首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Born TL  Franklin M  Blanchard JS 《Biochemistry》2000,39(29):8556-8564
The first unique step in bacterial and plant methionine biosynthesis involves the acylation of the gamma-hydroxyl of homoserine. In Haemophilus influenzae, acylation is accomplished via an acetyl-CoA-dependent acetylation catalyzed by homoserine transacetylase. The activity of this enzyme regulates flux of homoserine into multiple biosynthetic pathways and, therefore, represents a critical control point for cell growth and viability. We have cloned homoserine transacetylase from H. influenzae and present the first detailed enzymatic study of this enzyme. Steady-state kinetic experiments demonstrate that the enzyme utilizes a ping-pong kinetic mechanism in which the acetyl group of acetyl-CoA is initially transferred to an enzyme nucleophile before subsequent transfer to homoserine to form the final product, O-acetylhomoserine. The maximal velocity and V/K(homoserine) were independent of pH over the range of values tested, while V/K(acetyl)(-)(CoA) was dependent upon the ionization state of a single group exhibiting a pK value of 8.6, which was required to be protonated. Solvent kinetic isotope effect studies yielded inverse effects of 0.75 on V and 0.74 on V/K(CoA) on the reverse reaction and effects of 1.2 on V and 1.7 on V/K(homoserine) on the forward reaction. Direct evidence for the formation of an acetyl-enzyme intermediate was obtained using rapid-quench labeling studies. On the basis of these observations, we propose a chemical mechanism for this important member of the acyltransferase family and contrast its mechanism with that of homoserine transsuccinylase.  相似文献   

2.
3.
The methionine biosynthetic pathway found in bacteria is controlled at the first step, acylation of the γ-hydroxyl of homoserine. This reaction is catalyzed by one of two unique enzymes, homoserine transacetylase or homoserine transsuccinylase, which have no amino acid sequence similarity. We cloned, expressed, and purified homoserine transsuccinylase from the thermophilic bacterium Thermotoga maritima. Substrate specificity experiments demonstrated that acetyl-coenzyme A (CoA) is the preferred acyl donor and is used at least 30-fold more efficiently than succinyl-CoA. Steady-state kinetic experiments confirm that the enzyme utilizes a ping-pong kinetic mechanism in which the acetate group of acetyl-CoA is initially transferred to an enzyme nucleophile before subsequent transfer to homoserine. The maximal velocity, V/K acetyl-CoA and V/K homoserine, all exhibited bell-shaped pH curves with apparent pKs of 6.0–6.9 and 8.2–8.8. The enzyme was inactivated by iodoacetamide in a pH-dependent manner, with an apparent pK of 6.3, suggesting the presence of an active-site cysteine residue which forms an acetyl-enzyme thioester intermediate during catalytic turnover, similar to observations with other transsuccinylases. In addition, the enzyme is highly stable at elevated temperatures, maintaining full activity at 70°C. Taken together, these data suggest that the T. maritima enzyme functions biochemically as a transacetylase, despite having the sequence of a transsuccinylase.  相似文献   

4.
Homoserine trans-succinylase is the first enzyme in methionine biosynthesis of Escherichia coli and catalyzes the activation of homoserine via a succinylation reaction. The in vivo activity of this enzyme is subject to tight regulation by several mechanisms, including repression and activation of gene expression, feedback inhibition, temperature regulation and proteolysis. This complex regulation reflects the key role of this enzyme in bacterial metabolism. Here, we demonstrate--using proteomics and high-resolution mass spectrometry--that succinyl is covalently bound to one of the two adjacent lysine residues at positions 45 and 46. Replacing these lysine residues by alanine abolished the enzymatic activity. These findings position the lysine residues, one of which is conserved, at the active site.  相似文献   

5.
Substrate specificity within the family of enzymes designated as homoserine transsuccinylases is variable, with some organisms utilizing succinyl-CoA and other organisms utilizing acetyl-CoA. In this study it is shown that the enzyme from Bacillus cereus uses acetyl-CoA as its acyl donor, but its catalytic rate is significantly lower than other HTS family members. BcHTS is inactivated by both iodoacetamide and diethyl pyrocarbonate and the enzyme can be partially protected from inactivation by the presence of succinyl-CoA. This leads to the conclusion that BcHTS can bind both acetyl-CoA and succinyl-CoA and suggests that it may represent an intermediate between the succinate-transferring HTS family members and the acetate-transferring HTS family members. The B. cereus enzyme was unable to rescue growth of an Escherichia coli strain lacking a functional transsuccinylase, however.  相似文献   

6.
Isotope exchange kinetics at chemical equilibrium have been used to investigate the kinetic mechanism of homoserine dehydrogenase (EC 1.1.1.3) of the (Thr-sensitive) aspartokinase/homoserine dehydrogenase-I multifunctional enzyme from E. coli. For the reaction (L-ASA + NADPH + H+ = L-Hse + NADP+), at pH 9.0, 37 degrees C, Keq = 100 (+/- 20). Under these conditions, the rate for exchange of [14C]-L-homoserine (Hse) in equilibrium L-aspartate-beta-semialdehyde (ASA) is nearly twice that for the [3H]-NADP+ in equilibrium NADPH exchange. This indicates that covalent interconversion between reactants and products bound in the active site cannot be rate-limiting. Upon variation of the concentrations of all four substrates in constant ratio at equilibrium (to minimize dead-end complex formation), the Hse in equilibrium ASA exchange increased smoothly toward a maximum. In contrast, the NADP+ in equilibrium NADPH exchange rate increased to a maximum value at partial saturation, then decreased to approximately half the maximum rate. These data are consistent with a preferred-order random kinetic mechanism in which the dominant pathway involves association of NADPH prior to L-ASA and dissociation of L-Hse prior to NADP+.  相似文献   

7.
Homoserine transacetylase catalyzes one of the required steps in the biosynthesis of methionine in fungi and several bacteria. We have determined the crystal structure of homoserine transacetylase from Haemophilus influenzae to a resolution of 1.65 A. The structure identifies this enzyme to be a member of the alpha/beta-hydrolase structural superfamily. The active site of the enzyme is located near the end of a deep tunnel formed by the juxtaposition of two domains and incorporates a catalytic triad involving Ser143, His337, and Asp304. A structural basis is given for the observed double displacement kinetic mechanism of homoserine transacetylase. Furthermore, the properties of the tunnel provide a rationale for how homoserine transacetylase catalyzes a transferase reaction vs hydrolysis, despite extensive similarity in active site architecture to hydrolytic enzymes.  相似文献   

8.
Dimers of aspartokinase I/homoserine dehydrogenase I from Escherichia coli K 12 have been isolated under very mild conditions. The dimers which cannot be distinguished from the tetramers by their kinetic properties, reassociate in the presence of potassium ions or L-aspartate. The selective sensitivity of aspartokinase I/homoserine dehydrogenase I to mild proteolytic digestion of dimers has been used to probe the reassociation reaction under the conditions of aspartokinase assay. We demonstrate that rapid reassociation occurs and that the protein species present in the assay when dimers are used to test the activity is tetrameric. These results confirm the previously proposed model for the subunit association of aspartokinase I/homoserine dehydrogenase I.  相似文献   

9.
Characterization of ligand-induced states of maize homoserine dehydrogenase   总被引:1,自引:0,他引:1  
The threonine-sensitive homoserine dehydrogenase (L-homoserine: NAD(P)+ oxido-reductase), isolated from seedlings of Zea mays L., is characterized by variable kinetic and regulatory properties. Previous analysis of this enzyme suggested that it is capable of ligand-mediated interconversions among four kinetically distinct states (S. Krishnaswamy and J. K. Bryan (1983) Arch. Biochem. Biophys. 222, 449-463). These forms of the enzyme have been identified and found to differ in oligomeric configuration and conformation. In the presence of KCl and threonine a rapid equilibrium among three species of the enzyme (B, T, and K) is established. Each of these species can undergo a unique slow transition to a steady-state form under assay conditions. Results obtained from gel-filtration chromatography and sucrose density centrifugation indicate that the B and steady-state forms are tetramers and the T and K states are dimers. Evidence is presented to indicate that the rapid conversion from one dimeric species to the other can only occur via formation of the tetrameric B state. Chromatography under reacting-enzyme conditions provides direct support for the slow formation of a common steady-state species from any one of the other forms of the enzyme. The rate of transition is influenced by threonine, homoserine, NAD+, and, for transitions involving association reactions, by enzyme concentration. Small, reproducible differences in the apparent size of the T and K forms, and the B and steady-state species, are attributed to changes in conformation. This conclusion is supported by differential susceptibility of the enzymic states to proteolytic inactivation, by different rates of inactivation by dithio-bis-nitrobenzoate, and by alterations in their thermal stability. In addition, the B, T, and K states of the enzyme exhibit unique intrinsic fluorescence spectra. Spectral changes are shown to closely parallel changes in kinetic and hysteretic properties of the enzyme. The results of diverse methods of analysis are internally consistent, and provide considerable support for the conclusion that this pleiotropic regulatory enzyme can exist in any of several physically distinct states.  相似文献   

10.
Patel MP  Liu WS  West J  Tew D  Meek TD  Thrall SH 《Biochemistry》2005,44(50):16753-16765
Beta-ketoacyl-acyl carrier protein reductase (KACPR) catalyzes the NADPH-dependent reduction of beta-ketoacyl-acyl carrier protein (AcAc-ACP) to generate (3S)-beta-hydroxyacyl-ACP during the chain-elongation reaction of bacterial fatty acid biosynthesis. We report the evaluation of the kinetic and chemical mechanisms of KACPR using acetoacetyl-CoA (AcAc-CoA) as a substrate. Initial velocity, product inhibition, and deuterium kinetic isotope effect studies were consistent with a random bi-bi rapid-equilibrium kinetic mechanism of KACPR with formation of an enzyme-NADP(+)-AcAc-CoA dead-end complex. Plots of log V/K(NADPH) and log V/K(AcAc)(-)(CoA) indicated the presence of a single basic group (pK = 5.0-5.8) and a single acidic group (pK = 8.0-8.8) involved in catalysis, while the plot of log V vs pH indicated that at high pH an unprotonated form of the ternary enzyme complex was able to undergo catalysis. Significant and identical primary deuterium kinetic isotope effects were observed for V (2.6 +/- 0.4), V/K(NADPH) (2.6 +/- 0.1), and V/K(AcAc)(-)(CoA) (2.6 +/- 0.1) at pH 7.6, but all three values attenuated to values of near unity (1.1 +/- 0.03 or 0.91 +/- 0.02) at pH 10. Similarly, the large alpha-secondary deuterium kinetic isotope effect of 1.15 +/- 0.02 observed for [4R-(2)H]NADPH on V/K(AcAc)(-)(CoA) at pH 7.6 was reduced to a value of unity (1.00 +/- 0.04) at high pH. The complete analysis of the pH profiles and the solvent, primary, secondary, and multiple deuterium isotope effects were most consistent with a chemical mechanism of KACPR that is stepwise, wherein the hydride-transfer step is followed by protonation of the enolate intermediate. Estimations of the intrinsic primary and secondary deuterium isotope effects ((D)k = 2.7, (alpha)(-D)k = 1.16) and the correspondingly negligible commitment factors suggest a nearly full expression of the intrinsic isotope effects on (D)V/K and (alpha)(-D)V/K, and are consistent with a late transition state for the hydride transfer step. Conversely, the estimated intrinsic solvent effect ((D)2(O)k) of 5.3 was poorly expressed in the experimentally derived parameters (D)2(O)V/K and (D)2(O)V (both = 1.2 +/- 0.1), in agreement with the estimation that the catalytic commitment factor for proton transfer to the enolate intermediate is large. Such detailed knowledge of the chemical mechanism of KAPCR may now help guide the rational design of, or inform screening assay-design strategies for, potent inhibitors of this and related enzymes of the short chain dehydrogenase enzyme class.  相似文献   

11.
Zheng R  Blanchard JS 《Biochemistry》2000,39(13):3708-3717
Ketopantoate reductase (EC 1.1.1.169) catalyzes the NADPH-dependent reduction of alpha-ketopantoate to form D-(-)-pantoate in the pantothenate/coenzyme A biosynthetic pathway. The enzyme encoded by the panE gene from E. coli K12 was overexpressed and purified to homogeneity. The native enzyme exists in solution as a monomer with a molecular mass of 34 000 Da. The steady-state initial velocity and product inhibition patterns are consistent with an ordered sequential kinetic mechanism in which NADPH binding is followed by ketopantoate binding, and pantoate release precedes NADP(+) release. The pH dependence of the kinetic parameters V and V/K for substrates in both the forward and reverse reactions suggests the involvement of a single general acid/base in the catalytic mechanism. An enzyme group exhibiting a pK value of 8.4 +/- 0.2 functions as a general acid in the direction of the ketopantoate reduction, while an enzyme group exhibiting a pK value of 7.8 +/- 0.2 serves as a general base in the direction of pantoate oxidation. The stereospecific transfer of the pro-S hydrogen atom of NADPH to the C-2 position of ketopantoate was demonstrated by (1)H NMR spectroscopy. Primary deuterium kinetic isotope effects of 1.3 and 1.5 on V(for) and V/K(NADPH), respectively, and 2.1 and 1.3 on V(rev) and V/K(HP), respectively, suggest that hydride transfer is not rate-limiting in catalysis. Solvent kinetic isotope effects of 1.3 on both V(for) and V/K(KP), and 1.4 and 1.5 on V(rev) and V/K(HP), respectively, support this conclusion. The apparent equilibrium constant, K(eq)', of 676 at pH 7.5 and the standard free energy change, DeltaG, of -14 kcal/mol suggest that ketopantoate reductase reaction is very favorable in the physiologically important direction of pantoate formation.  相似文献   

12.
Homoserine acyltransferases catalyze the commitment step to methionine and other important biological precursors which make this class of enzymes essential for the survival of bacteria, plants and fungi. This class of enzymes is not found in humans, making them an attractive new target for antimicrobial design. Homoserine O-succinyltransferase (HST) is a representative from this class, with little known about the key amino acids involved in substrate specificity and catalysis. HST from Escherichia coli has been cloned, purified and kinetically characterized. Through site-directed mutagenesis and steady-state kinetic studies the residues that comprise a catalytic triad for HST, the catalytic cysteine nucleophile, an active site acid-base histidine, and the base orienting glutamate, have been identified and characterized. Several residues which confer substrate specificity for both homoserine and succinyl-CoA have also been identified and kinetically evaluated. Mutations of an active site glutamate to either aspartate or alanine drastically increase the K(m) for homoserine, assigning this glutamate to a binding role for the alpha-amino group of homoserine. An active site arginine orients the carboxyl moiety of homoserine, while the carboxyl moiety of succinyl-CoA is positioned for catalysis by a lysine residue. Removing functionality at either of these positions alters the enzyme's ability to effectively utilize homoserine or succinyl-CoA, respectively, reflected in an increased K(m) and decreased catalytic efficiency. The data presented here provides new details of the catalytic mechanism of succinyltransferases, resolves a controversy between alternative mechanistic hypotheses, and provides a starting point for the development of selective inhibitors of HST.  相似文献   

13.
Quorum sensing is the process by which bacteria alter gene regulation in response to their population density. The enzymatic inactivation of quorum signals has shown promise for use in genetically modified organisms resistant to pathogens. We recently characterized the ability of a cytochrome P450, P450BM-3, to oxidize the quorum sensing signals known as acyl homoserine lactones. The oxidation of the acyl homoserine lactones reduced their activity as quorum signals. The enzyme also oxidized the inactive lactonolysis products, acyl homoserines. The enzyme showed similar binding affinity for the acyl homoserine lactones and acyl homoserines. The latter reaction may lead to problems when lactonases and the P450-dependent system are used in tandem, as oxidation of the acyl homoserines produced by lactonolysis in vivo may compete with acyl homoserine lactone oxidation by the cytochrome P450. We report here that a single mutation (R47S) in P450BM-3 is capable of increasing the acyl homoserine lactone: acyl homoserine substrate binding selectivity of the enzyme nearly 250-fold, reducing the potential for competition by acyl homoserines and significantly enhancing the potential for use of P450BM-3 as part of a pathogen resistance system in genetically modified crops.  相似文献   

14.
Argyrou A  Blanchard JS 《Biochemistry》2001,40(38):11353-11363
The gene encoding dihydrolipoamide dehydrogenase from Mycobacterium tuberculosis, Rv0462, was expressed in Escherichia coli and the protein purified to homogeneity. The 49 kDa polypeptide forms a homodimer containing one tightly bound molecule of FAD/monomer. The results of steady-state kinetic analyses using several reduced pyridine nucleotide analogs and a variety of electron acceptors, and the ability of the enzyme to catalyze the transhydrogenation of NADH and thio-NAD(+) in the absence of D,L-lipoamide, demonstrated that the enzyme uses a ping-pong kinetic mechanism. Primary deuterium kinetic isotope effects on V and V/K at pH 7.5 using NADH deuterated at the C(4)-proS position of the nicotinamide ring are small [(D)(V/K)(NADH) = 1.12 +/- 0.15, (D)V(app) = 1.05 +/- 0.07] when D,L-lipoamide is the oxidant but large and equivalent [(D)(V/K)(NADH) = (D)V = 2.95 +/- 0.03] when 5-hydroxy-1,4-naphthoquinone is the oxidant. Solvent deuterium kinetic isotope effects at pH 5.8, using APADH as the reductant, are inverse with (D)(V/K)(APADH) = 0.73 +/- 0.03, (D)(V/K)(Lip(S))2 = 0.77 +/- 0.03, and (D)V(app) = 0.77 +/- 0.01. Solvent deuterium kinetic isotope effects with 4,4-dithiopyridine (DTP), the 4-thiopyridone product of which requires no protonation, are also inverse with (D)(V/K)(APADH) = 0.75 +/- 0.06, (D)(V/K)(DTP) = 0.71 +/- 0.02, and (D)V(app) = 0.56 +/- 0.15. All proton inventories were linear, indicating that a single proton is being transferred in the solvent isotopically sensitive step. Taken together, these results suggest that (1) the reductive half-reaction (hydride transfer from NADH to FAD) is rate limiting when a quinone is the oxidant, and (2) deprotonation of enzymic thiols, most likely Cys(46) and Cys(41), limits the reductive and oxidative half-reactions, respectively, when D,L-lipoamide is the oxidant.  相似文献   

15.
Bacteria and yeast rely on either homoserine transsuccinylase (HTS, metA) or homoserine transacetylase (HTA; met2) for the biosynthesis of methionine. Although HTS and HTA catalyze similar chemical reactions, these proteins are typically unrelated in both sequence and three-dimensional structure. Here we present the 2.0 A resolution x-ray crystal structure of the Bacillus cereus metA protein in complex with homoserine, which provides the first view of a ligand bound to either HTA or HTS. Surprisingly, functional analysis of the B. cereus metA protein shows that it does not use succinyl-CoA as a substrate. Instead, the protein catalyzes the transacetylation of homoserine using acetyl-CoA. Therefore, the B. cereus metA protein functions as an HTA despite greater than 50% sequence identity with bona fide HTS proteins. This result emphasizes the need for functional confirmation of annotations of enzyme function based on either sequence or structural comparisons. Kinetic analysis of site-directed mutants reveals that the B. cereus metA protein and the E. coli HTS share a common catalytic mechanism. Structural and functional examination of the B. cereus metA protein reveals that a single amino acid in the active site determines acetyl-CoA (Glu-111) versus succinyl-CoA (Gly-111) specificity in the metA-like of acyltransferases. Switching of this residue provides a mechanism for evolving substrate specificity in bacterial methionine biosynthesis. Within this enzyme family, HTS and HTA activity likely arises from divergent evolution in a common structural scaffold with conserved catalytic machinery and homoserine binding sites.  相似文献   

16.
Heparan sulfate (HS) glycosaminoglycans are a structurally diverse class of complex biomolecules that modulate many important events at the cell surface and within the extracellular matrix and whose structural heterogeneity derives largely from the sequence-specific N- and O-sulfations catalyzed by an extensive repertoire of sulfating enzymes. We have expressed the human heparan sulfate 3-OST-1 isoform in Escherichia coli and subsequently purified a soluble, active enzyme. To assess its functionality, we determined the kinetic parameters for the recombinant 3-O-sulfotransferase-1 using a radiochemical assay that directly measures the 3-O-sulfation of unlabeled bovine kidney heparan sulfate in vitro using [(35)S]PAPS as the sulfate donor. The apparent K(m) values measured were in the low micromolar range (K(HS)(m) = 4.3 microM; K(PAPS)(m) = 38.6 microM); V(max) values of 18 and 21 pmol sulfate/min/pmol of enzyme for HS and PAPS, respectively. These values were compared with kinetic parameters likewise measured for recombinant 3-OST-1 purified from baculovirus-infected sf9 cells. The two enzymes appear to modify heparan sulfate in vitro to roughly the same extent and with comparable specificities. The expression of 3-OST-1 in E. coli represents an important step in subsequent structure-function studies.  相似文献   

17.
Marks GT  Susler M  Harrison DH 《Biochemistry》2004,43(13):3802-3813
Two detailed mechanisms [Marks et al. (2001) Biochemistry 40, 6805] have been proposed to explain the activity of methylglyoxal synthase (MGS), a homohexameric allosterically regulated enzyme that catalyzes the elimination of phosphate from DHAP to form enol pyruvaldehyde. This enol then tautomerizes to methylglyoxal in solution. In one of these mechanisms His 98 plays an obligate role in the transfer of a proton from the O(3) oxygen of DHAP to the O2 oxygen. To test this hypothesized mechanism, the variants H98N and H98Q were expressed and purified. Relative to the wild-type enzyme, the H98N variant shows a 50-fold decrease in k(cat) with all other kinetic parameters (i.e., K(m), K(PGA), etc.) being nearly the same. By contrast, the apparent catalytic rate for the H98Q variant is 250-fold lower than that of the wild-type enzyme. Inorganic phosphate acts as a competitive inhibitor (with a K(i) of 15 microM) rather than as an allosteric-type inhibitor as it does in the wild-type enzyme, and the competitive inhibitor phosphoglyolate (PGA) acts as an activator of this variant. These facts are explained by a shift in the conformational equilibrium toward an "inactive" state. When activation by PGA is accounted for, the catalytic rate for the "active" state of H98Q is estimated to be only 6-fold less than that of the wild-type enzyme, and thus His 98 is not essential for catalysis. Primary deuterium isotope effect data were measured for the wild-type enzyme and the two variants. At pH 7.0, the (D)V isotope effect (1.5) and the absence of a (D)(V/K) isotope effect for the wild-type enzyme suggest that the rate for the isotopically sensitive step is partially rate limiting but greater than the rate of substrate dissociation. Both the (D)V (2.0) and (D)(V/K) (3.4) isotope effects were more pronounced in the H98N variant, while the H98Q variant displayed an unusual inverse (D)V (0.8) isotope effect and a normal (D)(V/K) (1.5) isotope effect. The X-ray crystal structures of PGA bound to the H98Q and H98N variants were both determined to a resolution of 2.2 A. These mutations had little effect on the overall structure. However, the pattern of hydrogen bonding in the active site suggests an explanation as to how in the wild-type and H98N mutated enzymes the "inactive to active" equilibrium lies toward the active state, while with the H98Q mutated enzyme the equilibrium lies toward the inactive state. Thus, the role of His 98 appears to be more as a regulator of the enzyme's conformation rather than as a critical contributor to the catalytic mechanism.  相似文献   

18.
抗坏血酸对酵母蔗糖酶的激活动力学研究   总被引:2,自引:0,他引:2  
采用甲苯自溶法从鲜酵母中提取了蔗糖酶,并用乙醇分级及DEAE-纤维素柱层析进行了纯化,用PAG凝胶电泳作了纯度鉴定,在pH5.0,30℃条件下进行了酶反应,用双倒数作图法测出其Km=2.1×10-2mol/L,Vmax=0.26(每分钟的光密度值).在此系统中,加入不同浓度的抗坏血酸(Vit.C),发现其具有激活作用并存在量效关系.双倒数作图显示:酶的表观Vmax(Vp)随抗坏血酸浓度的增加而增大,但其表观Km(Kp)不变(Kp=Km).经实验结果分析,推论出抗坏血酸激活作用的酶促反应方程式,并推导出反应速度公式  相似文献   

19.
Homoserine kinase was purified to apparent homogeneity from a derepressed strain of Escherichia coli K12, using standard fractionation techniques. It is a dimer (Mr = 60000) composed of apparently identical polypeptide chains (Mr = 29000). Its amino acid composition and N-terminal sequence have been determined. L-Threonine is a competitive inhibitor of the substrate L-homoserine; this inhibition is straighforward and shows no sign of co-operativity. Evidence is presented that homoserine and threonine bind to the same site of this non-allosteric enzyme. The binding of homoserine and threonine can also be studied by difference spectroscopy; the latter studies reveal an unexpected effect of magnesium ions, which might be the basis for the unusual high Mg2+ requirement for optimal enzyme reaction.  相似文献   

20.
Fillgrove KL  Anderson VE 《Biochemistry》2001,40(41):12412-12421
The chemical mechanism of the 2,4-dienoyl-CoA reductase (EC 1.3.1.34) from rat liver mitochondria has been investigated. This enzyme catalyzes the NADPH-dependent reduction of 2,4-dienoyl-coenzyme A (CoA) thiolesters to the resulting trans-3-enoyl-CoA. Steady-state kinetic parameters for trans-2,trans-4-hexadienoyl-CoA and 5-phenyl-trans-2,trans-4-pentadienoyl-CoA were determined and demonstrated that the dienoyl-CoA and NADPH bind to the 2,4-dienoyl-CoA reductase via a sequential kinetic mechanism. Kinetic isotope effect studies and the transient kinetics of substrate binding support a random order of nucleotide and dienoyl-CoA addition. The large normal solvent isotope effects on V/K ((D)(2)(O)V/K) and V ((D)(2)(O)V) for trans-2,trans-4-hexadienoyl-CoA reduction indicate that a proton transfer step is rate limiting for this substrate. The stability gained by conjugating the phenyl ring to the diene in PPD-CoA results in the reversal of the rate-determining step, as evidenced by the normal isotope effects on V/K(CoA) ((D)V/K(CoA)) and V/K(NADPH) ((D)V/K(NADPH)). The reversal of the rate-determining step was supported by transient kinetics where a burst was observed for the reduction of trans-2,trans-4-hexadienoyl-CoA but not for 5-phenyl-trans-2,trans-4-pentadienoyl-CoA reduction. The chemical mechanism is stepwise where hydride transfer from NADPH occurs followed by protonation of the observable dienolate intermediate, which has an absorbance maximum at 286 nm. The exchange of the C alpha protons of trans-3-decenoyl-CoA, catalyzed by the 2,4-dienoyl-CoA reductase, in the presence of NADP(+) suggests that formation of the dienolate is catalyzed by the enzyme active site.  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