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 93 毫秒
1.
Luminescent lanthanide (III) ions have been exploited for circularly polarized luminescence (CPL) for decades. However, very few of these studies have involved chiral samarium (III) complexes. Complexes are prepared by mixing axial chiral ligands (R/S))‐2,2’‐bis(diphenylphosphoryl)‐1,1′‐binaphthyl (BINAPO) with europium and samarium Tris (trifluoromethane sulfonate) (Eu (OTf)3 and Sm (OTf)3). Luminescence‐based titration shows that the complex formed is Ln((R/S)‐BINAPO)2(OTf)3, where Ln = Eu or Sm. The CPL spectra are reported for Eu((R/S)‐BINAPO)2(OTf)3 and Sm((R/S)‐BINAPO)2(OTf)3. The sign of the dissymmetry factors, gem, was dependent upon the chirality of the BINAPO ligand, and the magnitudes were relatively large. Of all of the complexes in this study, Sm((S)‐BINAPO)2(OTf)3 has the largest gem = 0.272, which is one of the largest recorded for a chiral Sm3+ complex. A theoretical three‐dimensional structural model of the complex that is consistent with the experimental observations is developed and refined. This report also shows that (R/S)‐BINAPO are the only reported ligands where gem (Sm3+) > gem (Eu3+).  相似文献   

2.
《Chirality》2017,29(6):273-281
Enantiomeric 1H and 13C NMR signal separation behaviors of various α‐amino acids and DL‐tartarate were investigated by using the samarium(III) and neodymium(III) complexes with (S ,S )‐ethylenediamine‐N ,N' ‐disuccinate as chiral shift reagents. A relatively smaller concentration ratio of the lanthanide(III) complex to substrates was suitable for the neodymium(III) complex compared with the samarium(III) one, striking a balance between relatively greater signal separation and broadening. To clarify the difference in the signal separation behavior, the chemical shifts of β‐protons for fully bound D‐ and L‐alanine (δb(D) and δb(L)) and their adduct formation constants (K s) were obtained for both metal complexes. Preference for D‐alanine was similarly observed for both complexes, while it was revealed that the difference between the δb(D) and δb(L) values is the significant factor to determine the enantiomeric signal separation. The neodymium(III) and samarium(III) complexes can be used complementarily for higher and smaller concentration ranges of substrates, respectively, because the neodymium(III) complex gives the larger difference between the δb(D) and δb(L) values with greater signal broadening compared to the samarium(III) complex.  相似文献   

3.
The charged, electroactive bipyridine‐helicene‐ruthenium(III) complex [ 4 ] . +,PF6? has been prepared from 3‐(2‐pyridyl)‐4‐aza[6]helicene and a Ru‐bis‐(β‐diketonato)‐bis‐acetonitrile precursor (β‐diketonato: 2,2,6,6‐tetramethyl‐3,5‐heptanedionato). Its chiroptical properties (electronic circular dichroism and optical rotation) were studied both experimentally and theoretically and suggest the presence of 2 diastereoisomers, namely (P,Δ)‐ and (P,Λ)‐[ 4 ] . +,PF6? (denoted jointly as (P,Δ*)‐[ 4 ] . +,PF6?) and their mirror‐images (M,Λ)‐ and (M,Δ)‐[ 4 ] . +,PF6? ((M,Δ*)‐[ 4 ] . +,PF6?). The electrochemical reduction of (P,Δ*)‐[ 4 ] . +,PF6? to neutral complex (P,Δ*)‐ 4 was performed and revealed strong changes in the UV‐vis and electronic circular dichroism spectra. A reversible redox‐triggered chiroptical switching process was then achieved.  相似文献   

4.
A chiral spin crossover iron(II) complex, fac-Λ-[FeII(HLR)3](ClO4)2·EtOH was synthesized and its crystal structures in both the high-spin (HS) and low-spin (LS) states were determined, where HLR denotes 2-methylimidazol-4-yl-methylideneamino-R-(+)-1-methylphenyl. The complex assumes octahedral coordination geometry of N6 donor atoms by three bidentate ligands HLR. The complex exists as the facial-Λ-isomer of fac-Λ-[FeII(HLR)3]2+ of the possible geometrical fac- and mer-isomers and the Δ- and Λ-enantiomorphs. The X-ray structural analyses revealed that the R-form of the ligand (HLR) induces the fac-Λ-isomer of fac-Λ-[FeII(HLR)3]2+ and the S-form of the ligand (HLS) induces the fac-Δ-isomer of fac-Δ-[Fe(HLS)3]2+. The complex fac-Λ-[FeII(HLR)3](ClO4)2·EtOH shows a complete steep spin crossover between the HS and the LS states at T1/2 = 195 K.  相似文献   

5.
A new tetracopper(II) complex bridged both by oxamido and carboxylato groups, namely [Cu4(dmaepox)2(bpy)2](NO3)2·2H2O, where H3dmaepox and bpy represent N‐benzoato‐N′‐ (3‐methylaminopropyl)oxamide and 2,2′‐bipyridine, was synthesized, and its structure reveals the presence of a centrosymmetric cyclic tetracopper(II) cation assembled by a pair of cis‐dmaepox3–‐ bridged dicopper(II) units through the carboxylato groups, in which the endo‐ and exo‐copper(II) ions bridged by the oxamido group have a square‐planar and a square‐pyramidal coordination geometries, respectively. The aromatic packing interactions assemble the complex molecules to a two‐dimensional supramolecular structure. The reactivity toward DNA and protein bovine serum albumin (BSA) indicates that the complex can interact with herring sperm DNA through the intercalation mode and the binding affinity is dominated by the hydrophobicity and chelate ring arrangement around copper(II) ions and quenches the intrinsic fluorescence of BSA via a static process. The cytotoxicity of the complex shows selective cancer cell antiproliferative activity.  相似文献   

6.
A profound influence of water has previously been detected in the complexation of the enantiomers of methyl 2‐chloropropanoate (MCP) and the chiral selector octakis(3‐O‐butanoyl‐2,6‐di‐O‐pentyl)‐γ‐cyclodextrin (Lipodex‐E) in NMR and sensor experiments. We therefore investigated the retention behavior of MCP enantiomers on Lipodex‐E by gas chromatography (GC) under hydrous conditions. Addition of water to the N2 carrier gas modestly reduced the retention factors k of the enantiomers, notably for the second eluted enantiomer (S)‐MCP. This resulted in an overall decrease of enantioselectivity ‐ΔS,R(ΔG) in the presence of water. The effect was fully reversible. Consequently, for a conditioned column in the absence of residual water, the determined thermodynamic data, i.e. ΔS,R(ΔH) = –12.64 ± 0.08 kJ mol‐1 and ΔS,R(ΔS) = –28.18 ± 0.23 J K‐1 mol‐1, refer to a true 1:1 complexation process devoid of hydrophobic hydration. Chirality 28:124–131, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

7.
A determination method for Co(II), Fe(II) and Cr(III) ions by luminol‐H2O2 system using chelating reagents is presented. A metal ion‐chelating ligand complex with a Co(II) ion and a chelating reagent like ethylenediaminetetraacetic acid (EDTA) produced highly enhanced chemiluminescence (CL) intensity as well as longer lifetime in the luminol‐H2O2 system compared to metals that exist as free ions. Whereas free Cu(II) and Pb(II) ions had a strong catalytic effect on the luminol‐H2O2 system, significantly, the complexes of Cu(II) and Pb(II) with chelating reagents lost their catalytic activity due to the chelating reagents acting as masking agents. Based on the observed phenomenon, it was possible to determine Co(II), Fe(II) and Cr(III) ions with enhanced sensitivity and selectivity using the chelating reagents of the luminol‐H2O2 system. The effects of ligand, H2O2 concentration, pH, buffer solution and concentrations of chelating reagents on CL intensity of the luminol‐H2O2 system were investigated and optimized for the determination of Co(II), Fe(II) and Cr(III) ions. Under optimized conditions, the calibration curve of metal ions was linear over the range of 2.0 × 10‐8 to 2.0 × 10‐5 M for Co(II), 1.0 × 10‐7 to 2.0 × 10‐5 M for Fe (II) and 2.0 × 10‐7 to 1.0 × 10‐4 M for Cr(III). Limits of detection (3σ/s) were 1.2 × 10‐8, 4.0 × 10‐8 and 1.2 × 10‐7 M for Co(II), Fe(II) and Cr(III), respectively. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

8.
9.
A new dinuclear copper(II) complex bridged by N‐[3‐(dimethylamino)propyl]‐N′‐ (2‐carbo‐xylatophenyl)oxamide (H3dmapob), and endcapped with 2,2′‐diamino‐4,4′‐bithiazole (dabt), namely [Cu2(dmapob)(dabt)(CH3OH)(pic)]·(DMF)0.75·(CH3OH)0.25 has been synthesized and characterized by elemental analysis, molar conductivity measurement, infrared and electronic spectra studies, and single‐crystal X‐ray diffraction. In the crystal structure, both copper(II) ions have square–pyramidal coordination geometries. The Cu···Cu separation through the oxamido bridge is 5.176(9) Å. A two‐dimensional supramolecular framework is formed through hydrogen bonds and π–π stacking interactions. The reactivities toward herring sperm DNA and bovine serum albumin (BSA) show that the complex can interact with the DNA via intercalation mode and bind to the BSA responsible for quenching of tryptophan fluorescence by the static quenching mechanism. The in vitro anticancer activities suggest that the copper(II) complex is active against the selected tumor cell lines. The influence of different bridging ligands in dinuclear complexes on the DNA‐ and BSA‐binding properties as well as anticancer activities is preliminarily discussed.  相似文献   

10.
The first successful enantioseparation of representative O,O‐diphenyl‐N‐arylthioureidoalkylphosphonates, (±)‐Ptc‐ValP(OPh)2 & (±)‐Ptc‐LeuP(OPh)2 and thiourylenedi(isobutyl phosphonate), Tcm[ValP(OPh)2]2 on analytical and semipreparative scale was achieved by high‐performance liquid chromatography using polysaccharide‐based chiral stationary phases (CPs). Atc‐AAP(OPh)2 was obtained using modified tricomponent condensations of the corresponding aldehydes, N‐arylthiourea and triphenyl phosphite whereas Tcm[ValP(OPh)2]2 by the condensations of aldehydes, thiourea, and triphenyl phosphite. The prepared, racemic (±)‐Atc‐AAP(OPh)2 [(±)‐Ptc‐ValP(OPh)2, (±)‐Ptc‐LeuP(OPh)2, (±)‐Ptc‐PglyP(OPh)2 and (±)‐Ntc‐PglyP(OPh)2] and racemic (±)‐Tcm[AAP(OPh)2]2 [(±)‐Tcm[NvaP(OPh)2]2 & (±)‐Tcm[ValP(OPh)2]2] were adequately characterized and used for chromatographic separations on high‐performance liquid chromatography–chiral stationary phases. The best results were obtained for (±)‐Ptc‐ValP(OPh)2, (±)‐Ptc‐LeuP(OPh)2 and (±)‐Tcm[ValP(OPh)2]2.  相似文献   

11.
A new oxamido‐bridged bicopper(II) complex, [Cu2(pdpox)(bpy)(CH3OH)](ClO4), where H3pdpox and bpy stand for N‐(2‐hydroxyphenyl)‐N′‐[3‐(diethylamino)propyl]oxamide and 2,2′‐bipyridine, respectively, has been synthesized and characterized by elemental analyses, molar conductivity measurements, infrared and electronic spectra studies, and X‐ray single crystal diffraction. In the crystal structure, the pdpox3? ligand bridges two copper(II) ions as cisoid conformation. The inner copper(II) ion has a {N3O} square‐planar coordination geometry, while the exo‐ one is in a {N2O3} square‐pyramidal environment. There are two sets of interpenetrating two‐dimensional hydrogen bonding networks parallel to the planes (2 1 0) and (), respectively, to form a three‐dimensional supramolecular structure. The bicopper(II) complex exhibits cytotoxic activity against the SMMC7721 and A549 cell lines. The reactivity toward herring sperm DNA and bovine serum albumin revealed that the bicopper(II) complex can interact with the DNA by intercalation mode, and the complex binds to protein BSA responsible for quenching of tryptophan fluorescence by static quenching mechanism. © 2013 Wiley Periodicals, Inc. J BiochemMol Toxicol 27:412‐424, 2013; View this article online at wileyonlinelibrary.com . DOI 10.1002/jbt.21504  相似文献   

12.
A new trinickel(II) complex bridged by N‐[3‐(dimethylamino)propyl]‐ N ′‐(2‐hydroxylphenyl)oxamido (H3pdmapo), namely [Ni3(pdmapo)2(H2O)2]?4CH3OH, was synthesized and characterized by X‐ray single‐crystal diffraction and other methods. In the molecule, two symmetric cis‐ pdmapo3? mononickel(II) complexes as a “complex ligand” using the carbonyl oxygen atoms coordinate to the center nickel(II) ion situated on an inversion point. The Ni···Ni distance through the oxamido bridge is 5.2624(4) Å. The center nickel(II) ion and the lateral ones have octahedral and square‐planar coordination geometries, respectively. In the crystal, a three‐dimensional supramolecular network dominated by hydrogen bonds is observed. The reactivity toward DNA/protein bovine serum albumin (BSA) revealed that the complex could interact with herring sperm DNA (HS‐DNA) through the intercalation mode and quench the intrinsic fluorescence of BSA via a static mechanism. The in vitro anticancer activities suggested that the complex is active against the selected tumor cell lines.  相似文献   

13.
The S-bridged trinuclear complexes composed of heavy d6 metal ions, [RhIII{M(aet)3}2]3+ (M=IrIII(1), RhIII(2); aet = 2-aminoethanethiolate), have been prepared by the reactions of fac(S)-[M(aet)3] with RhCl3 · 3H2O. The complexes were separated into meso (1a, 2a) and rac (1b, 2b) isomers by SP-Sephadex C-25 column chromatography. 1b and 2b were optically resolved by the column chromatographic method and characterized by CD spectroscopy. Crystal structures of 1a, 1b and 2a were determined by X-ray diffraction, and it was found that they consist of linear-type trinuclear structures. The central Rh(III) ion in the present complexes has d6 electronic configuration with the non-degenerated A-type cubic field term, and showed long Rh?M distances, acute S-M-S angles and obtuse Rh-S-M angles. These are in contrast with the complexes having the degenerated T-type cubic field term such as [M{M(aet)3}2]n+ (M=VIII, MoIV and ReIII, M=IrIII, RhIII, n=3 or 4). All the isomers have been comparatively characterized and discussed in solid state and the solution for spectrochemical and electrochemical properties.  相似文献   

14.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
We describe herein the synthesis of (rac)‐ or enantiopure (S)‐(?)‐(2‐MeBu)N(Pr)2MeI ammonium salts. These racemic and enantiopure ammonium salts were used as cationic templates to obtain new two‐dimensional (2D) ferromagnets [(rac)‐(2‐MeBu)N(Pr)2Me][MnCr(C2O4)3] and [(S)‐(?)‐(2‐MeBu)N(Pr)2Me][ΔMnΛ nCr(C2O4)3]. The absolute configuration of the hexacoordinated Cr(III) metallic ion in the enantiopure 2D network was determined by a circular dichroism measurement. The structure of [(2‐MeBu)N(Pr)2Me][MnCr(C2O4)3], established by single crystal X‐ray diffraction, belongs to the chiral P63 space group. According to direct current (dc) magnetic measurements, these compounds are ferrromagnets with a temperature Tc = 6°K. Chirality 25:444–448, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

16.
[Ni(C11H9N2O5)2(H2O)2]?3(C3H7NO) ( 1 ) and [Co(C11H9N2O5)2(H2O)2]?3(C3H7NO) ( 2 ) are synthesized and characterized by elemental analysis, FT‐IR spectra, magnetic susceptibility, and thermal analysis. In addition, the crystal structure of Ni(II) complex is presented. Both complexes show distorted octahedral geometry. In 1 and 2, metal ions are coordinated by two oxygen atoms of salicylic residue and two nitrogen atoms of maleic amide residue from two ligands, and two oxygen atoms from two water molecules. In this paper, both compounds showed excellent inhibitory effects against human carbonic anhydrase (hCA) isoforms I, and II, α‐glycosidase, acetylcholinesterase (AChE), and butyrylcholinesterase (BChE). Compounds 1 and 2 had Ki values of 18.36 ± 4.38 and 26.61 ± 7.54 nM against hCA I and 13.81 ± 3.02 and 29.56 ± 6.52 nM against hCA II, respectively. On the other hand, their Ki values were found to be 487.45 ± 54.18 and 453.81 ± 118.61 nM against AChE and 199.21 ± 50.35 and 409.41 ± 6.86 nM against BChE, respectively.  相似文献   

17.
A novel ternary complex, Tb2L4·L′·(ClO4)6·8H2O, has been synthesized using bis(benzylsulfinyl)methane as the first ligand L and 2,2′‐dipyridyl as the second ligand L′. The ternary complex was characterized by element analysis, molar conductivity, coordination titration analysis, infrared, thermogravimetric‐differential scanning calorimetric and ultraviolet spectra. The results indicated that the composition of the complex was Tb2L4·L′·(ClO4)6·8H2O (L = C6H5CH2SOCH2SOCH2C6H5; L′ = Dipy). Fourier transform infrared results revealed that the perchlorate group was bonded with the Tb(III) ion by the oxygen atom, and the coordination was bidentate. The fluorescent spectra illustrated that the complex displayed characteristic fluorescence in the solid state. After the introduction of the second ligand, 2,2‐dipyridyl, the relative emission intensity and fluorescence lifetime of the ternary complex Tb2L4·L′·(ClO4)6·8H2O were enhanced compared to the binary complex TbL2.5(ClO4)3·3H2O. This indicated that the presence of both organic ligand bis(benzylsulfinyl)methane and the second ligand 2,2‐dipyridyl could sensitize the fluorescence intensity of Tb(III) ion, and introduction of the 2,2‐dipyridyl group resulted in an enhancement of the fluorescence of the Tb(III) ternary rare earth complex. The strongest characteristic fluorescence emission intensity of the ternary complex was 9.36 times that of the binary complex. The phosphorescence spectra and fluorescence lifetime of the complex were also measured. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

18.
A novel ligand, 1‐(naphthalen‐2‐yl)‐2‐(phenylsulthio)ethanone was synthesized using a new method and its two europium (Eu) (III) complexes were synthesized. The compounds were characterized by elemental analysis, coordination titration analysis, molar conductivity, infrared, thermo gravimetric analyzer‐differential scanning calorimetry (TGA‐DSC), 1H NMR and UV spectra. The composition was suggested as EuL5 · (ClO4)3 · 2H2O and EuL4 · phen(ClO4)3 · 2H2O (L = C10H7COCH2SOC6H5). The fluorescence spectra showed that the Eu(III) displayed strong characteristic metal‐centered fluorescence in the solid state. The ternary rare earth complex showed stronger fluorescence intensity than the binary rare earth complex in such material. The strongest characteristic fluorescence emission intensity of the ternary system was 1.49 times as strong as that of the binary system. The phosphorescence spectra were also discussed. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

19.
Chemiluminescence (CL) of the rhodamine 6‐G‐diperiodatonickelate (IV) (Rh6‐G‐Ni(IV) complex) in the presence of Brij‐35 was examined in an alkaline medium and implemented using flow‐injection analysis to analyze Mn(II) in natural waters. Brij‐35 was identified as the surfactant of choice that enhanced CL intensity by about 62% of the reaction. The calibration curves were linear in the range 1.7 × 10?3 – 0.2 (0.9990, n = 7) and 8.0 × 10?4 – 0.1 μg ml?1 (0.9990, n = 7) with limits of detection (LODs) (S:N = 3) of 5.0 × 10?4 and 2.4 × 10?4 μg ml?1 without and with using an in‐line 8‐hydroxyquinoline (8‐HQ) resin mini‐column, respectively. The sample throughput and relative standard deviation were 200 h?1 and 1.7–2.2% in the range studied respectively. Mn(II) concentrations in certified reference materials and natural water samples was successfully determined. A brief discussion about the possible CL reaction mechanism is also given. In addition, analysis of V(III), Cr(III) and Fe(II) was also performed without and with using an in‐line 8–HQ column and selective elution of each metal ion was achieved by adjusting the pH of the sample carrier stream with aqueous HCl solution.  相似文献   

20.
One chiral L ‐valine (L ‐Val) was inserted into the C‐terminal position of achiral peptide segments constructed from α‐aminoisobutyric acid (Aib) and α,β‐dehydrophenylalanine (ΔZPhe) residues. The IR, 1H NMR and CD spectra indicated that the dominant conformations of the pentapeptide Boc‐Aib‐ΔPhe‐(Aib)2‐L ‐Val‐NH‐Bn (3) and the hexapeptide Boc‐Aib‐ΔPhe‐(Aib)3‐L ‐Val‐NH‐Bn (4) in solution were both right‐handed (P) 310‐helical structures. X‐ray crystallographic analyses of 3 and 4 revealed that only a right‐handed (P) 310‐helical structure was present in their crystalline states. The conformation of 4 was also studied by molecular‐mechanics calculations. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  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