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 728 毫秒
1.
First examples of tungsten aminocarbene complexes [(OC5)W{C(SiR1nR23-n)NH2}] 2a-d (R1 = Ph, R2 = Me) were synthesized via ammonolysis of the corresponding methoxycarbene complexes 1a-d. They were characterized by NMR spectroscopy, MS, IR, UV/Vis and elemental analysis, and in the case of the C-triphenylsilyl derivative 2a by single-crystal X-ray structure analysis. The reaction of P-chloro alkylidenephosphane 3 with complexes 2a-d, meant to give 2H-azaphosphirene complexes, was monitored by 31P NMR spectroscopy to reveal the formation of the products 4-7, which were presumably formed via decomposition of the transient complexes 10a-d.  相似文献   

2.
Reaction of N-(2′-hydroxyphenyl)-4-R-benzaldimines (L-R, R = OCH3, CH3, H, Cl and NO2) with [Os(PPh3)3Br2] in refluxing 2-methoxyethanol in the presence of triethylamine affords two families of organoosmium complexes (1-R and 2-R). In both 1-R and 2-R complexes a benzaldimine ligand is coordinated to the metal center as tridentate C,N,O-donor. In the 1-R complexes, a bidentate N,O-donor imionsemiquinonate ligand, derived from the hydrolysis of another benzaldimine, and a PPh3 ligand are also coordinated to osmium. In the 2-R complexes, a carbonyl, derived from decarbonylation of 4-R-benzaldehyde (derived from the same hydrolysis stated above), and two PPh3 ligands take up the remaining coordination sites on osmium. Structures of the 1-Cl and 2-OCH3 complexes have been determined by X-ray crystallography. All the 1-R and 2-R complexes are diamagnetic, and show characteristic 1H NMR signals and intense MLCT transitions in the visible region. Cyclic voltammetry on the 1-R complexes shows a reversible Os(III)-Os(IV) oxidation within 0.47-0.67 V (vs SCE), followed by an irreversible oxidation of the imionsemiquinonate ligand within 1.10-1.36 V. An irreversible Os(III)-Os(II) reduction is also displayed by the 1-R complexes within −1.02 to −1.14 V. Cyclic voltammetry on the 2-R complexes shows a reversible Os(II)-Os(III) oxidation within 0.29-0.51 V, followed by a quasi-reversible oxidation within 1.04-1.29 V, and an irreversible reduction of the coordinated benzaldimine ligand within −1.16 to −1.31 V.  相似文献   

3.
The 4-R-benzaldehyde thiosemicarbazones (L-R) are known to react with [Rh(PPh3)3Cl] in refluxing ethanol in the presence of a base (NEt3) to afford organorhodium complexes (2-R), where the thiosemicarbazones are coordinated to rhodium as tridentate CNS donors with the sulfur atom oxidized by aerial oxygen to sulfone. Two triphenylphosphines and a hydride are also coordinated to the metal center. From the reaction with 4-nitrobenzaldehyde thiosemicarbazone, a second organorhodium complex (1-NO2) is obtained, in which the sulfur atom is not oxidized. Reaction of the 4-R-benzaldehyde thiosemicarbazones with [Rh(PPh3)3Cl] in refluxing ethanol in the absence of NEt3 affords another group of organorhodium complexes (3-R), in which the thiosemicarbazones are coordinated to rhodium as tridentate CNS donors, along with two triphenylphosphines and a chloride. In these 3-R complexes also the sulfur atom is not oxidized. Structures of all the complexes have been optimized by DFT calculations and compared with the already known X-ray crystallographic structures. Also the experimentally observed electronic absorption bands have been assigned to specific transitions based on the TDDFT studies. Molecular electrostatic potential (MESP) topographical analysis performed to find the deepest MESP point on the coordinated sulfur atom (Vmin) is used as a probe for assessing the oxidizability of the coordinated sulfur in 1-R and 3-R complexes. Energy differences between the three sets of complexes have been estimated and based on the results obtained, 3-R has been experimentally transformed into 2-R, via formation of 1-R as the intermediate.  相似文献   

4.
Yuji Takashima 《Tetrahedron》2010,66(1):197-2519
A general approach to the (S)- and (R)-isoflavans was invented, and efficiency of the method was demonstrated by the synthesis of (S)-equol ((S)-3), (R)-sativan ((R)-4), and (R)-vestitol ((R)-5). The key step is the allylic substitution of (S)-6a (Ar1=2,4-(MeO)2C6H3) and (R)-6b (Ar1=2,4-(BnO)2C6H3) with copper reagents derived from CuBr·Me2S and Ar2-MgBr (7a, Ar2=4-MeOC6H4; 7b, 2,4-(MeO)2C6H3; 7c, 2-MOMO-4-MeOC6H3), furnishing anti SN2′ products (R)-8a and (S)-8b,c with 93-97% chirality transfer in 60-75% yields. The olefinic part of the products was oxidatively cleaved and the Me and Bn groups on the Ar1 moieties was then removed. Finally, phenol bromide 9a and phenol alcohols 9b,c underwent cyclization with K2CO3 and the Mitsunobu reagent to afford (S)-3 and (R)-4 and -5, respectively.  相似文献   

5.
Two types of perfluoro alkyl-containing amphiphilic sulfones 7-9 and 13-15, respectively, and sulfonate betaines 23-32 were prepared using 2-[(perfluoroalkyl)methyl]oxiranes (1-3, RF = C4F9, C6F13, C8F17) or 3-(perfluoroalkyl)propyl iodides (16 and 17, RF = C6F13, C8F17) as the starting compounds. The overall yields of two-step syntheses were above 90%. The compounds 7-9 were prepared by the reaction of oxiranes 1-3 with 2-sulfanylethan-l-ol and subsequent oxidation of intermediate sulfides. Similarly, the amphiphiles 13-15 were obtained by analogous reaction of oxiranes 1-3 with thiomorpholine and subsequent oxidation of the sulfur atom in the morpholine ring. In the syntheses of betaines 23-32, the starting compounds 1-3 or 16 and 17 were first reacted with dimethylamine followed by the ring-opening reaction of the intermediate fluoroalkyl(dimethyl)amines with propane-1,3- or butane-1,4-sultones.  相似文献   

6.
Heats of formation have been derived from G3(MP2)//B3LYP and G3MP2B3(+) atomization energies for tert-butyl radical (6R), cubyl radical, bicyclooctyl radical (1R), and tricyclo[3.3.n.03,7]alk-3(7)-yl (n=0-3, 2R-5R) radicals, and their respective anions (1A-6A) and hydrocarbons (1H-6H). The electron affinity (EA) of 6R is estimated at 1.5±2 kcal/mol and tert-butyl anion (6A) is likely to be bound. In the homologous series 2R-5R the EAs range from 3.4±2 to 13.5±2 kcal/mol. The computed enthalpies of the acidities of the tricyclic hydrocarbons 1H-5H are in the range 407-411 kcal/mol. Their C-H bond dissociation energies (BDEs) are in the range 97-110 kcal/mol. The increase of the BDEs in the homologous series 2H-5H and the increase of EAs of 2A-5A is attributed to the enhanced pyramidalization induced in radicals 2R-5R by the shortening of the methylene chain connecting carbons C3 and C7.  相似文献   

7.
Reaction of 1,3-diaryltriazenes (R-C6H4-NN-(NH)-C6H4-R, R = OCH3, CH3, H, Cl, NO2 at the para position) with [Rh(PPh3)3Cl] in ethanol in the presence of a base (NEt3) affords a family of yellow complexes (1-R) containing a PPh3, two de-protonated triazenes coordinated as bidentate N,N-donors, and an aryl (C6H4-R) fragment coordinated in the η1-fashion. A similar reaction in toluene yields a group of reddish-orange complexes (2-R) containing a PPh3, two N,N-coordinated triazenes, and a chloride. Structures of the 1-CH3 and 2-CH3 complexes have been determined by X-ray crystallography. All the 1-R and 2-R complexes are diamagnetic, and show characteristic 1H NMR signals and intense MLCT transitions in the visible region. The 1-R and 2-R complexes also fluoresce in the visible region under ambient condition while excited at around 400 nm. Cyclic voltammetry on these complexes shows a Rh(III)-Rh(IV) oxidation (within 0.76-1.68 vs. SCE), followed by an oxidation of the coordinated triazene ligand (except the R = NO2 complexes). An irreversible reduction of the coordinated triazene is also observed for all the complexes below −0.96 V vs. SCE. In the 1-R and 2-R complexes potential of the Rh(III)-Rh(IV) oxidation correlates linearly with the electron-withdrawing nature of the para-substituent (R).  相似文献   

8.
Methylation of α-disubstituted cyclopentanone 1 with Me3Al in CH2Cl2 at 0 °C for 30 min gave diastereoselectively a mixture of (1R*,2S*)-2 and (1R*,2R*)-2 in a 96:4 ratio and 83% total yield. When the same methylation was carried out at 0 °C for 1 h and then at room temperature for 120 h, a diastereomeric mixture of (1R*,2S*)-2 and (1R*,2R*)-2 was obtained in a 12:88 ratio and in 88% total yield. The stereochemistry of the two diastereomers was determined by the results of acetalization of their diol derivatives 3 and 5. Isomerization between the Me2Al-alkoxides of (1R*,2S*)-2 and (1R*,2R*)-2 and its possible mechanism were investigated by HPLC analysis of the methylation reaction process at 0 °C for 1 h and then at room temperature for 56 h and also by their mutual epimerization reactions.  相似文献   

9.
The resolution by Lipase PS of rac-5 (from reduction of ketone 6, obtained from dicyclopentadiene with a new environment-friendly synthesis) gives (2S)-5, which was further reduced to the endo(2R)-1a alcohol. The endo(2S)-1b alcohol was obtained from camphor with a multistep synthesis. Pinacol couplings of 3a,b, carried out with Mg/Hg or Corey's general procedure respectively, afforded with high diastereoselectivity the C2 symmetry diols (2R,2′R)-2a and (2S,2′S)-2b, with endo oriented OH functions. The enantiogenic power of the endo alcohol (2R)-1a and (2S)-1b and of the diols (2R,2′R)-2a and (2S,2′S)-2b was tested towards the LiAlH4 reduction of acetophenone. The C2 symmetry appears to play a fundamental role.  相似文献   

10.
Machiko Ono  Yuki Shida 《Tetrahedron》2007,63(41):10140-10148
(±)-(4,5-anti)-4-Benzyloxy-5-hydroxy-(2E)-hexenoic acid 6 was subjected to δ-lactonization in the presence of 2,4,6-trichlorobenzoyl chloride and pyridine to give the α,β-unsaturated-δ-lactone congener (±)-7 (87% yield) accompanied by trans-cis isomerization. This δ-lactonization procedure was applied to the chiral synthesis of (+)-(4S,5R)-7 or (−)-(4R,5S)-7 from the chiral starting material (+)-(4S,5R)-6 or (−)-(4R,5S)-6. Deprotection of the benzyl group in (+)-(4S,5R)-7 or (−)-(4R,5S)-7 by the AlCl3/m-xylene system gave the natural osmundalactone (+)-(4S,5R)-5 or (−)-(4R,5S)-5 in good yield, respectively. Condensation of (−)-(4R,5S)-5 and tetraacetyl-β-d-glucosyltrichloroimidate 22 in the presence of BF3·Et2O afforded the condensation product (−)-8 (97% yield), which was identical to tetra-O-acetylosmundalin (−)-8 derived from natural osmundalin 9.  相似文献   

11.
A series of aluminum and zinc complexes supported by functionalized phenolate ligands were synthesized and characterized. Reaction of 2-(3,5-R2C3N2)C6H4NH2 (R = Me, Ph) with salicylaldehyde or 3,5-di-tert-butylsalicylaldehyde afforded 2-((2-(1H-pyrazol-1-yl)phenylimino)methyl)phenol derivatives 2a-2d. Treatment of 2a-2d with an equiv. of AlR23 (R2 = Me, Et) gave corresponding aluminum aryloxides 3a-3e, while reaction with an equiv. of ZnEt2 afforded zinc aryloxides 4a-4d. Treatment of 2c with 0.5 equiv. of ZnEt2 formed diphenolato zinc complex 5. All new compounds were characterized by 1H and 13C NMR spectroscopy and elemental analyses. The structures of complexes 3a, 4a and 5 were further characterized by single crystal X-ray diffraction techniques. The catalytic activity of complexes 3-5 toward the ring-opening polymerization of ε-caprolactone was studied. The zinc complexes (4a-4d) exhibited higher catalytic activity than the aluminum complexes (3a-3e). The diphenolato zinc complex 5 showed lower catalytic activity than the ethylzinc complexes 4a-4d. The aluminum complex (3b) is inactive to initiate the ROP of rac-lactide, while the zinc complex (4d) is active initiator for the ROP of rac-lactide, giving atactic polylactide.  相似文献   

12.
We have reported that our new axially dissymmetric ligand with two chiral centers, (Ra)-2,2′-bis[(R)-1H-1-hydroxyperfluorooctyl]biphenyl ((Ra)-(R)2-1c, or tentatively called as (Ra)-(R)2-PFCAB-7), worked as a good asymmetric inducer for the reaction of benzaldehyde with diethylzinc. Now, a mixture of (Ra)-(R)2- and (Sa)-(R)2-PFCAB-7 even in 1:4 ratio (−60% de) was found to give nearly the same asymmetric induction as pure (Ra)-(R)2-PFCAB-7 of the corresponding molar percents. This result suggests that both isomers do not form complex and that (Ra)-(R)2-PFCAB-7 accelerates the reaction and induces high asymmetry, while (Sa)-(R)2-1c does not accelerate the reaction significantly and does not induce asymmetry at all. This ligand of low ee, (Ra)-(R)2-PFCAB-7 of 20% ee, did not show appreciable asymmetric amplification, suggesting no formation of heterochiral complex.  相似文献   

13.
Fanhong Wu  Fanhua Xiao  Yongjia Shen 《Tetrahedron》2006,62(43):10091-10099
Sodium dithionite initiated free-radical addition of polyfluoroalkyl iodides (2m-2s) with norbornene 1a and its derivatives, such as norbornene-2-carboxylates 1b and 1c, and norbornene-2-carboxylic acids 1d and 1e was investigated. In all the cases, the addition of RF group was stereoselectively delivered at exo-position and the predominant configuration of products was trans. Under the similar condition, norbornene-2-carboxylic ethyl ester 1b reacted with 2p to give 6-exo-RF-5-endo-iodo adduct 3bp and 5-exo-RF-6-endo-iodo adduct 5bp in the ratio of 4:1. While 1c, which has a heavy crowded group in the 2-endo-position, gave 6-exo-RF-5-endo-iodo adduct 3cp and polyfluoroalkylated product 4cp retaining the trans-configuration and the exo-orientation of RF group. The fluoroalkylation-lactonization reaction occurred in the reaction of norbornene-2-endo-carboxylic acids 1d and 1e with polyfluoroalkyl iodides to afford the corresponding fluoroalkylated γ-lactone products (7dp-7ds, and 7em-7er). The configuration of the products was further confirmed by 2D NMR and X-ray diffraction analyses for the first time.  相似文献   

14.
Stapled helical l-leucine-based heptapeptides were synthesized and used as catalysts for the enantioselective epoxidation of α,β-unsaturated ketones. All N-terminal free stapled peptides were successfully used as chiral catalysts. Among them, the use of H-hS3,7hS-10 gave epoxide products with high enantioselectivities of up to 99% ee. Furthermore, the dominant conformations of the N-terminal protected stapled peptides R3,7R-10 and hS3,7hS-10 were investigated by 1H NMR, IR, CD spectra, and X-ray crystallographic analysis. The peptide R3,7R-10 formed a right-handed (P) α-helix in solution and in the crystalline state, while hS3,7hS-10 formed a right-handed (P) 310-helix in solution.  相似文献   

15.
Epoxidations of trans-β-methylstyrene, trans-stilbene and trans-methyl p-methoxycinnamate using chiral dioxiranes derived from both enantiopure diastereomers of α-fluoro cyclohexanones, (2S, 5R)-3a-6a and (2R, 5R)-3e-6e are studied and compared. From ab initio calculations at the HF/6-31G level of conformational inter-conversion for (2S, 5R)-D5a and (2R, 5R)-D5e dioxiranes it was found that, due to the α-fluorine atom, conformer K1 is more stable in the case of (2S, 5R)-D5a while conformer K2 is more stable in the case of (2R, 5R)-D5e. However, in both cases, the more stable conformers, K1 and K2, undergo rapid inter-conversion. Therefore, based on slow epoxidation reactions and rapid ring inversion of six-membered ring dioxiranes the Curtin-Hammett principle holds. Conformation K2 with axial fluorine having been found to be more reactive, the inversion of configuration observed for the epoxides obtained with ketones 3e-6e (compared with ketones 3a-6a) could be rationalized from competitive reactions of K2 and K1 conformations leading to simultaneous production of both (−) and (+) epoxides in the case of ketones 3e-6e.  相似文献   

16.
A series of antimalarial chiral 1,2,4-trioxanes (1-8) were synthesised in high enantiomeric purities. Enantioselective addition of R2Zn reagent to 3-methyl-2-butenal catalysed by (+)-MIB or (−)-MIB yielded both the enantiomers of the chiral allylic alcohols 9-11 (90-98% ee), which were subjected to diastereoselective photooxygenation in the presence of tetraphenylporphine (TPP) to obtain (R,R)-threo- or (S,S)-threo-β-hydroperoxy alcohols (12-14). Reaction of β-hydroperoxy alcohols (12-14) with different cyclic ketones produced optically active trioxanes 1-8.  相似文献   

17.
Bis(silylamino)tin dichlorides 1 [X2SnCl2 with X=N(Me3Si)2 (a), N(9-BBN)SiMe3 (b), N(tBu)SiMe3 (c), and N(SiMe2CH2)2 (d)] were prepared from the reaction of two equivalents of the respective lithium amides (Li-a-d) with tin tetrachloride, SnCl4, or from the 1:1 reaction of the respective bis(amino)stannylene with SnCl4. The compounds 1 react with two equivalents of lithium alkynides LiCCR1 to give the di(1-alkynyl)-bis(silylamino)tin compounds X2Sn(CCR1)2, 2 (R1=Me), 3 (R1=tBu), and 4 (R1=SiMe3). Problems were encountered, mainly with LiCCtBu as well as with 1b, since side reactions also led to the formation of 1-alkynyl-bis(silylamino)tin chlorides 5-7 and tri(1-alkynyl)(silylamino)tin compounds 8 and 9. 1,1-Ethylboration of compounds 2-4 led to stannoles 10, 11, and in the case of propynides, also to 1,4-stannabora-2,5-cyclohexadiene derivatives 12. The molecular structure of the stannole 11b (R1=SiMe3) was determined by X-ray analysis. The reaction of 2a and d with triallylborane afforded novel heterocycles, the 1,3-stannabora-2-ethylidene-4-cyclopentenes 14. These reactions proceed via intermolecular 1,1-allylboration, followed by an intramolecular 1,2-allylboration to give 14, and a second intramolecular 1,2-allylboration leads to the bicyclic compounds 15.  相似文献   

18.
Condensation of the O-protected hydroxyferrocene carbaldehyde (Sp)-1 with suitable diamines, followed by liberation of the hydroxyferrocene moiety leads to a new type of ferrocene-based salen ligands (3). While the use of ethylenediamine in the condensation reaction yields the planar-chiral ethylene-bridged ligand [(Sp,Sp)-3a], reaction with the enantiomers of trans-1,2-cyclohexylendiamine gives rise to the corresponding diastereomeric cyclohexylene-bridged systems [(S,S,Sp,Sp)-3b and (R,R,Sp,Sp)-3c], which feature a combination of a planar-chiral ferrocene unit with a centrochiral diamine backbone. Starting with the ferrocene-aldehyde derivative (Rp)-1, the enantiomeric ligand series (3d/e/f) is accessible via the same synthetic route.The (Sp)-series of these newly developed N2O2-type ligands was used for the construction of the corresponding mononuclear bis(isopropoxy)titanium (4a/b/c), methylaluminum (5a/b/c) and chloroaluminum-complexes (6a/b/c), which were isolated in good yields and identified by X-ray diffraction in several cases. The aluminum complexes (5/6) were successfully used in the Lewis-acid catalyzed addition of trimethylsilylcyanide to benzaldehyde, yielding the corresponding cyanohydrins in 45-62% enantiomeric excess.  相似文献   

19.
The reaction of RN(CH2CH2OH)CHR1CR2R3OH (1-8) with a stoichiometric amount of tetrachloro(bromo)germane leads to the corresponding RN(CH2CH2O)(CHR1CR2R3O)GeHal2 (9-21). Difluorenylgermocane 22 was prepared by treatment of diethoxydifluorenylgermane with N-methyldiethanolamine. Different dialkanolamines were found to be successive precursors of dimethylgermocanes, RN(CH2CH2O)(CHR1CR2R3O)GeMe2 (23-26). The chemical properties of simple and easy to access germocanes RN(CH2CH2O)2GeX2 [X = OH, Br (28), Cl (29)] were studied and the difluoro (27), haloalkoxy (30-32), and dialkoxy (33, 34) derivatives were prepared. The structures of the compounds 16, 20-22, and 26 were confirmed by X-ray diffraction and the structural features in solution of 23 and 26 were studied by NMR spectroscopy (NOEs). The relationship between the nature of substituents at different positions of the germocane skeleton and the strength of the intramolecular Ge ← N bond is discussed.  相似文献   

20.
Gelation of malonamides was investigated for the first time. Bis(phenylglycinol)malonamide 1, and methyl-, dimethyl-, ethyl-, diethyl- and isopropylmalonamides 2, 3, 4, 5 and 6, respectively, exhibited profoundly different gelling properties. Monoalkyl malonamides are efficient organogelators, and their gelling properties strongly depend on their stereochemistry. In contrast, symmetrically substituted dialkymalonamides, that is, (R,R)-dimethylmalonamide 3 and (R,R)-diethylmalonamide 5 as well as the unsubstituted 1 lack any gelation ability. Methyl derivative (R,R)-2 is an excellent, and its ethyl analogue (R,R)-4 a moderate gelator of toluene, p-xylene and tetralin while the isopropyl derivative (R,R)-6 shows only very weak gelation of tetralin and some more polar solvents. Meso diastereoisomers (R,r,S)-2 and (R,s,S)-2, as well as (R,r,S)-4 and (R,s,S)-4), each possessing a pseudoasymmetric centre represent very rare examples of gelling meso-compounds. The racemate 4 (rac-4) showed more efficient gelation of some solvents than the pure enantiomer (R,R)-4, while rac-2 failed to gel any of the solvents which were efficiently gelled by (R,R)-2.  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