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 484 毫秒
1.
Four derivatives of acridine and acridinium compounds (L1, L2, L1H and L2H) comprised thiourea-binding sites were synthesised. The binding abilities of receptors L1, L2, L1H and L2H towards amino acids (l-Trp, l-Phe, l-Leu, l-Ala and l-Gly) were studied by 1H NMR spectroscopy, UV–vis and fluorescence spectrophotometry. Hydrogen bonding interactions between thiourea-binding site of the ligand and the carboxylate groups in zwitterionic amino acids were found to be the main interactions driving complexation to take place. The stoichiometry of 1:1 ligand to amino acid was observed in all cases. Neutral ligands L1 and L2 showed weak binding towards all studied amino acids. The cyclic ligand L1 showed better binding ability towards tryptophan (Trp) than the acyclic ligand L2 did (K for Trp is 307 and 266 M? 1 for L1 and L2, respectively). Interestingly, binding abilities of the protonated ligands, L1H and L2H, towards studied amino acids, especially Trp (K for Trp is 3157 and 2873 M? 1 for L1H and L2H, respectively), were increased due to R–COO…H…N+–acridinium interactions. Calculated structures of L1H·Trp and L2H·Trp showed that the polyglycol moiety in L1H provided a hydrophobic cavity for binding Trp resulting in a stronger binding affinity of L1H over L2H.  相似文献   

2.
A 1H NMR spectroscopic study of tetra-n-butylammonium halides (TBAX: X = Cl, Br or I) in CDCl3 solutions was conducted. Complexation studies of TBAX salts with different host molecules using 1H NMR in CDCl3 have previously revealed that the reference residual CHCl3 proton signal had been shifted downfield. The aim of the study was to quantify the extent of these chemical shift changes with TBAX salts. Linear concentration–chemical shift relationships in each case were obtained from the resulting titration plots obtained from the addition of the TBAX salts alone to CDCl3. Interactions in the solid state as determined by X-ray crystallography support the solution-state investigations indicating halide ion–chloroform proton interactions.  相似文献   

3.
The effect of cationic micelles of cetyltrimethyl ammonium bromide (CTAB) on the observed pseudo-first-order rate constant for the interaction of nickel dipeptide complex [Ni(II)-Gly-Gly]+ with ninhydrin has been studied spectrophotometrically. At constant temperature and pH, increase in the [CTAB] from 0.0 to 60.0 × 10?3 mol dm?3 caused nearly three-fold increase of the rate constant. The micellar catalysis is explained in terms of the pseudophase model. From the observed kinetic data, binding constants of micelle–[Ni(II)-Gly-Gly]+ (K S), and micelle–ninhydrin (K N) are evaluated, respectively, to be 5.3 mol?1 dm3 and 84.0 mol?1 dm3. The role of added inorganic (NaCl, NaBr, Na2SO4) and organic salts (NaBenz, NaSal) on the reaction rate has also been examined.  相似文献   

4.
The complexation of N-phthaloyl, N-formyl, and N,N-dimethyl derivatives of S-methylcysteine methyl ester (both racemic and optically pure) with three dimeric rhodium(II) salts, acetate Rh2AcO4, trifluoroacetate Rh2TFA4, and (R)-(+)-α-methoxy-α-trifluoromethylphenylacetate Rh2Mosh4 was investigated by nuclear magnetic resonance spectroscopy (NMR) at room and lower temperatures. The complexation was carried out in situ, in CDCl3 solution using titration procedure; the results were examined by the analysis of 1H and 13C NMR chemical shift change (Δδ). The complexation of free S-methyl cysteine and hydrochloride salt of its methyl ester was performed in D2O solution. For comparison, complexation of some derivatives of leucine, phenylalanine, and proline was examined.

N-phthaloyl and N-formyl derivatives of cysteine formed 1 : 1 and 1 : 2 axial complexes with all dirhodium salts. Rhodium substrates were bonded via sulfur. In one case, the complexation of Rh2TFA4 by both sulfur and N-formyl oxygen was noted. Similar complexation of Rh2TFA4, via CHO group, was found for N-formyl derivatives of leucine, phenylalanine, and proline. For N,N-dimethyl derivative of cysteine, both N and S atoms were involved in bonding. At room temperature, in all cases, ligand exchange was fast on the NMR timescale.  相似文献   

5.
Molecular salts and cocrystals of amino acids have potential applications as molecular materials with nonlinear optical, ferroelectric, piezoelectric, and other various target physical properties. The wide choice of amino acids and coformers makes it possible to design various crystal structures. The amino acid–maleic acid system provides a perfect example of a rich variety of crystal structures with different stoichiometries, symmetries and packing motifs built from the molecular building blocks, which are either exactly the same, or differ merely by protonation or as optical isomers. The present paper reports the crystal structures of two new salts of the dl ‐norvaline–maleic acid system with 1:1 and 2:1 stoichiometries, namely dl ‐norvalinium hydrogen maleate, C5H12NO2+·C4H3O4, (I), and dl ‐norvalinium hydrogen maleate–dl ‐norvaline, C5H12NO2+·C4H3O4·C5H11NO2, (II). These are the first examples of molecular salts of dl ‐norvaline with an organic anion. The crystal structure of (I) has the same C 22(12) structure‐forming motif which is common for hydrogen maleates of amino acids. The structure of (II) has dimeric cations. Of special interest is that the single crystals of (I) which are originally formed on crystallization from aqueous solution transform into single crystals of (II) if stored in the mother liquor for several hours.  相似文献   

6.
Novel N‐sulfonylated amino alcohols were synthesized from L‐amino acids and (+)‐camphor, and their application to asymmetric reduction of prochiral ketones with NaBH4–BF3 · Et2O is described.  相似文献   

7.
Two non-natural amino acids with fluorescent urea side-chains were prepared from Fmoc-protected aspartic and glutamic acids. In acetonitrile solution, the emission of the Asp derivative is strongly quenched by HCO3 or H2PO4 (K ≥ 104 M? 1) but not by less-basic Cl or NO3. Solutions containing excess bicarbonate ion appear peach-colored, with λabs at 394 and 495 nm ascribed to the anion complex and urea-deprotonated sensor, respectively. Corresponding fluorescence bands are observed at 475 and 579 nm. Dihydrogenphosphate is not sufficiently basic to remove H+ from the ground state of the fluorophore. However, deprotonation of the excited state occurs in the presence of>1 equiv of H2PO4em = 578 nm). According to 1H NMR in DMSO-d6, recognition of H2PO4 occurs at the urea N–H groups and the amino acid backbone N–H. DFT techniques further predict that the backbone C = O group accepts an H-bond from the anion. The Glu derivative has lower affinity for anions; the additional CH2 group in its side-chain apparently sets the backbone N–H and C = O too far from the urea to contribute significantly to binding. To demonstrate suitability for standard Fmoc-based solid-phase peptide synthesis, the Asp derivative was incorporated into a 12-residue peptide.  相似文献   

8.
Gao  Xiaokang  Li  Qian  Zhao  Xinfeng  Huang  Jingjing  Bian  Liujiao  Zheng  Jianbin  Li  Zijian  Zhang  Youyi  Zheng  Xiaohui 《Chromatographia》2014,77(17):1235-1239

The interaction between drugs and receptors is particularly important in revealing the drug acting mechanism and developing new leads. In this work, α 1-Adrenoceptor (α 1-AR) from HEK293 cell line is purified and immobilized on the surface of macro-pore silica gel to prepare an high-performance affinity chromatography stationary phase for the pursuit of drug–receptor interactions by competition zonal elution. Naftopidil is found to have only one type of binding site to α 1-AR with an association constant of 1.45 × 106 M−1 and a concentration of binding sites of 1.56 × 10−6 M, while terazosin hydrochloride proves to present two kinds of binding site on the receptor at which the association constants are determined to be 1.61 × 105 M−1 and 2.06 × 103 M−1, and the corresponding concentrations of the binding sites are 1.56 × 10−6 M and 1.11 × 10−3 M, respectively. It is concluded that the stationary phase containing attached α 1-AR can be used to realize the binding of a drug to the receptor.

  相似文献   

9.
A La(III) complex, [LaIIICl2(NOR)2]Cl (2), containing norfloxacin (NOR) (1), a synthetic fluoroquinolone antibacterial agent, has been synthesized and characterized by elemental analysis, IR, UV–vis spectra and 1H NMR spectroscopy, and molar conductance measurements. The interaction between 2 and CT-DNA was investigated by steady-state absorption and fluorescence techniques in different pH media, and showed that 2 could bind to CT-DNA presumably via non-intercalative mode and the La(III) complex showed moderate ability to bind CT-DNA compared to other La(III) complexes. The binding site number n, and apparent binding constant KA, corresponding thermodynamic parameters ΔG#, ΔH#, ΔS# at different temperatures were calculated. The binding constant (KA) values are 0.23 ± 0.05, 0.56 ± 0.05, and 0.18 ± 0.08 × 105 L mol?1 for pH 4, 7, and 11, respectively. It was also found that the fluorescence quenching mechanism of CT-DNA by La(III) complex was a static quenching process.  相似文献   

10.
Jiang  Hong  Ding  Jinghua  Chang  Pei  Chen  Zaixing  Sun  Guifan 《Chromatographia》2010,71(11):1075-1079

Study on the stoichiometry and affinity of the arsenicals bound to HSA is an important step toward a better understanding of arsenic toxic effects. After incubation of AsIII or AsV with HSA at the physiological conditions (pH 7.43 and 37 °C), the free arsenicals and arsenic-HSA complexes were separated and detected by the combined techniques of microdialysis and liquid chromatography with hydride generation atomic fluorescence spectroscopy (MD–LC–HGAFS). The decrease of AsIII peak response rather than AsV indicated that HSA reacted with AsIII but not AsV. The binding plots indicated that the binding between HSA and AsIII was in Scatchard pattern when the concentration ratios of AsIII to HSA were ≤1:1. The strong binding sites (n 1) were 1.6 and the stability constant (K 1) was 1.54 × 106 M−1. When the concentration ratios of AsIII to HSA were >1:1, the binding was in Plasvento pattern with the stability constant K 2 ≅ 0 and no specific binding of AsIII with HSA. On the contrary, AsV did not show binding with HSA. The results showed that AsIII reacted with HSA more readily than AsV, which provides a chemical basis for arsenic toxicity.

  相似文献   

11.
The antibiotic valinomycin is a potassium-selective ionophore, which increases the transport of potassium ions across cell membranes and thereby causes damage to bacteria cells. Valinomycin has been extensively studied as an ionophore for cations. We report for the first time the direct binding of anions to valinomycin using electrospray ionisation mass spectrometry and 1H nuclear magnetic resonance (NMR) spectroscopy. The binding selectivity for halide ions is found to be in the order Cl>Br~F?I based on electrospray ionisation mass spectrometry experiments in methanol. 1H NMR studies in acetone-d 6 and CD3CN reveal the binding selectivity of Cl>Br?F~I. NMR studies and density functional theory (DFT) calculations support a bracelet-like structure for the binding of a chloride ion to valinomycin. Association constants of 531 ± 45 M? 1 and 57 ± 2 M? 1 were obtained via NMR titrations in acetone-d 6 for chloride and bromide ions, respectively.  相似文献   

12.
The non-aggregational cyclo[6]aramide has demonstrated 2:1 host–guest complexation towards diquat with very strong binding ability (K1 = 5.41 × 104 M? 1, K2 = 4.33 × 106 M? 1). The donor–acceptor binding process of the macrocycle and the quaternary salt was investigated by 1H NMR, ESI mass spectrometry and UV–vis spectroscopy. The binding mode is supported by both experiments and theoretical simulations. This work provides the first example of using recently developed H-bonded aromatic oligoamide macrocycles for binding diquat in solution.  相似文献   

13.
Two iron(II) complexes, [FeII(pytBuN3)2](FeCl4) (1) and [FeII(pytBuMe2N3)Cl2] (2), with sterically constrained pytBuN3 and pytBuMe2N3 chelate ligands (pytBuN3 = 2,6-bis-(aldiimino)pyridyl; pytBuMe2N3 = 2,6-bis-(ketimino)pyridyl), have been synthesized and characterized by elemental analysis, IR, UV–vis spectra, and preliminary X-ray single-crystal diffraction. The latter revealed that Fe(II) in 1 is six-coordinate by six nitrogen donors from two bisiminopyridines in a distorted octahedron. Complex 2 reacts with thiourea with a second-order rate constant k2 = (2.50 ± 0.05) × 10?3 M?1 s?1 at 296 K, and the reaction seemed to be slow. In a similar way, the interaction of 2 and DNA was studied by fluorescence and absorption spectroscopy. The results revealed that 2 caused fluorescence quenching of DNA through a dynamic quenching procedure. The binding constants KA, Kapp, and KSV as well as the number of binding sites between 2 and DNA were determined.  相似文献   

14.
In this article, we describe the influence of structure on biological behavior of amino acid-Pd complex and compare it with oxalipalladium. A new water-soluble oxalipalladium analog with formula of [Pd(DACH)(isopentylgly)](NO3), where DACH is 1R,2R-diaminocyclohexane, has been synthesized and characterized by elemental analysis, conductivity measurements, IR, UV–Vis, and 1H NMR spectroscopies. The interactions of oxalipalladium and its amino acid derivative with highly polymerized calf-thymus DNA have been extensively studied by spectroscopic methods. The high binding constants of oxalipalladium (0.38 × 104 M?1) and new amino acid-Pd complex (0.65 × 104 M?1) were determined using absorption measurements. Also circular dichroism (CD) studies show that Pd complex causes more disturbances on DNA structure rather than oxalipalladium. The experimental results proposed that [Pd(DACH)(isopentylgly)](NO3) is bound to DNA by groove-binding mode as well as partially covalent interaction, while oxalate analog binds covalently to DNA after hydrolysis. Interaction of the two metal derivative complexes was studied by molecular docking simulation. The results showed that amino acid-Pd complex has higher negative docking energy and higher tendency for interaction with DNA, and exert more structural change on DNA. Finally, the anticancer and growth inhibitory activities of synthesized complexes were investigated against human colon cancer cell line of HCT116 after 24 h incubation time using MTT assay. Results show that the complex [Pd(DACH)(isopentylgly)](NO3) showed enhanced anticancer and growth inhibitory activities against human colon can cell line HCT116.  相似文献   

15.
Schiff base pyridin-2-ylimino methyl naphthanol (HL) was synthesized and characterized by spectroscopic (FTIR, ESIMS, and NMR) techniques. The ligand was reacted with perchlorate salts of Mn+2, Co+2, and Ni+2. ESIMS mass spectra indicate the formation of mononuclear complex ML2 for all three complexes. CoL2 crystallizes in P21/n space group, adopting a distorted tetrahedral geometry where Co is in a N2O2 donor environment. Structure of the Co complex was optimized by DFT calculation. Solution-phase complexation between the ligand and the three metals ions: Mn+2, Co+2, and Ni+2 (pH 7.2 in tris buffer), in CH3CN–H2O was performed spectrophotometrically by UV–vis spectral study. Job’s plot from each titration suggests a 1 : 2 metal to ligand combination. The association constants for the formation of ML2 are as follows: Mn (19.80 × 103 M?1), Co (14.54 × 103 M?1) and Ni (19.04 × 103 M?1).  相似文献   

16.
[Pd(BHEP)Cl2] (BHEp = 1,4-bis(2-hydroxyethyl)piperazine) was synthesized and characterized. The palladium center has a typical square-planar geometry with a tetrahedral distortion. The alcohol groups of the ligand do not participate in binding to Pd(II). The DFT/B3LYP method was used for geometric optimization of the ligand and the complex using the Gaussian 09 program and compared with experimental results. The stoichiometry and stability constants of the complexes formed between [Pd(BHEP)(H2O)2]2+ and some selected amino acids, peptides, and DNA constituents were investigated at 25 °C and 0.1 M ionic strength. The binuclear complex [(H2O)(BHEP)Pd(Bip)Pd(BHEP)(H2O)]4+ was detected, where Bip = 4,4′-bipiperidine. Inosine, uracil, and thymine interact with the binuclear complex via substitution of both coordinated water molecules. The potentiometric results were complimented by spectroscopic measurements. The concentration distribution diagrams of the various species formed were evaluated.  相似文献   

17.
1,3-Bis(α-picolyloxy)-p-tert-butylcalix[4]crown-5 in the cone conformation (2), 1,8-diiodoperfluorooctane or 1,6-diiodoperfluorohexane, and potassium iodide ternary mixtures undergo in solution self-sorting and afford crystalline “supramolecular salts”. These hybrid materials consist of supercation [K+ ? 2] and superanion [I–(CF2) n –I…I…I–(CF2) n –I…I…] (n = 6,8) components. In the supercations the potassium ion is embedded in the ionophoric pocket created by the heteroatoms present at the lower rim. In the superanions the iodide ions form infinite fluorous polyanionic chains as a result of a self-assembly process which relies on halogen bonding. Both cation encapsulation and anion-perfluorocarbon halogen bonding were detected in solution by 1H and 19F NMR, and in the gas phase by ESI MS.  相似文献   

18.
Two new complexes, [Co(L)2]Cl·(MeOH)2 (1) and [Ni(L)2]4·EtOH (2) (L?=?(E)-2-(amino((pyridin-2-ylmethylene)amino)methylene)maleonitrile), were synthesized and characterized by X-ray crystallography, IR, UV, and fluorescence spectroscopy. According to X-ray crystallographic studies, each metal was six-coordinate with six nitrogens from two ligands. Both complexes form two-dimensional supramolecular networks via hydrogen bonding and π–π interactions. Ultraviolet and visible spectra showed that absorptions arise from π–π ?, MLCT, and dd electron transitions. Fluorescence spectroscopy revealed moderate intercalative binding of these two complexes with EB–DNA, with apparent binding constant (K app) values of 9.14?×?105 and 3.20?×?105?M?1 for Co(III) and Ni(II) complexes, respectively. UV–visible absorption spectra showed that the absorption of DNA at 260?nm was quenched for 2 but quenched then improved for 1 with addition of complexes, tentatively attributed to the effect of the combined intercalative binding and electrostatic interaction for 1.  相似文献   

19.
The supramolecular interactions of the ocular drug tropicamide (TR) with cucurbit[7]uril (CB7) and cucurbit[8]uril (CB8) were investigated in aqueous solutions by using 1H NMR, ESI-MS and UV–vis spectroscopic techniques. The results indicate a 1:1 binding stoichiometry of TR with CB7 and CB8. The binding constants of TR in its protonated form were higher (e.g. K = 4 × 106 M? 1 with CB8) than in its neutral form (e.g. K = 1.4 × 104 M? 1 with CB8), which led to a complexation-induced increase in its pK a value of ca. 0.5 and 2 units with CB7 and CB8, respectively. In the presence of about 1% (w/v) CB8, the ionisation degree of 0.1% (w/v) TR was increased from 2% to 62% at neutral pH. The increase in the pK a value and thus stabilisation of the protonated TR species at neutral pH is discussed in the context of supramolecular drug delivery of ophthalmologic drugs.  相似文献   

20.
Complex formation of poly(ethylene oxide) (PEO) with divalent barium and strontium salts was investigated in methanol. In these systems the complexation was accompanied by a considerable degree of ionic association. An analytical model for the polymer-ion complexation based on a one-dimensional lattice model was proposed. According to this model, the electrostatic effects between the bound ions were separated from the total free energy change of the binding. Three binding constants, i.e., the ionic association constant K A, the cation binding constant, K c, and the anion binding constant, K a, could be estimated. K A for barium and strontium salts was comparable, and the effect of counteranions on K A was not large. K c for barium salts was almost independent of the kind of counteranion and larger than that for corresponding strontium salts, indicating stronger polymer-ion interaction for barium salts. The anion binding constant, K a, was strongly dependent upon the kind of anion, and the order was CI? ? ? 4 ?. The pronounced ion binding for larger anions may be explained by the more favorable free energy change of desolvation. Finally, the concentration of free and bound ionic species was determined as a function of PEO concentration.  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