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.
The reduction potentials for two Wurster’s crowns, aza crown ethers which incorporate the redox active N,N,N′,N′-tetraalkyl-1,4-phenylenediamine into the structure of 18-crown-6, were studied in the presence of the siderophore ferrioxamine B, FeHDFB+. Addition of FeHDFB+ resulted in a positive shift in the host reduction potential for both aza crown ethers studied. This shift is explained in terms of host-guest supramolecular assembly formation, which was independently verified by FAB-MS. An enhanced affinity for host-guest formation of the reduced aza crown ether was calculated for each aza crown ether-siderophore assembly using a thermochemical cycle. These differences in host binding affinity as a function of redox state can be harnessed for use in specific metal ion compartmentalization with application, for example, to environmental remediation.  相似文献   

2.
Stability constants for the 1:1 complexes of Na+, K+, Rb+, and Cs+ with dibenzo-18-crown-6 (DB18C6) and dibenzo-24-crown-8 (DB24C8) have been determined by conductometry at 25 °C in a poorly solvating solvent, nitromethane. For both the crown ethers, the stability constant decreases with increasing metal ion size, Na+ > K+ > Rb+ > Cs+, regardless of the size compatibility between the metal ions and the ligand cavities. A comparison of the results with those in several other solvents (S: acetonitrile, propylene carbonate, water, methanol, and N,N-dimethylformamide) leads to the conclusion that the selectivity sequence of these crown ethers in nitromethane agrees with the intrinsic one in the absence of a solvent. Transfer activity coefficients of the crown ethers and their complexes from nitromethane to S have been determined to evaluate the solute-solvent interactions. It is shown that DB24C8 shields the alkali metal ions more effectively from the solvents than DB18C6 because of the larger number of oxygen atoms and the more flexible structure of DB24C8. Regarding the complexation in nitromethane as a reference, the complex stability and selectivity in S are discussed. The selectivities of these crown ethers in water, methanol, and N,N-dimethylformamide, which apparently obey the size-fit concept, are largely due to the solvation of the free alkali metal ions.  相似文献   

3.
Colyophilization or codrying of subtilisin Carlsberg with the crown ethers 18-crown-6, 15-crown-5, and 12-crown-4 substantially improved enzyme activity in THF, acetonitrile, and 1,4-dioxane in the transesterification reactions of N-acetyl-L-phenylalanine ethylester and 1-propanol and that of (+/-)-1-phenylethanol and vinylbutyrate. The acceleration of the initial rate, V(0), ranged from less than 10-fold to more than 100-fold. All crown ethers activated subtilisin substantially, which excludes a specific macrocyclic effect from being responsible. The secondary structure of subtilisin was studied by Fourier-transform infrared (FTIR) spectroscopy. 18-Crown-6 and 15-crown-5 led to a more nativelike structure of subtilisin in the organic solvents employed when compared with that of the dehydrated enzyme obtained from buffer alone. However, the high level of activation with 12-crown-4 where this effect was not observed excluded overall structural preservation from being the primary cause of the observed enzyme activation. The conformational mobility of subtilisin was investigated by performing thermal denaturation experiments in 1,4-dioxane. Although only a small effect of temperature on subtilisin structure was observed for the samples prepared with or without 12-crown-4, both 18-crown-6 and 15-crown-5 caused the enzyme to denature at quite low temperatures (38 degrees C and 56 degrees C, respectively). No relationship between this property and V(0) was evident, but increased conformational mobility of the protein decreased its storage stability. The possibility of a "molecular imprinting" effect was also tested by removing 18-crown-6 from the subtilisin-18-crown-6 colyophilizate by washing. V(0) was only halved as a result of this procedure, an effect insignificant compared with the ca. 80-fold rate enhancement observed prior to washing in THF. This suggests that molecular imprinting is likely the primary cause of subtilisin activation by crown ethers, as recently suggested.  相似文献   

4.
Substituted η3-butadienyl complexes containing amide-armed crowns (X) of general formula [MoCl(CO)23-CH2C(COX)CCH2)(phen)]n (phen=1,10-phenanthroline) were prepared and investigated for their ability to extract alkali metal ions from a mixed phase system. Reaction of the chlorocarbonyl precursor (1) with 1-aza-15-crown-5, 4-aminobenzo-15-crown-5, 2-aminomethyl-15-crown-5, 4-aminobenzo-18-crown-6 or 2-aminomethyl-18-crown-6 gave monomeric complexes (n=1), and addition of sodium tetraphenylboron to the 15-crown-5-substituted complexes gave the corresponding sodium salts. Dinuclear complexes (n=2) were formed by reaction of 1 and 1,7-diaza-15-crown-5 or 4,4(5)-diaminobenzo-15-crown-5. Comparison of amidobenzo- and 2-amidomethyl-15-crown-5-substituted complexes showed enhanced sodium transport properties for the latter, and spectroscopic and molecular modeling studies suggested complexation occurred by concerted action of the amide and crown.  相似文献   

5.
Dibenzo-16-crown-4 (1) indicates high silver and thallium(I) ion selectivity over sodium, potassium, and rubidium ion evaluated from the solvent extraction of metal picrates, while its cation-binding ability is lower than those of dibenzo-18-crown-6 (2) and dibenzo-22-crown-6 (3). Taking account of the highest thallium(I) ion selectivity for 1 obtained from extraction experiments, PVC membrane thallium(I)-selective electrodes based on 1 are prepared. The electrode shows the best potentiometric selectivity coefficients for thallium(I) over potassium and rubidium than those of 2 and 3, and commercially available bis(crown ether)s (4).  相似文献   

6.
The fluorescence decay kinetics from a benzonitrile solution of a dibenzo-18-crown-6 ether bridged porphyrin-fullerene dyad has been studied in the presence of a range of metal ions. Dual-exponential fluorescence decay behaviour has been attributed to conformational flexibility of the molecule influencing quenching interactions between the photo-excited porphyrin and fullerene. Additions of sodium, potassium and lithium ions significantly modulate the observed fluorescence decay processes while the larger tetrabutylammonium ion has only a minor affect. The results are discussed in terms of ion inclusion within the crown ether affecting both the bridge conformational properties and donor-acceptor electronic interactions.  相似文献   

7.
Reaction of the zwitterionic imidazolium salt [(CH2COOH)(CH2COO)im] with K2CO3 or BaO in the presence of 18-crown-6 affords the salts [(CH2COO)2im][K(18-crown-6)] and [(CH2COO)2im]2[Ba(18-crown-6)], respectively. Recrystallisation of these crown complexes from the ionic liquid 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)amide, [emim][Tf2N], at a water interface, results in the formation of new salts in which the original anion is replaced by Tf2N. Single crystal X-ray diffraction has been performed on two of the salts. Notably, the potassium structure containing 18-crown-6 and Tf2N forms a linear chain coordination polymer that can be regarded as metal organic frameworks (MOFs). Moreover, this study provides insights into the separation of group I and II metal ions using crown ethers in combination with ionic liquids.  相似文献   

8.
《Inorganica chimica acta》2006,359(7):2107-2114
Phosphodiesterolytic activity of samarium complexes containing crown ethers and amino acids was systematically studied. Formation constants of mixed ligand Sm–crown ethers–amino acids complexes (crown ethers = 18-crown-6, 15-crown-5 and 12-crown-4 and amino acids = Gly and Arg) were determined at 37.0 °C and 0.50 M NMe4Cl. Kinetics of the hydrolysis of BNPP (bis(4-nitrophenyl)phosphate) mediated by lanthanide(III)-mixed ligands complexes was studied under the same experimental conditions. The rate of BNPP cleavage is sensitive to metal ion concentration, pH, and ligand to metal molar ratio. Hydrolysis follows Michaelis–Menten-type saturation kinetics. High pH values markedly increase the observed activity. Potentiometric titrations results together with kinetic data of all these systems, under identical conditions, allowed us to identify the active species towards hydrolysis. Complexes with phosphodiesterolytic activity are monomeric hydroxylated cationic species. In general, a good phosphodiesterolytic activity is observed for these complexes under similar conditions to the physiological ones.  相似文献   

9.
Studies on the mechanism of crown-ether-induced activation are described in this paper. Michaelis Menten kinetics of -chymotrypsin in toluene in the presence and absence of 18-crown-6 showed that only Vmax is increased upon crown ether treatment. Parallel Lineweaver–Burk plots indicate that crown ethers do not activate the enzyme by specific interactions in the active site, such as transition state stabilization or facilitated transport of water molecules. Increased Vmax values of crown-ether-treated enzyme most probably originate from conformational changes, which alter kcat as well as the amount of catalytically active enzyme.  相似文献   

10.
Some crown ethers translocate cations across the liposomal membrane either by a carrier mechanism or by forming ion channels. We report formation of ion channels in lipid bilayer membranes by bis[(benzo-15-crown-5)-15-yl methyl] pimelate, a crown ether known to form ion inclusion complexes with alkali metal cations. The channels have characteristic long openings lasting several seconds and a low conductance (4 pS in 500 mM KCl and 2.5 pS in 500 mM NaCl). A model of the crown ether channel formed by stacking of four monomers is proposed. A large database of structural information on crown ethers and their ion inclusion complexes as well as large family of crown ethers with a variety of substitutions in the ring are commercially available. Thus the crown ether channel is an attractive model system to study the role of various chemical moieties in ion conduction which may provide deeper insight into understanding the mechanism(s) of selectivity, ion transport, etc. in biological ion channels.  相似文献   

11.
A quick, efficient and convenient method for the regiospecific reductive ring opening of 4,6-O-benzylidene acetals of O-/S-alkyl/aryl glycosides of mono- and disaccharides, leading to the exclusive formation of the corresponding 6-O-benzyl ethers, using sodium cyanoborohydride in the presence of molecular iodine, is reported. It has been observed that common protecting groups such as ethers and esters are well tolerated under the conditions studied. The reaction was proved unsuccessful when applied to a glucosamine-derived benzylidene acetal.  相似文献   

12.
Structural analogues of phospholipidic platelet activating factor, (2-acetoxy-3-octadecyloxy)propyl-1-phosphonocholine and (2-methoxy-3-octadecyloxy)propyl-1-phosphonocholine, were synthesized. High efficiency of the polymer-bound dibenzo-18-crown-6-ether as the O-alkylation catalyst was demonstrated. Reaction of allyl-octadecyl ether with methanol and iodine in the presence of zinc oxide was shown to give a mixture of 1-iodo-2-methoxy- and 1-methoxy-2-iodoprop-3-yl ethers of octadecanol in the 3:1 ratio.  相似文献   

13.
The crowned coumarin complexes are well known compounds for their ion recognition abilities. They undergo photophysical changes upon cation binding. On the basis of density functional theory calculations, we examined the sodium cation (Na+) binding energies of coumarin-crown ethers based on 15-Crown-5 (15 C5) and 18-Crown-6 (18 C6) as well as the optical absorptions of coumarin-crown ethers based on 12-Crown-4 (12 C4), 15 C5 and 18 C6. We explored why the attachment of crown ether ring to coumarin affects the Na+ binding energies of coumarin-crown ethers and also why the optical absorption of coumarin is modified by the crown ethers. Our study reveals that the Na+ ion binding energies of coumarin-crown ethers depend strongly on the size of the crown ether ring and also on the attachment position of the ether ring on coumarin. These factors affect the intramolecular charge transfer and overall stability of the complexes. The absorptions of the coumarin and ether ring parts of coumarin-crown ether are red shifted from those of isolated coumarin and crown ether, respectively. The red-shift of the coumarin ester group absorption is much stronger depending on the attachment position of the ether ring to coumarin. The absorption intensity of the coumarin part in coumarin-crown ethers is reduced for the benzene group absorption, but is enhanced for the ester group absorption.
Figure
In this study, the several Na+ - coumarin-crown ether coordination complexes based on 12-Crown-4, 15-Crown-5 and 18-Crown-6 were examined by the density functional theory method. The structural effects on the Na+ binding energies and optical properties were studied.  相似文献   

14.
The role of relativistic effects (RE) in the structures of Cd(II) complexes with crown ethers, and the reason the ‘soft’ Cd(II) strongly prefers to bind to SCN through N, are considered. The synthesis and structures of [Cd(18-crown-6)(thiourea)2] (ClO4)2.18-crown-6 (1) and [Cd(Cy2-18-crown-6)(NCS)2] (2) are reported. (18-crown-6 = 1,4,7,10,13,16-hexaoxacyclooctadecane; Cy2-18-crown-6 = cis-anti-cis-2,5,8,15,18,21-hexaoxatricylo[20.4.0.0(9,14)]hexacosane). In 1 Cd is coordinated in the plane of the crown which has close to D3d symmetry, with long Cd-O bonds averaging 2.688 Å. The two thiourea molecules form relatively short Cd-S bonds that average 2.468 Å, with an S-Cd-S angle of 164.30°. This structure conforms with the idea that Cd(II) can adopt a near-linear structure involving two covalently-bound donor atoms (the S-donors) with short Cd-S bonds, which resembles gas-phase structures for species such as CdCl2. The structure of 2 is similar, with the two SCN ligands N-bonded to Cd, with short Cd-N bonds of 2.106 Å, and N-Cd-N angle of 180°. The crown in 2 forms long Cd-O bonds that average 2.698 Å. Molecular mechanics calculations suggest that a main reason Cd(II) prefers to bind to SCN through N is that when bound through S, the small Cd-S-C angle, which is typically close to 100°, brings the ligand into close contact with other ligands present, and causes steric destabilization. In contrast, the Cd-N-C angles for SCN coordinated through N are much larger, being 171.4° in 2, which keeps the SCN groups well clear of the crown ether. DFT (density functional theory) calculations are used to generate the structures of [Cd(18-crown-6)(H2O)2]2+ (3) and [Cd(18-crown-6)Cl2] (4). In 3, the Cd(II) is bound to only three O-donors of the macrocycle, with Cd-O bonds averaging 2.465 Å. The coordinated waters form an O-Cd-O angle of 139.47°, with Cd-O bonds of 2.295 Å. In contrast, for 4, the Cd is placed centrally in the cavity of the D3d symmetry crown, with long Cd-O bonds averaging 2.906 Å. The Cl groups form a Cl-Cd-Cl angle of 180°, with short Cd-Cl bonds of 2.412 Å. With ionically bound groups on the axial sites of[Cd(18-crown-6)X2] complexes, such as with X = H2O in 3, the Cd(II) does not adopt linear geometry involving the two X groups, with long Cd-O bonds to the O-donors of the macrocycle. With covalently-bound X = Cl in 4, short Cd-Cl bonds and a linear [Cl-Cd-Cl] unit results, with long Cd-O bonds to the crown ether.  相似文献   

15.
A series of crown ethers involving lauryl glucoside were synthesized and their assembly behavior in water was studied. The synthesis applied a simple protection scheme based on benzylidenation for the glycolipid, and cation templating for the macrocycle. A sequential build-up of the crown ether by bis-hydroxylethylation of the glucoside followed by reaction with di-, tri-, or tetraethylene glycol ditosylate provided better yields of the macrocycle compared to a single step cyclization with tetraethylene glycole ditosylate. The macrocycles containing up to six oxygens showed significantly higher affinity for sodium than for potassium, while more effective potassium complexation was found for the 21-crown-7 compound. The ion binding affinity leads to a slight but significant increase of the CMC of the crown ether containing surfactant in water upon the addition of sodium electrolyte.  相似文献   

16.
In order to apply the excellent chiral recognition ability of chiral pseudo-18-crown-6 ethers that we developed to chiral separation, we prepared a chiral stationary phase (CSP) by immobilizing a chiral pseudo-18-crown-6-type host on 3-aminopropyl silica gel. A chiral column was prepared by the slurry-packing method in a stainless steel HPLC column. A liquid chromatography system using this CSP combined with the detection by mass spectrometry was used for enantiomer separation of amino compounds. A normal mobile phase can be used on this CSP as opposed to conventional dynamic coating-type CSPs. Enantiomers of 18 common natural amino acids were efficiently separated. The chiral separation observed for amino acid methyl esters, amino alcohols, and lipophilic amines was fair using this HPLC system. In view of the correlation between the enantiomer selectivity observed in chromatography and the complexion in solution, the chiral recognition in host-guest interactions might contribute to this enantiomer separation.  相似文献   

17.
The use of crown ethers for a phase transfer-catalyzed synthesis of heteroaromatic glycosides of N-acetylglucosamine was studied. The solid-liquid system and catalysis by 15-crown-5 were found to provide for both the 100% conversion of α-D-glucosaminyl chloride peracetate and a high reaction rate. The interaction of α-D-glucosaminyl chloride peracetate and oxadiazole and triazole mercapto derivatives capable of thiol-thione tautomerism carried out at room temperature in acetonitrile in the presence of anhydrous potassium carbonate and crown ethers was shown to lead to both S- and N-glucosides. The structures of the compounds synthesized were confirmed by X-ray analysis and 13C and 1H NMR spectroscopy.  相似文献   

18.
The ionophores benzo-18-crown-6 (18-C-6), t-butylbenzo-18-crown-6(TBB) and di-t-butyldibenzo-30-crown-10 (30-C-10) were testedfor their effects on potassium ion absorption in onion rootsegments, and in wheat and mung bean seedlings. Potassium uptake,efflux and transport were progressively reduced in onion rootsegments and seedlings by 18-C-6 over the range 0.1–1.0mM. The effects of TBB (up to 0.3 mM) were more severe but otherwisegenerally similar to those of 18-C-6 in seedlings. Both ionophoresreduced growth, and at the highest concentrations, resultedin root potassium ion content falling below initial values after48 h treatment. The effects of 30-C-10 were evident at muchlower concentrations, inhibition of net potassium uptake occurringabove 10–4 M. Between 10–4 and 10–4 M 30-C-10,however, a modest but significant stimulation of potassium uptakewas observed in onion roots and seedlings; growth of seedlingswas largely unaffected. The reductions in potassium absorptionwere attributed to the promotion, by the ionophores, of facilitateddiffusion down the electrochemical diffusion gradient, counteringthe efficiency of the potassium ion influx pump. Stimulationof uptake at certain concentrations of 30-C-10 was consideredmore likely to be due to an inhibition of passive potassiumefflux, rather than a stimulation of active influx. The importanceof stability constants, bonding and lipophilicity, in determiningthe relative effectiveness of the ionophores, is discussed. Allium cepa L, onion, Triticum aestivum L, wheat, Phaseolus aureus L, mung bean, cyclic ‘crown’ polyethers, potassium fluxes, ion transport  相似文献   

19.
The effects of 18-crown-6 on the synthesis of peptides catalyzed by alpha-chymotrypsin are reported. Lyophilization of the enzyme in the presence of 50 equivalents of 18-crown-6 results in a 425-fold enhanced activity when the reaction between the 2-chloroethylester of N-acetyl-L-phenylalanine and L-phenylalaninamide is carried out in acetonitrile. Addition of crown ether renders the dipeptide synthesis in nonaqueous solvents catalyzed by alpha-chymotrypsin possible on a preparative scale. The acceleration is observed in different solvents and for various peptide precursors. Copyright 1998 John Wiley & Sons, Inc.  相似文献   

20.
New complexes LnI2·18-crown-6 (Ln-Sm, Tm, Dy, Nd) and LnJ2·dibenzo-18-crown-6 (Ln-Sm, Tm) were synthesized using the solutions of LnI2 in THF. The compounds obtained oxidize quickly in air, but are relatively stable in an inert atmosphere. The Tm2+ complex is decomposed by light. The compounds obtained are poorly soluble in THF, the Sm2+ and Tm2+ compounds are soluble in CH3CN, forming solutions with a period of half oxidation of 170 h and 6 min, respectively. Iodide ions of the complexes can be substituted for Cl? during treatment of the compounds by solution of LiCl in THF. The reflection spectra of the compounds synthesized are similar to the absorption spectra of Ln2+ in THF, although a shift of bands towards the short wave region is observed.The study of the Ln2+ oxidation kinetics in H2O, CH3CN, THF in the presence of crown ethers has shown that their stability is influenced not only by the type of solvent, relative solubility and stability of complexes Ln2+ and Ln3+, but also by phenyl groups, and by decreasing stability of Dy2+ and Nd2+.  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