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.
We have discovered an enzyme in the cytosol of Escherichia coli that generates lipid A disaccharides from monosaccharide precursors by the following route: 2,3-diacyl-GlcN-1-P + UDP-2,3-diacyl-GlcN---- 2,3-diacyl-GlcN (beta, 1----6) 2,3-diacyl-GlcN-1-P + UDP. Previous studies from our laboratory have documented the presence in vivo of the precursors 2,3-diacylglucosamine 1-phosphate (2,3-diacyl-GlcN-1-P) (lipid X of E. coli) and UDP-2,3-diacylglucosamine (UDP-2,3-diacyl-GlcN) (Bulawa, C.E., and Raetz, C.R.H.J. Biol. Chem. 259, 4846-4851). Both substrates are novel glucosamine-derived phospholipids, acylated with beta-hydroxymyristoyl moieties, and they accumulate in E. coli mutants defective in the pgsB gene. Synthetic ADP-, GDP-, and CDP-2,3-diacylglucosamines are inefficient substrates compared to the naturally occurring UDP derivative. The free-acid form of the tetraacyldisaccharide 1-phosphate product (C68H129N2O20P) that is generated in vitro has Mr = 1325.74 as judged by fast atom bombardment mass spectrometry. Mild acid hydrolysis (0.1 M HCl for 30 min at 100 degrees C) liberates greater than 95% of the phosphate moiety as Pi. Detailed analysis by 1H and 13C NMR spectroscopy confirms the presence of a phosphate residue at position 1 of the disaccharide, an alpha-anomeric configuration at the reducing end, and a beta, 1----6 linkage between the two glucosamines. Importantly the disaccharide 1-phosphate synthase is missing in extracts of E. coli strains harboring the pgsB1 mutation, consistent with the massive accumulation of 2,3-diacyl-GlcN-1-P and UDP-2,3-diacyl-GlcN in vivo. The enzymatic reaction reported here represents a major biosynthetic route for the formation of lipid A disaccharides in E. coli and other Gram-negative bacteria. An in vitro system for the biosynthesis of lipid A disaccharides has not been described previously.  相似文献   

2.
The lipid A disaccharide of the Escherichia coli envelope is synthesized from the two fatty acylated glucosamine derivatives UDP-N2,O3-bis[(R)-3-hydroxytetradecanoyl]-alpha-D- glucosamine (UDP-2,3-diacyl-GlcN) and N2,O3-bis[(R)-3-hydroxytetradecanoyl]-alpha-D-glucosamine 1-phosphate (2,3-diacyl-GlcN-1-P) [Ray, B. L., Painter, G., & Raetz, C. R. H. (1984) J. Biol. Chem. 259, 4852-4859]. We have previously shown that UDP-2,3-diacyl-GlcN is generated in extracts of E. coli by fatty acylation of UDP-GlcNAc, giving UDP-3-O-[(R)-3-hydroxymyristoyl]-GlcNAc as the first intermediate, which is rapidly converted to UDP-2,3-diacyl-GlcN [Anderson, M. S., Bulawa, C. E., & Raetz, C. R. H. (1985) J. Biol. Chem. 260, 15536-15541; Anderson, M. S., & Raetz, C. R. H. (1987) J. Biol. Chem. 262, 5159-5169]. We now demonstrate a novel enzyme in the cytoplasmic fraction of E. coli, capable of deacetylating UDP-3-O-[(R)-3-hydroxymyristoyl]-GlcNAc to form UDP-3-O-[(R)-3-hydroxymyristoyl]glucosamine. The covalent structure of the previously undescribed UDP-3-O-[(R)-3-hydroxymyristoyl] glucosamine intermediate was established by 1H NMR spectroscopy and fast atom bombardment mass spectrometry. This material can be made to accumulate in E. coli extracts upon incubation of UDP-3-O-[(R)-3- hydroxymyristoyl]-GlcNAc in the absence of the fatty acyl donor [(R)-3-hydroxymyristoyl]-acyl carrier protein. However, addition of the isolated deacetylation product [UDP-3-O-[(R)-3-hydroxymyristoyl] glucosamine] back to membrane-free extracts of E. coli in the presence of [(R)-3-hydroxymyristoyl]-acyl carrier protein results in rapid conversion of this compound into the more hydrophobic products UDP-2,3-diacyl-GlcN, 2,3-diacyl-GlcN-1-P, and O-[2-amino-2-deoxy-N2,O3- bis[(R)-3-hydroxytetradecanoyl]-beta-D-glucopyranosyl]-(1----6)-2-amino- 2-deoxy-N2,O3-bis[(R)-3-hydroxytetradecanoyl]-alpha-D- glucopyranose 1-phosphate (tetra-acyldisaccharide-1-P), demonstrating its competency as a precursor. In vitro incubations using [acetyl-3H]UDP-3-O-[(R)-3-hydroxymyristoyl]-GlcNAc confirmed release of the acetyl moiety in this system as acetate, not as some other acetyl derivative. The deacetylation reaction was inhibited by 1 mM N-ethylmaleimide, while the subsequent N-acylation reaction was not. Our observations provide strong evidence that UDP-3-O-[(R)-3-hydroxymyristoyl]glucosamine is a true intermediate in the biosynthesis of UDP-2,3-diacyl-GlcN and lipid A.  相似文献   

3.
The Gram-negative bacterium Escherichia coli has previously been shown to utilize two unique glucosamine (GlcN)-derived phospholipids in the biosynthesis of lipid A disaccharides (Bulawa, C.E., and Raetz, C. R.H. (1984) J. Biol. Chem. 259, 4846-4851; Ray, B. L., Painter, G.L., and Raetz, C.R.H. (1984) J. Biol. Chem. 259, 4852-4859. We now present evidence that these compounds, UDP-2,3-diacyl-GlcN and 2,3-diacyl-GlcN-1-phosphate (2,3-diacyl-GlcN-1-P), are generated in extracts of E. coli by fatty acylation of UDP-GlcNAc. The initial reaction is an O-acylation of the glucosamine ring, presumably of the 3-OH group, with (R)-beta-hydroxymyristate, followed by removal of the acetyl moiety, and further fatty acylation of the N atom with (R)-beta-hydroxymyristate to yield UDP-2,3-diacyl-GlcN. Hydrolysis of the pyrophosphate bridge in this molecule gives 2,3-diacyl-GlcN-1-P + UMP. In vivo pulse labeling with 32Pi supports this postulated pathway, since UDP-2,3-diacyl-GlcN is labeled prior to 2,3-diacyl-GlcN-1-P. UDP-glucosamine is inactive as a substrate in the initial acylation reaction. These acylations show an absolute specificity for fatty acyl moieties activated with acyl carrier protein. No reaction is detected with fatty acyl-CoA or free fatty acid. The fatty acylation of sugar nucleotides has not been reported previously in E. coli or any other organism.  相似文献   

4.
Escherichia coli mutants defective in the pgsB gene are phosphatidylglycerol-deficient in certain genetic settings and accumulate novel, glucosamine-derived phospholipids (Nishijima, M., and Raetz, C. R. H. (1979) J. Biol. Chem. 254, 7837-7844). The simplest of these compounds is 2,3-diacylglucosamine 1-phosphate (2,3-diacyl-GlcN-1-P) ("lipid X" of E. coli), in which beta-hydroxymyristoyl moieties are the sole fatty acid substituents (Takayama, K., Qureshi, N., Mascagni, P., Nashed, M. A., Anderson, L., and Raetz, C. R. H. (1983) J. Biol. Chem. 258, 7379-7385). We now report a sensitive radiochemical method for detection of 2,3-diacyl-GlcN-1-P in wild type E. coli and demonstrate that there are about 4000 molecules/cell (0.02% of the total CHCl3-soluble phosphorus). In mutants bearing the pgsB1 lesion, the levels are 100- to 300-fold higher. In addition, we have discovered a novel liponucleotide, UDP-2,3-diacyl-GlcN, that also accumulates in conjunction with the pgsb1 mutation. This material represents 0.005% of the wild type phospholipid and accumulates 50- to 100-fold in the mutant. The identification of UDP-2,3-diacyl-GlcN in E. coli is based on: 1) migration of a minor 32P-labeled lipid from wild type and mutant cells with a UDP-2,3-diacyl-GlCn standard during two-dimensional thin layer chromatography; 2) susceptibility of this 32P-labeled material to cleavage by a liponucleotide-specific pyrophosphatase; and 3) chromatographic identification of [32P]UMP and [32P]2,3-diacyl-GlcN-1-P (lipid X) as the sole products of the enzymatic degradation. As shown in the accompanying article, this novel nucleotide is crucial for biosynthesis of lipid A disaccharides in extracts of E. coli and Salmonella typhimurium.  相似文献   

5.
Extracts of Escherichia coli contain an enzyme that generates the beta,1----6 linkage of lipid A from fatty-acylated monosaccharide precursors, according to the reaction: 2,3-diacyl-GlcN-1-P + UDP-2,3-diacyl-GlcN----2,3-diacyl-GlcN (beta, 1----6)2,3-diacyl-GlcN-1-P + UDP (Ray, B. L., Painter, G., and Raetz, C. R. H. (1984) J. Biol. Chem. 259, 4852-4859). We now describe a membrane-bound kinase that phosphorylates the 4'-position of the above tetraacyldisaccharide 1-phosphate product. The lipid A 4'-kinase is distinct from the diglyceride kinase of E. coli. When crude membrane preparations are employed, several nucleoside triphosphates are able to support the phosphorylation of the tetraacyldisaccharide 1-phosphate, but ATP is the most efficient. The 4'-kinase requires Mg2+ and is stimulated by phospholipids, especially cardiolipin. Under optimal conditions the specific activity in crude extracts is 0.5 nmol/min/mg. The enzyme is rapidly inactivated by preincubation in the presence of detergents, such as Nonidet P-40 or octylglucoside, but phosphoenolpyruvate and glycerol stabilize the enzyme. The product generated in vitro has been characterized by fast atom bombardment mass spectrometry and by 1H and 31P NMR spectroscopy. Those analyses confirm that the 4' hydroxyl is the site of phosphorylation. The 4'-kinase reported here is likely to represent a key step in the de novo biosynthesis of lipid A.  相似文献   

6.
Ribose-1,5-bisphosphate is synthesized in a reaction that uses ribose-1(or 5)-P as the phosphoryl acceptor and the acyl-P of 3-phosphoglyceryl phosphate as the donor. Glucose-1,6-bisphosphate is synthesized in a similar reaction. The relative activity with the two substrates remains unchanged over almost 300-fold purification of the enzyme, indicating that glucose-1,6-bisphosphate synthase catalyzes both reactions. The relative V/Km values for alternative phosphoryl acceptors are ribose-1-P (1); glucose-1-P (0.30); mannose-1-P and ribose-5-P (0.11); glucose-6-P (0.10); 2-deoxyglucose-6-P (0.03); and 2-deoxyribose-5-P (0.02). Fructose-1- and 6-phosphates are not substrates. The synthesis of both ribose-1,5-bisphosphate and glucose-1,6-bisphosphate is inhibited by physiologically significant levels of fructose-1,6-bisphosphate, glycerate-2,3-bisphosphate, glycerate-3-phosphate, citrate, and inorganic phosphate. Ribose-1,5-bisphosphate is a strong activator of brain phosphofructokinase.  相似文献   

7.
The lpxH gene encodes the UDP-2,3-diacylglucosamine-specific pyrophosphatase that catalyzes the fourth step of lipid A biosynthesis in Escherichia coli. To confirm the function of lpxH, we constructed KB21/pKJB5. This strain contains a kanamycin insertion element in the chromosomal copy of lpxH, complemented by plasmid pKJB5, which is temperature-sensitive for replication and harbors lpxH(+). KB21/pKJB5 grows at 30 degrees C but loses viability at 44 degrees C, demonstrating that lpxH is essential. CDP-diglyceride hydrolase (Cdh) catalyzes the same reaction as LpxH in vitro but is non-essential and cannot compensate for the absence of LpxH. The presence of Cdh in cell extracts interferes with the LpxH assay. We therefore constructed KB25/pKJB5, which contains both an in-frame deletion of cdh and a kanamycin insertion mutation in lpxH, covered by pKJB5. When KB25/pKJB5 cells are grown at 44 degrees C, viability is lost, and all in vitro LpxH activity is eliminated. A lipid migrating with synthetic UDP-2,3-diacylglucosamine accumulates in KB25/pKJB5 following loss of the covering plasmid at 44 degrees C. This material was converted to the expected products, 2,3-diacylglucosamine 1-phosphate and UMP, by LpxH. Pseudomonas aeruginosa contains two proteins with sequence similarity to E. coli LpxH. The more homologous protein catalyzes UDP-2,3-diacylglucosamine hydrolysis in vitro. The corresponding gene complements KB25/pKJB5 at 44 degrees C, but the less homologous gene does not. The accumulation of UDP-2,3-diacylglucosamine in our lpxH mutant is consistent with the observation that the lipid A disaccharide synthase LpxB, the next enzyme in the pathway, cannot condense two UDP-2,3-diacylglucosamine molecules, but instead utilizes UDP-2,3-diacylglucosamine as its donor and 2,3-diacylglucosamine 1-phosphate as its acceptor.  相似文献   

8.
We recently reported the partial purification of a cAMP-independent and Ca2+-calmodulin-independent glycogen synthase kinase from porcine renal cortex (Schlender, K. K., Beebe, S. J., and Reimann, E. M. (1981) Cold Spring Harbor Conf. Cell Proliferation, 389-400). Subsequent purification indicated that the enzyme preparation consisted of at least three forms of glycogen synthase kinase which could be resolved by ATP gradient elution from aminoethylphosphate-agarose (AEP-agarose). The predominant form of glycogen synthase kinase, which eluted from AEP-agarose between 2 and 6 mM ATP, was purified approximately 800-fold and is designated GSK-A1. It had a molecular weight of 45,000-50,000 as determined by gel filtration and sucrose density gradient centrifugation. It catalyzed the transfer of 1 mol of 32P/mol of synthase subunit into a low molecular weight (10,000) CNBr peptide which was tentatively identified as Ser-7 (site 2) by high performance liquid chromatography. This phosphorylation decreased the activity ratio (activity in the absence of glucose-6-P divided by activity in the presence of 7.2 mM glucose-6-P) from 0.95 to about 0.55. GSK-A1 appeared to be specific for and had low s0.5 values for both substrates, ATP (13 microM) and glycogen synthase (0.3-0.4 microM). The enzyme could not use GTP as the phosphate donor. GSK-A1 was not affected by the protein kinase inhibitor, cAMP, cGMP, Ca2+-calmodulin, EGTA, or trifluoperazine and had a broad pH optimum (pH 7.0-8.5). A second form, GSK-A2, was eluted from AEP-agarose between 7 and 9 mM ATP. GSK-A2 could transfer a 2nd mol of 32P/mol of synthase subunit and decreased the activity ratio to 0.30. The interrelation among these multiple forms is not clear, but the data suggest that multiple kinases are required to form the highly inactivated glycogen synthase in renal tissues.  相似文献   

9.
The levels of glycerate 2,3-P2 and of 2,3-bisphosphoglycerate synthase and 2,3-bisphosphoglycerate phosphatase activities have been determined in isolated rat hepatocytes and adipocytes and in perfused rat tissues to discard blood contamination. The values obtained are much lower than those previously reported, ranging 0.50-40 nmol/g tissue. No relationship appears to exist between glycerate 2,3-P2 concentration and the levels of the enzymatic activities involved in glycerate 2,3-P2 metabolism. Assay of glycerate 2,3-P2 in tissue extracts constitute a very useful way to quantify blood contamination.  相似文献   

10.
The steady state kinetics and effects of salts on chicken breast phosphoglycerate mutase have been examined. The enzyme can catalyze three phosphoryl transfer reactions: mutase, bisphosphoglycerate phosphatase, and bisphosphoglycerate synthase. The mutase rate was measured in the favorable direction (Keq = glycerate-3-P/glycerate-2-P approximately equal to 12) using [2T]glycerate-2-P as substrate. The bisphosphoglycerate phosphatase activity was studied in the presence of the activator, glycolate-2-P. The latter is an analog of the glycerate-P's and appears to act as an abortive mutase substrate. The kinetic pattern obtained with both activities is that of a ping-pong mechanism with inhibition by the second substrate occurring at a lower concentration than the Km value for that substrate. The kinetic parameters for the mutase determined in 50 mM N-[tris(hydroxymethyl)methyl-2-amino]ethanesulfonate (TES)/sodium buffer containing 0.1 M KCl, pH 7.5, 25 degrees C are: Km glycerate-2,3-P2, 0.069 micron; Km glycerate-2-P, 14 micron; Km glycerate-3-P approximately 200 micron; Ki glycerate-2-P, 4 micron. The kinetic parameters for the phosphatase reaction in 50 mM triethanolamine/Cl- buffer, pH 7.5, 25 degrees C are: Km glycerate-2,3-P2, 0.065 micron:Km glycolate-2P, 479 micron; Ki glycolate-2-P, 135 micron. The enzyme is sensitive to changes in the ionic environment. Increasing salt concentrations activate the phosphatase in the presence of glycolate-2-P by decreasing the apparent Km of glycerate-2,3-P2. The effects are due to the anionic component and Cl- greater than acetate greater than TES. The same salts are competitive inhibitors with respect to glycolate-2-P. With high levels of KCl that produce a 30-fold decrease in the apparent maximal velocity due to competition with glycolate-2-P, the Km of glycerate-2,3-P2 remains low. These observations lead us to postulate that each monophosphoglycerate substrate has a separate site on the enzyme and that glycerate-2,3-P2 can bind to either site. The binding of anions to one site of the nonphosphorylated enzyme allows an increase in the on and off rates of glycerate-2,3-P2 at the alternate site. Salts inhibit the mutase reaction. The Km of glycerate-2,3-P2 is increased as is that of glycerate-2-P. The effect on the Km of glycerate-2,3-P2 is attributed to an increase in the off rate/on rate ratio for glycerate-2,3-P2. The bisphosphoglycerate synthase reaction is shown to require added glycerate-3-P. The equilibrium between enzyme and glycerate-1,3-P2 is favorable (Kdiss less than or equal 7 X 10(-8) M) and suggests that in the absence of a separate synthase this reaction may have functional significance.  相似文献   

11.
Ligand-binding studies with labelled triethyltin on yeast mitochondrial membranes showed the presence of high-affinity sites (KD = 0.6 micronM; 1.2 +/- 0.3 nmol/mg of protein) and low-affinity sites (KD less than 45 micronM; 70 +/- 20 nmol/mg of protein). The dissociation constant of the high-affinity site is in good agreement with the concentration of triethyltin required for inhibition of mitochondrial ATPase (adenosine triphosphatase) and oxidative phosphorylation. The high-affinity site is not competed for by oligomycin or venturicidin, indicating that triethyltin reacts at a different site from these inhibitors of oxidative phosphorylation. Fractionation of the mitochondrial membrane shows a specific association of the high-affinity sites with the ATP synthase complex. During purification of ATP synthase (oligomycin-sensitive ATPase) there is a 5-6-fold purification of oligomycin- and triethyltin-sensitive ATPase activity concomitant with a 7-9-fold increase in high-affinity triethyltin-binding sites. The purified yeast oligomycin-sensitive ATPase complex contains approximately six binding sites for triethyltin/mol of enzyme complex. It is concluded that specific triethyltin-binding sites are components of the ATP synthase complex, which accounts for the specific inhibition of ATPase and oxidative phosphorylation by triethyltin.  相似文献   

12.
Solubilization of phosphatidylinositol (PtdIns) synthase (CDP-diacylglycerol: myo-inositol 3-phosphatidyltransferase, EC 2.7.8.11) from rat pituitary (GH3) tumours was investigated. PtdIns synthase activity was partially extracted from crude membranes by 3 M-KCl. Prior separation of membranes revealed that a greater proportion of plasma-membrane PtdIns synthase activity was salt-extractable than was endoplasmic reticulum activity. The activity of the salt-extracted enzyme was maximized by low concentrations of 3-(3-cholamidopropyl) dimethylammonio-1-propanesulphonate (CHAPS; 0.5 mM), Triton X-100 (0.1 mM) or a phospholipid mixture (0.05 mg/ml), but higher concentrations of detergents were inhibitory. The activity of salt-extracted PtdIns synthase was 0.25 +/- 0.08 nmol/min per mg of protein. Salt-extracted PtdIns synthase activity was dependent on Mg2+ (maximal at 0.1 mM) and Mn2+ (maximal at 5 mM), and its pH optimum was in the range 7.0-7.5. The apparent Km for myo-inositol (in the presence of 0.1 mM-CDP-diacylglycerol) was 0.06 mM, and that for CDP-diacylglycerol (at 0.1 mM-myo-inositol) was 0.21 mM. Salt-extracted PtdIns synthase activity was potently inhibited by Ca2+ (50% inhibition at 1 microM), with over 90% inhibition at 10 microM-Ca2+. These data imply the existence of two forms of membrane-associated PtdIns synthase, namely salt-extractable and salt-resistant, with different intracellular localizations. The salt-extractable form of this enzyme may be a useful preparation for further characterization and purification of mammalian PtdIns synthase.  相似文献   

13.
Antiserum prepared from a rabbit which was immunized with human erythrocyte glycerate-2,3-P2 synthase was found to react with glycerate-2,3-P2 synthase in rabbit erythroid cells. By using this antiserum, it was proved that the specific activity of this enzyme was unchanged during the development of the rabbit erythroid cells. This leads us to conclude that the increased activity of glycerate-2,3-P2 synthase in developing erythroid cells (Narita, H., Ikura, K., Yanagawa, S., Sasaki, R., Chiba, H., Saimyoji, H., and Kumagai, N. (1980) J. Biol. Chem. 255, 5230-5235) is due to the accumulation of enzyme protein. There is at least a 16-fold increase in the level of this protein during development from bone marrow erythroid cells to erythrocytes. The synthesis of glycerate-2,3-P2 synthase was shown to occur in rabbit reticulocytes and bone marrow erythroid cells. These cells were incubated for protein synthesis and the protein synthesized was precipitated with the anti-glycerate-2,3-P2 synthase antiserum and separated on sodium dodecyl sulfate-polyacrylamide gels. The immunoprecipitated product was shown to produce fragments of the same molecular weight after digestion with V8 protease as did the pure glycerate-2,3-P2 synthase. The proportion of glycerate-2,3-P2 synthase synthesis in reticulocytes (0.04% of total protein synthesis) was comparable to the level of this protein in the cells (0.07% of the total protein).  相似文献   

14.
The kinetic properties of phosphofructokinase from muscle of the giant cirripede Austromegabalanus psittacus were characterized, after partial purification by ion exchange chromatography on DEAE-cellulose. This enzyme showed differences regarding PFKs from other marine invertebrates: the affinity for fructose 6-phosphate (Fru 6-P) was very low, with an S(0.5) of 22.6+/-1.4 mM (mean+/-S.D., n=3), and a high cooperativity (n(H) of 2.90+/-0.21; mean+/-S.D., n=3). The barnacle PFK showed hyperbolic saturation kinetics for ATP (apparent K(m ATP)=70 microM, at 5 mM Fru 6-P, in the presence of 2 mM ammonium sulfate). ATP concentrations higher than 1 mM inhibited the enzyme. Ammonium sulfate activated the PFK several folds, increasing the affinity of the enzyme for Fru 6-P and V(max). 5'-AMP (0.2 mM) increased the affinity for Fru 6-P (S(0.5) of 6.2 mM). Fructose 2,6-bisphosphate activated the PFK, with a maximal activation at concentrations higher than 2 microM. Citrate reverted the activation of PFK produced by 0.2 mM 5'-AMP (IC(50 citrate)=2.0 mM), producing a higher inhibition than that exerted on other invertebrate PFKs. Barnacle muscular PFK was activated in vitro after exposure to exogenous cyclic-AMP (0.1 mM) as well as by phosphatidylserine (50 microg/ml), indicating a possible control by protein kinase A and a phospholipid dependent protein kinase (PKC). The results suggest a highly regulated enzyme in vivo, by allosteric mechanisms and also by protein phosphorylation.  相似文献   

15.
Previously we reported that stable transfection of human UDP-glucose pyrophosphorylase (hUGP2) rescued galactose-1-phosphate uridyltransferase (GALT)-deficient yeast from "galactose toxicity." Here we test in human cell lines the hypothesis that galactose toxicity was caused by excess accumulation of galactose-1-phosphate (Gal-1-P), inhibition of hUGP2, and UDP-hexose deficiency. We found that SV40-transformed fibroblasts derived from a galactosemic patient accumulated Gal-1-P from 1.2+/-0.4 to 5.2+/-0.5 mM and stopped growing when transferred from 0.1% glucose to 0.1% galactose. Control fibroblasts accumulated little Gal-1-P and continued to grow. The GALT-deficient cells had 157+/-10 micromoles UDP-glucose/100 g protein and 25+/-5 micromoles UDP-galactose/100 g protein when grown in 0.1% glucose. The control cells had 236+/-25 micromoles UDP- glucose/100 g protein and 82+/-10 micromoles UDP-galactose/100 g protein when grown in identical medium. When we transfected the GALT-deficient cells with either the hUGP2 or GALT gene, their UDP-glucose content increased to 305+/-28 micromoles/100 g protein (hUGP2-transfected) and 210+/-13 micromoles/100 g protein (GALT-transfected), respectively. Similarly, UDP-galactose content increased to 75+/-12 micromoles/100 g protein (hUGP2-transfected) and 55+/-9 micromoles/100 g protein (GALT-transfected), respectively. Though the GALT-transfected cells grew in 0.1% galactose with little accumulation of Gal-1-P (0.2+/-0.02 mM), the hUGP2-transfected cells grew but accumulated some Gal-1-P (3.1+/-0.4 mM). We found that 2.5 mM Gal-1-P increased the apparent KM of purified hUGP2 for glucose-1-phosphate from 19.7 microM to 169 microM, without changes in apparent Vmax. The Ki of the reaction was 0.47 mM. Gal-1-P also inhibited UDP-N-acetylglucosamine pyrophosphorylase, which catalyzes the formation of UDP-N-acetylglucosamine. We conclude that intracellular concentrations of Gal-1-P found in classic galactosemia inhibit UDP-hexose pyrophosphorylases and reduce the intracellular concentrations of UDP-hexoses. Reduced Sambucus nigra agglutinin binding to glycoproteins isolated from cells with increased Gal-1-P is consistent with the resultant inhibition of glycoprotein glycosylation.  相似文献   

16.
The enzyme UDP-glucose pyrophosphorylase (UGPase) from potato (Solanum tuberosum L. cv Norchip) tubers was purified 177-fold to near homogeneity and to a specific activity of 1099 international units/mg of protein. The molecular mass of the purified enzyme was 53 kD as determined by SDS-PAGE and gel filtration. Immunological and activity assays detected UGPase at similar levels in potato stems, stolons, and tubers. Leaves and roots contained lower levels of UGPase activity and protein. Lineweaver-Burk plots for substrates inorganic pyrophosphate and UDP-glucose were linear in the pyrophosphorolytic direction, yielding Km values of 0.13 and 0.14 mM, respectively. However, Lineweaver-Burk plots for the substrates glucose-1-P and UTP were biphasic in nature when UGPase was assayed in the direction of UDP-glucose synthesis. At physiological substrate concentrations (i.e. from 0.05-0.20 mM), Km values of 0.08 mM (glucose-1-P) and 0.12mM (UTP) were obtained. When substrate concentrations increased above 0.20 mM, Km values increased to 0.68 mM (glucose-1-P) and 0.53 mM (UTP). These kinetic patterns of potato UGPase suggest a "negative cooperative effect" (A. Conway, D.E. Koshland, Jr. [1968] Biochemistry 7: 4011-4022) with respect to the substrates glucose-1-P and UTP. The biphasic substrate saturation curves were similar to the kinetics of the dimeric form of UGPase purified from Salmonella typhimurium (T. Nakae [1971] J Biol Chem 246: 4404-4411). The in vivo significance of the enzyme's "negative cooperativity" in the direction of UDP-glucose synthesis and potato sweetening is discussed.  相似文献   

17.
S L Li  P F Yan  I B Paz  Y Fujita-Yamaguchi 《Biochemistry》1992,31(49):12455-12462
We have expressed, purified, and characterized the insulin receptor protein tyrosine kinase (PTK) retaining the transmembrane and downstream domains. The proteins expressed in insect cells using a baculovirus expression system were identified as membrane-bound by immunofluorescence staining and biochemical characterization. One-step purification by immunoaffinity chromatography from Triton X-100 cell extracts resulted in a approximately 360-fold increase in the specific kinase activity with a yield of approximately 50%. An appMr = approximately 60,000 protein was the major component identified by both silver staining of the purified enzyme and immunostaining of the crude extracts after separation by sodium dodecyl sulfate-polyacrylamide gel electrophoresis under reducing conditions. Using nondenaturing conditions, the molecular weight was estimated to be approximately 250,000 and approximately 500,000 by glycerol gradient centrifugation and gel permeation chromatography, respectively, suggesting that oligomers of the beta-subunit domains such as tetramers and octamers are formed. The basal PTK activity of this enzyme was much higher than those of previously reported soluble-form insulin receptor PTKs expressed in insect cells or the native receptor. Km and Vmax for two substrates, src-related peptide and poly(Glu, Tyr) (4:1), were 2.4 mM and 2.5 mumol min-1 mg-1 and 0.26 mM and 1.2 mumol min-1 mg-1, respectively. Specific activities measured under two previously reported conditions using histone H2B as a substrate were 100 or 135 nmol min-1 mg-1, in contrast to those of soluble PTKs which were reported to be 20 or 70 nmol min-1 mg-1, respectively. The purified enzyme was autophosphorylated at Tyr residues. Autophosphorylation activated the enzyme approximately 3-fold.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
2,4-Dinitrotoluene (2,4-DNT) and 2,6-DNT are priority pollutants, and 2,4-DNT dioxygenase of Burkholderia sp. strain DNT (DDO) catalyzes the initial oxidation of 2,4-DNT to form 4-methyl-5-nitrocatechol and nitrite but has significantly less activity on other dinitrotoluenes and nitrotoluenes (NT). Hence, oxidation of 2,3-DNT, 2,4-DNT, 2,5-DNT, 2,6-DNT, 2NT, and 4NT were enhanced here by performing saturation mutagenesis on codon I204 of the alpha subunit (DntAc) of DDO and by using a membrane agar plate assay to detect catechol formation. Rates of degradation were quantified both by the formation of nitrite and by the formation of the intermediates with high performance liquid chromatography. The degradation of both 2,3-DNT and 2,5-DNT were achieved for the first time (no detectable activity with the wild-type enzyme) using whole Escherichia coli TG1 cells expressing DDO variants DntAc I204L and I204Y (0.70 +/- 0.03 and 0.22 +/- 0.02 nmol/min/mg protein for 2,5-DNT transformation, respectively). DDO DntAc variant I204L also transformed both 2,6-DNT and 2,4-DNT 2-fold faster than wild-type DDO (0.8 +/- 0.6 nmol/min/mg protein and 4.7 +/- 0.5 nmol/min/mg protein, respectively). Moreover, the activities of DDO for 2NT and 4NT were also enhanced 3.5-fold and 8-fold, respectively. Further, DntAc variant I204Y was also discovered with comparable rate enhancements for the substrates 2,4-DNT, 2,6-DNT, and 2NT but not 4NT. Sequencing information obtained during this study indicated that the 2,4-DNT dioxygenases of Burkholderia sp. strain DNT and B. cepacia R34 are more closely related than originally reported. This is the first report of engineering an enzyme for enhanced degradation of nitroaromatic compounds and the first report of degrading 2,5-DNT.  相似文献   

19.
Purification and characterization of human erythrocyte uridylyl transferase   总被引:1,自引:0,他引:1  
A new method for the purification of human erythrocyte uridylyl transferase (UDPglucose: alpha-D-galactose-1-phosphate uridylyltransferase EC 2.7.7.12) is described. It consists of a hydrophobic purification step associated with hydroxyapatite chromatography and provided for the first time a purification of more than 45 000-fold with a high activity (15 I.U/mg) and a yield of 32%. We show that the enzyme is a dimer and has a molecular weight of 88 000. It can be resolved into three bands by isoelectric focusing with an apparent pI between 5.0 and 5.4. It could be shown by steady-state initial rate measurements that the interconversion of the two substrates of human transferase (Gal-1-P and UDP-glucose) follows ping-pong bi-bi kinetics, with Km values of 0.2 and 0.065 mM, respectively.  相似文献   

20.
The ability of the chicken erythrocyte to accumulate 2,3-bisphosphoglycerate (2,3-P2-glycerate) and its effect upon the oxygen affinity (P50) of the cell suspensions have been determined. Erythrocytes from chick embryos, which contain 4-6 mM 2,3-P2-glycerate, and from chickens at various ages, which contain 3-4 mM inositol pentakisphosphate but no 2,3-P2-glycerate, were incubated with inosine, pyruvate, and inorganic phosphate. Red blood cells from 20-day chick embryos incubated in Krebs-Ringer, pH 7.45, containing 20 mM inosine and 20 mM pyruvate had an increase in 2,3-P2-glycerate from 4.3 to 11.9 mM after 4 h. Importantly, as 2,3-P2-glycerate concentration increased there was a corresponding increase in P50 of the cell suspension. Further, erythrocytes from 9- and 11-week, and 7-, 14-, 24-, and 28-month-old chickens when incubated similarly with inosine and pyruvate accumulated 2,3-P2-glycerate with corresponding increases in P50 of the cell suspensions. The ability of the red cell to accumulate this compound under the incubation conditions used apparently decreases with age of the bird (e.g., 11.9 mM in the 20-day embryo to 1.1 mM in the 28-month-old chicken after 4 h incubation). Despite the presence of significant amounts of inositol-P5, the accumulation of 2,3-P2-glycerate markedly decreases oxygen affinity of the cell suspensions. The delta P50/mumol increase in 2,3-P2-glycerate in the red cells of the 20-day chick embryo after 4 h incubation is 1.5 Torr; conversely, the delta P50/mumol decrease in 2,3-P2-glycerate in the red cells of the 17-day embryo after 6 h incubation in the presence of sodium bisulfite is 2.8 Torr. The demonstrated ability of the chicken erythrocyte to accumulate 2,3-P2-glycerate in response to certain substrates suggests that regulation of concentration of this compound could contribute significantly to regulation of blood oxygen affinity in birds.  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