3,4‐dimethylphenoxy, 3‐methylphenoxy, 4‐methylphenoxy, and a simple phenoxy group were resolved very well on the three CSPs even though the chiral recognition efficiencies vary with the CSPs. However, mexiletine analogs containing 2‐methylphenoxy group were not resolved at all or only slightly resolved. Among the three CSPs, CSP 3 was found to show the highest chiral recognition efficiencies for the resolution of mexiletine and its analogs, especially in terms of resolution (RS). Chirality 26:272–278, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

5.
Gene expression can be altered by small molecules that target DNA; sequence as well as shape selectivities are both extremely important for DNA recognition by intercalating and groove‐binding ligands. We have characterized a carbohydrate scaffold (1) exhibiting DNA “shape readout” properties. Thermodynamic studies with 1 and model duplex DNAs demonstrate the molecule's high affinity and selectivity towards B* form (continuous AT‐rich) DNA. Isothermal titration calorimetry (ITC), circular dichroism (CD) titration, ultraviolet (UV) thermal denaturation, and Differential Scanning Calorimetry were used to characterize the binding of 1 with a B* form AT‐rich DNA duplex d[5′‐G2A6T6C2‐3′]. The binding constant was determined using ITC at various temperatures, salt concentrations, and pH. ITC titrations were fit using a two‐binding site model. The first binding event was shown to have a 1:1 binding stoichiometry and was predominantly entropy‐driven with a binding constant of approximately 108 M?1. ITC‐derived binding enthalpies were used to obtain the binding‐induced change in heat capacity (ΔCp) of ?225 ± 19 cal/mol·K. The ionic strength dependence of the binding constant indicated a significant electrolytic contribution in ligand:DNA binding, with approximately four to five ion pairs involved in binding. Ligand 1 displayed a significantly higher affinity towards AT‐tract DNA over sequences containing GC inserts, and binding experiments revealed the order of binding affinity for 1 with DNA duplexes: contiguous B* form AT‐rich DNA (d[5′‐G2A6T6C2‐3′]) >B form alternate AT‐rich DNA (d[5′‐G2(AT)6C2‐3′]) > A form GC‐rich DNA (d[5′‐A2G6C6T2‐3′]), demonstrating the preference of ligand 1 for B* form DNA. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 720–732, 2014.  相似文献   

6.
The binding of the fluorescent probes 1-anilino-8-naphthalene sulfonate and dansyl cadaverine to the sodium salts of cholic, deoxycholic and dehydrocholic acids has been investigated. Enhanced probe solubilisation accompanies aggregation. Monitoring of fluorescence intensities as a function of bile salt concentration permits the detection of primary micelle formation, as well as secondary association. The transition concentrations obtained by fluorescence are in good agreement with values determined for the critical micelle concentrations, by other methods. Differences in the behaviour of cholate and deoxycholate have been noted. Fluorescence polarisation studies of 1,6-diphenyl-1,3,5-hexatriene solubilised in bile salt micelles suggest a higher microviscosity for the interior of the deoxycholate micelle as compared to cholate. 1H NMR studies of deoxycholate over the range 1–100 mg/ml suggest that micelle formation leads to a greater immobilisation of the C18 and C19 methyl groups as compared to the C21 methyl group. Well resolved 13C resonances are observed for all three steroids even at high concentration. Both fluorescence and NMR studies confirm that dehydrocholate does not aggregate.  相似文献   

7.
Phospholipase C catalyzed hydrolysis of dimyristoyl phosphatidylcholine (DMPC) in phospholipid-bile salt mixed micelles was studied with particular attention on the relationship between interfacial enzyme activity and the physicochemical properties of substrate aggregates. Steady state kinetics is observed and it is argued that conditions for steady state exist because the enzyme encounters a steady supply of substrate by hopping between micelles at a rate faster than the chemical reaction rate. An existing kinetic model is reformulated to a more usable form. This presents a new approach to treating the kinetic data and allows extraction of the kinetic parameters of the model from the activity dependence on micellar lipid substrate surface concentration. The kinetic parameters were found to depend on the physicochemical properties of substrate aggregates, but remain constant over a range of lipid and bile salt concentrations. The substrate aggregates were characterized by time-resolved fluorescence quenching (TRFQ). The activity values and the micelle sizes group into two sets: (i) larger micelles for bile salt/lipid 5 with lower activity and longer steady state ( approximately 10 min). At least two sets of parameters, for bile salt/lipid 5, characterize the kinetics. Higher enzyme-micelle dissociation constant and lower catalytic rate are found for the group of smaller micelles. An explanation supporting our finding is that as micelles become smaller the overlap area for enzyme-micelle binding decreases, leading to weaker binding. Consequently the enzyme dissociation constant increases. Extension of the present approach to other phospholipases and substrates to establish its generality and correlation between micelle size and the catalytic rate are areas for future investigations.  相似文献   

8.
Chiral polymeric nanoparticles are of prime importance, mainly due to their enantioselective potential, for many applications such as catalysis and chiral separation in chromatography. In this article we report on the preparation of chiral polymeric nanoparticles by miniemulsion polymerization. In addition, we describe the use of isothermal titration calorimetry (ITC) to measure the chiral interactions and the energetics of the adsorption of enantiomers from aqueous solutions onto chiral polymeric nanoparticles. The characterization of chirality in nano‐systems is a very challenging task; here, we demonstrate that ITC can be used to accurately determine the thermodynamic parameters associated with the chiral interactions of nanoparticles. The use of ITC to measure the energetics of chiral interactions and recognition at the surfaces of chiral nanoparticles can be applied to other nanoscale chiral systems and can provide further insight into the chiral discrimination processes of nanomaterials. Chirality 27:613–618, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

9.
Micelle formations of sodium glyco- and taurochenodeoxycholate (NaGCDC and NaTCDC) and sodium glyco- and tauroursodeoxycholates (NaGUDC and NaTUDC) was studied at 308.2 K for their critical micelle concentrations at various NaCl concentrations by pyrene fluorescence probe, and the degree of counterion binding to micelle was determined using the Corrin-Harkins plots. The degree of counterion binding was found to be 0.37-0.38 for chenodeoxycholate conjugates, while the determination of the degree was quite difficult for ursodeoxycholate conjugates. The change of I1/I3 values on the fluorescence spectrum with the conjugate bile salt concentration suggested two steps for their bile salt aggregation. The first step is a commencement of smaller aggregates, the first cmc, and the second one is a starting of stable aggregates, the second cmc. The aggregation number was determined at 308.2 K and 0.15 M NaCl concentration by static light scattering: 16.3 and 11.9 for sodium NaGCDC and NaTCDC, and 7.9 and 7.1 for NaGUDC and NaTUDC, respectively. The solubilization of cholesterol into the bile salt micelles in the presence of coexisting cholesterol phase and the maximum additive concentration (MAC) of cholesterol was determined against the bile salt concentration. The standard Gibbs energy change for the solubilization was evaluated, where the micelles were regarded as a chemical species. The solubilization was stabilized in the order of NaGUDC approximately = NaTUDC < NaTC < NaGC < NaTCDC < NaGCDC < NaTDC < NaGDC, where the preceding results were taken into the order.  相似文献   

10.
Many macromolecular interactions, including protein‐nucleic acid interactions, are accompanied by a substantial negative heat capacity change, the molecular origins of which have generated substantial interest. We have shown previously that temperature‐dependent unstacking of the bases within oligo(dA) upon binding to the Escherichia coli SSB tetramer dominates the binding enthalpy, ΔHobs, and accounts for as much as a half of the observed heat capacity change, ΔCp. However, there is still a substantial ΔCp associated with SSB binding to ssDNA, such as oligo(dT), that does not undergo substantial base stacking. In an attempt to determine the origins of this heat capacity change, we have examined by isothermal titration calorimetry (ITC) the equilibrium binding of dT(pT)34 to SSB over a broad pH range (pH 5.0–10.0) at 0.02 M, 0.2 M NaCl and 1 M NaCl (25°C), and as a function of temperature at pH 8.1. A net protonation of the SSB protein occurs upon dT(pT)34 binding over this entire pH range, with contributions from at least three sets of protonation sites (pKa1 = 5.9–6.6, pKa2 = 8.2–8.4, and pKa3 = 10.2–10.3) and these protonation equilibria contribute substantially to the observed ΔH and ΔCp for the SSB‐dT(pT)34 interaction. The contribution of this coupled protonation (∼ −260 to −320 cal mol−1 K−1) accounts for as much as half of the total ΔCp. The values of the “intrinsic” ΔCp,0 range from −210 ± 33 cal mol−1 °K−1 to −237 ± 36 cal mol−1K−1, independent of [NaCl]. These results indicate that the coupling of a temperature‐dependent protonation equilibria to a macromolecular interaction can result in a large negative ΔCp, and this finding needs to be considered in interpretations of the molecular origins of heat capacity changes associated with ligand‐macromolecular interactions, as well as protein folding. Proteins 2000;Suppl 4:8–22. © 2000 Wiley‐Liss, Inc.  相似文献   

11.
One of the most powerful techniques that are currently available to measure thermodynamic parameters such as enthalpy (ΔH), Gibbs free energy (ΔG), entropy changes (ΔS), and binding affinity in chemical reactions is isothermal titration calorimetry (ITC). Recent advances in instrumentation have facilitated the development of ITC as a very essential analytical tool in biology and chemistry. In this article, we will focus on a review of the literature on the application of ITC for the study of chiral systems and chiral interactions. We present studies in which the ITC technique is used to study chiral interactions, for instance in chiral solutions, chiral organometallic complexes, guest‐host chiral binding interactions, and biological macromolecules. Finally, we put strong emphasis on the most recent application of ITC for the study of chirality in nanosystems and at the nanoscale.  相似文献   

12.
13.
Micellization of sodium chenodeoxycholate (NaCDC) was studied for the critical micelle concentration (CMC), the micelle aggregation number, and the degree of counterion binding to micelle at 288.2, 298.2, 308.2, and 318.2 K. They were compared with those of three other unconjugated bile salts; sodium cholate (NaC), sodium deoxycholate (NaDC), and sodium ursodeoxycholate (NaUDC). The I(1)/I(3) ratio of pyrene fluorescence and the solubility dependence of solution pH were employed to determine the CMC values. As the results, a certain concentration range for the CMC and a stepwise molecular aggregation for micellization were found reasonable. Using a stepwise association model of the bile salt anions, the mean aggregation number (n) of NaCDC micelles was found to increase with the total anion concentration, while the n values decreased with increasing temperature; 9.1, 8.1, 7.4, and 6.3 at 288.2, 298.2, 308.2, and 318.2 K, respectively, at 50 mmol dm(-3). The results from four unconjugated bile salts indicate that the number, location, and orientation of hydroxyl groups in the steroid nucleus are quite important for growth of the micelles. Activity of the counterion (Na(+)) was determined by a sodium ion selective electrode in order to confirm the low counterion binding to micelles. The solubilized amount of cholesterol into the aqueous bile salt solutions increased in the order of NaUDC相似文献   

14.
Rapid and simple isocratic high‐performance liquid chromatographic methods with UV detection were developed and validated for the direct resolution of racemic mixtures of hyoscyamine sulfate and zopiclone. The method involved the use of αl‐acid glycoprotein (AGP) as chiral stationary phase. The stereochemical separation factor (?) and the stereochemical resolution factor (Rs) obtained were 1.29 and 1.60 for hyoscyamine sulfate and 1.47 and 2.45 for zopiclone, respectively. The method was used for determination of chiral switching (eutomer) isomers: S‐hyoscyamine sulfate and eszopiclone. Several mobile phase parameters were investigated for controlling enantioselective retention and resolution on the chiral AGP column. The influence of mobile phase, concentration and type of uncharged organic modifier, ionic strength, and column temperature on enantioselectivity were studied. Calibration curves were linear in the ranges of 1–10 µg mL‐1 and 0.5–5 µg mL‐1 for S‐hyoscyamine sulfate and eszopiclone, respectively. The method is specific and sensitive, with lower limits of detection and quantifications of 0.156, 0.515 and 0.106, 0.349 for S‐hyoscyamine sulfate and eszopiclone, respectively. The method was used to identify quantitatively the enantiomers profile of the racemic mixtures of the studied drugs in their pharmaceutical preparations. Thermodynamic studies were performed to calculate the enthalpic ΔH and entropic ΔS terms. The results showed that enantiomer separation of the studied drugs were an enthalpic process. Chirality 28:49–57, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

15.
The equilibrium of bile salt between aqueous phase and mixed micelle was studied in solutions of pure bile salt and lecithin comparing taurocholate and taurochenodeoxycholate. The relationship between bile salt concentration in the aqueous phase and the ratio of bile salt/lecithin in the mixed micelle was determined by equilibrium dialysis on serial dilutions of these solutions. Extrapolation of this relationship to zero mixed-micellar bile salt permitted calculation of the critical micelle concentration (CMC) of the mixed micelle. For taurocholate, taurochenodeoxycholate, and an equimolar mix of these two bile salts, the mixed micelle CMC's were 3.1 mM, 0.47 mM, and 0.89 mM respectively. In the most concentrated solutions, aqueous phase bile salt concentration surpassed the CMC of the simple bile salt micelle by more than four-fold indicating the presence of simple micelles as well as mixed micelles. At all dilutions taurochenodeoxycholate had a much greater affinity for the mixed micelle than did taurocholate. This last finding may be the reason for the superior cholesterol solubilizing capacity of taurochenodeoxycholate-lecithin solutions compared to taurocholate-lecithin solutions.  相似文献   

16.
Vibrational circular dichroism (VCD) spectroscopy was applied to gelation by a chiral low‐molecular mass weight gelator, N,N’‐diperfluoroalkanoyl‐1,2‐trans‐diaminocyclohexane. Attention was focused on the winding effects of (–CF2)n chains on the gelating ability. For this purpose, a series of gelators were synthesized with perfluoroalkyl chains of different length (n = 6–8). When gelation was studied using acetonitrile as a solvent, the fibrils took different morphologies, depending on the chain length: twisted saddle‐like ribbon or helical ribbon from fibril (n = 6) and a helical ribbon from platelet (n = 8). The signs of VCD peaks assigned to the couplet of C=O stretching and to the C‐F stretching were also dependent on n, indicating that a gelator molecule changed conformation on elongating perfluoroalkyl chains. A model is proposed for the aggregation modes in fibrils. Chirality 28:361–364, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

17.
Precipitation of calcium palmitate from bile salt-containing dispersions   总被引:1,自引:0,他引:1  
Addition of calcium chloride to mixed micellar systems composed of sodium salts of palmitic acid and high concentrations of different bile acids results in precipitation of Ca(palmitate)2 only when the palmitate concentration exceeds a critical value, which is dependent on the concentrations of Ca2+, Na+ and bile salt, and on the type of bile salt used. All these dependencies, as well as the complex and interrelated effects of the various parameters on the kinetics of Ca(palmitate)2 precipitation are consistent with the following mechanism: (i) calcium binds to palmitate-bile salt mixed micelles and promotes their aggregation, at a rate governed by the concentration ratio between bound calcium and micelles (here denoted "binding ratio"). (ii) Ca(palmitate)2 precipitation occurs within the aggregate of micelles only if those micelles include sufficient amounts of Ca2+ and palmitate to allow for the formation of large enough crystal units of Ca(palmitate)2 which can serve as nucleation "seeds". Both the concentrations of micelles and Na+ have dual effects on the rate of precipitation. Increasing micelle concentration, by itself, accelerates aggregation but at the same time leads to a decrease of the binding ratio, thus reducing the rate of precipitation. Na+ which reduces the binding ratio through competitive binding also reduces the surface charge, thus assisting micelle aggregation. Our model also explains the facilitation of precipitation observed when phosphatidylcholine is contained in the palmitate-bile salt mixed micelles and the inhibitory effect of the water soluble bovine serum albumin.  相似文献   

18.
A novel nickel(II) hexaaza macrocyclic complex, [Ni(LR,R)](ClO4)2 ( 1 ), containing chiral pendant groups was synthesized by an efficient one‐pot template condensation and characterized (LR,R═1,8‐di((R)‐α‐methylnaphthyl)‐1,3,6,8,10,13‐hexaazacyclotetradecane). The crystal structure of compound 1 was determined by single‐crystal X‐ray analysis. The complex was found to have a square‐planar coordination environment for the nickel(II) ion. Open framework [Ni(LR,R)]3[C6H3(COO)3]2 ( 2 ) was constructed from the self‐assembly of compound 1 with deprotonated 1,3,5‐benzenetricarboxylic acid, BTC3?. Chiral discrimination of rac‐1,1′‐bi‐2‐naphthol and rac‐2,2,2‐trifluoro‐1‐(9‐anthryl)ethanol was performed to determine the chiral recognition ability of the chiral complex ( 1 ) and its self‐assembled framework ( 2 ). Binaphthol showed a good chiral discrimination on the framework ( 2 ). The optimum experimental conditions for the chiral discrimination were examined by changing the weight ratio between the macrocyclic complex 1 or self‐assembled framework 2 and racemates. The detailed synthetic procedures, spectroscopic data including single‐crystal X‐ray analysis, and the results of the chiral recognition for the compounds are described. Chirality, 25:54‐58, 2013 © 2012 Wiley Periodicals, Inc.  相似文献   

19.
Protein‐nucleic acids binding driven by electrostatic interactions typically are characterized by the release of counter ions, and the salt‐inhibited binding association constant (Ka) and the magnitude of exothermic binding enthalpy (ΔH). Here, we report a non‐classical thermodynamics of streptavidin (SA)–aptamer binding in NaCl (140–350 mM) solutions near room temperatures (23–27 °C). By using isothermal titration calorimetry (ITC) and circular dichroism (CD)/fluorescence spectroscopy, we found that the binding was enthalpy driven with a large entropy cost (ΔH ?20.58 kcal mol?1, TΔS ?10.99 kcal mol?1, and Ka 1.08 × 107 M?1 at 140 mM NaCl 25 °C). With the raise of salt concentrations, the ΔH became more exothermic, yet the Ka was almost unchanged (ΔH ?26.29 kcal mol?1 and Ka 1.50 × 107 M?1 at 350 mM NaCl 25 °C). The data suggest that no counter Na+ was released in the binding. Spectroscopy data suggest that the binding, with a stoichiometry of 2, was accompanied with substantial conformational changes on SA, and the changes were insensitive to the variation of salt concentrations. To account for the non‐classical results, we propose a salt bridge exchange model. The intramolecular binding‐site salt bridge(s) of the free SA and the charged phosphate group of aptamers re‐organize to form the binding complex by forming a new intermolecular salt bridge(s). The salt bridge exchange binding process requires minimum amount of counter ions releasing but dehydration of the contacting surface of SA and the aptamer. The energy required for dehydration is reduced in the case of binding solution with higher salt concentration and account for the higher binding exothermic mainly. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

20.
We investigated the stereodynamics of 5,5’‐substituted tropos BIPHEP ligands (2,2’‐bis(diphenylphosphino)‐biphenyls) by enantioselective dynamic high‐performance liquid chromatography (DHPLC) to elucidate the influence of the substitution pattern and electronics of the substituents (methyl, methoxy, and hydroxyl groups). By temperature‐dependent dynamic HPLC measurements the activation parameters ΔG, ΔH, and ΔS could be determined with high precision, revealing that the activation barrier of these 5,5’‐substituted BIPHEP ligands ranges in a narrow band between 87.8 and 93.0 kJ mol–1, making them highly attractive as deracemizable dynamic chiral ligands in asymmetric catalysis. Interestingly, the activation parameters are highly influenced by a hydroxyl or methoxy group in the 5,5’‐position of the BIPHEP ligands. Chirality 25:126–132, 2013. © 2012 Wiley Periodicals, Inc.  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