首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 515 毫秒
1.
Escherichia coli inorganic pyrophosphatase (PPase) is a hexamer of identical subunits. This work shows that trimeric form of PPase exhibits the interaction of the active sites in catalysis. Some trimer subunits demonstrate high substrate binding affinity typical for hexamer whereas the rest of subunits reveal more than 300-fold substrate affinity decrease. This fact indicates the appearance of negative cooperativity of trimer subunits upon substrate binding. Association of the wild-type (WT) trimer with catalytically inactive, but still substrate binding mutant trimer into hexameric chimera restores the high activity of the first trimer, characteristic of trimer incorporated in the hexamer of WT PPase. Interaction of PPase active sites suggests that there are pathways for information transmission between the active sites, providing the perfect organization and concerted functioning of the hexameric active sites in catalysis.  相似文献   

2.
Kuo SY  Pan RL 《Plant physiology》1990,93(3):1128-1133
Tonoplast membrane of etiolated mung bean (Vinga radiata. L.) seedlings contained H+-translocating pyrophosphatase (PPase). Modification of tonoplast vesicles and partially purified PPase from etiolated mung bean seedlings with arginine-specific reagents, phenylglyoxal (PGO) and 2,3-butanedione (BD), resulted in a marked decline in H+-translocating PPase activity. The half-maximal inhibition was brought about by 20 millimolar PGO and 50 millimolar BD for membrane bound and 1.5 millimolar PGO and 5.0 millimolar BD for soluble PPase, respectively. The substrate, Mg2+-pyrophosphate, provided partial protection against inactivation by these reagents. Loss of activity of partially purified PPase followed pseudo-first order kinetics. The double logarithm plots of pseudo-first order rate constant versus reagent concentrations gave slopes of 0.88 (PGO) and 0.90 (BD), respectively, suggesting that the inactivation may possibly result from reaction of at least one arginyl residue at the active site of H+-translocating PPase.  相似文献   

3.
Soluble inorganic pyrophosphatases (PPases) comprise two evolutionarily unrelated families (I and II). These two families have different specificities for metal cofactors, which is thought to be because of the fact that family II PPases have three active site histidines, whereas family I PPases have none. Here, we report the structural and functional characterization of a unique family I PPase from Mycobacterium tuberculosis (mtPPase) that has two His residues (His21 and His86) in the active site. The 1.3-A three-dimensional structure of mtPPase shows that His86 directly interacts with bound sulfate, which mimics the product phosphate. Otherwise, mtPPase is structurally very similar to the well studied family I hexameric PPase from Escherichia coli, although mtPPase lacks the intersubunit metal binding site found in E. coli PPase. The cofactor specificity of mtPPase resembles that of E. coli PPase in that it has high activity in the presence of Mg2+, but it differs from the E. coli enzyme and family II PPases because it has much lower activity in the presence of Mn2+ or Zn2+. Replacements of His21 and His86 in mtPPase with the residues found in the corresponding positions of E. coli PPase had either no effect on the Mg2+- and Mn2+-supported reactions (H86K) or reduced Mg2+-supported activity (H21K). However, both replacements markedly increased the Zn2+-supported activity of mtPPase (up to 11-fold). In the double mutant, Zn2+ was a 2.5-fold better cofactor than Mg2+. These results show that the His residues in mtPPase are not essential for catalysis, although they determine cofactor specificity.  相似文献   

4.
Earlier it has been demonstrated that inactivation of inorganic pyrophosphatase (PPase) of S. cerevisiae by 7-chloro-4-nitronbenzofurasane is due to modification of Tyr89. The effect of pH and active center ligands on this reaction has been studied. It was found that pK for Tyr89 does not exceed 8.5; the phosphate-metal complex binding to the high affinity center protects Tyr89 from inactivation. Activating ions (Mg2+ and Zn2+) do not influence the inactivation, whereas the PPase inhibitor, Ca2+, enhances this process after saturation of the high affinity binding site. Saturation of two binding sites with Ca2+ has a protective effect on the enzyme. An increase in the rate of Tyr89 binding to the inhibitor in the presence of low concentrations of Ca2+ is due to the decrease of Tyr89 pK. The data obtained suggest that Tyr89 is located near the high affinity binding site for phosphate. The high reactivity of Tyr89 and its tight binding in the active center point to the presence of a hydrogen bondage with the substrate and suggest a role of a proton donor whose acceptor is the product of the enzymatic reaction, i.e., phosphate.  相似文献   

5.
Inorganic pyrophosphatase from Mycobacterium tuberculosis (Mt-PPase) is one of the possible targets for the rational design of anti-tuberculosis agents. In this paper, functional properties of this enzyme are characterized in the presence of the most effective activators--Mg2+ and Mn2+. Dissociation constants of Mt-PPase complexed with Mg2+ or Mn2+ are essentially similar to those of Escherichia coli PPase. Stability of a hexameric form of Mt-PPase has been characterized as a function of pH both for the metal-free enzyme and for Mg2+- or Mn2+-enzyme. Hexameric metal-free Mt-PPase has been shown to dissociate, forming monomers at pH below 4 or trimers at pH from 8 to 10. Mg2+ or Mn2+ shift the hexamer-trimer equilibrium found for the apo-Mt-PPase at pH 8-10 toward the hexameric form by stabilizing intertrimeric contacts. The pK(a) values have been determined for groups that control the observed hexamer-monomer (pK(a) 5.4), hexamer-trimer (pK(a) 7.5), and trimer-monomer (pK(a) 9.8) transitions. Our results demonstrate that due to the non-conservative amino acid residues His21 and His86 in the active site of Mt-PPase, substrate specificity of this enzyme, in contrast to other typical PPases, does not depend on the nature of the metal cofactor.  相似文献   

6.
Family II pyrophosphatases (PPases), recently found in bacteria and archaebacteria, are Mn(2+)-containing metalloenzymes with two metal-binding subsites (M1 and M2) in the active site. These PPases can use a number of other divalent metal ions as the cofactor but are inactive with Zn(2+), which is known to be a good cofactor for family I PPases. We report here that the Mg(2+)-bound form of the family II PPase from Streptococcus gordonii is nearly instantly activated by incubation with equimolar Zn(2+), but the activity thereafter decays on a time scale of minutes. The activation of the Mn(2+)-form by Zn(2+) was slower but persisted for hours, whereas activation was not observed with the Ca(2+)- and apo-forms. The bound Zn(2+) could be removed from PPase by prolonged EDTA treatment, with a complete recovery of activity. On the basis of the effect of Zn(2+) on PPase dimerization, the Zn(2+) binding constant appeared to be as low as 10(-12) M for S. gordonii PPase. Similar effects of Zn(2+) and EDTA were observed with the Mg(2+)- and apo-forms of Streptococcus mutans and Bacillus subtilis PPases. The effects of Zn(2+) on the apo- and Mg(2+)-forms of HQ97 and DE15 B. subtilis PPase variants (modified M2 subsite) but not of HQ9 variant (modified M1 subsite) were similar to that for the Mn(2+)-form of wild-type PPase. These findings can be explained by assuming that (a) the PPase tightly binds Mg(2+) and Mn(2+) at the M2 subsite; (b) the activation of the corresponding holoenzymes by Zn(2+) results from its binding to the M1 subsite; and (c) the subsequent inactivation of Mg(2+)-PPase results from Zn(2+) migration to the M2 subsite. The inability of Zn(2+) to activate apo-PPase suggests that Zn(2+) binds more tightly to M2 than to M1, allowing direct binding to M2. Zn(2+) is thus an efficient cofactor at subsite M1 but not at subsite M2.  相似文献   

7.
Mature nonstructural protein-15 (nsp15) from the severe acute respiratory syndrome coronavirus (SARS-CoV) contains a novel uridylate-specific Mn2+-dependent endoribonuclease (NendoU). Structure studies of the full-length form of the obligate hexameric enzyme from two CoVs, SARS-CoV and murine hepatitis virus, and its monomeric homologue, XendoU from Xenopus laevis, combined with mutagenesis studies have implicated several residues in enzymatic activity and the N-terminal domain as the major determinant of hexamerization. However, the tight link between hexamerization and enzyme activity in NendoUs has remained an enigma. Here, we report the structure of a trimmed, monomeric form of SARS-CoV nsp15 (residues 28 to 335) determined to a resolution of 2.9 A. The catalytic loop (residues 234 to 249) with its two reactive histidines (His 234 and His 249) is dramatically flipped by approximately 120 degrees into the active site cleft. Furthermore, the catalytic nucleophile Lys 289 points in a diametrically opposite direction, a consequence of an outward displacement of the supporting loop (residues 276 to 295). In the full-length hexameric forms, these two loops are packed against each other and are stabilized by intimate intersubunit interactions. Our results support the hypothesis that absence of an adjacent monomer due to deletion of the hexamerization domain is the most likely cause for disruption of the active site, offering a structural basis for why only the hexameric form of this enzyme is active.  相似文献   

8.
Replisome DNA primases are responsible for the synthesis of short RNA primers required for the initiation of repetitive Okazaki fragment synthesis on the lagging strand during DNA replication. In bacteriophage T4, the primase (gp61) interacts with the helicase (gp41) to form the primosome complex, an interaction that greatly stimulates the priming activity of gp61. Because gp41 is hexameric, a question arises as to whether gp61 also forms a hexameric structure during replication. Several results from this study support such a structure. Titration of the primase/single-stranded DNA binding followed by fluorescence anisotropy implicated a 6:1 stoichiometry. The observed rate constant, k(cat), for priming was found to increase with the primase concentration, implicating an oligomeric form of the primase as the major functional species. The generation of hetero-oligomeric populations of the hexameric primase by controlled mixing of wild type and an inactive mutant primase confirmed the oligomeric nature of the most active primase form. Mutant primases defective in either the N- or C-terminal domains and catalytically inactive could be mixed to create oligomeric primases with restored catalytic activity suggesting an active site shared between subunits. Collectively, these results provide strong evidence for the functional oligomerization of gp61. The potential roles of gp61 oligomerization during lagging strand synthesis are discussed.  相似文献   

9.
The amino-terminal cysteine of glucosamine-6-phosphate synthase (GlmS) acts as a nucleophile to release and transfer ammonia from glutamine to fructose 6-phosphate through a channel. The crystal structure of the C1A mutant of Escherichia coli GlmS, solved at 2.5 Å resolution, is organized as a hexamer, where the glutaminase domains adopt an inactive conformation. Although the wild-type enzyme is active as a dimer, size exclusion chromatography, dynamic and quasi-elastic light scattering, native polyacrylamide gel electrophoresis, and ultracentrifugation data show that the dimer is in equilibrium with a hexameric state, in vitro and in cellulo. The previously determined structures of the wild-type enzyme, alone or in complex with glucosamine 6-phosphate, are also consistent with a hexameric assembly that is catalytically inactive because the ammonia channel is not formed. The shift of the equilibrium toward the hexameric form in the presence of cyclic glucosamine 6-phosphate, together with the decrease of the specific activity with increasing enzyme concentration, strongly supports product inhibition through hexamer stabilization. Altogether, our data allow us to propose a morpheein model, in which the active dimer can rearrange into a transiently stable form, which has the propensity to form an inactive hexamer. This would account for a physiologically relevant allosteric regulation of E. coli GlmS. Finally, in addition to cyclic glucose 6-phosphate bound at the active site, the hexameric organization of E. coli GlmS enables the binding of another linear sugar molecule. Targeting this sugar-binding site to stabilize the inactive hexameric state is therefore suggested for the development of specific antibacterial inhibitors.  相似文献   

10.
11.
Modification of Saccharomyces cerevisiae inorganic pyrophosphatase (PPase) with 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide is known to lead to a loss of enzymatic activity, the rate of which is decreased in the presence of ligands binding to the active site [Cooperman, B. S., & Chiu, N. Y. (1973) Biochemistry 12, 1676-1682; Heitman, P., & Uhlig, H. J. (1974) Acta Biol. Med. Ger. 32, 565-594]. In this work we show that, when such inactivation is carried out in the presence of [14C]glycine ethyl ester (GEE), GEE is covalently incorporated into PPase, incorporation into the most highly labeled tryptic peptide is site-specific, as evidenced by the reduction of such incorporation in the presence of the active site ligands Zn2+ and Pi, the extent of formation of this specifically labeled peptide correlates with the fractional loss of PPase activity, and the specifically labeled peptide corresponds to residues 145-153 and the position of incorporation within this peptide is Glu-149. The significance of our findings for the location of the active site and for the catalytic mechanism of PPase is briefly considered in the light of the 3-A X-ray crystallographic structure of Arutyunyun and his colleagues [Arutyunyun, E. G., et al. (1981) Dokl. Akad. Nauk SSSR 258, 1481-1485; Kuranova, I. P., et al. (1983) Bioorg. Khim. 9, 1611-1919; Terzyan, S. S., et al. (1984) Bioorg. Khim. 10, 1469-1482].  相似文献   

12.
On helicases and other motor proteins   总被引:1,自引:0,他引:1  
Helicases are molecular machines that utilize energy derived from ATP hydrolysis to move along nucleic acids and to separate base-paired nucleotides. The movement of the helicase can also be described as a stationary helicase that pumps nucleic acid. Recent structural data for the hexameric E1 helicase of papillomavirus in complex with single-stranded DNA and MgADP has provided a detailed atomic and mechanistic picture of its ATP-driven DNA translocation. The structural and mechanistic features of this helicase are compared with the hexameric helicase prototypes T7gp4 and SV40 T-antigen. The ATP-binding site architectures of these proteins are structurally similar to the sites of other prototypical ATP-driven motors such as F1-ATPase, suggesting related roles for the individual site residues in the ATPase activity.  相似文献   

13.
Green crab (Scylla Serrata) alkaline phosphatase (EC 3.1.3.1.) is a metalloenzyme, the each active site in which contains a tight cluster of two zinc ions and one magnesium ion. The kinetic theory of the substrate reaction during irreversible inhibition of enzyme activity previously described by Tsou has been applied to a study on the kinetics of the course of inactivation of the enzyme by ethylenediaminetetraacetic acid disodium (EDTA). The kinetics of the substrate reaction with different concentrations of the substrate p-nitrophenyl phosphate (PNPP) and inactivator EDTA suggested a complexing mechanism for inactivation by, and substrate competition with, EDTA at the active site. The inactivation kinetics are single phasic, showing the initial formation of an enzyme-EDTA complex is a relatively rapid reaction, followed a slow inactivation step that probably involves a conformational change of the enzyme. Zinc ions are finally removed from the enzyme. The presence of metal ions apparently stabilizes an active-site conformation required for enzyme activity.  相似文献   

14.
The alignments of the amino acid sequences of inorganic pyrophosphatase (PPase) from Saccharomyces cerevisiae (Y1-PPase, 286 amino acids) and Escherichia coli (E-PPase, 175 amino acids) are examined in the light of crystallographic and chemical modification results placing specific amino acid residues at the active site of the yeast enzyme. The major results are: (1) the full E-PPase sequence aligns within residues 28-225 of Y1-PPase, raising the possibility that the N-terminal and C-terminal portions of Y1-PPase may not be essential for activity, and (2) that whereas the overall identity between the two sequences is only modest (22-27% depending on the choice of alignment parameters), of some 17 putative active site residues, 14-16 are identical between Y-PPase and E-PPase. PPase thus appears to be an example of enzymes from widely divergent species that conserve common functional elements within the context of substantial overall sequence variation.  相似文献   

15.

Background

Plants, fungi, bacteria and the apicomplexan parasite Plasmodium falciparum are able to synthesize vitamin B6 de novo, whereas mammals depend upon the uptake of this essential nutrient from their diet. The active form of vitamin B6 is pyridoxal 5-phosphate (PLP). For its synthesis two enzymes, Pdx1 and Pdx2, act together, forming a multimeric complex consisting of 12 Pdx1 and 12 Pdx2 protomers.

Methodology/Principal Findings

Here we report amino acid residues responsible for stabilization of the structural and enzymatic integrity of the plasmodial PLP synthase, identified by using distinct mutational analysis and biochemical approaches. Residues R85, H88 and E91 (RHE) are located at the Pdx1:Pdx1 interface and play an important role in Pdx1 complex assembly. Mutation of these residues to alanine impedes both Pdx1 activity and Pdx2 binding. Furthermore, changing D26, K83 and K151 (DKK), amino acids from the active site of Pdx1, to alanine obstructs not only enzyme activity but also formation of the complex. In contrast to the monomeric appearance of the RHE mutant, alteration of the DKK residues results in a hexameric assembly, and does not affect Pdx2 binding or its activity. While the modelled position of K151 is distal to the Pdx1:Pdx1 interface, it affects the assembly of hexameric Pdx1 into a functional dodecamer, which is crucial for PLP synthesis.

Conclusions/Significance

Taken together, our data suggest that the assembly of a functional Pdx1:Pdx2 complex follows a defined pathway and that inhibition of this assembly results in an inactive holoenzyme.  相似文献   

16.
(Na+ + K+)-ATPase from beef brain and pig kidney are slowly inactivated by chromium(III) complexes of nucleotide triphosphates in the absence of added univalent and divalent cations. The inactivation of (Na+ + K+)-ATPase activity was accompanied by a parallel decrease of the associated K+-activated p-nitrophenylphosphatase and a parallel loss of the capacity to form, Na+-dependently, a phosphointermediate from [gamma-32P]ATP. The kinetics of inactivation and of phosphorylation with [gamma-32P]CrATP and [alpha-32P]CrATP are consistent with the assumption of the formation of a dissociable complex of CrATP with the enzyme (E) followed by phosphorylation of the enzyme: formula: (see text). The dissociation constant of the CrATP complex of the pig kidney enzyme at 37 degrees C was 43 microM. The inactivation rate constant (k + 2 = 0.033 min-1) was in the range of the dissociation rate constant kd of ADP from the enzyme of 0.011 min-1. The phosphoenzyme was unreactive towards ADP as well as to K+. No hydrolysis of the native isolated phosphoenzyme was observed within 6 h under a variety of conditions, but high concentrations of Na+ reactivated it slowly. The capacity of the Cr-phosphoenzyme of 121 +/- 18 pmol/unit enzyme is identical with the capacity of the unmodified enzyme to form, Na+-dependently, a phosphointermediate. The Cr-phosphoenzyme behaved after acid denaturation like an acylphosphate towards hydroxylamine, but the native phosphoenzyme was not affected by it. ATP protected the enzyme against the inactivation by CrATP (dissociation constant of the enzyme ATP complex = 2.5 microM) as well as low concentrations of K+. CrATP was a competitive inhibitor of (Na+ + K+)-ATPase. It is concluded that CrATP is slowly hydrolyzed at the ATP-binding site of (Na+ + K+)-ATPase and inactivates the enzyme by forming an almost non-reactive phosphoprotein at the site otherwise needed for the Na+-dependent proteinkinase reaction as the phosphate acceptor site.  相似文献   

17.
A phosphoprotein phosphatase (PPase: EC 3.1.3.2) was recently purified from rat epidermis. The enzyme dephosphorylates phosphoprotein, and its properties, such as pH optimum, inhibitor spectrum, and Fe2+ activation, differ from those of other soluble phosphatases. We investigated in 2-day-old rat skin the distribution of immunologically detectable PPase and intracellular localization of PPase activity. The reaction of rabbit monospecific anti-PPase IgG was identified in granular and cornified cells by the avidin-biotin complex method. For activity staining, basic principles of the Gomori lead-salt method and azo dye technique with the substrates p-nitrophenylphosphate (p-NPP) and alpha-naphthyl phosphate (NP), respectively, were modified according to the biochemical properties of PPase activity which is resistant to formalin, Na tartrate, and NaF. Activity was detectable in granular cells including keratohyalin granules and the lower strata of cornified cells. The activity was inhibited by 1 mM CuSO4 and enhanced by a mixture of 0.5 mM FeSO4 and 1 mM ascorbic acid. We consider that PPase may be involved in dephosphorylation of histidine-rich proteins in granular and cornified cells and may play a key role in intracellular catabolism associated with epidermal cell differentiation.  相似文献   

18.
The acetyl-CoA:acetoacetate CoA-transferase of Escherichia coli was reversibly inactivated by pyridoxal 5′-phosphate. The residual activity of the enzyme was dependent on the concentration of the modifying reagent to a concentration of 5 mm. The maximum level of inactivation was 89%. Kinetic and equilibrium analyses of inactivation were consistent with a two-step process (Chen and Engel, 1975, Biochem. J.149, 619) in which the extent of inactivation was limited by the ratio of first-order rate constants for the reversible formation of an inactive Schiff base of pyridoxal 5′-phosphate and the enzyme from a noncovalent, dissociable complex of the enzyme and modifier. The calculated minimum residual activity was in close agreement with the experimentally determined value. The conclusion that the loss of catalytic activity resulted from modification of a lysine residue at the active site was based on the following data, (a) After incubation with 5 mm pyridoxal 5′-phosphate, 3.95 mol of the reagent was incorporated per mole of free enzyme with 89% loss of activity, while 2.75 mol of pyridoxal 5′-phosphate was incorporated into the enzyme-CoA intermediate with a loss of 10% of catalytic activity; the intermediate was formed in the presence of acetoacetyl-CoA; (b) acid hydrolysis of the modified, reduced enzyme-CoA intermediate yielded a single fluorescent compound that was identified as N6-pyridoxyllysine by chromatography in two solvent systems; (c) the enzyme was also protected from inactivation by saturating concentrations of free CoA and ADP but not by adenosine. The results suggested that a lysine residue is involved in the electrostatic binding of the pyrophosphate group of CoA. Carboxylic acid substrate did not protect the enzyme from inactivation.  相似文献   

19.
When ferric ion was added to solutions of the enzyme dextransucrase, first-order followed by second-order inactivation behavior was observed. The initial rapid activity loss was attributed to a ferric ion interacting with the thiol group of the native monomer to form a less active enzyme-ion complex; the second inactivation stage involved enzyme-ion complex aggregation and disulfide cross-link formation. In contrast, Cu2+ ion inactivation demonstrated simple first-order kinetics. As with Fe3+, Cu2+ ions can form complexes with enzyme thiol groups. However, unlike ferric ions, cupric ions can also strongly interact with the imidazole ring of histidine. Since the dextransucrase active site contains two key histidines, imidazole-cupric-ion interactions could potentially inhibit enzymatic activity. Thus, it was hypothesized that first-order Cu2+ inactivation kinetics involved the adsorption of this ion to the enzyme's activity site. The addition of a reducing agent such as dithiothreitol can inhibit the second enzyme aggregation stage by breaking disulfide cross-links but cannot restrict the initial formation of metal-enzyme complexes.  相似文献   

20.
The exosome is a 3' --> 5' exoribonuclease complex involved in RNA processing. We report the crystal structure of the RNase PH core complex of the Sulfolobus solfataricus exosome determined at a resolution of 2.8 A. The structure reveals a hexameric ring-like arrangement of three Rrp41-Rrp42 heterodimers, where both subunits adopt the RNase PH fold common to phosphorolytic exoribonucleases. Structure-guided mutagenesis reveals that the activity of the complex resides within the active sites of the Rrp41 subunits, all three of which face the same side of the hexameric structure. The Rrp42 subunit is inactive but contributes to the structuring of the Rrp41 active site. The high sequence similarity of this archaeal exosome to eukaryotic exosomes and its high structural similarity to the bacterial mRNA-degrading PNPase support a common basis for RNA-degrading machineries in all three domains of life.  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-23

京公网安备 11010802026262号