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 31 毫秒
1.
The rapid expansion of the amount of genomic and structural data has provided many examples of enzymes with evolutionarily related active sites that catalyze different reactions. Functional comparisons of these active sites can provide insight into the origins of the enormous catalytic proficiency of enzymes and the evolutionary changes that can lead to different enzyme activities. The alkaline phosphatase (AP) superfamily is an ideal system to use in making such comparisons given the extensive data available on both nonenzymatic and enzymatic phosphoryl transfer reactions. Some superfamily members, such as AP itself, preferentially hydrolyze phosphate monoesters, whereas others, such as nucleotide pyrophosphatase/phosphodiesterase (NPP), preferentially hydrolyze phosphate diesters. We have measured rate constants for NPP-catalyzed hydrolysis of phosphate diesters and monoesters. NPP preferentially catalyzes diester hydrolysis by factors of 10(2)-10(6), depending on the identity of the diester substrate. To identify features of the NPP active site that could lead to preferential phosphate diester hydrolysis, we have determined the structure of NPP in the absence of ligands and in complexes with vanadate and AMP. Comparisons to existing structures of AP reveal bimetallo cores that are structurally indistinguishable, but there are several distinct structural features outside of the conserved bimetallo site. The structural and functional data together suggest that some of these distinct functional groups provide specific substrate binding interactions, whereas others tune the properties of the bimetallo active site itself to discriminate between phosphate diester and monoester substrates.  相似文献   

2.
O'Brien PJ  Herschlag D 《Biochemistry》2001,40(19):5691-5699
Escherichia coli alkaline phosphatase (AP) is a proficient phosphomonoesterase with two Zn(2+) ions in its active site. Sequence homology suggests a distant evolutionary relationship between AP and alkaline phosphodiesterase/nucleotide pyrophosphatase, with conservation of the catalytic metal ions. Furthermore, many other phosphodiesterases, although not evolutionarily related, have a similar active site configuration of divalent metal ions in their active sites. These observations led us to test whether AP could also catalyze the hydrolysis of phosphate diesters. The results described herein demonstrate that AP does have phosphodiesterase activity: the phosphatase and phosphodiesterase activities copurify over several steps; inorganic phosphate, a strong competitive inhibitor of AP, inhibits the phosphodiesterase and phosphatase activities with the same inhibition constant; a point mutation that weakens phosphate binding to AP correspondingly weakens phosphate inhibition of the phosphodiesterase activity; and mutation of active site residues substantially reduces both the mono- and diesterase activities. AP accelerates the rate of phosphate diester hydrolysis by 10(11)-fold relative to the rate of the uncatalyzed reaction [(k(cat)/K(m))/k(w)]. Although this rate enhancement is substantial, it is at least 10(6)-fold less than the rate enhancement for AP-catalyzed phosphate monoester hydrolysis. Mutational analysis suggests that common active site features contribute to hydrolysis of both phosphate monoesters and phosphate diesters. However, mutation of the active site arginine to serine, R166S, decreases the monoesterase activity but not the diesterase activity, suggesting that the interaction of this arginine with the nonbridging oxygen(s) of the phosphate monoester substrate provides a substantial amount of the preferential hydrolysis of phosphate monoesters. The observation of phosphodiesterase activity extends the previous observation that AP has a low level of sulfatase activity, further establishing the functional interrelationships among the sulfatases, phosphatases, and phosphodiesterases within the evolutionarily related AP superfamily. The catalytic promiscuity of AP could have facilitated divergent evolution via gene duplication by providing a selective advantage upon which natural selection could have acted.  相似文献   

3.
Comparisons among evolutionarily related enzymes offer opportunities to reveal how structural differences produce different catalytic activities. Two structurally related enzymes, Escherichia coli alkaline phosphatase (AP) and Xanthomonas axonopodis nucleotide pyrophosphatase/phosphodiesterase (NPP), have nearly identical binuclear Zn2+ catalytic centers but show tremendous differential specificity for hydrolysis of phosphate monoesters or phosphate diesters. To determine if there are differences in Zn2+ coordination in the two enzymes that might contribute to catalytic specificity, we analyzed both x-ray absorption spectroscopic and x-ray crystallographic data. We report a 1.29-Å crystal structure of AP with bound phosphate, allowing evaluation of interactions at the AP metal site with high resolution. To make systematic comparisons between AP and NPP, we measured zinc extended x-ray absorption fine structure for AP and NPP in the free-enzyme forms, with AMP and inorganic phosphate ground-state analogs and with vanadate transition-state analogs. These studies yielded average zinc–ligand distances in AP and NPP free-enzyme forms and ground-state analog forms that were identical within error, suggesting little difference in metal ion coordination among these forms. Upon binding of vanadate to both enzymes, small increases in average metal–ligand distances were observed, consistent with an increased coordination number. Slightly longer increases were observed in NPP relative to AP, which could arise from subtle rearrangements of the active site or differences in the geometry of the bound vanadyl species. Overall, the results suggest that the binuclear Zn2+ catalytic site remains very similar between AP and NPP during the course of a reaction cycle.  相似文献   

4.
Calcineurin (CN) is a Ca(2+)/calmodulin (CaM)-dependent protein serine/threonine phosphatase that contains Zn(2+) in its catalytic domain and can be stimulated by divalent ions such as Mn(2+) and Ni(2+). In this study, the role of exogenous Zn(2+) in the regulation of CN activity and its relevance to the role of Ni(2+) was investigated. Zn(2+) at a concentration range of 10nM-10 micro M inhibited Ni(2+)-stimulated CN-activity in vitro in a dose-dependent manner and approximately 50% inhibition was attained with 0.25 micro M Zn(2+). Kinetic analysis showed that Zn(2+) inhibited the activity of CN by competing with Ni(2+). Interaction of CN and CaM was not inhibited with Zn(2+) at 10 micro M. Zn(2+) never affected the activity of cAMP phosphodiesterase 1 or myosin light-chain kinase (CaM-dependent enzymes) and rather activated alkaline phosphatase. The present results indicate that Zn(2+) should be a potent inhibitor for CN activity although this ion is essential for CN.  相似文献   

5.
Quantifying the rate of phosphate monoester dianion hydrolysis under physiological conditions has implications for designing transition state mimics and understanding how catalysis is facilitated. Catalysis is energetically most efficient if the mechanistic pathway in solution is stabilised. Monoesters are believed to have a "dissociative" transition state that has little bonding to the nucleophile and leaving group. However, in many instances, it is suggested that enzymes catalyse monoester transfer through an associative (diester-like) pathway. This is perhaps easier to rationalise in terms of the active site residues available. For example, in the catalytic subunit of protein phosphatase 1 (PP1), these are metal ions and cationic side chains which might be expected to stabilise developing negative charge. By using multiple interactions simultaneously, cooperativity in catalysis may be achieved. However, this idea is difficult to demonstrate unambiguously in large, complex natural systems. This contribution examines the background reactivity of phosphate esters, and reports data showing that the substrates for serine/threonine phosphatases have slower intrinsic rate constants than any other enzyme substrates. Using model complexes, the characteristics of alternative (associative) mechanisms that have been proposed for the metallophosphatase catalysed reaction are explored. Finally, complementary catalytic groups are combined with this core complex to look for experimental evidence for possible cooperativity in this context.  相似文献   

6.
During the mating season the female mallards produce sex pheromones, diesters of 3-hydroxy fatty acids, in their uropygial glands. Subcellular fractionation by sucrose and Nycodenz density gradient centrifugations and electron microscopic examination of the fractions showed that diesters of 3-hydroxy acids and the enzymes that catalyze the formation and esterification of the 3-hydroxy fatty acids are located in the catalase-containing fractions, probably peroxisomes, whereas monoester synthesizing activities are located in the endoplasmic reticulum. Fatty acyl-CoA reductase that would provide fatty alcohol needed for the synthesis of monoester and diester waxes was found both in the peroxisomal and endoplasmic reticulum fraction. Upon daily intramuscular injection of estradiol into the females in the nonmating season, the short chain monoester waxes of the uropygial glands were replaced by long chain monoester waxes, and subsequently the monoester waxes were replaced by diester waxes. Injection of thyroxine with estradiol hastened the induction of the compositional changes including diester synthesis. Similar changes, including the synthesis of the female pheromones, were induced in the uropygial glands by the hormone treatment of males that do not normally produce diesters at any time during their life cycle. The structure and composition of the diesters induced by hormone treatment of both males and females were identical to those of the female pheromones produced during their mating season. Electron microscopic examination of diaminobenzidine-treated glands showed that peroxisomes proliferated in the gland of the females in the mating season and in the estradiol-treated males that produce the diesters.  相似文献   

7.
Haematococcus pluvialis cultivated in a N-deficient medium produced astaxanthin (1 % of total carotenoids), the monoester (76 %) and diester (7 %) of astaxanthin, β,β-carotene (1 %), an adonirubin ester (3 %), (3R,3′R,6′R)-lutein (7 %), violaxanthin (2 %) and neoxanthin (1 %). The CD values of the mono- and diesters of astaxanthin, the HPLC properties of astaxanthin monoester further esterified with (?)-camphanic acid and the optical purity of astaxanthin [determined by HPLC analysis of the diester of(?)-camphanic acid] produced by saponification of the natural mono- and diesters of astaxanthin in the absence of oxygen showed that this green alga synthesizes pure (3S,3′S)-astaxanthin esters.  相似文献   

8.
4-Nitrophenyl and 2-napthyl monoesters of phenylphosphonic acid have been synthesized, and an enzyme catalyzing their hydrolysis was resolved from alkaline phosphatase of a commerical calf intestinal alkaline phosphatase preparation by extensive ion-exchange chromatography, chromatography on L-phenylalanyl-Sepharose with a decreasing gradient of (NH4) 2SO4, and gel filtration. Detergent-solubilized enzyme from fresh bovine intestine was purified after (NH4)2SO4 fractionation by the same technique. The purified enzyme is homogeneous by polyacrylamide gel electrophoresis and sedimentation equilibrium centrifugation. It has a molecular weight of 108,000, contains approximately 21% carbohydrate, and has an amino acid composition considerably different from that reported from alkaline phosphatase from the same tissue. The homogeneous intestinal enzyme, an efficient catalyst of phosphonate ester hydoolysis but not of phosphate monoester hydrolysis, was identified as a 5'-nucleotide phosphodiesterase by its ability to hydrolyze 4-nitrophenyl esters of 5'-TMP but not of 3'-TMP. Also consistent with this identification was the ability of the enzyme to hydrolyze 5'-ATP to 5'-AMP and PPi, NAD+ to 5'-AMP and NMN, TpT to 5'-TMP and thymidine, pApApApA to 5'-AMP, and only the single-stranded portion of tRNA from the 3'-OH end. Snake venom 5'-nucleotide phosphodiesterase also hydrolyzes phosphonate esters, but 3'-nucleotide phosphodiesterase of spleen and cyclic 3',5'-AMP phosphodiesterase do not. Thus, types of phosphodiesterases can be conveniently distinguished by their ability to hydrolyze phosphonate esters. As substrates for 5'-nucleotide phosphodiesterases, phosphonate esters are preferable to the more conventional esters of nucleotides and bis(4-nitrophenyl) phosphate because of their superior stability and ease of synthesis. Furthermore, the rate of hydrolysis of phosphonate esters under saturating conditions is greater than that of the conventional substrates. At substrate concentrations of 1 mM the rates of hydrolysis of phosphonate esters and of nucleotide esters are comparable and both superior to that of bis(4-nitrophenyl) phosphate.  相似文献   

9.
Catalytic properties of alkaline phosphatase from pig kidney   总被引:6,自引:5,他引:1       下载免费PDF全文
The enzymic properties of alkaline phosphatase (EC 3.1.3.1) from pig kidney brush-border membranes were studied. 1. It hydrolyses ortho- and pyro-phosphate esters, the rate limiting step (V(max.)) being independent of the substrate. It transphosphorylates to Tris at concentrations above 0.1m-Tris. 2. The pH optimum for hydrolysis was between 9.8 and 10. The pK of the enzyme-substrate complex is 8.7 for p-nitrophenyl phosphate and beta-glycerophosphate. Excess of substrate inhibits the enzymic activity with decreasing pH. The pK of the substrate-inhibited enzyme-substrate complex, 8.7, is very similar to that for the enzyme-substrate complex. The pK values of the free enzyme appear to be 8.7 and 7.9. 3. Inactivation studies suggest that there is an essential tyrosine residue at the active centre of the enzyme. 4. The energy of activation (E) and the heat of activation (DeltaH) at pH9.5 showed a transition at 24.8 degrees C that was unaffected by Mg(2+). 5. Kinetic and atomic-absorption analysis indicated the essential role of two Zn(2+) ions/tetrameric enzyme for an ordered association of the monomers. Zn(2+) in excess and other bivalent ions compete for a second site with Mg(2+). Mg(2+) enhances only the rate-limiting step of substrate hydrolysis. 6. Amino acid inhibition studies classified the pig kidney enzyme as an intermediate type of previously described alkaline phosphatases. It has more similarity with the enzyme from liver and bone than with that from placenta.  相似文献   

10.
Sites of alkaline phosphatase activity have been located by an electron microscopic histochemical (Gomori) technique in vegetative cells of a repressible strain SB15 of Bacillus subtilis, derepressed and repressed by inorganic phosphate, and in a mutant SB1004 which forms alkaline phosphatase in a medium high in phosphate. The sites of enzyme activity were revealed as discrete, dense, and largely spherical bodies of varying sizes (20 to 150 nm). Cells of both repressible and repression-resistant strains acted on a wide variety of phosphate esters (p-nitrophenylphosphate, beta-glycerophosphate, adenosine-5'-phosphate, glucose-6-phosphate, glucose-l-phosphate, adenosine triphosphate, and sodium pyrophosphate) to produce inorganic phosphorus under conditions of alkaline phosphatase assay [0.05 m tris(hydroxymethyl)aminomethane buffer (pH 8.4) containing 2 mm MgCl(2)]. The purified alkaline phosphatase also acted on all these esters, although much less effectively on adenosine triphosphate and sodium pyrophosphate than did the cells. Comparison of the relative utilization of the various substrates by repressed and derepressed cells and purified enzyme suggested the presence of multiple enzymes in the cells. Thus, the cytochemical method of trapping the newly generated inorganic phosphorus determines the location of an alkaline phosphatase of broad substrate profile, and in addition locates the sites of other enzymes generating inorganic phosphorus under identical conditions of assay. It is intriguing that all of these enzymes usually exist in a few clusters attached to the peripheral plasma membrane. In addition to this predominant location, there were a few sites of enzyme activity in the cytoplasm unattached to any discernible structure, and also in the cell wall of the repression-resistant and of the derepressed, repressible strains.  相似文献   

11.
Ultraviolet difference spectra are produced by the binding of divalent metal ions to metal-free alkaline phosphatase (EC 3.1.3.1). The interaction of the apoprotein with Zn2+, Mn2+, Co2+ and Cd2+, which induce the tight binding of one phosphate ion per dimer, give distinctly different ultraviolet spectra changes from Ni2+ and Hg2+ which do not induce phosphate binding. Spectrophotometric titrations at alkaline pH of various metallo-enzymes reveal a smaller number of ionizable tyrosines and a greater stability towards alkaline denaturation in the Zn2+- and Mn2+-enzymes than in the Ni2+-, Hg2+- and apoenzymes. The Zn2+- and Mn2+-enzymes have CD spectra in the region of the aromatic transitions that are different from the CD spectra of the Ni2+-, Hg2+- and apoenzymes. Modifications of arginines with 2,3-butanedione show that a smaller number of arginine residues are modified in the Zn2+-enzyme than in the Hg2+-enzyme. The presented data indicate that alkaline phosphatase from Escherichia coli must have a well-defined conformation in order to bind phosphate. Some metal ions (i.e. Zn2+, Co2+, Mn2+ and Cd2+), when interacting with the apoenzyme, alter the conformation of the protein molecule in such a way that it is able to interact with substrate molecules, while other metal ions (i.e. Ni2+ and Hg2+) are incapable of inducing the appropriate conformational change of the apoenzyme. These findings suggest an important structural function of the first two tightly bound metal ions in enzyme.  相似文献   

12.
Enhanced chemiluminescent assays for hydrolase enzymes have been developed using proehancer and pro-anti-enhancer substrates. Alkaline phosphatase is measured using disodium para-iodophenyl phosphate (proenhancer) which is converted to para-iodophenol and this in turn enhances the light emission from the horseradish peroxidase catalysed chemiluminescent oxidation of luminol by peroxide. An alternative strategy uses para-nitrophenyl phosphate which is converted by alkaline phosphatase to para-nitrophenol which inhibits the enhanced chemiluminescent reaction. The detection limit for the enzyme using the proenhancer and pro-anti-enhancer assays was 100 attomoles and 1 picornole, respectively. The proenhancer strategy was effective in assays for beta-D-galactosidase, beta-D -glucosidase and aryl sulfatase. A limited comparison of the proenhancer and a conventional colorimetric assay for an alkaline phosphatase label in an enzyme immunoassay for alpha-fetoprotein showed good agreement.  相似文献   

13.
Callus calcifying cartilage alkaline phosphatase was resolved by DEAE-cellulose column chromatography into two distinct phsophatase activities. The phosphatase activity which was eluted first from the column, (phosphatase I), was active towards a variety of phosphate esters, sodium pyrophosphatase and several linear polyphosphates, while the second phosphatase activity , (phosphatase II), was active toward simple phosphate esters but not towards sodium pyrophosphate and linear oligo or polyphosphates. All the phosphate esters, sodium pyrophosphate and polyphosphates at higher concentrations were inhibitory for phosphatase I. The modulating effects of magnesium, calcium, zinc and other phosphatase modulators have been investigated. Both phosphatases from callus calcifying cartilage were found to be substrates of neuraminidase with sialic acid as the product. Besides the difference in their specificity, the phosphatases were found to be immunologically different and to have different molecular weights, strong indication that they are different enzymes.  相似文献   

14.
The values of K(i) for methylphosphate dianion (MP(2-)) inhibition of the cleavage of 2-hydroxypropyl-4-nitrophenyl phosphate (HpPNP) catalyzed by 1,3-bis(1,4,7-triazacyclonon-1-yl)-2-hydroxypropane (Zn(2)(1)(H(2)O)) approach a small limiting value of 6 microM at pH相似文献   

15.
Analysis of sequence alignments of alkaline phosphatases revealed a correlation between metal specificity and certain amino acid side chains in the active site that are metal-binding ligands. The Zn(2+)-requiring Escherichia coli alkaline phosphatase has an Asp at position 153 and a Lys at position 328. Co(2+)-requiring alkaline phosphatases from Thermotoga maritima and Bacillus subtilis have a His and a Trp at these positions, respectively. The mutations D153H, K328W, and D153H/K328W were induced in E. coli alkaline phosphatase to determine whether these residues dictate the metal dependence of the enzyme. The wild-type and D153H enzymes showed very little activity in the presence of Co(2+), but the K328W and especially the D153H/K328W enzymes effectively use Co(2+) for catalysis. Isothermal titration calorimetry experiments showed that in all cases except for the D153H/K328W enzyme, a possible conformation change occurs upon binding Co(2+). These data together indicate that the active site of the D153H/K328W enzyme has been altered significantly enough to allow the enzyme to utilize Co(2+) for catalysis. These studies suggest that the active site residues His and Trp at the E. coli enzyme positions 153 and 328, respectively, at least partially dictate the metal specificity of alkaline phosphatase.  相似文献   

16.
The effect of citrinin poisoning on rabbit kidney alkaline phosphatase was investigated. After seven days administration of citrinin (2 mg/kg body weight daily) the animals were sacrificed and the level of enzymes estimated in serum and kidney. Serum enzymes showed no variation in activity in the citrinin-treated animals, but in kidney, alkaline phosphatase activity decreased significantly. The decreased activity was mainly associated with the cytoplasmic fraction and in fractions Ib and II. The enzyme II obtained from citrinin-treated animal showed no kinetic difference in substrate specificity, inhibition by phenylalanine, phosphate, sodium-EDTA and Zn2+ ions, activation by Mg2+ ions, thermal inactivation and electrophoretic mobility to that of control Enzyme II. Immunological studies showed that the decrease in enzyme activity was due to existence of inactive enzyme protein. Hormones like cyclic AMP, prostaglandin E1 and parathyroid hormone reversed the decreased enzyme activity due to citrinin poisoning in mouse and rabbit. This study favours the possible existence of active and inactive forms of alkaline phosphatase in the system.  相似文献   

17.
黍子过氧化物酶(proso millet peroxidase,PmPOD)具有磷酸酶活性,可以断裂DNA中磷酸二酯键及脱氧核糖核苷酸(dNMPs)中磷酸单酯键。在此反应过程中,Mg^2+显著增强PmPOD的磷酸酶活性,但其具体的机制尚不明确。本文采用紫外-可见分光光谱法和荧光光谱法,研究了以dNMPs为底物时,Mg^2+对PmPOD磷酸酶活性的影响,并对其反应机制进行了初步的探究。紫外-可见分光光度法结果表明:Mg^2+介导了PmPOD与底物的相互作用,但Mg^2+并未直接与PmPOD发生相互作用。荧光光谱进一步表明,在Mg^2+存在的情况下,dNMPs对PmPOD内源荧光淬灭方式发生变化,由动态淬灭转变为静态淬灭。同时还发现,dNMPs与PmPOD的结合常数Ka增加约2~10倍(与不存在Mg^2+条件相比),依次为:KadCMP>KadGMP>KadTMP>KadAMP。高效液相色谱表明,Mg^2+可增强PmPOD水解dNMPs的速率3~13倍,且水解速率VdCMP>VdGMP>VdTMP>VdAMP,与结合常数的变化一致。因此,我们得出结论,PmPOD发挥磷酸酶活性时,Mg^2+首先与dNMPs形成中间产物,这一中间产物更适合与PmPOD形成复合物,增大了底物dNMPs与PmPOD结合常数,进而加速了PmPOD水解dNMPs。本研究为Mg^2+在过氧化物酶催化DNA水解的机制提供了相关依据,为研究金属离子增强蛋白酶活性的机制提供了理论基础。  相似文献   

18.
Interaction of purified human liver and placental alkaline phosphatases (orthophosphoric-monoester phosphohydrolase (alkaline optimum), EC 3.1.3.1) with sulfhydryl groups, sulfhydryl reagents, and Mg2+ were studied. L-Cysteine (0.1 mmol/l) or Mg2+ activated the liver enzyme 4-5-fold and the placental enzyme 2-3-fold, with optimal pH 7.5-8.0; these activations were not additive. L-Cysteine (2 mmol/l) inhibited both enzymes maximally at pH greater than 9.0; phosphate protected the enzymes. S-Methylcysteine had little effect, with or without Mg2+. Inhibition by sulfur-containing compounds paralleled their ability to bind Zn2+. Fluoresceine mercury acetate (specific for sulfhydryl groups) inhibited the isoenzymes, whereas iodoacetic acid, iodoacetamide, dithionitrobenzoic acid, and p-chloromercuribenzoate had little effect. The inhibition was reversed by L-cysteine and only slightly protected by inorganic phosphate. Thus, there are two sites on human liver and placental alkaline phosphatase that interact with L-cysteine; a Mg2+-binding site, which results in activation, and a site that involves one or both of the bound Zn2+ ions and results in inactivation. Both enzymes have a protected essential thiol group.  相似文献   

19.
The binding mechanism of Mg(2+) at the M3 site of human placental alkaline phosphatase was found to be a slow-binding process with a low binding affinity (K(Mg(app.)) = 3.32 mM). Quenching of the intrinsic fluorescence of the Mg(2+)-free and Mg(2+)-containing enzymes by acrylamide showed almost identical dynamic quenching constant (K(sv) = 4.44 +/- 0.09 M(-1)), indicating that there is no gross conformational difference between the M3-free and the M3-Mg(2+) enzymes. However, Zn(2+) was found to have a high affinity with the M3 site (K(Zn(app.)) = 0.11 mM) and was observed as a time-dependent inhibitor of the enzyme. The dependence of the observed transition rate from higher activity to lower activity (k(obs)) at different zinc concentrations resulted in a hyperbolic curve suggesting that zinc ion induces a slow conformational change of the enzyme, which locks the enzyme in a conformation (M3'-Zn) having an extremely high affinity for the Zn(2+) (K*(Zn(app.)) = 0.33 microM). The conformation of the M3'-Zn enzyme, however, is unfavorable for the catalysis by the enzyme. Both Mg(2+) activation and Zn(2+) inhibition of the enzyme are reversible processes. Structural information indicates that the M3 site, which is octahedrally coordinated to Mg(2+), has been converted to a distorted tetrahedral coordination when zinc ion substitutes for magnesium ion at the M3 site. This conformation of the enzyme has a small dynamic quenching constant for acrylamide (K(sv) = 3.86 +/- 0.04 M(-1)), suggesting a conformational change. Both Mg(2+) and phosphate prevent the enzyme from reaching this inactive structure. GTP plays an important role in reactivating the Zn-inhibited enzyme activity. We propose that, under physiological conditions, magnesium ion may play an important modulatory role in the cell for protecting the enzyme by retaining a favorable geometry of the active site needed for catalysis.  相似文献   

20.
1. The differential effects of adding Zn(2+) and Mg(2+) on the orthophosphatase and inorganic pyrophosphatase activities of human intestinal alkaline phosphatase were studied. 2. In the presence of excess of Zn(2+), inorganic pyrophosphatase activity is inhibited. At higher concentrations of pyrophosphate, hydrolysis of this substrate takes place, but is inhibited competitively by the Zn(2+)-pyrophosphate complex. This complex also acts as a competitive inhibitor of orthophosphate hydrolysis. 3. Excess of Mg(2+) also inhibits pyrophosphatase action by removal of substrate; at low concentrations, this ion activates pyrophosphatase, as is the case with orthophosphatase. 4. It is concluded that, when interactions between metal ions and pyrophosphate are taken into account, the effects of these ions are consistent with the view that alkaline phosphatases possess both orthophosphatase and inorganic pyrophosphatase activities.  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