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 921 毫秒
1.
Lipoxygenases (LOXs) are a key part of several signaling pathways that lead to inflammation and cancer. Yet, the mechanisms of substrate binding and allosteric regulation by the various LOX isoforms remain speculative. Here we report the 2.47-Å resolution crystal structure of the arachidonate 11R-LOX from Gersemia fruticosa, which sheds new light on the mechanism of LOX catalysis. Our crystallographic and mutational studies suggest that the aliphatic tail of the fatty acid is bound in a hydrophobic pocket with two potential entrances. We speculate that LOXs share a common T-shaped substrate channel architecture that gives rise to the varying positional specificities. A general allosteric mechanism is proposed for transmitting the activity-inducing effect of calcium binding from the membrane-targeting PLAT (polycystin-1/lipoxygenase/α-toxin) domain to the active site via a conserved π-cation bridge.  相似文献   

2.
α‐Dioxygenases (α‐DOX) are heme‐containing enzymes found predominantly in plants and fungi, where they generate oxylipins in response to pathogen attack. α‐DOX oxygenate a variety of 14–20 carbon fatty acids containing up to three unsaturated bonds through stereoselective removal of the pro‐R hydrogen from the α‐carbon by a tyrosyl radical generated via the oxidation of the heme moiety by hydrogen peroxide (H2O2). We determined the X‐ray crystal structures of wild type α‐DOX from Oryza sativa, the wild type enzyme in complex with H2O2, and the catalytically inactive Y379F mutant in complex with the fatty acid palmitic acid (PA). PA binds within the active site cleft of α‐DOX such that the carboxylate forms ionic interactions with His‐311 and Arg‐559. Thr‐316 aids in the positioning of carbon‐2 for hydrogen abstraction. Twenty‐five of the twenty eight contacts made between PA and residues lining the active site occur within the carboxylate and first eight carbons, indicating that interactions within this region of the substrate are responsible for governing selectivity. Comparison of the wild type and H2O2 structures provides insight into enzyme activation. The binding of H2O2 at the distal face of the heme displaces residues His‐157, Asp‐158, and Trp‐159 ~2.5 Å from their positions in the wild type structure. As a result, the Oδ2 atom of Asp‐158 interacts with the Ca atom in the calcium binding loop, the side chains of Trp‐159 and Trp‐213 reorient, and the guanidinium group of Arg‐559 is repositioned near Tyr‐379, poised to interact with the carboxylate group of the substrate.  相似文献   

3.
Lipoxygenases (LOX) play critical roles in mammalian biology in the generation of potent lipid mediators of the inflammatory response; consequently, they are targets for the development of isoform-specific inhibitors. The regio- and stereo-specificity of the oxygenation of polyunsaturated fatty acids by the enzymes is understood in terms of the chemistry, but structural observation of the enzyme-substrate interactions is lacking. Although several LOX crystal structures are available, heretofore the rapid oxygenation of bound substrate has precluded capture of the enzyme-substrate complex, leaving a gap between chemical and structural insights. In this report, we describe the 2.0 Å resolution structure of 8R-LOX in complex with arachidonic acid obtained under anaerobic conditions. Subtle rearrangements, primarily in the side chains of three amino acids, allow binding of arachidonic acid in a catalytically competent conformation. Accompanying experimental work supports a model in which both substrate tethering and cavity depth contribute to positioning the appropriate carbon at the catalytic machinery.  相似文献   

4.
Conversion of the primary bile acids cholic acid (CA) and chenodeoxycholic acid (CDCA) to the secondary bile acids deoxycholic acid (DCA) and lithocholic acid (LCA) is performed by a few species of intestinal bacteria in the genus Clostridium through a multistep biochemical pathway that removes a 7α‐hydroxyl group. The rate‐determining enzyme in this pathway is bile acid 7α‐dehydratase (baiE). In this study, crystal structures of apo‐BaiE and its putative product‐bound [3‐oxo‐Δ4,6‐lithocholyl‐Coenzyme A (CoA)] complex are reported. BaiE is a trimer with a twisted α + β barrel fold with similarity to the Nuclear Transport Factor 2 (NTF2) superfamily. Tyr30, Asp35, and His83 form a catalytic triad that is conserved across this family. Site‐directed mutagenesis of BaiE from Clostridium scindens VPI 12708 confirm that these residues are essential for catalysis and also the importance of other conserved residues, Tyr54 and Arg146, which are involved in substrate binding and affect catalytic turnover. Steady‐state kinetic studies reveal that the BaiE homologs are able to turn over 3‐oxo‐Δ4‐bile acid and CoA‐conjugated 3‐oxo‐Δ4‐bile acid substrates with comparable efficiency questioning the role of CoA‐conjugation in the bile acid metabolism pathway. Proteins 2016; 84:316–331. © 2016 Wiley Periodicals, Inc.  相似文献   

5.
Rhomboids are intramembrane proteases that use a catalytic dyad of serine and histidine for proteolysis. They are conserved in both prokaryotes and eukaryotes and regulate cellular processes as diverse as intercellular signalling, parasitic invasion of host cells, and mitochondrial morphology. Their widespread biological significance and consequent medical potential provides a strong incentive to understand the mechanism of these unusual enzymes for identification of specific inhibitors. In this study, we describe the structure of Escherichia coli rhomboid GlpG covalently bound to a mechanism‐based isocoumarin inhibitor. We identify the position of the oxyanion hole, and the S1‐ and S2′‐binding subsites of GlpG, which are the key determinants of substrate specificity. The inhibitor‐bound structure suggests that subtle structural change is sufficient for catalysis, as opposed to large changes proposed from previous structures of unliganded GlpG. Using bound inhibitor as a template, we present a model for substrate binding at the active site and biochemically test its validity. This study provides a foundation for a structural explanation of rhomboid specificity and mechanism, and for inhibitor design.  相似文献   

6.
7.
The ubiquitously distributed peroxiredoxins (Prxs) have been shown to have diverse functions in cellular defense‐signaling pathways. They have been largely classified into three Prx classes, 2‐Cys Prx, atypical 2‐Cys Prx and 1‐Cys Prx, which can be distinguished by how many Cys residues they possess and by their catalytic mechanisms. Proteins belonging to the typical 2‐Cys Prx group containing the N‐terminal peroxidatic Cys residue undergo a cycle of peroxide‐dependent oxidation to sulfenic acid and thiol‐dependent reduction during H2O2 catalysis. However, in the presence of high concentrations of H2O2 and catalytic components, including thioredoxin (Trx), Trx reductase and NADPH, the sulfenic acid can be hyperoxidized to cysteine sulfinic acid. The overoxidized 2‐Cys Prxs are slowly reduced by the action of the adenosine 5′‐triphosphate‐dependent enzyme, sulfiredoxin. Upon exposure of cells to strong oxidative or heat‐shock stress conditions, 2‐Cys Prxs change their protein structures from low‐molecular weight to high‐molecular weight complexes, which trigger their functional switching from peroxidases to molecular chaperones. The C‐terminal region of 2‐Cys Prx also plays an essential role in this structural conversion. Thus, proteins with truncated C‐termini are resistant to overoxidation and cannot regulate their structures or functions. These reactions are primarily guided by the active site peroxidatic Cys residue, which serves as an ‘H2O2‐sensor’ in cells. The reversible structural and functional switching of 2‐Cys Prxs provides cells with a means to adapt to external stresses by presumably activating intracellular defense‐signaling systems. In particular, plant 2‐Cys Prxs localized in chloroplasts have dynamic protein structures that undergo major conformational changes during catalysis, forming super‐complexes and reversibly attaching to thylakoid membranes in a redox‐dependent manner.  相似文献   

8.
The cyclooxygenases (COX-1 and COX-2) are membrane-associated heme-containing homodimers that generate prostaglandin H2 from arachidonic acid (AA). Although AA is the preferred substrate, other fatty acids are oxygenated by these enzymes with varying efficiencies. We determined the crystal structures of AA, eicosapentaenoic acid (EPA), and docosahexaenoic acid (DHA) bound to Co3+-protoporphyrin IX-reconstituted murine COX-2 to 2.1, 2.4, and 2.65 Å, respectively. AA, EPA, and docosahexaenoic acid bind in different conformations in each monomer constituting the homodimer in their respective structures such that one monomer exhibits nonproductive binding and the other productive binding of the substrate in the cyclooxygenase channel. The interactions identified between protein and substrate when bound to COX-1 are conserved in our COX-2 structures, with the only notable difference being the lack of interaction of the carboxylate of AA and EPA with the side chain of Arg-120. Leu-531 exhibits a different side chain conformation when the nonproductive and productive binding modes of AA are compared. Unlike COX-1, mutating this residue to Ala, Phe, Pro, or Thr did not result in a significant loss of activity or substrate binding affinity. Determination of the L531F:AA crystal structure resulted in AA binding in the same global conformation in each monomer. We speculate that the mobility of the Leu-531 side chain increases the volume available at the opening of the cyclooxygenase channel and contributes to the observed ability of COX-2 to oxygenate a broad spectrum of fatty acid and fatty ester substrates.  相似文献   

9.
p‐Coumaric acid decarboxylases (PDCs) catalyze the nonoxidative decarboxylation of hydroxycinnamic acids to generate the corresponding vinyl derivatives. Despite the biotechnological relevance of PDCs in food industry, their catalytic mechanism remains largely unknown. Here, we report insights into the structural basis of catalysis for the homodimeric PDC from Lactobacillus plantarum (LpPDC). The global fold of LpPDC is based on a flattened β‐barrel surrounding an internal cavity. Crystallographic and functional analyses of single‐point mutants of residues located within this cavity have permitted identifying a potential substrate‐binding pocket and also to provide structural evidences for rearrangements of surface loops so that they can modulate the accessibility to the active site. Finally, combination of the structural and functional data with in silico results enables us to propose a two‐step catalytic mechanism for decarboxylation of p‐coumaric acid by PDCs where Glu71 is involved in proton transfer, and Tyr18 and Tyr20 are involved in the proper substrate orientation and in the release of the CO2 product. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

10.
Plants produce short‐chain aldehydes and hydroxy fatty acids, which are important industrial materials, through the lipoxygenase pathway. Based on the information that lipoxygenase activity is up‐regulated in tobacco leaves upon infection with tobacco mosaic virus (TMV), we introduced a melon hydroperoxide lyase (CmHPL) gene, a tomato peroxygenase (SlPXG) gene and a potato epoxide hydrolase (StEH) into tobacco leaves using a TMV‐based viral vector system to afford aldehyde and hydroxy fatty acid production. Ten days after infiltration, tobacco leaves infiltrated with CmHPL displayed high enzyme activities of 9‐LOX and 9‐HPL, which could efficiently transform linoleic acid into C9 aldehydes. Protein extracts prepared from 1 g of CmHPL‐infiltrated tobacco leaves (fresh weight) in combination with protein extracts prepared from 1 g of control vector‐infiltrated tobacco leaves (as an additional 9‐LOX source) produced 758 ± 75 μg total C9 aldehydes in 30 min. The yield of C9 aldehydes from linoleic acid was 60%. Besides, leaves infiltrated with SlPXG and StEH showed considerable enzyme activities of 9‐LOX/PXG and 9‐LOX/EH, respectively, enabling the production of 9,12,13‐trihydroxy‐10(E)‐octadecenoic acid from linoleic acid. Protein extracts prepared from 1 g of SlPXGinfiltrated tobacco leaves (fresh weight) in combination with protein extracts prepared from 1 g of StEH‐infiltrated tobacco leaves produced 1738 ± 27 μg total 9,12,13‐trihydroxy‐10(E)‐octadecenoic acid isomers in 30 min. The yield of trihydroxyoctadecenoic acids from linoleic acid was 58%. C9 aldehydes and trihydroxy fatty acids could likely be produced on a larger scale using this expression system with many advantages including easy handling, time‐saving and low production cost.  相似文献   

11.
  • The green microalga Lobosphaera incisa is an oleaginous eukaryotic alga that is rich in arachidonic acid (20:4). Being rich in this polyunsaturated fatty acid (PUFA), however, makes it sensitive to oxidation. In plants, lipoxygenases (LOXs) are the major enzymes that oxidise these molecules.
  • Here, we describe, to our best knowledge, the first characterisation of a cDNA encoding a LOX (LiLOX) from a green alga. To obtain first insights into its function, we expressed it in E. coli, purified the recombinant enzyme and analysed its enzyme activity.
  • The protein sequence suggests that LiLOX and plastidic LOXs from bryophytes and flowering plants may share a common ancestor. The fact that LiLOX oxidises all PUFAs tested with a consistent oxidation on the carbon n‐6, suggests that PUFAs enter the substrate channel through their methyl group first (tail first). Additionally, LiLOX form the fatty acid hydroperoxide in strict S configuration.
  • LiLOX may represent a good model to study plastid LOX, because it is stable after heterologous expression in E. coli and highly active in vitro. Moreover, as the first characterised LOX from green microalgae, it opens the possibility to study endogenous LOX pathways in these organisms.
  相似文献   

12.
Dicamba (2-methoxy-3,6-dichlorobenzoic acid) O-demethylase (DMO) is the terminal Rieske oxygenase of a three-component system that includes a ferredoxin and a reductase. It catalyzes the NADH-dependent oxidative demethylation of the broad leaf herbicide dicamba. DMO represents the first crystal structure of a Rieske non-heme iron oxygenase that performs an exocyclic monooxygenation, incorporating O2 into a side-chain moiety and not a ring system. The structure reveals a 3-fold symmetric trimer (α3) in the crystallographic asymmetric unit with similar arrangement of neighboring inter-subunit Rieske domain and non-heme iron site enabling electron transport consistent with other structurally characterized Rieske oxygenases. While the Rieske domain is similar, differences are observed in the catalytic domain, which is smaller in sequence length than those described previously, yet possessing an active-site cavity of larger volume when compared to oxygenases with larger substrates. Consistent with the amphipathic substrate, the active site is designed to interact with both the carboxylate and aromatic ring with both key polar and hydrophobic interactions observed. DMO structures were solved with and without substrate (dicamba), product (3,6-dichlorosalicylic acid), and either cobalt or iron in the non-heme iron site. The substitution of cobalt for iron revealed an uncommon mode of non-heme iron binding trapped by the non-catalytic Co2+, which, we postulate, may be transiently present in the native enzyme during the catalytic cycle. Thus, we present four DMO structures with resolutions ranging from 1.95 to 2.2 Å, which, in sum, provide a snapshot of a dynamic enzyme where metal binding and substrate binding are coupled to observed structural changes in the non-heme iron and catalytic sites.  相似文献   

13.
Reactive oxygen species (ROS) production by respiratory Complex I from Escherichia coli was studied in bacterial membrane fragments and in the isolated and purified enzyme, either solubilized or incorporated in proteoliposomes. We found that the replacement of a single amino acid residue in close proximity to the nicotinamide adenine dinucleotide (NADH)‐binding catalytic site (E95 in the NuoF subunit) dramatically increases the reactivity of Complex I towards dioxygen (O2). In the E95Q variant short‐chain ubiquinones exhibit strong artificial one‐electron reduction at the catalytic site, also leading to a stronger increase in ROS production. Two mechanisms can contribute to the observed kinetic effects: (a) a change in the reactivity of flavin mononucleotide (FMN) towards dioxygen at the catalytic site, and (b) a change in the population of the ROS‐generating state. We propose the existence of two (closed and open) states of the NAD+‐bound enzyme as one feature of the substrate‐binding site of Complex I. The analysis of the kinetic model of ROS production allowed us to propose that the population of Complex I with reduced FMN is always low in the wild‐type enzyme even at low ambient redox potentials, minimizing the rate of reaction with O2 in contrast to E95Q variant.  相似文献   

14.
The enzyme 4‐oxalocrotonate tautomerase catalyzes the ketonization of dienols, which after further processing become intermediates in the Krebs cycle. The enzyme uses a general acid–base mechanism for proton transfer: the amino‐terminal proline has been shown to function as the catalytic base and Arg39 has been implicated as the catalytic acid. We report the results of molecular docking simulations of 4‐oxalocrotonate tautomerase with two substrates, 2‐hydroxymuconate and 5‐carboxymethyl‐2‐hydroxymuconate. pKa calculations are also performed for the free enzyme. The predicted binding mode of 2‐hydroxymuconate is in agreement with experimental data. A model for the binding mode of 5‐carboxymethyl‐2‐hydroxymuconate is proposed which explains the lower catalytic efficiency of the enzyme toward this substrate. The pKa predictions and docking simulations support residue Arg39 as the general acid for the enzyme catalysis. © 1999 John Wiley & Sons, Inc. Biopoly 50: 319–328, 1999  相似文献   

15.
Pyrroloquinoline quinone [4,5‐dihydro‐4,5‐dioxo‐1H‐pyrrolo[2,3‐f]quinoline‐2,7,9‐tricarboxylic acid (PQQ)] is a bacterial cofactor in numerous alcohol dehydrogenases including methanol dehydrogenase and glucose dehydrogenase. Its biosynthesis in Klebsiella pneumoniae is facilitated by six genes, pqqABCDEF and proceeds by an unknown pathway. PqqC is one of two metal free oxidases of known structure and catalyzes the last step of PQQ biogenesis which involves a ring closure and an eight‐electron oxidation of the substrate [3a‐(2‐amino‐2‐carboxyethyl)‐4,5‐dioxo‐4,5,6,7,8,9‐hexahydroquinoline‐7,9‐dicarboxylic acid (AHQQ)]. PqqC has 14 conserved active site residues, which have previously been shown to be in close contact with bound PQQ. Herein, we describe the structures of three PqqC active site variants, H154S, Y175F, and the double mutant R179S/Y175S. The H154S crystal structure shows that, even with PQQ bound, the enzyme is still in the “open” conformation with helices α5b and α6 unfolded and the active site solvent accessible. The Y175F PQQ complex crystal structure reveals the closed conformation indicating that Y175 is not required for the conformational change. The R179S/Y175S AHQQ complex crystal structure is the most mechanistically informative, indicating an open conformation with a reaction intermediate trapped in the active site. The intermediate seen in R179S/Y175S is tricyclic but nonplanar, implying that it has not undergone oxidation. These studies implicate a stepwise process in which substrate binding leads to the generation of the closed protein conformation, with the latter playing a critical role in O2 binding and catalysis. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

16.
The β-hydroxyacyl-acyl carrier protein dehydratase of Plasmodium falciparum (PfFabZ) catalyzes the third and important reaction of the fatty acid elongation cycle. The crystal structure of PfFabZ is available in hexameric (active) and dimeric (inactive) forms. However, PfFabZ has not been crystallized with any bound inhibitors until now. We have designed a new condition to crystallize PfFabZ with its inhibitors bound in the active site, and determined the crystal structures of four of these complexes. This is the first report on any FabZ enzyme with active site inhibitors that interact directly with the catalytic residues. Inhibitor binding not only stabilized the substrate binding loop but also revealed that the substrate binding tunnel has an overall shape of “U”. In the crystal structures, residue Phe169 located in the middle of the tunnel was found to be in two different conformations, open and closed. Thus, Phe169, merely by changing its side chain conformation, appears to be controlling the length of the tunnel to make it suitable for accommodating longer substrates. The volume of the substrate binding tunnel is determined by the sequence as well as by the conformation of the substrate binding loop region and varies between organisms for accommodating fatty acids of different chain lengths. This report on the crystal structures of the complexes of PfFabZ provides the structural basis of the inhibitory mechanism of the enzyme that could be used to improve the potency of inhibitors against an important component of fatty acid synthesis common to many infectious organisms.  相似文献   

17.
Secreted LOX from Pseudomonas aeruginosa (PA-LOX) has previously been identified as arachidonic acid 15S-lipoxygenating enzyme. Here we report that the substitution of Ala420Gly in PA-LOX leads to an enzyme variant with pronounced dual specificity favoring arachidonic acid 11R-oxygenation. When compared with other LOX-isoforms the molecular oxygen affinity of wild-type PA-LOX is 1–2 orders of magnitude lower (Km O2 of 0.4 mM) but Ala420Gly exchange improved the molecular oxygen affinity (Km O2 of 0.2 mM). Experiments with stereo-specifically deuterated linoleic acid indicated that the formation of both 13S- and 9R-HpODE involves abstraction of the proS-hydrogen from C11 of the fatty acid backbone. To explore the structural basis for the observed functional changes (altered specificity, improved molecular oxygen affinity) we solved the crystal structure of the Ala420Gly mutant of PA-LOX at 1.8 Å resolution and compared it with the wild-type enzyme. Modeling of fatty acid alignment at the catalytic center suggested that in the wild-type enzyme dioxygen is directed to C15 of arachidonic acid by a protein tunnel, which interconnects the catalytic center with the protein surface. Ala420Gly exchange redirects intra-enzyme O2 diffusion by bifurcating this tunnel so that C11 of arachidonic acid also becomes accessible for O2 insertion.  相似文献   

18.
Creatininase is a binuclear zinc enzyme and catalyzes the reversible conversion of creatinine to creatine. It exhibits an open-closed conformational change upon substrate binding, and the differences in the conformations of Tyr121, Trp154, and the loop region containing Trp174 were evident in the enzyme-creatine complex when compared to those in the ligand-free enzyme. We have determined the crystal structure of the enzyme complexed with a 1-methylguanidine. All subunits in the complex existed as the closed form, and the binding mode of creatinine was estimated. Site-directed mutagenesis revealed that the hydrophobic residues that show conformational change upon substrate binding are important for the enzyme activity.We propose a catalytic mechanism of creatininase in which two water molecules have significant roles. The first molecule is a hydroxide ion (Wat1) that is bound as a bridge between the two metal ions and attacks the carbonyl carbon of the substrate. The second molecule is a water molecule (Wat2) that is bound to the carboxyl group of Glu122 and functions as a proton donor in catalysis. The activity of the E122Q mutant was very low and it was only partially restored by the addition of ZnCl2 or MnCl2. In the E122Q mutant, kcat is drastically decreased, indicating that Glu122 is important for catalysis. X-ray crystallographic study and the atomic absorption spectrometry analysis of the E122Q mutant-substrate complex revealed that the drastic decrease of the activity of the E122Q was caused by not only the loss of one Zn ion at the Metal1 site but also a critical function of Glu122, which most likely exists for a proton transfer step through Wat2.  相似文献   

19.
We report the cocrystal structures of a computationally designed and experimentally optimized retro-aldol enzyme with covalently bound substrate analogs. The structure with a covalently bound mechanism-based inhibitor is similar to, but not identical with, the design model, with an RMSD of 1.4 Å over active-site residues and equivalent substrate atoms. As in the design model, the binding pocket orients the substrate through hydrophobic interactions with the naphthyl moiety such that the oxygen atoms analogous to the carbinolamine and β-hydroxyl oxygens are positioned near a network of bound waters. However, there are differences between the design model and the structure: the orientation of the naphthyl group and the conformation of the catalytic lysine are slightly different; the bound water network appears to be more extensive; and the bound substrate analog exhibits more conformational heterogeneity than typical native enzyme–inhibitor complexes. Alanine scanning of the active-site residues shows that both the catalytic lysine and the residues around the binding pocket for the substrate naphthyl group make critical contributions to catalysis. Mutating the set of water-coordinating residues also significantly reduces catalytic activity. The crystal structure of the enzyme with a smaller substrate analog that lacks naphthyl ring shows the catalytic lysine to be more flexible than in the naphthyl–substrate complex; increased preorganization of the active site would likely improve catalysis. The covalently bound complex structures and mutagenesis data highlight the strengths and weaknesses of the de novo enzyme design strategy.  相似文献   

20.
We have previously shown that a coenzyme-B12 analog, adenosylcobalamin (AdoCbl)-(e-OH), with the e-propionamide group converted to a carboxylic acid, serves as a poor coenzyme for dioldehydrase. During the course of the catalytic process, the enzyme AdoCbl-(e-OH) complex becomes catalytically inactive (T. Toraya, E. Krodel, A. S. Mildvan, and R. H. Abeles, 1979, Biochemistry18, 417–426). We have now examined the mechanism of this inactivation further. Inactivation only occurs in the presence of substrate. The dioldehydrase coenzyme analog complex is stable in the absence of substrate. In the inactivated complex, the coenzyme analog was stoichiometrically converted to a cob(II)alamin species. The cob-(II)alamin formed remained irreversibly bound at the active site of the enzyme and resisted oxidation by O2 even in the presence of CN?. Stoichiometric formation of 5′-deoxyadenosine from the 5′-deoxy-5′-adenosyl moiety of the coenzyme analog was demonstrated with [8-14C]-AdoCbl(e-OH). This nucleoside also remained tightly bound to the enzyme and was not exchangeable with free 5′-deoxyadenosine nor was it removed by Sephadex chromatography. The rate of inactivation showed no deuterium isotope effect when the inactivation occurred in the presence of l,2-propanediol-l-d2. The inactivated complex was resolved by acid ammonium sulfate treatment into the intact apoenzyme and the hydroxocobalamin derivative. This indicates that the apoenzyme itself is not modified in the inactivation process. These results suggest that the inactivation reaction occurs from one of the intermediates in the normal catalysis. We propose that the inactivation is due to incorrect binding of the modified coenzyme in an intermediate of the catalytic process. This incorrect binding leads to the loss of the substrate radical, and consequently, to loss of catalytic 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