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 406 毫秒
1.
Four enol lactones, bearing phenyl or 1-naphthyl substituents on the alpha or beta positions [3-phenyl-6-methylenetetrahydro-2-pyranone (alpha Ph6H, IIc), 3-(1-naphthyl)-6-methylenetetrahydro-2-pyranone (alpha Np6H, IId), 4-phenyl-6-methylenetetrahydro-2-pyranone (beta Ph6H, IIIc), and 4-(1-naphthyl)-6-methylenetetrahydro-2-pyranone (beta Np6H, IIId)], available as pure R and S enantiomers, have been studied as alternate substrate inhibitors of chymotrypsin. Kinetic constants for substrate binding (Ks) and acylation (ka) were determined by a competitive substrate assay, using succinyl-L-Ala-L-Ala-L-Pro-L-Phe p-nitroanilide; the deacylation rate constant (kd) was determined by the proflavin displacement assay. All lactones undergo rapid acylation (ka varies from 17 to 170 min-1) that shows little enantioselectivity; there is, however, pronounced enantioselectivity in substrate binding for three of the lactones [Ks(R/S) = 40-110]. In each case it is the enantiomer with the S configuration that has the higher affinity. In all cases, deacylation rates are slow, and in two cases, acyl enzymes with half-lives of 4.0 and 12.5 h at pH 7.2, 25 degrees C, are obtained (for beta Ph6H and alpha Np6H, respectively). In these cases, high deacylation enantioselectivity is observed [kd(S/R) = 60-70], and the lactone more weakly bound as a substrate (R enantiomer) gives the more stable acyl enzyme. Two hypotheses, involving hindrance of the attack of water or an exchange of the ester and ketone carbonyl groups in the acyl enzyme, are advanced as possible explanations for the high stability of these acyl enzymes.  相似文献   

2.
Solution NMR studies of a (15)N-labeled G-protein alpha-subunit (G(alpha)) chimera ((15)N-ChiT)-reconstituted heterotrimer have shown previously that G-protein betagamma-subunit (G(betagamma)) association induces a "pre-activated" conformation that likely facilitates interaction with the agonist-activated form of a G-protein-coupled receptor (R*) and guanine nucleotide exchange (Abdulaev, N. G., Ngo, T., Zhang, C., Dinh, A., Brabazon, D. M., Ridge, K. D., and Marino, J. P. (2005) J. Biol. Chem. 280, 38071-38080). Here we demonstrated that the (15)N-ChiT-reconstituted heterotrimer can form functional complexes under NMR experimental conditions with light-activated, detergent-solubilized rhodopsin (R*), as well as a soluble mimic of R*. NMR methods were used to track R*-triggered guanine nucleotide exchange and release of guanosine 5'-O-3-thiotriphosphate (GTPgammaS)/Mg(2+)-bound ChiT. A heteronuclear single quantum correlation (HSQC) spectrum of R*-generated GTPgammaS/Mg(2+)-bound ChiT revealed (1)HN, (15)N chemical shift changes relative to GDP/Mg(2+)-bound ChiT that were similar, but not identical, to those observed for the GDP.AlF(4)(-)/Mg(2+)-bound state. Line widths observed for R*-generated GTPgammaS/Mg(2+)-bound (15)N-ChiT, however, indicated that it is more conformationally dynamic relative to the GDP/Mg(2+)- and GDP.AlF(4)(-)/Mg(2+)-bound states. The increased dynamics appeared to be correlated with G(betagamma) and R* interactions because they are not observed for GTPgammaS/Mg(2+)-bound ChiT generated independently of R*. In contrast to R*, a soluble mimic that does not catalytically interact with G-protein (Abdulaev, N. G., Ngo, T., Chen, R., Lu, Z., and Ridge, K. D. (2000) J. Biol. Chem. 275, 39354-39363) is found to form a stable complex with the GTPgammaS/Mg(2+)-exchanged heterotrimer. The HSQC spectrum of (15)N-ChiT in this complex displays a unique chemical shift pattern that nonetheless shares similarities with the heterotrimer and GTPgammaS/Mg(2+)-bound ChiT. Overall, these results demonstrated that R*-induced changes in G(alpha) can be followed by NMR and that guanine nucleotide exchange can be uncoupled from heterotrimer dissociation.  相似文献   

3.
The biological activities of transforming growth factor-beta isoforms (TGF-beta(1,2)) are known to be modulated by alpha(2)-macroglobulin (alpha(2)M). alpha(2)M forms complexes with numerous growth factors, cytokines, and hormones, including TGF-beta. Identification of the binding sites in TGF-beta isoforms responsible for high affinity interaction with alpha(2)M many unravel the molecular basis of the complex formation. Here we demonstrate that among nine synthetic pentacosapeptides with overlapping amino acid sequences spanning the entire TGF-beta(1) molecule, the peptide (residues 41-65) containing Trp-52 exhibited the most potent activity in inhibiting the formation of complexes between (125)I-TGF-beta(1) and activated alpha(2)M (alpha(2)M*) as determined by nondenaturing polyacrylamide gel electrophoresis and by plasma clearance in mice. TGF-beta(2) peptide containing the homologous sequence and Trp-52 was as active as the TGF-beta(1) peptide, whereas the corresponding TGF-beta(3) peptide lacking Trp-52, was inactive. The replacement of the Trp-52 with alanine abolished the inhibitory activities of these peptides. (125)I-TGF-beta(3), which lacks Trp-52, bound to alpha(2)M* with an affinity lower than that of (125)I-TGF-beta(1). Furthermore, unlabeled TGF-beta(3) and the mutant TGF-beta(1)W52A, in which Trp-52 was replaced with alanine, were less potent than unlabeled TGF-beta(1) in blocking I(125)-TGF-beta(1) binding to alpha(2)M*. TGF-beta(1) and TGF-beta(2) peptides containing Trp-52 were also effective in inhibiting I(125)-nerve growth factor binding to alpha(2)M*. Tauhese results suggest that Trp-52 is involved in high affinity binding of TGF-beta to alpha(2)M*. They also imply that TGF-beta and other growth factors/cytokines/hormones may form complexes with alpha(2)M* via a common mechanism involving the interactions between topologically exposed Trp and/or other hydrophobic residues and a hydrophobic region in alpha(2)M*.  相似文献   

4.
A group of unnatural 1-(2-deoxy-beta-D-ribofuranosyl)-2,4-difluorobenzenes possessing a 5-I or 5-CF3 substituent, that were originally designed as thymidine mimics, were coupled via their 5'-OH group to a cyclosaligenyl (cycloSal) ring system having a variety of C-3 substituents (Me, OMe, H). The 5'-O-cycloSal-pronucleotide concept was designed to effect a thymidine kinase-bypass, thereby providing a method for the intracellular delivery and generation of the 5'-O-monophosphate for nucleosides that are poorly phosphorylated. The 5'-O-cycloSal pronucleotide phosphotriesters synthesized in this study were obtained as a 1:1 mixture of two diastereomers that differ in configuration (S(P) or R(P)) at the asymmetric phosphorous center. The (S(P))- and (R(P))-diastereomers for the 5'-O-3-methylcycloSal- and 5'-O-3-methoxycycloSal derivatives of 1-(2-deoxy-beta-D-ribofuranosyl)-2,4-difluoro-5-iodobenzene were separated by silica gel flash column chromatography. This class of cycloSal pronucleotide compounds generally exhibited weak cytotoxic activities in a MTT assay (CC50 values in the 10(-3) to 10(-4) M range), against a number of cancer cell lines (143B, 143B-LTK, EMT-6, Hela, 293), except for cyclosaligenyl-5'-O-[1'-(2,4-difluoro-5-iodophenyl)-2'-deoxy-beta-D-ribofuranosyl]phosphate that was more potent (CC50 values in the 10(-5) to 10(-6) M range), than the reference drug 5-iodo-2'-deoxyuridine (IUDR) which showed CC50 values in the 10(-3) to 10(-5) M range.  相似文献   

5.
One- and two-color, mass selected R2PI spectra of the S(1)<--S(0) transitions in the bare (+)-(R)-1-phenyl-1-ethanol (E(R)) and its complexes with different solvent molecules (solv) (-)-(R)-2-butanol (B(R)) or (+)-(S)-2-butanol (B(S)) and (-)-(R)-2-butylamine (A(R)) or (+)-(S)-2-butylamine (A(S)), have been recorded after a supersonic molecular beam expansion. The one-color R2PI excitation spectra of the diastereomeric complexes are characterized by significant shifts of their band origin relative to that of bare E(R). The extent and the direction of these spectral shifts are found to depend on the structure and the configuration of solv and are attributed to different short-range interactions in the ground and excited states of the complexes. In analogy with other diastereomeric complexes, the phenomenological binding energy of the homochiral cluster is found to be greater than that of the heterochiral one.  相似文献   

6.
A series of carbazole-functionalized carboxylate ligands (N-carbazolylacetic acid (L1), 4-carbazol-9-yl-benzoic acid (L2), and 3-(4-carbazol-9-yl-phenyl)-acrylic acid (L3)) and their corresponding silver complexes were designed and synthesized and the structures were determined by single crystal X-ray diffractions. The silver atoms in the complexes are in tetrahedral geometry coordinated with two oxygen from carboxylic and two phosphorous atoms from triphenylphosphine. The complexes exhibit excellent electrochemical characters in solution and strong photoluminescence in the solid state. The emission wavelengths of the compounds can be tuned (from ultraviolet to visible region) by introducing of the second coordinating ligand triphenylphosphine and by elongation of the conjugation moieties.  相似文献   

7.
Diorganotin(IV) complexes R(2)Sn(cap) (capH(2)=N-[(S)-3-mercapto-2-methylpropionyl]-L-proline; R=Me, Et, n-Bu and t-Bu) were prepared and characterised. The FTIR and Raman spectra demonstrated that the organotin(IV) moieties interact with the [S] atom of the ligand, while the other coordination sites are the carboxylate and the amide -CO groups. M?ssbauer Delta data showed that the diorganotin(IV) compounds adopt slightly distorted trigonal-bipyramidal (tbp) geometry. A single-crystal X-ray study was performed on the compound Me(2)Sn(cap): the Sn atom is five-coordinated in a distorted tbp environment, with two [O] atoms in the axial positions and the [S] and two [C] atoms in the equatorial (eq) plane. Each cap ligand coordinates to two different Sn atoms, and infinite zigzag chains are formed, directed parallel to each other and to the b axis of the unit cell. NMR (CDCl(3)) of the Me(2)Sn(IV) and n-Bu(2)Sn(IV) complexes indicated the presence of different oligomeric species.  相似文献   

8.
The oral glucose minimal model (OMM) measures insulin sensitivity (S(I)) and the glucose rate of appearance (R(a)) of ingested glucose in the presence of physiological changes of insulin and glucose concentrations. However, S(I) of OMM measures the overall effect of insulin on glucose utilization and glucose production. In this study we show that, by adding a tracer to the oral dose, e.g., of a meal, and by using the labeled version of OMM, OMM* to interpret the data, one can measure the selective effect of insulin on glucose disposal, S(I)*. Eighty-eight individuals underwent both a triple-tracer meal with the tracer-to-tracee clamp technique, providing a model-independent reference of the R(a) of ingested glucose (R(a meal)(ref)) and an insulin-modified labeled intravenous glucose tolerance test (IVGTT*). We show that OMM* provides not only a reliable means of tracing the R(a) of ingested glucose (R(a meal)) but also accurately measures S(I)*. We do so by comparing OMM* R(a meal) with the model-independent R(a meal)(ref) provided by the tracer-to-tracee clamp technique, while OMM* S(I)* is compared with both S(I)(* ref), obtained by using as known input R(a meal)(ref), and with S(I)* measured during IVGTT*.  相似文献   

9.
Substituent effects on the enantioselectivity for the lipase-catalyzed esterifications in organic solvents were studied by use of 2-(4-substituted phenoxy)propionic acids as the substrates with various substituents of H, F, Cl, CF(3), CH(3), CH(3)CH(2), and CH(3)O. The distinction in the behavior of their enantioselectivity was primarily responsible for the size effects of the substituents, although the substituents are far away from the stereocenter of the substrates. For the similar substituents in size, CH(3) and CF(3), however, their electronic effects played an important role in controlling the enantioselectivity. This variation of the enantioselectivity due to the electronic effects is also supported by the discussion based on the value of the Michaelis constant (K(m)) obtained. In addition, by raising the reaction temperature with enough water added to isopropyl ether as the reaction medium, the enantioselectivity is found to be dramatically enhanced for the substrate bearing CH(3)O group due to the strong electron-donating effect.  相似文献   

10.
The known complex trans-RuCl2(NBD)Py2 (1, NBD is norbornadiene, Py is pyridine) reacts with either (R)-BINAP ((R)-2, 2'-bis(diphenylphosphino)-1,1'-binaphthyl), (S;S)-Chiraphos ((2S;3S-(-)-2,3-bis(diphenylphosphino)butane), (S;S)-Skewphos ((2S;4S)-(-)-2,4-bis(diphenylphosphino)pentane), (R)-(S)-Josiphos ((R)-(-)-1-[(S)-2-(diphenylphosphino)ferrocenyl]ethyl-dicyclohexylpho sphine), (R;R)-Norphos ((2R;3R)-(-)-2, 3-bis(diphenylphosphino)bicyclo[2.2.1]hept-5-ene), or (R;R)-Me-DUPHOS ((-)-1,2-bis((2R;5R)-2, 5-dimethylphospholano)benzene) to generate in high yields the crystalline complexes trans-RuCl2(P-P*)Py2 (P-P* is the corresponding chiral bis(phosphine)). The complexes trans-RuCl2(P-P*)Py2 are active enantioselective hydrogenation catalysts for ketoesters and noncarboxylic olefins in the presence of small amounts of HBF4 (aq.). They are active for hydrogenation of carboxylic substrates in the presence of Et3N. Reaction of trans-RuCl2(P-P*)Py2 with (rac)-1,2-diphenylethylene-diamine (N-N*, either enantiomer) forms in good yields the corresponding compounds trans-RuCl2(P-P*)(N-N*). Representative hydrogenations with these catalysts are presented.  相似文献   

11.
A novel combination of experimental data and extensive computational modeling was used to explore probable protein-protein interactions between photoactivated rhodopsin (R*) and experimentally determined R*-bound structures of the C-terminal fragment of alpha-transducin (Gt(alpha)(340-350)) and its analogs. Rather than using one set of loop structures derived from the dark-adapted rhodopsin state, R* was modeled in this study using various energetically feasible sets of intracellular loop (IC loop) conformations proposed previously in another study. The R*-bound conformation of Gt(alpha)(340-350) and several analogs were modeled using experimental transferred nuclear Overhauser effect data derived upon binding R*. Gt(alpha)(340-350) and its analogs were docked to various conformations of the intracellular loops, followed by optimization of side-chain spatial positions in both R* and Gt(alpha)(340-350) to obtain low-energy complexes. Finally, the structures of each complex were subjected to energy minimization using the OPLS/GBSA force field. The resulting residue-residue contacts at the interface between R* and Gt(alpha)(340-350) were validated by comparison with available experimental data, primarily from mutational studies. Computational modeling performed for Gt(alpha)(340-350) and its analogs when bound to R* revealed a consensus of general residue-residue interactions, necessary for efficient complex formation between R* and its Gt(alpha) recognition motif.  相似文献   

12.
Visy J  Fitos I  Mády G  Urge L  Krajcsi P  Simonyi M 《Chirality》2002,14(8):638-642
The binding of bimoclomol enantiomers to human plasma, its components, as well as to plasma from monkey, dog, rat, and mouse was investigated by ultrafiltration and equilibrium dialysis. The considerably stronger binding of the (-)-(S)-enantiomer found in human plasma is due to the alpha(1)-acid glycoprotein (AAG) component. The binding parameters for AAG (n(R)K(R) = 1.3 x 10(4) M(-1) and n(S)K(S) = 1.0 x 10(5) M(-1)) revealed high enantioselectivity, while the binding to human serum albumin was found to be weak (nK = 5 x 10(3) M(-1)) and not stereoselective. (-)-(S)-Bimoclomol was extensively displaced in the presence of specific marker ligands for the "FIS" subfraction of human AAG. Comparative binding studies indicated considerable differences between plasma of the five species investigated.  相似文献   

13.
A series of Iron (Fe(II)) and manganese (Mn(II)) complexes of 1,3-substituted 5-(2-benzothiazolyl)formazans are reported. The crystal structures of the Fe(II) and Mn(II) complexes of 1,3-diphenyl-5-(2-benzothiazolyl)formazan are very similar and both contain a coordination sphere of four five-membered rings involving the N1 and N3 nitrogen atoms of the formazan chain and the N2 of the heterocyle resulting in a distorted octahedral structure in both cases. The distortions arise primarily from the spatial requirements of the bulky phenyl substituents.  相似文献   

14.
Sequence-specific hydrogen-1 NMR assignments were made to all of the 29 amino acid residues of reactive-site-hydrolyzed Cucurbita maxima trypsin inhibitor I (CMTI-I*) by the application of two-dimensional NMR (2D NMR) techniques, and its secondary structural elements (two tight turns, a 3(10)-helix, and a triple-stranded beta-sheet) were identified on the basis of short-range NOESY cross peaks and deuterium-exchange kinetics. These secondary structural elements are present in the intact inhibitor [Holak, T. A., Gondol, D., Otlewski, J., & Wilusz, T. (1989) J. Mol. Biol. 210, 635-648] and are unaffected by the hydrolysis of the reactive-site peptide bond between Arg5 and Ile6, in accordance with the earlier conclusion reached for CMTI-III* [Krishnamoorthi, R., Gong, Y.-X., Lin, C. S., & VanderVelde, D. (1992) Biochemistry 31, 898-904]. Chemical shifts of backbone hydrogen atoms, peptide NH's, and C alpha H's, of CMTI-I* were compared with those of the intact inhibitor, CMTI-I, and of the reactive-site-hydrolyzed, natural, E9K variant, CMTI-III*. Cleavage of the Arg5-Ile6 peptide bond resulted in changes of chemical shifts of most of the backbone atoms of CMTI-I, in agreement with the earlier results obtained for CMTI-III. Comparison of chemical shifts of backbone hydrogen atoms of CMTI-I* and CMTI-III* revealed no changes, except for residues Glu9 and His25. However, the intact forms of the same two proteins, CMTI-I and CMTI-III, showed small but significant perturbations of chemical shifts of residues that made up the secondary structural elements of the inhibitors.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Esterases, lipases, and serine proteases have been applied as versatile biocatalysts for preparing a variety of chiral compounds in industry via the kinetic resolution of their racemates. In order to meet this requirement, three approaches of enzyme engineering, medium engineering, and substrate engineering are exploited to improve the enzyme activity and enantioselectivity. With the hydrolysis of (R,S)-mandelates in biphasic media consisting of isooctane and pH 6 buffer at 55 degrees C as the model system, the strategy of combined substrate engineering and covalent immobilization leads to an increase of enzyme activity and enantioselectivity from V(S)/(E(t)) = 1.62 mmol/h g and V(S)/V(R) = 43.6 of (R,S)-ethyl mandelate (1) for a Klebsiella oxytoca esterase (named as SNSM-87 from the producer) to 16.7 mmol/h g and 867 of (R,S)-2-methoxyethyl mandelate (4) for the enzyme immobilized on Eupergit C 250L. The analysis is then extended to other (R,S)-2-hydroxycarboxylic acid esters, giving improvements of the enzyme performance from V(S)/(E(t)) = 1.56 mmol/h g and V(S)/V(R) = 41.9 of (R,S)-ethyl 3-chloromandelate (9) for the free esterase to 39.4 mmol/h g and 401 of (R,S)-2-methoxyethyl 3-chloromandelate (16) for the immobilized enzyme, V(S)/(E(t)) = 5.46 mmol/h g and V(S)/V(R) = 8.27 of (R,S)-ethyl 4-chloromandelate (10) for free SNSM-87 to 33.5 mmol/h g and 123 of (R,S)-methyl 4-chloromandelate (14) for the immobilized enzyme, as well as V(S)/(E(t)) = 3.0 mmol/h g and V(S)/V(R) = 7.94 of (R,S)-ethyl 3-phenyllactate (11) for the free esterase to 40.7 mmol/h g and 158 of (R,S)-2-methoxyethyl 3-phenyllactate (18) for the immobilized enzyme. The great enantioselectivty enhancement is rationalized from the alteration of ionization constants of imidazolium moiety of catalytic histidine for both enantiomers and conformation distortion of active site after the covalent immobilization, as well as the selection of leaving alcohol moiety via substrate engineering approach.  相似文献   

16.
To probe the interaction between transducin (G(t)) and photoactivated rhodopsin (R*), 14 analog peptides were designed and synthesized restricting the backbone of the R*-bound structure of the C-terminal 11 residues of G(t)alpha derived by transferred nuclear Overhauser effect (TrNOE) NMR. Most of the analogs were able to bind R*, supporting the TrNOE structure. Improved affinities of constrained peptides indicated that preorganization of the bound conformation is beneficial. Cys347 was found to be a recognition site; particularly, the free sulfhydryl of the side chain seems to be critical for R* binding. Leu349 was another invariable residue. Both Ile and tert-leucine (Tle) mutations for Leu349 significantly reduced the activity, indicating that the Leu side chain is in intimate contact with R*. The structure of R* was computer generated by moving helix 6 from its position in the crystal structure of ground-state rhodopsin (R) based on various experimental data. Seven feasible complexes were found when docking the TrNOE structure with R* and none with R. The analog peptides were modeled into the complexes, and their binding affinities were calculated. The predicted affinities were compared with the measured affinities to evaluate the modeled structures. Three models of the R*/G(t)alpha complex showed strong correlation to the experimental data.  相似文献   

17.
The extent of substrate enantioselectivity and regioselectivity of a series of aliphatic 2,3-dialkyl- and trialkylsubstituted oxiranes in their in vitro epoxide-hydrolase-catalyzed hydrolysis depends on the size of the alkyl residues and on the substitution pattern of the oxirane ring. The enzyme-catalyzed hydrolysis of cis-oxiranes, containing at least one methyl substituent, shows complete or nearly complete substrate enantioselectivity and regioselectivity with nucleophilic attack by water occurring with inversion of configuration at the methylsubstituted ring carbon atom of (S)-configuration. In the hydrolysis of the isomeric trans-oxiranes, both enantiomers are metabolized with a higher rate for the (2S;3S)-enantiomer. The conversion of trimethyloxirane occurs with high substrate enantioselectivity in favor of the (S)-enantiomer and with complete regioselectivity at the monomethylsubstituted ring carbon atom. The differentiation of the enantiotopic ring carbon atoms (product enantioselectivity) in the smallest aliphatic meso-oxirane, cis-2,3-dimethyloxirane, leads to (2R;3R)-butane-2,3-diol with ee = 86%. cis-2-Ethyl-3-propyloxirane, possessing alkyl residues larger than methyl, represents an extremely poor substrate in the epoxide-hydrolase-catalyzed hydrolysis process.  相似文献   

18.
Halohydrin dehalogenase from Agrobacterium radiobacter AD1 (HheC) is a valuable tool in the preparation of R enantiomers of epoxides and β-substituted alcohols. In contrast, the halohydrin dehalogenase from Arthrobacter sp. AD2 (HheA) shows a low S enantioselectivity toward most aromatic substrates. Here, three amino acids (V136, L141, and N178) located in the two neighboring active-site loops of HheA were proposed to be the key residues for controlling enantioselectivity. They were subjected to saturation mutagenesis aimed at evolving an S-selective enzyme. This led to the selection of two outstanding mutants (the V136Y/L141G and N178A mutants). The double mutant displayed an inverted enantioselectivity (from S enantioselectivity [E(S)] = 1.7 to R enantioselectivity [E(R)] = 13) toward 2-chloro-1-phenylethanol without compromising enzyme activity. Strikingly, the N178A mutant showed a large enantioselectivity improvement (E(S) > 200) and a 5- to 6-fold-enhanced specific activity toward (S)-2-chloro-1-phenylethanol. Further analysis revealed that those mutations produced some interference for the binding of nonfavored enantiomers which could account for the observed enantioselectivities. Our work demonstrated that those three active-site residues are indeed crucial in modulating the enantioselectivity of HheA and that a semirational design strategy has great potential for rapid creation of novel industrial biocatalysts.  相似文献   

19.
Halohydrin dehalogenase (HheC) from Agrobacterium radiobacter AD1 catalyzes the reversible intramolecular nucleophilic displacement of a halogen by a hydroxyl group in vicinal haloalcohols, producing the corresponding epoxides. The enzyme displays high enantioselectivity toward some aromatic halohydrins. To understand the kinetic mechanism and enantioselectivity of the enzyme, steady-state and pre-steady-state kinetic analysis was performed with p-nitro-2-bromo-1-phenylethanol (PNSHH) as a model substrate. Steady-state kinetic analyses indicated that the k(cat) of the enzyme with the (R)-enantiomer (22 s(-1)) is 3-fold higher than with the (S)-enantiomer and that the K(m) for the (R)-enantiomer (0.009 mM) is about 45-fold lower than that for the (S)-enantiomer, resulting in a high enantiopreference for the (R)-enantiomer. Product inhibition studies revealed that HheC follows an ordered Uni Bi mechanism for both enantiomers, with halide as the first product to be released. To identify the rate-limiting step in the catalytic cycle, pre-steady-state experiments were performed using stopped-flow and rapid-quench methods. The results revealed the existence of a pre-steady-state burst phase during conversion of (R)-PNSHH, whereas no such burst was observed with the (S)-enantiomer. This indicates that a product release step is rate-limiting for the (R)-enantiomer but not for the (S)-enantiomer. This was further examined by doing single-turnover experiments, which revealed that during conversion of the (R)-enantiomer the rate of bromide release is 21 s(-1). Furthermore, multiple turnover analyses showed that the binding of (R)-PNSHH is a rapid equilibrium step and that the rate of formation of product ternary complex is 380 s(-1). Taken together, these findings enabled the formulation of an ordered Uni Bi kinetic mechanism for the conversion of (R)-PNSHH by HheC in which all of the rate constants are obtained. The high enantiopreference for the (R)-enantiomer can be explained by weak substrate binding of the (S)-enantiomer and a lower rate of reaction at the active site.  相似文献   

20.
Kao MF  Lu PY  Kao JY  Wang PY  Wu AC  Tsai SW 《Chirality》2012,24(1):60-66
The best reaction condition of Candida antartica lipase B as biocatalyst, 3-(2-pyridyl)pyrazole as leaving azole, and water-saturated methyl t-butyl ether as reaction medium at 45°C were first selected for performing the hydrolytic resolution of (R,S)-2-(4-chlorophenoxyl) azolides (1-4). In comparison with the kinetic resolution of (R,S)-2-phenylpropionyl 3-(2-pyridyl)pyrazolide or (R,S)-α-methoxyphenylacetyl 3-(2-pyridyl)pyrazolide at the same reaction condition, excellent enantioselectivity with more than two order-of-magnitudes higher activity for each enantiomer was obtained. The resolution was then extended to other (R,S)-3-(2-pyridyl)pyrazolides (5-7) containing 2-chloro, 3-chloro, or 2,4-dichloro substituent, giving good (E > 48) to excellent (E > 100) enantioselectivity. The thermodynamic analysis for 1, 2, and 4-7 demonstrates profound effects of the acyl or leaving moiety on varying enthalpic and entropic contributions to the difference of Gibbs free energies. A thorough kinetic analysis further indicates that on the basis of 6, the excellent enantiomeric ratio for 4 and 7 is due to the higher reactivity of (S)-4 and lower reactivity of (R)-7, respectively.  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