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.
We have developed and validated a high‐performance liquid chromatography method that uses monolithic silica disk‐packed spin columns and a monolithic silica column for the simultaneous determination of NG‐monomethyl‐l ‐arginine, NG,NG‐dimethyl‐l ‐arginine, and NG,NG′‐dimethyl‐l ‐arginine in human plasma. For solid‐phase extraction, our method employs a centrifugal spin column packed with monolithic silica bonded to propyl benzenesulfonic acid as a cation exchanger. After pretreatment, the methylated arginines are converted to fluorescent derivatives with 4‐fluoro‐7‐nitro‐2,1,3‐benzoxadiazole, and then the derivatives are separated on a monolithic silica column. l ‐Arginine concentration was also determined in diluted samples. Standard calibration curves revealed that the assay was linear in the concentration range 0.2–1.0 μM for methylated arginines and 40–200 μM for l ‐arginine. Linear regression of the calibration curve yielded equations with correlation coefficients of 0.999 (r2). The sensitivity was satisfactory, with a limit of detection ranging from 3.75 to 9.0 fmol for all four compounds. The RSDs were 4.3–4.8% (intraday) and 3.0–6.8% (interday). When this method was applied to samples from six healthy donors, the detected concentrations of NG‐monomethyl‐l ‐arginine, NG,NG‐dimethyl‐l ‐arginine, NG,NG′‐dimethyl‐l ‐arginine and l ‐arginine were 0.05 ± 0.01, 0.41 ± 0.07, 0.59 ± 0.11, and 83.8 ± 30.43 μM (n = 6), respectively.  相似文献   

2.
Treatment of amines with 1-(4-nitrophenol)-N-(O-benzylhydroxy)carbamate yields the O-benzyl protected N-hydroxyureas. Hydrogenation of the O-benzyl protected N-hydroxyureas over 5% Pd/BaSO4 cleanly gives the N-hydroxyureas in good yield. In addition to primary and secondary aliphatic and aromatic amines, this method converts amino sugars to the corresponding N-hydroxyureas without extensive protecting group chemistry.  相似文献   

3.
NG,NG‐dimethyl‐l ‐arginine (asymmetric dimethylarginine, ADMA),NG‐monomethyl‐l ‐arginine (l ‐NMMA) and NG,NG‐dimethyl‐l ‐arginine (symmetric dimethylarginine, SDMA) are released during hydrolysis of proteins containing methylated arginine residues. ADMA and l ‐NMMA inhibit nitric oxide synthase by competing with l ‐arginine substrate. All three methylarginine derivatives also inhibit arginine transport. To enable investigation of methylarginines in diseases involving impaired nitric oxide synthesis, we developed a high‐performance liquid chromatography (HPLC) assay to simultaneously quantify arginine, ADMA, l ‐NMMA and SDMA. Our assay requires 12 μL of plasma and is ideal for applications where sample availability is limited. We extracted arginine and methylarginines with mixed‐mode cation‐exchange columns, using synthetic monoethyl‐l ‐arginine as an internal standard. Metabolites were derivatized with ortho‐phthaldialdeyhde and 3‐mercaptopropionic acid, separated by reverse‐phase HPLC and quantified with fluorescence detection. Standard curve linearity was ≥0.9995 for all metabolites. Inter‐day coefficient of variation (CV) values were ≤5% for arginine, ADMA and SDMA in human plasma and for arginine and ADMA in mouse plasma. The CV value for l ‐NMMA was higher in human (10.4%) and mouse (15.8%) plasma because concentrations were substantially lower than ADMA and SDMA. This assay provides unique advantages of small sample volume requirements, excellent separation of target metabolites from contaminants and validation for both human and mouse plasma samples. © 2015 The Authors Biomedical Chromatography published by John Wiley & Sons, Ltd.  相似文献   

4.
The synthesis of oligonucleotides containing 8-aza-2′-deoxyguanosine (z8Gd; 1 ) or its N8-regioisomer z8Gd* ( 2 ) instead of 2′-deoxyguanosine (Gd) is described. For this purpose, the NH2 group of 1 and 2 was protected with a (dimethylamino)methylidene residue (→ 5, 6 ), a 4,4′-dimethoxytrityl group was introduced at 5′-OH (→ 7, 8 ), and the phosphonates 3a and 4 as well as the phosphoramidite 3b were prepared. These building blocks were used in solid-phase oligonucleotide synthesis. The oligonucleotides were characterized by enzymatic hydrolysis and melting curves (Tm values). The thermodynamic data of the oligomers 12–15 indicate that duplexes were stabilized when 1 was replacing Gd. The aggregation of d(T-G-G-G-G-T) ( 18 ) was studied by RP 18 HPLC, gel electrophoresis and CD spectroscopy and compared with that of oligonucleotides containing an increasing number of z8Gd residues instead of Gd. Similarly to [d(C-G)]3 ( 12a ), the hexamer d(C-z8G-C-z8G-C-G) ( 14 ) underwent salt-dependent B-Z transition.  相似文献   

5.
The synthesis and characterisation of new arborescent architectures of poly(L ‐lysine), called lysine dendrigraft (DGL) polymers, are described. DGL polymers were prepared through a multiple‐generation scheme (up to generation 5) in a weakly acidic aqueous medium by polycondensing Nε‐trifluoroacetyl‐L ‐lysine‐N‐carboxyanhydride (Lys(Tfa)‐NCA) onto the previous generation G(n?1) of DGL, which was used as a macroinitiator. The first generation employed spontaneous NCA polycondensation in water without a macroinitiator; this afforded low‐molecular‐weight, linear poly(L ‐lysine) G1 with a polymerisation degree of 8 and a polydispersity index of 1.2. The spontaneous precipitation of the growing Nε‐Tfa‐protected polymer (GnP) ensures moderate control of the molecular weight (with unimodal distribution) and easy work‐up. The subsequent alkaline removal of Tfa protecting groups afforded generation Gn of DGL as a free form (with 35–60 % overall yield from NCA precursor, depending on the DGL generation) that was either used directly in the synthesis of the next generation (G(n+1)) or collected for other uses. Unprotected forms of DGL G1–G5 were characterised by size‐exclusion chromatography, capillary electrophoresis and 1H NMR spectroscopy. The latter technique allowed us to assess the branching density of DGL, the degree of which (ca. 25 %) turned out to be intermediate between previously described dendritic graft poly(L ‐lysines) and lysine dendrimers. An optimised monomer (NCA) versus macroinitiator (DGL G(n?1)) ratio allowed us to obtain unimodal molecular weight distributions with polydispersity indexes ranging from 1.3 to 1.5. Together with the possibility of reaching high molecular weights (with a polymerisation degree of ca. 1000 for G5) within a few synthetic steps, this synthetic route to DGL provides an easy, cost‐efficient, multigram‐scale access to dendritic polylysines with various potential applications in biology and in other domains.  相似文献   

6.
The base-pairing properties of N7-(2-deoxy-β-D -erythro-pentofuranosyl)guanine (N7Gd; 1 ) are investigated. The nucleoside 1 was obtained by nucleobase-anion glycosylation. The glycosylation reaction of various 6-alkoxy-purin-2-amines 3a - i with 2-deoxy-3,5-di-O-(4-toluoyl)-α-D -erythro-pentofuranosyl chloride ( 8 ) was studied. The N9/N7-glycosylation ratio was found to be 1:1 when 6-isopropoxypurin-2-amine ( 3d ) was used, whereas 6-(2-methoxyethoxy)purin-2-arnine ( 3i ) gave mainly the N9-nucleoside (2:1). Oligonucleotides containing compound 1 were prepared by solid-phase synthesis and hybridized with complementary strands having the four conventional nucleosides located opposite to N7Gd. According to Tm values and enthalpy data of duplex formation, a base pair between N7Gd and dG is suggested. From the possible N7Gd dG base pair motives, Hoogsteen pairing can be excluded as 7-deaza-2′-deoxyguanosine forms the same stable base pair with N7Gd as dG.  相似文献   

7.
A novel synthesis of the bovine insulin B chain in the blocked form (7) applying Nim-Trt-, Cys-SEt-, NG-H⊕- and Benzyl-protection in the residual positions is described starting from the partial fragments 1–4. By choice the deprotection is possible by HF, HBr/TFE with C2H5SH addition and also under special conditions by catalytic hydrogenolysis. HBr/TFE acidolysis and hydrogenolysis lead to the best resultates in respect to yield and homogeneity of the final product. Thus Bunte salts are received in good yield related to the protected B chain and with the same insulin forming potency as such of native provenience.  相似文献   

8.
The concept of overall connectivity of a graph G was extended here to the definition of the overall hyper-Wiener index OWW(G) of a graph G, defined as the sum of the hyper-Wiener indexes in all subgraphs of G, as well as the sum of eth-order terms, e OWW(G), with e being the number of edges in the subgraph. The potential usefulness of the overall hyper-Wiener index in QSAR/QSPR is evaluated by its correlation with a number of properties of C3-C8 alkanes and cycloalkanes.  相似文献   

9.
The release of NO by [Fe(NO)(Et2NpyS4)], where (Et2NpyS4)2? = 2,6-bis(2-mercaptophenylthiomethyl)-4-substituted pyridine(2-), has been studied in the absence and presence of a trapping agent. The results show that [Fe(NO)(Et2NpyS4)] releases NO spontaneously in solution with a slow rate, k-NO = 1.7 × 10?4 s?1 at 23 °C, in a reversible reaction. NO release becomes faster when the reaction intermediate [Fe(Et2 NpyS4)] was trapped by CO, thereby preventing the back reaction. The release of NO was studied as a function of CO concentration and temperature. The reported activation parameters, especially the positive activation entropy values for the release of NO, favor the operation of a dissociative interchange (Id) mechanism. Thus, [Fe(NO)(Et2NpyS4)] can serve as a NO deliverer.  相似文献   

10.
The first total synthesis of sphingolipid (2S,3R,4E)‐N2‐octadecanoyl‐4‐tetradecasphingenine ( 1a ), a natural sphingolipid isolated from Bombycis Corpus 101A, and of its styryl analogue 1b was achieved in good overall yield (Schemes 1 and 2). The key step involved the installation with (E) stereoselectivity of a long lipophilic chain or phenyl group on allyl alcohol derivative 3 via a cross‐metathesis reaction (→ 5a or 5b ). The N‐Boc protected 3 was easily accessible from (S)‐Garner aldehyde.  相似文献   

11.
The application of the improved phosphoramidite strategy for the synthese of oligonucleotides using β-eliminating protecting groups to phospholipid chemistry offers the possibility to synthesize phospholipid conjugates of AZT ( 6 ) and cordycepin. The synthesis of 3′-azido-3′-deoxythymidine ( 6 ) was achieved by a new isolation procedure without chromatographic purification steps in an overall yield of 50%. Protected cordycepin ( = 3′-de-oxyadenosine) derivatives, the N6,2′-bis[2-(4-nitrophenyl)ethoxycarbonyl]cordycepin ( 12 ) and the N6,5′-bis[2-(4-nitrophenyl)ethoxycarbonyl]cordycepin ( 13 ) wre prepared by known methods and direct acylation of N6-[2-(4-nitrophenyl)ethoxycarbonyl]cordycepin ( 9 ), respectively. These protected nucleosides and the 3′-azido-3′-de-oxythymidine ( 6 ) reacted with newly synthesized and properly characterized lipid-phosphoramidites 21–25 , catalyzed by 1H-tetrazole, to the corresponding nucleoside-phospholipid conjugates 26–38 in high yield. The deprotection was accomplished via β-elimination with 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) in aprotic solvents to give analytically pure nucleoside-phospholipid diesters 39–51 as triethylammonium or sodium salts. The newly synthesized compounds were characterized by elemental analyses and UV and 1H-NMR spectra.  相似文献   

12.
The protected 2′‐deoxyguanosine derivatives 5a – c undergo N9N7 isomerization in the melt and in solution. The rate of isomerization is much faster than in the case of the corresponding ribonucleosides and occurs even in the absence of a catalyst. In the melt (195°, 2 min), the N2,3′‐O,5′‐O‐tris(4‐toluoyl) derivative 5b and the N2‐acetyl‐3′,5′‐bis‐O‐[(tert‐butyl)dimethylsilyl] derivative 5c gave anomeric mixtures of the N7‐isomers 9b / 10b (43%) and 9c / 10c (55%), respectively. In addition, the N9α‐D ‐anomers 8b and 8c were obtained. Different from 5b , the isomerization of peracetylated 5a resulted in low yields. Compound 5b was also prone to isomerization performed in solution (toluene, 100°, 5 min; chlorobenzene, 120°, 5 min), furnishing the N7‐regioisomers in 24–53% yield. The highest yield of the N9N7 isomerization occurred in the presence of 2‐deoxy‐3,5‐di‐O‐(4‐toluoyl)‐α‐D ‐erythro‐pentofuranosyl chloride.  相似文献   

13.
The N,N-diisopropylphosphoramidites 10a and 10b of appropriately protected chiral diastereoisomers of d(T[P-18O]-A) ( 8a and 8b , resp.), chiral by virtue of the isotope 18O at the P-atom, have been synthesized. The 18O-isotope was incorporated by oxidation of the phosphite triester 3 with H2[18O]/I2. Separation of the diastereoisomers was accomplished by flash chromatography of the O-3′-deprotected phosphate triesters 5a/b . The absolute configuration at the chiral P-atom was deduced from the methylation products of the fully deprotected diastereoisomers 8a and 8b . Phosphinylation of 5a and 5b yielded the configurationally pure phosphoramidites 10a and 10b , respectively, which were then employed in solid-phase synthesis to yield the self-complementary oligomers d(G-A-G-T-(Rp)-[P-18O]-A-C-T-C) ( 13 ) and d(G-A-G-T-(SP)-[P-18O]-A-C-T-C) ( 14 ), respectively.  相似文献   

14.
The starting material O‐protected glycosyl isothiocyanate ( 1?3 ) was refluxed with 1,4‐diaminobenzene in CHCl3 under nitrogen atmosphere to give 1,4‐bis(N‐glycosyl)thioureidobenzene ( 4?6 ). Then 1,4‐bis[N‐(4/6‐substituted benzothiazole‐2‐yl)‐N′‐glycosylguanidino]benzenes ( 8a?8e , 9a?9e , 10a?10e ) were obtained in good yield by reaction of compounds ( 4?6 ) with 2‐amino‐4/6‐benzothizoles ( 7a?7e ) and HgCl2 in the presence of TEA in DMF. The structures of all 18 new compounds were confirmed by IR, 1H NMR, LC‐MS and elemental analysis. The bioactivity of anti‐HIV‐1 protease (HIV‐1 PR) and against angiotensin converting enzyme (ACE) have been evaluated.  相似文献   

15.
The first results of a study aiming at an efficient preparation of a large variety of 2′‐O‐[(triisopropylsilyl)oxy]methyl(= tom)‐protected ribonucleoside phosphoramidite building blocks containing modified nucleobases are reported. All of the here presented nucleosides have already been incorporated into RNA sequences by several other groups, employing 2′‐O‐tbdms‐ or 2′‐O‐tom‐protected phosphoramidite building blocks (tbdms = (tert‐butyl)dimethylsilyl). We now optimized existing reactions, developed some new and shorter synthetic strategies, and sometimes introduced other nucleobase‐protecting groups. The 2′‐O‐tom, 5′‐O‐(dimethoxytrityl)‐protected ribonucleosides N2‐acetylisocytidine 5 , O2‐(diphenylcarbamoyl)‐N6‐isobutyrylisoguanosine 8 , N6‐isobutyryl‐N2‐(methoxyacetyl)purine‐2,6‐diamine ribonucleoside (= N8‐isobutyryl‐2‐[(methoxyacetyl)amino]adenosine) 11 , 5‐methyluridine 13 , and 5,6‐dihydrouridine 15 were prepared by first introducing the nucleobase protecting groups and the dimethoxytrityl group, respectively, followed by the 2′‐O‐tom group (Scheme 1). The other presented 2′‐O‐tom, 5′‐O‐(dimethoxytrityl)‐protected ribonucleosides inosine 17 , 1‐methylinosine 18 , N6‐isopent‐2‐enyladenosine 21 , N6‐methyladenosine 22 , N6,N6‐dimethyladenosine 23 , 1‐methylguanosine 25 , N2‐methylguanosine 27 , N2,N2‐dimethylguanosine 29 , N6‐(chloroacetyl)‐1‐methyladenosine 32 , N6‐{{{(1S,2R)‐2‐{[(tert‐butyl)dimethylsilyl]oxy}‐1‐{[2‐(4‐nitrophenyl)ethoxy]carbonyl}propyl}amino}carbonyl}}adenosine 34 (derived from L ‐threonine) and N4‐acetyl‐5‐methylcytidine 36 were prepared by nucleobase transformation reactions from standard, already 2′‐O‐tom‐protected ribonucleosides (Schemes 2–4). Finally, all these nucleosides were transformed into the corresponding phosphoramidites 37 – 52 (Scheme 5), which are fully compatible with the assembly and deprotection conditions for standard RNA synthesis based on 2′‐O‐tom‐protected monomeric building blocks.  相似文献   

16.
Abstract

Synthesis of biologically active oligosaccharides, haptens and their protein conjugates is a major area of interest because of their role in antigen-antibody interaction and receptor effects1. A number of these molecules contain α-or β-linked 2-acetamido-2-deoxy-D-glucosamine (GlcNAc) moieties. Most commonly, during the oligosaccharide synthesis, introduction of the β-glycosidically linked GlcNAc residue is achieved by either the oxazoline2 or the phthalimido method3. Of these, the latter is preferred because 2-N-phthalimido protected glycosamine units having a halogen or a thioalkyl group at C-1 have consistently proved to be more efficient donors than are the oxazolines. However, time and again, subsequent conversion of the N-phthalimido to amine by hydrazinolysis has proved inadequate. This has often resulted in a poor overall yield after an otherwise efficient synthesis. Recently it was shown that the phthalimido function could be removed under mild conditions from a number of amino acids4. We now report that this technique can be efficiently used for the deprotection of the phthalimido function in suitably protected carbohydrate compounds (2,3 and 5).  相似文献   

17.
Two new ω‐(1H‐pyrrol‐3‐yl)alkanoic acids 8a , b and the corresponding amines 9a , b were prepared on large scale in 41–51 and 27–39% overall yield, respectively, starting from N‐phenylsulfonyl‐protected pyrrole. The target compounds contain the desired functional groups for attachment of biomolecules such as proteins. During synthesis, an unprecedented partial reduction of the pyrrole ring with NaBH3CN in glacial AcOH was observed, for which a plausible mechanism is proposed (Scheme 3).  相似文献   

18.
The free-radical nitric oxide is now considered to play an important role in mammalian physiology and pathology. Enzymatic studies have shown that nitric oxide biosynthesis is initiated by an NADPH-dependentN-hydroxlation ofl-arginine, formingN -hydroxy-l-arginine as an intermediate. However, the subsequent enzymatic steps that generate nitric oxide fromN -hydroxy-l-arginine are unknown. We have used ab initio quantum chemical calculations to investigate a mechanism that forms nitric oxide fromN-hydroxyguanidine, used as a model forN -hydroxy-l-arginine. Our calculations indicate that mechanisms of nitric oxide formation involving nucleophilic attack by hydroperoxy anion at theN-hydroxyguanidine carbon are energetically feasible.  相似文献   

19.
Mononitrosyl and trans ‐Dinitrosyl Complexes of Phthalocyaninates of Manganese and Rhenium Tetra(n‐butyl)ammonium or di(triphenylphosphane)iminium nitrosylacidophthalocyaninato(2–)manganate, (cat)[Mn(NO)(X)pc2–] (X = ONO, NCO, N3; cat = nBu4N, PNP) is prepared from acidophthalocyaninato(2–)manganese, [Mn(X)pc2–], (cat)NO2 and (nBu4N)BH4 in CH2Cl2 or from nitrosylphthalocyaninato(2–)manganese, [Mn(NO)pc2–] and (nBu4N)X (X = ONO, NCO, N3, NCS) at T < 120 °C, respectively. [Mn(NO)(X)pc2–] dissociates in methanol, and [Mn(NO)pc2–] precipitates. Nitrito(O)phthalocyaninato(2–)manganese, (cat)NO2 and hydrogensulfide yield trans‐di(nitrosyl)phthalocyaninato(2–)manganate, trans[Mn(NO)2pc2–], isolated as red violet (PNP) and (nBu4N) complex salt. Nitrosyl(triphenylphosphane oxide)phthalocyaninato(2–)manganese, [Mn(NO)(OPPh3)pc2–] is obtained by addition of OPPh3 to [Mn(NO)pc2–] at 200 °C. Di(triphenylphosphane)phthalocyaninato(2–)rhenium(II) and (PNP)NO2 in CH2Cl2 or in molten (PNP)NO2 and PPh3 at 100 °C yields green blue l‐di(triphenylphosphane)iminium nitrosylnitrito(O)phthalocyaninato(2–)rhenate, l(PNP)[Re(NO)(ONO)pc2–]. Similarly, but with (nBu4N)NO2 red plates of tetra‐(n‐butyl)ammonium trans‐di(nitrosyl)phthalocyaninato(2–)rhenate, (nBu4N)trans[Re(NO)2pc2–] is isolated. Addition of (PNP)Br or (PNP)PF6 to a concentrated solution of (nBu4N)trans[Re(NO)2pc2–] in pyridine precipitates l(PNP)trans[Re(NO)2pc2–]. (nBu4N)trans[Re(NO)2pc2–] and PPh3 at 300 °C yield blue green nitrosyl(triphenylphosphane oxide)phthalocyaninato(2–)‐ rhenium, [Re(NO)(OPPh3)pc2–], that is oxidised with iodine precipitating nitrosyl(triphenylphosphane oxide)phthalocyaninato(2–)rhenium triiodide, [Re(NO)(OPPh3)pc2–]I3. The crystal structures of l(PNP)[Mn(NO)(ONO)pc2–] ( 1 ), l(PNP)‐ [Mn(NO)(NCO)pc2–] ( 2 ), l(PNP)trans[Mn(NO)2pc2–] ( 3 ), l(PNP)trans[Re(NO)2pc2–] ( 4 ) [Mn(NO)(OPPh3)pc2–] ( 5 ), [Re(NO)(OPPh3)pc2–] ( 6 ), and [Re(NO)(OPPh3)pc2–]I3 · CH2Cl2 ( 7 ) have been determined. The M–N(NO) distance varies between 1.623(12) Å in 5 and 1.846(3) Å in 3 . The M–N–O moiety is almost linear. The UV‐Vis spectra with the B band at ca. 14500 cm–1and the Q band at 30400 cm–1 do not dependent significantly on the axial ligand and the metal atom and its oxidation state. N–O stretching vibrations are observed in the IR spectra between 1701 cm–1 in 3 and 1753 cm–1 in [Mn(NO)pc2–] or for the Re series between 1571 cm–1 in 4 and 1724 cm–1 in 7 . M–N(NO) stretching and M–N–O deformation vibrations are assigned in the IR spectra and resonance Raman spectra between 486 cm–1 in 4 and 620 cm–1 in 1 .  相似文献   

20.
The degradation pathways of highly active [Cp*Ir(κ2-N,N-R-pica)Cl] catalysts (pica=picolinamidate; 1 R=H, 2 R=Me) for formic acid (FA) dehydrogenation were investigated by NMR spectroscopy and DFT calculations. Under acidic conditions (1 equiv. of HNO3), 2 undergoes partial protonation of the amide moiety, inducing rapid κ2-N,N to κ2-N,O ligand isomerization. Consistently, DFT modeling on the simpler complex 1 showed that the κ2-N,N key intermediate of FA dehydrogenation ( INH ), bearing a N-protonated pica, can easily transform into the κ2-N,O analogue ( INH2 ; ΔG≈11 kcal mol−1, ΔG ≈−5 kcal mol−1). Intramolecular hydrogen liberation from INH2 is predicted to be rather prohibitive (ΔG≈26 kcal mol−1, ΔG≈23 kcal mol−1), indicating that FA dehydrogenation should involve mostly κ2-N,N intermediates, at least at relatively high pH. Under FA dehydrogenation conditions, 2 was progressively consumed, and the vast majority of the Ir centers (58 %) were eventually found in the form of Cp*-complexes with a pyridine-amine ligand. This likely derived from hydrogenation of the pyridine-carboxiamide via a hemiaminal intermediate, which could also be detected. Clear evidence for ligand hydrogenation being the main degradation pathway also for 1 was obtained, as further confirmed by spectroscopic and catalytic tests on the independently synthesized degradation product 1 c . DFT calculations confirmed that this side reaction is kinetically and thermodynamically accessible.  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