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 359 毫秒
1.
The reaction of 2,6-dimethoxypyridine-3-carboxylic acid (DMPH) with different precursors [Ti(η5-C5H5)2Cl2], [Ti(η5-C5H4Me)2Cl2], [Ti(η5-C5H4SiMe3)(η5-C5H5)Cl2], [Ti(η5-C5Me5)Cl3], SnMe3Cl and GatBu3 yielded the complexes [Ti(η5-C5H5)2(DMP-κO)2] (1), [Ti(η5-C5H4Me)2(DMP-κO)2] (2), [Ti(η5-C5H4SiMe3)(η5-C5H5)(DMP-κO)2] (3), [Ti(η5-C5Me5)(DMP-κ2O,O′)3] (4), [SnMe3(μ-DMP-κOO′)] (5), and [GatBu2(μ-DMP-κOO′)]2 (6). 1-6 have been characterized by spectroscopic methods and the molecular structure of the complexes 1, 2, 3, 5 and 6 have been determined by X-ray diffraction studies. The cytotoxic activity of 1-6 was tested against the tumour cell lines human adenocarcinoma HeLa, human myelogenous leukaemia K562, human malignant melanoma Fem-x and human breast carcinoma MDA-MB-361. The results of this study show a higher cytotoxicity of the tin(IV) and gallium(III) derivatives in comparison to their titanium(IV) counterparts. Furthermore, the different titanium compounds showed differences in their cytotoxicities with a higher activity of complex 4 (mono-(cyclopentadienyl) derivative) compared to that of 1-3 (bis-(cyclopentadienyl) complexes). A qualitative UV-vis study of the interactions of these complexes with DNA has also been carried out.  相似文献   

2.
Reversed-phase HPLC analysis of the methanol extract of the seeds of Centaurea montana afforded a flavanone, montanoside (4), six epoxylignans, berchemol (7), berchemol 4′-O-β-d-glucoside (5), pinoresinol (10), pinoresinol 4-O-β-d-glucoside (8), pinoresinol 4,4′-di-O-β-d-glucoside (6), pinoresinol 4-O-apiose-(1→2)-β-d-glucoside (9), two quinic acid derivatives, trans-3-O-p-coumaroylquinic acid (1), cis-3-O-p-coumaroylquinic acid (2), and eight indole alkaloids, tryptamine (3), N-(4-hydroxycinnamoyl)-5-hydroxytryptamine (11), cis-N-(4-hydroxycinnamoyl)-5-hydroxytryptamine (12), centcyamine (16), cis-centcyamine (17), moschamine (13), cis-moschamine (14) and a dimeric indole alkaloid, montamine (15). While the structures of two new compounds, montanoside (4) and montamine (15), were established unequivocally by UV, IR, MS and a series of 1D and 2D NMR analyses, all known compounds were identified by comparison of their spectroscopic data with literature data. The antioxidant properties of these compounds were assessed by the DPPH assay, and their toxicity towards brine shrimps and cytotoxicity against CaCo-2 colon cancer cells were evaluated by the brine shrimp lethality and the MTT cytotoxicity assays, respectively. The novel dimer, montamine (15), showed significant in vitro anticolon cancer activity (IC50=43.9 μM) while that of the monomer, moschamine (13), was of a moderate level (IC50=81.0 μM).  相似文献   

3.
The use of succinamic acid (H2sucm) in CuII/N,N′,N″-donor [2,2′:6′,2″-terpyridine (terpy), 2,6-bis(3,5-dimethylpyrazol-1-yl)pyridine (dmbppy)] reaction mixtures yielded compounds [Cu(Hsucm)(terpy)]n(ClO4)n (1), [Cu(Hsucm)(terpy)(MeOH)](ClO4) (2), [Cu2(Hsucm)2(terpy)2](ClO4)2 (3), [Cu(ClO4)2(terpy)(MeOH)] (4), [Cu(Hsucm)(dmbppy)]n(NO3)n·3nH2O (5.3nH2O), and [CuCl2(dmbppy)]·H2O (6·H2O). The succinamate(−1) ligand exists in four different coordination modes in the structures of 13 and 5, i.e., the μ2OO′:κO″ in 1 and 5 which involves asymmetric chelating coordination of the carboxylato group and ligation of the amide O-atom leading to 1D coordination polymers, the μ22OO′ in 3 which involves asymmetric chelating and bridging coordination of the carboxylato group, and the asymmetric chelating mode in 2. The primary amide group, either coordinated in 1 and 5, or uncoordinated in 2 and 3, participate in hydrogen bonding interactions, leading to interesting crystal structures. Characteristic IR bands of the complexes are discussed in terms of the known structures and the coordination modes of the Hsucm ligands. The thermal decomposition of complex 5·3nH2O was monitored by TG/DTG and DTA measurements.  相似文献   

4.
Oxygenation of 2,5,9,12-tetra(tert-butyl)diacenaphtho[1,2-b:1′,2′-d]-thiophene (1, C40H44S) by peracids gave the cyclic sulfonic ester 4 (2,7,10,13-tetra(tert-butyl)diacenaphtho[1,2-c:1′,2′-e]oxathiin 5,5-dioxide, C40H44O3S) which, when heated in nitrobenzene, is converted into a complex, macrocyclic anhydride 3 (C80H88O3), which is derived from two molecules of 4. Further investigation found a likely intermediate in this reaction, 4,4′,7,7′-tetra(tert-butyl)-1,1′-biacenaphthylenylidene-2,2′-dione (5, C40H44O2), apparently formed from 4 by additional oxidation. Anhydride 3 plausibly arises by Diels-Alder reaction of 4 and 5 followed by several ring fragmentations. The structures of 3, 4, and 5 were unambiguously established by X-ray crystallography.  相似文献   

5.
Eight new compounds including 9′-[2-amino-3-(4″-O-methyl-α-rhamnopyranosyloxy) phenyl]nonanoic acid (1), 9′-[2-amino-3-(4″-O-methyl-α-ribopyranosyloxy)phenyl] nonanoic acid (2), 11′-[2-amino-3-(4″-O-methyl-α-rhamnopyranosyloxy)phenyl]undecanoic acid (3), 11′-[2-amino-3-(4″-O-methyl-α-ribopyranosyloxy)phenyl]undecanoic acid (4), 8-(4′-O-methyl-α-rhamnopyranosyloxy)-3,4-dihydroquinolin-2(1H)-one (5), 8-(4′-O-methyl-α-ribopyranosyloxy)-3,4-dihydroquinolin-2(1H)-one (6), 8-(4′-O-methyl-α-rhamnopyranosyloxy)-2-methyquinoline (7), and 8-(4′-O-methyl-α-ribopyranosyloxy)-2-methylquinoline (8) were isolated from Actinomadura sp. BCC27169. The chemical structures of these compounds were determined based on NMR and high-resolution mass spectroscopy. The absolute configurations of these monosaccharides were revealed by the hydrolysis of compounds 7 and 8. Compounds 3 and 8 exhibited antitubercular activity at MIC 50 μg/mL. Only compound 3 showed cytotoxicity against KB cell at IC50 18.63 μg/mL, while other isolated compounds were inactive at tested maximum concentration (50 μg/mL).  相似文献   

6.
The complex [(η6-C6Me6)Ru(μ-Cl)Cl]21 react with sodium salts of β-diketonato ligands in methanol to afford the oxygen bonded neutral complexes of the type [(η6-C6Me6)Ru(κ2-O,O′-R1COCHCOR2)Cl] {R1, R2 = CH3 (2), CH3, C6H5 (3), C6H5 (4), OCH3 (5), OC2H5 (6)}. Complex 4 with AgBF4 yields the γ-carbon bonded ruthenium dimeric complex 7. Complex 4 also reacts with tertiary phosphines and bridging ligands to yield complexes of the type [(η6-C6Me6)Ru(κ2-O,O′-C6H5COCHCOC6H5)(L)]+ (L = PPh3 (8), PMe2Ph (9)) and [{η6-C6Me6)Ru(κ2-O,O′-C6H5COCHCOC6H5)}2(μ-L)] L = 4,4′-bipyridine (4,4′-bipy) (11), 1,4-dicyanobenzene (DCB) (12) and pyrazine (Pz) (13). Complexes 2-4 react with sodium azide to yield neutral complexes [(η6-C6Me6)Ru(κ2-O,O′-R1COCHCOR2)N3] {R1, R2 = CH3 (10a), CH3, C6H5 (10b), C6H5 (10c). All these complexes were characterized by FT-IR and FT-NMR spectroscopy as well as analytical data. The molecular structures of complexes [(η6-C6Me6)Ru(κ2-O,O′CH3COCH-COC6H5)Cl] (3) and [(η6-C6Me6)Ru(κ2-O,O′-C6H5COCHCOC6H5] (4) were established by single crystal X-ray diffraction studies. The complex 3 crystallizes in the triclinic space group, [a = 7.9517(4), b = 9.0582(4) and c = 14.2373(8) Å, α = 88.442(3)°, β = 76.6.8(3)° and γ = 81.715(3)°. V = 987.17(9) Å3, Z = 2]. Complex 4 crystallizes in the monoclinic space group, P21/c [a = 7.5894(8), b = 20.708(2) and c = 29.208(3) Å,β = 92.059(3)° V = 4587.5(9) Å3, Z = 8].  相似文献   

7.
Lithiation of 2-dimethylaminoindene followed by quenching with [(R)-(1,1′-binaphthalene-2,2′-diyl)]chlorophosphite and treatment with triethylamine afforded the crystallographically characterized enantiopure P,N-indene 3 in 71% isolated yield. In the course of rhodium coordination chemistry studies involving 3, the formation of the isolable complex [(κ2-P,N-3)(κ1-P,N-3)RhCl] (7) (81%) was observed, thereby confirming the propensity of this new ligand to form LnRh(3)2 complexes. Such coordination chemistry behavior may contribute in part to the generally poor catalytic performance exhibited by mixtures of 3 and rhodium precursor complexes in the asymmetric hydrogenation and hydrosilylation studies described herein.  相似文献   

8.
Two polar phosphinoferrocene ligands, 1′-(diphenylphosphino)ferrocene-1-carboxamide (1) and 1′-(diphenylphosphino)ferrocene-1-carbohydrazide (2), were synthesized in good yields from 1′-(diphenylphosphino)ferrocene-1-carboxylic acid (Hdpf) via the reactive benzotriazole derivative, 1-[1′-(diphenylphosphino)ferrocene-1-carbonyl]-1H-1,2,3-benzotriazole (3). Alternatively, the hydrazide was prepared by the conventional reaction of methyl 1′-(diphenylphosphino)ferrocene-1-carboxylate with hydrazine hydrate, and was further converted via standard condensation reactions to three phosphinoferrocene heterocycles, viz 2-[1′-(diphenylphosphino)ferrocen-1-yl]-1,3,4-oxadiazole (4), 1-[1′-(diphenylphosphino)ferrocen-1-carbonyl]-3,5-dimethyl-1,2-pyrazole (5), and 1-[1′-(diphenylphosphino)ferrocene-1-carboxamido]-3,5-dimethylpyrrole (6). Compounds 1 and 2 react with [PdCl2(cod)] (cod = η22-cycloocta-1,5-diene) to afford the respective bis-phosphine complexes trans-[PdCl2(L-κP)2] (7, L = 1; 8, L = 2). The dimeric precursor [(LNC)PdCl]2 (LNC = 2-[(dimethylamino-κN)methyl]phenyl-κC1) is cleaved with 1 to give the neutral phosphine complex [(LNC)PdCl(1P)] (9), which is readily transformed into a ionic bis-chelate complex [(LNC)PdCl(12O,P)][SbF6] (10) upon removal of the chloride ligand with Ag[SbF6]. Pyrazole 5 behaves similarly affording the related complexes [(LNC)PdCl(5P)] (12) and [(LNC)PdCl(52O,P)][SbF6] (13), in which the ferrocene ligand coordinates as a simple phosphine and an O,P-chelate respectively, while oxadiazole 4 affords the phosphine complex [(LNC)PdCl(4P)] (11) and a P,N-chelate [(LNC)PdCl(42N3,P)][SbF6] (14) under similar conditions. All compounds were characterized by elemental analysis and spectroscopic methods (multinuclear NMR, IR and MS). The solid-state structures of 1⋅½AcOEt, 2, 7⋅3CH3CN, 8⋅2CHCl3, 9⋅½CH2Cl2⋅0.375C6H14, 10, and 14 were determined by single-crystal X-ray crystallography.  相似文献   

9.
Kin-ichi Oyama 《Tetrahedron》2004,60(9):2025-2034
We have succeeded in the first total synthesis of apigenin 7,4′-di-O-β-d-glucopyranoside (1a), a component of blue pigment, protodelphin, from naringenin (2). Glycosylation of 2 according to Koenigs-Knorr reaction provided a monoglucoside 4a in 80% yield, and this was followed by DDQ oxidation to give apigenin 7-O-glucoside (12a). Further glycosylation of 4′-OH of 12a with 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl fluoride (5a) was achieved using a Lewis acid-and-base promotion system (BF3·Et2O, 2,6-di-tert-butyl-4-methylpyridine, and 1,1,3,3-tetramethylguanidine) in 70% yield, and subsequent deprotection produced 1a. Synthesis of three other chiral isomers of 1a, with replacement of d-glucose at 7 and/or 4′-OH by l-glucose (1b-d), and four chiral isomers of apigenin 7-O-β-glucosides (6a,b) and 4′-O-β-glucosides (7a,b) also proved possible.  相似文献   

10.
The current paper describes the synthesis and spectral investigations on the adducts of [Zn(dbzdtc)2] (1) with 1,10-phen (2), tmed (3), 2,2′-bipy (4) and 4,4′-bipy (5) (where, dbzdtc = dibenzyldithiocarbamate anion, 1,10-phen = 1,10-phenanthroline, tmed = tetramethylethylenediamine, 2,2′-bipy = 2,2′-bipyridine, 4,4′-bipy = 4,4′-bipyridne) and single crystal X-ray structures of [Zn(dbzdtc)2(1,10-phen)] (2) and [Zn(dbzdtc)2(tmed)] (3) and [Zn(dbzdtc)2(4,4′-bipy)] (5). 1H and 13C NMR spectra of 1,10-phen, tmed, 2,2′-bipy and 4,4′-bipy adducts were recorded. 1H NMR spectra of the complexes show the drift of electrons from the nitrogen of the substituents forcing a high electron density towards sulfur via the thioureide π-system. In the 13C NMR spectra, the most important thioureide (N13CS2) carbon signals are observed in the region: 206–210 ppm. Fluorescence spectra of complexes (2) and (4) show intense fluorescence due to the presence of rigid conjugate systems such as 1,10-phenanthroline and 2,2′-bipyridine. The observed fluorescence maxima for complexes with an MS4N2 chromophore in the visible region are assigned to the metal-to-ligand charge transfer (MLCT) processes. Single crystal X-ray structural analysis of (2) and (3) showed that the zinc atom is in a distorted octahedral environment. Bond Valence Sum was found to be equivalent to 1.865 for (2), 1.681 for (3) supporting the correctness of the determined structure. BVS of (3) deviates from the formal oxidation number of zinc due to the non-aromatic, sterically hindering tetramethyl bonding end of tmed. Thermal studies on the compounds show the formation of Zn(NCS)2 as an intermediate during the decay.  相似文献   

11.
The reaction of the electronically unsaturated platina-β-diketone [Pt2{(COMe)2H}2(μ-Cl)2] (1a) with N?N donors led to the formation of diacetyl(hydrido)platinum(IV) complexes [Pt(COMe)2Cl(H)(N?N)] (2). By the reaction of these complexes with NaOH in a two-phase system (H2O/CH2Cl2) diacetylplatinum(II) complexes [Pt(COMe)2(N?N)] (N?N = bpy, 4a; 4,4′-Me2-bpy, 4b; 4,4′-t-Bu2-bpy, 4c; 4,4′-Ph2-bpy, 4d; 4,4′-t-Bu2-6-n-Bu-bpy, 4e; bpym, 4f; bpyr, 4g; phen, 4h; 4-Me-phen, 4i; 5-Me-phen, 4j) were obtained. All complexes were characterized by microanalysis, IR and 1H and 13C NMR spectroscopy. Additionally, complexes 4a, 4c, 4d and 4e were characterized by single-crystal X-ray diffraction analysis. The observed variety of packing patterns resulting from π-π stacking and hydrogen bonding is discussed.  相似文献   

12.
Six polymeric metal(II)-benzoate complexes of formula [Co2(O2CPh)4(4,4′-bpy)2]n (1-Co), [Ni(O2CPh)4(H2O)2(4,4′-bpy)]n (2-Ni), [Cu2(O2CPh)4(4,4′-bpy)]n (3-Cu), [Zn2(O2CPh)2(OH)2(4,4′-bpy)2]n (4-Zn), [Zn3(O2CPh)4(μ-OH)2(4,4′-bpy)2]n (5-Zn), and [Cd2(O2CPh)4(4,4′-bpy)2]n (6-Cd) have been synthesized and characterized (4,4′-bpy = 4,4′-bipyridine). 1-Co and 6-Cd show ladder-type double chains, 2-Ni does a helical structure, 3-Cu does a one-dimensional chain containing paddle-wheel units, 4-Zn does a zigzag chain, and 5-Zn does two-dimensional sheets. Since different structures provide different coordination geometry of each metal ion, it is clear that selection of appropriate metal ions can control the coordination geometry of each metal ion to form different crystal structures. Reactivity study of the compounds 17 for the transesterification of a variety of esters has shown that 4-Zn and 5-Zn are very efficient and the best among them. The catalyst 6-Cd containing Cd ion, well known as an inert metal ion for the ligand substitution, also catalyzed efficiently the transesterification of a variety of esters, and its reactivity is comparable to 4-Zn and 5-Zn. Moreover, the redox-active metal-containing polymers, 1-Co, 3-Cu, and 7-Mn, have shown efficient catalytic reactivities for the transesterification reactions, while 2-Ni has displayed a very slow conversion. The reactivities of the compounds used in this study are in the order of 5-Zn > 4-Zn > 6-Cd > 7-Mn ∼ 3-Cu > 1-Co > 2-Ni, indicating that the non-redox metal-containing compounds (5-Zn, 4-Zn, and 6-Cd) show better activity than the redox-active metal-containing compounds (7-Mn, 3-Cu, 1-Co, and 2-Ni). These results suggest that it is possible to tune the catalytic activities by changing from Zn to those metals such as Cd, a kinetically inert metal, or Cu, Mn, and Co, the redox-active metals.  相似文献   

13.
The synthetic investigation of the Cu(ClO4)2·6H2O/fumaric acid (H2fum)/N,N’-chelates (1,10-phen, 2,2′-bpy) tertiary reaction systems has yielded mononuclear, dinuclear and tetranuclear complexes, and three coordination polymers. The chemical and structural identity of the products depends on the solvent, the absence or presence of external hydroxides in the reaction mixtures and the N,N’-donor. Three fumarato(−2) complexes, i.e. compounds [Cu2(fum)(phen)4](ClO4)2·2H2O (1·2H2O), [Cu(fum)(phen)(H2O)]n (3) and [Cu2(fum)(bpy)2(H2O)2]n(ClO4)2n (6), were isolated and structurally characterized, and four non-fumarato complexes, i.e. compounds [Cu43-ΟΗ)22-ΟΗ)2(phen)4(H2O)2](ClO4)4·2H2O (2·2H2O), [Cu(ClO4)(phen) (MeCN)2(H2O)](ClO4) (4), [Cu(ClO4)(phen)(MeCN)2]n(ClO4)n (5) and [Cu(ClO4)2(bpy)(MeCN)2] (7), were simultaneously obtained from the reaction systems investigated. The coordination versatility of the fumarato(−2) ligand is reflected to the three different coordination modes observed in 1·2H2O, 3 and 6; the monodentate bridging μ2OO′ mode in 3, the asymmetric chelating bridging μ2OO′:κO′′:κO′′′ mode in 1·2H2O and 3, and the syn,syn bridging μ4OO′:κO′′:κO′′′ mode in 6. The crystal structures of the complexes are stabilized by intra- and inter-molecular hydrogen bonding and π–π stacking interactions leading to interesting supramolecular architectures. Characteristic IR bands of the complexes are discussed in terms of the known structures, and the coordination modes of the fum2− ligands.  相似文献   

14.
Reactions of Mo(II)-tetraphosphine complex [MoCl24-P4)] (2; P4 = meso-o-C6H4(PPhCH2CH2PPh2)2) with a series of small molecules have been investigated. Thus, treatment of 2 with alkynes RCCR′ (R = Ph, R′ = H; R = p-tolyl, R′ = H; R = Me, R′ = Ph) in benzene or toluene gave neutral mono(alkyne) complexes [MoCl2(RCCR′)(κ3-P4)] containing tridentate P4 ligand, which were converted to cationic complexes [MoCl(RCCR′)(κ4-P4)]Cl having tetradentate P4 ligand upon dissolution into CDCl3 or CD2Cl2. The latter complexes were available directly from the reactions of 2 with the alkynes in CH2Cl2. On the other hand, treatment of 2 with 1 equiv. of XyNC (Xy = 2,6-Me2C6H3) afforded a seven-coordinate mono(isocyanide) complex [MoCl2(XyNC)(κ4-P4)] (7), which reacted further with XyNC to give a cationic bis(isocyanide) complex [MoCl(XyNC)24-P4)]Cl (8). From the reaction of 2 with CO, a mono(carbonyl) complex [MoCl2(CO)(κ4-P4)] (9) was obtained as a sole isolable product. Reaction of 9 with XyNC afforded [MoCl(CO)(XyNC)(κ4-P4)]Cl (10a) having a pentagonal-bipyramidal geometry with axial CO and XyNC ligands, whereas that of 7 with CO resulted in the formation of a mixture of 10a and its isomer 10b containing axial CO and Cl ligands. Structures of 7 and 9 as well as [MoCl(XyNC)24-P4)][PF6](8′) and [MoCl(CO)(XyNC)(κ4-P4)][PF6] (10a′) derived by the anion metathesis from 8 and 10a, respectively, were determined in detail by the X-ray crystallography.  相似文献   

15.
We report herein the synthesis of appropriately protected 2′-deoxy-2′-fluoro-4′-thiouridine (5), -thiocytidine (7), and -thioadenosine (35) derivatives, substrates for the synthesis of novel modified RNAs. The synthesis of 5 and 7 was achieved via the reaction of 2,2′-O-anhydro-4′-thiouridine (3) with HF/pyridine in a manner similar to that of its 4′-O-congener whereas the synthesis of 35 from 4′-thioadenosine derivatives was unsuccessful. Accordingly, 35 was synthesized via the glycosylation of the fluorinated 4-thiosugar 25 with 6-chloropurine. The X-ray crystal structural analysis revealed that 2′-deoxy-2′-fluoro-4′-thiocytidine (8) adopted predominately the same C3′-endo conformation as 2′-deoxy-2′-fluorocytidine.  相似文献   

16.
The readily available 3-O-benzoyl-4-O-benzyl-1,2-O-isopropylidene-β-d-fructopyranose (6) was straightforwardly transformed into 5-azido-3-O-benzoyl-4-O-benzyl-5-deoxy-1,2-O-isopropylidene-β-d-fructopyranose (8), after treatment under modified Garegg's conditions followed by reaction of the resulting 3-O-benzoyl-4-O-benzyl-5-deoxy-5-iodo-1,2-O-isopropylidene-α-l-sorbopyranose (7) with lithium azide in DMF. O-debenzoylation at C(3) in 8, followed by oxidation and reduction caused the inversion of the configuration to afford the corresponding β-d-psicopyranose derivative 11 that was transformed into the related 3,4-di-O-benzyl derivative 12. Cleavage of the acetonide of 12 to give 13 followed by O-tert-butyldiphenylsilylation afforded a resolvable mixture of 14 and 15. Compound 14 was transformed into (2R,3R,4S,5R)- (17) and (2R,3R,4S,5S)-3,4-dibenzyloxy-2′,5′-di-O-tert-butyldiphenylsilyl-2,5-bis(hydroxymethyl)pyrrolidine (18) either by a tandem Staudinger/intramolecular aza-Wittig process and reduction of the resulting intermediate Δ2-pyrroline (16), or only into 18 by a high stereoselective catalytic hydrogenation. When 15 was subjected to the same protocol, (2S,3S,4R,5R)- (21) and (2R,3S,4R,5R)-3,4-dibenzyloxy-2′-O-tert-butyldiphenylsilyl-2,5-bis(hydroxymethyl)pyrrolidine (22) were obtained, respectively.  相似文献   

17.
Amide coupling between [2-(diphenylphosphino)phenyl]methylamine and 1′-(diphenylphosphino)ferrocene-1-carboxylic acid (Hdpf) afforded a novel diphosphine-amide, 1-{N-[(2-(diphenylphosphino)phenyl)methyl]carbamoyl}-1′-(diphenylphosphino)ferrocene (1), which was subsequently studied as a ligand for palladium(II) complexes. Depending on the metal precursor, the following complexes were isolated: [PdCl2(12P,P′)] (2), [PdCl(Me)(12P,P′)] (3), [(μ-1){PdCl2(PBu3)}2] (4) and [(μ-1){PdCl(LNC)}2] (LNC = 2-[(dimethylamino-κN)methyl]phenyl-κC1), featuring this ligand either as a trans-chelating or as a P,P′-bridging donor. The crystal structure of 2·1.25CH2Cl2 was established by X-ray crystallography, corroborating that 1 coordinates as a trans-spanning diphosphine without any significant distortion to the coordination sphere. Complex 2 together with a catalyst prepared in situ from 1 and palladium(II) acetate were tested in Suzuki-Miyaura reaction of aryl bromides with phenylboronic acid in dioxane.  相似文献   

18.
Sensitized photocycloaddition reactions of 6,6′-dimethyl-4,4′-[1,3-bis(methylenoxy)phenylene]-di-2-pyrone (1) with electron-poor α,ω-diolefins such as ethylene diacrylate (2a) and polyoxyethylene dimethacrylates (2b-d) afforded site- and stereoselective macrocyclic dioxatetralactones (3a-d) and (4b) having 18- to 25-membered rings across the C5-C6 and C5′-C6′ double bonds, or C5-C6 and C3′-C4′ double bonds in 1, respectively. Similar photoreactions of 1 with electron-rich α,ω-diolefins such as poly(ethylene glycol)divinyl ether (2e and 2f) afforded crown ether-type macrocyclic compounds (5e and 5f) having 18- and 21-membered rings across the C3-C4 and C3′-C4′ double bonds in 1, respectively. The stereochemical features of 3b, 5e-xx, and 5e-nn were determined by the X-ray crystal analysis. The reaction mechanism was inferred by MO methods.  相似文献   

19.
The use of succinamic acid (H2sucm)/N,N′-chelate (2,2′-bipyridine, bpy; 4,4′-dimethyl-2,2′-bipyridine, dmbpy; 1,10-phenanthroline, phen) ‘ligand blends’ in CuX2·yH2O (X = NO3, y = 3; X = Cl, y = 0) chemistry has yielded the new complexes [Cu2(Hsucm)3(bpy)2](NO3)·0.5MeOH (1·0.5MeOH), [Cu2(Hsucm)(OH)Cl(bpy)2](OH)·3.6H2O (5·3.6H2O) and [Cu2(Hsucm)2Cl2(phen)2] (6). The succinamate(−1) ion behaves as a carboxylate ligand and exists in two different coordination modes in the structures of the above complexes, i.e., the common syn, syn μ2OO′ in 1, 5 and 6, and the μ22OO′ in 1. The primary amide group of Hsucm remains uncoordinated and participates in intermolecular hydrogen bonding interactions leading to 1D, 2D and 3D networks. Characteristic IR bands of the complexes are discussed in terms of the known structures and the coordination modes of the Hsucm ligands.  相似文献   

20.
Giuseppe Faita 《Tetrahedron》2010,66(16):3024-5854
The asymmetric Friedel-Crafts reaction between methyl (E)-2-oxo-4-aryl-3-butenoates (1a-c) and activated benzenes (2a-d) has been efficiently catalyzed by the ScIII triflate complex of (4′S,5′S)-2,6-bis[4′-(triisopropylsilyl) oxymethyl-5′-phenyl-1′,3′-oxazolin-2′-yl]pyridine (pybox 3). The 4,4-diaryl-2-oxo-butyric acid methyl esters (4) are usually formed in good yields and the enantioselectivity is up to 99% ee. The sense of the stereoinduction can be rationalized with the same octahedral complex (10) between 1, pybox 3 and Sc triflate already proposed for other reactions involving pyruvates, and catalyzed by the same complex.  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