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 62 毫秒
1.
Four chiral OsII arene anticancer complexes have been isolated by fractional crystallization. The two iodido complexes, (SOs,SC)‐[Os(η6p‐cym)(ImpyMe)I]PF6 (complex 2 , (S)‐ImpyMe: N‐(2‐pyridylmethylene)‐(S)‐1‐phenylethylamine) and (ROs,RC)‐[Os(η6p‐cym)(ImpyMe)I]PF6 (complex 4 , (R)‐ImpyMe: N‐(2‐pyridylmethylene)‐(R)‐1‐phenylethylamine), showed higher anticancer activity (lower IC50 values) towards A2780 human ovarian cancer cells than cisplatin and were more active than the two chlorido derivatives, (SOs,SC)[Os(η6p‐cym)(ImpyMe)Cl]PF6, 1 , and (ROs,RC)‐[Os(η6p‐cym)(ImpyMe)Cl]PF6, 3 . The two iodido complexes were evaluated in the National Cancer Institute 60‐cell‐line screen, by using the COMPARE algorithm. This showed that the two potent iodido complexes, 2 (NSC: D‐758116/1) and 4 (NSC: D‐758118/1), share surprisingly similar cancer cell selectivity patterns with the anti‐microtubule drug, vinblastine sulfate. However, no direct effect on tubulin polymerization was found for 2 and 4 , an observation that appears to indicate a novel mechanism of action. In addition, complexes 2 and 4 demonstrated potential as transfer‐hydrogenation catalysts for imine reduction.  相似文献   

2.
Two tetraphenylethylene (TPE) bridged tetraimidazolium salts, [H4 L ‐Et](PF6)4 and [H4 L ‐Bu](PF6)4, were used as precursors for the synthesis of the dinuclear AgI and AuI tetracarbene complexes [Ag2( L ‐Et)](PF6)2, [Ag2( L ‐Bu)](PF6)2, [Au2( L ‐Et)](PF6)2, and [Au2( L ‐Bu)](PF6)2. The tetraimidazolium salts show almost no fluorescence (Φ F<1 %) in dilute solution while their NHC complexes display fluorescence “turn‐on” (Φ F up to 47 %). This can be ascribed to rigidification mediated by the restriction of intramolecular rotation within the TPE moiety upon complexation. DFT calculations confirm that the metals are not involved in the lowest excited singlet and triplet states, thus explaining the lack of phosphorescence and fast intersystem crossing as a result of heavy atom effects. The rigidification upon complexation for fluorescence turn‐on constitutes an alternative to the known aggregation‐induced emission (AIE).  相似文献   

3.
The first heterodinuclear ruthenium(II) complexes of the 1,6,7,12‐tetraazaperylene (tape) bridging ligand with iron(II), cobalt(II), and nickel(II) were synthesized and characterized. The metal coordination sphere in this complexes is filled by the tetradentate N,N′‐dimethyl‐2,11‐diaza[3.3](2,6)‐pyridinophane (L‐N4Me2) ligand, yielding complexes of the general formula [(L‐N4Me2)Ru(µ‐tape)M(L‐N4Me2)](ClO4)2(PF6)2 with M = Fe {[ 2 ](ClO4)2(PF6)2}, Co {[ 3 ](ClO4)2(PF6)2}, and Ni {[ 4 ](ClO4)2(PF6)2}. Furthermore, the heterodinuclear tape ruthenium(II) complexes with palladium(II)‐ and platinum(II)‐dichloride [(bpy)2Ru(μ‐tape)PdCl2](PF6)2 {[ 5 ](PF6)2} and [(dmbpy)2Ru(μ‐tape)PtCl2](PF6)2 {[ 6 ](PF6)2}, respectively were also prepared. The molecular structures of the complex cations [ 2 ]4+ and [ 4 ]4+ were discussed on the basis of the X‐ray structures of [ 2 ](ClO4)4 · MeCN and [ 4 ](ClO4)4 · MeCN. The electrochemical behavior and the UV/Vis absorption spectra of the heterodinuclear tape ruthenium(II) complexes were explored and compared with the data of the analogous mono‐ and homodinuclear ruthenium(II) complexes of the tape bridging ligand.  相似文献   

4.
Chloride abstraction from the half‐sandwich complexes [RuCl2(η6p‐cymene)(P*‐κP)] ( 2a : P* = (Sa,R,R)‐ 1a = (1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl bis[(1R)‐1‐phenylethyl)]phosphoramidite; 2b : P* = (Sa,R,R)‐ 1b = (1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl bis[(1R)‐(1‐(1‐naphthalen‐1‐yl)ethyl]phosphoramidite) with (Et3O)[PF6] or Tl[PF6] gives the cationic, 18‐electron complexes dichloro(η6p‐cymene){(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl {(1R)‐1‐[(1,2‐η)‐phenyl]ethyl}[(1R)‐1‐phenylethyl]phosphoramidite‐κP}ruthenium(II) hexafluorophosphate ( 3a ) and [Ru(S)]‐dichloro(η6p‐cymene){(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl {(1R)‐1‐[(1,2‐η)‐naphthalen‐1‐yl]ethyl}[(1R)‐1‐(naphthalen‐1‐yl)ethyl]phosphoramidite‐κP)ruthenium(II) hexafluorophosphate ( 3b ), which feature the η2‐coordination of one aryl substituent of the phosphoramidite ligand, as indicated by 1H‐, 13C‐, and 31P‐NMR spectroscopy and confirmed by an X‐ray study of 3b . Additionally, the dissociation of p‐cymene from 2a and 3a gives dichloro{(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl [(1R)‐(1‐(η6‐phenyl)ethyl][(1R)‐1‐phenylethyl]phosphoramidite‐κP)ruthenium(II) ( 4a ) and di‐μ‐chlorobis{(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl [(1R)‐1‐(η6‐phenyl)ethyl][(1R)‐1‐phenylethyl]phosphoramidite‐κP}diruthenium(II) bis(hexafluorophosphate) ( 5a ), respectively, in which one phenyl group of the N‐substituents is η6‐coordinated to the Ru‐center. Complexes 3a and 3b catalyze the asymmetric cyclopropanation of α‐methylstyrene with ethyl diazoacetate with up to 86 and 87% ee for the cis‐ and the trans‐isomers, respectively.  相似文献   

5.
The protonation and ZnII/CuII complexation constants of tripodal polyamine ligand N1‐(2‐aminoethyl)‐N1‐(1H‐imidazol‐4‐ylmethyl)‐ethane‐1,2‐diamine (HL) were determined by potentiometric titration. Three new compounds, i.e. [H3(HL)](ClO4)3 ( 5 ), [Zn(HL)Cl](ClO4) ( 6 ) and {[Zn(L)](ClO4)}n ( 7 ) were obtained by reactions of HL · 4HCl with Zn(ClO4)2 · 6H2O under different reaction pH, and they were compared with the corresponding CuII complexes reported previously. The results indicate that the reaction pH and metal ions have remarkable influence on the formation and structure of the complexes.  相似文献   

6.
Chiral cyclic α,α‐disubstituted amino acids, (3S,4S)‐ and (3R,4R)‐1‐amino‐3,4‐(dialkoxy)cyclopentanecarboxylic acids ((S,S)‐ and (R,R)‐Ac5cdOR; R: methyl, methoxymethyl), were synthesized from dimethyl L ‐(+)‐ or D ‐(?)‐tartrate, and their homochiral homoligomers were prepared by solution‐phase methods. The preferred secondary structure of the (S,S)‐Ac5cdOMe hexapeptide was a left‐handed (M) 310 helix, whereas those of the (S,S)‐Ac5cdOMe octa‐ and decapeptides were left‐handed (M) α helices, both in solution and in the crystal state. The octa‐ and decapeptides can be well dissolved in pure water and are more α helical in water than in 2,2,2‐trifluoroethanol solution. The left‐handed (M) helices of the (S,S)‐Ac5cdOMe homochiral homopeptides were exclusively controlled by the side‐chain chiral centers, because the cyclic amino acid (S,S)‐Ac5cdOMe does not have an α‐carbon chiral center but has side‐chain γ‐carbon chiral centers.  相似文献   

7.
Novel optically active substituted acetylenes HC? CCH2CR1(CO2CH3)NHR2 [(S)‐/(R)‐ 1 : R1 = H, R2 = Boc, (S)‐ 2 : R1 = CH3, R2 = Boc, (S)‐ 3 : R1 = H, R2 = Fmoc, (S)‐ 4 : R1 = CH3, R2 = Fmoc (Boc = tert‐butoxycarbonyl, Fmoc = 9‐fluorenylmethoxycarbonyl)] were synthesized from α‐propargylglycine and α‐propargylalanine, and polymerized with a rhodium catalyst to provide the polymers with number‐average molecular weights of 2400–38,900 in good yields. Polarimetric, circular dichroism (CD), and UV–vis spectroscopic analyses indicated that poly[(S)‐ 1 ], poly[(R)‐ 1 ], and poly[(S)‐ 4 ] formed predominantly one‐handed helical structures both in polar and nonpolar solvents. Poly[(S)‐ 1a ] carrying unprotected carboxy groups was obtained by alkaline hydrolysis of poly[(S)‐ 1 ], and poly[(S)‐ 4b ] carrying unprotected amino groups was obtained by removal of Fmoc groups of poly[(S)‐ 4 ] using piperidine. Poly[(S)‐ 1a ] and poly[(S)‐ 4b ] also exhibited clear CD signals, which were different from those of the precursors, poly[(S)‐ 1 ] and poly[(S)‐ 4 ]. The solution‐state IR measurement revealed the presence of intramolecular hydrogen bonding between the carbamate groups of poly[(S)‐ 1 ] and poly[(S)‐ 1a ]. The plus CD signal of poly[(S)‐ 1a ] turned into minus one on addition of alkali hydroxides and tetrabutylammonium fluoride, accompanying the red‐shift of λmax. The degree of λmax shift became large as the size of cation of the additive. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

8.
Four novel chiral phenylacetylenes having an L ‐amino alcohol residue and two hydroxymethyl groups were synthesized and polymerized by an achiral catalyst ((nbd)Rh+6‐(C6H5)B?(C6H5)3]) or a chiral catalytic system ([Rh(nbd)Cl]2/(S)‐ or (R)‐phenylethylamine ((S)‐ or (R)‐PEA)). The two resulting polymers having an L ‐valinol or L ‐phenylalaninol residue showed Cotton effects at wavelengths around 430 nm. This observation indicated that they had an excess of one‐handed helical backbones. Positive and negative Cotton effects were observed only for the polymers having an L ‐valinol residue produced by using (R)‐ and (S)‐PEA as a cocatalyst, respectively, although the monomer had the same chirality. Even when the achiral catalyst was used, the two resulting polymers having an L ‐valinol or L ‐phenylalaninol residue showed Cotton effects despite the long distance between the chiral groups and the main chain. We have found the first example of a new type of chiral monomer, that is, a chiral phenylacetylene monomer having an L ‐amino alcohol residue and two hydroxy groups that was suitable for both modes of asymmetric polymerization, that is, the helix‐sense‐selective polymerization ( HSSP ) with the chiral catalytic system and the asymmetric‐induced polymerization ( AIP ) with the achiral catalyst. The other two monomers having L ‐alaninol and L ‐tyrosinol were found to be unsuitable to neither HSSP nor AIP because of their polymers' low solubility. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

9.
Various ligands, such as (Z)‐1‐phenyl‐2‐[(4S)‐4‐phenyl‐4,5‐dihydro‐1,3‐oxazol‐2‐yl]ethen‐1‐ol ((S)‐ 1a ) and (Z)‐1‐phenyl‐2‐[(4S)‐4‐phenyl‐4,5‐dihydro‐1,3‐thiazol‐2‐yl]ethen‐1‐ol ((S)‐ 1c ), were investigated as auxiliaries for the asymmetric synthesis of chiral ruthenium(II) complexes. The reaction of these chiral auxiliary ligands with [RuCl2(dmso)4], 2,2′‐bipyridine (bpy, 2.2 equiv), and triethylamine (10 equiv) in DMF/PhCl (1:8) at 140 °C for several hours diastereoselectively provided the complexes Λ‐[Ru(bpy)2{(S)‐ 1a ? H}] (Λ‐(S)‐ 2a , 52 % yield, 56:1 d.r.) and Λ‐[Ru(bpy)2{(S)‐ 1c ? H}] (Λ‐(S)‐ 2c , 48 % yield, >100:1 d.r.) in a single step after purification. Both Λ‐(S)‐ 2a and Λ‐(S)‐ 2c could be converted into Λ‐[Ru(bpy)3](PF6)2 by replacing the bidentate enolato ligands with bpy, under retention of configuration, induced by either NH4PF6 as a weak acid (from Λ‐(S)‐ 2a : 73 % yield, 22:1 e.r.; from Λ‐(S)‐ 2c : 77 % yield, 22:1 e.r.), TFA as a strong acid (from Λ‐(S)‐ 2a : 72 % yield, 52:1 e.r.; from Λ‐(S)‐ 2c : 85 % yield, 25:1 e.r.), methylation with Meerwein′s salt (from Λ‐(S)‐ 2a : 59 % yield, 46:1 e.r.; from Λ‐(S)‐ 2c : 86 % yield, 37:1 e.r.), ozonolysis (from Λ‐(S)‐ 2a : 56 % yield, 22:1 e.r.; from Λ‐(S)‐ 2c : 43 % yield, 6.3:1 e.r.), or oxidation with a peroxy acid (from Λ‐(S)‐ 2a : 72 % yield, 45:1 e.r.; from Λ‐(S)‐ 2c : 79 % yield, 8.5:1 e.r.). This study shows that, except for the reaction with NH4PF6, oxazoline‐enolato complex Λ‐(S)‐ 2a provides Λ‐[Ru(bpy)3](PF6)2 with higher enantioselectivities than analogous thiazoline‐enolato complex Λ‐(S)‐ 2c , which might be due to the higher coordinative stability of the thiazoline‐enolato complex, thus requiring more prolonged reaction times. Thus, this study provides attractive new avenues for the asymmetric synthesis of non‐racemic ruthenium(II)‐polypyridyl complexes without the need for using a strong acid or a strong methylating reagent, as has been the case in all previously reported auxiliary methods from our group.  相似文献   

10.
Organometallic 5d6 Transition Metal Complexes of 1‐Methyl‐(2‐alkylthiomethyl)‐1H‐benzimidazole Ligands: Structures and Electrochemical Oxidation The complexes [(mmb)Re(CO)3Cl], [(mtb)Re(CO)3Cl], [(mmb)OsCl(Cym)](PF6) and [(Cym)OsCl(mtb)](PF6) where Cym = p‐cymene, mmb = 1‐methyl‐(2‐methylthiomethyl)‐1H‐benzimidazole and mtb = 1‐methyl‐(2‐tert‐butylthiomethyl)‐1H‐benzimidazole were synthesized and, except for the latter, structurally characterized. In comparison with other late transition metal compounds of these N‐S chelate ligands the rhenium(I) systems exhibit a balanced coordination to both N and S donor atoms. Anodic one‐electron oxidation produces EPR‐silent rhenium(II) states whereas the osmium(III) species [(mmb)OsCl(Cym)]2+ could be identified via EPR and UV/VIS spectroelectrochemistry.  相似文献   

11.
A series of chiral diphosphinoferrocene ligands 3a – i , derived from josiphos (=(2R)‐1‐[(1R)‐1‐(dicyclohexylphosphino)ethyl]‐2‐(diphenylphosphino)ferrocene, formerly called {(R)‐1‐[(S)‐2‐(diphenylphosphino)ferrocenyl]ethyl}dicycloxexylphosphine) where the electronic properties of the ligand are systematically varied, were prepared. X‐Ray studies of five of these new ligands confirmed that these compounds display very similar conformations in the solid state and that no structural criteria could be found indicating the modified electronic properties. These ligands find application in the Pd‐catalyzed highly regio‐ and stereoselective CO/propene copolymerization reaction, where the electronic properties of the ligand show a great impact on the catalyst activity. Coordination‐chemical aspects of these diphosphinoferrocenes relevant to the CO/propene copolymerization reaction were addressed by the preparation and characterization of Pd‐ and Pt‐complexes of the general formula [PdCl2(P−P)] ( 5 ), [PdMe2(P−P)] ( 6 ), [PdClMe(P−P)] ( 7 ), [PdMe(MeCN)(P−P)]PF6 ( 8 ), and [PtClMe(P−P)] ( 9 ) (P−P=chiral diphosphinoferrocene ligand ( 3a – h ), four of which were characterized by X‐ray crystallography.  相似文献   

12.
Based on the a ligand BDPPZ [(9a,13a‐dihydro‐4,5,9,14‐tetraaza‐benzo[b]triphenylene‐11‐yl)‐phenyl‐methanone] (1) and its polypyridyl hetero‐ and homoleptic Ru(II) metal complexes, [Ru(bpy)2L](PF6)2 (2), [Ru(phen)2L](PF6)2 (3), [Ru(dafo)2L](PF6)2 (4), [Ru(dcbpy)2L](PF6)2 (5) and [RuL3](PF6)2 (6) (where, L = ligand, bpy = 2,2′‐bipyridine, phen = 1,10‐phenantroline, dafo = 4,5‐diazafluoren‐9‐one and dcbpy = 3,3′‐dicarboxy‐2,2′‐bipyridine), have been synthesized and characterized by elemental analysis, UV–vis, FT‐IR, 1H and 13C‐NMR spectra (for ligand), molar conductivity measurements and X‐ray powder techniques. The electrochemical parameters of the substituted ligand and its polypyridyl hetero‐ and homoleptic Ru(II) metal complexes are reported by cyclic voltammetry. UV–vis spectroscopy is used to compare the differences between the conjugated π systems in this ligand and its Ru(II) metal complexes. The polypyridyl hetero‐ and homoleptic Ru(II) metal complexes also tested as catalysts for the formation of cyclic organic carbonates from carbon dioxide and liquid epoxides which served as both reactant and solvent. The results showed that the [Ru(L)3](PF6)2 (6) complex is more efficient than the other Ru(II) complexes for the formation of cyclic organic carbonates from carbon dioxide. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.
Alkynyl gold(I) metallaligands [(AuC≡Cbpyl)2(μ‐diphosphine)] (bpyl=2,2′‐bipyridin‐5‐yl; diphosphine=Ph2P(CH2)nPPh2, [n=3 (LPr), 4 (LBu), 5 (LPent), 6 (LHex)], dppf (LFc), Binap (LBinap) and Diop (LDiop)) react with MX2 (M=Fe, Zn, X=ClO4; M=Co, X=BF4) to give triple helicates [M2(LR)3]X4. These complexes, except those containing the semirigid LBinap metallaligand, present similar hydrodynamic radii (determined by diffusion NMR spectroscopy measurements) and a similar pattern in the aromatic region of their 1H NMR spectra, which suggests that in solution they adopt a compact structure where the long and flexible organometallic strands are folded. The diastereoselectivity of the self‐assembly process was studied by using chiral metallaligands, and the absolute configuration of the iron(II) complexes with LBinap and LDiop was determined by circular dichroism spectroscopy (CD). Thus, (R)‐LBinap or (S)‐LBinap specifically induce the formation of (Δ,Δ)‐[Fe2((R)‐LBinap)3](ClO4)4 or (Λ,Λ)‐[Fe2((S)‐LBinap)3](ClO4)4, respectively, whereas (R,R)‐ or (S,S)‐LDiop give mixtures of the ΔΔ‐ and ΛΛ‐diastereomers. The ΔΔ helicate diastereomer is dominant in the reaction of FeII with (R,R)‐LDiop, whereas the ΛΛ isomer predominates in the analogous reaction with (S,S)‐LDiop. The photophysical properties of the new dinuclear alkynyl complexes and the helicates have been studied. The new metallaligands and the [Zn2(LR)3]4+ helicates present luminescence from [π→π*] excited states mainly located in the C≡Cbpyl units.  相似文献   

14.
In the title coordination polymer, catena‐poly[[dichloridomanganese(II)]‐μ‐1,1‐diphenyl‐3,3′‐[(1R,2R)‐cyclohexane‐1,2‐diylbis(azaniumylylidene)]dibut‐1‐en‐1‐olate‐κ2O:O′], [MnCl2(C26H30N2)]n, synthesized by the reaction of the chiral Schiff base ligand 1,1‐diphenyl‐3,3′‐[(1R,2R)‐cyclohexane‐1,2‐diylbis(azanediyl)]dibut‐2‐en‐1‐one (L) with MnCl2·4H2O, the asymmetric unit contains one crystallographically unique MnII ion, one unique spacer ligand, L, and two chloride ions. Each MnII ion is four‐coordinated in a distorted tetrahedral coordination environment by two O atoms from two L ligands and by two chloride ligands. The MnII ions are bridged by L ligands to form a one‐dimensional chain structure along the a axis. The chloride ligands are monodentate (terminal). The ligand is in the zwitterionic enol form and displays intramolecular ionic N+—H...O hydrogen bonding and π–π interactions between pairs of phenyl rings which strengthen the chains.  相似文献   

15.
Three tripodal ligands H3L1–3 containing imidazole rings were synthesized by the reaction of 1,10‐phenanthroline‐5,6‐dione with 1,3,5‐tris[(3‐formylphenoxy)methyl]benzene, 1,3,5‐tris[(3‐formylphenoxy)methyl]‐2,4,6‐trimethylbenzene, and 2,2′,2"‐tris[(3‐formylphenoxy)ethyl]amine, respectively. Trinuclear RuII polypyridyl complexes [(bpy)6Ru3H3L1–3](PF6)6 were prepared by the condensation of Ru(bpy)2Cl2 · 2H2O with ligands H3L1–3. The pH effects on the UV/Vis absorption and fluorescence spectra of the three complexes were studied, and ground‐ and excited‐state ionization constants of the three complexes were derived. The three complexes act as “off‐on‐off” fluorescence pH switch through protonation and deprotonation of imidazole ring with a maximum on‐off ratio of 5 in buffer solution at room temperature.  相似文献   

16.
In this work, we have successfully synthesized a new family of chiral Schiff base–phosphine ligands derived from chiral binaphthol (BINOL) and chiral primary amine. The controllable synthesis of a novel hexadentate and tetradentate N,O,P ligand that contains both axial and sp3‐central chirality from axial BINOL and sp3‐central primary amine led to the establishment of an efficient multifunctional N,O,P ligand for copper‐catalyzed conjugate addition of an organozinc reagent. In the asymmetric conjugate reaction of organozinc reagents to enones, the polymer‐like bimetallic multinuclear Cu? Zn complex constructed in situ was found to be substrate‐selective and a highly excellent catalyst for diethylzinc reagents in terms of enantioselectivity (up to >99 % ee). More importantly, the chirality matching between different chiral sources, C2‐axial binaphthol and sp3‐central chiral phosphine, was crucial to the enantioselective induction in this reaction. The experimental results indicated that our chiral ligand (R,S,S)‐ L1 ‐ and (R,S)‐ L4 ‐based bimetallic complex catalyst system exhibited the highest catalytic performance to date in terms of enantioselectivity and conversion even in the presence of 0.005 mol % of catalyst (S/C=20 000, turnover number (TON)=17 600). We also studied the tandem silylation or acylation of enantiomerically enriched zinc enolates that formed in situ from copper‐ L4 ‐complex‐catalyzed conjugate addition, which resulted in the high‐yield synthesis of chiral silyl enol ethers and enoacetates, respectively. Furthermore, the specialized structure of the present multifunctional N,O,P ligand L1 or L4 , and the corresponding mechanistic study of the copper catalyst system were investigated by 31P NMR spectroscopy, circular dichroism (CD), and UV/Vis absorption.  相似文献   

17.
Based on chiral, enantiomerically pure 7‐[(S)‐phenylethylurea]‐8‐hydroxyquinoline ( 1 ‐H), trinuclear helicate‐type complexes 2 – 5 are formed with divalent transition‐metal cations. X‐ray structural analyses reveal the connection of two monomeric complex units [M( 1 )3]? (M=Zn, Mn, Co, Ni) by a central metal ion to form a “dimer”. Due to the enantiopurity of the ligand, the complexes are obtained as pure enantiomers, resulting in pronounced circular dichroism (CD) spectra. Single‐ion effects and intra‐ and intermolecular coupling are observed with dominating ferromagnetic coupling in the case of the cobalt(II) and nickel(II) and dominating antiferromagnetic coupling in the case of the manganese(II) complex.  相似文献   

18.
To investigate how the central metalloligand geometry influences distant or vicinal metal‐to‐metal charge‐transfer (MMCT) properties of polynuclear complexes, cis‐ and trans‐isomeric heterotrimetallic complexes, and their one‐ and two‐electron oxidation products, cis/trans‐ [Cp(dppe)FeIINCRuII(phen)2CN‐FeII(dppe)Cp][PF6]2 (cis/trans‐ 1 [PF6]2), cis/trans‐[Cp(dppe)FeIINCRuII(phen)2CNFeIII‐(dppe)Cp][PF6]3 (cis/trans‐ 1 [PF6]3) and cis/trans‐[Cp(dppe)FeIIINCRuII(phen)2CN‐FeIII(dppe)Cp][PF6]4 (cis/trans‐ 1 [PF6]4) have been synthesized and characterized. Electrochemical measurements show the presence of electronic interactions between the two external FeII atoms of the cis‐ and trans‐isomeric complexes cis/trans‐ 1 [PF6]2. The electronic properties of all these complexes were studied and compared by spectroscopic techniques and TDDFT//DFT calculations. As expected, both mixed valence complexes cis/trans‐ 1 [PF6]3 exhibited different strong absorption signals in the NIR region, which should mainly be attributed to a transition from an MO that is delocalized over the RuII‐CN‐FeII subunit to a FeIII d orbital with some contributions from the co‐ligands. Moreover, the NIR transition energy in trans‐ 1 [PF6]3 is lower than that in cis‐ 1 [PF6]3, which is related to the symmetry of their molecular orbitals on the basis of the molecular orbital analysis. Also, the electronic spectra of the two‐electron oxidized complexes show that trans‐ 1 [PF6]4 possesses lower vicinal RuII→FeIII MMCT transition energy than cis‐ 1 [PF6]4. Moreover, the assignment of MMCT transition of the oxidized products and the differences of the electronic properties between the cis and trans complexes can be well rationalized using TDDFT//DFT calculations.  相似文献   

19.
Herein, a unique coordination system that exhibits multiple chiral inversions and molecular dimerization in response to a subtle pH change is reported. Treatment of (Δ)2‐H3[Au3Co2(L ‐cys)6] (H3[ 1 a ]) with [Co3(aet)6](NO3)3 (aet=2‐aminoethanethiolate) in water at pH 7 gave a 1:1 complex salt of [Co3(aet)6]3+ and [ 1 a ]3?, retaining the AuI3CoIII2 structure and chiral configurations of [ 1 a ]3?. Similar treatment at pH 9 led to not only the inversion of all of the chiral CoIII and S centers but also the dimerization of [ 1 a ]3?, giving a 2:1 complex salt of [Co3(aet)6]3+ and (Λ)4(R)12‐[Au6Co4(L ‐cys)12]6? ([ 2 ]6?). When dissociated from [Co3(aet)6]3+ in solution, [ 2 ]6? was converted to (Λ)2(R)6‐[Au3Co2(L ‐cys)6]3? ([ 1 b ]3?) with retention of the chiral configurations.  相似文献   

20.
The title cocrystal contains two chiral conformational diastereomers, viz. (1S,2R,RN)‐ and (1S,2R,SN)‐, of [2,4‐di‐tert‐butyl‐6‐{[(1‐oxido‐1‐phenylpropan‐2‐yl)(methyl)amino]methyl}phenolato](methanol)‐cis‐dioxidomolybdenum(VI), [Mo(C25H35NO2)O2(CH3OH)], representing the first example of a structurally characterized molybdenum complex with enantiomerically pure ephedrine derivative ligands. The MoVI cations exhibit differently distorted octahedral coordination environments, with two oxide ligands positioned cis to each other. The remainder of the coordination comprises phenoxide, alkoxide and methanol O atoms, with an amine N atom completing the octahedron. The distinct complexes are linked by strong intermolecular O—H...O hydrogen bonds, resulting in one‐dimensional molecular chains. Furthermore, the phenyl rings are involved in weak T‐shaped/edge‐to‐face π–π interactions with each other.  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