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 31 毫秒
1.
We have synthesized zinc complexes of H2ENTPP (5-(8-ethoxycarbonyl-1-naphthyl)-10,15,20-triphenyl porphyrin) as a model to study hydrogen-bonding interactions. When water or methanol is a ligand, crystals of [Zn(ENTPP)(CH3OH)] or [Zn(ENTPP)(H2O)]?·?C6H5CH3 were obtained. In both structures, the ligand has hydrogen-bonding interactions, but in different patterns. In [Zn(ENTPP)(CH3OH)], the methanol oxygen and carboxylate oxygen in the naphthyl group form an intermolecular hydrogen bond. In [Zn(ENTPP)(H2O)]?·?C6H5CH3, there are two independent molecules A and B. In molecule B, there is an intramolecular hydrogen bond between the water oxygen and the carboxylate oxygen, while in molecule A, besides the intramolecular hydrogen bond, there is an intermolecular hydrogen bond between the water oxygen and the carboxylate oxygen. 1H NMR spectra suggest the binding of methanol or water to zinc are equilibrium processes in solution. Equilibrium constant has been determined by UV-Vis measurements, and it suggests the binding affinity of zinc to methanol has been moderately increased.  相似文献   

2.
 A solvated and cross-linked copolymer of N-isopropylacrylamide (IPAAm) and 2-(acrylamido)-2-methyl propane sulfonic acid (AMPS) was synthesized and its interaction with cationic surfactant lauryl-pyridinium chloride (C12PyCl) was investigated. The solvated copolymer exhibited a lower critical solution temperature (LCST) in water, which was extensively shifted to a higher temperature due to the increase of hydrophilicity introduced by AMPS. In C12PyCl solution, LCST of the copolymer was dramatically decreased due to the binding of C12PyCl to AMPS unit, forming a stoichiometric complex. However, in the concentrated C12PyCl solution, its LCST increased due to the non-stoichiometric complex formation. This phenomenon was further examined in the cross-linked copolymer, analyzed by binding isotherms. Two-step binding of surfactant was demonstrated followed by gel shrinking and re-swelling. This binding mechanism was further discussed regarding the effect of charge density and the hydro-phobicity of the main-chain backbone in terms of electrostatic and hydrophobic interactions. Received: 13 May 1997 Accepted: 13 August 1997  相似文献   

3.
Water molecules in the binding pocket of a protein and their role in ligand binding have increasingly raised interest in recent years. Displacement of such water molecules by ligand atoms can be either favourable or unfavourable for ligand binding depending on the change in free enthalpy. In this study, we investigate the displacement of water molecules by an apolar probe in the binding pocket of two proteins, cyclin-dependent kinase 2 and tRNA-guanine transglycosylase, using the method of enveloping distribution sampling (EDS) to obtain free enthalpy differences. In both cases, a ligand core is placed inside the respective pocket and the remaining water molecules are converted to apolar probes, both individually and in pairs. The free enthalpy difference between a water molecule and a CH3 group at the same location in the pocket in comparison to their presence in bulk solution calculated from EDS molecular dynamics simulations corresponds to the binding free enthalpy of CH3 at this location. From the free enthalpy difference and the enthalpy difference, the entropic contribution of the displacement can be obtained too. The overlay of the resulting occupancy volumes of the water molecules with crystal structures of analogous ligands shows qualitative correlation between experimentally measured inhibition constants and the calculated free enthalpy differences. Thus, such an EDS analysis of the water molecules in the binding pocket may give valuable insight for potency optimization in drug design.  相似文献   

4.
New oxygen carriers have been synthesized by the interaction of the CuII/NiII derivative of the bis(5-nitroindazolyl)methane complex with 14-membered 1,8-dihydro-1,3,6,8,10,13-hexaazacyclotetradecane (M-mac), where M=FeIII, NiII, CoII, to yield binuclear complexes. These complexes have been characterized by physico-chemical methods: elemental analysis, i.r., 1H-n.m.r., 13C-n.m.r., 2D cosy n.m.r., e.p.r., u.v.–vis. spectroscopy and cyclic voltammetry. A representative binuclear FeIII–CuII complex was chosen to interact with H2O2 to elucidate the mechanistic pathway of oxygen binding in solution spectrophotometrically, and also by cyclic voltammetry. Hydrogen peroxide exhibits two mechanisms for binding, either (i) homolysis or (ii) heterolytic cleavage. The mode of H2O2 binding can be hazardous in (i) due to the threat of oxidative damage to the cellular structure, proteins and metabolites, or eco-friendly as in (ii) which produces innocuous products such as water and dioxygen. This study aims to combat the problems associated with H2O2 binding in nature by producing a parallel study on model compounds.  相似文献   

5.
The solubilities of glycine, alanine, phenylalanine, and proline in H2O and D2O from 283 K to 335 K were determined. It was found that glycine and alanine are less soluble in heavy water than in light water but proline is more soluble in heavy water over the whole temperature range studied. Phenylalanine is more soluble in H2O than D2O below 310 K but above that temperature heavy water becomes a better solvent. An influence of H/D isotope substitution on the enthalpies of solution is also observed. In the case of glycine and alanine enthalpies of solution in heavy water increase by a small amount and in the same time the solution enthalpy for phenylalanine in D2O increases markedly. No change in the solution enthalpy for proline was observed. The isotope effects on solubility and the solution enthalpy are qualitatively discussed.  相似文献   

6.
A porous hollow-fiber membrane containing a phenyl group as a hydrophobic ligand was prepared by radiation-induced graft polymerization of glycidyl methacrylate, followed by successive ring-opening reactions with phenol and water. Bovine serum albumin (BSA) was bound to the ligand during permeation of a BSA solution in phosphate buffer containing 2M (NH4)2SO4 through the pores of the hollow fiber. Subsequent elution with an (NH4)2SO4-free buffer exhibited an elution percentage of 82%. Repeated cycles of adsorption and elution caused the accumulation of BSA on the pore surface, resulting in a decrease in the binding capacity of BSA with increasing number of cycles. In contrast, by permeating 1 M NaOH after each elution, the binding capacity of BSA was maintained even after ten cycles. This alkaline regeneration was found to be effective in ensuring repeated use of the phenyl-group-containing porous membrane for recovery of proteins.  相似文献   

7.
《Solid State Sciences》2001,3(4):407-415
Concerning the generation of manganese ferrite by coprecipitation from MnO2 and FeSO47H2O aqueous solution, based on TG, DTG and DTA measurements, IR and mass spectra, SEM and X-ray diffraction data, the authors present an investigation on the elimination of water from the non-aged, aged and air-calcined coprecipitates. Some powders microstructure changes generated by water release through aging in solution or heating in air are presented.  相似文献   

8.
顺铂化合物与鸟嘌呤异构体相互作用的理论研究   总被引:1,自引:0,他引:1  
章志强  周立新  和芹 《中国化学》2005,23(10):1327-1332
The influence of binding of cisplatin adducts on tautomeric equilibrium of guanine was investigated using quantum chemical method. The monoaqua adduct [Pt(NH3)2Cl(H2O)]^+ and the diaqua adduct [Pt(NHa)2(H2O)2]^2+ were chosen for coordination to the N(7) site of guanine tautomers. The results demonstrate that the platinum adducts influence moderately on tautomeric equilibrium, but do not change the relative stability of tautomers whether in gas phase or in aqueous solution. The keto form having H atom at N(1) and N(9) was always the predominant structure when cisplatin adducts were bound to guanine. However, other forms could coexist in water. Meanwhile, our calculations suggest that the tautomeric equilibrium should be via the same intermediate.  相似文献   

9.
The diffusion-limited binding kinetics of antigen (or antibody) in solution to antibody (or antigen) immobilized on a biosensor surface is analyzed within a fractal framework. The fit obtained by a dual-fractal analysis is compared with that obtained from a single-fractal analysis. In some cases, the dual-fractal analysis provides an improved fit when compared with a single-fractal analysis. This was indicated by the regression analysis provided by Sigmaplot (San Rafael, CA). These examples are presented. It is of interest to note that the state of disorder (or the fractal dimension) and the binding rate coefficient both increase (or decrease, a single example is presented for this case) as the reaction progresses on the biosensor surface. For example, for the binding of monoclonal antibody MAb 49 in solution to surface-immobilized antigen, a 90.4% increase in the fractal dimension (Df1 toD f2 ) from 1.327 to 2.527 leads to an increase in the binding rate coefficient (k1 to k2) by a factor of 9.4 from 11.74 to 110.3. The different examples analyzed and presented together provide a means by which the antigen-antibody reactions may be better controlled by noting the magnitude of the changes in the fractal dimension and in the binding rate coefficient as the reaction progresses on the biosensor surface.  相似文献   

10.
A mathematical model of convective mass exchange between dispersed flows in a valve-pulsatory apparatus with permeable partition for carrying out processes with finely dispersed solid phase is proposed. The mass exchange between flows of aqueous Na2CO3 solution and distilled water, or pearlite suspension in water and aqueous solution of NaCl was studied on a pilot apparatus.  相似文献   

11.
Density functional theory calculations are reported on a set of three model structures of the Mn4Ca cluster in the water‐oxidizing complex of Photosystem II (PSII), which share the structural formula [CaMn4C9H10N2O16]q+ ? (H2O)n (q=?1, 0, 1, 2, 3; n=0–7). In these calculations we have explored the preferred hydration sites of the Mn4Ca cluster across five overall oxidation states (S0 to S4) and all feasible magnetic‐coupling arrangements to identify the most likely substrate–water binding sites. We have also explored charge‐compensated structures in which the overall charge on the cluster is maintained at q=0 or +1, which is consistent with the experimental data on sequential proton loss in the real system. The three model structures have skeletal arrangements that are strongly reminiscent, in their relative metal‐atom positions, of the 2.9‐, 3.7‐, and 3.5 Å‐resolution crystal structures, respectively, whereas the charge states encompassed in our study correspond to an assignment of (MnIII)3MnII for S0 and up to (MnIV)3MnIII for S4. The three models differ principally in terms of the spatial relationship between one Mn (Mn(4)) and a generally robust Mn3Ca tetrahedron that contains Mn(1), Mn(2), and Mn(3). Oxidation‐state distributions across the four manganese atoms, in most of the explored charge states, are dependent on details of the cluster geometry, on the extent of assumed hydration of the clusters, and in some instances on the imposed magnetic‐coupling between adjacent Mn atoms. The strongest water‐binding sites are generally those on Mn(4) and Ca. However, one structure type displays a high‐affinity binding site between Ca and Mn(3), the S‐state‐dependent binding‐energy pattern of which is most consistent with the substrate water‐exchange kinetics observed in functional PSII. This structure type also permits another water molecule to access the cluster in a manner consistent with the substrate–water interaction with the Mn cluster, seen in electron spin‐echo envelope modulation (ESEEM) studies of the functional enzyme in the S0 and S2 states. It also rationalizes the significant differences in hydrogen‐bonding interactions of the substrate water observed in the FTIR measurements of the S1 and S2 states. We suggest that these two water‐binding sites, which are molecularly close, model the actual substrate‐binding sites in the enzyme.  相似文献   

12.
Studies of protein−ligand binding often rely on dissolving the ligand in dimethyl sulfoxide (DMSO) to achieve sufficient solubility, and then titrating the ligand solution into the protein solution. As a result, the final protein−ligand solution contains small amounts of DMSO in the buffer. Here we report how the addition of DMSO impacts studies of protein conformational dynamics. We used 15N NMR relaxation to compare the rotational diffusion correlation time (τC) of proteins in aqueous buffer with and without DMSO. We found that τC scales with the viscosity of the water−DMSO mixture, which depends sensitively on the amount of DMSO and varies by a factor of 2 across the relevant concentration range. NMR relaxation studies of side chains dynamics are commonly interpreted using τC as a fixed parameter, obtained from backbone 15N relaxation data acquired on a separate sample. Model-free calculations show that errors in τC, arising from mismatched DMSO concentration between samples, lead to significant errors in order parameters. Our results highlight the importance of determining τC for each sample or carefully matching the DMSO concentrations between samples.  相似文献   

13.
Valinomycin is a naturally occurring cyclic dodecadepsipeptide with the formula cyclo‐[d ‐HiVA→l ‐Val →l ‐LA→l ‐Val]3 (d ‐HiVA is d ‐α‐hydroxyisovaleic acid, Val is valine and LA is lactic acid), which binds a K+ ion with high selectively. In the past, several cation‐binding modes have been revealed by X‐ray crystallography. In the K+, Rb+ and Cs+ complexes, the ester O atoms coordinate the cation with a trigonal antiprismatic geometry, while the six amide groups form intramolecular hydrogen bonds and the network that is formed has a bracelet‐like conformation (Type 1 binding). Type 2 binding is seen with the Na+ cation, in which the valinomycin molecule retains the bracelet conformation but the cations are coordinated by only three ester carbonyl groups and are not centrally located. In addition, a picrate counter‐ion and a water molecule is found at the center of the valinomycin bracelet. Type 3 binding is observed with divalent Ba2+, in which two cations are incorporated, bridged by two anions, and coordinated by amide carbonyl groups, and there are no intramolecular amide hydrogen bonds. In this paper, we present a new Type 4 cation‐binding mode, observed in valinomycin hexaaquamagnesium bis(trifluoromethanesulfonate) trihydrate, C54H90N6O18·[Mg(H2O)6](CF3SO3)2·3H2O, in which the valinomycin molecule incorporates a whole hexaaquamagnesium ion, [Mg(H2O)6]2+, via hydrogen bonding between the amide carbonyl groups and the hydrate water H atoms. In this complex, valinomycin retains the threefold symmetry observed in Type 1 binding, but the amide hydrogen‐bond network is lost; the hexaaquamagnesium cation is hydrogen bonded by six amide carbonyl groups. 1H NMR titration data is consistent with the 1:1 binding stoichiometry in acetonitrile solution. This new cation‐binding mode of binding a whole hexaaquamagnesium ion by a cyclic polypeptide is likely to have important implications for the study of metal binding with biological models under physiological conditions.  相似文献   

14.
Metal coordination to N9‐substituted adenines, such as the model nucleobase 9‐methyladenine (9MeA), under neutral or weakly acidic pH conditions in water preferably occurs at N1 and/or N7. This leads, not only to mononuclear linkage isomers with N1 or N7 binding, but also to species that involve both N1 and N7 metal binding in the form of dinuclear or oligomeric species. Application of a trans‐(NH3)2PtII unit and restriction of metal coordination to the N1 and N7 sites and the size of the oligomer to four metal entities generates over 50 possible isomers, which display different feasible connectivities. Slowly interconverting rotamers are not included in this number. Based on 1H NMR spectroscopic analysis, a qualitative assessment of the spectroscopic features of N1,N7‐bridged species was attempted. By studying the solution behavior of selected isolated and structurally characterized compounds, such as trans‐[PtCl(9MeA‐N7)(NH3)2]ClO4 ? 2H2O or trans,trans‐[{PtCl(NH3)2}2(9MeA‐N1,N7)][ClO4]2 ? H2O, and also by application of a 9MeA complex with an (NH3)3PtII entity at N7, [Pt(9MeA‐N7)(NH3)3][NO3]2, which blocks further cross‐link formation at the N7 site, basic NMR spectroscopic signatures of N1,N7‐bridged PtII complexes were identified. Among others, the trinuclear complex trans‐[Pt(NH3)2{μ‐(N1‐9MeA‐N7)Pt(NH3)3}2][ClO4]6 ? 2H2O was crystallized and its rotational isomerism in aqueous solution was studied by NMR spectroscopy and DFT calculations. Interestingly, simultaneous PtII coordination to N1 and N7 acidifies the exocyclic amino group of the two 9MeA ligands sufficiently to permit replacement of one proton each by a bridging heterometal ion, HgII or CuII, under mild conditions in water.  相似文献   

15.
The size distribution and molecular structure of water clusters play a critical role in the chemical, biological and atmospheric process. The common experimental study of water clusters in aqueous solution is challenged due to the influence of local Hbonding environments on vibration spectroscopies or vacuum requirements for most mass spectrometry technologies. Here, the time-of-flight secondary ion mass spectrometry (ToF-SIMS) combining with a microfluidic chip has been applied to achieve the in-situ discrimination of the size distribution for water clusters in liquid water at room temperature. The results demonstrated that the presented method is highly system stable, reproducible and accurate. The comparison of heavy water with pure water was made to further demonstrate the accuracy of this technique. These results showed that (H2O)3H+ and (D2O)4D+ are the most dominant clusters in pure and heavy water, respectively. This one water molecule difference in the dominant cluster size may due to the nuclear quantum effects on water’s hydrogen bonded network. It is the first time to experimentally show the size distribution of water clusters over a wide range (n=1–30) for pure (H2O) and heavy (D2O) water from molecular level. This technique provides an achievable method for liquid water, which offers a bridge to close the gap between theoretical and experimental study of water cluster in aqueous solution.  相似文献   

16.
Calorimetric measurements have been made of the differential enthalpies of solution as a function of composition of both components in the binary systems tetraethyleneglycol octylether (C8E4)-water and polyethyleneglycol 400 (PEG)-water, as a function of composition, at three different temperatures. Heat capacity changes for dissolution were calculated from the temperature variation of the solution enthalpies. Excess enthalpies and excess heat capacities of mixing were calculated from the differential enthalpies of solution. All measurements on C8E4 were made above the critical micelle concentration (c.m.c.) so the results relate to C8E4 in aggregated form. The thermochemical properties of the C8E4 and PEG systems with water are similar. The differential solution enthalpy of the organic solute in pure water is fairly exothermic and then increases smoothly with increasing solute content. Likewise the solution enthalpy of water in pure C8E4 or PEG is fairly exothermic, but increases steadily to become zero at a water content corresponding to more than five water molecules per ethyleneoxide group. The measurements on the C8E4 system at 40°C were made close to the demixing temperature. The results are compared with previously reported results on the 2-butoxyethanol (BE)-water system.  相似文献   

17.
Absolute bond dissociation energies of water to sodium glycine cations and glycine to hydrated sodium cations are determined experimentally by competitive collision-induced dissociation (CID) of Na+Gly(H2O)x, x = 1–4, with xenon in a guided ion beam tandem mass spectrometer. The cross sections for CID are analyzed to account for unimolecular decay rates, internal energy of reactant ions, multiple ion–molecule collisions, and competition between reaction channels. Experimental results show that the binding energies of water and glycine to the complexes decrease monotonically with increasing number of water molecules. Ab initio calculations at four different levels show good agreement with the experimental bond energies of water to Na+Gly(H2O)x, x = 0–3, and glycine to Na+(H2O), whereas the bond energies of glycine to Na+(H2O)x, x = 2–4, are systematically higher than the experimental values. These discrepancies may provide some evidence that these Na+Gly(H2O)x complexes are trapped in excited state conformers. Both experimental and theoretical results indicate that the sodiated glycine complexes are in their nonzwitterionic forms when solvated by up to four water molecules. The primary binding site for Na+ changes from chelation at the amino nitrogen and carbonyl oxygen of glycine for x = 0 and 1 to binding at the C terminus of glycine for x = 2–4. The present characterization of the structures upon sequential hydration indicates that the stability of the zwitterionic form of amino acids in solution is a consequence of being able to solvate all charge centers.  相似文献   

18.
Photovoltaic conversion has been achieved by use of chloroplasts (photosynthetic organs) from spinach adsorbed on a nanocrystalline TiO2 film on an indium tin oxide (ITO) glass electrode (chloroplast/TiO2 electrode). The shape of the absorption spectrum of the chloroplast/TiO2 electrode is almost the same that of a dispersion of the chloroplasts. Absorption maxima of the chloroplast/TiO2 electrode observed at 430, 475, and 670 nm were attributed to carotenoid and chlorophyll molecules, suggesting that chloroplasts have been adsorbed by the nanocrystalline TiO2 film on the ITO electrode. The photocurrent responses of chloroplast/TiO2 electrodes were measured by using a solution of 0.1 M tetrabutylammonium hexafluorophosphate in acetonitrile as redox electrolyte in the presence or absence of water and 100 mW cm?2 irradiation. The photocurrent of the chloroplast/TiO2 electrode was increased by adding water to the redox electrolyte. The photocurrent responses of chloroplast/TiO2 electrodes irradiated with monochromatic light (680 nm, the absorption band of photosystem II complexed with evolved oxygen) were measured by use of a solution of 0.1 M tetrabutylammonium hexafluorophosphate in acetonitrile as redox electrolyte in the presence or absence of water. A chloroplast/TiO2 electrode photocurrent was observed only when the redox electrolyte containing water was used, indicating that the oxygen evolved from water by photosystem II in chloroplasts adsorbed by a nanocrystalline TiO2 film on an ITO electrode irradiated at 680 nm is reduced to water by the catalytic activity of the platinum electrode. The maximum incident photon-to-current conversion efficiency (IPCE) was 0.8 % on irradiation at 670 nm.  相似文献   

19.
钟克利  邓隆隆  郭佳  张强  侯淑华  边延江  汤立军 《化学通报》2018,81(12):1110-114,1120
本文利用2-氨基吡啶与4-二乙胺基水杨醛反应合成了5-二乙胺基-2-(吡啶-2’-亚氨甲基)苯酚(探针L),对其结构进行了表征。在DMSO/Tris(6:4, v/v, pH =7.4)溶液中,探针L高选择性荧光“关-开”识别Zn2+,在365 nm紫外灯照射下,由无荧光变成蓝色荧光。实验表明,探针L与Zn2+的结合比为1:1,结合常数为2.6×103 L. mol-1,检测限为 9.39×10-7mol/L,pH适用范围为7-11,并可检测水样中的Zn2+。  相似文献   

20.
Calculations are presented for the structure and the isomerization reaction of various conformers of the bare serine, neutral serine–(H2O)n and serine zwitterion–(H2O)n (n = 1, 2) clusters. The effects of binding water molecules on the relative stability and the isomerization processes are examined. Hydrogen bonding between serine and the water molecule(s) may significantly affect the relative stability of conformers of the neutral serine–(H2O)n (n = 1, 2) clusters. The sidechain (OH group) in serine is found to have a profound effect on the structure and isomerization of serine–(H2O)n (n = 1, 2) clusters. Conformers with the hydrogen bonding between water and the hydroxyl group of serine are predicted. A detailed analysis is presented of the isomerization (proton transfer) pathways between the neutral serine–(H2O)2 and serine zwitterion–(H2O)2 clusters by carrying out the intrinsic reaction coordinate analysis. At least two water molecules need to bind to produce the stable serine zwitterion–water cluster in the gas phase. The isomerization for the serine–(H2O)2 cluster proceeds by the concerted double and triple proton transfer mechanism occurring via the binding water molecules, or via the hydroxyl group. © 2004 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  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