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 281 毫秒
1.
Development of a feasible method for studying the competitive interaction between a pair of antagonists is essential for understanding the antagonism of trace metals in biological systems. Herein, we report the application of CE on‐line coupled with ICP mass spectroscopy (CE‐ICP‐MS) to investigate the competitive binding of Zn2+ against Cd2+ for glutathione (GSH), which is related to the detoxification of Cd2+ in biological system, and introduce a method to evaluate the kinetics and thermodynamics for the competitive binding of Zn2+ against Cd2+ for GSH. The CE‐ICP‐MS hybrid technique allows easy and sensitive probing of the competitive binding of Zn2+ against Cd2+ for GSH and quantitative determination of the important thermodynamic and kinetic parameters of the competitive binding of Zn2+ against Cd2+ for GSH. Owing to the high sensitivity and element selectivity with multi‐elements detection capacity of ICP‐MS, we detailed the evaluation of the kinetics and thermodynamics describing the competition of Zn2+ against Cd2+ for GSH from the systematic data obtained by CE‐ICP‐MS. The competitive binding of Zn2+ against Cd2+ for GSH was demonstrated exothermic and thermodynamically favorable (ΔG=?7.2 kJ/mol) and driven entirely by a large favorable enthalpy decrease (ΔH=?15.1 kJ/mol) but with an unfavorable entropy decrease (ΔS=?25.6 J/mol/K). The kinetic data were fit to a second‐order equation with the reaction rate constant (k) of (2.18±0.10)×102 L/(mol·s) under the simulated physiological condition.  相似文献   

2.
A kinetic study is reported for SNAr reactions of 2,4,6‐tris(trifluoromethanesulfonyl) anisole 1a with a series of para‐X‐substituted anilines 2a–e in a methanol solution at various temperatures. The substituent effects on free energy (ΔG), enthalpy (ΔH), and entropy (ΔS) of activation are examined. Aniline addition to triflone 1a is characterized by a βX=0.57, αZ=0.31, and an imbalance of I = αZ–βX=?0.26. The imbalance shows that resonance development lags behind C? N bond formation at the transition state. Interestingly, analysis of the results in terms of Marcus theory reveals that these SNAr are associated with some extremely low intrinsic reactivity (log ko=?1.25& © 2011 Wiley Peiodicals, Inc. Int J Chem Kinet 43: 255–262, 2011  相似文献   

3.
郭清莲a 李冉b  c 周新a  c  刘义b  c 《中国化学》2008,26(12):2207-2215
用荧光光谱和紫外吸收光谱法研究了酮康唑与牛血清白蛋白和人血清白蛋白的相互作用。实验进行于pH = 7.40±0.1的0.1 mol∙L-1PBS磷酸缓冲溶液。实验结果表明,酮康唑与牛血清白蛋白和人血清白蛋白的结合常数均会随着温度的升高而降低,酮康唑可以有规律地使血清白蛋白内源荧光猝灭,其猝灭机理可认为是酮康唑与白蛋白形成复合物的静态猝灭。并且获得了不同温度下,酮康唑与白蛋白作用的结合常数以及∆G、∆H和∆S等热力学参数。根据所得结果可推断酮康唑与白蛋白的作用力主要为静电作用力和疏水作用力,同时由FRET能量转移理论计算得出了酮康唑与白蛋白结合位置的距离r。  相似文献   

4.
Cadmium(II) imidazole (IMI) azide [Cd(IMI)2(N3)2]n (1) was synthesized using imidazole and azide, and was characterized by the elemental analysis and FTIR spectrum. The crystal structure was determined by X-ray single crystal diffraction, and the crystallographic data show that the crystal belongs to orthorhombic, Pba2 space group, α?=?10.780(4) Å, b?=?13.529(5) Å, and c?=?3.6415(12) Å. Its crystal density is 2.080?g·cm–3. Cd(II) is a six-coordinate with six nitrogens from four imidazoles and two azides with μ–1,1 coordination. The thermal decomposition mechanism was determined based on differential scanning calorimetry (DSC) and thermogravimetry-derivative thermogravimetry (TG-DTG) analysis, and the kinetic parameters of the first exothermic process were studied using Kissinger’s method and Ozawa’s method, respectively. The energy of combustion, enthalpy of formation, critical temperature of thermal explosion, entropy of activation (ΔS ), enthalpy of activation (ΔH ), and free energy of activation (ΔG ) were measured and calculated. In the end, impact sensitivity was also determined by standard method.  相似文献   

5.
In the experimental part of this study, thermal properties of the Al and NiO composites in micro‐ and nano‐sized Al are investigated. Differential scanning calorimetry (DSC) analysis of the onset temperatures of ignition, activation energy (Ea), frequency factor (A), rate constant (k), critical ignition temperature of thermal explosion (Tb), and self‐accelerating decomposition temperature (TSADT), as well as the thermodynamic parameters (ΔS , ΔH , and ΔG ) are used to explore the thermal behavior and analyze the kinetics. Thermal analysis suggests that the mechanism is based on solid–solid diffusion and liquid–gas for the nano‐ and micro‐Al/NiO composite, respectively. Our results indicate that the incorporation of nano‐Al particles can significantly reduce the ignition temperature, Ea, A, k, Tb, and TSADT. In the second part of this work, molecular dynamics (MD) simulation is used to investigate the behavior of Al/NiO thermite reaction using the Reaxff force field to evaluate the experimental results. Theoretically, MD results show 1,154 K as the reaction ignition temperature, which is in reasonably good agreement with experimental temperature of 893°C (1,166 K). The radial distribution function (RDF) shows that no reaction occurs at 500 K but it is complete at 1,200 K.  相似文献   

6.
《Analytical letters》2012,45(16):2693-2710
Abstract

The binding interaction of the terazosin hydrochloride and prazosin hydrochloride with bovine serum albumin (BSA) was studied by spectrofluorimetry. Both of these two compounds quenched the fluorescence of BSA. The thermodynamic parameters (ΔH 0, ΔS 0 and ΔG 0) obtained from the fluorescence data measured at two different temperatures showed that the binding of terazosin hydrochloride to BSA involved hydrogen bonds and that of prazosin hydrochloride to BSA involved hydrophobic and electrostatic interactions. In this work, the competitive interaction of the terazosin hydrochloride and prazosin hydrochloride with BSA was studied by three-way excitation-emission fluorescence with the aid of parallel factor analysis (PARAFAC).  相似文献   

7.
γ-tubulin plays crucial role in the nucleation and organization of microtubules during cell division. Recent studies have also indicated its role in the regulation of microtubule dynamics at the plus end of the microtubules. Moreover, γ-tubulin has been found to be over-expressed in many cancer types, such as carcinomas of the breast and glioblastoma multiforme. These studies have led to immense interest in the identification of chemical leads that might interact with γ-tubulin and disrupt its function in order to explore γ-tubulin as potential chemotherapeutic target. Recently a colchicine-interacting cavity was identified at the interface of γ-tubulin dimer that might also interact with other similar compounds. In the same direction we theoretically investigated binding of a class of compounds, noscapinoids (noscapine and its derivatives) at the interface of the γ-tubulin dimer. Molecular interaction of noscapine and two of its derivatives, amino-noscapine and bromo-noscapine, was investigated by molecular docking, molecular dynamics simulation and binding free energy calculation. All noscapinoids displayed stable interaction throughout simulation of 25 ns. The predictive binding free energy (ΔGbind) indicates that noscapinoids bind strongly with the γ-tubulin dimer. However, bromo-noscapine showed the best binding affinity (ΔGbind = –37.6 kcal/mol) followed by noscapine (ΔGbind = –29.85 kcal/mol) and amino-noscapine (ΔGbind = –23.99 kcal/mol) using the MM-PBSA method. Similarly using the MM-GBSA method, bromo-noscapine showed highest binding affinity (ΔGbind = –43.64 kcal/mol) followed by amino-noscapine (ΔGbind = –37.56 kcal/mol) and noscapine (ΔGbind = –34.57 kcal/mol). The results thus generate compelling evidence that these noscapinoids may hold great potential for preclinical and clinical evaluation.  相似文献   

8.
张志凌  左超  庞代文 《化学学报》2005,63(22):2069-2076
采用自己建立的DNA表面电化学研究微量方法, 研究了单双链DNA与两种锇配合物(联吡啶锇和二氯菲咯啉锇)的相互作用. 研究发现, 两种锇配合物都是通过静电作用与DNA结合, 其作用方式不受溶液离子强度的影响. 并计算得到了联吡啶锇和二氯菲咯啉锇与dsDNA和ssDNA相互作用的多个热力学和动力学参数, 如结合常数K3+K2+, 结合常数比K3+/K2+, 离子强度为零时的极限比 , 结合自由能ΔGb, 解离速度常数k, 结合位点数s.  相似文献   

9.
Modelling of proton and metal exchange in the alginate biopolymer   总被引:1,自引:0,他引:1  
Acid–base behaviour of a commercial sodium alginate extracted from brown seaweed (Macrocystis pyrifera) has been investigated at different ionic strengths (0.1≤I/mol l?1≤1.0) and in different supporting electrolytes (Et4NI, NaCl, KCl, LiCl, NaCl+MgCl2), with the aim of examining the influence of ionic medium on the proton-binding capacity and of quantifying the strength of interaction with light metal ions in the perspective of speciation studies in natural aqueous systems. Potentiometric ([H+]-glass electrode) and titration calorimetric data were expressed as a function of the dissociation degree (α) using different models (Henderson–Hasselbalch modified, Högfeldt three parameters and linear equations). The dependence on ionic strength of the protonation constants was taken into account by a modified specific interaction theory model. Differences among different media were explained in terms of the interaction between polyanion and metal cations of the supporting electrolytes. Quantitative information on the proton-binding capacity, together with the stabilities of different species formed, is reported. Protonation thermodynamic parameters, at α=0.5, are log K H=3.686±0.005, ΔG 0=?21.04±0.03 kJ mol?1, ΔH 0=4.8±0.2 kJ mol?1 and TΔS 0=35.7±0.3 kJ mol?1, at infinite dilution. Protonation enthalpies indicate that the main contribution to proton binding arises from the entropy term. A strict correlation between ΔG and TΔS was found, TΔS=?9.5–1.73 ΔG. Results are reported in light of building up a chemical complexation model of general validity to explain the binding ability of naturally occurring polycarboxylate polymers and biopolymers. Speciation profiles of alginate in the presence of sodium and magnesium ions, naturally occurring cations in natural waters, are also reported.  相似文献   

10.
Triazolophanes are used as the venue to compete an aliphatic propylene CH hydrogen‐bond donor against an aromatic phenylene one. Longer aliphatic C? H ??? Cl? hydrogen bonds were calculated from the location of the chloride within the propylene‐based triazolophane. The gas‐phase energetics of chloride binding (ΔGbind, ΔHbind, ΔSbind) and the configurational entropy (ΔSconfig) were computed by taking all low‐energy conformations into account. Comparison between the phenylene‐ and propylene‐based triazolophanes shows the computed gas‐phase free energy of binding decreased from ΔGbind=?194 to ?182 kJ mol?1, respectively, with a modest enthalpy–entropy compensation. These differences were investigated experimentally. An 1H NMR spectroscopy study on the structure of the propylene triazolophane’s 1:1 chloride complex is consistent with a weaker propylene CH hydrogen bond. To quantify the affinity differences between the two triazolophanes in dichloromethane, it was critical to obtain an accurate binding model. Four equilibria were identified. In addition to 1:1 complexation and 2:1 sandwich formation, ion pairing of the tetrabutylammonium chloride salt (TBA+ ? Cl?) and cation pairing of TBA+ with the 1:1 triazolophane–chloride complex were observed and quantified. Each complex was independently verified by ESI‐MS or diffusion NMR spectroscopy. With ion pairing deconvoluted from the chloride–receptor binding, equilibrium constants were determined by using 1H NMR (500 μM ) and UV/Vis (50 μM ) spectroscopy titrations. The stabilities of the 1:1 complexes for the phenylene and propylene triazolophanes did not differ within experimental error, ΔG=(?38±2) and (?39±1) kJ mol?1, respectively, as verified by an NMR spectroscopy competition experiment. Thus, the aliphatic CH donor only revealed its weaker character when competing with aromatic CH donors within the propylene‐based triazolophane.  相似文献   

11.
Acetyl-coenzyme A carboxylases (ACCs) is the first committed enzyme of fatty acid synthesis pathway. The inhibition of ACC is thought to be beneficial not only for diseases related to metabolism, such as type-2 diabetes, but also for infectious disease like bacterial infection disease. Soraphen A, a potent allosteric inhibitor of BC domain of yeast ACC, exhibit lower binding affinities to several yeast ACC mutants and the corresponding drug resistance mechanisms are still unknown. We report here a theoretical study of binding of soraphen A to wild type and yeast ACC mutants (including F510I, N485G, I69E, E477R, and K73R) via molecular dynamic simulation and molecular mechanics/generalized Born surface area free energy calculations methods. The calculated binding free energies of soraphen A to yeast ACC mutants are weaker than to wild type, which is highly consistent with the experimental results. The mutant F510I weakens the binding affinity of soraphen A to yeast ACC mainly by decreasing the van der Waals contributions, while the weaker binding affinities of Soraphen A to other yeast ACC mutants including N485G, I69E, E477R, and K73R are largely attributed to the decreased net electrostatic (ΔEele?+?ΔGGB) interactions. Our simulation results could provide important insights for the development of more potent ACC inhibitors.  相似文献   

12.
As determined by both 1H NMR and UV/Vis spectroscopic titration, ESI‐MS, isothermal titration calorimetry, and DFT molecular modeling, advanced glycation end products (AGE) breaker alagebrium (ALA) formed 1:1 guest–host inclusion complexes with cucurbit[7]uril (CB[7]), with a binding affinity, Ka, in the order of magnitude of 105 m ?1, thermodynamically driven by both enthalpy (ΔH=?6.79 kcal mol?1) and entropy (TΔS=1.21 kcal mol?1). For the first time, a dramatic inhibition of keto–enol tautomerism of the carbonyl α‐hydrogen of ALA has been observed, as evidenced by over an order of magnitude decrease of both the first step rate constant, k1, and the second step rate constant, k2, during hydrogen/deuterium exchange in D2O. Meanwhile, as expected, the reactivity of C2‐hydrogen was also inhibited significantly, with an upshift of 2.09 pKa units. This discovery will not only provide an emerging host molecule to modulate keto–enol tautomerism, but also potentially lead to a novel supramolecular formulation of AGE‐breaker ALA for improved stability and therapeutic efficacy.  相似文献   

13.
The binding of kaempferol‐3,7‐αL‐rhamnopyranoside (KRR) with bovine serum albumin (BSA) was investigated by different spectroscopic methods under simulative physiological conditions. Analysis of ?uorescence quenching data of BSA by KRR at different temperatures using Stern‐Volmer methods revealed the formation of a ground state KRR‐BSA complex with moderate binding constant of the order 104 L·mol?1. The existence of some metal ions could weaken the binding of KRR on BSA. The changes in the van't Hoff enthalpy (ΔH0) and entropy (ΔS0) of the interaction were estimated to be ?26.53 kJ·mol?1 and 3.33 J·mol?1·K?1 and both hydrophobic and electrostatic forces contributed to stabilizing the BSA‐KRR complex. According to the F?ster theory of non‐radiation energy transfer, the distance r between the donor (BSA) and the acceptor (KRR) was obtained (r=2.83 nm). Site marker competitive experiments showed that KRR could bind to Site I of BSA. In addition, synchronous fluorescence, UV‐Vis absorption and circular dichroism (CD) results indicated that the KRR binding could cause conformational changes of BSA.  相似文献   

14.
The interaction between mercuric ion and apoCopC in the absence or presence of cupric ion was investigated through difference UV spectra in Hepes buffer (10 mmol·L^-1) at pH 7.4. The results suggest that mercuric ion can bind to C- and N-terminal binding sites of apoCopC, and the conditional binding constants were calculated to be kN=(6.79± 1.12)× 10^6 mol^-1·L and kc=(3.06±0.05)× 10^5 mol^-1·L. Using urea as a chemical agent, the conformational stabilities of apoCopC and HgN^2+ -CopC-Hgc^2+ were monitored by fluorescence spectrum in Hepes buffer (50 mmol·L^-1) at pH 7.4. The free energy of stabilization is (14.69±0.85) and (16.66±0.55) kJ.mol^-1, respectively. HgN^2+ -CopC-Hgc^2+ is more stable than apoCopC.  相似文献   

15.
The benzoylformyl Pd(II) complex, Pd(PPh3)2(Cl)(COCOPh), thermally decomposes to the corresponding benzoyl complex by the loss of CO. The predominant route of decarbonylation is led by a reversible dissociation of a phosphine ligand. The disappearance of the benzoylformyl complex in solutions follows first order kinetics not only in the existence of excess PPh3 but also in the absence of added PPh3. Through the treatments of both preequilibrium and the steady state approximations to the kinetic data, the rate constants of the intramolecular acyl migration, k1 and k2; as well as the equilibrium constant and the individual rate constants of the reversible phosphine dissociation step, K, kd and kd, were determined in CHCl3. The activation parameters for kd, being ΔH? = 25.4 Kcal Mol?1, ΔS? = 15.9 eu, ΔG? = 20.7 Kcal Mol?1;and for k?d, being ΔH? = 13.0 Kcal Mol?1, ΔS? = ?7.9 eu, ΔG? = 15.4 Kcal Mol?1, were evaluated.  相似文献   

16.
An optical biosensor with a stirred cuvette has been used to monitor the interaction between immobilized human serum albumin (HSA) and three water-soluble cationic porphyrins. The binding constants at 25℃ obtained from biosensor analysis were compared with those from fluorescence spectroscopy. The interactions were further investigated at temperatures from 15℃ to 30℃. The thermodynamics parameters, changes of free energy (△G), enthalpy (△H) and entropy (△S), were evaluated from equilibrium data. It appeared that the binding process was governed primarily by electrostatic forces.  相似文献   

17.
The conductances of cetylpyridinium chloride (CPyCl) + tetradecyltrimethylammonium bromide (TTAB) mixtures over the entire mole fraction range of CPyCl (xCPyCl) were measured in aqueous binary mixtures of polyethylene glycol 1000 (PEG 1000) containing 0.5, 1, 2, 5 and 10 wt% of PEG 1000 at 30°C. From the conductivity data, the critical micellar concentration (cmc), degree of counter ion association (χ) and the standard free energy of micelle formation (ΔGo M ) of CPyCl and TTAB were computed. The DGo M value is further divided into the hydrophobic contribution of free energy of transfer of the surfactant hydrocarbon chain from the medium to the micelle (ΔGo HP ), and the energy associated with the surface contributions (ΔGo s ) consisting of electrostatic interactions between the head groups and counter ions. Both contributions show a linear dependence on the amount of PEG additive. These results have been explained on the basis of the medium effects of aqueous PEG.

The mixed micelle formation by CPyCl+TTAB show non-ideal behavior which is quite similar in the absence, as well as in the presence, of PEG additive as evaluated by using the regular solution theory. The interaction parameter, b, is always negative and remains almost constant with respect to the amount of PEG additive. This indicates that the mixed micelle formation occurs mainly due to the synergistic interactions between the unlike surfactant monomers only.  相似文献   

18.
The synthesis and the solid state magnetic properties of (nitronyl nitroxide)‐substituted trioxytriphenylamine radical cation tetrachlorogallate, NNTOT+·GaCl4? , are reported. In the temperature region between 300 and 3 K, the magnetic behavior is characterized by the strong intramolecular ferromagnetic interaction (J/kB=+400 K) between the radical ( NN ) and the radical cation ( TOT +) and the weak intermolecular antiferromagnetic interaction (J/kB=?1.9 K) between NNTOT+ ions. Below 3 K, a 3D‐type long‐range magnetic ordering into a weak ferromagnet was observed (TN=2.65 K). The magnetic entropy (Smag=8.97 J K?1 mol?1) obtained by the heat capacity measurement is in good agreement with the theoretical value of R ln3=9.13 J K?1 mol?1 based on the S=1 state.  相似文献   

19.
A family of seven cationic gold complexes that contain both an alkyl substituted π‐allene ligand and an electron‐rich, sterically hindered supporting ligand was isolated in >90 % yield and characterized by spectroscopy and, in three cases, by X‐ray crystallography. Solution‐phase and solid‐state analysis of these complexes established preferential binding of gold to the less substituted C?C bond of the allene and to the allene π face trans to the substituent on the uncomplexed allenyl C?C bond. Kinetic analysis of intermolecular allene exchange established two‐term rate laws of the form rate=k1[complex]+k2[complex][allene] consistent with allene‐independent and allene‐dependent exchange pathways with energy barriers of ΔG1=17.4–18.8 and ΔG2=15.2–17.6 kcal mol?1, respectively. Variable temperature (VT) NMR analysis revealed fluxional behavior consistent with facile (ΔG=8.9–11.4 kcal mol?1) intramolecular exchange of the allene π faces through η1‐allene transition states and/or intermediates that retain a staggered arrangement of the allene substituents. VT NMR/spin saturation transfer analysis of [{P(tBu)2o‐binaphthyl}Au(η2‐4,5‐nonadiene) ]+SbF6? ( 5 ), which contains elements of chirality in both the phosphine and allene ligands, revealed no epimerization of the allene ligand below the threshold for intermolecular allene exchange (ΔG298K=17.4 kcal mol?1), which ruled out the participation of a η1‐allylic cation species in the low‐energy π‐face exchange process for this complex.  相似文献   

20.
We report that 2,6‐lutidine?trichloroborane (Lut?BCl3) reacts with H2 in toluene, bromobenzene, dichloromethane, and Lut solvents producing the neutral hydride, Lut?BHCl2. The mechanism was modeled with density functional theory, and energies of stationary states were calculated at the G3(MP2)B3 level of theory. Lut?BCl3 was calculated to react with H2 and form the ion pair, [LutH+][HBCl3?], with a barrier of ΔH=24.7 kcal mol?1G=29.8 kcal mol?1). Metathesis with a second molecule of Lut?BCl3 produced Lut?BHCl2 and [LutH+][BCl4?]. The overall reaction is exothermic by 6.0 kcal mol?1rG°=?1.1). Alternate pathways were explored involving the borenium cation (LutBCl2+) and the four‐membered boracycle [(CH2{NC5H3Me})BCl2]. Barriers for addition of H2 across the Lut/LutBCl2+ pair and the boracycle B?C bond are substantially higher (ΔG=42.1 and 49.4 kcal mol?1, respectively), such that these pathways are excluded. The barrier for addition of H2 to the boracycle B?N bond is comparable (ΔH=28.5 and ΔG=32 kcal mol?1). Conversion of the intermediate 2‐(BHCl2CH2)‐6‐Me(C5H3NH) to Lut?BHCl2 may occur by intermolecular steps involving proton/hydride transfers to Lut/BCl3. Intramolecular protodeboronation, which could form Lut?BHCl2 directly, is prohibited by a high barrier (ΔH=52, ΔG=51 kcal mol?1).  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