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 234 毫秒
1.
Both cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase were recovered mainly from the supernatant fractions of guinea-pig pancreas, but a higher proportion of the activity of the former was associated with the pellet fractions. The activities in the supernatant were not separated by gel filtration, but were clearly separated by subsequent chromatography on an anion-exchange resin. The activities of cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase had high-affinity (Km 6.5±1.1μm and 31.9±3.9μm respectively) and low-affinity (Km 0.56±0.05mm and 0.32±0.03mm respectively) components. The activity of neither enzyme was affected by the pancreatic secretogens, cholecystokinin-pancreozymin, secretin and carbachol. Removal of ions by gel filtration resulted in a marked reduction in cyclic nucleotide phosphodiesterase activity, which could be restored by addition of Mg2+. Mn2+ (3mm) was as effective as Mg2+ (3mm) in the case of cyclic AMP phosphodiesterase, but was less than half as effective in the case of cyclic GMP phosphodiesterase. The metal-ion chelators, EDTA and EGTA, also decreased activity. Ca2+ (1mm) did not affect the activity of cyclic nucleotide phosphodiesterase when the concentration of Mg2+ was 3mm. At concentrations of Mg2+ between 0.1 and 1mm, 1mm-Ca2+ was activatory, and at concentrations of Mg2+ below 0.1mm, 1mm-Ca2+ was inhibitory. These results are discussed in terms of the possible significance of cyclic nucleotide phosphodiesterase in the physiological control of cyclic nucleotide concentrations during stimulus–secretion coupling.  相似文献   

2.
1. The polychaete worm Marphysa sanguinea has a circulating erythrocruorin of mol.wt. about 2·4×106 (S020,w 58·2s, D20,w 2·06×10−7 cm.2/sec). This is the predominant form existing at pH 6–8 and (non-protein) I 0·10–0·21, and also at approx. pH 6·7 and I 0·15–3·00. 2. The pigment contains 2·24% of protohaem. 3. The 58s protein has an electrophoretic mobility of 8·08×10−5 cm.2/v/sec. at pH 8·12, I 0·21 and 0°. The isoelectric point of suspended particles is 4·63 at I 0·16 and 21·5°. 4. At very low ionic strength and pH 6·7 (unbuffered) the 58s pigment associates reversibly to 97s and 150s forms, which are probably dimer and tetramer species. 5. At pH 10·0 and I 0·025, it dissociates irreversibly to give a small amount of 2–4s non-haem-containing protein and much 9s haem-enriched protein. These and the 58s pigment may correspond to structures found in Levin's (1963) electron-microscope studies of other erythrocruorins. 6. Absorption spectra of the 58s oxygenated erythrocruorin and the deoxygenated and carbon monoxide derivatives have been obtained.  相似文献   

3.
1. Rat-liver mitochondria showed a decrease in amino acid production after preparation in 0·25m-sucrose containing EDTA (1mm), but an increase in water content. When EDTA was replaced by Mn2+ (1mm) or succinate (1mm), both amino acid production and water content were lowered, whereas preparation in 0·9% potassium chloride caused an increase in both. 2. Amino acid production by rat-liver homogenates prepared in 0·9% potassium chloride or 0·25m-sucrose was similar (qamino acid 0·047 and 0·042 respectively aerobically). After freezing-and-thawing qamino acid values were approximately doubled, and approached that of a homogenate prepared in water. 3. All cations tested inhibited amino acid production by mitochondria, Hg2+ and Zn2+ being the most effective in tris–hydrochloric acid buffer. In phosphate buffer Mg2+ and Mn2+ had no effect. Of the anions tested only pyrophosphate and arsenate had any inhibitory effect at final concn. 1mm. 4. Iodosobenzoate (1mm) and p-chloromercuribenzenesulphonate (1mm) inhibited mitochondrial amino acid production by 70–80%, whereas soya-bean trypsin inhibitor, EDTA and di-isopropyl phosphorofluoridate inhibited by a maximum of 30%. Respiratory inhibitors had no effect. 5. Rat-liver homogenate and subcellular fractions each showed an individual pattern of inhibition when a series of inhibitors was tested. 6. Amino acid production by mitochondria was decreased by up to 50% in the presence of oxidizable substrate, apart from α-glycerophosphate and palmitate, which had no effect. CoA stimulated amino acid production in tris–hydrochloric acid but not in phosphate buffer, α-oxoglutarate abolishing the stimulation. 7. Cysteine and glutathione stimulated amino acid production by whole mitochondria by 30%, but only reduced glutathione stimulated production in broken mitochondria. 8. Adrenocorticotrophic hormone and growth hormone stimulated mitochondrial amino acid production by 21–24%, whereas insulin inhibited production by 25%. 9. Coupled oxidative phosphorylation increased amino acid production by up to 154% at 25° and 40°. The increase was abolished by 2,4-dinitrophenol. 10. Amino acid incorporation in mitochondria was accompanied by an increase in amino acid production, both being decreased by chloramphenicol. 11. Mitochondrial production of ninhydrin-positive material was increased in the presence of albumin. The biggest increase was noted for the soluble fraction of broken mitochondria. No increase was found in the presence of 14C-labelled algal protein or denatured mitochondrial protein.  相似文献   

4.
1. Rat tissue homogenates convert dl-1-aminopropan-2-ol into aminoacetone. Liver homogenates have relatively high aminopropanol-dehydrogenase activity compared with kidney, heart, spleen and muscle preparations. 2. Maximum activity of liver homogenates is exhibited at pH9·8. The Km for aminopropanol is approx. 15mm, calculated for a single enantiomorph, and the maximum activity is approx. 9mμmoles of aminoacetone formed/mg. wet wt. of liver/hr.at 37°. Aminoacetone is also formed from l-threonine, but less rapidly. An unidentified amino ketone is formed from dl-4-amino-3-hydroxybutyrate, the Km for which is approx. 200mm at pH9·8. 3. Aminopropanol-dehydrogenase activity in homogenates is inhibited non-competitively by dl-3-hydroxybutyrate, the Ki being approx. 200mm. EDTA and other chelating agents are weakly inhibitory, and whereas potassium chloride activates slightly at low concentrations, inhibition occurs at 50–100mm. 4. It is concluded that aminopropanol-dehydrogenase is located in mitochondria, and in contrast with l-threonine dehydrogenase can be readily solubilized from mitochondrial preparations by ultrasonic treatment. 5. Soluble extracts of disintegrated mitochondria exhibit maximum aminopropanol-dehydrogenase activity at pH9·1 At this pH, Km values for the amino alcohol and NAD+ are approx. 200 and 1·3mm respectively. Under optimum conditions the maximum velocity is approx. 70mμmoles of aminoacetone formed/mg. of protein/hr. at 37°. Chelating agents and thiol reagents appear to have little effect on enzyme activity, but potassium chloride inhibits at all concentrations tested up to 80mm. dl-3-Hydroxybutyrate is only slightly inhibitory. 6. Dehydrogenase activities for l-threonine and dl-4-amino-3-hydroxybutyrate appear to be distinct from that for aminopropanol. 7. Intraperitoneal injection of aminopropanol into rats leads to excretion of aminoacetone in the urine. Aminoacetone excretion proportional to the amount of the amino alcohol administered, is complete within 24hr., but represents less than 0·1% of the dose given. 8. The possible metabolic role of amino alcohol dehydrogenases is discussed.  相似文献   

5.
Vessal M  Hassid WZ 《Plant physiology》1973,51(6):1055-1060
d-Glucosamine-6-P N-acetyltransferase (EC 2.3.1.4) from mung bean seeds (Phaseolus aureus) was purified 313-fold by protamine sulfate and isoelectric precipitation, ammonium sulfate and acetone fractionation, and CM Sephadex column chromatography. The partially purified enzyme was highly specific for d-glucosamine-6-P. Neither d-glucosamine nor d-galactosamine could replace this substrate. The partially purified enzyme preparation was inhibited up to 50% by 2 × 10−2m EDTA, indicating the requirement of a divalent cation. Among divalent metal ions tested, Mg2+ was required for maximum activity of the enzyme. Mn2+ and Zn2+ were inhibitory, while Co2+ had no effect on the enzyme activity. The pH optimum of the enzyme in sodium acetate and sodium citrate buffers was found to be 5.2. The effect of Mg2+ on the enzyme in sodium acetate and sodium citrate buffers was particularly noticeable in the range of optimum pH. Km values of 15.1 × 10−4m and 7.1 × 10−4m were obtained for d-glucosamine-6-P and acetyl CoA, respectively. The enzyme was completely inhibited by 1 × 10−4mp-hydroxymercuribenzoate, and this inhibition was partially reversed by l-cysteine; indicating the presence of sulfhydryl groups at or near the active site of the enzyme.  相似文献   

6.
1. The kinetic properties of the soluble and particulate hexokinases from rat heart have been investigated. 2. For both forms of the enzyme, the Km for glucose was 45μm and the Km for ATP 0·5mm. Glucose 6-phosphate was a non-competitive inhibitor with respect to glucose (Ki 0·16mm for the soluble and 0·33mm for the particulate enzyme) and a mixed inhibitor with respect to ATP (Ki 80μm for the soluble and 40μm for the particulate enzyme). ADP and AMP were competitive inhibitors with respect to ATP (Ki for ADP was 0·68mm for the soluble and 0·60mm for the particulate enzyme; Ki for AMP was 0·37mm for the soluble and 0·16mm for the particulate enzyme). Pi reversed glucose 6-phosphate inhibition with both forms at 10mm but not at 2mm, with glucose 6-phosphate concentrations of 0·3mm or less for the soluble and 1mm or less for the particulate enzyme. 3. The total activity of hexokinase in normal hearts and in hearts from alloxan-diabetic rats was 21·5μmoles of glucose phosphorylated/min./g. dry wt. of ventricle at 25°. The temperature coefficient Q10 between 22° and 38·5° was 1·93; the ratio of the soluble to the particulate enzyme was 3:7. 4. The kinetic data have been used to predict rates of glucose phosphorylation in the perfused heart at saturating concentrations of glucose from measured concentrations of ATP, glucose 6-phosphate, ADP and AMP. These have been compared with the rates of glucose phosphorylation measured with precision in a small-volume recirculation perfusion apparatus, which is described. The correlation between predicted and measured rates was highly significant and their ratio was 1·07. 5. These findings are consistent with the control of glucose phosphorylation in the perfused heart by glucose 6-phosphate concentration, subject to certain assumptions that are discussed in detail.  相似文献   

7.
1. Human uterine cervical stroma was found to contain a Ca2+-independent neutral proteinase against casein and N-benzoyl-dl-arginine p-nitroanilide (Bz-dl-Arg-Nan). This enzyme was tightly bound to an insoluble material (20000g pellet) and was solubilized by high concentrations of NaCl or KCl. High concentrations of them in the reaction system, however, inhibited reversibly the activity of this enzyme. 2. The neutral proteinase was partially purified by extraction with NaCl, gel filtration on Sephadex G-200 and affinity chromatography on casein–Sepharose. 3. The optimal pH of this partially purified enzyme was 7.4–8.0 against casein and Bz-dl-Arg-Nan. The molecular weight of the enzyme was found to be about 1.4×105 by gel filtration on Sephadex G-200. 4. The enzyme was significantly inhibited by di-isopropyl phosphorofluoridate (0.1mm). High concentration of phenylmethanesulphonyl fluoride (5mm), 7-amino-1-chloro-3-l-tosylamidoheptan-2-one (0.5mm), antipain (10μm) or leupeptin (10μm) was also found to be inhibitory, but chymostatin (40μg/ml), soya-bean trypsin inhibitor (2.5mg/ml), human plasma (10%, v/v), p-chloromercuribenzoate (1mm), EDTA (10mm) and 1-chloro-4-phenyl-3-l-tosylamidobutan-2-one (1mm) had no effect on the enzyme. 5. The neutral proteinase hydrolysed casein, Bz-dl-Arg-Nan and heat-denatured collagen, but was inactive towards native collagen and several synthetic substrates, such as 4-phenylazobenzyloxycarbonyl-Pro-Leu-Gly-Pro-d-Arg, 3-carboxypropionyl-Ala-Ala-Ala p-nitroanilide and 2,4-dinitrophenyl-Pro-Gln-Gly-Ile-Ala-Gly-Gln-d-Arg, and also proteoglycan. The enzyme did not act as a plasminogen activator. 6. These properties suggested that a neutral proteinase in the human uterine cervix was different from enzymes previously reported.  相似文献   

8.
Enzymatic processes are useful for industrially important sugar production, and in vitro two-step isomerization has proven to be an efficient process in utilizing readily available sugar sources. A hypothetical uncharacterized protein encoded by ydaE of Bacillus licheniformis was found to have broad substrate specificities and has shown high catalytic efficiency on d-lyxose, suggesting that the enzyme is d-lyxose isomerase. Escherichia coli BL21 expressing the recombinant protein, of 19.5 kDa, showed higher activity at 40 to 45°C and pH 7.5 to 8.0 in the presence of 1.0 mM Mn2+. The apparent Km values for d-lyxose and d-mannose were 30.4 ± 0.7 mM and 26 ± 0.8 mM, respectively. The catalytic efficiency (kcat/Km) for lyxose (3.2 ± 0.1 mM−1 s−1) was higher than that for d-mannose (1.6 mM−1 s−1). The purified protein was applied to the bioproduction of d-lyxose and d-glucose from d-xylose and d-mannose, respectively, along with the thermostable xylose isomerase of Thermus thermophilus HB08. From an initial concentration of 10 mM d-lyxose and d-mannose, 3.7 mM and 3.8 mM d-lyxose and d-glucose, respectively, were produced by two-step isomerization. This two-step isomerization is an easy method for in vitro catalysis and can be applied to industrial production.  相似文献   

9.
Isolation and characterization of ribosomal ribonucleic acid   总被引:5,自引:3,他引:2  
1. Ribosomal RNA has been prepared by extracting tissues with a phenol–cresol mixture, and ribosomal RNA can be selectively precipitated with m-cresol. No rapidly labelled RNA was associated with this material. 2. However, if RNA and DNA are extracted with 4-aminosalicylate and phenol–cresol mixture and the nucleic acids precipitated, DNA, glycogen and s-RNA (transfer RNA) can be extracted with 3m-sodium acetate and in this case rapidly labelled RNA remains associated with the ribosomal RNA. 3. The ribosomal RNA is stable in the presence of concentrated salt solution and, although the secondary structure is lost by heating at 70° in 10mm-sodium acetate, it can be re-formed in the presence of 200mm-sodium acetate. 4. The 28s and 18s components have been separated and their base compositions determined.  相似文献   

10.
Cytoglobin (Cygb) was investigated for its capacity to function as a NO dioxygenase (NOD) in vitro and in hepatocytes. Ascorbate and cytochrome b5 were found to support a high NOD activity. Cygb-NOD activity shows respective Km values for ascorbate, cytochrome b5, NO, and O2 of 0.25 mm, 0.3 μm, 40 nm, and ∼20 μm and achieves a kcat of 0.5 s−1. Ascorbate and cytochrome b5 reduce the oxidized Cygb-NOD intermediate with apparent second order rate constants of 1000 m−1 s−1 and 3 × 106 m−1 s−1, respectively. In rat hepatocytes engineered to express human Cygb, Cygb-NOD activity shows a similar kcat of 1.2 s−1, a Km(NO) of 40 nm, and a kcat/Km(NO) (kNOD) value of 3 × 107 m−1 s−1, demonstrating the efficiency of catalysis. NO inhibits the activity at [NO]/[O2] ratios >1:500 and limits catalytic turnover. The activity is competitively inhibited by CO, is slowly inactivated by cyanide, and is distinct from the microsomal NOD activity. Cygb-NOD provides protection to the NO-sensitive aconitase. The results define the NOD function of Cygb and demonstrate roles for ascorbate and cytochrome b5 as reductants.  相似文献   

11.
β-Galactosidase-catalysed hydrolysis of β-d-galactopyranosyl azide   总被引:3,自引:3,他引:0  
1. β-d-Galactopyranosyl azide is hydrolysed by the β-galactosidase of Escherichia coli to galactose and azide ion at a mechanistically significant rate. 2. Methyl 1-thio-β-d-galactopyranoside is a competitive inhibitor of the hydrolysis of the azide and of o-nitrophenyl β-d-galactopyranoside with Ki 1.8mm. 3. β-Galactosidase can thus hydrolyse a range of substrates of general structure β-d-galactopyranosyl-X(Y), where the atom X has a lone pair of electrons on which the enzyme may act as a Lewis or Brønsted acid, but in which the length of the bond cleaved varies significantly, which is inconsistent with the orbital steering hypothesis.  相似文献   

12.
1. Phosphomevalonate kinase and 5-pyrophosphomevalonate decarboxylase have been purified from the freeze-dried latex serum of the commercial rubber tree Hevea brasiliensis. 2. The phosphomevalonate kinase was acid- and heat-labile and required the presence of a thiol to maintain activity. 3. The 5-pyrophosphomevalonate decarboxylase was relatively acid-stable and more heat-stable than the phosphokinase. 4. Maximum activity of the phosphokinase was achieved at pH 7.2 with 0.2mm-5-phosphomevalonate (Km 0.042mm), 2.0mm-ATP (Km 0.19mm) and 8mm-Mg2+ at 40°C. The apparent activation energy was 14.8kcal/mol. 5. Maximum activity of 5-pyrophosphomevalonate decarboxylase was achieved at pH5.5–6.5 with 0.1mm-5-pyrophosphomevalonate (Km 0.004mm), 1.5mm-ATP (Km 0.12mm) and 2mm-Mg2+. The apparent activation energy was 13.7kcal/mol. The enzyme was somewhat sensitive to inhibition by its products, isopentenyl pyrophosphate and ADP.  相似文献   

13.
KIF3AB is an N-terminal processive kinesin-2 family member best known for its role in intraflagellar transport. There has been significant interest in KIF3AB in defining the key principles that underlie the processivity of KIF3AB in comparison with homodimeric processive kinesins. To define the ATPase mechanism and coordination of KIF3A and KIF3B stepping, a presteady-state kinetic analysis was pursued. For these studies, a truncated murine KIF3AB was generated. The results presented show that microtubule association was fast at 5.7 μm−1 s−1, followed by rate-limiting ADP release at 12.8 s−1. ATP binding at 7.5 μm−1 s−1 was followed by an ATP-promoted isomerization at 84 s−1 to form the intermediate poised for ATP hydrolysis, which then occurred at 33 s−1. ATP hydrolysis was required for dissociation of the microtubule·KIF3AB complex, which was observed at 22 s−1. The dissociation step showed an apparent affinity for ATP that was very weak (K½,ATP at 133 μm). Moreover, the linear fit of the initial ATP concentration dependence of the dissociation kinetics revealed an apparent second-order rate constant at 0.09 μm−1 s−1, which is inconsistent with fast ATP binding at 7.5 μm−1 s−1 and a Kd,ATP at 6.1 μm. These results suggest that ATP binding per se cannot account for the apparent weak K½,ATP at 133 μm. The steady-state ATPase Km,ATP, as well as the dissociation kinetics, reveal an unusual property of KIF3AB that is not yet well understood and also suggests that the mechanochemistry of KIF3AB is tuned somewhat differently from homodimeric processive kinesins.  相似文献   

14.
1. The aerobic transport of d-glucose and d-galactose in rabbit kidney tissue at 25° was studied. 2. In slices forming glucose from added substrates an accumulation of glucose against its concentration gradient was found. The apparent ratio of intracellular ([S]i) and extracellular ([S]o) glucose concentrations was increased by 0·4mm-phlorrhizin and 0·3mm-ouabain. 3. Slices and isolated renal tubules actively accumulated glucose from the saline; the apparent [S]i/[S]o fell below 1·0 only at [S]o higher than 0·5mm. 4. The rate of glucose oxidation by slices was characterized by the following parameters: Km 1·16mm; Vmax. 4·5μmoles/g. wet wt./hr. 5. The active accumulation of glucose from the saline was decreased by 0·1mm-2,4-dinitrophenol, 0·4mm-phlorrhizin and by the absence of external Na+. 6. The kinetic parameters of galactose entry into the cells were: Km 1·5mm; Vmax 10μmoles/g. wet wt./hr. 7. The efflux kinetics from slices indicated two intracellular compartments for d-galactose. The galactose efflux was greatly diminished at 0°, was inhibited by 0·4mm-phlorrhizin, but was insensitive to ouabain. 8. The following mechanism of glucose and galactose transport in renal tubular cells is suggested: (a) at the tubular membrane, these sugars are actively transported into the cells by a metabolically- and Na+-dependent phlorrhizin-sensitive mechanism; (b) at the basal cell membrane, these sugars are transported in accordance with their concentration gradient by a phlorrhizin-sensitive Na+-independent facilitated diffusion. The steady-state intracellular sugar concentration is determined by the kinetic parameters of active entry, passive outflow and intracellular utilization.  相似文献   

15.
We investigated 3,3′,5-tri-iodo-l-thyronine transport by human erythrocytes and by `ghosts'' prepared from these cells. Uptake of tri-iodothyronine by erythrocytes at 37°C was time-dependent with a maximum reached after 60min. Tracer analysis after incubation for 1min revealed only one saturable binding site, with Km 128±19nm (mean±s.e.m.; n=7) and Vmax. 4.6±0.7pmol of tri-iodothyronine/min per 6×107 cells. After 10min incubation Km 100±16nm (n=10) was found with Vmax. 7.7±1.2pmol of tri-iodothyronine/10min per 6×107 cells. At 0°C the uptake system is still active, with Km 132±26nm and Vmax. 1.8±0.3pmol of tri-iodothyronine/10min per 6×107 cells. The Vmax. with intact cells is 5-fold greater than the Vmax. with membranes derived from the same amount of cells when uptake studies are performed in media with similar free tri-iodothyronine concentrations. This indicates that at least 80% of tri-iodothyronine taken up by the intact erythrocytes enters the cell. This saturable uptake system can be inhibited by X-ray-contrast agents in a dose-dependent fashion. (±)-Propranolol, but not atenolol, has the same effect, indicating that the membrane-stabilizing properties of (±)-propranolol are involved. Furthermore, there is no inhibition by ouabain or vanadate, which indicates that tri-iodothyronine uptake is not dependent on the activity of Na++K+-dependent adenosine triphosphatase. We have prepared erythrocyte `ghosts'', resealed after 2.5min with 0mm-, 2mm- or 4mm-ATP inside. Inclusion of ATP and integrity of the membrane of the erythrocyte `ghosts'' were verified on the basis of an ATP-concentration-dependent functioning of the Ca2+ pump. No difference was found in the uptake of tri-iodothyronine by erythrocyte `ghosts'' with or without ATP included, indicating that uptake of tri-iodothyronine is not ATP-dependent. The following conclusions are drawn. (1) Tri-iodothyronine enters human erythrocytes. (2) There is only one saturable uptake system present for tri-iodothyronine, which is neither energy (i.e. ATP)-dependent nor influenced by the absence of an Na+ gradient across the plasma membrane. This mode of uptake of tri-iodothyronine by human erythrocytes is in sharp contrast with that of rat hepatocytes, which uptake system is energy-dependent and ouabain-sensitive [Krenning, Docter, Bernard, Visser & Hennemann (1978) FEBS Lett. 91, 113–116; Krenning, Docter, Bernard, Visser & Hennemann (1980) FEBS Lett. 119, 279–282]. (3) X-ray-contrast agents inhibit tri-iodothyronine uptake by erythrocytes in a similar fashion to that by which they inhibit the uptake of tri-iodothyronine by rat hepatocytes [Krenning, Docter, Bernard, Visser & Hennemann (1982) FEBS Lett. 140, 229–233].  相似文献   

16.
1. Diaminopimelate decarboxylase from a soluble extract of Escherichia coli A.T.C.C. 9637 was purified 200-fold by precipitation of nucleic acids, fractionation with acetone and then with ammonium sulphate, adsorption on calcium phosphate gel and chromatography on DEAE-cellulose or DEAE-Sephadex. 2. The purified enzyme showed only one component in the ultracentrifuge, with a sedimentation coefficient of 5·4s. One major peak and three much smaller peaks were observed on electrophoresis of the enzyme at pH8·9. 3. The mol.wt. of the enzyme was approx. 200000. The catalytic constant was 2000mol. of meso-diaminopimelic acid decomposed/min./mol. of enzyme, at 37°. The relative rates of decarboxylation at 25°, 37° and 45° were 0·17:1·0:1·6. At 37° the Michaelis constant was 1·7mm and the optimum pH was 6·7–6·8. 4. There was an excess of acidic amino acids over basic amino acids in the enzyme, which was bound only on basic cellulose derivatives at pH6·8. 5. The enzyme had an absolute requirement for pyridoxal phosphate as a cofactor; no other derivative of pyridoxine had activity. A thiol compound (of which 2,3-dimercaptopropan-1-ol was the most effective) was also needed as an activator. 6. In the presence of 2,3-dimercaptopropan-1-ol (1mm), heavy-metal ions (Cu2+, Hg2+) did not inhibit the enzyme, but there was inhibition by several amino acids with analogous structures to diaminopimelate, generally at high concentrations relative to the substrate. Penicillamine was inhibitory at relatively low concentrations; its action was prevented by pyridoxal phosphate.  相似文献   

17.
Yeast App1p is a phosphatidate phosphatase (PAP) that associates with endocytic proteins at cortical actin patches. App1p, which catalyzes the conversion of phosphatidate (PA) to diacylglycerol, is unique among Mg2+-dependent PAP enzymes in that its reaction is not involved with de novo lipid synthesis. Instead, App1p PAP is thought to play a role in endocytosis because its substrate and product facilitate membrane fission/fusion events and regulate enzymes that govern vesicular movement. App1p PAP was purified from yeast and characterized with respect to its enzymological, kinetic, and regulatory properties. Maximum PAP activity was dependent on Triton X-100 (20 mm), PA (2 mm), Mg2+ (0.5 mm), and 2-mercaptoethanol (10 mm) at pH 7.5 and 30 °C. Analysis of surface dilution kinetics with Triton X-100/PA-mixed micelles yielded constants for surface binding (KsA = 11 mm), interfacial PA binding (KmB = 4.2 mol %), and catalytic efficiency (Vmax = 557 μmol/min/mg). The activation energy, turnover number, and equilibrium constant were 16.5 kcal/mol, 406 s−1, and 16.2, respectively. PAP activity was stimulated by anionic lipids (cardiolipin, phosphatidylglycerol, phosphatidylserine, and CDP-diacylglycerol) and inhibited by zwitterionic (phosphatidylcholine and phosphatidylethanolamine) and cationic (sphinganine) lipids, nucleotides (ATP and CTP), N-ethylmaleimide, propranolol, phenylglyoxal, and divalent cations (Ca2+, Mn2+, and Zn2+). App1p also utilized diacylglycerol pyrophosphate and lyso-PA as substrates with specificity constants 4- and 7-fold lower, respectively, when compared with PA.  相似文献   

18.
Two major peaks of RNA polymerase activity have been routinely separated by diethylaminoethyl cellulose chromatography following solubilization from soybean (Glycine max L. var. Wayne) chromatin. The relative amounts of these two peaks depend upon the manner in which the chromatin is purified. Pelleting the chromatin through dense sucrose solutions results in not only a loss of total solubilized RNA polymerase activity but also a selective loss of the α-amanitin-sensitive form of the enzyme. Peak I elutes from a diethylaminoethyl cellulose column at a KCl concentration of approximately 0.27 m, is insensitive to α-amanitin and rifamycin, and has Mg2+ + Mn2+ optima of 5 mm and 1.25 mm, respectively. The enzyme is inhibited by KCl concentrations of about 0.03 m or greater. Peak II elutes from the column at a KCl concentration of approximately 0.35 m, is sensitive to α-amanitin, insensitive to rifamycin, and has Mg2+ + Mn2+ optima of 2 mm and 1.0 mm, respectively. Activity is inhibited by KCl concentrations of about 0.06 m or greater. Both enzymes prefer denatured calf thymus DNA, but peak II exhibits a stronger preference.  相似文献   

19.
1. all-trans-Retinoic acid at concentrations greater than 10−7m stimulated the incorporation of d-[3H]glucosamine into 8m-urea/5% (w/v) sodium dodecyl sulphate extracts of 1m-CaCl2-separated epidermis from pig ear skin slices cultured for 18h. The incorporation of 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected. 2. Electrophoresis of the solubilized epidermis showed increased incorporation of d-[3H]glucosamine into a high-molecular-weight glycosaminoglycan-containing peak when skin slices were cultured in the presence of 10−5m-all-trans-retinoic acid. The labelling of other epidermal components with d-[3H]glucosamine, 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected by 10−5m-all-trans-retinoic acid. 3. Trypsinization dispersed the epidermal cells and released 75–85% of the total d-[3H]glucosamine-labelled material in the glycosaminoglycan peak. Thus most of this material was extracellular in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 4. Increased labelling of extracellular epidermal glycosaminoglycans was also observed when human skin slices were treated with all-trans-retinoic acid, indicating a similar mechanism in both tissues. Increased labelling was also found when the epidermis was cultured in the absence of the dermis, suggesting a direct effect of all-trans-retinoic acid on the epidermis. 5. Increased incorporation of d-[3H]-glucosamine into extracellular epidermal glycosaminoglycans in 10−5m-all-trans-retinoic acid-treated skin slices was apparent after 4–8h in culture and continued up to 48h. all-trans-Retinoic acid (10−5m) did not affect the rate of degradation of this material in cultures `chased' with 5mm-unlabelled glucosamine after 4 or 18h. 6. Cellulose acetate electrophoresis at pH7.2 revealed that hyaluronic acid was the major labelled glycosaminoglycan (80–90%) in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 7. The labelling of epidermal plasma membranes isolated from d-[3H]glucosamine-labelled skin slices by sucrose density gradient centrifugation was similar in control and 10−5m-all-trans-retinoic acid-treated tissue. 8. The results indicate that increased synthesis of mainly extracellular glycosaminoglycans (largely hyaluronic acid) may be the first response of the epidermis to excess all-trans-retinoic acid.  相似文献   

20.
[3H]Triamcinolone acetonide (15nm) was incubated with cytosol (150000g fraction) prepared from oviducts of egg-laying hens. The extent of steroid binding, as determined by charcoal assays, was greatest between 2–4h at 4°C. A similar time curve was obtained when cytosol preparations were first fractionated with (NH4)2SO4 before labelling. The addition of 10mm-Na2MoO4 or 10mm-ATP during the incubation of hen oviduct cytosol with [3H]triamcinolone acetonide lowered the extent of steroid binding. The presence of glycerol (20%), however, increased the extent of [3H]triamcinolone acetonide binding in cytosol fractions from chick (330%) and hen (160%) oviducts. The [3H]triamcinolone acetonide–receptor complex was stable for over 4h at 4°C, but dissociated rapidly at 37°C, exhibiting a half-life of about 10min. The presence of 10mm-Na2MoO4 and 10mm-ATP or both had a small protective effect on the dissociation of [3H]triamcinolone acetonide–receptor complex. The receptor from hen oviduct showed significant affinity for unlabelled triamcinolone acetonide, cortisol, compound R5020 and dihydrotestosterone and, to a lesser extent, for oestradiol, oestrone and progesterone. Diethylstilboestrol treatment of immature chicks appeared to induce a more specific binder, which showed affinity for unlabelled triamcinolone acetonide, cortisol and compound R5020 only. Scatchard analysis of [3H]triamcinolone acetonide binding in hen oviduct cytosol revealed a Kd value of 6.4nm. The steroid–receptor complex sedimented as a 7–8S and a 4S entity on low-salt (0.01m-KCl)- and high-salt (0.3m-KCl)-containing sucrose gradients respectively. The cytosol [3H]triamcinolone acetonide–receptor complex showed no affinity for ATP–Sepharose or DNA–cellulose, but acquired this ability on heat activation (23°C, 40min). The data indicate the avian oviduct possesses a high-affinity binding molecule that fulfils the criteria of a glucocorticoid receptor.  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