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.
Phosphoglucose isomerase (EC 5.3.1.9) catalyzes the interconversion of D-glucopyranose-6-phosphate and D-fructofuranose-6-phosphate by promoting an intrahydrogen transfer between C1 and C2. A conserved histidine exists throughout all phosphoglucose isomerases and was hypothesized to be the base catalyzing the isomerization reaction. In the present study, this conserved histidine, His311, of the enzyme from Bacillus stearothermophilus was subjected to mutational analysis, and the mutational effect on the inactivation kinetics by N-bromoacetylethanolamine phosphate was investigated. The substitution of His311 with alanine, asparagine, or glutamine resulted in the decrease of activity, in k(cat)/K(M), by a factor of 10(3), indicating the importance of this residue. N-bromoacetylethanolamine phosphate inactivated irreversibly the activity of wild-type phosphoglucose isomerase; however, His311 --> Ala became resistant to this inhibitor, indicating that His311 is located in the active site and is responsible for the inactivation of the enzyme by this active site-directed inhibitor. The pKa of His311 was estimated to be 6.31 according to the pH dependence of the inactivation. The proximity of this value with the pKa value of 6.35, determined from the pH dependence of k(cat)/K(M), supports a role of His311 as a general base in the catalysis.  相似文献   

2.
Pyrococcus furiosus phosphoglucose isomerase (PfPGI) is a metal-containing enzyme that catalyses the interconversion of glucose 6-phosphate (G6P) and fructose 6-phosphate (F6P). The recent structure of PfPGI has confirmed the hypothesis that the enzyme belongs to the cupin superfamily and identified the position of the active site. This fold is distinct from the alphabetaalpha sandwich fold commonly seen in phosphoglucose isomerases (PGIs) that are found in bacteria, eukaryotes and some archaea. Whilst the mechanism of the latter family is thought to proceed through a cis-enediol intermediate, analysis of the structure of PfPGI in the presence of inhibitors has led to the suggestion that the mechanism of this enzyme involves the metal-dependent direct transfer of a hydride between C1 and C2 atoms of the substrate. To gain further insight in the reaction mechanism of PfPGI, the structures of the free enzyme and the complexes with the inhibitor, 5-phospho-d-arabinonate (5PAA) in the presence and absence of metal have been determined. Comparison of these structures with those of equivalent complexes of the eukaryotic PGIs reveals similarities at the active site in the disposition of possible catalytic residues. These include the presence of a glutamic acid residue, Glu97 in PfPGI, which occupies the same position relative to the inhibitor as that of the glutamate that is thought to function as the catalytic base in the eukaryal-type PGIs. These similarities suggest that aspects of the catalytic mechanisms of these two structurally unrelated PGIs may be similar and based on an enediol intermediate.  相似文献   

3.
Phosphoglucose isomerase (PGI) catalyzes the interconversion between glucose 6-phosphate and fructose 6-phosphate in the glycolysis pathway. In mammals, the enzyme is also identical to the extracellular proteins neuroleukin, tumor-secreted autocrine motility factor (AMF) and differentiation and maturation mediator for myeloid leukemia. Hereditary deficiency of the enzyme causes non-spherocytic hemolytic anemia in human. In the present study, a novel interaction between GTP and human PGI was corroborated by UV-induced crosslinking, affinity purification and kinetic study. GTP not only inhibits the isomerization activity but also compromises the AMF function of the enzyme. Kinetic studies, including the Yonetani-Theorell method, suggest that GTP is a competitive inhibitor with a Ki value of 63 μM and the GTP-binding site partially overlaps with the catalytic site. In addition, GTP stabilizes the structure of human PGI against heat- and detergent-induced denaturation. Molecular modelling and dynamic simulation suggest that GTP is bound in a syn-conformation with the γ-phosphate group located near the phosphate-binding loop and the ribose moiety positioned away from the active-site residues.  相似文献   

4.
5.
Structural genomics demonstrates that despite low levels of structural similarity of proteins comprising a metabolic pathway, their substrate binding regions are likely to be conserved. Herein based on the 3D-structures of the α/β-fold proteins involved in the ara operon, we attempted to predict the substrate binding residues of thermophilic Geobacillus stearothermophilusl-arabinose isomerase (GSAI) with no 3D-structure available. Comparison of the structures of l-arabinose catabolic enzymes revealed a conserved feature to form the substrate-binding modules, which can be extended to predict the substrate binding site of GSAI (i.e., D195, E261 and E333). Moreover, these data implicated that proteins in the l-arabinose metabolic pathway might retain their substrate binding niches as the modular structure through conserved molecular evolution even with totally different structural scaffolds.  相似文献   

6.
The interactions of phosphoglucose isomerase with solutions of rapidly interconverting oxovanadates are described. Quantitative speciation analyses of vanadate monomer, dimer, tetramer and pentamer were performed under the conditions of the enzyme assay. 2D EXSY51V NMR spectroscopy was used to demonstrate that the oligoanions are sufficiently long-lived to be recognized by the enzyme as different species. The observations are consistent with the vanadate tetramer being responsible for the observed inhibition. Selected oxomolybdates ([(CH3)2AsMo4OH]2– and [(NH3C2H4P)2Mo5O21]2–) were also found to inhibit 6-phosphogluconate dehydrogenase, glucose-6-phosphate dehydrogenase, glycerol-3-phosphate dehydrogenase and fructose-1,6-bisphosphate aldolase. The selectivities and affinities of these proteins for the oxometalates are discussed.  相似文献   

7.
The active site loop of triosephosphate isomerase (TIM) exhibits a hinged-lid motion, alternating between the two well defined "open" and "closed" conformations. Until now the closed conformation had only been observed in protein complexes with substrate analogues. Here, we present the first rabbit muscle apo TIM structure, refined to 1.5A resolution, in which the active site loop is either in the open or in the closed conformation in different subunits of the enzyme. In the closed conformation described here, the lid loop residues participate in stabilizing hydrogen bonds characteristic of holo TIM structures, whereas chemical interactions observed in the open loop conformation are similar to those found in the apo structures of TIM. In the closed conformation, a number of water molecules are observed at the projected ligand atom positions that are hydrogen bonded to the active site residues. Additives used during crystallization (DMSO and Tris molecules and magnesium atoms) were modeled in the electron density maps. However, no specific binding of these molecules is observed at, or close to, the active site and the lid loop. To further investigate this unusual closed conformation of the apo enzyme, two more rabbit muscle TIM structures, one in the same and another in a different crystal form, were determined. These structures present the open lid conformation only, indicating that the closed conformation cannot be explained by crystal contact effects. To rationalize why the active site loop is closed in the absence of ligand in one of the subunits, extensive comparison with previously solved TIM structures was carried out, supported by the bulk of available experimental information about enzyme kinetics and reaction mechanism of TIM. The observation of both open and closed lid conformations in TIM crystals might be related to a persistent conformational heterogeneity of this protein in solution.  相似文献   

8.
O S Smart  J Akins  D M Blow 《Proteins》1992,13(2):100-111
A proposed reaction mechanism for the enzyme D-xylose isomerase involves the ring opening of the cyclic substrate with a subsequent conformational rearrangement to an extended open-chain form. Restrained energy minimization was used to simulate the rearrangement. In the ring-opening step, the substrate energy function was gradually altered from a cyclic to an open-chain form, with energy minimization after each change. The protein/sugar contact energy did not increase significantly during the process, showing that there was no steric hindrance to ring opening. The conformational rearrangement involves an alteration in the coordination of the substrate to metal ion [1], which was induced by gradually changing restraints on metal/ligand distances. By allowing varying amounts of flexibility in the protein and examining a simplified model system, the interactions of the sugar with metal ion [1] and its immediate ligands were found to be the most important contributors to the energy barrier for the change. Only small changes in the positions of protein atoms were required. The energy barrier to the rearrangement was estimated to be less than the Arrhenius activation energy for the enzymatic reaction. This is in accordance with experimental indications that the isomerization step is rate determining.  相似文献   

9.
A gene encoding galactose 1-phosphate uridylyltransferase (GalT) was identified in the hyperthermophilic archaeon Pyrobaculum aerophilum. The gene was overexpressed in Escherichia coli, after which its product was purified and characterized. The expressed enzyme was highly thermostable and retained about 90% of its activity after incubation for 10 minutes at temperatures up to 90°C. Two different crystal structures of P. aerophilum GalT were determined: the substrate-free enzyme at 2.33 Å and the UDP-bound H140F mutant enzyme at 1.78 Å. The main-chain coordinates of the P. aerophilum GalT monomer were similar to those in the structures of the E. coli and human GalTs, as was the dimeric arrangement. However, there was a striking topological difference between P. aerophilum GalT and the other two enzymes. In the E. coli and human enzymes, the N-terminal chain extends from one subunit into the other and forms part of the substrate-binding pocket in the neighboring subunit. By contrast, the N-terminal chain in P. aerophilum GalT extends to the substrate-binding site in the same subunit. Amino acid sequence alignment showed that a shorter surface loop in the N-terminal region contributes to the unique topology of P. aerophilum GalT. Structural comparison of the substrate-free enzyme with UDP-bound H140F suggests that binding of the glucose moiety of the substrate, but not the UDP moiety, gives rise to a large structural change around the active site. This may in turn provide an appropriate environment for the enzyme reaction.  相似文献   

10.
11.
BACKGROUND: Holo-(acyl carrier protein) synthase (AcpS), a member of the phosphopantetheinyl transferase superfamily, plays a crucial role in the functional activation of acyl carrier protein (ACP) in the fatty acid biosynthesis pathway. AcpS catalyzes the attachment of the 4'-phosphopantetheinyl moiety of coenzyme A (CoA) to the sidechain of a conserved serine residue on apo-ACP. RESULTS: We describe here the first crystal structure of a type II ACP from Bacillus subtilis in complex with its activator AcpS at 2.3 A. We also have determined the structures of AcpS alone (at 1.8 A) and AcpS in complex with CoA (at 1.5 A). These structures reveal that AcpS exists as a trimer. A catalytic center is located at each of the solvent-exposed interfaces between AcpS molecules. Site-directed mutagenesis studies confirm the importance of trimer formation in AcpS activity. CONCLUSIONS: The active site in AcpS is only formed when two AcpS molecules dimerize. The addition of a third molecule allows for the formation of two additional active sites and also permits a large hydrophobic surface from each molecule of AcpS to be buried in the trimer. The mutations Ile5-->Arg, Gln113-->Glu and Gln113-->Arg show that AcpS is inactive when unable to form a trimer. The co-crystal structures of AcpS-CoA and AcpS-ACP allow us to propose a catalytic mechanism for this class of 4'-phosphopantetheinyl transferases.  相似文献   

12.
The prevalence of glucose-6-phosphate dehydrogenase (G6PD) deficiency and its gene mutations were studied in the Achang population from Lianghe County in Southwestern China. We found that 7.31% (19 of 260) males and 4.35% (10 of 230) females had G6PD deficiency. The molecular analysis of G6PD gene exons 2–13 was performed by a PCR-DHPLC-Sequencing or PCR-Sequencing. Sixteen independent subjects with G6PD Mahidol (487G>A) and the new polymorphism IVS5-612 (G>C), which combined into a novel haplotype, were identified accounting for 84.2% (16/19). And 100% Achang G6PD Mahidol were linked to the IVS5-612 C. The percentage of G6PD Mahidol in the Achang group is close to that in the Myanmar population (91.3% 73/80), which implies that there are some gene flows between Achang and Myanmar populations. Interestingly, G6PD Canton (1376G>T) and G6PD Kaiping (1388G>A), which were the most common G6PD variants from other ethnic groups in China, were not found in this Achang group, suggesting that there are different G6PD mutation profiles in the Achang group and other ethnic groups in China. Our findings appear to be the first documented report on the G6PD genetics of the AChang people, which will provide important clues to the Achang ethnic group origin and will help prevention and treatment of malaria in this area.  相似文献   

13.
The prevalence of glucose-6-phosphate dehydrogenase (G6PD) deficiency and its gene mutations were studied in the Achang population from Lianghe County in Southwestern China. We found that 7.31% (19 of 260) males and 4.35% (10 of 230) females had G6PD deficiency. The molecular analysis of G6PD gene exons 2―13 was performed by a PCR-DHPLC-Sequencing or PCR-Sequencing. Sixteen inde-pendent subjects with G6PD Mahidol (487G>A) and the new polymorphism IVS5-612 (G>C), which combined into a novel haplotype, were identified accounting for 84.2% (16/19). And 100% Achang G6PD Mahidol were linked to the IVS5-612 C. The percentage of G6PD Mahidol in the Achang group is close to that in the Myanmar population (91.3% 73/80), which implies that there are some gene flows between Achang and Myanmar populations. Interestingly, G6PD Canton (1376G>T) and G6PD Kaiping (1388G>A), which were the most common G6PD variants from other ethnic groups in China, were not found in this Achang group, suggesting that there are different G6PD mutation profiles in the Achang group and other ethnic groups in China. Our findings appear to be the first documented report on the G6PD genetics of the AChang people, which will provide important clues to the Achang ethnic group origin and will help prevention and treatment of malaria in this area.  相似文献   

14.
This paper documents the maximal activities of the glycolytic enzymes in the red blood cells of normal mice and mice infected with Plasmodium berghei. There appears to be sufficient parasite-related activity of each glycolytic enzyme to support the increased glycolytic rate, i.e., increased glucose consumption, of the parasite-infected red blood cell. The relative proportions of glycolytic enzyme activities in parasite-infected red cells are different from the proportions in either normal or reticulocyte-rich blood, indicating that the increased enzyme activities associated with infected cells are not due to contaminating host red cells or reticulocytes. A comparison of maximal enzyme activities to the rate of whole cell glucose consumption indicates that different glycolytic control mechanisms are operating in the infected RBC from those in the uninfected cells.  相似文献   

15.
Transient receptor potential melastatin-1 channel (TRPM1) is an important mediator of calcium influx into the cell that is expressed in melanoma and ON-bipolar cells. Similar to other members of the TRP channel family, the intracellular N- and C- terminal domains of TRPM1 are expected to play important roles in the modulation of TRPM1 receptor function. Among the most commonly occurring modulators of TRP channels are the cytoplasmically expressed calcium binding proteins calmodulin and S100 calcium-binding protein A1 (S100A1), but the interaction of TRPM1 with S100A1 has not been described yet. Here, using a combination of biophysical and bioinformatics methods, we have determined that the N-terminal L242-E344 region of TRPM1 is a S100A1 binding domain. We show that formation of the TRPM1/S100A1 complex is calcium-dependent. Moreover, our structural model of the complex explained data obtained from fluorescence spectroscopy measurements revealing that the complex formation is facilitated through interactions of clusters positively charged (K271A, R273A, R274A) and hydrophobic (L263A, V270A, L276A) residues at the N-terminus of TRPM1. Taken together, our data suggest a molecular mechanism for the potential regulation of TRPM1 by S100A1.  相似文献   

16.
17.
A K Barnes  C H Wynn 《Proteins》1988,4(3):182-189
Recently developed computer programs, including secondary structure and epitopic site predictions, have been used to align lysosomal proteins for maximum homology, based on conservative interchanges, and the aligned sequences have been searched for potential sites for posttranslational modification, glycosylation, and binding and catalysis of substrate. The homology and prediction of the posttranslational modification of the alpha- and beta-subunits of hexosaminidase is in good agreement with previous observations, and an explanation of the differing substrate specificities of the two subunits is advanced. We show that the striking homology between alpha-glucosidase and isomaltase is reflected in the apparent conservation of the active site in both enzymes. Nonhomologous regions have been examined in detail in a search for binding sites for glycogen and maltose, and two such sites have been tentatively identified. A highly redundant consensus sequence for the phosphorylation of mannose in lysosomal proteins, YXX(Y, W, or F), is suggested.  相似文献   

18.
Glucose-6-phosphate isomerase (GPI) is a glycolytic enzyme with ill-defined oligomeric state. In order to obtain insight into the correlation between oligomerization and the catalytic function of this enzyme, the crystal structure of GPI from the extreme thermophile Thermus thermophilus HB8 (TtGPI) has been determined at 1.95 Å resolution. The crystallographic asymmetric unit contains an apparent dimer. The core fold of protomer and the interprotomer spatial arrangement of the dimer are similar to those of already reported crystal structures of other GPIs. The active site is located on the dimer interface, and putative catalytic residues are well conserved among the GPIs. These results suggest that the observed dimeric state of TtGPI in the crystal is biologically relevant and that this enzyme uses a common catalytic mechanism for the isomerase reaction. Gel-filtration chromatography, chemical cross-linking, sedimentation equilibrium by analytical ultracentrifugation, and dynamic light-scattering experiments indicate that TtGPI exists in a dynamic equilibrium between monomeric and dimeric states in solution. Several factors potentially contributing to the thermal stability of TtGPI protomer were identified: (i) a decrease in denaturation entropy by the shorter polypeptide length and by amino acid composition, including the increased number of proline residues and a higher arginine-to-lysine ratio; (ii) a larger number of ion pairs; and (iii) a reduction in cavity volume. From these results, it is suggested that transient dimer formation is sufficient for the catalytic function and that the TtGPI protomer itself has intrinsically higher thermal stability.  相似文献   

19.

Background

Fructose-1,6-bisphosphatase, a major enzyme of gluconeogenesis, is inhibited by AMP, Fru-2,6-P2 and by high concentrations of its substrate Fru-1,6-P2. The mechanism that produces substrate inhibition continues to be obscure.

Methods

Four types of experiments were used to shed light on this: (1) kinetic measurements over a very wide range of substrate concentrations, subjected to detailed statistical analysis; (2) fluorescence studies of mutants in which phenylalanine residues were replaced by tryptophan; (3) effect of Fru-2,6-P2 and Fru-1,6-P2 on the exchange of subunits between wild-type and Glu-tagged oligomers; and (4) kinetic studies of hybrid forms of the enzyme containing subunits mutated at the active site residue tyrosine-244.

Results

The kinetic experiments with the wild-type enzyme indicate that the binding of Fru-1,6-P2 induces the appearance of catalytic sites with lower affinity for substrate and lower catalytic activity. Binding of substrate to the high-affinity sites, but not to the low-affinity sites, enhances the fluorescence emission of the Phe219Trp mutant; the inhibitor, Fru-2,6-P2, competes with the substrate for the high-affinity sites. Binding of substrate to the low-affinity sites acts as a “stapler” that prevents dissociation of the tetramer and hence exchange of subunits, and results in substrate inhibition.

Conclusions

Binding of the first substrate molecule, in one dimer of the enzyme, produces a conformational change at the other dimer, reducing the substrate affinity and catalytic activity of its subunits.

General significance

Mimics of the substrate inhibition of fructose-1,6-bisphosphatase may provide a future option for combatting both postprandial and fasting hyperglycemia.  相似文献   

20.
Staphylococcus aureus is a significant human pathogen. Among its large repertoire of secreted toxins is a group of staphylococcal superantigen-like proteins (SSLs). These are homologous to superantigens but do not have the same activity. SSL5 is shown here to bind to human granulocytes and to the cell surface receptors for human IgA (FcαRI) and P-selectin [P-selectin glycoprotein ligand-1 (PSGL-1)] in a sialic acid (Sia)-dependent manner. Co-crystallization of SSL5 with the tetrasaccharide sialyl Lewis X (sLeX), a key determinant of PSGL-1 binding to P-selectin, led to crystal structures of the SSL5–sLeX complex at resolutions of 1.65 and 2.75 Å for crystals at two pH values. In both structures, sLeX bound to a specific site on the surface of the C-terminal domain of SSL5 in a conformation identical with that bound by P-selectin. Conservation of the key carbohydrate binding residues indicates that this ability to bind human glycans is shared by a substantial subgroup of the SSLs, including SSL2, SSL3, SSL4, SSL5, SSL6, and SSL11. This indicates that the ability to target human glycans is an important property of this group of toxins. Structural comparisons also showed that the Sia binding site in SSL5 contains a substructure that is shared by other Sia binding proteins from bacteria as well as viruses and represents a common binding motif.  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