首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
An NADP-dependent dehydrogenase catalyzing the conversion of l-sorbosone to l-ascorbic acid has been isolated from Phaseolus vulgaris L. and Spinacia oleracea L. and partially purified. It is stable at −20°C for up to 8 months. Molecular masses, as determined by gel filtration, were 21 and 29 kilodaltons for bean and spinach enzymes, respectively. Km for sorbosone were 12 ± 2 and 18 ± 2 millimolar and for NADP+, 0.14 ± 0.05 and 1.2 ± 0.5 millimolar, for bean and spinach, respectively. Lycorine, a purported inhibitor of l-ascorbic acid biosynthesis, had no effect on the reaction.  相似文献   

2.
Accumulation of d-leucine, d-allo-isoleucine, and d-valine was observed in the growth medium of a lactic acid bacterium, Lactobacillus otakiensis JCM 15040, and the racemase responsible was purified from the cells and identified. The N-terminal amino acid sequence of the purified enzyme was GKLDKASKLI, which is consistent with that of a putative γ-aminobutyrate aminotransferase from Lactobacillus buchneri. The putative γ-aminobutyrate aminotransferase gene from L. buchneri JCM 1115 was expressed in recombinant Escherichia coli and then purified to homogeneity. The enzyme catalyzed the racemization of a broad spectrum of nonpolar amino acids. In particular, it catalyzed at high rates the epimerization of l-isoleucine to d-allo-isoleucine and d-allo-isoleucine to l-isoleucine. In contrast, the enzyme showed no γ-aminobutyrate aminotransferase activity. The relative molecular masses of the subunit and native enzyme were estimated to be about 49 kDa and 200 kDa, respectively, indicating that the enzyme was composed of four subunits of equal molecular masses. The Km and Vmax values of the enzyme for l-isoleucine were 5.00 mM and 153 μmol·min−1·mg−1, respectively, and those for d-allo-isoleucine were 13.2 mM and 286 μmol·min−1·mg−1, respectively. Hydroxylamine and other inhibitors of pyridoxal 5′-phosphate-dependent enzymes completely blocked the enzyme activity, indicating the enzyme requires pyridoxal 5′-phosphate as a coenzyme. This is the first evidence of an amino acid racemase that specifically catalyzes racemization of nonpolar amino acids at the C-2 position.  相似文献   

3.
A fourth molecular from of α-galactosidase, designated LIV, an alkaline α-galactosidase, was isolated from leaves of Cucurbita pepo and purified 165-fold. It was active over a narrow pH range with optimal hydrolysis of p-nitrophenyl-α-d-galactoside and stachyose at pH 7.5. The rate of stachyose hydrolysis was 10 times that of raffinose. Km determinations in McIlvaine buffer (200 millimolar Na2-phosphate, 100 millimolar citric acid, pH 7.5) for p-nitrophenyl-α-d-galactoside, stachyose, and raffinose were 1.40, 4.5, and 36.4 millimolar, respectively. LIV was partially inhibited by Ca2+, Mg2+, and Mn2+, more so by Ni2+, Zn2+, and Co2+, and highly so by Cu2+, Ag2+, Hg2+ and by p-chloromercuribenzoate. It was not inhibited by high concentrations of the substrate p-nitrophenyl-α-d-galactoside or by myo-inositol, but α-d-galactose was a strong inhibitor. As observed for most other forms of α-galactosidase, LIV only catalyzed the hydrolysis of glycosides possessing the α-d-galactose configuration at C1, C2, and C4, and did not hydrolyze p-nitrophenyl-α-d-fucoside (α-d-galactose substituted at C6). The enzyme was highly sensitive to buffers and chelating agents. Maximum hydrolytic activity for p-nitrophenyl-α-d-galactoside was obtained in McIlvaine buffer (pH 7.5). In 10 millimolar triethanolaminehydrochloride-NaOH (pH 7.5) or 10 millimolar Hepes-NaOH (pH 7.5), hydrolytic activity was virtually eliminated, but the addition of low concentrations of either ethylenediaminetetraacetate or citrate to these buffers restored activity almost completely. Partial restoration of activity was also observed, but at higher concentrations, with pyruvate and malate. Similar effects were found for stachyose hydrolysis, but in addition some inhibition of LIV in McIlvaine buffer, possibly due to the high phosphate concentration, was observed with this substrate. It is questionable whether the organic acid anions possess any regulatory control of LIVin vivo. It was possible that the results reflected the ability of these anions, and ethylene-diaminetetraacetate, to restore LIV activity through coordination with some toxic cation introduced as a buffer contaminant.  相似文献   

4.
δ-Aminolevulinic acid (ALA), a key precursor of the tetrapyrroles heme and chlorophyll, is capable of being synthesized by two different routes in cells of the unicellular green alga Euglena gracilis: from the intact carbon skeleton of glutamate, and via the condensation of glycine and succinyl CoA, mediated by the enzyme ALA synthase. The regulatory properties of ALA synthase were examined in order to establish its role in Euglena.

Partially purified Euglena ALA synthase, unlike the case with the bacterial or animal-derived enzyme, does not exhibit allosteric inhibition by the tetrapyrrole pathway products heme, protoporphyrin IX, and porphobilinogen, at concentrations up to 100 micromolar.

In aplastidic mutant cells, extractable ALA synthase activity is constant during exponential growth, and decreases to low levels as the cells reach the stationary state. Rapid exponential decline of ALA synthase (t1/2 = 55 min) occurs after administration of 43 micromolar cycloheximide, but not 6.2 millimolar chloramphenicol. These results suggest that, as in other eukaryotic cells, ALA synthase is synthesized on cytoplasmic ribosomes and is subject to rapid turnover in vivo.

Extractable ALA synthase activity increases 2.5-fold within 6 hours after administration of 100 millimolar ethanol, a stimulator of mitochondrial development, and 4.5-fold within 12 hours after administration of 1 millimolar 4,6-dioxoheptanoic acid, which blocks ALA utilization, suggesting that activity is controlled in vivo by a feedback induction-repression mechanism, coupled with rapid enzyme turnover.

In heterotrophically grown wild-type cells, low levels of ALA synthase rapidly increase 4.5-fold within 12 hours after cells are transferred from the light to the dark, and decrease exponentially (t1/2 = 75 min) when cells are transferred from the dark to light. The dark levels are equal to those in light- or dark-grown aplastidic mutant cells. The low level occurring in light-grown wild-type cells is not altered by the presence of 10 micromolar 3-(3,4-dichlorophenyl)-1,1-dimethylurea, which blocks photosynthetic O2 production. The decrease that occurs on dark-to-light transfer can be diminished by 12- or 24-hour prior incubation with 6.2 millimolar chloramphenicol, which also retards chlorophyll synthesis after the transfer to light.

The positive relationship of ALA synthase activity to degree of mitochondrial expression, and the inverse relationship to plastid development and chlorophyll synthesis, suggests that ALA synthase functions to provide precursors to nonplastid tetrapyrroles in Euglena. In light-grown, wild-type cells, the diminished levels of ALA synthase may be due to the ability of developing plastids to export heme or a heme precursor to other cellular regions, which thereby supplants the necessity for ALA formation via the ALA synthase route.

  相似文献   

5.
An enzyme catalyzing the formation of δ-aminolevulinic acid by transamination of γ,δ-dioxovaleric acid with l-α-alanine, l-glutamic acid, or l-phenylalanine has been detected in extracts of Chlorella vulgaris. The activity of this enzyme does not appear to parallel changes in chlorophyll content in a Chlorella mutant which requires light for chlorophyll production. The role of this enzyme in δ-aminolevulinic acid metabolism in plants is not clearly understood.  相似文献   

6.
Beale SI  Foley T 《Plant physiology》1982,69(6):1331-1333
N-Methyl mesoporphyrin IX, an inhibitor of heme synthesis, increases extractable δ-aminolevulinic acid (ALA) synthase activity when administered to growing cultures of Euglena gracilis Klebs strain Z Pringsheim in micromolar concentrations. Wild-type light-grown green cells and white aplastidic cells exhibited 2.8-fold and 1.8-fold increases, respectively, in ALA synthase activity within five to six hours after incubation with 4 × 10−6 molar N-methyl mesoporphyrin IX. Protoheme levels were decreased and 59Fe incorporation into heme was inhibited by N-methyl mesoporphyrin IX, indicating that, as in animal cells, N-methyl mesoporphyrin IX acts specifically to block iron insertion into protoporphyrin IX. Chlorophyll synthesis in wild-type cells was not affected within the first 6 hours after administration of N-methyl mesoporphyrin IX.  相似文献   

7.
1. The formation of adenosine 5′-phosphate, guanosine 5′-phosphate and inosine 5′-phosphate from [8-14C]adenine, [8-14C]guanine and [8-14C]hypoxanthine respectively in the presence of 5-phosphoribosyl pyrophosphate and an extract from Ehrlich ascites-tumour cells was assayed by a method involving liquid-scintillation counting of the radioactive nucleotides on diethylaminoethylcellulose paper. The results obtained with guanine were confirmed by a spectrophotometric assay which was also used to assay the conversion of 6-mercaptopurine and 5-phosphoribosyl pyrophosphate into 6-thioinosine 5′-phosphate in the presence of 6-mercaptopurine phosphoribosyltransferase from these cells. 2. At pH 7·8 and 25° the Michaelis constants for adenine, guanine and hypoxanthine were 0·9 μm, 2·9 μm and 11·0 μm in the assay with radioactive purines; the Michaelis constant for guanine in the spectrophotometric assay was 2·6 μm. At pH 7·9 the Michaelis constant for 6-mercaptopurine was 10·9 μm. 3. 25 μm-6-Mercaptopurine did not inhibit adenine phosphoribosyltransferase. 6-Mercaptopurine is a competitive inhibitor of guanine phosphoribosyltransferase (Ki 4·7 μm) and hypoxanthine phosphoribosyltransferase (Ki 8·3 μm). Hypoxanthine is a competitive inhibitor of guanine phosphoribosyltransferase (Ki 3·4 μm). 4. Differences in kinetic parameters and in the distribution of phosphoribosyltransferase activities after electrophoresis in starch gel indicate that different enzymes are involved in the conversion of adenine, guanine and hypoxanthine into their nucleotides. 5. From the low values of Ki for 6-mercaptopurine, and from published evidence that ascites-tumour cells require supplies of purines from the host tissues, it is likely that inhibition of hypoxanthine and guanine phosphoribosyltransferases by free 6-mercaptopurine is involved in the biological activity of this drug.  相似文献   

8.
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   

9.
Streptococcus intermedius is a known human pathogen and belongs to the anginosus group (S. anginosus, S. intermedius, and S. constellatus) of streptococci (AGS). We found a large open reading frame (6,708 bp) in the lac operon, and bioinformatic analysis suggested that this gene encodes a novel glycosidase that can exhibit β-d-galactosidase and N-acetyl-β-d-hexosaminidase activities. We, therefore, named this protein “multisubstrate glycosidase A” (MsgA). To test whether MsgA has these glycosidase activities, the msgA gene was disrupted in S. intermedius. The msgA-deficient mutant no longer showed cell- and supernatant-associated β-d-galactosidase, β-d-fucosidase, N-acetyl-β-d-glucosaminidase, and N-acetyl-β-d-galactosaminidase activities, and all phenotypes were complemented in trans with a recombinant plasmid carrying msgA. Purified MsgA had all four of these glycosidase activities and exhibited the lowest Km with 4-methylumbelliferyl-linked N-acetyl-β-d-glucosaminide and the highest kcat with 4-methylumbelliferyl-linked β-d-galactopyranoside. In addition, the purified LacZ domain of MsgA had β-d-galactosidase and β-d-fucosidase activities, and the GH20 domain exhibited both N-acetyl-β-d-glucosaminidase and N-acetyl-β-d-galactosaminidase activities. The β-d-galactosidase and β-d-fucosidase activities of MsgA are thermolabile, and the optimal temperature of the reaction was 40°C, whereas almost all enzymatic activities disappeared at 49°C. The optimal temperatures for the N-acetyl-β-d-glucosaminidase and N-acetyl-β-d-galactosaminidase activities were 58 and 55°C, respectively. The requirement of sialidase treatment to remove sialic acid residues of the glycan branch end for glycan degradation by MsgA on human α1-antitrypsin indicates that MsgA has exoglycosidase activities. MsgA and sialidase might have an important function in the production and utilization of monosaccharides from oligosaccharides, such as glycans for survival in a normal habitat and for pathogenicity of S. intermedius.  相似文献   

10.
The uptake of phenylalanine was studied with vacuole isolated from barley mesophyll protoplasts. The phenylalanine transport exhibited saturation kinetics with apparent Km-values of 1.2 to 1.4 millimolar for ATP- or PPi-driven uptake; Vmax app was 120 to 140 nanomoles Phe per milligram of chlorophyll per hour (1 milligram of chlorophyll corresponds to 5 × 106 vacuoles). Half-maximal transport rates driven with ATP or PPi were reached at 0.5 millimolar ATP or 0.25 millimolar PPi. ATP-driven transport showed a distinct pH optimum at 7.3 while PPi-driven transport reached maximum rates at pH 7.8. Direct measurement of the H+-translocating enzyme activities revealed Km app values of 0.45 millimolar for ATPase (EC 3.6.1.3) and 23 micromolar for pyrophosphatase (PPase) (EC 3.6.1.1). In contrast to the coupled amino acid transport, ATPase and PPase activities had relative broad pH optima between 7 to 8 for ATPase and 8 to 9 for PPase. ATPase as well as ATP-driven transport was markedly inhibited by nitrate while PPase and PPi-coupled transport was not affected. The addition of ionophores inhibited phenylalanine transport suggesting the destruction of the electrochemical proton potential difference Δ μH+ while the rate of ATP and PPi hydrolysis was stimulated. The uptake of other lipophilic amino acids like l-Trp, l-Leu, and l-Tyr was also stimulated by ATP. They seem to compete for the same carrier system. l-Ala, l-Val, d-Phe, and d-Leu did not influence phenylalanine transport suggesting a stereospecificity of the carrier system for l-amino acids having a relatively high hydrophobicity.  相似文献   

11.
The metabolism of protocatechuate by Pseudomonas testosteroni   总被引:5,自引:5,他引:0  
1. Protocatechuate 4,5-oxygenase, purified 21-fold from extracts of Pseudomonas testosteroni, was examined in the ultracentrifuge and assigned a mol.wt. of about 140000. 2. When diluted, the enzyme rapidly lost activity during catalysis. Inactivation was partially prevented by l-cysteine. 3. With a saturating concentration of protocatechuate (1·36mm), Km for oxygen was 0·303mm. This value is greater than the concentration of oxygen in water saturated with air at 20°. 4. Cell extracts converted protocatechuate into γ-carboxy-γ-hydroxy-α-oxovalerate, which was isolated as its lactone. 5. γ-Carboxy-γ-hydroxy-α-oxovalerate pyruvate-lyase activity was stimulated by Mg2+ ions and mercaptoethanol. Cells grown with p-hydroxybenzoate as carbon source contained higher concentrations of this enzyme than those grown with succinate.  相似文献   

12.
Three of the four kynurenine aminotransferases (KAT I, II, and IV) that synthesize kynurenic acid, a neuromodulator, are identical to glutamine transaminase K (GTK), α-aminoadipate aminotransferase, and mitochondrial aspartate aminotransferase, respectively. GTK/KAT I and aspartate aminotransferase/KAT IV possess cysteine S-conjugate β-lyase activity. The gene for the former enzyme, GTK/KAT I, is listed in mammalian genome data banks as CCBL1 (cysteine conjugate beta-lyase 1). Also listed, despite the fact that no β-lyase activity has been assigned to the encoded protein in the genome data bank, is a CCBL2 (synonym KAT III). We show that human KAT III/CCBL2 possesses cysteine S-conjugate β-lyase activity, as does mouse KAT II. Thus, depending on the nature of the substrate, all four KATs possess cysteine S-conjugate β-lyase activity. These present studies show that KAT III and glutamine transaminase L are identical enzymes. This report also shows that KAT I, II, and III differ in their ability to transaminate methyl-l-selenocysteine (MSC) and l-selenomethionine (SM) to β-methylselenopyruvate (MSP) and α-ketomethylselenobutyrate, respectively. Previous studies have identified these seleno-α-keto acids as potent histone deacetylase inhibitors. Methylselenol (CH3SeH), also purported to have chemopreventive properties, is the γ-elimination product of SM and the β-elimination product of MSC catalyzed by cystathionine γ-lyase (γ-cystathionase). KAT I, II, and III, in part, can catalyze β-elimination reactions with MSC generating CH3SeH. Thus, the anticancer efficacy of MSC and SM will depend, in part, on the endogenous expression of various KAT enzymes and cystathionine γ-lyase present in target tissue coupled with the ability of cells to synthesize in situ either CH3SeH and/or seleno-keto acid metabolites.  相似文献   

13.
α-dl-Difluoromethylornithine (DFMO), a specific enzyme-activated inhibitor of ornithine decarboxylase, at 0.5 to 2.0 millimolar significantly inhibited mycelial growth and especially sporulation of Helminthosporium maydis in the dark; its inhibitory effect on sporulation was greatly increased under light conditions. Putrescine at 0.25 millimolar fully prevented the inhibitory effects of DFMO; the inhibition caused by the latter could not be prevented by cadaverine or CaCl2. α-dl-Difluoromethylarginine, a specific enzyme-activated inhibitor of arginine decarboxylase, at 0.1 to 2.0 millimolar had a weak inhibitory effect on the fungus. The effect was not dependent on the inhibitor concentration and there was no detectable arginine decarboxylase activity in the fungus.  相似文献   

14.
The composition of the cell wall of Fusicoccum amygdali   总被引:1,自引:1,他引:0       下载免费PDF全文
1. The cell wall of Fusicoccum amygdali consisted of polysaccharides (85%), protein (4–6%), lipid (5%) and phosphorus (0.1%). 2. The main carbohydrate constituent was d-glucose; smaller amounts of d-glucosamine, d-galactose, d-mannose, l-rhamnose, xylose and arabinose were also identified, and 16 common amino acids were detected. 3. Chitin, which accounted for most of the cell-wall glucosamine, was isolated in an undegraded form by an enzymic method. Chitosan was not detected, but traces of glucosamine were found in alkali-soluble and water-soluble fractions. 4. Cell walls were stained dark blue by iodine and were attacked by α-amylase, with liberation of glucose, maltose and maltotriose, indicating the existence of chains of α-(1→4)-linked glucopyranose residues. 5. Glucose and gentiobiose were liberated from cell walls by the action of an exo-β-(1→3)-glucanase, giving evidence for both β-(1→3)- and β-(1→6)-glucopyranose linkages. 6. Incubation of cell walls with Helix pomatia digestive enzymes released glucose, N-acetyl-d-glucosamine and a non-diffusible fraction, containing most of the cell-wall galactose, mannose and rhamnose. Part of this fraction was released by incubating cell walls with Pronase; acid hydrolysis yielded galactose 6-phosphate and small amounts of mannose 6-phosphate and glucose 6-phosphate as well as other materials. Extracellular polysaccharides of a similar nature were isolated and may be formed by the action of lytic enzymes on the cell wall. 7. About 30% of the cell wall was resistant to the action of the H. pomatia digestive enzymes; the resistant fraction was shown to be a predominantly α-(1→3)-glucan. 8. Fractionation of the cell-wall complex with 1m-sodium hydroxide gave three principal glucan fractions: fraction BB had [α]D +236° (in 1m-sodium hydroxide) and showed two components on sedimentation analysis; fraction AA2 had [α]D −71° (in 1m-sodium hydroxide) and contained predominantly β-linkages; fraction AA1 had [α]D +40° (in 1m-sodium hydroxide) and may contain both α- and β-linkages.  相似文献   

15.
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   

16.
1. Cholesteryl 3β-sulphate is oxidized in vitro by preparations of bovine adrenal-cortex mitochondria to pregnenolone sulphate and isocaproic acid (4-methyl-pentanoic acid) without hydrolysis of the ester linkage. 2. Free cholesterol is the preferred substrate for adrenal-cortex cholesterol oxidase; the apparent Km for cholesteryl sulphate is 500μm and for free cholesterol 50μm under the same conditions. 3. Cholesteryl 3β-acetate is hydrolysed by bovine adrenal-cortex mitochondria in vitro to free cholesterol, which is subsequently oxidized to more polar steroids and isocaproic acid. Evidence was obtained that other cholesterol esters behave similarly. Cholesterol esters may thus act as precursors of steroid hormones. 4. Cholest-4-en-3-one is only poorly oxidized to isocaproic acid and more polar steroids and thus is probably not a significant precursor of steroid hormones. 5. Cholesteryl esters inhibit the oxidation of cholesterol competitively (Ki for cholesteryl phosphate 28μm, for cholesteryl sulphate 110μm, for cholesteryl acetate 65μm) but pregnenolone esters do not inhibit this system. 6. Pregnenolone and 20α-hydroxycholesterol (both metabolites of cholesterol in this system) inhibit the oxidation of cholesterol non-competitively. Ki for pregnenolone is 130μm and Ki for 20α-hydroxycholesterol is 17μm. 7. 25-Oxo-27-norcholesterol inhibits cholesterol oxidation non-competitively (Ki16μm). A number of other Δ5-3β-hydroxy steroids inhibit cholesterol oxidation and evidence was obtained that the 3β-hydroxyl group was necessary for inhibitory activity. 8. Pregnenolone, 20α-hydroxycholesterol and 25-oxo-27-norcholesterol inhibit oxidation of cholesteryl sulphate by this system but their sulphates do not. 9. 3β-Hydroxychol-5-enoic acid, 3α-hydroxy-5β-cholanic acid and 3β-hydroxy-22,23-bisnorchol-5-enoic acid stimulated formation of isocaproic acid from cholesterol. 10. No evidence was obtained that phosphorylation or sulphation are obligatory steps in cholesterol oxidation by adrenal-cortex mitochondria. 11. The cholesteryl 3β-sulphate sulphatase of bovine adrenal cortex was found mostly in the microsomal fraction and was inhibited by inorganic phosphate.  相似文献   

17.
Lipase (triacylglycerol acylhydrolase [EC 3.1.1.3.]) was extracted from the microsomal fraction of cotyledons of dark grown seedlings of Canola (Brassica napus L. cv Westar) by treatment with Triton X-100. The enzyme was partially purified by chromatography on Sephacryl S-300 and DEAE Bio-Gel and was stable when stored at −20°C in 50% (v/v) glycerol. The lipase aggregated readily but the distribution of species present in solution could be controlled by nonionic detergents. A species with an apparent Mr of about 250,000 was obtained by gel filtration chromatography in the presence of 1% (v/v) Triton X-100. Lipase activity was optimal near neutral pH, and the reaction approached maximum velocity at a concentration of 0.5 to 1 millimolar emulsified triolein. The reaction rate responded linearly to temperature up to about 40°C and the hydrolytic process had an activation energy of 18 kilocalories per mole. Microsomal lipase lost about 20% and 80% activity when heat-treated for 1 hour at 40°C and 60°C, respectively. At appropriate concentrations, the detergents Triton X-100, n-octyl-β-d-glucopyranoside, (3-[(3-cholamidopropyl-O-dimethylammonio]-1-propanesulfonate, cetyl trimethylammonium bromide, and sodium dodecyl sulfate all inhibited lipase activity. n-Octyl-β-d-glucopyranoside, however, was stimulatory in the 2 to 8 millimolar concentration range. The inhibitory effects of Triton X-100 were reversible.  相似文献   

18.
Activated factor XII (FXIIa) is selectively inhibited by corn Hageman factor inhibitor (CHFI) among other plasma proteases. CHFI is considered a canonical serine protease inhibitor that interacts with FXIIa through its protease-binding loop. Here we examined whether the protease-binding loop alone is sufficient for the selective inhibition of serine proteases or whether other regions of a canonical inhibitor are involved. Six CHFI mutants lacking different N- and C-terminal portions were generated. CHFI-234, which lacks the first and fifth disulfide bonds and 11 and 19 amino acid residues at the N and C termini, respectively, exhibited no significant changes in FXIIa inhibition (Ki = 3.2 ± 0.4 nm). CHFI-123, which lacks 34 amino acid residues at the C terminus and the fourth and fifth disulfide bridges, inhibited FXIIa with a Ki of 116 ± 16 nm. To exclude interactions outside the FXIIa active site, a synthetic cyclic peptide was tested. The peptide contained residues 20–45 (Protein Data Bank code 1BEA), and a C29D substitution was included to avoid unwanted disulfide bond formation between unpaired cysteines. Surprisingly, the isolated protease-binding loop failed to inhibit FXIIa but retained partial inhibition of trypsin (Ki = 11.7 ± 1.2 μm) and activated factor XI (Ki = 94 ± 11 μm). Full-length CHFI inhibited trypsin with a Ki of 1.3 ± 0.2 nm and activated factor XI with a Ki of 5.4 ± 0.2 μm. Our results suggest that the protease-binding loop is not sufficient for the interaction between FXIIa and CHFI; other regions of the inhibitor also contribute to specific inhibition.  相似文献   

19.
Gibberellic acid (GA3) induces invertase activity within 6 hours in Avena stem segments that are incubated in the dark at 23°. The maximum amount of promotion is about 5 times that of invertase activity in untreated segments. GA3 causes significant promotion of invertase activity at concentrations as low as 3 × 10−5 μm GA3. The increase in invertase activity elicited by GA3 between 3 × 10−5 μm and 300 μm closely parallels the growth promotion that is caused by GA3 over this concentration range. In control segments, invertase activity rises steeply during the first 6 hours of incubation, then decays slowly between 12 and 48 hours. In GA3-treated segments, the invertase activity also rises during the first 6 hours, parallel to that in control segments and continues to rise during the next 42 hours. These changes in invertase activity during 48-hour incubation periods do not parallel the changes in growth that occur in control and GA3-treated segments. Cycloheximide at 10 μg/ml abolishes all GA3-promoted growth and invertase activity in these segments. Actinomycin D at 40 and 80 μg/ml decreases GA3-promoted growth by 20% and invertase activity by 38 and 44%, respectively. The data clearly support the idea that protein synthesis is necessary for GA3-promoted growth and invertase activity in Avena stem segments.  相似文献   

20.
1. The total activity of adenine phosphoribosyltransferase/liver of mice remained constant from 1 to 16 days after birth despite a fourfold increase in liver weight. The total activity of this enzyme increased fivefold from 16 to 36 days and then remained relatively constant at least until 96 days after birth. Total hypoxanthine-phosphoribosyltransferase activity/liver steadily increased between 1 and 57 days after birth. 2. The mean Km of 5-phosphoribosyl pyrophosphate with adenine phosphoribosyltransferase was 10·1μm between 3 and 11 days, at 64 days and at 96 days after birth. Between 17 and 51 days the mean Km value was 3·0μm. The Km of 5-phosphoribosyl pyrophosphate with hypoxanthine phosphoribosyltransferase remained constant at 28·2μm between 2 and 64 days. 3. Adenine-phosphoribosyltransferase activity was stimulated between 15 and 83% by 60μm-ATP when extracts were made between 3 and 11 days, at 64 days or at 96 days after birth. Between 17 and 51 days ATP had little stimulatory effect on the activity of this enzyme. 4. AMP competed with 5-phosphoribosyl pyrophosphate in the reaction catalysed by adenine phosphoribosyltransferase. Liver extracts containing enzyme with a low value of Km for 5-phosphoribosyl pyrophosphate (3μm) had a Km/Ki ratio approximately half that of extracts with a high value of Km (10μm). 5. The results indicate that two different forms of adenine phosphoribosyltransferase can exist in mouse liver at different stages of development. The physiological significance of these findings is discussed.  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