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 28 毫秒
1.
It is widely recognized that the two major forms of GAD present in adult vertebrate brains are each composed of two major sequence domains that differ in size and degree of similarity. The amino-terminal domain is smaller and shows little sequence identity between the two forms. This domain is thought to mediate the subcellular targeting of the two GADs. Substantial parts of the amino-terminal domain appear to be exposed and flexible, as shown by proteolysis experiments and the locations of posttranslational modifications. The carboxyl-terminal sequence domain contains the catalytic site and shows substantial sequence similarity between the forms. The interaction of GAD with its cofactor, pyridoxal-5' phosphate (pyridoxal-P), plays a key role in the regulation of GAD activity. Although GAD(65) and GAD(67) interact differently with pyridoxal-P, their cofactor-binding sites contain the same set of nine putative cofactor-binding residues and have the same basic structural fold. Thus the cofactor-binding differences cannot be attributed to fundamental structural differences between the GADs but must result from subtle modifications of the basic cofactor-binding fold. The presence of another conserved motif suggests that the carboxyl-terminal domain is composed of two functional domains: the cofactor-binding domain and a small domain that closes when the substrate binds. Finally, GAD is a dimeric enzyme and conserved features of GADs superfamily of pyridoxal-P proteins indicate the dimer-forming interactions are mediated mainly by the carboxyl-terminal domain.  相似文献   

2.
We used sequence and structural comparisons to determine the fold for eukaryotic ornithine decarboxylase, which we found is related to alanine racemase. These enzymes have no detectable sequence identity with any protein of known structure, including three pyridoxal phosphate-utilizing enzymes. Our studies suggest that the N-terminal domain of ornithine decarboxylase folds into a beta/alpha-barrel. Through the analysis of known barrel structures we developed a topographic model of the pyridoxal phosphate-binding domain of ornithine decarboxylase, which predicts that the Schiff base lysine and a conserved glycine-rich sequence both map to the C-termini of the beta-strands. Other residues in this domain that are likely to have essential roles in catalysis, substrate, and cofactor binding were also identified, suggesting that this model will be a suitable guide to mutagenic analysis of the enzyme mechanism.  相似文献   

3.
More than 50% of glutamate decarboxylase (GAD) in brain is present as apoenzyme. Recent work has opened the possibility that apoGAD can be studied in brain by labeling with radioactive cofactor. Such studies would be aided by a compound that inhibits specific binding. One possibility is 4-deoxy-pyridoxine 5-phosphate, a close structural analog of the cofactor pyridoxal 5-phosphate. The effects of deoxypyridoxine-P on the cyclic series of reactions that interconverts apo- and holoGAD was investigated and found to be consistent with simple competitive inhibition of the activation of apoGAD by pyridoxal-P. As expected from the cycle GAD was inactivated when incubated with glutamate and deoxypyridoxine-P even though cofactor was present, but no inactivation was observed with deoxypyridoxine-P in the absence of glutamate. Deoxypyridoxine-P also stabilized apoGAD against heat denaturation. These effects were quantitatively accounted for by a kinetic model of the apo-holoGAD cycle. Deoxypyridoxine-P inhibited the labeling by [32P]pyridoxal-P of GAD isolated from rat brain. Hippocampal extracts were labeled with [32P]pyridoxal-P and analyzed by SDS-polyacrylamide gel electrophoresis. Remarkably few bands were strongly labeled. The major labeled band (at 63 kDa) corresponded to one of the forms of GAD. Other strongly-labeled bands were observed at 65 kDa (corresponding to the higher molecular weight form of GAD) and at 69–72 kDa. Labeling of the 63- and 65-kDa bands was inhibited by deoxypyridoxine-P, but the 69–72 kDa bands were unaffected, suggesting that the latter were non-specifically labeled. The results suggest that the 63-kDa form of GAD makes up the majority of apoGAD in hippocampus.Special issue dedicated to Dr. Eugene Roberts.  相似文献   

4.
Cofactor and tryptophan accessibility of the 65-kDa form of rat brain glutamate decarboxylase (GAD) was investigated by fluorescence quenching measurements using acrylamide, I-, and Cs+ as the quenchers. Trp residues were partially exposed to solvent. I- was less able and Cs+ was more able to quench the fluorescence of Trp residues in the holoenzyme of GAD (holoGAD) than the apoenzyme (apoGAD). The fraction of exposed Trp residues were in the range of 30-49%. In contrast, pyridoxal-P bound to the active site of GAD was exposed to solvent. I- was more able and Cs+ was less able to quench the fluorescence of pyridoxal-P in holoGAD. The cofactor was present in a positively charged microenvironment, making it accessible for interactions with anions. A difference in the exposure of Trp residues and pyridoxal-P to these charged quenchers suggested that the exposed Trp residues were essentially located outside of the active site. Changes in the accessibility of Trp residues upon pyridoxal-P binding strongly supported a significant conformational change in GAD. Fluorescence intensity measurements were also carried out to investigate the unfolding of GAD using guanidine hydrochloride (GdnHCl) as the denaturant. At 0.8-1.5 M GdnHCl, an intermediate step was observed during the unfolding of GAD from the native to the denatured state, and was not found during the refolding of GAD from the denatured to native state, indicating that this intermediate step was not a reversible process. However, at >1.5 M GdnHCl for holoGAD and >2.0 M GdnHCl for apoGAD, the transition leading to the denatured state was reversible. It was suggested that the intermediate step involved the dissociation of native dimer of GAD into monomers and the change in the secondary structure of the protein. Circular dichroism revealed a decrease in the alpha-helix content of GAD from 36 to 28%. The unfolding pattern suggested that GAD may consist of at least two unfolding domains. Unfolding of the lower GdnHCl-resisting domain occurred at a similar concentration of denaturant for apoGAD and holoGAD, while unfolding of the higher GdnHCl-resisting domain occurred at a higher concentration of GdnHCl for apoGAD than holoGAD.  相似文献   

5.
The conformation of bovine serum albumin (BSA), a pyridoxal phosphate (pyridoxal-P) carrier, was investigated by using uv/visible spectrophotometry, fluorescence spectroscopy, circular dichroism, and differential scanning microcalorimetry. Upon interacting with pyridoxal-P, the uv/visible absorption spectrum of BSA exhibits peaks at 330 and 392 nm due to the formation of a Schiff base. Pyridoxal-P quenches the fluorescence emission intensity (excited at 295 or 280 nm) by 24% and enhances fluorescence steady-state polarization of BSA by 20%. These observations suggest a conformational change in BSA when it interacts with pyridoxal-P. However, this conformational change appears to be small since circular dichroism showed only a 2-4% decrease in the alpha-helical content of BSA and no change in the beta-sheet content, and differential scanning microcalorimetry yielded only a 10% change in the enthalpy of thermal unfolding of BSA. 2-Aminoethylisothiouronium bromide, an antioxidant, causes no effect on either uv/visible absorption spectrum or fluorescence emission intensity of BSA, suggesting that BSA lacks sensitive sulfhydryl groups. To help in understanding BSA as a carrier for pyridoxal-P, the results were compared with those for glutamate decarboxylase (GAD), a pyridoxal-P-dependent protein, which requires pyridoxal-P as the cofactor for activity. Although BSA and GAD exhibit comparable molecular weights (66430 versus 65300), numbers of amino acid residues (582 versus 585), and binding affinity (>10(6) M-1), distinct conformational alterations occur between the two proteins upon interacting with pyridoxal-P: a small conformational change for BSA versus a large conformational change for GAD. In contrast to the case of BSA, AET causes significant effects on both the uv/visible spectrum and fluorescence emission intensity of GAD, because GAD contains sensitive sulfhydryl groups. Factors such as disulfide bond and active site sequence were discussed to understand BAS as a carrier for pyridoxal-P and a pyridoxal-P-independent protein.  相似文献   

6.
Glutamate-Dependent Active-Site Labeling of Brain Glutamate Decarboxylase   总被引:3,自引:3,他引:0  
A major regulatory feature of brain glutamate decarboxylase (GAD) is a cyclic reaction that controls the relative amounts of holoenzyme and apoenzyme [active and inactive GAD with and without bound pyridoxal 5'-phosphate (pyridoxal-P, the cofactor), respectively]. Previous studies have indicated that progression of the enzyme around the cycle should be stimulated strongly by the substrate, glutamate. To test this prediction, the effect of glutamate on the incorporation of pyridoxal-P into rat-brain GAD was studied by incubating GAD with [32P]pyridoxal-P, followed by reduction with NaBH4 to link irreversibly the cofactor to the enzyme. Adding glutamate to the reaction mixture strongly stimulated labeling of GAD, as expected. 4-Deoxypyridoxine 5'-phosphate (deoxypyridoxine-P), a close structural analogue of pyridoxal-P, was a competitive inhibitor of the activation of glutamate apodecarboxylase by pyridoxal-P (Ki = 0.27 microM) and strongly inhibited glutamate-dependent labeling of GAD. Analysis of labeled GAD by sodium dodecyl sulfate (SDS)-polyacrylamide gel electrophoresis showed two labeled proteins with apparent molecular masses of 59 and 63 kDa. Both proteins could be purified by immunoaffinity chromatography on a column prepared with a monoclonal antibody to GAD, and both were labeled in a glutamate-dependent, deoxypyridoxine-P-sensitive manner, indicating that both were GAD. Three peaks of GAD activity (termed peaks I, II, and III) were separated by chromatography on phenyl-Sepharose, labeled with [32P]pyridoxal-P, purified by immunoaffinity chromatography, and analyzed by SDS-polyacrylamide gel electrophoresis. Peak I contained only the 59-kDa labeled protein. Peaks II and III contained the both the 59- and 63-kDa proteins, but in differing proportions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Birkholtz L  Joubert F  Neitz AW  Louw AI 《Proteins》2003,50(3):464-473
The ornithine decarboxylase (ODC) component of the bifunctional S-adenosylmethionine decarboxylase/ornithine decarboxylase enzyme (PfAdoMetDC-ODC) of Plasmodium falciparum was modeled on the crystal structure of the Trypanosoma brucei enzyme. The homology model predicts a doughnut-shaped active homodimer that associates in a head-to-tail manner. The monomers contain two distinct domains, an N-terminal alpha/beta-barrel and a C-terminal modified Greek-key domain. These domains are structurally conserved between eukaryotic ODC enzymes and are preserved in distant analogs such as alanine racemase and triosephosphate isomerase-like proteins. Superimposition of the PfODC model on the crystal structure of the human enzyme indicates a significant degree of deviation in the carbon alpha-backbone of the solvent accessible loops. The surface locality of the ab initio modeled 38 amino acid parasite-specific insert suggests a role in the stabilization of the large bifunctional protein complex. The active site pockets of PfODC at the interface between the monomers appear to be conserved regarding the binding sites of the cofactor and substrate, but each contains five additional malaria-specific residues. The predicted PfODC homology model is consistent with mutagenesis results and biochemical studies concerning the active site residues and areas involved in stabilizing the dimeric form of the protein. Two competitive inhibitors of PfODC could be shown to interact with several parasite-specific residues in comparison with their interaction with the human ODC. The PfODC homology model contributes toward a structure-based approach for the design of novel malaria-specific inhibitors.  相似文献   

8.
The association between glutamate decarboxylase (GAD) and its cofactor, pyridoxal-5′-phos-phate (pyridoxal-P), was studied using 20,0000 supernatant of rat brain. In this preparation GAD required added pyridoxal-P to maintain a linear reaction rate beyond 5 min of incubation. Following exhaustive dialysis the enzyme was more than 83% saturated with cofactor indicating that the cofactor was tightly bound to the enzyme. When incubations were performed in the presence of glutamate and without added pyridoxal-P there was a progressive inactivation of the enzyme which was dependent on the glutamate concentration. This lost activity was almost completely recovered by addition of pyridoxal-P to the dialyzed glutamate-inactivated enzyme. The results suggest that glutamate inactivates GAD by promoting the dissociation of pyridoxal-P from the enzyme thereby producing inactive apoen-zyme which can be reactivated by combining with available pyridoxal-P. This interpretation is supported by the finding that progress curves for the reaction were accurately described over a 30 min incubation period and 10-fold glutamate concentration range by an integrated rate equation which takes the glutamate-promoted dissociation of cofactor into account. The progressive inactivation could not be attributed to denaturation of the enzyme, impurities in the substrate, effects of pH, depletion of substrate, protein concentration, sulfhydryl reagents or product inhibition. The results presented here also show that certain precautions must be adopted to accurately measure GAD activity in the absence of added pyridoxal-P as has been widely done in studies of drug action. Specifically, measurements must be made at short times of incubation and low concentrations of glutamate to minimize the glutamate-promoted inactivation of the enzyme.  相似文献   

9.
Glutamate decarboxylase (GAD, EC 4.1.1.15) catalyses the alpha-decarboxylation of glutamate to produce gamma-aminobutyrate (GABA). The nucleotide sequences of two divergent GADs (designated GAD1 and GAD3) were isolated from a Nicotiana tabacum L. cv. Samsun NN leaf cDNA library. Open reading frames indicated that GAD1 encodes a polypeptide of 496 amino acids and has greater than 99% identity with known tobacco GADs, whereas GAD3 encodes a polypeptide of 491 amino acids and has about 14% divergence from known tobacco GADs. Genomic DNA analysis suggested that there are at least four tobacco GAD genes, existing in pairs of highly identical genes. An in vitro assay at pH 7.3 revealed that activities of the recombinant proteins, after isolation from Escherichia coli and partial purification by nickel-affinity chromatography, are 57-133 times the control levels in the presence of 0.5 mM calcium and 0.2 micro M bovine calmodulin.  相似文献   

10.
Two genes encode distinct glutamate decarboxylases   总被引:54,自引:0,他引:54  
gamma-Aminobutyric acid (GABA) is the most widely distributed known inhibitory neurotransmitter in the vertebrate brain. GABA also serves regulatory and trophic roles in several other organs, including the pancreas. The brain contains two forms of the GABA synthetic enzyme glutamate decarboxylase (GAD), which differ in molecular size, amino acid sequence, antigenicity, cellular and subcellular location, and interaction with the GAD cofactor pyridoxal phosphate. These forms, GAD65 and GAD67, derive from two genes. The distinctive properties of the two GADs provide a substrate for understanding not only the multiple roles of GABA in the nervous system, but also the autoimmune response to GAD in insulin-dependent diabetes mellitus.  相似文献   

11.
Abstract— The level of saturation of glutamate decarboxylase (GAD) by cofactor, pyridoxal-5'-phosphate (pyridoxal-P), determined in synaptosomes prepared from substantia nigra tissue, was reduced from 45 to 28%; when ATP was included in the homogenizing medium to prevent nonspecific activation of GAD by endogenous pyridoxal-P. When the synaptosomes were incubated for 5–20 min at 37°C in Krebs-Ringer phosphate buffer (KRP), the level of saturation of GAD by cofactor decreased further, from 28 to 20%. Depolarization of the nigral synaptosomes by either high K+ (55 mM) or veratridine resulted in a significant increase in the level of GAD saturation by cofactor, to 32 and 41%. respectively. Omitting Ca2+ from the incubation medium blocked the depolarization-induced rise in the level of saturation. Depolarization with high K+ and veratridine also caused a significant decrease in the ATP concentration in the synaptosomes. No difference in ATP concentration was observed when the samples were incubated at 37°C for 5–20min or incubated in the absence of added Ca2+ with high K+. Results provide further evidence that in vivo brain GAD is largely unsaturated by cofactor and support the possibility that increased release and utilization of GABA may be associated with increases in the amount of pyridoxal-P endogenously bound to GAD in nerve terminals.  相似文献   

12.
Glutamic acid decarboxylase is responsible for synthesizing GABA, the major inhibitory neurotransmitter, and exists in two isoforms—GAD65 and GAD67. The enzyme is cleaved under excitotoxic conditions, but the mechanisms involved and the functional consequences are not fully elucidated. We found that excitotoxic stimulation of cultured hippocampal neurons with glutamate leads to a time-dependent cleavage of GAD65 and GAD67 in the N-terminal region of the proteins, and decrease the corresponding mRNAs. The cleavage of GAD67 was sensitive to the proteasome inhibitors MG132, YU102 and lactacystin, and was also abrogated by the E1 ubiquitin ligase inhibitor UBEI-41. In contrast, MG132 and UBEI-41 were the only inhibitors tested that showed an effect on GAD65 cleavage. Excitotoxic stimulation with glutamate also increased the amount of GAD captured in experiments where ubiquitinated proteins and their binding partners were isolated. However, no evidences were found for direct GADs ubiquitination in cultured hippocampal neurons, and recombinant GAD65 was not cleaved by purified 20S or 26S proteasome preparations. Since calpains, a group of calcium activated proteases, play a key role in GAD65/67 cleavage under excitotoxic conditions the results suggest that GADs are cleaved after ubiquitination and degradation of an unknown binding partner by the proteasome. The characteristic punctate distribution of GAD65 along neurites of differentiated cultured hippocampal neurons was significantly reduced after excitotoxic injury, and the total GAD activity measured in extracts from the cerebellum or cerebral cortex at 24h postmortem (when there is a partial cleavage of GADs) was also decreased. The results show a role of the UPS in the cleavage of GAD65/67 and point out the deregulation of GADs under excitotoxic conditions, which is likely to affect GABAergic neurotransmission. This is the first time that the UPS has been implicated in the events triggered during excitotoxicity and the first molecular target of the UPS affected in this cell death process.  相似文献   

13.
14.
The interactions between glutamate decarboxylase (GAD) and its cofactor pyridoxal phosphate (PLP) play a key role in the regulation of GAD activity. The enzyme has two isoforms, GAD65 and GAD67. A comparison of binding constants, rate constants, and kinetic profiles for the formation of holoenzyme (holoGAD65 and holoGAD67) revealed that the two isoforms interact distinctively with the cofactor. GAD67 exhibits a higher binding constant for PLP binding, making it more difficult to dissociate PLP from holoGAD67 than holoGAD65. Meanwhile, PLP binding occurs at a much slower rate for GAD67 than GAD65, as evidenced by lower rate constants and a slower initial rate of the holoenzyme formation. Job's plots revealed a stoichiometry of 1:1 for PLP binding to GAD65 before and after the saturation level of PLP, while 1:2 for PLP binding to GAD67 prior to the saturation of PLP and 1:1 at the saturation level of PLP. These results suggested that the two binding sites of GAD65 exhibit similar affinities for PLP. In contrast, one binding site of GAD67 exhibits a significantly higher affinity for PLP than the other binding site. Based on these findings, it was proposed that a slower PLP binding to GAD67 than GAD65 and a less ease to dissociate PLP from holoGAD67 than holoGAD65 are important underlying factors. This attributes to GAD67 being more highly saturated by PLP and GAD65 being less saturated by PLP. A larger conformation change constant for GAD67 than GAD65 supported a significant conformational change induced by the initial PLP binding to GAD67, which affects the other binding site affinity of GAD67. The present studies provided valuable insights into distinctive properties between the two isoforms of GAD.  相似文献   

15.
The smaller isoform of the GABA-synthesizing enzyme, glutamic acid decarboxylase 65 (GAD65), is unusually susceptible to becoming a target of autoimmunity affecting its major sites of expression, GABA-ergic neurons and pancreatic beta-cells. In contrast, a highly homologous isoform, GAD67, is not an autoantigen. We used homolog-scanning mutagenesis to identify GAD65-specific amino acid residues which form autoreactive B-cell epitopes in this molecule. Detailed mapping of 13 conformational epitopes, recognized by human monoclonal antibodies derived from patients, together with two and three-dimensional structure prediction led to a model of the GAD65 dimer. GAD65 has structural similarities to ornithine decarboxylase in the pyridoxal-5'-phosphate-binding middle domain (residues 201-460) and to dialkylglycine decarboxylase in the C-terminal domain (residues 461-585). Six distinct conformational and one linear epitopes cluster on the hydrophilic face of three amphipathic alpha-helices in exons 14-16 in the C-terminal domain. Two of those epitopes also require amino acids in exon 4 in the N-terminal domain. Two distinct epitopes reside entirely in the N-terminal domain. In the middle domain, four distinct conformational epitopes cluster on a charged patch formed by amino acids from three alpha-helices away from the active site, and a fifth epitope resides at the back of the pyridoxal 5'-phosphate binding site and involves amino acid residues in exons 6 and 11-12. The epitopes localize to multiple hydrophilic patches, several of which also harbor DR*0401-restricted T-cell epitopes, and cover most of the surface of the protein. The results reveal a remarkable spectrum of human autoreactivity to GAD65, targeting almost the entire surface, and suggest that native folded GAD65 is the immunogen for autoreactive B-cells.  相似文献   

16.
The S-adenosylmethionine-dependent methyltransferase enzymes share little sequence identity, but incorporate a highly conserved structural fold. Surprisingly, residues that bind the common cofactor are poorly conserved, although the binding site is localised to the same region of the fold. The substrate-binding region of the fold varies enormously. Over the past two years, there has been a significant increase in the number of structures that are known to incorporate this fold, including several uncharacterized proteins and two proteins that lack methyltransferase activity.  相似文献   

17.
Hoffman DW  Carroll D  Martinez N  Hackert ML 《Biochemistry》2005,44(35):11777-11785
Antizyme and its isoforms are members of an unusual yet broadly conserved family of proteins, with roles in regulating polyamine levels within cells. Antizyme has the ability to bind and inhibit the enzyme ornithine decarboxylase (ODC), targeting it for degradation at the proteasome; antizyme is also known to affect the transport of polyamines and interact with the antizyme inhibitor protein (AZI), as well as the cell-cycle protein cyclin D1. In the present work, NMR methods were used to determine the solution structure of a stable, folded domain of mammalian antizyme isoform-1 (AZ-1), consisting of amino acid residues 87-227. The protein was found to contain eight beta strands and two alpha helices, with the strands forming a mixed parallel and antiparallel beta sheet. At the level of primary sequence, antizyme is not similar to any protein of known structure, and results show that antizyme exhibits a novel arrangement of its strands and helices. Interestingly, however, the fold of antizyme is similar to that found in a family of acetyl transferases, as well as translation initiation factor IF3, despite a lack of functional relatedness between these proteins. Structural results, combined with amino acid sequence comparisons, were used to identify conserved features among the various homologues of antizyme and their isoforms. Conserved surface residues, including a cluster of acidic amino acids, were found to be located on a single face of antizyme, suggesting this surface is a possible site of interaction with target proteins such as ODC. This structural model provides an essential framework for an improved future understanding of how the different parts of antizyme play their roles in polyamine regulation.  相似文献   

18.
Conserved lysines of mouse ornithine decarboxylase were individually mutated to arginines. The mutations at amino acid residues 69, 115, and 169 greatly reduced or abolished enzymatic activity. Lysine 69 is the site of Schiff base formation with the cofactor pyridoxal phosphate; the functional role of the other two lysines essential for activity is not known. Coexpression of wild type ornithine decarboxylase along with the lysine 115 to arginine mutant reduced the activity of the former without diminishing the amount of wild type protein. This form of negative complementation was seen when wild type and mutant protein were coexpressed either by in vitro translation or in bacteria. The data are consistent with the conclusion that a wild type and mutant subunit form a heterodimer that is enzymatically inactive.  相似文献   

19.
Reactivation of substrate-inactivated brain glutamate decarboxylase   总被引:8,自引:0,他引:8  
The effects of ATP and inorganic phosphate (Pi) on the reactivation of glutamate apodecarboxylase by its cofactor pyridoxal-5'-phosphate (pyridoxal-P) was studied. Apoenzyme was prepared by preincubation with glutamate. Apoenzyme prepared with glutamate alone was reactivated slowly and incompletely by adding a saturating concentration of pyridoxal-P (20 microM). Reactivation was slightly enhanced by 1-10 mM Pi. Reactivation by pyridoxal-P plus Pi was greatly enhanced by the presence of low concentrations (less than 100 microM) of ATP during the preparation of apoenzyme with glutamate. Reactivation was much lower if Pi was omitted. Enhancement of reactivation by ATP was due to its effect during apoenzyme formation, since ATP did not enhance reactivation if added only during reactivation and since the enhancing effect persisted after the removal of free ATP by chromatography on Sephadex G-25 after apoenzyme preparation and before reactivation. Reactivation was inhibited by high concentrations of ATP (greater than 100 microM), possibly by competition of ATP for the cofactor binding site. Four factors (glutamate, pyridoxal-P, ATP, and Pi) control a cycle of inactivation and reactivation that appears to be important in the regulation of brain glutamate decarboxylase.  相似文献   

20.
Schubert HL  Hill CP 《Biochemistry》2006,45(51):15188-15196
Mutations in the gene encoding human ATP:cobalamin adenosyltransferase (hATR) can result in the metabolic disorder known as methylmalonic aciduria (MMA). This enzyme catalyzes the final step in the conversion of cyanocobalamin (vitamin B12) to the essential human cofactor adenosylcobalamin. Here we present the 2.5 A crystal structure of ATP bound to hATR refined to an Rfree value of 25.2%. The enzyme forms a tightly associated trimer, where the monomer comprises a five-helix bundle and the active sites lie on the subunit interfaces. Only two of the three active sites within the trimer contain the bound ATP substrate, thereby providing examples of apo- and substrate-bound-active sites within the same crystal structure. Comparison of the empty and occupied sites indicates that twenty residues at the enzyme's N-terminus become ordered upon binding of ATP to form a novel ATP-binding site and an extended cleft that likely binds cobalamin. The structure explains the role of 20 invariant residues; six are involved in ATP binding, including Arg190, which hydrogen bonds to ATP atoms on both sides of the scissile bond. Ten of the hydrogen bonds are required for structural stability, and four are in positions to interact with cobalamin. The structure also reveals how the point mutations that cause MMA are deficient in these functions.  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