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 15 毫秒
1.
Fourteen independent mutants of Saccharomyces cerevisiae defective in sn-glycerol-3-phosphate acyltransferase activity were isolated using a colony autoradiographic screening technique. All 14 mutants were similarly defective in dihydroxyacetone phosphate acyltransferase activity. The mutations were recessive and fell into a single complementation group. Tetrad analysis gave results consistent with mutations in a single nuclear gene affecting both activities. sn-Glycerol-3-phosphate acyltransferase activity from different mutant strains exhibited different substrate dependencies and differing responses to temperature, detergent, and pH. In each case, the response of the dihydroxyacetone phosphate acyltransferase activity was similar to that of the sn-glycerol-3-phosphate acyltransferase. These results are consistent with the mutations occurring in the structural gene. The data also establish that the predominant dihydroxyacetone phosphate acyltransferase activity in yeast is a second activity of the sn-glycerol-3-phosphate acyltransferase.  相似文献   

2.
Kroemer M  Merkel I  Schulz GE 《Biochemistry》2003,42(36):10560-10568
The structure of L-rhamnulose-1-phosphate aldolase has been established at 1.35 A resolution in a crystal form that was obtained by a surface mutation and has one subunit of the C(4)-symmetric tetramer in the asymmetric unit. It confirms an earlier 2.7 A resolution structure which was determined in a complicated crystal form with 20 subunits per asymmetric unit. The chain fold and the active center are similar to those of L-fuculose-1-phosphate aldolase and L-ribulose-5-phosphate 4-epimerase. The active center similarity is supported by a structural comparison of all three enzymes and by the binding mode of the inhibitor phosphoglycolohydroxamate at the site of the product dihydroxyacetone phosphate for the two aldolases. The sensitivity of the catalytic rate to several mutations and a comparison with the established mechanism of the related aldolase give rise to a putative catalytic mechanism. This mechanism involves the same binding mode of the second product L-lactaldehyde in both aldolases, except for a 180 degrees flip of the aldehyde group distinguishing between the two epimers rhamnulose and fuculose. The N-terminal domain exhibits a correlated anisotropic mobility that channels the isotropic Brownian motion into a directed movement of the catalytic base and the substrate phosphate on the N-domain toward the zinc ion and the lactaldehyde on the C-terminal domain. We suggest that this movement supports the catalysis mechanically.  相似文献   

3.
The inhibition of triosephosphate isomerase (TPI) in glycolysis by the pyruvate kinase (PK) substrate phosphoenolpyruvate (PEP) results in a newly discovered feedback loop that counters oxidative stress in cancer and actively respiring cells. The mechanism underlying this inhibition is illuminated by the co-crystal structure of TPI with bound PEP at 1.6 Å resolution, and by mutational studies guided by the crystallographic results. PEP is bound to the catalytic pocket of TPI and occludes substrate, which accounts for the observation that PEP competitively inhibits the interconversion of glyceraldehyde-3-phosphate and dihydroxyacetone phosphate. Replacing an isoleucine residue located in the catalytic pocket of TPI with valine or threonine altered binding of substrates and PEP, reducing TPI activity in vitro and in vivo. Confirming a TPI-mediated activation of the pentose phosphate pathway (PPP), transgenic yeast cells expressing these TPI mutations accumulate greater levels of PPP intermediates and have altered stress resistance, mimicking the activation of the PK–TPI feedback loop. These results support a model in which glycolytic regulation requires direct catalytic inhibition of TPI by the pyruvate kinase substrate PEP, mediating a protective metabolic self-reconfiguration of central metabolism under conditions of oxidative stress.  相似文献   

4.
Electrophilic catalysis by histidine-95 in triosephosphate isomerase has been probed by using Fourier transform infrared spectroscopy and X-ray crystallography. The carbonyl stretching frequency of dihydroxyacetone phosphate bound to the wild-type enzyme is known to be 19 cm-1 lower (at 1713 cm-1) than that of dihydroxyacetone phosphate free in solution (at 1732 cm-1), and this decrease in stretching frequency has been ascribed to an enzymic electrophile that polarizes the substrate carbonyl group toward the transition state for the enolization. Infrared spectra of substrate bound to two site-directed mutants of yeast triosephosphate isomerase in which histidine-95 has been changed to glutamine or to asparagine show unperturbed carbonyl stretching frequencies between 1732 and 1742 cm-1. The lack of carbonyl polarization when histidine-95 is removed suggests that histidine-95 is indeed the catalytic electrophile, at least for dihydroxyacetone phosphate. Kinetic studies of the glutamine mutant (H95Q) have shown that the enzyme follows a subtly different mechanism of proton transfers involving only a single acid-base catalytic group. These findings suggest an additional role for histidine-95 as a general acid-base catalyst in the wild-type enzyme. The X-ray crystal structure of the H95Q mutant with an intermediate analogue, phosphoglycolohydroxamate, bound at the active site has been solved to 2.8-A resolution, and this structure clearly implicates glutamate-165, the catalytic base in the wild-type isomerase, as the sole acid-base catalyst for the mutant enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Methylglyoxal synthetase, which catalyzes the conversion of dihydroxyacetone phosphate to methylglyoxal and inorganic phosphate, has been isolated and crystalized in good yields from Proteus vulgaris. The enzyme was shown to be homogeneous by a variety of criteria and was found to be a dimer (Mr = 135,000; s20,w = 7.2 S) composed of two apparently identical catalytic and physical properties and their interconvertible nature suggest that they do not represent true isozymes. The enzyme is specific for dihydroxyacetone phosphate and does not form methylglyoxal from glyceraldehyde 3-phophate, glyceraldehyde, or dihydroxyacetone. Nonphosphorylated analogs are neither substrates nor competive inhibitors, but a variety of phosphorylated analogs are competitive with respect to dihydroxyacetone phosphate. The enzyme is inhibited by inorganic orthophosphate in a complex manner which is overcome by dihydroxyacetone phosphate in a signoidal manner  相似文献   

6.
A detailed understanding of the catalytic mechanism of enzymes is an important step toward improving their activity for use in biotechnology. In this paper, crystal soaking experiments and X-ray crystallography were used to analyse the mechanism of the Agrobacterium radiobacter phosphotriesterase, OpdA, an enzyme capable of detoxifying a broad range of organophosphate pesticides. The structures of OpdA complexed with ethylene glycol and the product of dimethoate hydrolysis, dimethyl thiophosphate, provide new details of the catalytic mechanism. These structures suggest that the attacking nucleophile is a terminally bound hydroxide, consistent with the catalytic mechanism of other binuclear metallophosphoesterases. In addition, a crystal structure with the potential substrate trimethyl phosphate bound non-productively demonstrates the importance of the active site cavity in orienting the substrate into an approximation of the transition state.  相似文献   

7.
In the yeast Saccharomyces cerevisiae lipid particles harbor two acyltransferases, Gat1p and Slc1p, which catalyze subsequent steps of acylation required for the formation of phosphatidic acid. Both enzymes are also components of the endoplasmic reticulum, but this compartment contains additional acyltransferase(s) involved in the biosynthesis of phosphatidic acid (K. Athenstaedt and G. Daum, J. Bacteriol. 179:7611-7616, 1997). Using the gat1 mutant strain TTA1, we show here that Gat1p present in both subcellular fractions accepts glycerol-3-phosphate and dihydroxyacetone phosphate as a substrate. Similarly, the additional acyltransferase(s) present in the endoplasmic reticulum can acylate both precursors. In contrast, yeast mitochondria harbor an enzyme(s) that significantly prefers dihydroxyacetone phosphate as a substrate for acylation, suggesting that at least one additional independent acyltransferase is present in this organelle. Surprisingly, enzymatic activity of 1-acyldihydroxyacetone phosphate reductase, which is required for the conversion of 1-acyldihydroxyacetone phosphate to 1-acylglycerol-3-phosphate (lysophosphatidic acid), is detectable only in lipid particles and the endoplasmic reticulum and not in mitochondria. In vivo labeling of wild-type cells with [2-3H, U-14C]glycerol revealed that both glycerol-3-phosphate and dihydroxyacetone phosphate can be incorporated as a backbone of glycerolipids. In the gat1 mutant and the 1-acylglycerol-3-phosphate acyltransferase slc1 mutant, the dihydroxyacetone phosphate pathway of phosphatidic acid biosynthesis is slightly preferred as compared to the wild type. Thus, mutations of the major acyltransferases Gat1p and Slc1p lead to an increased contribution of mitochondrial acyltransferase(s) to glycerolipid synthesis due to their substrate preference for dihydroxyacetone phosphate.  相似文献   

8.
A detailed understanding of the catalytic mechanism of enzymes is an important step toward improving their activity for use in biotechnology. In this paper, crystal soaking experiments and X-ray crystallography were used to analyse the mechanism of the Agrobacterium radiobacter phosphotriesterase, OpdA, an enzyme capable of detoxifying a broad range of organophosphate pesticides. The structures of OpdA complexed with ethylene glycol and the product of dimethoate hydrolysis, dimethyl thiophosphate, provide new details of the catalytic mechanism. These structures suggest that the attacking nucleophile is a terminally bound hydroxide, consistent with the catalytic mechanism of other binuclear metallophosphoesterases. In addition, a crystal structure with the potential substrate trimethyl phosphate bound non-productively demonstrates the importance of the active site cavity in orienting the substrate into an approximation of the transition state.  相似文献   

9.
l-Glycerol 3-phosphate dehydrogenase has been isolated and partially purified from the endosperm of developing castor beans. The enzyme is entirely cytosolic and is not found in the plastid fraction. No activity was found in germinating castor beans. The pH optimum for the reduction of dihydroxyacetone phosphate is 8.1 and is 9.6 for the reverse reaction. The molecular weight determined by gel filtration chromatography is between 71,000 and 83,000. Both substrates show substrate inhibition at concentrations about 13 μm for NADH and 400 μm for dihydroxyacetone phosphate. Substrate interaction kinetics gave limiting Km values of 2.7 and 35.5 μm for NADH and dihydroxyacetone phosphate, respectively. Substrate interaction and product inhibition kinetics were consistent with an ordered sequential mechanism with NADH being the first substrate to bind and NAD+ being the last product to dissociate.  相似文献   

10.
The structure of a class II fructose-1,6-bisphosphate aldolase in complex with the substrate analogue and inhibitor phosphoglycolohydroxamate (PGH) has been determined using X-ray diffraction terms to a resolution of 2.0 A (1 A=0.1 nm). The crystals are trigonal, space group P3121 with a=b=78.24 A, c=289.69 A. The asymmetric unit is a homodimer of (alpha/beta)8 barrels and the model has refined to give R-work 19.2 %, R-free (based on 5 % of the data) 23.0 %. PGH resembles the ene-diolate transition state of the physiological substrate dihydroxyacetone phosphate. It is well ordered and bound in a deep polar cavity at the C-terminal end of the (alpha/beta)8 barrel, where it chelates the catalytic zinc ion using hydroxyl and enolate oxygen atoms. Trigonal bipyramidal coordination of the zinc ion is completed by three histidine residues. The complex network of hydrogen bonds at the catalytic centre is required to organise the position of key functional groups and metal ion ligands. A well-defined monovalent cation-binding site is observed following significant re-organisation of loop structures. This assists the formation of a phosphate-binding site on one side of the barrel that tethers PGH in the catalytic site. The positions of functional groups of substrate and putative interactions with key amino acid residues are identified. Knowledge of the complex structure complements the results of spectroscopic and site-directed mutagenesis studies, and contributes to our understanding of the mechanism and substrate specificity of this family of enzymes. A reaction mechanism distinct from that proposed for other class II aldolases is discussed. The results suggest that the class II aldolases should be sub-divided into two groups on the basis of both distinct folds and mechanism.  相似文献   

11.
Rates of phosphatidate synthesis from dihydroxyacetone phosphate via acyl dihydroxyacetone phosphate or glycerol phosphate are compared in homogenates of 13 tissues, most of which are deficient in glycerol phosphate dehydrogenase (EC 1.1.1.8). In all tissues examined, dihydroxyacetone phosphate entered phosphatidate more rapidly via acyl dihydroxyacetone phosphate than via glycerol phosphate. Tissues with a relatively low rate of phosphatidate synthesis via glycerol phosphate, showed no compensating increase in the rate of synthesis via acyl dihydroxyacetone phosphate. The rates at which tissue homogenates synthesize phosphatidate from dihydroxyacetone phosphate via glycerol phosphate increase as glycerol phosphate dehydrongenase increase. Both glycerol phosphate dehydrogenase and glycerol phosphate: acyl CoA acyltransferase (EC 2.3.1.15) are more active than dihydroxyacetone phosphate : acyl CoA acyltransferase (EC 2.3.1.42). Thus, all the tissue homogenates possessed an apparently greater capability to synthesize phosphatidate via glycerol phosphate than via acyl dihydroxyacetone phosphate, but did not express this potential. This result is discussed in relation to in vivo substrate limitations.  相似文献   

12.
Lorentzen E  Siebers B  Hensel R  Pohl E 《Biochemistry》2005,44(11):4222-4229
The glycolytic enzyme fructose-1,6-bisphosphate aldolase (FBPA) catalyzes the reversible cleavage of fructose 1,6-bisphosphate to glyceraldehyde 3-phosphate and dihydroxyacetone phosphate. Catalysis of Schiff base forming class I FBPA relies on a number of intermediates covalently bound to the catalytic lysine. Using active site mutants of FBPA I from Thermoproteus tenax, we have solved the crystal structures of the enzyme covalently bound to the carbinolamine of the substrate fructose 1,6-bisphosphate and noncovalently bound to the cyclic form of the substrate. The structures, determined at a resolution of 1.9 A and refined to crystallographic R factors of 0.148 and 0.149, respectively, represent the first view of any FBPA I in these two stages of the reaction pathway and allow detailed analysis of the roles of active site residues in catalysis. The active site geometry of the Tyr146Phe FBPA variant with the carbinolamine intermediate supports the notion that in the archaeal FBPA I Tyr146 is the proton donor catalyzing the conversion between the carbinolamine and Schiff base. Our structural analysis furthermore indicates that Glu187 is the proton donor in the eukaryotic FBPA I, whereas an aspartic acid, conserved in all FBPA I enzymes, is in a perfect position to be the general base facilitating carbon-carbon cleavage. The crystal structure of the Trp144Glu, Tyr146Phe double-mutant substrate complex represents the first example where the cyclic form of beta-fructose 1,6-bisphosphate is noncovalently bound to FBPA I. The structure thus allows for the first time the catalytic mechanism of ring opening to be unraveled.  相似文献   

13.
A gene having high sequence homology (45-49%) with the glycerol-1-phosphate dehydrogenase gene from Methanobacterium thermoautotrophicum was cloned from the aerobic hyperthermophilic archaeon Aeropyrum pernix K1 (JCM 9820). This gene expressed in Escherichia coli with the pET vector system consists of 1113 nucleotides with an ATG initiation codon and a TAG termination codon. The molecular mass of the purified enzyme was estimated to be 38 kDa by SDS/PAGE and 72.4 kDa by gel column chromatography, indicating presence as a dimer. The optimum reaction temperature of this enzyme was observed to be 94-96 degrees C at near neutral pH. This enzyme was subjected to two-substrate kinetic analysis. The enzyme showed substrate specificity for NAD(P)H-dependent dihydroxyacetone phosphate reduction and NAD(+)-dependent glycerol-1-phosphate (Gro1P) oxidation. NADP(+)-dependent Gro1P oxidation was not observed with this enzyme. For the production of Gro1P in A. pernix cells, NADPH is the preferred coenzyme rather than NADH. Gro1P acted as a noncompetitive inhibitor against dihydroxyacetone phosphate and NAD(P)H. However, NAD(P)(+) acted as a competitive inhibitor against NAD(P)H and as a noncompetitive inhibitor against dihydroxyacetone phosphate. This kinetic data indicates that the catalytic reaction by glycerol- 1-phosphate dehydrogenase from A. pernix follows a ordered bi-bi mechanism.  相似文献   

14.
In the presence of triose phosphate isomerase, the substrate dihydroxyacetone phosphate is reduced stereoselectively by NaBH(4). The reduction of enzyme-bound substrate is almost completely or completely stereoselective and occurs about one order of magnitude faster than that in free solution. This acceleration implies a polarization of the carbonyl group when dihydroxyacetone phosphate is bound.  相似文献   

15.
Mature boar spermatozoa oxidized glycerol to carbon dioxide in the absence of any detectable activity of glycerol kinase. With triosephosphate isomerase and glyceraldehyde-3-phosphate dehydrogenase inhibited by the presence of 3-chloro-1-hydroxypropanone (CHOP), dihydroxyacetone phosphate accumulated in incubates when glycerol-3-phosphate was the substrate, but not when it was glycerol. Both dihydroxyacetone and glyceraldehyde could be used as substrates; in the presence of CHOP, dihydroxyacetone phosphate and fructose-1,6-bisphosphate accumulated when dihydroxyacetone was the substrate, but not when it was glyceraldehyde. The metabolic pathways glycerol----glyceraldehyde----glyceraldehyde 3-phosphate and dihydroxyacetone----dihydroxyacetone phosphate have been shown to operate in these cells.  相似文献   

16.
17.
The equivalence of the four dihydroxyacetone phosphate binding sites of aldolase was abolished by lowering the temperature. At pH 6.2 and ?13 2C, four binding sites were detected by gel filtration; two sites with a Kdiss ?0.1 μm, and a second set of sites with a Kdiss = 4 μm. The alteration of the binding was accompanied by the alteration of the catalytic activity. The low-affinity sites were incapable of catalyzing the cleavage of the (3S) CH bond of dihydroxyacetone phosphate, and form only the ketimine phosphate intermediate. The high-affinity sites were still able to cleave the (3S) CH bond of dihydroxyacetone phosphate; however, the eneamine phosphate intermediate formed was almost fully converted into the eneamine-aldehyde … phosphate intermediate, which was the prevailing species at the equilibrium. The mechanism of the half-of-the sites reactivity of aldolase at low temperature has been explained and the nonequivalence of sites in promoting catalysis has been utilized to dissect and characterize the individual partial reactions of the enzyme. In the course of these studies it has been shown that the rate of hydration-dehydration of dihydroxyacetone phosphate at ?24 °C was too slow to measure.  相似文献   

18.
The crystal structures of l-fuculose-1-phosphate aldolase (FucA) with and without a ligated analogue of dihydroxyacetone phosphate (DHAP) and of a number of active center mutants have resulted in a model of the catalytic mechanism. This model has now been confirmed by structural analyses of further mutations at the zinc coordination sphere and at the phosphate site. In addition, these mutants have revealed new aspects of the catalysis: the hydroxyl group of Tyr113' (from a neighboring subunit), which sits just outside the zinc coordination sphere, steers DHAP towards a productive binding mode at the zinc ion; Glu73 contacts zinc in between the two ligand positions intended for the DHAP oxygen atoms and thus avoids blocking of these positions by a tetrahedrally coordinated hydroxy ion; the FucA polypeptide does not assume its minimum energy state but oscillates between two states of elevated energy as demonstrated by a mutant in a minimum energy state. The back and forth motion involves a mobile loop connecting the phosphate site with intersubunit motions and thus with the Brownian motion of the solvent. The phosphate group is bound strongly at a given distance to the zinc ion, which prevents the formation of too tight a DHAP:zinc complex. This observation explains our failure to find mutants that accept phosphate-free substitutes for DHAP. The FucA zinc coordination sphere is compared with that of carbonic anhydrase.  相似文献   

19.
We present molecular dynamics simulations on the active site region of dimeric triose phosphate isomerase (TIM) using the co-ordinates of native chicken muscle TIM as a starting point and performing simulations with no substrate, with dihydroxyacetone phosphate (DHAP), the natural substrate, and with dihydroxyacetone sulfate (DHAS), a substrate analog. Whereas most of the protein moves less than 1 A during the simulation, some residues in the active site loop move more than 8 A during the 10.5 picoseconds of dynamics for each of the simulations. Most interestingly, the nature of the loop motion depends on the substrate, with the largest motion found in the presence of DHAP, and only in the presence of DHAP does the loop move to "close off" the active site pocket. The final structure found for the DHAP-chicken TIM complex is qualitatively similar to that described by Alber et al. for DHAP-yeast TIM. Simulations on the monomeric protein gives insight into why the molecule is active only as a dimer.  相似文献   

20.
An important active-site residue in the glycolytic enzyme triosephosphate isomerase is His-95, which appears to act as an electrophilic component in catalyzing the enolization of the substrates. With the techniques of site-directed mutagenesis, His-95 has been replaced by Gln in the isomerase from Saccharomyces cerevisiae. The mutant isomerase has been expressed in Escherichia coli strain DF502 and purified to homogeneity. The specific catalytic activity of the mutant enzyme is less than that of wild type by a factor of nearly 400. The mutant enzyme can be resolved from the wild-type isomerase on nondenaturing isoelectric focusing gels, and an isomerase activity stain shows that the observed catalytic activity indeed derives from the mutant protein. The inhibition constants for arsenate and for glycerol phosphate with the mutant enzyme are similar to those with the wild-type isomerase, but the substrate analogues 2-phosphoglycolate and phosphoglycolohydroxamate bind 8- and 35-fold, respectively, more weakly to the mutant isomerase. The mutant enzyme shows the same stereospecificity of proton transfer as the wild type. Tritium exchange experiments similar to those used to define the free energy profile for the wild-type yeast isomerase, together with a new method of analysis involving 14C and 3H doubly labeled substrates, have been used to investigate the energetics of the mutant enzyme catalyzed reaction. When the enzymatic reaction is conducted in tritiated solvent, the mutant isomerase does not catalyze any appreciable exchange between protons of the remaining substrate and those of the solvent either in the forward reaction direction (using dihydroxyacetone phosphate as substrate) or in the reverse direction (using glyceraldehyde phosphate as substrate). However, the specific radioactivity of the product glyceraldehyde phosphate formed in the forward reaction is 31% that of the solvent, while that of the product dihydroxyacetone phosphate formed in the reverse reaction is 24% that of the solvent. The deuterium kinetic isotope effects observed with the mutant isomerase using [1(R)-2H]dihydroxyacetone phosphate and [2-2H]glyceraldehyde 3-phosphate are 2.15 +/- 0.04 and 2.4 +/- 0.1, respectively. These results lead to the conclusion that substitution of Gln for His-95 so impairs the ability of the enzyme to stabilize the reaction intermediate that there is a change in the pathways of proton transfer mediated by the mutant enzyme. The data allow us more closely to define the role of His-95 in the reaction catalyzed by the wild-type enzyme, while forcing us to be alert to subtle changes in mechanistic pathways when mutant enzymes are generated.  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