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.
《Tetrahedron: Asymmetry》2000,11(13):2765-2779
The ligands 6-[(diphenylphosphanyl)methoxy]-4,8-di-tert-butyl-2,10-dimethoxy-5,7-dioxa-6-phosphadibenzo[a,c]cycloheptene, 1, (S)-4-[(diphenylphosphanyl)methoxy]-3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4a′]dinaphthalene, (S)-2, and (S)-4-[(diphenylphosphanyl)methoxy]-2,6-bis-trimethylsilanyl-3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4-a′]dinaphthalene, (S)-3, (S)-2-(3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4-a′]dinaphthalen-4-yloxymethyl)pyridine, (S)-4, and (S)-2-(3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4-a′]dinaphthalen-4-yloxy)pyridine, (S)-5, have been easily prepared.The cationic complexes [Pd(η3-C3H5)(L-L′)]CF3SO3 (L–L′=1–(S)-5) and [Pd(η3-PhCHCHCHPh)(L–L′)]CF3SO3 (L–L′=(S)-2–(S)-4) were synthesized by conventional methods starting from the complexes [Pd(η3-C3H5)Cl]2 and [Pd(η3-PhCHCHCHPh)Cl]2, respectively. The behavior in solution of all the π-allyl- and π-phenylallyl-(L–L′)palladium derivatives 614 was studied by 1H, 31P{1H}, 13C{1H} NMR and 2D-NOESY spectroscopy. As concerns the ligands (S)-4 and (S)-5, a satisfactory analysis of the structures in solution was possible only for palladium–allyl complexes [Pd(η3-C3H5)((S)-4)]CF3SO3, 11, and [Pd(η3-C3H5)((S)-5)]CF3SO3, 12, since the corresponding species [Pd(η3-PhCHCHCHPh)((S)-4)]CF3SO3, 13, and [Pd(η3-PhCHCHCHPh)((S)-5)]CF3SO3, 14, revealed low stability in solution for a long time. The new ligands (S)-2–(S)-5 were tested in the palladium-catalyzed enantioselective substitution of (1,3-diphenyl-1,2-propenyl)acetate by dimethylmalonate. The precatalyst [Pd(η3-C3H5)((S)-2)]CF3SO3 afforded the allyl substituted product in good yield (95%) and acceptable enantioselectivities (71% e.e. in the S form). A similar result was achieved with the precatalyst [Pd(η3-C3H5)((S)-3)]CF3SO3. The nucleophilic attack of the malonate occurred preferentially at allylic carbon far from the binaphthalene moiety, namely trans to the phosphite group. When the complexes containing ligands (S)-4 and (S)-5 were used as precatalysts, the product was obtained as a racemic mixture in high yield. The number of the configurational isomers of the Pd-allyl intermediates present in solution in the allylic alkylation and the relative concentrations are considered a determining factor for the enantioselectivity of the process.  相似文献   

2.
《Tetrahedron: Asymmetry》2007,18(4):464-475
In CDCl3 solution, enantiopure (S)-1-benzyl-6-methylpiperazine-2,5-dione (S)-1a formed diastereomeric COH–N hydrogen-bonded associates with racemic (RS,Z)-1-benzyl-3-[(dimethylamino)methylidene]piperazine-2,5-diones 2a and 2b, (RS)-tert-butyl pyroglutamate (RS)-2c and (RS)-N-benzoylalanine methyl ester (RS)-2d. This resulted in splitting (doubling) of the characteristic signals in the 1H NMR and 13C spectra of racemic compounds 2ad in the presence of 1 equiv of (S)-1a. The formation of hydrogen-bonded dimers in CDCl3 solution was studied by 1H NMR, 13C NMR and 2D NMR and confirmed by the intermolecular NOE observed between the hydrogen-bonded amide protons from each of the monomeric units, (S)-1a and 2ac. On the other hand, a slightly different binding mode was proposed for association of (S)-1a with alaninamide (RS)-2d. Enantiomer compositions of known (weighed) mixtures of both enantiomers of tert-butyl pyroglutamate 2c were re-determined by 1H NMR in the presence of (S)-1a in CDCl3. The experimental values were in good agreement with the theoretical values, thus indicating the potential applicability of (S)-1a and related diketopiperazines as chiral solvating agents in NMR spectroscopy.  相似文献   

3.
(1R,2S,6R)-2-Amino-6-hydroxycyclooctanecarboxylic acid (?)-10 was synthesized from (1R,2S)-2-aminocyclooct-5-enecarboxylic acid (+)-2 via an iodolactone intermediate, while (1R,2S,3R,4S)-2-amino-5,6-dihydroxycyclooctanecarboxylic acid (?)-12 was prepared by using the OsO4-catalyzed oxidation of Boc-protected amino ester (?)-5. The stereochemistry and relative configurations of the synthesized compounds were determined by 1D and 2D NMR spectroscopy (based on 2D NOE cross-peaks and 3J(H,H) coupling constants) and X-ray crystallography.  相似文献   

4.
《Tetrahedron: Asymmetry》1999,10(18):3615-3626
Seven new enantiomerically pure chiral pyridino-crown ethers (S,S)-4–(R,R)-10 were prepared. Three of them [(S,S)-4, (S,S)-7 and (R,R)-10] contain one, and two of them [(S,S)-5 and (S,S)-8] contain two linker chains with a terminal double bond. These linker chains were connected to the carbon atom at position 9 (opposite the pyridine moiety) of the macrocycle. The terminal double bond of the linker makes it possible to attach these ligands to silica gel to obtain chiral stationary phases (CSPs). The enantioselectivity of the new ligands toward the enantiomers of α-(1-naphthyl)ethylammonium perchlorate (NEA) was also determined by a titration 1H NMR method.  相似文献   

5.
《Tetrahedron: Asymmetry》2017,28(6):762-782
A convenient approach for the absolute configuration assignment of secondary alcohols in the (8R,1′R,2′S,5′R)-15,25, (8S,1′R,2′S,5′R)-15,25, (8R,1′R)-2124, and (8S,1′R)-2124 ester series, and of primary amines in the (8R,1′R)-3237 and (8S,1′R)-3237 amide series, by means of 1H NMR and VCD spectroscopy, using 2-cyano-2-indolylpropanoic acid as a chiral derivatizing agent is presented. DFT calculations were carried out to demonstrate the anisotropic effect of the indole skeleton on the chiral alcohol or the amine fragment. Vibrational circular dichroism (VCD) measurements of the above series indicated a VCD bisignated couplet resulting from the interaction of the ester carbonyl group and the CN group. The absolute configuration assignments were further tested by X-ray diffraction analysis.  相似文献   

6.
Reactions of the trans-PdCl2(PPh3)2 precursor with furan-2-carbaldehyde thiosemicarbazone (Hftsc) and thiophene-2-carbaldehyde thiosemicarbazone (Httsc), in 1:1 molar ratios in the presence of Et3N base, removed one Cl and one PPh3 group from the PdII center, and yielded the complexes [Pd(η2-N3,S-ftsc)(PPh3)Cl] (1) and [Pd(η2-N3,S-ttsc)(PPh3)Cl] (2), respectively. However, when a 1:2 molar ratio (M:L) was used, both Cl and PPh3 ligands were removed, yielding the complexes trans-[Pd(η2-N3,S-ftsc)2] (3) and trans-[Pd(η2-N3,S-ttsc)2] (4). Complexes 14 have been characterized with the help of analytical data, spectroscopic techniques (IR, 1H and 31P NMR) and single crystal X-ray crystallography. The thiosemicarbazone ligands behave as uninegative N3,S-chelating ligands in complexes 14. In contrast, pyrrole-2-carbaldehyde thiosemicarbazone (H2ptsc) and salicylaldehyde thiosemicarbazone (H2stsc) invariably formed the complexes [Pd(η3-N4,N3,S-ptsc)(PPh3)] (5) and [Pd(η3–O, N3,S-stsc)(PPh3)] (6), respectively, and the ligands acted as binegative tridentate donors (N4, N3, S, 5; O, N3, S, 6).  相似文献   

7.
Planar chiral alkenylferrocene phosphanes, viz. (Sp)-[Fe(η5-C5H3-1-PPh2-2-CHCR2)(η5-C5H5)] (R = H, (Sp)-2; Ph, (Sp)-5) and (Sp)-[Fe(η5-C5H3-1-PPh2-2-(E)-CHCHR)(η5-C5H5)] (R = Ph, (Sp)-3; C(O)CH3, (Sp)-6; and CO2CH2CH3, (Sp)-7) have been prepared by alkenylation of (Sp)-2-(diphenylphosphanyl)ferrocenecarboxaldehyde and tested as ligands for enantioselective palladium-catalysed allylic alkylation of 1,3-diphenyprop-2-en-1-yl acetate with dimethyl malonate. All phosphanylalkenes formed active catalysts. However, the induced enantioselectivity was only poor to moderate [12-43% ee after 20 h at room temperature], with the ee’s and configuration of the preferred product strongly depending on the ligand structure. The catalytic results have been related to solution properties (NMR, ESI MS) and the solid-state structural data (X-ray diffraction) of [Pd(η3-1,3-Ph2C3H3){(Sp)-22P}]ClO4 ((Sp)-12), which represent a model of the plausible reaction intermediate.  相似文献   

8.
A set of new diastereopure unsymmetrical α-diimine ligands 2a-d derived from methylglyoxal and optically pure primary amines 1a-d afforded the new chiral Pd(II)-complexes (S,S)-3a, (S,S)-3b, (S,S)-3c, and (1S, 2S, 3S, 5R)-3d. All compounds have been characterized by IR, 1H, and 13C NMR spectroscopies along with MS-FAB+ spectrometry. The crystal and molecular structure for the complexes 3a, 3b and 3d have been fully confirmed by single-crystal X-ray studies. Likewise, complexes 3a-d have also been screened for their in vitro cytotoxicity against different classes of cancer: leukemia (K-562 CML), colon cancer (HCT-15), human breast adenocarcinoma (MCF-7), central nervous system (U-251 Glio) and prostate cancer (PC-3) cell lines.  相似文献   

9.
《Tetrahedron: Asymmetry》1998,9(4):563-574
Homochiral crown ether (S,S)-1 containing 1-naphthyl groups as chiral barriers together with the phenol moiety was prepared by using (S)-3 as a chiral subunit which was resolved in enantiomerically pure form by lipase-catalyzed enantioselective acylation of (±)-3. Homochiral phenolic crown ether (S,S)-2, containing phenyl groups as chiral barriers, was also prepared from (S)-5 which was derived from (S)-mandelic acid. The association constants for their complexes with chiral amines in CHCl3 were determined at various temperatures by the UV–visible spectroscopic method demonstrating that the crown ethers (S,S)-1 and (S,S)-2 displayed the large ΔRSΔG values of 6.2 and 6.4 kJ mol−1, respectively, towards the amine 21 at 15°C. Thermodynamic parameters for complex formation were also determined and a linear correlation between TΔRSΔS and ΔRSΔH values was observed.  相似文献   

10.
《Tetrahedron: Asymmetry》1999,10(18):3493-3505
Perhydropyrimidinone (S)-1 is alkylated with very high diastereoselectivity to give trans products (2S,5R)-3, (2S,5R)–4 and (2S,5R)-5. Dialkylation of (S)-1 also proceeds with complete stereoselectivity to afford adducts (2S,5R)-6, (2S,5S)-6, (2S,5R)-7 and (2S,5S)-7. Hydrolysis (6N HCl, 100°C) of monoalkylated derivative (2S,5R)-3 gives enantiopure α-substituted β-amino acid (R)-8. Hydrolysis of dialkylated adducts 6 and 7 affords enantiopure α,α-disubstituted β-amino acids (R)- or (S)-9 and (R)- or (S)-10. Related iminoester (2S,6S)-2 is alkylated with complete diastereoselectivity to give products (2S,6S)-1113 whose hydrolysis under relatively mild conditions (2N CF3CO2H, CH3OH, 100°C) affords enantiopure N-benzoylated β,β-disubstituted β-amino acid esters (S)-1416, with intact double bonds in the olefinic substituents.  相似文献   

11.
Shuhei Higashibayashi 《Tetrahedron》2004,60(51):11977-11982
The NMR profiles (13C-δ, 1H-δ, 1H(OH)-δ, and 3JH,H) of syn- and anti-diols—3a,b in achiral solvents were found to be very similar to each other. Contrarily, their Δδδ=δ(R,R)-2δ(S,S)-2) behaviors in chiral bidentate NMR solvent (R,R)- and (S,S)-BMBA-p-Me (2) were found to be significantly different. On the basis of this NMR characteristic, a method has been developed to predict both the relative and absolute configurations of acyclic secondary 1,2-diols.  相似文献   

12.
C2-Symmetrical chiral thioureas (S,S)-1 and (S,S)-2 were prepared in good yield by the reaction of 2 equiv of inexpensive (S)-1-phenylethylamine, or the corresponding naphthyl analog, with 1 equiv of thiophosgene in the presence of excess triethylamine. The presence of asymmetric elements in (S,S)-1 and (S,S)-2, and their capacity to act as receptors for anionic species via hydrogen bonding were exploited in the development of 1H NMR spectroscopic enantiodiscrimination of chiral carboxylic acids. In particular, the diastereomeric complexes derived from thioureas (S,S)-1 and (S,S)-2 with ammonium salts of the chiral acids gave rise to well separated signals of the α-hydrogens and simple integration provides the corresponding enantiomeric ratios. Furthermore, it was observed that Cα-H in the (R) enantiomers of the chiral α-hydroxy and α-amino carboxylic acids studied in this work consistently appears downfield relative to the same signals in the (S) enantiomers.  相似文献   

13.
Twelve OCO bisacetamide ligands 4aa-4dc were synthesized after condensation of isophenylenediamines 1a-1d and anhydride/acyl chlorides. The corresponding Pd(II)–OCO–H(5aa-5ac), Pd(II)–OCO–Me(5ba-5bc), Pd(II)–OCO–OMe(5ca-5?cc), Pd(II)–OCO–NO2(5da-5dc) pincer complexes were prepared via C-H activation of precursors and Pd(OAc)2, and characterized by IR, 1H NMR, 13C NMR and elemental analysis. The α-arylation of ketones and aryl bromides catalyzed by 5 under low catalyst loadings (0.1?mol%) show that 5da exhibits the highest catalytic activity, resulting in a 98% isolated yield.  相似文献   

14.
《Tetrahedron: Asymmetry》1998,9(11):1891-1897
The absolute conformation and configuration of diastereomeric amides (4A,B6A,B) of (1S,3R)-camphanic acid (lactone of 1-hydroxy-2,2,3-trimethylcyclopentan-1,3-dicarboxylic acid, (−)-camphanic acid 9) with α-arylethylamines 13 are deduced from 1H NMR data and MM2 calculations. The α-arylethyl group in diastereomers A and B adopt nearly opposite absolute conformations, stabilized by hydrogen bonding in the syn-oriented O–C(1)–C(6)–N–H unit, and repulsive interaction between the 1′C–Me group and the amide CO group. The absolute configuration (1′S) is assigned to the 4A6A diastereomers, and the (1′R)-configuration to the 4B6B diastereomers; this assignment is confirmed by the preparation of 4A and 5A from enantiomerically pure (1′S)-α-arylethylamines 1 and 2, respectively. These results also enabled the assignment of pro-R (HR) and pro-S (HS) protons in the benzyl derivative 7.  相似文献   

15.
《Tetrahedron: Asymmetry》1999,10(21):4075-4078
Enantiopure 2-hydroxy-2-(1-naphthyl)propionic acid (+)-2 was prepared by the stereoselective Grignard reaction of 1-naphthylmagnesium bromide with (1R,3R,4S)-menthyl pyruvate 3 or (1R,3R,4S)-8-phenylmenthyl pyruvate 4, and the absolute configuration of acid (+)-2 was unambiguously determined to be S by the 1H NMR anisotropy method.  相似文献   

16.
《Tetrahedron: Asymmetry》2000,11(6):1249-1253
Racemic 2-methoxy-2-(1-naphthyl)propionic acid (1, MαNP acid) was enantioresolved as its esters derived from various chiral alcohols. For example, a diastereomeric mixture of esters prepared from (±)-1 and (1R,3R,4S)-(−)-menthol was easily separated by HPLC on silica gel yielding esters (−)-2a and (−)-2b, the separation factor α=1.83 being unusually large. The 1H NMR chemical shift differences, Δδ=δ(R)–δ(S), between diastereomers 2a and 2b, are much larger than those of conventional chiral auxiliaries, e.g. Mosher’s MTPA and Trost’s MPA acids. This acid 1 is therefore very powerful for determining the absolute configuration of chiral alcohols by the 1H NMR anisotropy method. Solvolysis of the separated esters yielded enantiopure acids (S)-(+)-1 and (R)-(−)-1, which are useful for enantioresolution of racemic alcohols.  相似文献   

17.
A series of rigid and chiral C2-symmetric 18-crown-6 type macrocycles (S,S)-4, (S,S)-5, (S,S)-6 and (R,R)-2 bearing diamide–ester groups were synthesized. The binding properties of these macrocycles were examined for α-(1-naphthyl)ethylammonium perchlorates salts by an 1H NMR titration method. Taking into account the host employed, important differences were observed in the Ka values of (R)- and (S)-enantiomers of guests for macrocycles (S,S)-4 and (S,S)-6, KS/KR = 3.6, and KS/KR = 0.1 (KR/KS = 10.3) ΔΔG = 3.19 and ΔΔG = ?5.77 kJ mol?1, respectively. The results indicated excellent enantioselectivity of macrocyclic (S,S)-6 towards the enantiomers of α-(1-naphthyl)ethylammonium perchlorate salts.  相似文献   

18.
《Tetrahedron: Asymmetry》2006,17(21):2976-2980
The insect pheromone (2S,3R,7R)-3,7-dimethyltridec-2-yl acetate 1-Ac was prepared from diastereomerically enriched (2S1,3R1,7R)-1, which in turn was obtained by the coupling of racemic 3,4-dimethyl-γ-butyrolactone with (7S)-2-methyloctyllithium, followed by a Wolff–Kishner reduction of the resulting ketone. Conversion of (2S1,3R1,7R)-1 to the corresponding alkyl hydrogen phthalate and diastereomer salt formation with (S)-PhCHMeNH2 provided after several crystallizations individual diastereomer, which was later transformed into target 1-Ac after hydrolysis and acylation.  相似文献   

19.
This paper summarizes our recent studies on the mechanism of O-atom transfer to a benzylic C–H bond promoted by dopamine β-hydroxylase (DBH) and its biomimetic models. We demonstrate that it is possible to carry out parallel and comparative studies on enzyme (DBH) and its biomimetic models with the same substrate: 2-aminoindane (1). It was chosen because its two stereogenic centers, both in benzylic positions, make it very powerful for studying the stereochemistry of an O-atom transfer reaction. DBH-catalyzed hydroxylation of 1 exclusively produced 14 % of trans-(1S,2S)-2-amino-1-indanol 4a (93 % e.e.). Studies with stereospecifically deuterium labeled 2-aminoindanes (1R,2S)-2 and (1S,2S)-3 showed that the formation of 4a was the result of an overall process with retention of configuration where an O-atom is stereospecifically inserted in the trans pro-S position of the substrate. Addition of 1 and (±)2 to 2-vinylpyridine gave 2-X-IndPY2 ligands 5 and 6, which were transformed into copper(I) and (II) complexes {7}(PF6) and {8}(CF3SO3), respectively. Reaction with dioxygen led to new complexes in which an O-atom transfer to a benzylic C–H bond has been performed in the same manner as that of DBH. With deuterium labeled cis-2-d-IndPY2 ligand 6, we demonstrated that the reaction occurs by a stereospecific process with retention of configuration. In both cases (enzymatic vs. biomimetic) the O-atom transfers occur in a two-step process involving radical intermediates.  相似文献   

20.
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