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 46 毫秒
1.
Enantiopreferential energy transfer processes between dissymmetric lanthanide and transition metal complexes dissolved in acetonitrile are studied using chiroptical luminescence techniques. The energy donors (luminophores) in this study are a racemic mixture of Ln(dpa)3 (3-) complexes (where Ln = Eu3+ or Tb3+ and dpa = 2,6-pyridinedicarboxylate), and the energy acceptors (quenchers) are an enantiomerically-resolved population of Co(R,R-chxn)3 3+ (where R,R-chxn = trans-1R,2R-diaminocyclohexane) complexes. The luminophores are dissolved in acetonitrile as (NEt4)3[Ln(dpa)3] (where NEt(4) = tetraethlylammonium) and (NBu4)[Ln(dpa)3] (where NBu4 = tetrabutylammonium) salts. The unquenched luminescence lifetimes are reported for both Eu(dpa)3 (3-) and Tb(dpa)3 (3-) in acetonitrile over the range 263-333 K, and these results are compared to luminescence lifetimes in aqueous solution. Time-resolved chiroptical luminescence measurements of enantiopreferential quenching kinetics are reported for samples with Eu(dpa)3 (3-) and Co(R,R-chxn)3 3+ in acetonitrile over 263-333 K range. These results are analyzed using a phenomenological quenching kinetics model, and the results are compared to results in aqueous solution. These comparisons show that the overall Eu-Co luminescence quenching efficiency is reduced in acetonitrile vs. aqueous samples, because the salts of (NX4)3[Eu(dpa)3] are not completely dissociated in acetonitrile. However, the enantiopreference exhibited is identical in acetonitrile vs. aqueous solution.  相似文献   

2.
Time-resolved chiroptical luminescence (TR-CL) measurements are used to study chirality-dependent intermolecular interactions in dynamic excited-state quenching processes. The measurements are carried out on solution samples that contain a racemic mixture of chiral luminophore molecules (with enantiomeric structures denoted by LambdaL and DeltaL) and a small, optically resolved concentration of chiral quencher (CQ) molecules. The luminophores are excited with a pulse of linearly polarized laser radiation to produce an initially racemic excited-state population of LambdaL* and DeltaL* enantiomers, and TR-CL measurements are then used to monitor the differential decay kinetics of the LambdaL* and DeltaL* subpopulations. Observed differences between the LambdaL* and DeltaL* decay kinetics reflect differential rate processes and efficiencies for LambdaL*-CQ vs. DeltaL*-CQ quenching actions, and they are diagnostic of chiral discriminatory interactions between the luminophore and quencher molecules. Twelve different luminophore-quencher systems are examined, in both H(2)O and D(2)O solutions, and in each case the quenching kinetics are measured over the 273-308 K temperature range. In all of the systems examined here, quenching occurs via electronic energy-transfer processes in transient (LambdaL*-CQ) and (DeltaL*-CQ) encounter complexes, and the chiral discriminatory rate parameters reflect the relative stabilities and lifetimes of these complexes as well as their structures and internal (electronic and nuclear) dynamics. All of the luminophore and quencher molecules examined in this study have three-bladed propeller-like structures that are very similar in overall shape and size. However, they exhibit small differences in the structural details of their propeller blades, and it is found that these small differences in structure can produce both qualitative and very substantial quantitative differences in their chiral recognition and discrimination properties.  相似文献   

3.
The initial rates of reactivity of oxaliplatin, its metabolites Pt(dach)Cl2 and Pt(dach)(OH2)2(2+) with guanosine and L-met in water, NaCl and phosphate were compared. Versus guanosine, the most reactive molecule was Pt(dach)(OH2)2(2+), about 40 fold that of oxaliplatin, the least reactive was Pt(dach)Cl2, Versus L-met, Pt(dach)(OH2)2(2+), was also the most reactive species but only about 2 fold more reactive than Pt(dach)Cl2 and oxaliplatin. Pt(dach)(OH2)2(2+) was approximately 3 fold less reactive versus methionine than guanosine whereas oxaliplatin and Pt(dach)Cl2 were about seven fold more reactive versus methionine than guanosine. Thus, the three platinum compounds oxaliplatin, Pt(dach)Cl2 and Pt(dach)(OH2)2(2+) react with L-met but only the Pt(dach)(OH2)2(2+) has a high reactivity with guanosine. Oxaliplatin, which is stable in water, has to be transformed in the presence of chloride in chloro-derivatives which are aquated to become active particularly versus guanosine. These data demonstrate that oxaliplatin has similarities with cisplatin in terms of chloride versus water coordination and in terms of dependence on chloride concentration for transformations.  相似文献   

4.
Crystallisation of [Co(CN)(6)](3-) or [Cr(CN)(6)](3-) with Ln(iii) salts (Ln = Nd, Gd, Yb) from aqueous dmf afforded the cyanide-bridged d/f systems [Ln(dmf)(4)(H(2)O)(3)(micro-CN)Co(CN)(5)] (-, discrete dinuclear species) and {[Cr(CN)(4)(micro-CN)(2)Ln(H(2)O)(2)(dmf)(4)]}(infinity) (-, infinite cyanide-bridged chains with alternating Cr and Ln centres). With Ln = Gd the characteristic long-lived phosphorescence from d-d excited states of the [M(CN)(6)](3-) units was apparent in the red region of the spectrum, with lifetimes of the order of 1 micros, since the heavy atom effect of the Gd(iii) promotes inter-system crossing at the [M(CN)(6)](3-) units to generate the phosphorescent spin-forbidden excited states. With Ln = Yb or Nd however, the d-block luminescence was completely quenched due to fast (>10(8) s(-1)) energy-transfer to the Ln(iii) centre, resulting in the characteristic sensitised emission from Yb(iii) and Nd(iii) in the near-IR region. For both - and -, calculations based on spectroscopic overlap between emission of the donor (Co) and absorption of the acceptor (Ln) suggest that the Dexter energy-transfer mechanism is responsible for the complete quenching that we observe.  相似文献   

5.
Koide H  Uemura M 《Chirality》2000,12(5-6):352-359
Axially chiral N,N-diethyl 2,6-disubstituted benzamides were stereo-selectively prepared utilizing planar chiral (arene)chromium complexes as an enantiomerically active form by following two methods. Ortho-lithiation of the enantiomerically pure planar chiral tricarbonyl(N,N-diethyl 2-methylbenzamide)chromium complex followed by electrophilic quenching gave axially chiral 2-methyl-6-substituted N,N-diethyl benzamide chromium complexes. Photo-oxidative demetalation produced the chromium-free axially chiral benzamides as optically active compounds. An alternative method for the preparation of axial chiral benzamides is an enantioselective lithiation at the benzylic methyl of meso tricarbonyl(N,N-diethyl 2,6-dimethylbenzamide)chromium with appropriate chiral lithium amide base followed by quenching with alkyl halides.  相似文献   

6.
Many genomic assays rely on a distance-dependent interaction between luminescent labels, such as luminescence quenching or resonance energy transfer. We studied the interaction between electrochemically excited Ru(bpy)(3) (2+) and Cy5 in a hybridization assay on a chip. The 3' end of an oligonucleotide was labelled with Ru(bpy)(3) (2+) and the 5' end of a complementary strand with Cy5. Upon the hybridization, the electrochemiluminescence (ECL) of Ru(bpy)(3) (2+) was efficiently quenched by Cy5 with a sensitivity down to 30 nmol/L of the Cy5-labelled complementary strand. The quenching efficiency is calculated to be 78%. A similar phenomenon was observed in a comparative study using laser-excitation of Ru(bpy)(3) (2+). The hybridization with the non-labelled complementary or labelled non-complementary strand did not change the intensity of the ECL signal. Resonance energy transfer, electron transfer and static quenching mechanisms are discussed. Our results suggest that static quenching and/or electron transfer are the most likely quenching mechanisms.  相似文献   

7.
Extended metal atom chains (EMACs) contain a linear metal chain wrapped by various ligands. Most complexes are of the form M(3)(dpa)(4)X(2), where M = metal, dpa = 2,2'-dipyridylamide, and X = various anions. The ligands form helical coils about the metal chain, which results in chiral EMAC complexes. The EMACs containing the metals Co and Cu were partially separated in polar organic mode using a vancomycin-based chiral stationary phase. Under similar conditions, two EMACs with Ni metal and varying anions could be baseline separated. The polar organic mode was used because of the instability of the compounds in aqueous mobile phases. Also, these conditions are more conducive to preparative separations. Polarimetric measurements on the resolved enantiomers of Ni(3)(dpa)(4)Cl(2) indicate that they have extraordinarily high specific rotations (on the order of 5000 deg cc/g dm).  相似文献   

8.
Nine new metal complexes of the quinolone antibacterial agent N-propyl-norfloxacin, pr-norfloxacin, with VO(2+), Mn(2+), Fe(3+), Co(2+), Ni(2+), Zn(2+), MoO(2)(2+), Cd(2+) and UO(2)(2+) have been prepared and characterized with physicochemical and spectroscopic techniques while molecular mechanics calculations for Fe(3+), VO(2+) and MoO(2)(2+) complexes have been performed. In all complexes, pr-norfloxacin acts as a bidentate deprotonated ligand bound to the metal through the pyridone and one carboxylate oxygen atoms. All complexes are six-coordinate with slightly distorted octahedral geometry. For the complex VO(N-propyl-norfloxacinato)(2)(H(2)O) the axial position, trans to the vanadyl oxygen, is occupied by one pyridone oxygen atom. The investigation of the interaction of the complexes with calf-thymus DNA has been performed with diverse spectroscopic techniques and has shown that the complexes can be bound to calf-thymus DNA resulting to a B-->A DNA transition. The antimicrobial activity of the complexes has been tested on three different microorganisms. The complexes show equal or decreased biological activity in comparison to the free pr-norfloxacin except UO(2)(pr-norf)(2) which shows better inhibition against S. aureus.  相似文献   

9.
Circularly polarized luminescence (CPL) from selected transitions of Eu(III) in resolved single crystals of Na3[Eu(ODA)3].2NaClO4.6H2O are compared to CPL results obtained from solutions containing perturbed racemic mixtures of Eu(2,6-pyridine-dicarboxylate)3 (3-) and enantiomerically pure d-f helicate LambdaLambda-(-)EuCr(L8)3] in order to determine an empirical relationship between helicity and CPL spectra. Comparison of the CPL results, even for the magnetic dipole allowed transitions of Eu(III) where one measures large chiral discrimination, shows that the signs and magnitudes do not correlate with the overall helicity of the Eu(III) site. It is concluded that the symmetry of the Eu(III) site in LambdaLambda-(-)EuCr(L8)3 is not close enough to D3 to allow for the confirmation of the presumed spectra:structure correlation.  相似文献   

10.
The highly enantioselective hydrolytic kinetic resolution (HKR) of racemic terminal epoxides by new bimetallic chiral (salen)Co provides a operationally very simple protocol for the synthesis of enantiomerically enriched terminal epoxides (>99% ee) and diols. Optically pure chlorohydrins have been synthesized in one step by ring‐opening reactions of terminal epoxides with HCl using kinetic resolution. © 2005 Wiley‐Liss, Inc. Chirality  相似文献   

11.
The yield of hole injection into guanines of different oligonucleotide duplexes by a photooxidizing tethered Ru(II) complex is examined by measuring the luminescence quenching of the excited complex. This yield is investigated as a function of the anchoring site of the complex (on a thymine nucleobase in the middle of the sequence or on the 5' terminal phosphate) and the number and position of the guanine bases as compared with the site of attachment of the Ru(II) compound. In contrast to other studies, the tethered complex, [Ru(tap)(2)(dip)](2+), is a non-intercalating compound and has been shown previously to produce an irreversible photocrosslinking between the two strands as the ultimate step of hole injection. The study of luminescence quenching of the anchored complex by emission intensity and lifetime measurements for the different duplexes indicates that a direct contact between the complex and the guanine nucleobase is needed for the electron transfer to take place. Moreover, for none of the sequences a clear contribution of a static quenching is evidenced independently of the two types of attachment of the [Ru(tap)(2)(dip)](2+) complex to the oligonucleotide. A comparison of the fastest hole-injection process by electron transfer to the excited anchored [Ru(tap)(2)(dip)](2+), with the rate of the photo-electron transfer between the same complex free in solution and guanosine-5'-monophosphate, indicates that the hole injection by the anchored complex is slower by a factor of 10 at least. A bad overlap between donor and acceptor orbitals is probably the cause of this slow rate, which could be attributed to some steric hindrance induced by the complex linker.  相似文献   

12.
A new Ru(II) complex of [Ru(bpy)(2)(Hpip)](2+) {bpy = 2,2'bipyridine; Hppip = 2-(4-(pyridin-2-yl)phenyl)-1H-imidazo[4,5-f][1,10]phenanthroline} has been synthesized by grafting of 2-pyridyl to parent complex [Ru(bpy)(2)(Hpip)](2+) {Hppip = 2-(4-phenyl)-1H-imidazo[4,5-f][1,10]phenanthroline}. The acid-base properties of [Ru(bpy)(2)(Hpip)](2+) studied by UV-visible and luminescence spectrophotometric pH titrations, revealed off-on-off luminescence switching of [Ru(bpy)(2)(Hpip)](2+) that was driven by the protonation/deprotonation of the imidazolyl and the pyridyl moieties. The complex was demonstrated to be a DNA intercalator with an intrinsic DNA binding constant of (5.56 ± 0.2) x 10(5) M-1 in buffered 50 mM NaCl, as evidenced by UV-visible and luminescence titrations, reverse salt effect, DNA competitive binding with ethidium bromide, steady-state emission quenching by [Fe(CN)6]4-, DNA melting experiments and viscosity measurements. The density functional theory method was also used to calculate geometric/electronic structures of the complex in an effort to understand the DNA binding properties. All the studies indicated that the introduction of 2-pyridyl onto Hpip ligand is more favorable for extension of conjugate plane of the main ligand than that of phenyl, and for greatly enhanced ct-DNA binding affinity accordingly.  相似文献   

13.
Moussa A  Pham C  Bommireddy S  Muller G 《Chirality》2009,21(5):497-506
The perturbation of the racemic equilibrium of luminescent D3 terbium(III) complexes with chelidamic acid (CDA), a hydroxylated derivative of 2,6-pyridine-dicarboxylic acid (DPA), by added chiral biomolecules such as L-amino acids has been studied using circularly polarized luminescence and 13C NMR spectroscopy. It is shown in this work that the chiral-induced equilibrium shift of [Tb(CDA)3](6-) by L-amino acids (i.e. L-proline or L-arginine) was largely influenced by the hydrogen-bonding networks formed between the ligand interface of racemic [Tb(CDA)3](6-) and these added chiral agents. The capping of potential hydrogen-bonding sites by acetylation in L-proline led to a approximately 100-fold drop in the induced optical activity of the [Tb(CDA)3](6-):N-acetyl-L-proline system. This result suggested that the hydrogen-bonding networks serve as the basis for further noncovalent discriminatory interactions between racemic [Tb(CDA)3](6-) and added L-amino acids.  相似文献   

14.
Fluorescence quenching and resonance energy transfer have been studied by steady-state fluorescence spectroscopy. The experimental and theoretical values for the rate constants of the electronic energy transfer (kET) and critical radius (R0) were determined for prodan and laurdan as donors and octadecyl rhodamine B as acceptor. The spectroscopic data show, that prodan and laurdan in solution create an inhomogeneous spectroscopic medium in which multi-channel luminescence phenomena take place. This finding indicated that the modified form of the Stern-Volmer relation should be used for analyzing fluorescence quenching data. Results of performed studies point out, that dipole-dipole interaction is responsible for the resonance energy transfer from prodan and laurdan to octadecyl rhodamine B. The relative quenching efficiencies of both dyes depend on polarity of the medium and are higher for more polar solvent (AcN).  相似文献   

15.
The absolute configuration of the title compound has been determined after its enzymatic hydrolysis to 3-hydroxy-5-phenylpentanoic acid, esterification, and identification of the enantiomerically pure methyl (3R)-hydroxy-5-phenylpentanoate by HPLC on Chiralcel(R)OD-H. For reasons of inconsistent literature data, enantioselective reductions of methyl 3-oxo-5-phenylpentanoate have been reinspected and the stereochemical outcome unequivocally confirmed by both chiroptical and HPLC retention data.  相似文献   

16.
The achiral hydrocarbon tetraphenylethylene crystallizes in enantiomorphous forms (chiral space group: P21) to afford right- and left-handed hemihedral crystals, which can be recognized by solid-state circular dichroism spectroscopic analysis. Chiral organic crystals of tetraphenylethylene mediated enantioselective addition of diisopropylzinc to pyrimidine-5-carbaldehyde to give, in conjunction with asymmetric autocatalysis with amplification of chirality, almost enantiomerically pure (S)- and (R)-5-pyrimidyl alkanols whose absolute configurations were controlled efficiently by the crystalline chirality of the tetraphenylethylene substrate. Tetrakis(p-chlorophenyl)ethylene and tetrakis(p-bromophenyl)ethylene also show chirality in the crystalline state, which can also act as a chiral substrate and induce enantioselectivity of diisopropylzinc addition to pyrimidine-5-carbaldehyde in asymmetric autocatalysis to give enantiomerically enriched 5-pyrimidyl alkanols with the absolute configuration correlated with that of the chiral crystals. Highly enantioselective synthesis has been achieved using chiral crystals composed of achiral hydrocarbons, tetraphenylethylenes, as chiral inducers. This chemical system enables significant amplification of the amount of chirality using spontaneously formed chiral crystals of achiral organic compounds as the seed for the chirality of asymmetric autocatalysis.  相似文献   

17.
The enantioselective photolysis of a cold gas‐phase noncovalent complex of tryptophan with alkali‐metalized L‐serine, M+(L‐Ser)(Trp) (M = Na and Li), was examined using a tandem mass spectrometer containing a variable‐temperature ion trap. CO2 loss from Trp in the clusters was enantiomerically selective in ultraviolet excitation with linearly polarized light. M+(L‐Ser) promoted the enantioselective photolysis of Trp as a chiral auxiliary. The enantioselective photolysis of the D‐enantiomer was applied to a quantitative chiral analysis, in which the optical purity of tryptophan could be determined by measuring the relative abundance ratio R of the enantioselective CO2 loss to the chiral‐independent evaporation of L‐Ser in a single photodissociation mass spectrum of M+(L‐Ser)(Trp). Chirality 27:349–352, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

18.
The partitioning and biotransformations of oxaliplatin [trans-l-1,2-diaminocyclohexaneoxalatoplatinum(II)] were investigated in the blood of Wistar male rats in vitro. [3-H]-Oxaliplatin was incubated with rat blood at 37 degrees C in 5% CO2 and the concentrations of all Pt complexes containing the [3-H]-dach carrier ligand were followed for up to 12 hours. Decay for both oxaliplatin and Pt-dach in the plasma ultrafiltrate (PUF) was rapid (t 1/2 oxaliplatin = 0.68 h and t 1/2 for Pt-dach in the PUF = 0.85 h). After 9 hours, the concentration of oxaliplatin fell below the detection limit. By 4 hours, the PUF-Pt-dach reached a plateau, which was 12% of total Pt-dach. The binding of Pt-dach to red blood cells (RBCs) and plasma proteins was also very rapid (t 1/2 RBCs = 0.58 h and t 1/2 plasma proteins = 0.78 h) and reached equilibrium by 4 hours. At equilibrium, 35% of total Pt-dach was bound to plasma proteins, 12% was in the plasma ultrafiltrate, and 53% was found associated with RBCs. Of the Pt-dach associated with RBCs, 23% was bound to the RBC membrane, 58% was bound to RBC cytosolic proteins, and 19% was in the RBC cytosol ultrafiltrate. Thus, these studies confirm previous observations of oxaliplatin accumulation by rat RBCs. To better characterize the determinants of this accumulation, oxaliplatin and other Pt-dach complexes were compared with respect to both their uptake by rat RBCs and their partition coefficients in octanol and water. The rank order for the rate of uptake was ormaplatin approximately Pt(dach)Cl2 > oxaliplatin > Pt(dach)(mal); while the rank order for hydrophobicity was ormaplatin > Pt(dach)Cl2 > Pt(dach)(mal) > oxaliplatin. Thus, in general, Pt-dach complexes appeared to be taken up better by RBCs than cisplatin or carboplatin, and the hydrophobicity of most of the Pt-dach complexes appeared to correlate with uptake. However, factors other than the dach carrier ligand and hydrophobicity clearly influence uptake. The biotransformations of oxaliplatin in rat blood were characterized utilizing reverse-phase high-pressure liquid chromatography (HPLC). In the RBC cytosol, both oxaliplatin and Pt(dach)Cl2 were observed at early times, while Pt(dach)(GSH)2, Pt(dach)(Cys)2, Pt(dach)(GSH), and free dach accumulated and reached steady-state levels by 4 hours. Thus, in the RBC cytosol, only chemically unreactive biotransformation products such as free dach and Pt-dach complexes with cysteine and glutathione accumulated in significant amounts. Furthermore, only Pt(dach)(Cys)2 and free dach appeared to efflux from RBCs. Thus, RBCs do not appear to serve as a reservoir for cytotoxic Pt-dach complexes. Finally, the biotransformation products of oxaliplatin in the plasma were identified as Pt(dach)Cl2, Pt(dach)(Cys)2, Pt(dach)(GSH), Pt(dach)(Met), Pt(dach)(GSH)2, and free dach. Among these compounds, Pt(dach)Cl2 formed transiently, while Pt(dach)(Cys)2, Pt(dach)(Met), and free dach accumulated and were the major biotransformation products by 4 hours. Thus, this study has identified the major inert and reactive biotransformation products of oxaliplatin in both plasma and RBCs and thus provides the information required for detailed pharmacokinetic and biotransformation studies of oxaliplatin. [figure in text]  相似文献   

19.
Y Liu  K Zhang  Y Wu  J Zhao  J Liu 《Chemistry & biodiversity》2012,9(8):1533-1544
8-Hydroxyquinoline-7-carboxaldehyde (8-HQ-7-CA), Schiff-base ligand 8-hydroxyquinoline-7-carboxaldehyde benzoylhydrazone, and binuclear complexes [LnL(NO(3) )(H(2) O)(2) ](2) were prepared from the ligand and equivalent molar amounts of Ln(NO(3) )?6 H(2) O (Ln=La(3+) , Nd(3+) , Sm(3+) , Eu(3+) , Gd(3+) , Dy(3+) , Ho(3+) , Er(3+) , Yb(3+) , resp.). Ligand acts as dibasic tetradentates, binding to Ln(III) through the phenolate O-atom, N-atom of quinolinato unit, and C?N and ?O?C?N? groups of the benzoylhydrazine side chain. Dimerization of this monomeric unit occurs through the phenolate O-atoms leading to a central four-membered (LnO)(2) ring. Ligand and all of the Ln(III) complexes can strongly bind to CT-DNA through intercalation with the binding constants at 10(5) -10(6) M(-1) . Moreover, ligand and all of the Ln(III) complexes have strong abilities of scavenging effects for hydroxyl (HO(.) ) radicals. Both the antioxidation and DNA-binding properties of Ln(III) complexes are much better than that of ligand.  相似文献   

20.
The ligand ODHIP (3,4-dihydroxyl-imidazo[4,5-f][1,10]phenanthroline) and its cobalt(III) complex [Co(bpy)(2)(ODHIP)](3+) were synthesized and characterized. Binding of this complex with calf thymus DNA has been investigated by spectroscopic methods and viscosity. The experimental results indicated that the complex bound to DNA by intercalation. In Tris buffer, the complex could emit relatively weak luminescence. After binding to DNA, the notable enhancement was observed. However, when the Cu(2+) was further added, the luminescence decreased gradually and disappeared after the equimolar concentrations of Cu(2+) was added, which exhibited the "off-on-off" properties of molecular light switch.  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