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 640 毫秒
1.
Trapp O  Schurig V 《Chirality》2002,14(6):465-470
The axially chiral allenes dimethyl-1,3-allenedicarboxylate 1 and diethyl-1,3-allenedicarboxylate 2 show characteristic plateau formation during enantioselective GC separation on the chiral stationary liquid phase Chirasil-beta-Dex. The elution profiles, obtained from temperature-dependent dynamic GC (DGC) experiments (1: 100-140 degrees C; 2: 110-150 degrees C) were evaluated with the recently derived approximation function (AF) k1(approx) = f(t(R)(A),t(R)(B),w(h)(A),h(plateau), N) to yield the enantiomerization rate constant directly k(1). These values were compared with those obtained by computer-aided simulation with ChromWin. The Eyring activation parameters of the experimental interconversion profiles were determined to be: DeltaG(#)(298.15 K) = 103.6 +/- 0.9 kJ mol(-1), DeltaH(#) = 44.7 +/- 0.4 kJ mol(-1), DeltaS(#) = -198 +/- 7 J K(1) mol(-1) for dimethyl-1,3-allenedicarboxylate 1, and DeltaG(#)(298.15 K) = 103.5 +/- 1.1 kJ mol(-1), DeltaH(#) = 44.7 +/- 0.5 kJ mol(-1), DeltaS(#) = -197 +/- 9 J K(-1) mol(-1) for diethyl-1,3-allenedicarboxylate 2. The approximation function (AF) presented here allows the fast determination of rate constants k(1) and activation barriers of enantiomerization DeltaG(#) from chromatographic parameters without extensive computer simulation.  相似文献   

2.
It has been found that the interconversion of enantiomers on a chromatographic column during the separation process can be studied by the first-order kinetic equations derived both for reversible and irreversible reactions in a stationary system if the extent of interconversion is not too high. The equation derived for irreversible reactions gives, however, results also for higher degrees of enantiomerization while that derived for reversible interconversion failed. The irreversible equation was used to determine the enantiomerization barrier of N-(p-methoxybenzyl)-l,3,2-benzodithiazol-l-oxide enantiomers by supercritical fluid chromatography. The racemate of N-(p-methoxybenzyl)-l,3,2-benzodithiazol-l-oxide was separated by supercritical fluid chromatography on the (R,R)-Whelk-Ol column with supercritical carbon dioxide containing 20% methanol as a mobile phase. Peak areas of enantiomers prior to and after the separation used for the calculation of the enantiomerization barrier were determined by computer-assisted peak deconvolution of peak clusters registered on chromatograms using commercial software.  相似文献   

3.
An on-column stopped-flow multidimensional HPLC (sfMDHPLC) procedure using two chiral stationary phases (CSPs) and one achiral C18 column was developed for the determination of rate constants and free energy barriers of enantiomerization of (+/-)(R,S)-2,3,3a,4-tetrahydro-1H-pyrrolo[2,1-c][1,2,4]benzothiadiazine 5,5-dioxide. Moreover, a stopped-flow HPLC (sfHPLC) method previously developed was applied to the determination of kinetic parameters of enantiomerization of the above compound in the presence of a CSP. The individual enantiomers of the studied compound were isolated in parallel by preparative HPLC and the rate constants and free energy barriers of enantiomerization were determined in different solvents (off-column method). The data obtained by sfMDHPLC, sfHPLC and off-column methods were compared. The (S) enantiomer of the studied compound (S18986) was prepared by asymmetric synthesis and subsequently purified by preparative HPLC, followed by the determination of rate constants and free energy barriers of enantiomerization in different buffer solutions at pH 2-9.3.  相似文献   

4.
The software program DCXplorer is introduced to directly access interconversion rate constants in dynamic chromatography and electrophoresis. The program utilizes the unified equation of chromatography which can evaluate reaction rate constants of all kinds of first order reactions of processes taking place during a separation process. Evaluations with DCXplorer are facilitated by a graphical user interface which allows zooming into the area of interest of an interconversion profile and calculating reaction rate constants without a time consuming simulation process. DCXplorer was applied to determine the enantiomerization barrier of the diuretic drug chlorthalidone by pressure supported dynamic capillary electrokinetic chromatography (DEKC) under acidic conditions at pH 5.00 and pH 3.75. Activation parameters DeltaH(++) and DeltaS(++) were obtained from temperature dependent measurements between 15.0 and 35.0 degrees C in 5 K steps at pH 3.75 and between 30.0 and 50.0 degrees C in 10K steps at pH 5.00.  相似文献   

5.
Trapp O 《Chirality》2006,18(7):489-497
An analytical solution of the unified equation to evaluate elution profiles of interconverting enantiomers in dynamic chromatography is presented. Rate constants k1 and k(-1) and Gibbs activation energies are directly obtained from the chromatographic parameters (retention times tR A and tR A of the interconverting enantiomers, the peak widths at half height wA and wB, and the relative plateau height hp), and the initial amounts A0 and B0 of the enantiomers without any iterative and time consuming computational step. Therefore, this equation is no longer limited to racemic analytes. The analytical solution presented here was validated by comparison with a dataset of 125,000 simulated elution profiles of enantiomerizations. Furthermore, it was found that the recovery rate from a defined dataset is on average 40% higher using the unified equation compared to evaluation methods based on iterative computer simulation. The new equation was applied to determine the enantiomerization rate constant of 1-n-butyl-2-tert-butyldiaziridine by enantioselective gas chromatography. The activation parameters (DeltaH(double dagger) = 112.6 +/- 2.5 kJ/mol and DeltaS(double dagger) = -27 +/- 2 J/(K mol) were obtained from temperature-dependent measurements between 100 degrees C and 140 degrees C in 10K steps.  相似文献   

6.
The effect of association-dissociation on the sulphydryl reactivity of human hemoglobin A is reported. The reactivity of CysF9(93)beta towards the sulphydryl reagent, 5,5'-dithiobis(2-nitrobenzoate), is higher at lower concentrations of hemoglobin at all pH values. This is because hemoglobin dimers have higher sulphydryl reactivity than tetramers and it is known that the proportion of dimers increases as the hemoglobin concentration decreases. This study takes advantage of this observation to determine the tetramer-dimer dissociation constant, K(4,2), of hemoglobin A and subsequently the proton uptake and the proton release during this process. The concentration dependence profiles of the apparent second-order rate constants, k(app), show that (between 2 and 20 microM heme) k(app) decreases with increasing hemoglobin concentration. Above 30 M heme k(app) remains fairly constant for all hemoglobin derivatives (oxy, carbonmonoxy and aquomethemoglobin) used. The pH dependence of the negative logarithm of tetramer-dimer dissociation constant, pK(4,2), for oxy- (and for carbonmonoxy-) hemoglobin exhibits a biphasic character with a maximum near pH 7.4 (and 6.6). For aquomethemoglobin, pK(4,20 decreases with increasing pH. The tetramer-dimer dissociation of human oxyhemoglobin A at an ionic strength of 200 mM uptakes 0.87 +/- 0.09 mole of protons between pH 6.2 to 7.4 phase and releases 0.84 0.09 mole of protons between pH 7.4 and 9.0 phase. Under a similar condition carbonmonoxyhemoglobin uptakes 0.54 +/- 0.05 mole of protons between pH 5.8 and 6.6 phase and releases 0.48 +/- 0.05 mole of protons between pH 6.6 and 9.0 phase. Aquomethemoglobin has only a single phase, it releases 0.39 +/- 0.05 mole of protons during tetramer-dimer dissociation.  相似文献   

7.
A series of chiral tris-(1,10)-phenanthroline iron(II) complexes have been resolved by HPLC on chiral stationary phases based on either cellulose tris-(3,5-dimethylphenylcarbamate) or teicoplanin. At sub ambient temperatures, baseline separation of the enantiomers was observed for five different iron(II) complexes featuring substituted phenanthroline ligands. Dynamic HPLC profiles were observed near or above room temperature, indicating on-column Delta/Lambda enantiomerization. Rate constants for the Delta/Lambda interconversion in free solution and during chromatography were obtained by thermal racemization experiments and by computer simulation of the HPLC dynamic plots, respectively.  相似文献   

8.
An on-column HPLC procedure using a chiral stationary phase (CSP) was developed for the determination of rate constants and free energy barriers of enantiomerization of (+/-)IDRA21. Subsequently, the HPLC method was applied for investigation of two structurally related chiral compounds. The individual enantiomers of the studied compounds were isolated in parallel by preparative HPLC and rate constants and free energy barriers of enantiomerization were determined in different solvents. The on-column enantiomerization data revealed that CSP induces different rate constants for the two enantiomers. The results generated off-line were used to determine the influence of solvents on the racemization of (+) and (-) IDRA21 and to gain further insight into the enantiomerization mechanism of chiral 3,4-dihydro-1,2,4-benzothiadiazine 1,1-dioxide type compounds.  相似文献   

9.
Cyclometalated ruthenium(II) complexes, [Ru(II)(C~N)(N~N)(2)]PF(6) [HC~N=2-phenylpyridine (Hphpy) or 2-(4'-tolyl)pyridine; N~N=2,2'-bipyridine, 1,10-phenanthroline, or 4,4'-dimethyl-2,2'-bipyridine], are rapidly oxidized by H(2)O(2) catalyzed by plant peroxidases to the corresponding Ru(III) species. The commercial isoenzyme C of horseradish peroxidase (HRP-C) and two recently purified peroxidases from sweet potato (SPP) and royal palm tree (RPTP) have been used. The most favorable conditions for the oxidation have been evaluated by varying the pH, buffer, and H(2)O(2) concentrations and the apparent second-order rate constants ( k(app)) have been measured. All the complexes studied are oxidized by HRP-C at similar rates and the rate constants k(app) are identical to those known for the best substrates of HRP-C (10(6)-10(7) M(-1) s(-1)). Both cationic (HRP-C) and anionic (SPP and RPTP) peroxidases show similar catalytic efficiency in the oxidation of the Ru(II) complexes. The mediating capacity of the complexes has been evaluated using the SPP-catalyzed co-oxidation of [Ru(II)(phpy)(bpy)(2)]PF(6) and catechol as a poor peroxidase substrate as an example. The rate of enzyme-catalyzed oxidation of catechol increases more than 10000-fold in the presence of the ruthenium complex. A simple routine for calculating the rate constant k(c) for the oxidation of catechol by the Ru(III) complex generated enzymatically from [Ru(II)(phpy)(bpy)(2)](+) is proposed. It is based on the accepted mechanism of peroxidase catalysis and involves spectrophotometric measurements of the limiting Ru(II) concentration at different concentrations of catechol. The calculated k(c) value of 0.75 M(-1) s(-1) shows that the cyclometalated Ru(II) complexes are efficient mediators in peroxidase catalysis.  相似文献   

10.
报导了一种新型单氮杂 18—冠— 6键合硅胶固定相 ( BCN 18— C— 6)在小分子肽分析中的应用。研究了流动相 p H值、Cu2 +浓度、缓冲溶液浓度、甲醇体积比、氯化钠浓度等因素对容量因子 ( k)的影响 ,并分析了保留机理。在优化条件下 ,成功地分离了 4种小分子肽。肽的色谱峰面积与进样量间有良好的线性关系 ( r均大于 0 .99) ,各肽的最小检出量 ( mmin)均可达到 10 - 10 ~ 10 - 11m ol。 BCN 18— C— 6柱对小分子肽的分离选择性优于 ODS柱和硅胶柱  相似文献   

11.
Muradov KG  Artemyev NO 《Biochemistry》2000,39(14):3937-3942
The N-terminal regions of the heterotrimeric G-protein alpha-subunits represent one of the major Gbetagamma contact sites and have been implicated in an interaction with G-protein-coupled receptors. To probe the role of the N-terminal domain of transducin-alpha in G-protein function, a chimeric Gtialpha subunit with the 31 N-terminal Gtalpha residues replaced by the corresponding 42 residues of Gsalpha (Ns-Gtialpha) has been examined for the interaction with light-activated rhodopsin (R). Gtialpha displayed a somewhat higher R-stimulated rate of GTPgammaS binding relative to Ns-Gtialpha, suggesting modest involvement of the Gtalpha N-terminal sequence in recognition of the receptor. However, the intrinsic rate of nucleotide exchange in Ns-Gtialpha was significantly faster (k(app) = 0.014 min(-)(1)) than that in Gtialpha (k(app) = 0.0013 min(-1)) as judged by the GTPgammaS binding rates. Substitution of 42 N-terminal residues of Gsalpha by the Gtalpha residues in a reciprocal chimera, Nt-Gsalpha, had an opposite effect-notable reduction in the intrinsic GTPgammaS-binding rate (k(app) = 0.0075 min(-)(1)) in comparison with Gsalpha (k(app) = 0.028 min(-)(1)). Residue Val30 (His41 in Gsalpha) within the N-terminal region of Gtalpha interacts with the C-terminal residue, Ile339. To test the hypothesis that observed changes in the intrinsic nucleotide exchange rate in chimeric Galpha subunits might be attributed to this interaction, GtialphaVal30His, GtialphaIle339Ala, and Ns-GtialphaHis41Val mutants have been made and analyzed for basal GTPgammaS binding. GtialphaVal30His and GtialphaIle339Ala had increased GTPgammaS binding rates (k(app) = 0. 010 and 0.009 min(-)(1), respectively), whereas Ns-GtialphaHis41Val had a decreased GTPgammaS binding rate (k(app) = 0.0011 min(-)(1)) relative to their parent proteins. These results suggest that the coupling between the N-terminal and C-terminal domains of Gtalpha is important for maintaining a low nucleotide exchange rate in unstimulated transducin.  相似文献   

12.
Hardy LW  Nishida CH  Kirsch JF 《Biochemistry》1984,23(6):1288-1294
The pH dependence of k(cat) for the Bacillus cereus beta-lactamase I catalyzed hydrolysis of carbenicillin(VI), which differs from benzylpenicillin (I) in having a carboxylic moiety alpha to the phenyl ring, exhibits a profile consistent with a model in which the alpha-COOH and alpha-COO forms of the ES complex turn over with respective rate constants of 2152 s(-1) and 384 s(-1). The pK(a)(app) for the alpha-COOH is shifted from 3.2 in solution to 6.1 in the ES complex. The normalized k(cat)/K(m) vs. pH profile for VI is not superimposable on that of I, indicating that both the neutral and anionic forms of the carboxyl moiety of VI combine with the enzyme to give the first irreversibly formed complex, presumably the acyl-enzyme. Quantitative accord with the kinetic data is achieved only through fitting to a model where kinetically significant proton transfer in the ES complex is permitted. The second-order rate constants for the reaction of the enzyme with the alpha-COOH and alpha-COO forms of VI are 2.2 x 10(8) M(-1) s(-1) and 3.8 x 10(6) M(-1) s(-1), respectively. The high value for the alpha-COOH form suggests that this reaction may be in part diffusion controlled. This conjecture is borne out by the observation that the sensitivity of k(cat)/K(m) to eta(rel) decreases with increasing pH for VI, whereas this sensitivity is pH independent for I. These conclusions are further supported by the results of a kinetic investigation of the pH dependence of sulbenicillin (VII) where an alpha-SO3H replaces the alpha-COOH of VI. The strongly acidic sulfonic acid moiety of VII is fully ionized throughout nearly the entire pH range of interest, and its kinetics, as a function of pH, are very similar to those observed and calculated for the alpha-COO form of VI. Solvent deuterium kinetic isotope effects are reported for k(cat) and k(cat)/K(m) for both VI and VII.  相似文献   

13.
Thrombin (T) inactivation by the serpin, heparin cofactor II (HCII), is accelerated by the glycosaminoglycans (GAGs) dermatan sulfate (DS) and heparin (H). Equilibrium binding and thrombin inactivation kinetics at pH 7.8 and ionic strength (I) 0.125 m demonstrated that DS and heparin bound much tighter to thrombin (K(T(DS)) 1-5.8 microm; K(T(H)) 0.02-0.2 microm) than to HCII (K(HCII(DS)) 236-291 microm; K(HCII(H)) 25-35 microm), favoring formation of T.GAG over HCII.GAG complexes as intermediates for T.GAG.HCII complex assembly. At [GAG] < K(HCII(GAG)) the GAG and HCII concentration dependences of the first-order inactivation rate constants (k(app)) were hyperbolic, reflecting saturation of T.GAG complex and formation of the T.GAG.HCII complex from T.GAG and free HCII, respectively. At [GAG] > K(HCII(GAG)), HCII.GAG complex formation caused a decrease in k(app). The bell-shaped logarithmic GAG dependences fit an obligatory template mechanism in which free HCII binds GAG in the T.GAG complex. DS and heparin bound fluorescently labeled meizothrombin(des-fragment 1) (MzT(-F1)) with K(MzT(-F1)(GAG)) 10 and 20 microm, respectively, demonstrating a binding site outside of exosite II. Exosite II ligands did not attenuate the DS-accelerated thrombin inactivation markedly, but DS displaced thrombin from heparin-Sepharose, suggesting that DS and heparin share a restricted binding site in or nearby exosite II, in addition to binding outside exosite II. Both T.DS and MzT(-F1).DS interactions were saturable at DS concentrations substantially below K(HCII(DS)), consistent with DS bridging T.DS and free HCII. The results suggest that GAG template action facilitates ternary complex formation and accommodates HCII binding to GAG and thrombin exosite I in the ternary complex.  相似文献   

14.
The kinetics of the decomposition of H(2)O(2) catalyzed by Cu(II) has been studied by the initial-rate method in aqueous phosphate media at near physiological pH. The activity of the catalyst is increased by [Fe(CN)(6)](3-) and decreased by VO(3)(-), CrO(4)(2-) and Zn(II). Three reaction pathways are involved in the Cu(II)-H(2)O(2) reaction, the kinetic orders of the catalyst being 1 (rate constant k1), 2 (rate constant k2) and 3 (rate constant k3). The three pathways present fractional apparent orders (>1) in H(2)O(2) and base catalysis. The apparent activation energies associated to rate constants k1, k2 and k3 are 102+/-4, 65+/-8 and 61+/-5 kJ mol(-1). Free-radical chain mechanisms are proposed for the three pathways.  相似文献   

15.
The reversible enantiomerization of axially chiral 2'-dodecyloxy-6-nitrobiphenyl-2-carboxylic acid was studied in the presence of a brush type chiral stationary phase based on O-(tert-butylcarbamoyl) quinine as chiral selector unit by stopped-flow high-performance liquid chromatography (sfHPLC) and capillary electrochromatography (sfCEC). After initial separation of the enantiomers in the first section of the column, the flow was stopped and the resolved species allowed to enantiomerize on-column. From this conversion, which could be determined from the enantiomeric ratios at different enantiomerization times, kinetic rate constants were calculated. By sfHPLC at a constant temperature of 15 degrees C, kinetic rate constants in the presence of the CSP were found to be 4.1 x 10(-5) s(-1) and 2.2 x 10(-5) s(-1) for the (-) and (+)-enantiomers, respectively, corresponding to half-lives of 279 and 530 min. Thus, apparent activation energies of enantiomerization were calculated to be 93.0 and 94.6 kJ mol(-1) for the (-) and (+)-enantiomers. On the macroscopic level, the apparent difference of rotational energy barriers and kinetic rate constants for both enantiomers is reflected as deracemization. For example, starting from a racemic mixture, an enantiomeric excess (ee) of 14% was seen in the stopped-flow HPLC experiment described after an enantiomerization time of 220 min at 15 degrees C, and a maximal ee of 17% can be approximated after infinite enantiomerization time. There is good agreement between HPLC and CEC results as well as their experimental errors, confirming that the new sfCEC technique may be a valuable and convenient tool to study interconversion processes.  相似文献   

16.
Human vascular adhesion protein-1 (VAP-1) is an endothelial copper-dependent amine oxidase involved in the recruitment and extravasation of leukocytes at sites of inflammation. VAP-1 is an important therapeutic target for several pathological conditions. We expressed soluble VAP-1 in HEK293 EBNA1 cells at levels suitable for detailed mechanistic studies with model substrates. Using the model substrate benzylamine, we analyzed the steady-state kinetic parameters of VAP-1 as a function of solution pH. We found two macroscopic pK(a) values that defined a bell-shaped plot of turnover number k(cat,app) as a function of pH, representing ionizable groups in the enzyme-substrate complex. The dependence of (k(cat)/K(m))(app) on pH revealed a single pK(a) value (~9) that we assigned to ionization of the amine group in free benzylamine substrate. A kinetic isotope effect (KIE) of 6 to 7.6 on (k(cat)/K(m))(app) over the pH range of 6 to 10 was observed with d(2)-benzylamine. Over the same pH range, the KIE on k(cat) was found to be close to unity. The unusual KIE values on (k(cat)/K(m))(app) were rationalized using a mechanistic scheme that includes the possibility of multiple isotopically sensitive steps. We also report the analysis of quantitative structure-activity relationships (QSAR) using para-substituted protiated and deuterated phenylethylamines. With phenylethylamines we observed a large KIE on k(cat,app) (8.01 ± 0.28 with phenylethylamine), indicating that C-H bond breakage is limiting for 2,4,5-trihydroxyphenylalanine quinone reduction. Poor correlations were observed between steady-state rate constants and QSAR parameters. We show the importance of combining KIE, QSAR, and structural studies to gain insight into the complexity of the VAP-1 steady-state mechanism.  相似文献   

17.
Kozlov AG  Lohman TM 《Biochemistry》2002,41(39):11611-11627
The kinetic mechanism of transfer of the homotetrameric Escherichia coli SSB protein between ssDNA molecules was studied using stopped-flow experiments. Dissociation of SSB from the donor ssDNA was monitored after addition of a large excess of unlabeled acceptor ssDNA by using either SSB tryptophan fluorescence or the fluorescence of a ssDNA labeled with an extrinsic fluorophore [fluorescein (F) or Cy3]. The dominant pathway for SSB dissociation occurs by a "direct transfer" mechanism in which an intermediate composed of two DNA molecules bound to one SSB tetramer forms transiently prior to the release of the acceptor DNA. When an initial 1:1 SSB-ssDNA complex is formed with (dT)(70) in the fully wrapped (SSB)(65) mode so that all four SSB subunits are bound to (dT)(70), the formation of the ternary intermediate complex occurs slowly with an apparent bimolecular rate constant, k(2,app), ranging from 1.2 x 10(3) M(-1) s(-1) (0.2 M NaCl) to approximately 5.1 x 10(3) M(-1) s(-1) (0.4 M NaBr), and this rate limits the overall rate of the transfer reaction (pH 8.1, 25 degrees C). These rate constants are approximately 7 x 10(5)- and approximately 7 x 10(4)-fold lower, respectively, than those measured for binding of the same ssDNA to an unligated SSB tetramer to form a singly ligated complex. However, when an initial SSB-ssDNA complex is formed with (dT)(35) so that only two SSB subunits interact with the DNA in an (SSB)(35) complex, the formation of the ternary intermediate occurs much faster with a k(2,app) ranging from >6.3 x 10(7) M(-1) s(-1) (0.2 M NaCl) to 2.6 x 10(7) M(-1) s(-1) (0.4 M NaBr). For these experiments, the rate of dissociation of the donor ssDNA determines the overall rate of the transfer reaction. Hence, an SSB tetramer can be transferred from one ssDNA molecule to another without proceeding through a free protein intermediate, and the rate of transfer is determined by the availability of free DNA binding sites within the initial SSB-ssDNA donor complex. Such a mechanism may be used to recycle SSB tetramers between old and newly formed ssDNA regions during lagging strand DNA replication.  相似文献   

18.
Tripeptidyl-peptidase II (TPP II) is a subtilisin-like serine protease which forms a large enzyme complex (>4MDa). It is considered a potential drug target due to its involvement in specific physiological processes. However, information is scarce concerning the kinetic characteristics of TPP II and its active site features, which are important for design of efficient inhibitors. To amend this, we probed the active site by determining the pH dependence of TPP II catalysis. Access to pure enzyme is a prerequisite for kinetic investigations and herein we introduce the first efficient purification system for heterologously expressed mammalian TPP II. The pH dependence of kinetic parameters for hydrolysis of two different chromogenic substrates, Ala-Ala-Phe-pNA and Ala-Ala-Ala-pNA, was determined for murine, human and Drosophila melanogaster TPP II as well as mutant variants thereof. The investigation demonstrated that TPP II, in contrast to subtilisin, has a bell-shaped pH dependence of k(cat)(app)/K(M) probably due to deprotonation of the N-terminal amino group of the substrate at higher pH. Since both the K(M) and k(cat)(app) are lower for cleavage of AAA-pNA than for AAF-pNA we propose that the former can bind non-productively to the active site of the enzyme, a phenomenon previously observed with some substrates for subtilisin. Two mutant variants, H267A and D387G, showed bell-shaped pH-dependence of k(cat)(app), possibly due to an impaired protonation of the leaving group. This work reveals previously unknown differences between TPP II orthologues and subtilisin as well as features that might be conserved within the entire family of subtilisin-like serine peptidases.  相似文献   

19.
Trapp O  Schoetz G  Schurig V 《Chirality》2001,13(8):403-414
In recent years, dynamic chromatography and stopped-flow chromatographic techniques have become versatile tools for the determination of enantiomerization and isomerization barriers. Increasing demands for the stereochemical safety of chiral drugs contributed to the rapid development of new techniques. New computer-aided evaluation systems allow the on-line determination of interconversion barriers from the experimental chromatograms. Both dynamic chromatography and stopped-flow chromatography have been applied to the entire range of chromatographic methods (GC, SFC, HPLC, CE).  相似文献   

20.
The diacylglycerol (DG)/phorbol ester-dependent translocation of conventional protein kinase C (PKC) isozymes is mediated by the C1 domain, a membrane-targeting module that also selectively binds phosphatidylserine (PS). Using stopped-flow spectroscopy, we dissect the contribution of DG/phorbol esters (C1 ligand) and PS in driving the association and dissociation of the C1 domain from membranes. Specifically, we examine the binding to membranes of the C1B domain of PKCbeta with a substituted Trp (Y123W) whose fluorescence is quenched upon binding to membranes. Binding of this construct (C1Bbeta-Y123W) to phospholipid vesicles is cooperative with respect to PS content and dependent on C1 ligand, as previously characterized. Stopped-flow analysis reveals that the apparent association rate (k(on)(app)), but not the apparent dissociation rate (k(off)(app)), is highly sensitive to PS content: the 60-fold increase in membrane affinity for vesicles containing no PS compared with 40 mol % PS results primarily from a robust (30-fold) increase in k(on)(app) with little effect (2-fold) on k(off)(app). Membrane affinity is also controlled by the content and structure of the C1 ligand. In contrast to PS, these ligands markedly alter k(off)(app) with smaller effects on k(on)(app). We also show that the affinity for phorbol ester-containing membranes is 2 orders of magnitude higher than that for DG-containing membranes primarily resulting from differences in k(off)(app). Our data are consistent with a model in which the C1 domain is recruited to the membrane via an initial weak electrostatic interaction with PS, followed by a rapid two-dimensional search for ligand, the binding of which retains the domain at the membrane. Thus, PS drives the initial encounter, and DG/phorbol esters retain the domain on membranes. The decreased effectiveness of DG compared with phorbol esters in retaining the C1 domain on membranes contributes to the molecular dichotomy of the rapid, transient nature of DG-dependent PKC signaling versus the chronic hyperactivity of phorbol ester-activated PKC.  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