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 15 毫秒
1.
Four chiral polymers P-1, P-2, P-3 and P-4 were synthesized by the polymerization of (S)-2,2'-dioctoxy-1,1'- binaphthyl-6,6'-boronic acid (S-M-3) with (S)-6,6'-dibromo-1,1'-binaphthol (S-M-1), (R)-6,6'-dibromo-1,1'- binaphthol (R-M-1), (S)-3,3'-diiodo-1,1'-binaphthol (S-M-2) and (R)-3,3'-diiodo-1,1'-binaphthol (R-M-2) under Pd-catalyzed Suzuki reaction, respectively. All four polymers can show good solubility in some common solvents due to the nonplanarity of the polymers in the main chain backbone and flexible alkyl groups in the side chain. The analysis results indicate that specific rotation and circular dichroism (CD) spectral signals of the alternative S-S chiral polymers P-1 and P-3 are larger than those of S-R chiral polymers P-2 and P-4, but their UV-Vis and fluorescence spectra are almost similar. The results of asymmetric enantioselectivity of four polymers for diethylzinc addition to benzaldehyde indicate that catalytically active center is (R) or (S)-1, 1'-binaphthol moieties.  相似文献   

2.
This paper describes the development of novel aromatic platforms for supramolecular construction. By the Suzuki cross‐coupling protocol, a variety of functionalized m‐terphenyl derivatives were prepared (Schemes 1–4). Macrolactamization of bis(ammonium salt) (S,S)‐ 6 with bis(acyl halide) 7 afforded the macrocyclic receptor (S,S)‐ 2 (Scheme 1), which was shown by 1H‐NMR titration studies to form ‘nesting' complexes of moderate stability (Ka between 130 and 290 M ?1, 300 K) with octyl glucosides 13 – 15 (Fig. 2) in the noncompetitive solvent CDCl3. Suzuki cross‐coupling starting from 3,3′,5,5′‐tetrabromo‐1,1′‐biphenyl provided access to a novel series of extended aromatic platforms (Scheme 5) for cleft‐type (Fig. 1) and macrotricyclic receptors such as (S,S,S,S)‐ 1 . Although mass‐spectral evidence for the formation of (S,S,S,S)‐ 1 by macrolactamization between the two functionalized 3,3′,5,5′‐tetraaryl‐1,1′‐biphenyl derivatives (S,S)‐ 33 and 36 was obtained, the 1H‐ and 13C‐NMR spectra of purified material remained rather inconclusive with respect to both purity and constitution. The versatile access to the novel, differentially functionalized 3,3′,5,5′‐tetrabromo‐1,1′‐biphenyl platforms should ensure their wide use in future supramolecular construction.  相似文献   

3.
Vibrational‐circular‐dichroism (VCD) studies of the solution conformations of three 2,2′‐substituted biphenyls are reported. Biphenyls with only two substituents at the peri‐position normally show rotation about their central axis of chirality at room temperature in solution. We previously found no evidence for rotation of (P,4S)‐2‐[4,5‐dihydro‐4‐(1‐methylethyl)oxazol‐2‐yl]‐2′‐(hydroxymethyl)‐1,1′‐biphenyl ( 1 ) in CDCl3 about its 1,1′‐ axis, due to stabilization by intramolecular H‐bonding and the presence of the i‐Pr substituent, but two conformers were found in solution that result from rotation of the heterocycle in 1 between OH⋅⋅⋅N (the form present in the solid state) and the OH⋅⋅⋅O H‐bonded forms, with no rotation of the i‐Pr group or (P) → (M) twist [1]. For (P,S)‐ 2 , where the i‐Pr substituent of 1 has been replaced by a Ph group, rotation of the heterocycle takes place in CDCl3 solution. For (P,S)‐ 3 and (M,S)‐ 4 , where Me substitution at the two 6,6′‐positions of (P,S)‐ 1 prevents rotation about the central axis of chirality, rotation of the heterocycle is observed for the (P)‐configuration ((P,S)‐ 3 ), but not for the (M)‐configuration ((M,S)‐ 4 ). Only one rotamer involving the i‐Pr group, which was found in the solid state of (P,S)‐ 3 , was also observed in solution, but (M,S)‐ 4 , obtained as an oil, was found to be a mixture of three rotamers of the i‐Pr group.  相似文献   

4.
1,1′‐Biphenyl derivatives with amino acid/peptide substitution at C(2) and C(2′) (‘peptide‐biphenyl hybrids', 6 – 8 ) have been prepared by direct N‐acylation of amino acid/peptide derivatives with 1,1′‐biphenyl‐2,2′‐dicarbonyl dichloride ( 5 ). Both conformers, which arise from the rotation around the aryl aryl bond, have been detected by 1H‐NMR spectroscopy. Single atropisomers of each 6 ((R)‐configuration at the stereogenic axis) and 7 ((S)‐configuration at the stereogenic axis) have been obtained in quantitative yield by slow evaporation of methanolic solutions. The procedures are dynamic atropselective resolutions (asymmetric transformations of the second kind). The crystal structures of the peptide‐biphenyl hybrids 6 and 7 show highly ordered molecular and supramolecular structures with extensive intramolecular and intermolecular H‐bonding.  相似文献   

5.
In this paper,a series of new optically active MeO-BIPHEP-type ligands,(S)-6,6'-dimethoxy-2,2'-bis(di-p-alkoxyphenyl-phosphine)-1,1'-biphenyl[(S)-5b—(S)-5e]were prepared and characterized.Starting from thecommercially available triethyl phosphorite and m-bromoanisole,an optically active(S)-6,6'-dimethoxybiphenyl-2,2'-diyl-bis(phosphonic acid diester)was prepared by an improved way and converted to the corresponding dichlo-rides,which was used as a key intermediate to react with p-alkoxybenzenemagnesium bromide or p-alkoxyphenyllithium to directly give the enantiomerically pure diphosphines 5.  相似文献   

6.
Two chiral phosphoramidates,(R)-(-)-1,1'-binaphthyl-2,2'-dihydroxy-N-[α-(S)-methylbenzyl] phosphoramidate and (-)-1,1'-biphenyl-2,2'-dihydroxy-N-[α-(S)-methylbenzyl]-phosphoramidate were synthesized.Their crystal structures were determined by X-ray single crystal diffraction analysis.The phosphoramidate molecules are self-associated by inter-molecular N-H...O = P hydrogen bonds and aromatic edge to face interactions.  相似文献   

7.
Biphenyls with only two substituents at the ‘peri'‐position normally show rotation about their chiral axis at room temperature. Using vibrational circular dichroism (VCD), we found no evidence for rotation of (P)‐2′‐[(4S)‐4,5‐dihydro‐4‐(1‐methylethyl)oxazol‐2‐yl][1,1′‐biphenyl]‐2‐methanol ((P,S)‐ 1 ) in CDCl3 about its chiral axis due to stabilization by intramolecular H‐bonding. All rotamers of 1 were calculated at the DFT level, and, from these optimized structures, the VCD spectra were calculated and compared to the measured VCD spectra. The best agreement between calculated and measured spectra is obtained when two rotamers are present in solution. These rotamers differ primarily in their intramolecular H‐bonding interactions, having either OH???N (the form present in the solid state) or OH???O H‐bonds, i.e., a rotation of the heterocycle in 1 takes place in solution.  相似文献   

8.
A new family of optically active cyclophane receptors for the complexation of mono‐ and disaccharides in competitive protic solvent mixtures is described. Macrocycles (−)‐(R,R,R,R)‐ 1 – 4 feature preorganized binding cavities formed by four 1,1′‐binaphthalene‐2,2′‐diyl phosphate moieties bridged in the 3,3′‐positions by acetylenic or phenylacetylenic spacers. The four phosphodiester groups converge towards the binding cavity and provide efficient bidentate ionic H‐bond acceptor sites (Fig. 2). Benzyloxy groups in the 7,7′‐positions of the 1,1′‐binaphthalene moieties ensure solubility of the nanometer‐sized receptors and prevent undesirable aggregation. The construction of the macrocyclic framework of the four cyclophanes takes advantage of Pd0‐catalyzed aryl—acetylene cross‐coupling by the Sonogashira protocol, and oxidative acetylenic homo‐coupling methodology (Schemes 2 and 8 – 10). Several cleft‐type receptors featuring one 1,1′‐binaphthalene‐2,2′‐diyl phosphate moiety were also prepared (Schemes 1, 6, and 7). An undesired side reaction encountered during the synthesis of the target compounds was the formation of naptho[b]furan rings from 3‐ethynylnaphthalene‐2‐ol derivatives, proceeding via 5‐endo‐dig cyclization (Schemes 35). Computer‐assisted molecular modeling indicated that the macrocycles prefer nonplanar puckered, cyclobutane‐type conformations (Figs. 7 and 8). According to these calculations, receptor (−)‐(R,R,R,R)‐ 1 has, on average, a square binding site, which is complementary in size to one monosaccharide. The three other cyclophanes (−)‐(R,R,R,R)‐ 2 – 4 feature, on average, wider rectangular cavities, providing a good fit to one disaccharide, while being too large for the complexation of one monosaccharide. This substrate selectivity was fully confirmed in 1H‐NMR binding titrations. The chiroptical properties of the cyclophanes and their nonmacrocyclic precursors were investigated by circular dichroism (CD) spectroscopy. The CD spectra of the acyclic precursors showed a large dependence from the number of 1,1′‐binaphthalene moieties (Fig. 9), and those of the cyclophanes were remarkably influenced by the nature of the functional groups lining the macrocyclic cavity (Fig. 11). Profound differences were also observed between the CD spectra of linear and macrocyclic tetrakis(1,1′‐binaphthalene) scaffolds, which feature very different molecular shapes (Fig. 10). In 1H‐NMR binding titrations with mono‐ and disaccharides (Fig. 13), concentration ranges were chosen to favor 1 : 1 host−guest binding. This stoichiometry was experimentally established by the curve‐fitting analysis of the titration data and by Job plots. The titration data demonstrate conclusively that the strength of carbohydrate recognition is enhanced with an increasing number of bidentate ionic host−guest H‐bonds (Table 1) in the complex formed. As a result of the formation of these highly stable H‐bonds, carbohydrate complexation in competitive protic solvent mixtures becomes more favorable. Thus, cleft‐type receptors (−)‐(R)‐ 7 and (−)‐(R)‐ 38 with one phosphodiester moiety form weak 1 : 1 complexes only in CD3CN. In contrast, macrocycle (−)‐(R,R,R,R)‐ 1 with four phosphodiester groups undergoes stable inclusion complexation with monosaccharides in CD3CN containing 2% CD3OD. With their larger number of H‐bonding sites, disaccharide substrates bind even more strongly to the four phosphodiester groups lining the cavity of (−)‐(R,R,R,R)‐ 2 and complexation becomes efficient in CD3CN containing 12% CD3OD. Finally, the introduction of two additional methyl ester residues further enhances the receptor capacity of (−)‐(R,R,R,R)‐ 3 , and efficient disaccharide complexation occurs already in CD3CN containing 20% CD3OD.  相似文献   

9.
Photochromic 6‐bromomethyl‐6′‐methyl‐[2,2′‐bi‐1H‐indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 2 ), 6,6′‐ bis(bromomethyl)‐[2,2′‐bi‐1H‐indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 3 ) and 6,6′‐bis(dibromomethyl)‐[2,2′‐ bi‐1H‐indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 4 ) have been synthesized from 6,6′‐dimethyl‐[2,2′‐bi‐1H‐ indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 1 ). The single crystal of 4 was obtained and its crystal structure was analyzed. The results indicate that in crystal 4 , molecular arrangement is defective tightness compared with its precursor 1 . Besides, UV‐Vis absorption spectra in CH2Cl2 solution, photochromic and photomagnetic properties in solid state of 2 , 3 and 4 were also investigated. The results demonstrate that when the hydrogen atoms in the methyl group on the benzene rings of biindenylidenedione were substituted by bromines, its properties could be affected considerably.  相似文献   

10.
The (3R,5′R,6′R)‐ and (3R,5′R,6′S)‐capsanthol‐3′‐one (=3,6′‐dihydroxy‐β,κ‐caroten‐3′‐one; 4 and 5 , resp.) were reduced by different complex metal hydrides containing organic ligands. The ratio of the thus obtained diastereoisomeric (3′S)‐capsanthols 2 and 3 or (3′R)‐capsanthols 6 and 7 , respectively, was investigated. Four complex hydrides showed remarkable stereoselectivity and produced the (3′R,6′S)‐capsanthol ( 6 ) in 80 – 100% (see Table 1). The starting materials and the products were characterized by UV/VIS, CD, 1H‐ and 13C‐NMR, and mass spectra.  相似文献   

11.
By introducing slight structural modifications to a D4‐symmetric coordination capsule, we succeeded in isolating the nearly enantiopure capsules (P)‐ and (M)‐ 2 a (BF4)4. Chiral guest, dibenzyl 4,4′‐diacetoxy‐6,6′‐dimethyl‐[1,1′‐biphenyl]‐2,2′‐dicarboxylate ( 3 ) was encapsulated within the dissymmetric cavity of 2 a (BF4)4, resulting in a high diastereoselectivity of >99 % de. The encapsulated guest was successfully removed from the complex without racemization through precipitation of the empty capsule. CD spectra confirmed that the chirality of the capsule was maintained in THF and 1,4‐dioxane for long periods, whereas a small amount of acetonitrile accelerated racemization of the empty capsule. The activation parameters of the racemization reaction were determined in dichloromethane and 1,2‐dichloroethane, resulting in positive enthalpic contributions and large negative entropic contributions, respectively. Accordingly, the racemization fits a first‐order kinetic model. Mechanically coupled Cu+‐2,2′‐bipyridine coordination centers were responsible for the high‐energy barrier of racemization and led to the unique chiral memory of the dissymmetric cavity, which was turned off by the addition of acetonitrile.  相似文献   

12.
New chiral binaphthyl‐based polyarylenes [(S)‐ 3a and (S)‐ 3b ] with appendant Fréchet‐type poly(aryl ether) dendrons (first generation and second generation) were synthesized with Suzuki polycondensation from chiral (S)‐6,6′‐dibromo‐2,2′‐didendron‐substituted 1,1′‐binaphthyl derivatives and p‐phenylene diboronic acid. The polymers were studied with circular dichroism, fluorescence, and ultraviolet–visible spectra. Laser light scattering measurements of (S)‐ 3a and (S)‐ 3b showed that their weight‐average molecular weights were 2.39 × 105 and 1.09 × 104, respectively. The specific optical rotation [α]D was ?59.6 for (S)‐ 3a and ?62.7 for (S)‐ 3b . These dendronized conjugated polymers exhibited good thermal stability. The glass‐transition temperatures and the initial decomposition temperatures were 187.5 and 265.3 °C for (S)‐ 3a and 173.8 and 308.9 °C for (S)‐ 3b , respectively. (S)‐ 3a and (S)‐ 3b had high fluorescence quantum efficiencies, 87 and 91%, respectively, in tetrahydrofuran. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 1167–1172, 2002  相似文献   

13.
The 2,2′‐methylenebis[furan] ( 1 ) was converted to 1‐{(4R,6S))‐6‐[(2R)‐2,4‐dihydroxybutyl]‐2,2‐dimethyl‐1,3‐dioxan‐4‐yl}‐3‐[(2R,4R)‐tetrahydro‐4,6‐dihydroxy‐2H‐pyran‐2‐yl)propan‐2‐one ((+)‐ 18 ) and its (4S)‐epimer (?)‐ 19 with high stereo‐ and enantioselectivity (Schemes 13). Under acidic methanolysis, (+)‐ 18 yielded a single spiroketal, (3R)‐4‐{(1R,3S,4′R,5R,6′S,7R)‐3′,4′,5′,6′‐tetrahydro‐4′‐hydroxy‐7‐methoxyspiro[2,6‐dioxabicyclo[3.3.1]nonane‐3,2′‐[2H]pyran]‐6′‐yl}butane‐1,3‐diol ((?)‐ 20 ), in which both O‐atoms at the spiro center reside in equatorial positions, this being due to the tricyclic nature of (?)‐ 20 (methyl pyranoside formation). Compound (?)‐ 19 was converted similarly into the (4′S)‐epimeric tricyclic spiroketal (?)‐ 21 that also adopts a similar (3S)‐configuration and conformation. Spiroketals (?)‐ 20 , (?)‐ 21 and analog (?)‐ 23 , i.e., (1R,3S,4′R,5R,6′R)‐3′,4′,5′,6′‐tetrahydro‐6′‐[(2S)‐2‐hydroxybut‐3‐enyl]‐7‐methoxyspiro[2,6‐dioxabicyclo[3.3.1]nonane‐3,2′‐[2H]pyran]‐4′‐ol, derived from (?)‐ 20 , were assayed for their cytotoxicity toward murine P388 lymphocytic leukemia and six human cancer cell lines. Only racemic (±)‐ 21 showed evidence of cancer‐cell‐growth inhibition (P388, ED50: 6.9 μg/ml).  相似文献   

14.
The synthesis of novel 2,2‐disubstituted 2H‐azirin‐3‐amines with a chiral amino group is described. Chromatographic separation of the diastereoisomer mixture yielded the pure diastereoisomers (1′R,2R)‐ 4a – e and (1′R,2S)‐ 4a – e (Scheme 1, Table 1), which are synthons for the (R)‐ and (S)‐isomers of isovaline, 2‐methylvaline, 2‐cyclopentylalanine, 2‐methylleucine, and 2‐(methyl)phenylalanine, respectively. The configuration at C(2) of the synthons was determined by X‐ray crystallography relative to the known configuration of the chiral auxiliary group. The reaction of 4 with thiobenzoic acid, benzoic acid, and the dipeptide Z‐Leu‐Aib‐OH ( 12 ) yielded the monothiodiamides 10 , the diamides 11 (Scheme 2, Table 3), and the tripeptides 13 (Scheme 3, Table 4), respectively.  相似文献   

15.
((?)‐Menthyl (S)‐6′‐acrylyl‐2′‐methyloxy‐1,1′‐binaphthalene‐2‐carboxylate ( 3 ) was synthesized and anionically polymerized using n‐BuLi as an initiator in toluene. The monomer 3 was levorotatory and had an [α]D25 value of ?72.4, but its corresponding polymer poly‐ 3 was dextrorotatory and showed an [α]D25 value of +162.0. Poly‐ 3 was confirmed to exist in the form of one‐handed helical structure in solution by means of comparing the specific optical rotation and the CD spectra with that of 3 and the model compounds such as (?)‐menthyl (S)‐6′‐propionyl‐2′‐methyloxy‐1,1′‐binaphthalene‐2‐carboxylate 2b and (?)‐menthyl (S)‐6′‐heptanoyl‐2′‐methyloxy‐1,1′‐binaphthalene‐2‐carboxylate 2c . This conclusion was also confirmed by the fact that the g‐value of poly‐ 3 is about 11 times of that of monomer 3 .  相似文献   

16.
The title compound and its corresponding etheno‐ and ethano‐bridged compounds were successfully synthesized in enantiomerically pure form by McMurry coupling of 2,2′‐dimethoxy‐(R)‐ or ‐(S)‐1,1′‐binaphthyl‐6,6′‐dicarbaldehydes as the key reaction. The reaction proceeded in a highly diastereoselective manner; the reaction of the racemic dialdehyde did not afford the meso coupling product but gave only the racemic one in poor yield. The diyne crystallized in the chiral monoclinic space group P21 from toluene/hexane. Structural analysis reveals that it has a considerably twisted double‐helical structure in crystal form. The spectral properties (NMR, UV/Vis, and CD) clearly indicate the highly strained nature of the molecule. In particular, its UV/Vis and CD spectra exhibit a bathochromic shift of about 20 nm for the naphthyl π–π* transitions.  相似文献   

17.
Multiply enantioselective enzyme-catalyzed transacylation reactions are described. Two instances of triply enantioselective enzyme-catalyzed transacylations are 1) the reaction of rac-1-indanol with rac-1,1′-bi-2-naphthy]-2,2′-dibutyrate to afford (S)-1-indanoL (R)-1-indanylacetate, (S)-1,1′-bi-2-naphthyl-2,2′-diol, and (R)-1,1′-bi-2-naphthyl-2,2′-dibutyrate and 2) the reaction of rac-1-indanol with rac-2,2′-bis(butyroxymethyl)biphenyl to afford (S)-1-indanol, (R)-1-indanylbutyrate, (S)-2,2′-biphenyldimethanol, and (R)-2,2′-bis(butyroxy-methyl)biphenyl. Doubly enantioselective enzyme-catalyzed transacylations are described according to two instances: 1) the reaction of rac-1-indanol with rac-1,1′-bi-2-naphthyl-2-ol-2′-butyrate afforded (S)-1-indanol, (R)-1-indanylacetate, (S)-1,1′-bi-2-naphthyl-2,2′-diol, and (R)-1,1′-bi-2-naphthyl-2-ol-2′-butyrate, and 2) the reaction of rac-1-indanol with 1,3,5-O-methylidne-2,4,6-tri-O-butyrate-myo-inositol to afford (S)-1 -indanol, (R)-l-indanylbutyrate, and 1,3,5-O-methylidne-2,6-di-O-butyrate-myo-inositol. Multiply enantioselective enzyme-catalyzed reactions have a merit of the enhancement of enantiomeric excess over singly enantioselective ones.  相似文献   

18.
Diaquapalladium(2+) trifluoromethanesulfonates modified with (4R,4′S)‐ or (4S,4′S)‐2,2′‐bis(4‐benzyl‐4,5‐dihydrooxazole) (Cs‐ and C2‐ligands) produce isotactic poly(1‐oxo‐2‐phenylpropane‐1,3‐diyl) through copolymerization of styrene with carbon monoxide. However, the same meso‐catalyst in the presence of the free ligand leads to prevailingly syndiotactic growth of the copolymer, whereas the optically active catalyst, when used in the presence of the free enantiomeric ligand, gives an atactic copolymer.  相似文献   

19.
The planar 3,4‐diphosphinidenecyclobutene (DPCB) can be remarkably twisted into a C2‐type helical structure by dual coordination of a AuCl moiety. A prompt chirality control of the twisted DPCB skeleton ligated by the digold units affords the enantiopure structure by exchanging the chloride ligands for chiral [1,1′‐binaphthalene]‐2,2′‐dicarboxylate. The chirality of the diaurated 2,2′‐bis(diphenylphosphanyl)‐1,1′‐biphenyl (BIPHEP) system can be controlled prior to that of DPCB. Mixing of a DPCB‐bis(chlorogold) complex with the chiral silver salt dynamically leads to a single diastereomer, which was characterized by the 31P NMR spectrum and the CD couplet patterns in the visible (DPCB) area. The absolute configuration of the singly induced helical structure was assigned by the theoretical CD spectra determined by TD‐DFT calculations. Intramolecular alkoxycyclization of hexa‐4,5‐dien‐1‐ol catalyzed by the asymmetric DPCB‐digold structure were also attempted.  相似文献   

20.
Partially acetylated carotenoids were prepared from fully acetylated carotenoids by reaction with NaBH4, and were characterized by UV/VIS, CD, 1H‐NMR and mass spectra. The 3,6′‐diacetate, 3′,6′‐diacetate, and 6′‐acetate 10 – 12 , respectively, of (6′R)‐capsanthol (=(3R,3′S,5′R,6′R)‐β,κ‐carotene‐3,3′,6′‐triol; 4 ) were obtained from (6′R)‐capsanthol‐3,3′,6′‐triacetate ( 9 ), and the 3‐ and 3′‐acetates 13 and 14 , respectively, of 4 from (6′R)‐capsanthol 3,3′‐diacetate ( 8 ). The utility of this method was also demonstrated by the preparation of zeaxanthin and lutein monoacetates 16, 19 , and 20 .  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