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 219 毫秒
1.
The JmjC-domain-containing 2-oxoglutarate-dependent oxygenases catalyze protein hydroxylation and Nε-methyllysine demethylation via hydroxylation. A subgroup of this family, the JmjC lysine demethylases (JmjC KDMs) are involved in histone modifications at multiple sites. There are conflicting reports as to the substrate selectivity of some JmjC oxygenases with respect to KDM activities. In this study, a panel of modified histone H3 peptides was tested for demethylation against 15 human JmjC-domain-containing proteins. The results largely confirmed known Nε-methyllysine substrates. However, the purified KDM4 catalytic domains showed greater substrate promiscuity than previously reported (i.e., KDM4A was observed to catalyze demethylation at H3K27 as well as H3K9/K36). Crystallographic analyses revealed that the Nε-methyllysine of an H3K27me3 peptide binds similarly to Nε-methyllysines of H3K9me3/H3K36me3 with KDM4A. A subgroup of JmjC proteins known to catalyze hydroxylation did not display demethylation activity. Overall, the results reveal that the catalytic domains of the KDM4 enzymes may be less selective than previously identified. They also draw a distinction between the Nε-methyllysine demethylation and hydroxylation activities within the JmjC subfamily. These results will be of use to those working on functional studies of the JmjC enzymes.  相似文献   

2.
《Epigenetics》2013,8(12):1596-1603
The JmjC-domain-containing 2-oxoglutarate-dependent oxygenases catalyze protein hydroxylation and N?-methyllysine demethylation via hydroxylation. A subgroup of this family, the JmjC lysine demethylases (JmjC KDMs) are involved in histone modifications at multiple sites. There are conflicting reports as to the substrate selectivity of some JmjC oxygenases with respect to KDM activities. In this study, a panel of modified histone H3 peptides was tested for demethylation against 15 human JmjC-domain-containing proteins. The results largely confirmed known N?-methyllysine substrates. However, the purified KDM4 catalytic domains showed greater substrate promiscuity than previously reported (i.e., KDM4A was observed to catalyze demethylation at H3K27 as well as H3K9/K36). Crystallographic analyses revealed that the N?-methyllysine of an H3K27me3 peptide binds similarly to N?-methyllysines of H3K9me3/H3K36me3 with KDM4A. A subgroup of JmjC proteins known to catalyze hydroxylation did not display demethylation activity. Overall, the results reveal that the catalytic domains of the KDM4 enzymes may be less selective than previously identified. They also draw a distinction between the N?-methyllysine demethylation and hydroxylation activities within the JmjC subfamily. These results will be of use to those working on functional studies of the JmjC enzymes.  相似文献   

3.
Pyrrolysyl-tRNA synthetase (PylRS), an aminoacyl-tRNA synthetase (aaRS) recently found in some methanogenic archaea and bacteria, recognizes an unusually large lysine derivative, l-pyrrolysine, as the substrate, and attaches it to the cognate tRNA (tRNAPyl). The PylRS-tRNAPyl pair interacts with none of the endogenous aaRS-tRNA pairs in Escherichia coli, and thus can be used as a novel aaRS-tRNA pair for genetic code expansion. The crystal structures of the Methanosarcina mazei PylRS revealed that it has a unique, large pocket for amino acid binding, and the wild type M. mazei PylRS recognizes the natural lysine derivative as well as many lysine analogs, including N?-(tert-butoxycarbonyl)-l-lysine (Boc-lysine), with diverse side chain sizes and structures. Moreover, the PylRS only loosely recognizes the α-amino group of the substrate, whereas most aaRSs, including the structurally and genetically related phenylalanyl-tRNA synthetase (PheRS), strictly recognize the main chain groups of the substrate. We report here that wild type PylRS can recognize substrates with a variety of main-chain α-groups: α-hydroxyacid, non-α-amino-carboxylic acid, Nα-methyl-amino acid, and d-amino acid, each with the same side chain as that of Boc-lysine. In contrast, PheRS recognizes none of these amino acid analogs. By expressing the wild type PylRS and its cognate tRNAPyl in E. coli in the presence of the α-hydroxyacid analog of Boc-lysine (Boc-LysOH), the amber codon (UAG) was recoded successfully as Boc-LysOH, and thus an ester bond was site-specifically incorporated into a protein molecule. This PylRS-tRNAPyl pair is expected to expand the backbone diversity of protein molecules produced by both in vivo and in vitro ribosomal translation.  相似文献   

4.
The N-terminal α-amino groups of β1-bungarotoxin (β1-Bgt) fromBungarus multicinctus venom were modified with trinitrobenzene sulfonic acid and the modified derivative was separated by high performance liquid chromatography. The trinitrophenylated (TNP) derivative contained two TNP groups at the α-amino groups of A chain and B chain and showed a marked decrease in enzymatic activity. Methionine residues at positions 6 and 8 of the A chain were oxidized with chloramine T or cleaved with cyanogen bromide to remove the N-terminal octapeptide. Oxidation of methionine residues and removal of the N-terminal octapeptide caused a precipitous decrease in enzymatic activity, whereas antigenicity remained unchanged. The presence of dihexanoyllecithin influenced the interaction between β1-Bgt and 8-antilinonaphthalene sulfonate (ANS) and revealed that β1-Bgt consists of two types of ANS-binding sites, one at the substrate binding site of the A chain and the other might be at the B chain. The modified derivatives still retained their affinity for Ca2+ and ANS, indicating that the N-terminal region is not involved in Ca2+ and substrate binding. A fluorescence study revealed that the α-amino group of the A chain was in the vicinity of substrate binding site and that the TNP α-amino groups were in proximity to Trp-19 of the A chain. In addition, the study showed that the N-terminal region is important for stabilizing the architectural environment of Trp-19. The results, together with the proposal that Trp-19 of the A chain is involved in substrate binding, suggest that the N-terminal region of the A chain plays a crucial role in maintaining a functional active site for β1-Bgt.  相似文献   

5.
N2-(2-Carboxyethyl)arginine synthase (CEAS), an unusual thiamin diphosphate (ThDP)-dependent enzyme, catalyses the committed step in the biosynthesis of the β-lactamase inhibitor clavulanic acid in Streptomyces clavuligerus. Crystal structures of tetrameric CEAS-ThDP in complex with the substrate analogues 5-guanidinovaleric acid (GVA) and tartrate, and a structure reflecting a possible enol(ate)-ThDP reaction intermediate are described. The structures suggest overlapping binding sites for the substrates d-glyceraldehyde-3-phosphate (d-G3P) and l-arginine, and are consistent with the proposed CEAS mechanism in which d-G3P binds at the active site and reacts to form an α,β-unsaturated intermediate, which subsequently undergoes (1,4)-Michael addition with the α-amino group of l-arginine. Additional solution studies are presented which probe the amino acid substrate tolerance of CEAS, providing further insight into the l-arginine binding site. These findings may facilitate the engineering of CEAS towards the synthesis of alternative β-amino acid products.  相似文献   

6.
Six analogs of tryptophanyl-adenylate, which is an important intermediate in the enzymatic synthesis of Trp-tRNATrp, have been prepared. Four compounds, tryptophanyl-8-bromoadenylate, tryptophanyl-2-chloroadenylate, tryptophanyl-7-deazaadenylate and tryptophanyl-(N6-methyl)adenylate, contain modifications in the nucleobase moiety, while tryptophanyl-2′ deoxyadenylate and tryptophanyl-3′-deoxyadenylate were modified in the carbohydrate part of the molecule. Three of these analogs (2-chloro, 7-deaza, 2′-deoxy analogs) as well as ATP analogs with the same modifications were substrates in the aminoacylation reaction; three analogs (8-bromo, N6-methyl, 3′-deoxy analogs) were inactive as well as the corresponding ATP analogs. In contrast, in the ATPPPi pyrophosphate exchange in the absence of tRNA all ATP analogs except 8-bromo-ATP were substrates. However, the presence of tRNA reduced the number of ATP analogs being substrates to that number of substrates observed in the aminoacylation. Therefore, it can be concluded that the presence of tRNA is responsible for an increase of specificity. The diastereomers of adenosine 5′-O-(3-thiotriphosphate) (ATPαS), adenosine 5′-O-(2-thiotriphosphate) (ATPβS), and adenosine 5′-O-(3-thiotriphosphate) (ATPγS) were tested with various divalent metals as substrates in the pyrophosphate exchange reaction. The Sp diastereomer of ATPαS is a substrate with Mg2+, whereas the Rp diastereomer is inactive. Both diastereomers are inactive in the presence of Zn2+. Since Zn2+ binds preferentially to the sulfur atom, an explanation of these results is that the Mg2+ ion is not bound to the α-phosphate. Only the Sp isomer of the diastereomers of ATPβS acts as substrate in the presence of Mg2+. The stereospecificity becomes reversed in the presence of Zn2+. ATPγS acts as substrate with both Mg2+ and Zn2+. These results suggest that the Δ isomer of the β,γ-bidentate ATP-Mg2+ complex is the substrate for this enzyme. From these results a molecular model of the ATP-Mg2+ complex in the active site can be derived in which the nucleotide is attached to the enzyme by interactions in which the 3′-OH and 6-NH2 group, one oxygen atom of the α-phosphorus atom, and the coordinated magnesium cation are all involved.  相似文献   

7.
Carbon-13 spin-lattice relaxation times, T1, have been measured for aqueous solutions of L-aspartic acid, L-alanine, O-phospho-L-serine, and 2-mercapto-L-succinic acid in the presence of the paramagnetic metal ions, Cu2+ and Mn2+, and Mg2+ as a diamagnetic control, at ambient temperature and neutral pH. Nitrogen-15, oxygen-17 and proton relaxation times were also obtained for L-aspartic acid and phosphorus-31 relaxation times for O-phospho-L-serine under similar conditions. The structures of these complexes in solution were determined from the various metal ion-nuclei distances calculated from the paramagaetically-induced relaxation. These results indicate that the Cu2+ interaction with L-aspartic acid is through α-amino and β-carboxyl groups while Mn2+ coordinates most strongly through α-and β-carboxyl groups, with the possibility of a weak interaction through the amino group.An examination of the coordination of these divalent metal ions to an analog of L-aspartic acid in which the β-carboxyl group is replaced by a phosphate group (O-phospho-L-serine) indicated that Cu2+ coordination is now probably through the α-amino and phosphate groups, while this analog is a monodentate ligand for Mn2+ coordinating through the phosphate group. Removal of the β-carboxyl group (L-alanine) also results in Cu2+ coordination through the α-carboxyl and α-amino groups, and the same ligand interactions are observed with Mn2+. Replacement of the α-amino group of L-aspartic acid with an - SH group (2-mercapto-L-succinate) is sufficient to eliminate any specific coordination with either Cu2+ or Mn2+.  相似文献   

8.
The β-hydroxyacid dehydrogenases are a structurally conserved family of enzymes that catalyze the NAD+ or NADP+-dependent oxidation of specific β-hydroxyacid substrates like β-hydroxyisobutyrate. These enzymes share distinct domains of amino acid sequence homology, most of which now have assigned putative functions. 6-phosphogluconate dehydrogenase and β-hydroxyisobutyrate dehydrogenase, the most well-characterized members, both appear to be readily inactivated by chemical modifiers of lysine residues, such as 2,4,6-trinitrobenzene sulfonate (TNBS). Peptide mapping by ESI-LCMS showed that inactivation of β-hydroxyisobutyrate dehydrogenase with TNBS occurs with the labeling of a single lysine residue, K248. This lysine residue is completely conserved in all family members and may have structural importance relating to cofactor binding. The structural framework of the β-hydroxyacid dehydrogenase family is shared by many bacterial homologues. One such homologue from E. coli has been cloned and expressed as recombinant protein. This protein was found to have enzymatic activity characteristic of tartronate semialdehyde reductase, an enzyme required for bacterial biosynthesis of d-glycerate. A homologue from H. influenzae was also cloned and expressed as recombinant protein. This protein was active in the oxidation of d-glycerate, but showed approximately ten-fold higher activity with four carbon substrates like β-d-hydroxybutyrate and d-threonine. This enzyme might function in H. influenzae, and other species, in the utilization of polyhydroxybutyrates, an energy storage form specific to bacteria. Cloning and characterization of these bacterial β-hydroxyacid dehydrogenases extends our knowledge of this enzyme family.  相似文献   

9.
Cuticular phenoloxidase(s) from Sarcophaga bullata larvae oxidized a variety of o-diphenolic compounds. While catechol, 3,4-dihydroxybenzoic acid, dopa, dopamine, and norepinephrine were converted to their corresponding quinone derivatives, other catechols such as 3,4-dihydroxyphenylacetic acid, 3,4-dihydroxyphenethyl alcohol, 3,4-dihydroxyphenyl glycol, 3,4-dihy-droxymandelic acid, and N-acetyldopamine were oxidized to their side-chain oxygenated products. In addition, the enzyme-catalyzed oxidation of the latter group of compounds accompanied the formation of colorless catecholcuticle adducts consistent with the operation of β-sclerotization. Radioactive trapping experiments failed to support the participation of 1,2-dehydro-N-acetyldopamine as a freely formed intermediate during phenoloxidase-mediated oxidation of N-acetyldopamine. When specifically tritiated substrates were provided, cuticular enzyme selectively removed tritium from [7-3H]N-acetyldopamine and not from either [8-3H] or [ring-3H]N-acetyldopamine during the initial phase of oxidation. The above results are consistent with the generation and subsequent reactions of quinone methides as the initial products of enzyme-catalyzed N-acetyldopamine oxidation and confirm our hypothesis that quinone methides and not 1,2-dehydro-N-acetyldopamine are the reactive intermediate of β-sclerotization of sarcophagid cuticle. Quinone methide sclerotization resolves a number of conflicting observations made by previous workers in this field.  相似文献   

10.
Magnetic resonance and kinetic studies of the catalytic subunit of a Type II cAMP-dependent protein kinase from bovine heart have established the active complex to be an enzyme-ATP-metal bridge. The metal ion is β,γ coordinated with Δ chirality at the β-phosphorous atom. The binding of a second metal ion at the active site which bridges the enzyme to the three phosphoryl groups of ATP, partially inhibits the reaction. Binding of the metal-ATP substrate to the enzyme occurs in a diffusion-controlled reaction followed by a 40 ° change in the glycosidic torsional angle. This conformational change results from strong interaction of the nucleotide base with the enzyme. NMR studies of four ATP-utilizing enzymes show a correlation between such conformational changes and high nucleotide base specificity. Heptapeptide substrates and substrate analogs bind to the active site of the catalytic subunit at a rate significantly lower than collision frequency indicating conformational selection by the enzyme or a subsequent slow conformational change. NMR studies of the conformation of the enzyme-bound peptide substrates have ruled out α-helical and β-pleated sheet structures. The results of kinetic studies of peptide substrates in which the amino acid sequence was systematically varied were used to rule out the obligatory requirement for all possible β-turn conformations within the heptapeptide although an enzymatic preference for a β2–5 or β3–6 turn could not be excluded. Hence if protein kinase has an absolute requirement for a specific secondary structure, then this structure must be a coil. In the enzyme-substrate complex the distance along the reaction coordinate between the γ-P of ATP and the serine oxygen of the peptide substrate (5.3 ± 0.7 Å) allows room for a metaphosphate intermediate. This finding together with kinetic observations as well as the location of the inhibitory metal suggest a dissociative mechanism for protein kinase, although a mechanism with some associative character remains possible. Regulation of protein kinase is accomplished by competition between the regulatory subunit and peptide or protein substrates at the active site of the catalytic subunit. Thus, the regulatory subunit is found by NMR to block the binding of the peptide substrate to the active site of protein kinase but allows the binding of the nucleotide substrate and divalent cations. The dissociation constant of the regulatory subunit from the active site (10?10m) is increased ~10-fold by phosphorylation and ~104-fold by the binding of cAMP, to a value (10?5m) which exceeds the intracellular concentration of the R2C2 holoenzyme complex (10?6m). The resulting dissociation of the holoenzyme releases the catalytic subunit, permitting the active site binding of peptide or protein substrates.  相似文献   

11.
In the absence of any other oxidizable substrate, the perfused rat heart oxidizes [1-14C]leucine to 14CO2 at a rapid rate and releases only small amounts of α-[1-14C]ketoisocaproate into the perfusion medium. The branched-chain α-keto acid dehydrogenase complex, assayed in extracts of mitochondria prepared from such perfused hearts, is very active. Under such perfusion conditions, dichloroacetate has almost no effect on [1-14C]leucine oxidation, α-[1-14C]ketoisocaproate release, or branched-chain α-keto acid dehydrogenase activity. Perfusion of the heart with some other oxidizable substrate, e.g., glucose, pyruvate, ketone bodies, or palmitate, results in an inhibition of [1-14C]leucine oxidation to 14CO2 and the release of large amounts of α-[1-14C]ketoisocaproate into the perfusion medium. The branched-chain α-keto acid dehydrogenase complex, assayed in extracts of mitochondria prepared from such hearts, is almost completely inactivated. The enzyme can be reactivated, however, by incubating the mitochondria at 30 °C without an oxidizable substrate. With hearts perfused with glucose or ketone bodies, dichloroacetate greatly increases [1-14C]leucine oxidation, decreases α-[1-14C]ketoisocaproate release into the perfusion medium, and activates the branched-chain α-keto acid dehydrogenase complex. Pyruvate may block dichloroacetate uptake because dichloroacetate neither stimulates [1-14C]leucine oxidation nor activates the branched-chain α-keto acid dehydrogenase complex of pyruvate-perfused hearts. It is suggested that leucine oxidation by heart is regulated by the activity of the branched-chain α-keto acid dehydrogenase complex which is subject to interconversion between active and inactive forms. Oxidizable substrates establish conditions which inactivate the enzyme. Dichloroacetate, known to activate the pyruvate dehydrogenase complex by inhibition of pyruvate dehydrogenase kinase, causes activation of the branched-chain α-keto acid dehydrogenase complex, suggesting the existence of a kinase for this complex.  相似文献   

12.
Lysosomal carboxypeptidase B has been purified from rabbit lung acetone powder by acid precipitation and ammonium sulfate fractionation followed by further purification on Sephadex G-100, DEAE-Sephadex, Organomercurial-Sepharose, preparative isoelectric focusing, Sephadex G-75, and carboxymethyl-Sephadex. This procedure resulted in a homogeneous preparation as determined by polyacrylamide gel electrophoresis at pH 4.5, 8.3 and with sodium dodecyl sulfate. This enzyme has a molecular weight of 52,000, is composed of two subunits of approximately equivalent molecular weight, and is a glycoprotein with a carbohydrate content estimated to be 10% by weight. The amino acid composition is also reported. The enzyme is active on two synthetic substrates, α-N-benyoyl-l-arginineamide and hippuryl-l-arginine. With these two substrates, respectively, lysosomal carboxypeptidase B has pH optima of 5.7 and 5.0, temperature optima of 40 and 50 °C, and Km values of 10 and 16 mm. With each substrate, the enzyme requires the presence of a reducing agent for maximal activity and is inhibited to the same extent with several inhibitors. The most potent inhibitors were leupeptin and antipain at low concentrations (1 μm). Iodoacetate and Ac-(Ala)3-Ala-chloro-methyl ketone also inhibited at higher concentrations (10 μm). However, compounds such as leucyl-chloromethyl ketone, bestatin, pepstatin, phenylmethylsulfonyl fluoride, soybean trypsin inhibitor, and α-1-antitrypsin did not inhibit. When tested with peptides as substrates, this proteinase exhibited strong carboxypeptidase activity on the tetrapeptide, ThrProArgLys, and on angiotensin I, AspArgValTyrIle HisProPheHisLeu, liberating Lys, and Leu, respectively. Substance P (containing 11 amino acids plus a C-terminal amide group) was virtually inactive as a substrate for this enzyme. However, with oxidized insulin B chain as substrate, lysosomal carboxypeptidase B exhibited significant carboxypeptidase and endopeptidase activities.  相似文献   

13.
Specificity for the glucose-6-P inhibition site of hexokinase   总被引:4,自引:0,他引:4  
Inhibition of the three animal hexokinase isozymes by the following glucose-6-P analogs has been determined: α-glucose-1,6-P2, α- and β-methyl glucose-6-P, α- and β-glucose-6-P, 2-Cl- and 4F-glucose-6-P, 5-deoxyglucose-6-P, glucose-6-sulfate, and δ-gluconolactone-6-P. Although both anomers of glucose-6-P were about equally active inhibitors, the β-methyl derivative was inactive. Generally the α-methyl and α-PO3? derivatives were good inhibitors though weaker than glucose-6-P except in the case of hexokinase II for which α-glucose-1,6-P2 was an excellent inhibitor.  相似文献   

14.
α-Amylases have been found to convert starch and glycogen, in part, to products other than hemiacetal-bearing entities (maltose, maltodextrins, etc.)—hitherto, the only products obtained from natural α-glucans by α-amylolysis. Glycosides of maltosaccharides were synthesized by purified α-amylases acting on starch or bacterial glycogen in the presence of p-nitrophenyl α- or β-d-glucoside. From a digest with crystallized B. subtilis var. amyloliquefaciens α-amylase, containing 4 mg/ml of [14C]glycogen and 40 mmp-NP β-d-glucoside, three pairs of correspondingly labeled glycosides and sugars were recovered: p-NP α-d-[14C]glucopyranosyl (1 → 4) β-d-glucopyranoside, and [14C]glucose; p-NP α-[14C]maltosyl (1 → 4) β-d-glucopyranoside, and [14C]maltose; p-NP α-[14C]maltotriosyl (1 → 4) β-d-glucopyranoside, and [14C]maltotriose. The three glycosides accounted for 11.4% of the [14C]glycogen donor substrate; the three comparable sugars, for 30.4%; higher maltodextrins, for 58.2%. Calculations based on the molar yields of all reaction products show that [14C]glycosyl moieties were transferred from donor to p-NP β-d-glucoside with a frequency of 0.234 relative to all transfers to water. This is a very high value considering the minute molar ratio (0.0007) of β-d-glucoside-to-water concentration. Less striking but similar findings were obtained with cryst. hog pancreatic and Aspergillus oryzae α-amylases. The results extend earlier findings (Hehre et al., Advan. Chem. Ser. (1973) 117, 309) in showing that α-amylases have a substantial capacity to utilize the C4-carbinols of certain d-glucosyl compounds as acceptor sites.  相似文献   

15.
Two pure peroxidase isoenzymes B1 and D4 were isolated from the upper parts of 10-day-old wheat seedlings by means of gel and ion-exchange chromatography. Their MWs were 85000 and 24000 respectively. B1 was unstable and under various conditions it was converted to another isoenzyme, electrophoretically identical with D4. B1 contains about 40% of neutral sugars: 17.2% arabinose, 15.3% galactose, 5% glucose and traces of mannose. D4 is free of neutral sugars. None of the isoenzymes contained amino sugars. B1 oxidizes ferulic and p-coumaric acids. This oxidation has two pH optima of 4.4 and 5.4–5.6 and is inhibited by high concentrations of substrates, cyanide and azide. B1 oxidizes IAA in the presence of phenolic cofactor and Mn2+ ions. IAA oxidation has two pH optima of 4.5 and 5.6 and is inhibited by high substrate concentration, cyanide and azide, and by a number of indole derivatives. The main products of IAA oxidation are 3-methyleneoxindole and indole-3-methanol. o- and p- diphenols induce a lag period prior to IAA oxidation. Ferulic acid is oxidized during this lag period, probably to a dimer. B1 is able to produce H2O2 from oxygen. Mn2+ ions, a phenolic cofactor and an electron donor (IAA or NADH) are needed. B1 oxidizes α-keto-γ- methylmercaptobutyric acid to ethylene. D4 has a low peroxidatic activity and is inactive as an IAA oxidase. Thus B1 is probably an active cell wall-bound peroxidase isoenzyme, whereas D4 is its decomposition product.  相似文献   

16.
Tissue transglutaminase (TG2) catalyzes the Ca2+-dependent posttranslational modification of proteins via formation of isopeptide bonds between their glutamine and lysine residues. Although substrate specificity of TG2 has been studied repeatedly at the sequence level, no clear consensus sequences have been determined so far. With the use of the extensive structural information on TG2 substrate proteins listed in TRANSDAB Wiki database†, a slight preference of TG2 for glutamine and lysine residues situated in turns could be observed. When the spatial environment of the favored glutamine and lysine residues was analyzed with logistic regression, the presence of specific amino acid patterns was identified. By using the occurrence of the predictor amino acids as selection criteria, several polypeptides were predicted and later identified as novel in vitro substrates for TG2. By studying the sequence of TG2 substrate proteins lacking available crystal structure, the strong favorable influence on substrate selection of the presence of substrate glutamine and lysine residues in intrinsically disordered regions could also be revealed. The collected structural data have provided novel understanding of how this versatile enzyme selects its substrates in various cell compartments and tissues.  相似文献   

17.
Lysine monooxygenase, a pseudomonad flavoprotein, was almost inactive with an α-monoamino acid as substrate, but the addition of an alkylamine, a counterpart of the fragmented lysine molecule, caused marked oxygen consumption. The rate of oxygen consumption was examined and compared with various combinations of α-monoamino acids and alkylamines. Hydrogen peroxide was formed and the corresponding α-keto acid was produced from each amino acid. In most cases, the hydrogen peroxide formation was nearly equal to the oxygen consumption. In some other cases, however, the oxygen consumption exceeded the hydrogen peroxide formation, and the production of an acid amide in addition to the α-keto acid was qualitatively demonstrated. Thus, both an oxidative deamination and an oxygenative decarboxylation of the same substrate occurred concomitantly, although the ratio of the two types of reaction varied with different combinations of amino acids and alkylamines. Alkylamines at higher concentrations competitively inhibited the lysine oxygenation, whereas lower concentrations of alkylamines stimulated the lysine oxygenation by decreasing the sigmoidicity of the saturation curve for lysine. These results suggest a dual function of the alkylamine; i.e., the interaction with the active site as a fragment of substrate and the effect on the regulatory property of the enzyme.  相似文献   

18.
The X-ray crystal structure is presented for a nitrogenase MoFe protein where the alpha subunit residue at position 70 (α-70Val) has been substituted by the amino acid isoleucine (α-70Ile). Substitution of α-70Val by α-70Ile results in a MoFe protein that is hampered in its ability to reduce a range of substrates including acetylene and N2, yet retains normal proton reduction activity. The 2.3 Å structure of the α-70Ile MoFe protein is compared to the α-70Val wild-type MoFe protein, revealing that the δ methyl group of α-70Val is positioned over Fe6 within the active site FeMo-cofactor. This work provides strong crystallographic support for the previously proposed model that substrates bind and are reduced at a single 4Fe-4S face of the FeMo-cofactor and that when α-70Val is substituted by α-70Ile access of substrates to Fe6 of this face is effectively blocked. Furthermore the detailed examination of the structure provides the basis for understanding the ability to trap and characterize hydrides in the variant, contributing significantly to our understanding of substrate access and substrate reduction at the FeMo-cofactor active site of nitrogenase.  相似文献   

19.
Phosphoribosylpyrophosphate synthetase from Salmonella typhimurium contains nine lysine residues per subunit and can be inactivated by reagents specific for this amino acid. Pyridoxal-P reversibly inhibited the enzyme by about 70% by forming a Schiff base derivative with lysine. Reduction with NaBH4 made this inactivation irreversible. Kinetic experiments indicated that the failure to inactivate the enzyme completely in a single treatment with pyridoxal-P reflects a reversible equilibrium between inactive Schiff base and a noncovalent complex. Modification of one lysine residue per subunit correlated with apparently total loss of activity. The rate of inactivation of the enzyme was decreased fourfold by saturating concentrations of ATP and was decreased at least 20-fold by formation of a quaternary complex of the enzyme with Mg2+, α,β-methylene ATP, and ribose-5-P. Trinitrobenzenesulfonate also irreversibly inactivated the enzyme, but this reagent was less specific in that the loss of activity corresponded to the modification of four to five lysine residues. These results suggest that an essential lysine is near the active site of Phosphoribosylpyrophosphate synthetase.  相似文献   

20.
Angiosperms possess a retaining trans-α-xylosidase activity that catalyses the inter-molecular transfer of xylose residues between xyloglucan structures. To identify the linkage of the newly transferred α-xylose residue, we used [Xyl-3H]XXXG (xyloglucan heptasaccharide) as donor substrate and reductively-aminated xyloglucan oligosaccharides (XGO–NH2) as acceptor. Asparagus officinalis enzyme extracts generated cationic radioactive products ([3H]Xyl·XGO–NH2) that were Driselase-digestible to a neutral trisaccharide containing an α-[3H]xylose residue. After borohydride reduction, the trimer exhibited high molybdate-affinity, indicating xylobiosyl-(1→6)-glucitol rather than a di-xylosylated glucitol. Thus the trans-α-xylosidase had grafted an additional α-[3H]xylose residue onto the xylose of an isoprimeverose unit. The trisaccharide was rapidly acetolysed to an α-[3H]xylobiose, confirming the presence of an acetolysis-labile (1→6)-bond. The α-[3H]xylobiitol formed by reduction of this α-[3H]xylobiose had low molybdate-affinity, indicating a (1→2) or (1→4) linkage. In NaOH, the α-[3H]xylobiose underwent alkaline peeling at the moderate rate characteristic of a (1→4)-disaccharide. Finally, we synthesised eight non-radioactive xylobioses [α and β; (1↔1), (1→2), (1→3) and (1→4)] and found that the [3H]xylobiose co-chromatographed only with (1→4)-α-xylobiose. We conclude that Asparagus trans-α-xylosidase activity generates a novel xyloglucan building block, α-d-Xylp-(1→4)-α-d-Xylp-(1→6)-d-Glc (abbreviation: ‘V’). Modifying xyloglucan structures in this way may alter oligosaccharin activities, or change their suitability as acceptor substrates for xyloglucan endotransglucosylase (XET) activity.  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