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.
Aspartate kinase and two homoserine dehydrogenases were partially purified from 4-day-old pea seedlings. A sensitive method for measuring aspartate kinase activity is described. Aspartate kinase activity was dependent upon ATP, Mg2+ or Mn2+, and aspartate. The aspartate kinase was inhibited in a sigmoidal manner by threonine and Ki for threonine was 0·57 mM. The enzyme could be desensitized to the inhibitor and threonine protected the enzyme against thermal inactivation. Aspartate kinase activity was enhanced by isoleucine, valine and alanine. Homoserine, methionine and lysine were without effect. The homoserine dehydrogenase activity which was associated with aspartate kinase during purification could be resolved into two peaks by gel filtration. The activity of both peaks was inhibited by aspartate and cysteine and one was inhibited by threonine.  相似文献   

2.
In eukaryotes, two isozymes (I and II) of methionine aminopeptidase (MetAP) catalyze the removal of the initiator methionine if the penultimate residue has a small radius of gyration (glycine, alanine, serine, threonine, proline, valine, and cysteine). Using site-directed mutagenesis, recombinant yeast MetAP I derivatives that are able to cleave N-terminal methionine from substrates that have larger penultimate residues have been expressed. A Met to Ala change at 329 (Met206 in Escherichia coli enzyme) produces an average catalytic efficiency 1.5-fold higher than the native enzyme on normal substrates and cleaves substrates containing penultimate asparagine, glutamine, isoleucine, leucine, methionine, and phenylalanine. Interestingly, the native enzyme also has significant activity with the asparagine peptide not previously identified as a substrate. Mutation of Gln356 (Gln233 in E. coli MetAP) to alanine results in a catalytic efficiency about one-third that of native with normal substrates but which can cleave methionine from substrates with penultimate histidine, asparagine, glutamine, leucine, methionine, phenylalanine, and tryptophan. Mutation of Ser195 to alanine had no effect on substrate specificity. None of the altered enzymes produced cleaved substrates with a fully charged residue (lysine, arginine, aspartic acid, or glutamic acid) or tyrosine in the penultimate position.  相似文献   

3.
J K Wright  M Takahashi 《Biochemistry》1977,16(8):1541-1548
The aspartokinase activity of the aspartokinase-homoserine dehydrogenase complex of Escherichia coli was affinity labeled with substrates ATP, aspartate, and feedback inhibitor threonine. Exchange-inert ternary adducts of Co(III)-aspartokinase and either ATP, aspartate or threonine were formed by oxidation of corresponding Co(II) ternary complexes with H2O2. The ternary enzyme-Co(III)-threonine adduct (I) had 3.8 threonine binding sites per tetramer, one-half that of the native enzyme. The binding of threonine to I was still cooperative as determined by equilibrium dialysis (nH = 2.2) or by studying inhibition of residual dehydrogenase activity (nH = 2.7). Threonine still protected the SH groups of I against 5,5'-dithiobis(2-nitrobenzoate) (DTNB) reaction but the number of SH groups reacting with thiol reagents (DTNB) was reduced by 1-2 per subunit in the absence of threonine. This suggests either that Co(III) is bound to the enzyme via sulfhydryl groups or that 1-2SH groups are buried or rendered inaccessible in I. The binding of threonine to sites not blocked by the affinity labeling produced changes in the circular dichroism of the complex comparable to changes produced by threonine binding to native enzyme and also protected against proteolytic digestion. The major conformational changes produced by threonine are thus ascribable to binding at this one class of regulatory sites. The interactions of kinase substrates with various aspartokinase-Co(III) complexes containing ATP, aspartate, or threonine and a threonine-insensitive homoserine dehydrogenase produced by mild proteolysis were studied. The inhibition of homoserine dehydrogenase by kinase substrates is not due to binding of these inhibitors at the kinase active site but was shown to be due to binding to sites within the dehydrogenase domain of the enzyme. L-alpha-Aminobutyrate, a presumed threonine analogue, also inhibits the dehydrogenase by binding at the same or similar sites in the dehydrogenase domain and not at threonine regulatory site.  相似文献   

4.
The regulatory domain of the bifunctional threonine-sensitive aspartate kinase homoserine dehydrogenase contains two homologous subdomains defined by a common loop-alpha helix-loop-beta strand-loop-beta strand motif. This motif is homologous with that found in the two subdomains of the biosynthetic threonine-deaminase regulatory domain. Comparisons of the primary and secondary structures of the two enzymes allowed us to predict the location and identity of the amino acid residues potentially involved in two threonine-binding sites of Arabidopsis thaliana aspartate kinase-homoserine dehydrogenase. These amino acids were then mutated and activity measurements were carried out to test this hypothesis. Steady-state kinetic experiments on the wild-type and mutant enzymes demonstrated that each regulatory domain of the monomers of aspartate kinase-homoserine dehydrogenase possesses two nonequivalent threonine-binding sites constituted in part by Gln(443) and Gln(524). Our results also demonstrated that threonine interaction with Gln(443) leads to inhibition of aspartate kinase activity and facilitates the binding of a second threonine on Gln(524). Interaction of this second threonine with Gln(524) leads to inhibition of homoserine dehydrogenase activity.  相似文献   

5.
James CL  Viola RE 《Biochemistry》2002,41(11):3720-3725
The bifunctional enzyme aspartokinase-homoserine dehydrogenase I from Escherichia coli catalyzes non-consecutive reactions in the aspartate pathway of amino acid biosynthesis. Both catalytic activities are subject to allosteric regulation by the end product amino acid L-threonine. To examine the kinetics and regulation of the enzymes in this pathway, each of these catalytic domains were separately expressed and purified. The separated catalytic domains remain active, with each of their catalytic activities enhanced in comparison to the native enzyme. The allosteric regulation of the kinase activity is lost, and regulation of the dehydrogenase activity is dramatically decreased in these separate domains. To create a new bifunctional enzyme that can catalyze consecutive metabolic reactions, the aspartokinase I domain was fused to the enzyme that catalyzes the intervening reaction in the pathway, aspartate semialdehyde dehydrogenase. A hybrid bifunctional enzyme was also created between the native monofunctional aspartokinase III, an allosteric enzyme regulated by lysine, and the catalytic domain of homoserine dehydrogenase I with its regulatory interface domain still attached. In this hybrid the kinase activity remains sensitive to lysine, while the dehydrogenase activity is now regulated by both threonine and lysine. The dehydrogenase domain is less thermally stable than the kinase domain and becomes further destabilized upon removal of the regulatory domain. The more stable aspartokinase III is further stabilized against thermal denaturation in the hybrid bifunctional enzyme and was found to retain some catalytic activity even at temperatures approaching 100 degrees C.  相似文献   

6.
Metabolism of aspartate in Mycobacterium smegmatis   总被引:2,自引:0,他引:2  
Mycobacterium smegmatis grows best on L-asparagine as a sole nitrogen source; this was confirmed. [14C]Aspartate was taken up rapidly (46 nmol.mg dry cells-1.h-1 from 1 mM L-asparagine) and metabolised to CO2 as well as to amino acids synthesised through the aspartate pathway. Proportionately more radioactivity appeared in the amino acids in bacteria grown in medium containing low nitrogen. Activities of aspartokinase and homoserine dehydrogenase, the initial enzymes of the aspartate pathway, were carried by separate proteins. Aspartokinase was purified as three isoenzymes and represented up to 8% of the soluble protein of M. smegmatis. All three isoenzymes contained molecular mass subunits of 50 kDa and 11 kDa which showed no activity individually; full enzyme activity was recovered on pooling the subunits. Km values for aspartate were: aspartokinases I and III, 2.4 mM; aspartokinase II, 6.4 mM. Aspartokinase I was inhibited by threonine and homoserine and aspartokinase III by lysine, but aspartokinase II was not inhibited by any amino acids. Aspartokinase activity was repressed by methionine and lysine with a small residue of activity attributable to unrepressed aspartokinase I. Homoserine dehydrogenase activity was 96% inhibited by 2 mM threonine; isoleucine, cysteine and valine had lesser effects and in combination gave additive inhibition. Homoserine dehydrogenase was repressed by threonine and leucine. Only amino acids synthesised through the aspartate pathway were tested for inhibition and repression. Of these, only one, meso-diaminopimilate, had no discernable effect on either enzyme activity.  相似文献   

7.
We screened a gene trap library of Arabidopsis thaliana and isolated a line in which a gene encoding a homologue of monofunctional aspartate kinase was trapped by the reporter gene. Aspartate kinase (AK) is a key enzyme in the biosynthsis of aspartate family amino acids such as lysine, threonine, isoleucine, and methionine. In plants, two types of AK are known: one is AK which is sensitive to feedback inhibition by threonine and carries both AK and homoserine dehydrogenase (HSD) activities. The other one is monofunctional, sensitive to lysine and synergistically S-adenosylmethionine, and has only AK activity. We concluded that the trapped gene encoded a monofunctional aspartate kinase and designated as AK-lys3, because it lacked the HSD domain and had an amino acid sequence highly similar to those of the monofunctional aspartate kinases ofA. thaliana. AK-lys3 was highly expressed in xylem of leaves and hypocotyls and stele of roots. Significant expression of this gene was also observed in trichomes after bolting. Slight expression of AK-lys3 was detected in vascular bundles and mesophyll cells of cauline leaves, inflorescence stems, sepals, petals, and stigmas. These results indicated that this aspartate kinase gene was not expressed uniformly but in a spatially specific manner.  相似文献   

8.
9.
Baker P  Carere J  Seah SY 《Biochemistry》2012,51(22):4558-4567
BphJ, a nonphosphorylating acylating aldehyde dehydrogenase, catalyzes the conversion of aldehydes to form acyl-coenzyme A in the presence of NAD(+) and coenzyme A (CoA). The enzyme is structurally related to the nonacylating aldehyde dehydrogenases, aspartate-β-semialdehyde dehydrogenase and phosphorylating glyceraldehyde-3-phosphate dehydrogenase. Cys-131 was identified as the catalytic thiol in BphJ, and pH profiles together with site-specific mutagenesis data demonstrated that the catalytic thiol is not activated by an aspartate residue, as previously proposed. In contrast to the wild-type enzyme that had similar specificities for two- or three-carbon aldehydes, an I195A variant was observed to have a 20-fold higher catalytic efficiency for butyraldehyde and pentaldehyde compared to the catalytic efficiency of the wild type toward its natural substrate, acetaldehyde. BphJ forms a heterotetrameric complex with the class II aldolase BphI that channels aldehydes produced in the aldol cleavage reaction to the dehydrogenase via a molecular tunnel. Replacement of Ile-171 and Ile-195 with bulkier amino acid residues resulted in no more than a 35% reduction in acetaldehyde channeling efficiency, showing that these residues are not critical in gating the exit of the channel. Likewise, the replacement of Asn-170 in BphJ with alanine and aspartate did not substantially alter aldehyde channeling efficiencies. Levels of activation of BphI by BphJ N170A, N170D, and I171A were reduced by ≥3-fold in the presence of NADH and ≥4.5-fold when BphJ was undergoing turnover, indicating that allosteric activation of the aldolase has been compromised in these variants. The results demonstrate that the dehydrogenase coordinates the catalytic activity of BphI through allostery rather than through aldehyde channeling.  相似文献   

10.
Regulation of enzymes of lysine biosynthesis in Corynebacterium glutamicum   总被引:9,自引:0,他引:9  
The regulation of the six enzymes responsible for the conversion of aspartate to lysine, together with homoserine dehydrogenase, was studied in Corynebacterium glutamicum. In addition to aspartate kinase activity, the synthesis of diaminopimelate decarboxylase was also found to be regulated. The specific activity of this enzyme was reduced to one-third in extracts of cells grown in the presence of lysine. Aspartate-semialdehyde dehydrogenase, dihydrodipicolinate synthase, dihydrodipicolinate reductase, and diaminopimelate dehydrogenase were neither influenced in their specific activity, nor inhibited, by any of the aspartate family of amino acids. Homoserine dehydrogenase was repressed by methionine (to 15% of its original activity) and inhibited by threonine (4% remaining activity). Inclusion of leucine in the growth medium resulted in a twofold increase of homoserine dehydrogenase specific activity. The flow of aspartate semialdehyde to either lysine or homoserine was influenced by the activity of homoserine dehydrogenase or dihydrodipicolinate synthase. Thus, the twofold increase in homoserine dehydrogenase activity resulted in a decrease in lysine formation accompanied by the formation of isoleucine. In contrast, repression of homoserine dehydrogenase resulted in increased lysine formation. A similar increase of the flow of aspartate semialdehyde to lysine was found in strains with increased dihydrodipicolinate synthase activity, constructed by introducing the dapA gene of Escherichia coli (coding for the synthase) into C. glutamicum.  相似文献   

11.
1. In order to assess whether the potential ability of heart ventricular muscle and liver to metabolise substrates such as alanine, aspartate and lactate varies as the sheep matures and its nutrition changes, the activities of the following enzymes were determined in tissues of lambs obtained at varying intervals between 50 days after conception to 16 weeks after birth and in livers from adult pregnant ewes: lactate dehydrogenase (EC 1.1.1.27), alanine aminotransferase (EC 2.6.1.2), pyruvate kinase (EC 2.7.1.40), pyruvate carboxylase (EC 6.4.1.1), phosphoenolpyruvate carboxykinase (GTP)(EC 4.1.1.32), malate dehydrogenase (EC 1.1.1.37), aspartate aminotransferase (EC 2.6.1.1) and citrate (si)-synthase (EC 4.1.3.7). 2. In the heart a most marked increase in alanine aminotransferase activity was found throughout development. During this period the activities of citrate (si)-synthase, lactate dehydrogenase and pyruvate carboxylase also increased. There were no substantial changes in the activities of aspartate aminotransferase, malate dehydrogenase or pyruvate kinase. Pyruvate kinase activities were five times greater in the heart compared with those found in the liver. No significant activity of phosphoenolpyruvate carboxykinase (GTP) was detected in heart muscle. 3. In the liver the activities of both alanine aminotransferase and aspartate aminotransferase increased immediately following birth although the activity of alanine aminotransferase was lower in livers of pregnant ewes than in any of the lambs. As with alanine aminotransferase the highest activities of lactate dehydrogenase were found during the period of postnatal growth. No marked changes were observed in malate dehydrogenase or citrate (si)-synthase activities during development. A small decline in pyruvate kinase activity occurred whilst the activities of pyruvate carboxylase and phosphoenolpyruvate carboxykinase (GTP) tended to rise during development.  相似文献   

12.
J K Wright  J Feldman  M Takahashi 《Biochemistry》1976,15(17):3704-3710
The kinase active site of the aspartokinase-homoserine dehydrogenase enzyme complex of Excherichia coli has been affinity labeled both with substrates aspartate and adenosine triphosphate and feedback inhibitor threonine. Co(III) exchange-inert adducts of aspartokinase and inhibitor or substrates were produced in situ by oxidation of Co(II) with H2O2. Emzyme-Co(III)-adenosine 5'-triphosphate (ATP), enzyme-Co(III)-aspartate, and enzyme-Co(III)-threonine ternary adducts were produced in this manner. The formation of the enzyme-Co(III)-threonine adduct leads us to conclude that threonine inhibits the kinase activity of this enzyme complex by binding in the first coordination sphere of the catalytic metal ion cofactor, a conclusion which is consistent with evidence derived from previous nuclear magnetic resonance data obtained in this laboratory. The quaternary adducts formed by H2O2 oxidation in the presence of aspartokinase, Co(II), ATP, aspartate, and threonine comprised a mixture of both ezyme-Co(III)-ATP-aspartate and enzyme-Co(III)-ATP-threonine adducts. The formation of the quaternary aspartate-containing adduct was unexpected, since the presence of threonine was expected to prevent access of the aspartate to the active site; most significantly however, the the sum of the numbers of aspartate plus threonine molecules incorporated per active site is one. We believe that this shows direct steric overlap between the metal-adjacent binding sites for aspartate and threonine. Aspartate or threonine can not occupy the kinase active site simultaneously; this conclusion is consistent with the direct competitive inhibition of aspartate by threonine observed in steady-state kinetic studies.  相似文献   

13.
In our effort to elucidate the systems biology of the model organism, Escherichia coli, we have developed a mathematical model that simulates the allosteric regulation for threonine biosynthesis pathway starting from aspartate. To achieve this goal, we used kMech, a Cellerator language extension that describes enzyme mechanisms for the mathematical modeling of metabolic pathways. These mechanisms are converted by Cellerator into ordinary differential equations (ODEs) solvable by Mathematica. In this paper, we describe a more flexible model in Cellerator, which generalizes the Monod, Wyman, Changeux (MWC) model for enzyme allosteric regulation to allow for multiple substrate, activator and inhibitor binding sites. Furthermore, we have developed a model that describes the behavior of the bifunctional allosteric enzyme aspartate kinase I-homoserine dehydrogenase I (AKI-HDHI). This model predicts the partition of enzyme activities in the steady state which paves the way for a more generalized prediction of the behavior of bifunctional enzymes.  相似文献   

14.
The activity of three enzymes, aspartokinase, homoserine dehydrogenase, and homoserine kinase, has been studied in the industrial strainSaccharomyces cerevisiae IFI256 and in the mutants derived from it that are able to overproduce methionine and/or threonine. Most of the mutants showed alteration of the kinetic properties of the enzymes aspartokinase, which was less inhibited by threonine and increased its affinity for aspartate, and homoserine dehydrogenase and homoserine kinase, which both lost affinity for homoserine. Furthermore, they showed in vitro specific activities for aspartokinase and homoserine kinase that were higher than those of the wild type, resulting in accumulation of aspartate, homoserine, threonine, and/or methionine/S-adenosyl-methionine (Ado-Met). Together with an increase in the specific activity of both aspartokinase and homoserine kinase, there was a considerable and parallel increase in methionine and threonine concentration in the mutants. Those which produced the maximal concentration of these amino acids underwent minimal aspartokinase inhibition by threonine. This supports previous data that identify aspartokinase as the main agent in the regulation of the biosynthetic pathway of these amino acids. The homoserine kinase in the mutants showed inhibition by methionine together with a lack or a reduction of the inhibition by threonine that the wild type undergoes, which finding suggests an important role for this enzyme in methionine and threonine regulation. Finally, homoserine dehydrogenase displayed very similar specific activity in the mutants and the wild type in spite of the changes observed in amino acid concentrations; this points to a minor role for this enzyme in amino acid regulation.  相似文献   

15.
The enzyme threonine deaminase (TD) is a key regulatory enzyme in the pathway for the biosynthesis of isoleucine. TD is inhibited by its end product, isoleucine, and this effect is countered by valine, the product of a competing biosynthetic pathway. Sequence and structure analyses have revealed that the protomers of many TDs have C-terminal regulatory domains, composed of two ACT-like subdomains, which bind isoleucine and valine, while others have regulatory domains of approximately half the length, composed of only a single ACT-like domain. The regulatory responses of TDs from both long and short sequence varieties appear to have many similarities, but there are significant differences. We describe here the allosteric properties of Bacillus subtilis TD ( bsTD), which belongs to the short variety of TD sequences. We also examine the effects of several mutations in the regulatory domain on the kinetics of the enzyme and its response to effectors. The behavior of bsTD can be analyzed and rationalized using a modified Monod-Wyman-Changeux model. This analysis suggests that isoleucine is a negative effector, and valine is a very weak positive effector, but that at high concentrations valine inhibits activity by competing with threonine for binding to the active site. The behavior of bsTD is contrasted with the allosteric behavior reported for TDs from Escherichia coli and Arabidopsis thaliana, TDs with two subdomains. We suggest a possible evolutionary pathway to the more complex regulatory effects of valine on the activity of TDs of the long sequence variety, e.g., E. coli TD.  相似文献   

16.
Lysine, threonine, methionine and isoleucine are synthesized through the aspartate metabolic pathway. The concentrations of soluble lysine and threonine in cereal seeds are very low. Coix lacryma-jobi (coix) is a maize-related grass and the enzymological aspects of the aspartate metabolic pathway are completely unknown. In order to obtain information on lysine metabolism in this plant species, two enzymes involved in the biosynthesis of these amino acids (aspartate kinase 〚AK, EC 2.7.2.4〛 and homoserine dehydrogenase 〚HSDH, EC 1.1.1.3〛) and two enzymes involved in lysine degradation (lysine 2-oxoglutarate reductase 〚LOR, EC 1.5.1.8〛 and saccharopine dehydrogenase 〚SDH, EC 1.5.1.9〛) were isolated and partially characterized in coix seeds. AK activity was inhibited by threonine and lysine separately, suggesting the presence of two isoenzymes, one sensitive to lysine and the other sensitive to threonine, with the latter corresponding to approximately 60% of the total AK activity. In contrast to previous results from other plant species, the threonine-sensitive AK eluted from an ion exchange chromatography column at higher KCl concentration than the lysine-sensitive form. The HSDH activity extracted from the seeds was partially inhibited by threonine, indicating the presence of threonine-sensitive and threonine-resistant isoenzymes. LOR and SDH activities were detected only in the endosperm tissue and co-purified on an anion exchange chromatography column, suggesting that the two activities may be linked on a single bifunctional polypeptide, as observed for other plant species. One single SDH activity band was observed on non-denaturing PAGE gels. The Km for saccharopine of SDH was determined as 0.143 mM and the Km for NAD as 0.531 mM. Although SDH activity was shown to be stable, LOR, AK and HSDH were extremely unstable, under all buffer systems tested.  相似文献   

17.
The aspartate‐derived amino acid pathway in plants is an intensively studied metabolic pathway, because of the biosynthesis of the four essential amino acids lysine, threonine, isoleucine and methionine. The pathway is mainly controlled by the key regulatory enzymes aspartate kinase (AK; EC 2.7.2.4), homoserine dehydrogenase (HSDH; EC 1.1.1.3) and 4‐hydroxy‐tetrahydrodipicolinate synthase (EC 4.3.3.7), formerly referred to as dihydrodipicolinate synthase (DHDPS). They are encoded by isoenzyme families and it is not known why such families are evolutionarily maintained. To gain more insight into the specific roles and regulation of the isoenzymes, we inhibited DHDPS in Arabidopsis thaliana with the chemical compound (N,N‐dimethylglycinatoboranyloxycarbonylmethyl)‐dimethylamine‐borane (DDAB) and compared the short‐term effects on the biochemical and biomolecular level to the long‐term adaptations in dhdps knockout mutants. We found that DHDPS2 plays a crucial role in controlling lysine biosynthesis, thereby stabilizing flux through the whole aspartate pathway. Moreover, DHDPS2 was also shown to influence the threonine level to a large extent. In addition, the lysine‐sensitive AKs, AKLYS1 and AKLYS3 control the short‐ and long‐term responses to perturbed lysine biosynthesis in Arabidopsis thaliana.  相似文献   

18.
Biochemical characterization of rat brain protein kinase C isozymes   总被引:18,自引:0,他引:18  
Biochemical characteristics of three rat brain protein kinase C isozymes, types I, II, and III, were compared with respect to their protein kinase and phorbol ester-binding activities. All three isozymes appeared to be alike in their phorbol ester-binding activities as evidenced by their similar Kd for phorbol 12,13-dibutyrate and requirements for Ca2+ and phospholipids. However, differences with respect to the effector-mediated stimulation of protein kinase activity were detectable among these isozymes. The type I enzyme could be stimulated by cardiolipin to a greater extent than those of the type II and III enzymes. In the presence of cardiolipin, the concentrations of dioleoylglycerol or phorbol 12,13-dibutyrate required for half-maximal activation (A1/2) of the type I enzyme were nearly an order of magnitude lower than those for the type II and III enzymes. In the presence of phosphatidylserine, differences in the A1/2 of dioleoylglycerol and phorbol 12,13-dibutyrate for the three isozymes of protein kinase C were less significant than those measured in the presence of cardiolipin. Nevertheless, the A1/2 of these two activators for the type I enzyme were lower than those for the type II and III enzymes. At high levels of phosphatidylserine (greater than 15 mol %), binding of phorbol 12,13-dibutyrate to the type I enzyme evoked a corresponding stimulation of the kinase activity, whereas binding of this phorbol ester to the type II and III enzymes produced a lesser degree of kinase stimulation. For all three isozymes, the concentrations of phosphatidylserine required for half-maximum [3H]phorbol 12,13-dibutyrate binding were almost an order of magnitude less than those for kinase stimulation. Consequently, neither isozyme exhibited a significant kinase activity at lower levels of phosphatidylserine (less than 5 mol %) and phorbol 12,13-dibutyrate (50 nM), a condition sufficient to promote near maximal phorbol ester binding. In addition to their different responses to the various activators, the three protein kinase C isozymes also have different Km values for protein substrates. The type I enzyme appeared to have lower Km values for histone IIIS, myelin basic protein, poly(lysine, serine) (3:1) polymer, and protamine than those for the type II and III enzymes. These results documented that the three protein kinase C isozymes were distinguishable in their biochemical properties. In particular, the type I enzyme, which is a brain-specific isozyme, is distinct from the type II and III enzymes, both have a widespread distribution among different tissues.  相似文献   

19.
Amino acid biosynthesis: new architectures in allosteric enzymes.   总被引:1,自引:0,他引:1  
This review focuses on the allosteric controls in the Aspartate-derived and the branched-chain amino acid biosynthetic pathways examined both from kinetic and structural points of view. The objective is to show the differences that exist among the plant and microbial worlds concerning the allosteric regulation of these pathways and to unveil the structural bases of this diversity. Indeed, crystallographic structures of enzymes from these pathways have been determined in bacteria, fungi and plants, providing a wonderful opportunity to obtain insight into the acquisition and modulation of allosteric controls in the course of evolution. This will be examined using two enzymes, threonine synthase and the ACT domain containing enzyme aspartate kinase. In a last part, as many enzymes in these pathways display regulatory domains containing the conserved ACT module, the organization of ACT domains in this kind of allosteric enzymes will be reviewed, providing explanations for the variety of allosteric effectors and type of controls observed.  相似文献   

20.
Seah SY  Britton KL  Rice DW  Asano Y  Engel PC 《Biochemistry》2002,41(38):11390-11397
Homology-based modeling of phenylalanine dehydrogenases (PheDHs) from various sources, using the structures of homologous enzymes Clostridium symbiosum glutamate dehydrogenase and Bacillus sphaericus leucine dehydrogenase as a guide, revealed that an asparagine residue at position 145 of B. sphaericus PheDH was replaced by valine or alanine in PheDHs from other sources. This difference was proposed to be the basis for the poor discrimination by the B. sphaericus enzyme between the substrates L-phenylalanine and L-tyrosine. Residue 145 of this enzyme was altered, by site-specific mutagenesis, to hydrophobic residues alanine, valine, leucine, and isoleucine, respectively. The resultant mutants showed a high discrimination, above 50-fold, between L-phenylalanine and L-tyrosine. This higher specificity toward L-phenylalanine was due to K(m) values for L-phenylalanine lowered more than 20-fold compared to the values for L-tyrosine. The greater specificity for L-phenylalanine in the wild-type Bacillus badius enzyme, which has a valine residue in the corresponding position, was also found to be largely due to a lower K(m) for this substrate. Activities were also measured with a range of six amino acids with aliphatic, nonpolar side chains, and with the corresponding oxoacids, and in all cases the specificity constants for these substrates were increased in the mutant enzymes. As with phenylalanine, these increases are mainly attributable to large decreases in K(m) values.  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