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 78 毫秒
1.
The mechanism of complexation of alkali metal cations with macrocyclic ligands such as the simple crown ethers and the role of desolvation vs. ligand rearrangement are discussed. The unique role of water solvent in the rate-determining step of complexations in aqueous solutions is brought into focus. The competitive role of the anion, which becomes of paramount importance in solvents of low permittivity, is reiterated. Monoazo crown ethers are shown to possess isomeric equilibria in methanol solvent. The rate-determining process for the first step of complexation of these macrocycles with Na+ in methanol appears to be the rearrangement of the ligand through inversion to an exo position of the nitrogen lone electron pair. The rate-determining step of the overall complexation is the entrance of the Na+ into the ring with (possibly) concomitant rotation of the lone electron of the nitrogen to an endo configuration.  相似文献   

2.
The detailed mechanisms of the hydrolysis of carbonyl sulfide (OCS) by nucleophilic water and hydroxide ion in both the gas phase and bulk water solvent have been investigated using density functional theory. Various reaction channels on the potential surface have been identified. The thermodynamic results demonstrate that the hydrolysis of OCS by nucleophilic water and hydroxide ion should proceed more favorably at low temperature. The hydrolysis of OCS by the hydroxide ion is the main reaction channel from thermodynamic and kinetic perspectives, and the bulk solvent can influence the rate-determining step in this channel. However, the solvent barely modifies the activation energy of the rate-determining step. For the hydrolysis of OCS by nucleophilic water, the solvent does not modify the rate-determining step, and the corresponding activation energy of the rate-determining step barely changes. This bulk solvent effect suggests that most of the contribution of the solvent is accounted for by considering one water molecule and a hydroxide ion.  相似文献   

3.
H Kyushiki  A Ikai 《Proteins》1990,8(3):287-293
The overall activity of an animal fatty acid synthetase at the saturation level of substrate concentration decreased when the solvent viscosity, eta, of the reaction mixture was increased with viscogens such as glycerol, sucrose, and polyethylene glycol. The activity of the enzyme changed roughly proportional to eta-P, where p = 1.0 for glycerol, p = 0.66 for sucrose, and p less than 0.6 for polyethylene glycol with different molecular sizes. The thioesterase activity, which catalyzes the final partial reaction in the multifunctional enzyme, was not affected by 5-fold increase of solvent viscosity with sucrose. These results suggested that the rate-determining step of the enzyme other than the thioesterase reaction involves a microscopic transport step, the rate of which is influenced by the solvent viscosity. The microscopic transport step may be related to the transfer of the reaction intermediate from one active site to another or to the motion of a larger part of the enzyme requisite for the catalytic reaction. In the solution containing glycerol, the rate-determining motion was primarily diffusion limited since the inverse of the initial rate was proportional to eta, i.e., p = 1. Since the substrate concentration was at a saturation level in this experiment, the viscosity-dependent step cannot be the encounter between the enzyme and substrates, but must be intramolecular in origin, most probably the reaction catalyzed by beta-ketoacyl synthetase. In solutions containing other viscogens, however, p was less than 1.0, indicating a significant involvement of chemical steps in the rate-determining step as well. Bovine serum albumin, when used as a proteinic viscogen, also decreased the initial rate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
A representative of a new class of dyes with dual fluorescence due to an excited state intramolecular proton transfer (ESIPT) reaction, namely 1-methyl-2-(4-methoxy)phenyl-3-hydroxy-4(1H)-quinolone (QMOM), has been studied in a series of solvents covering a large range of polarity and basicity. A linear dependence of the logarithm of its two bands intensity ratio, log(I(N*)/I(T*)), upon the solvent polarity expressed as a function of the dielectric constant, (epsilon- 1)/(2epsilon + 1), is observed for a series of protic solvents. A linear dependence for log(I(N*)/I(T*)) is also found in aprotic solvents after taking into account the solvent basicity. In contrast, the positions of the absorption and the two emission bands of QMOM do not noticeably depend on the solvent polarity and basicity, indicating relatively small changes in the transition moment of QMOM upon excitation and emission. Time-resolved experiments in acetonitrile, ethyl acetate and dimethylformamide suggest an irreversible ESIPT reaction for this dye. According to the time-resolved data, an increase of solvent basicity results in a dramatic decrease of the ESIPT rate constant, probably due to the disruption of the intramolecular H-bond of the dye by the basic solvent. Due to this new sensor property, 3-hydroxyquinolones are promising candidates for the development of a new generation of environment-sensitive fluorescence dyes for probing interactions of biomolecules.  相似文献   

5.
The fluorescence properties of 1-alkyl(and aryl)thio-2-alkylisoindoles, formed by the reaction of o-phthalaldehyde (OPTA) and thiols with primary amines, are reported. Variations in thiol and amine substituents and solvent polarity have large effects on the isoindole fluorescence spectra. These parameters, in addition to 3-thiol substitution of the isoindoles, pH, and the use of phosphate vs borate aqueous buffers, were found to have dramatic effects on the corrected relative fluorescence intensity. Low concentrations and nonaqueous solvents apparently stabilized most adducts while aqueous solutions, especially at low pH, caused pseudo-first-order decomposition, probably via hydrolysis to the corresponding 2,3-dihydro-1H-isoindole-1-one. However, 3.3 × 10−8 solutions of the more intensely fluorescent adducts (total adduct 5 pmol) were readily detected if the fluorescence was determined shortly after adding the isoindole to pH 9.2 borate buffer. The adduct formed using ethanethiol and n-propylamine possessed spectral properties which were the most responsive to changes in solvent polarity and was the most stable under the various conditions employed. Finally, arguments are presented that these isoindoles are the products in several other fluorogenic assays using OPTA.  相似文献   

6.
Thermophilic enzymes tend to be less catalytically-active at lower temperatures relative to their mesophilic counterparts, despite having very similar crystal structures. An often cited hypothesis for this general observation is that thermostable enzymes have evolved a more rigid tertiary structure in order to cope with their more extreme, natural environment, but they are also less flexible at lower temperatures, leading to their lower catalytic activity under mesophilic conditions. An alternative hypothesis, however, is that complementary thermophilic-mesophilic enzyme pairs simply operate through different evolutionary-optimized catalytic mechanisms. In this communication, we present evidence that while the steps of the catalytic mechanisms for mesophilic and thermophilic indole-3-glycerol phosphate synthase (IGPS) enzymes are fundamentally similar, the identity of the rate-determining step changes as a function of temperature. Our findings indicate that while product release is rate-determining at 25°C for thermophilic IGPS, near its adaptive temperature (75°C), a proton transfer event, involving a general acid, becomes rate-determining. The rate-determining steps for thermophilic and mesophilic IGPS enzymes are also different at their respective, adaptive temperatures with the mesophilic IGPS-catalyzed reaction being rate-limited before irreversible CO2 release, and the thermophilic IGPS-catalyzed reaction being rate limited afterwards.  相似文献   

7.
A kinetically homogeneous anti-phosphate catalytic antibody preparation was shown to catalyse the hydrolysis of a series of O-aryl N-methyl carbamates containing various substituents in the 4-position of the O-phenyl group. The specific nature of the antibody catalysis was demonstrated by the adherence of these reactions to the Michaelis-Menten equation, the complete inhibition by a hapten analogue, and the failure of the antibody to catalyse the hydrolysis of the 2-nitrophenyl analogue of the 4-nitrophenylcarbamate substrate. Hammett sigma-rho analysis suggests that both the non-catalysed and antibody-catalysed reactions proceed by mechanisms in which development of the aryloxyanion of the leaving group is well advanced in the transition state of the rate-determining step. This is probably the ElcB (elimination-addition) mechanism for the non-catalysed reaction, but for the antibody-catalysed reaction might be either ElcB or B(Ac)2 (addition-elimination), in which the elimination of the aryloxy group from the tetrahedral intermediate has become rate-determining. This result provides evidence of the dominance of recognition of phenolate ion character in the phosphate hapten in the elicitation process, and is discussed in connection with data from the literature that suggest a B(Ac)2 mechanism, with rate-determining formation of the tetrahedral intermediate for the hydrolysis of carbamate substrates catalysed by an antibody elicited by a phosphonamidate hapten in which phenolate anion character is minimized. The present paper contributes to the growing awareness that small differences in the structure of haptens can produce large differences in catalytic characteristics.  相似文献   

8.
The photophysical properties of 2-amino-9,10-anthraquinone (2AAQ) have been investigated in different solvents and solvent mixtures and correlated with the Lippert-Mataga solvent polarity parameter, Deltaf. In the low solvent polarity region with Deltaf < ca. 0.1, the dye shows unusually high fluorescence quantum yields (Phif) and lifetimes (tauf) in comparison to those in other solvents of medium to high polarities. Similarly, the radiative rate constants (kf) are relatively lower and the non-radiative rate constants (knr) are relatively higher in the low polarity solvents in comparison to those in the medium to high polarity solvents. The current results have been rationalized assuming that the dye adopts different structural forms below and above the Deltaf value of approximately 0.1. It is inferred that in the low solvent polarity region the dye exists in a non-planar structure, with its 2-NH2 plane away from that of the 9,10-anthraquinone moiety in the ground state. In solvents of medium to high polarities, the dye exists in a polar intramolecular charge transfer (ICT) structure, where the amino lone pair of the 2-NH2 group is in strong resonance with the anthraquinone pi-cloud in the ground state. In all the solvents, however the dye is inferred to exist in the ICT structure in its excited (S1) state. Supportive evidence for the above hypothesis has been obtained from the solvent polarity effect on the Stokes' shifts for the dye. Quantum chemical studies on the structures of 2AAQ dye in the gas phase also give qualitative support for the inferences drawn from the photophysical properties of the dye in different solvents.  相似文献   

9.
Biocatalytic transesterification of methylmethacrylate is possible in many different solvents. The reaction rate is readily controlled by variation in solvent physical properties. The reaction proceeds better in hydrophobic solvents, and activity can be restored in hydrophilic solvents by the addition of water. We have now demonstrated that supercritical carbon dioxide is not a good solvent for the reaction between 2-ethlhexanol and methylmethacrylate. It apperars that the supercritical carbon dioxide may either alter the pH of the microaqueous environment associated with the protein or reversibly form covalent complexes with free amine groups on the surface of the enzyme. Although supercritical carbon dioxide is a poor solvent for acrylate transesterification, many other supercritical fluids (ethane, ethylene, sulfur hexafluoride, and fluoroform) are better than most conventional solvents. In supercritical ethane it is possible to control the activity of the enzyme by changing pressure, and the enzyme appears to follow Michaelis-Menten Kinetics. We find that sulfur hexafluoride, the first anhydrous inorganic solvent in which biocatalytic activity has been reported, is a better solvent than any conventional or supercritical organic fluid tested.  相似文献   

10.
A comparative study was made of Novozym 435-catalyzed regioselective acylation of 1-beta-D-arabinofuranosylcytosine with vinyl propionate for the preparation of the 5'-O-monoester in eleven co-solvent mixtures and three pure polar solvents. Novozym 435 displayed low or no acylation activity toward 1-beta-D-arabinofuranosylcytosine in pure polar solvents, although those solvents can dissolve the nucleosides well. When a hexane-pyridine co-solvent system was adopted, both the initial rate and the substrate conversion were enhanced markedly. The polarity of co-solvent mixtures had significant effect on the reaction. Among the solvent mixtures investigated, the higher the polarity of the solvent mixture, the lower the initial reaction rate and the substrate conversion. It was also found that the acylation was dependent on the hydrophobic solvent content, the water activity and the reaction temperature. The most suitable co-solvent, initial water activity, and reaction temperature were hexane-pyridine (28:72, v/v), 0.07, and 50 degrees C, respectively. Under these conditions, the initial rate, the substrate conversion and the regioselectivity were as high as 91.1 mM h(-1), >97% and >98%, respectively, after a reaction time of 6 h. Among the reaction mediums examined, the lowest apparent activation energy was achieved with hexane-pyridine (28:72, v/v), in which Novozym 435 also exhibited good thermal stability.  相似文献   

11.
Kinetic and spectroscopic studies were carried out to study the role of hydrophobic effect on the activity of bovine serum amine oxidase (BSAO). Increasing the chain length of the substrates (linear aliphatic primary monoamines), the affinity for the active site increases while the catalytic constant decreases in accordance with a relative low value of dielectric constant (about 10) estimated for the microenvironment of BSAO active site using a fluorescent probe sensitive to solvent polarity. The aliphatic chain of 1-aminononane induces a shift in the pK(a) of the product Schiff base, the hydrolysis of which appears to be a rate-determining step of the reaction. Furthermore, circular dichroism studies highlighted the "flexibility" of BSAO secondary structure that can explain the wide substrate specificity of this enzyme. These results should be useful to elucidate the substrate/inhibitor preferences of CuAOs, in particular of the human enzyme.  相似文献   

12.
The reaction of soybean lipoxygenase-1 with linoleic acid has been extensively studied and displays very large kinetic isotope effects. In this work, substrate and solvent kinetic isotope effects as well as the viscosity dependence of the oxidation of arachidonic acid were investigated. The hydrogen atom abstraction step was rate-determining at all temperatures, but was partially masked by a viscosity-dependent step at ambient temperatures. The observed KIEs on k(cat) were large ( approximately 100 at 25 degrees C).  相似文献   

13.
Effects of intra- and intermolecular hydrogen-bonds on the photophysical properties of 2'-aminoacetophenone derivatives (X-C6H4-COCH3) having a substituted amino group (X) with different hydrogen-bonding ability to the carbonyl oxygen (X: NH2(AAP), NHCH3(MAAP), N(CH3)2(DMAAP), NHCOCH3(AAAP), NHCOCF3(TFAAP)) are investigated by means of steady-state and time-resolved fluorescence spectroscopy and time-resolved thermal lensing. Based on the photophysical parameters obtained in aprotic solvents with different polarity and protic solvents with different hydrogen-bonding ability, the characteristic photophysical behavior of the 2'-aminoacetophenone derivatives is discussed in terms of hydrogen-bonding and n,pi*-pi,pi* vibronic coupling. The dominant deactivation process of AAP and MAAP in nonpolar aprotic solvents is the extremely fast internal conversion (k(ic)= 1.0 x 10(11) s(-1) for AAP and 3.9 x 10(10) s(-1) for MAAP in n-hexane). The internal conversion rates of both compounds decrease markedly with increasing solvent polarity, suggesting that vibronic interactions between close-lying S1(pi,pi*) and S2(n,pi*) states lead to the large increase in the non-radiative decay rate of the lowest excited singlet state. It is also suggested that for MAAP, which has a stronger hydrogen-bond as compared to AAP, an intramolecular hydrogen-bonding induced deactivation is involved in the dissipation of the S1 state. For DMAAP, which cannot possess an intramolecular hydrogen-bond, the primary relaxation mechanism of the S1 state in nonpolar aprotic solvents is the intersystem crossing to the triplet state, whereas in protic solvents very efficient internal conversion due to intermolecular hydrogen-bonding is induced. In contrast, the fluorescence spectra of AAAP and TFAAP, which have an amino group with a much stronger hydrogen-bonding ability, give strongly Stokes-shifted fluorescence, indicating that these compounds undergo excited-state intramolecular proton transfer reaction upon electronic excitation.  相似文献   

14.
The reaction between methionine and enneamolybdomanganate(IV) in perchloric acid was carried out under pseudo-first-order conditions keeping large excess of methionine. The orders in oxidant and substrate were found to be unity and 0.91, respectively. The reaction proceeds with rapid formation of complex between the reactants followed by its decomposition in a rate determining step. The accelerating effect of hydrogen ions on the reaction is due to the formation of active hexaprotonated oxidant species. The product of the reaction was found to be methionine sulfoxide. The reaction involves direct two-electron transfer step without any free radical intervention. The effect of ionic strength, solvent polarity and the activation parameters were also in support of the mechanism proposed.  相似文献   

15.
Thermo-kinetics of lipase-catalyzed synthesis of 6-O-glucosyldecanoate   总被引:1,自引:0,他引:1  
Lipase-catalyzed synthesis of 6-O-glucosyldecanoate from d-glucose and decanoic acid was performed in dimethyl sulfoxide (DMSO), a mixture of DMSO and tert-butanol and tert-butanol alone with a decreasing order of polarity. The highest conversion yield (> 65%) of decanoic acid was obtained in the blended solvent of intermediate polarity mainly because it could dissolve relatively large amounts of both the reactants. The reaction obeyed Michaelis-Menten type of kinetics. The affinity of the enzyme towards the limiting substrate (decanoic acid) was not affected by the polarity of the solvent, but increased significantly with temperature. The esterification reaction was endothermic with activation energy in the range of 60-67 kJ mol?1. Based on the Gibbs energy values, in the solvent blend of DMSO and tert-butanol the position of the equilibrium was shifted more towards the products compared to the position in pure solvents. Monoester of glucose was the main product of the reaction.  相似文献   

16.
Shapes of curves of pH-dependence of reactions   总被引:14,自引:14,他引:0  
A simple case is considered in which the rate of a two-step reaction depends on pH because the intermediate formed in the first step has to gain (or lose) a proton before it can react in the second step, and in which the rate-determining step therefore changes with pH. The curves of reaction rate against pH are shown to be symmetrical, and the sharpest peak possible has a width at half its height of 1.53pH units, i.e. of 2log(3+2 radical2). Any particular curve for this situation proves to be identical with a curve that could be generated for the pH-dependence of a single-step reaction in which the rate is proportional to the concentration of a particular ionic form of a reactant. Curves for the latter situation, however, can have forms impossible for the former case in which the rate-determining step changes, but only if the protonations that activate and deactivate the reactant are co-operative. The peak can then become even sharper, and its width at half its height can fall to 1.14pH units, i.e. to 2log(2+ radical3).  相似文献   

17.
Kinch LN  Phillips MA 《Biochemistry》2000,39(12):3336-3343
Trypanosoma cruzi S-adenosylmethionine decarboxylase (AdoMetDC) catalyzes the pyruvoyl-dependent decarboxylation of S-adenosylmethionine (AdoMet), which is an important step in the biosynthesis of polyamines. The time course of the AdoMetDC reaction under single-turnover conditions was measured to determine the rate of the slowest catalytic step up to and including decarboxylation. Analysis of this single-turnover data yields an apparent second-order rate constant for this reaction of 3300 M(-1) s(-1) in the presence of putrescine, which corresponds to a catalytic rate of >6 s(-1). This rate is minimally 100-fold faster than the steady-state rate suggesting that product release, which includes Schiff base hydrolysis, limits the overall reaction. AdoMetDC exhibits an inverse solvent isotope effect on the single-turnover kinetics, and the pH profile predicts a pK(a) of 8.9 for the basic limb. These results are consistent with a Cys residue functioning as a general acid in the rate-determining step of the single-turnover reaction. Mutation of Cys-82 to Ala reduces the rate of the single turnover reaction to 11 M(-1) s(-1) in the presence of putrescine. Further, a solvent isotope effect is not observed for the mutant enzyme. Reduction of the wild-type enzyme with cyanoborohydride traps the Schiff base between the enzyme and decarboxylated substrate, while little Schiff base species of either substrate or product was trapped with the C82A mutant. These data suggest that Cys-82 functions as a general acid/base to catalyze Schiff base formation and hydrolysis. The solvent isotope and pH effects are mirrored in single-turnover analysis of reactions without the putrescine activator, yielding an apparent second-order rate constant of 150 M(-1) s(-1). The presence of putrescine increases the single-turnover rate by 20-fold, while it has relatively little effect on the affinity of the enzyme for product. Therefore, putrescine likely activates the T. cruzi AdoMetDC enzyme by accelerating the rate of Schiff base exchange.  相似文献   

18.
Laabe于1987年提出了生物催化剂工程(Biocatalyst engineering)和介质工程(Medium engineering)的概念[1]。有机相生物催化中溶剂的选择也是介质工程的内容之一。纯酶在有机相中的催化作用已有大量报道[2],但对完整细胞研究甚步。本文以甲基单胞菌(Methylomonas Z201)完整细胞为生物催化剂.丙烯环氧化为指标反应.研究有机溶剂对活性的影响并对催化活性——溶剂疏水性进行了相关性分析。研究了水一十六烷两相体系中十六烷含量和搅拌速度对丙烯环氧化速度的影响和细胞的操作稳定性。  相似文献   

19.
Forces between type I collagen triple helices are studied in solvents of varying hydrogen-bonding ability. The swelling of collagen fibers in reconstituted films is controlled by the concentration of soluble polymers that are excluded from the fibers and that compete osmotically with collagen for available solvent. The interaxial spacing between the triple helices as a function of the polymer concentration is measured by x-ray diffraction. Exponential-like changes in the spacing with increasing osmotic stress, qualitatively similar to the forces previously found in aqueous solution, are also seen in formamide and ethylene glycol. These are solvents that, like water, are capable of forming three-dimensional hydrogen-bond networks. In solvents that either cannot form a network or have a greatly impaired ability to form a hydrogen-bonded network, strikingly different behavior is observed. A hard-wall repulsion is seen with collagen solvated by ethanol, 2-propanol, and N,N-dimethylformamide. The spacing between helices hardly changes with increasing polymer concentration until the stress exceeds some threshold where removal of the solvent becomes energetically favorable. No solvation of collagen is observed in dimethoxyethane. In solvents with an intermediate ability to form hydrogen-bonded networks, methanol, 2-methoxyethanol, or N-methylformamide, the change in spacing with polymer concentration is intermediate between exponential-like and hard-wall. These results provide direct evidence that the exponential repulsion observed between collagen helices at 0-8-A surface separations in water is due to the energetic cost associated with perturbing the hydrogen-bonded network of solvent molecules between the collagen surfaces.  相似文献   

20.
Designing enzymes for use in organic solvents.   总被引:1,自引:0,他引:1  
Enzymes are routinely used in organic solvents where numerous reactions of interest to synthetic and polymer chemists can be performed with high selectivity. Recently, it has become apparent that the catalytic properties of an enzyme can be tailored to a specific catalytic requirement by the use of solvent and protein engineering. The former involves altering the polarity, hydrophobicity, water content, etc., of the organic milieu, while the later applies site-directed mutagenesis to alter the physicochemical properties of the biocatalyst. The dominant effects of organic solvents on enzyme structure and function, and the potential of solvent and protein engineering to design enzymes to function optimally in organic media, are the major foci of this review.  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