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.
1. Rat-liver supernatant preparations are capable of achieving the biological sulphation of l-tyrosine methyl ester, the reaction proceeding maximally at a substrate concentration of 30 mm and at pH 7·0. 2. Two sulphated products are formed, one of which has been identified as l-tyrosine O-sulphate. On the basis of indirect evidence the other product can be assumed to be l-tyrosine O-sulphate methyl ester. 3. An enzyme present in rat-liver supernatant preparations is capable of converting l-tyrosine O-sulphate methyl ester into l-tyrosine O-sulphate. This enzyme is inhibited by l-tyrosine methyl ester. 4. l-Tyrosine ethyl ester also yields two sulphated products when used as an acceptor in the liver sulphating system. One of these has been identified chromatographically as l-tyrosine O-sulphate and the other may be presumed to be l-tyrosine O-sulphate ethyl ester.  相似文献   

2.
The first enzyme in the pathway for l-arabinose catabolism in eukaryotic microorganisms is a reductase, reducing l-arabinose to l-arabitol. The enzymes catalyzing this reduction are in general nonspecific and would also reduce d-xylose to xylitol, the first step in eukaryotic d-xylose catabolism. It is not clear whether microorganisms use different enzymes depending on the carbon source. Here we show that Aspergillus niger makes use of two different enzymes. We identified, cloned, and characterized an l-arabinose reductase, larA, that is different from the d-xylose reductase, xyrA. The larA is up-regulated on l-arabinose, while the xyrA is up-regulated on d-xylose. There is however an initial up-regulation of larA also on d-xylose but that fades away after about 4 h. The deletion of the larA gene in A. niger results in a slow growth phenotype on l-arabinose, whereas the growth on d-xylose is unaffected. The l-arabinose reductase can convert l-arabinose and d-xylose to their corresponding sugar alcohols but has a higher affinity for l-arabinose. The Km for l-arabinose is 54 ± 6 mm and for d-xylose 155 ± 15 mm.  相似文献   

3.

Background and Aims

Nitric oxide (NO) has been demonstrated to stimulate the activity of nitrate reductase (NR) in plant roots supplied with a low level of nitrate, and to affect proteins differently, depending on the ratio of NO to the level of protein. Nitrate has been suggested to regulate the level of NO in plants. This present study examined interactive effects of NO and nitrate level on NR activity in roots of tomato (Solanum lycocarpum).

Methods

NR activity, mRNA level of NR gene and concentration of NR protein in roots fed with 0·5 mm or 5 mm nitrate and treated with the NO donors, sodium nitroprusside (SNP) and diethylamine NONOate sodium (NONOate), and the NO scavenger, 2-(4-carboxyphenyl)-4,4,5,5-tetramethyl-imidazoline-1-oxyl-3-oxide (cPTIO), were measured in 25-d-old seedlings.

Key Results

Addition of SNP and NONOate enhanced but cPTIO decreased NR activity in the roots fed with 0·5 mm nitrate. The opposite was true for the roots fed with 5 mm nitrate. However, the mRNA level of the NR gene and the protein concentration of NR enzyme in the roots were not affected by SNP treatment, irrespective of nitrate pre-treatment. Nevertheless, a low rate of NO gas increased while cPTIO decreased the NR activities of the enzyme extracts from the roots at both nitrate levels. Increasing the rate of NO gas further increased NR activity in the enzyme extracts of the roots fed with 0·5 mm nitrate but decreased it when 5 mm nitrate was supplied. Interestingly, the stimulative effect of NO gas on NR activity could be reversed by NO removal through N2 flushing in the enzyme extracts from the roots fed with 0·5 mm nitrate but not from those with 5 mm nitrate.

Conclusions

The effects of NO on NR activity in tomato roots depend on levels of nitrate supply, and probably result from direct interactions between NO and NR protein.Key words: Nitric oxide, nitrate, nitrate reductase, post-translational regulation, tomato, Solanum lycocarpum  相似文献   

4.
Endogenous respiration of spores of the fungus Myrothecium verrucaria can be stimulated up to over-10 fold by diverse chemicals or by physical treatments. Greatest effects were caused by azide (12-fold at 250 μm) and by 2,4-dinitrophenol (7-fold at 300 μm). Marked stimulation was also caused by 10 μm silver (5-fold), 30 μm pentachlorophenol (6-fold), 10 μm carbonyl cyanide m-chlorophenyl hydrazone (4.5-fold) and 10 μm merthiolate (4-fold). Physical treatments such as heat (50 C), freezing, and sonication at sublethal levels were also stimulatory. Stimulation by azide or dinitrophenol was much greater in young than in old spores, whereas response to other chemicals and to freezing was relatively unaffected by spore age. In older spores the effect of azide was no greater than some other inhibitors. During incubation with azide, the endogenous trehalose reserves decreased and changes in free amino acids occurred, both increases and decreases. Thus anabolic as well as catabolic changes occur as evidenced also by the germination of a few (up to 5%) spores. The mechanisms of stimulation must be varied and complex. Permeability changes in the membrane confining endogenous reserves are proposed as a common initial cause. Additional changes in characteristics of membranes of other subcellular particles, as well as enzymic phenomena such as uncoupling of oxidative phosphorylation, are presumably involved in instances where greater stimulation occurs. The data are consistent with the hypothesis that dormancy in these spores results from separation of substrates from metabolic enzymes and more specifically that metabolites are sequestered rather than enzymes.  相似文献   

5.
l-Glutamine d-fructose 6-phosphate amidotransferase (EC 2.6.1.16) was extracted and purified 600-fold by acetone fractionation and diethylaminoethyl cellulose column chromatography from mung bean seeds (Phaseolus aureus). The partially purified enzyme was highly specific for l-glutamine as an amide nitrogen donor, and l-asparagine could not replace it. The enzyme showed a pH optimum in the range of 6.2 to 6.7 in phosphate buffer. Km values of 3.8 mm and 0.5 mm were obtained for d-fructose 6-phosphate and l-glutamine, respectively. The enzyme was competitively inhibited with respect to d-fructose 6-phosphate by uridine diphosphate-N-acetyl-d-glucosamine which had a Ki value of 13 μm. Upon removal of l-glutamine and its replacement by d-fructose 6-phosphate and storage over liquid nitrogen, the enzyme was completely desensitized to inhibition by uridine diphosphate-N-acetyl-d-glucosamine. This indicates that the inhibitor site is distinct from the catalytic site and that uridine diphosphate-N-acetyl-d-glucosamine acts as a feedback inhibitor of the enzyme.  相似文献   

6.
1. Aminoacyl-transfer-RNA synthetase activity in extracts prepared from tobacco leaf was increased 3–5-fold when sodium thioglycollate (30mm) and magnesium chloride (16mm) were included in the extraction medium. Omitting sucrose (0·45m) from the extraction medium did not alter the activity. 2. Activity was a linear function of enzyme concentration up to 1 disk (30mg. fresh wt.)/ml. and was not affected by dialysis at any concentration. 3. Activity increased about 13-fold above control values when a mixture of 21 amino acids and amides (1mm) was added to the reaction mixture. 4. Under the conditions used in the standard assay for aminoacyl-transfer-RNA synthetase activity Km (ATP) was 0·65mm and Km (l-amino acids) was 70μm. 5. Activity above the control value was found with all amino acids and amides tested except alanine, arginine, glutamic acid, glutamine and hydroxyproline. Activity was highest with leucine, isoleucine, valine, cysteine and histidine. Total activity with a mixture of 21 amino acids and amides was 20% lower than the total activity of the enzymes assayed separately.  相似文献   

7.
The biological sulphation of l-tyrosyl peptides   总被引:3,自引:3,他引:0       下载免费PDF全文
1. A rat-liver supernatant preparation can achieve the biological O-sulphation of l-tyrosylglycine and l-tyrosyl-l-alanine at pH7·0. 2. The optimum concentrations of l-tyrosylglycine and l-tyrosyl-l-alanine in this system are 50mm and 60mm respectively. 3. l-Tyrosylglycine yields two sulphated products, whereas l-tyrosyl-l-alanine yields three sulphated products, when used as acceptor for sulphate in the rat-liver system. 4. With both substrates, one of the sulphated products has been identified as the O-sulphate ester of the corresponding parent peptide.  相似文献   

8.
The fuel of respiration of rat kidney cortex   总被引:15,自引:15,他引:0       下载免费PDF全文
1. In kidney-cortex slices from the well-fed rat, glucose (5mm) supplied 25–30% of the respiratory fuel; in the starved state, the corresponding value was 10%. These results are based on measurements of the net uptake of glucose and of the specific radioactivity of labelled carbon dioxide formed in the presence of [U-14C]-glucose. 2. Added acetoacetate (5mm) or butyrate (10mm) provided up to 80%, and added oleate (2mm) up to 50% of the fuel of respiration. The oxidation of endogenous substrates was suppressed correspondingly. 3. More [U-14C]oleate was removed by the tissue than could be oxidized by the amount of oxygen taken up; less than 25% of the oleate removed was converted into respiratory carbon dioxide and about two-thirds was incorporated into the tissue lipids. The rate of oleate incorporation into the neutral-lipid fraction was calculated to be equivalent to the rate of oxidation of endogenous fat, which provided the chief remaining fuel. 4. The contribution of endogenous substrates to the respiration (50%) in the presence of added oleate is taken to reflect either a high turnover rate of the endogenous neutral lipids (approx. half-life 2·5hr.) or a raised rate of lipolysis caused by the experimental conditions in vitro. 5. Added l-α-glycerophosphate (2·5mm) increased oleate incorporation into the neutral-lipid fraction by up to 40% (i.e. caused a net synthesis of triglyceride). 6. Lactate (2·5mm) added as sole substrate supplied 30% of the respiratory fuel, but with added oleate (2mm) lactate was converted quantitatively into glucose. Oleate stimulated the rate of gluconeogenesis from lactate by 45%. 7. The oxidation of both long-chain and short-chain even-numbered fatty acids was accompanied by ketone-body formation. Ketone-body synthesis from oleate, but not from butyrate, increased six- to seven-fold after 48hr. of starvation. The maximum rates of renal ketogenesis (80μmoles/hr./g. dry wt., with butyrate) were about 20% of the maximum rates observed in the liver (on a weight-for-weight basis) and accounted for, at most, 35% of the fatty acid removed. 8. dl-Carnitine (1·0mm) had no effect on the rates of uptake of acetate, butyrate or oleate or on the rate of radioactive carbon dioxide formation from [U-14C]oleate, but increased ketone-body formation from oleate by more than 100%. Ketone-body formation from butyrate was not increased. 9. There is evidence supporting the assumption that there are cells in which gluconeogenesis and ketogenesis occur together, characterized by equal labelling of [U-14C]oleate and the ketone bodies formed, and other cells that oxidize fat and do not form ketone bodies. 10. Inhibitory effects of unlabelled acetoacetate on the oxidation of [1-14C]butyrate and of unlabelled butyrate on [4-14C]acetoacetate oxidation show that fatty acids and ketone bodies compete as fuels on the basis of their relative concentrations. 11. The pathway of ketogenesis in renal cortex must differ from that of the liver, as β-hydroxy-β-methylglutaryl-CoA synthetase is virtually absent from the kidney. In contrast with the liver the kidney possesses 3-oxo acid CoA-transferase (EC 2.8.3.5), and the ready reversibility of this reaction and that of thiolase (EC 2.3.1.9) provide a mechanism for ketone-body formation from acetyl-CoA. This mechanism may apply to extrahepatic tissues generally, with the possible exception of the epithelium of the rumen and intestines.  相似文献   

9.
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   

10.
1. Studies on the incorporation of [3-14C]pyruvate and d-3-hydroxy[3-14C]butyrate into the brain lipid fraction by brain homogenates of the suckling (7-day-old) rat have been carried out. 2. Whereas approximately twice as much CO2 was evolved from pyruvate compared with 3-hydroxybutyrate metabolism, similar amounts of the radioactivity of these two precursors were incorporated into the lipid fraction. Furthermore, in both cases the incorporation into lipid was almost tripled when glucose (10mm) or NADPH (2.5mm) was added to the incubation media. 3. If 5mm-(—)-hydroxycitrate, an ATP–citrate lyase inhibitor, was added to the incubation the incorporation of carbon from pyruvate was inhibited to 39% of the control and from 3-hydroxybutyrate to 73% of the control, whereas CO2 production from both precursors was not affected. 4. The incorporation from pyruvate or 3-hydroxybutyrate into lipids was not affected by the presence of 10mm-glutamate in the medium (to encourage N-acetylaspartate production). However, incorporation from pyruvate was inhibited by 21% in the presence of 5mm-amino-oxyacetate (a transaminase inhibitor) and by 83% in the presence of both hydroxycitrate (5mm) and amino-oxyacetate. 5. Incorporation from 3-hydroxybutyrate into brain lipids was inhibited by 20% by amino-oxyacetate alone, but by 55% in the presence of both hydroxycitrate and amino-oxyacetate. 6. It is concluded that the mechanism of carbon transfer from pyruvate into lipids across the mitochondrial membrane in the suckling rat brain is mainly via citrate and N-acetylaspartate. 3-Hydroxybutyrate, in addition to using these routes, may also be incorporated via acetoacetate formation and transport to the cytosol.  相似文献   

11.
Biosynthesis of tropane alkaloids is thought to proceed by way of the diamine putrescine, followed by its methylation by putrescine N-methyltransferase (PMT; EC 2.1.1.53). High PMT activities were found in branch roots and/or cultured roots of several solanaceous plants. PMT was partially purified and characterized from cultured roots of Hyoscyamus albus that contain hyoscyamine as the main alkaloid. Initial velocity studies and product inhibition patterns of PMT are consistent with an ordered bi-bi mechanism, in which the Km values for putrescine and S-adenosyl-l-methionine are 277 and 203 μm, respectively, and the Ki value for S-adenosyl-l-homocysteine is 110 μm. PMT efficiently N-methylated amines that have at least two amino groups separated by three or four methylene groups. Monoamines were good competitive inhibitors of PMT, among which n-butylamine, cyclohexylamine, and exo-2-aminonorbornane were most inhibitory, with respective Ki values of 11.0, 9.1, and 10.0 μm. When n-butylamine was fed to root cultures of H. albus, the alkamine intermediates (tropinone, tropine, and pseudotropine) drastically decreased at 1 mm of the exogenous monoamine, and the hyoscyamine content decreased by 52% at 6 mm, whereas the contents of 6β-hydroxyhyoscyamine and scopolamine did not change. Free and conjugated forms of polyamines were also measured. The n-butylamine treatment caused a large increase in the putrescine content (especially in the conjugated pool), and the spermine content also increased slightly, whereas the spermidine content decreased slightly. The increase in the putrescine pool size (approximately 40 nmol/mg dry weight) was large enough to account for the decrease in the total alkaloid pool size. Similar results were also obtained in root cultures of Datura stramonium. These studies further support the role of PMT as the first committed enzyme specific to alkaloid biosynthesis.  相似文献   

12.
Fructokinase has been purified from developing potato (Solanum tuberosum L.) tubers by a combination of hydrophobic interaction, affinity chromatography, and gel filtration. The protein has a native molecular mass of approximately 70 kD but is apparently a dimer. Ion-exchange chromatography and two-dimensional western blots resolved three major fructokinases, designated FK-I, FK-II, and FK-III in order of their elution from a Mono-Q column. Fructokinase activity proved labile when proteins were purified in the absence of fructose. Kinetically, FKs I, II, and III all have broad pH optima with peaks at about pH 8.5. The enzymes have a high specificity for fructose (Km values ranging from 0.041 to 0.128 mm), and can utilize a range of nucleoside triphosphates. Unlike FKs I and II, FK-III is not inhibited by fructose concentrations in excess of 1 mm. MgADP inhibited activity of the three FKs (between 68 and 75% inhibition at 1.0 mm), whereas fructose 6-P caused inhibition at concentrations of 10 mm. There were no regulatory effects observed with a range of other metabolites. K+ (10 mm) activated FK-I by 4-fold and FKs II and III by only about 50%.  相似文献   

13.
Here we performed structural and biochemical analyses on the TK2285 gene product, an uncharacterized protein annotated as a member of the ribokinase family, from the hyperthermophilic archaeon Thermococcus kodakarensis. The three-dimensional structure of the TK2285 protein resembled those of previously characterized members of the ribokinase family including ribokinase, adenosine kinase, and phosphofructokinase. Conserved residues characteristic of this protein family were located in a cleft of the TK2285 protein as in other members whose structures have been determined. We thus examined the kinase activity of the TK2285 protein toward various sugars recognized by well characterized ribokinase family members. Although activity with sugar phosphates and nucleosides was not detected, kinase activity was observed toward d-allose, d-lyxose, d-tagatose, d-talose, d-xylose, and d-xylulose. Kinetic analyses with the six sugar substrates revealed high Km values, suggesting that they were not the true physiological substrates. By examining activity toward amino sugars, sugar alcohols, and disaccharides, we found that the TK2285 protein exhibited prominent kinase activity toward myo-inositol. Kinetic analyses with myo-inositol revealed a greater kcat and much lower Km value than those obtained with the monosaccharides, resulting in over a 2,000-fold increase in kcat/Km values. TK2285 homologs are distributed among members of Thermococcales, and in most species, the gene is positioned close to a myo-inositol monophosphate synthase gene. Our results suggest the presence of a novel subfamily of the ribokinase family whose members are present in Archaea and recognize myo-inositol as a substrate.  相似文献   

14.
1. Threonine aldolase was purified about 200-fold in 10% yield from Clostridium pasteurianum and its properties were examined. The final preparation gave three bands after ionophoresis on polyacrylamide gel. 2. The purified enzyme was shown to produce glycine and acetaldehyde in stoicheiometric amounts from threonine. The reverse reaction was demonstrated qualitatively. 3. The enzyme has a broad pH optimum at 6.5–7.0. 4. The enzyme is highly specific for l-threonine. 5. The enzyme is completely inhibited by 1mm concentrations of hydroxylamine and semicarbazide. Activity is decreased to 20% of the original by treatment with cysteine plus mercaptoethanol; most of the loss is regained on incubation with pyridoxal phosphate. It is concluded that pyridoxal phosphate is a prosthetic group. 6. The relationship between velocity and substrate concentration is atypical but indicates a Km value of 0.42mm. 7. The enzyme was demonstrated in several other strictly anaerobic bacteria.  相似文献   

15.
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   

16.
The effects of dopamine on pituitary prolactin secretion and pituitary cyclic AMP accumulation were studied by using anterior pituitary glands from adult female rats, incubated in vitro. During 2h incubations, significant inhibition of prolactin secretion was achieved at concentrations between 1 and 10nm-dopamine. However, 0.1–1μm-dopamine was required before a significant decrease in pituitary cyclic AMP content was observed. In the presence of 1μm-dopamine, pituitary cyclic AMP content decreased rapidly to reach about 75% of the control value within 20min and there was no further decrease for at least 2h. Incubation with the phosphodiesterase inhibitors theophylline (8mm) or isobutylmethylxanthine (2mm) increased pituitary cyclic AMP concentrations 3- and 6-fold respectively. Dopamine (1μm) had no effect on the cyclic AMP accumulation measured in the presence of theophylline, but inhibited the isobutylmethylxanthine-induced increase by 50%. The dopamine inhibition of prolactin secretion was not affected by either inhibitor. Two derivatives of cyclic AMP (dibutyryl cyclic AMP and 8-bromo cyclic AMP) were unable to block the dopamine (1μm) inhibition of prolactin secretion, although 8-bromo cyclic AMP (2mm) significantly stimulated prolactin secretion and both compounds increased somatotropin (growth hormone) release. Cholera toxin (3μg/ml for 4h) increased pituitary cyclic AMP concentrations 4–5-fold, but had no effect on prolactin secretion. The inhibition of prolactin secretion by dopamine was unaffected by cholera toxin, despite the fact that dopamine had no effect on the raised pituitary cyclic AMP concentration caused by this factor. Dopamine had no significant effect on either basal or stimulated somatotropin secretion under any of the conditions tested. We conclude that the inhibitory effects of dopamine on prolactin secretion are probably not mediated by lowering of cyclic AMP concentration, although modulation of the concentration of this nucleotide in some other circumstances may alter the secretion of the hormone.  相似文献   

17.
By the use of the proteolytic substrates benzoyl-dl-arginine-p-nitroanilide and benzoyl-l-arginine ethyl ester the enzyme arachain has been purified 325-fold from acetone powders of ungerminated peanuts. The pH optimum for the hydrolysis of benzoyl-dl-arginine-p-nitroanilide was 8.1 in tris buffer, and for benzoyl-l-arginine ethyl ester was 7.5 using N - 2 - hydroxyethylpiperazine - N′ - 2 - ethanesulfonic acid buffer. The purest fraction showed one main band with one to three minor bands on disc gel electrophoresis. The major protein component had an S20,w of 6.20. The energy of activation for the hydrolysis of benzoyl-dl-arginine-p-nitroanilide was calculated to be 16 kilocalories. The Michaelis constant for benzoyl-dl-arginine-p-nitroanilide was 10 micromolar and for benzoyl-l-arginine ethyl ester was 110 micromolar. The enzyme showed essentially no activity with casein, dimethyl casein, or bovine serum albumin as substrates. A large number of peptides were hydrolyzed by the enzyme, only l-leucyl-l-tyrosine being resistant of the peptides tested. The results suggest that arachain is not a “trypsin-like” protease but is a peptide hydrolase.  相似文献   

18.
C-reactive protein (CRP) is an acute phase protein of the pentraxin family that binds ligands in a Ca2+-dependent manner, and activates complement. Knowledge of its oligomeric state in solution and at surfaces is essential for functional studies. Analytical ultracentrifugation showed that CRP in 2 mm Ca2+ exhibits a rapid pentamer-decamer equilibrium. The proportion of decamer decreased with an increase in NaCl concentration. The sedimentation coefficients s20,w0 of pentameric and decameric CRP were 6.4 S and in excess of 7.6 S, respectively. In the absence of Ca2+, CRP partially dissociates into its protomers and the NaCl concentration dependence of the pentamer-decamer equilibrium is much reduced. By x-ray scattering, the radius of gyration RG values ranged from 3.7 nm for the pentamer to above 4.0 nm for the decamer. An averaged KD value of 21 μm in solution (140 mm NaCl, 2 mm Ca2+) was determined by x-ray scattering and modeling based on crystal structures for the pentamer and decamer. Surface plasmon resonance showed that CRP self-associates on a surface with immobilized CRP with a similar KD value of 23 μm (140 mm NaCl, 2 mm Ca2+), whereas CRP aggregates in low salt. It is concluded that CRP is reproducibly observed in a pentamer-decamer equilibrium in physiologically relevant concentrations both in solution and on surfaces. Both 2 mm Ca2+ and 140 mm NaCl are essential for the integrity of CRP in functional studies and understanding the role of CRP in the acute phase response.  相似文献   

19.
1. Male rats were fed for 14 days on powdered diets containing (by weight) 53% of starch, or on diets in which 20g of starch per 100g of diet was replaced by lard or corn oil. They were then fed acutely by stomach tube with a single dose of glucose, fructose or ethanol of equivalent energy contents, or with 0.15m-NaCl. The serum concentrations of corticosterone, insulin, glucose, glycerol, triacylglycerol and cholesterol were measured up to 6h after this treatment. 2. Feeding saline (0.9% NaCl) acutely to the rats maintained on the three powdered diets produced a small transient increase in circulating corticosterone that was similar to that in rats maintained on the normal 41B pelleted diet. 3. Feeding glucose acutely to the rats on the powdered diets produced peak concentrations of corticosterone that were 2–3-fold higher than those seen in rats maintained on the 41B diet. The duration of this response increased in the order starch diet<lard diet<corn-oil diet. This abnormal corticosterone response to glucose feeding appeared to be responsible for an increased activity in phosphatidate phosphohydrolase in the livers of rats fed the starch and lard diets of 2.9- and 4.9-fold respectively. The latter increase was similar to that produced by ethanol, whereas glucose did not increase the phosphohydrolase activity in the liver of rats maintained on the 41B diet. 4. Feeding fructose acutely produced even more marked increases than glucose in the concentrations of circulating corticosterone in rats given the powdered diets, but unlike glucose did not increase circulating insulin. The duration of the corticosterone response again increased in the order starch diet<lard diet<corn-oil diet. The concentrations of circulating glucose were increased by fructose feeding in rats maintained on these diets, but they were not altered in the rats maintained on the 41B pellets. A prolonged increase in serum corticosterone concentrations was also observed when fructose was fed to rats maintained on pelleted diets enriched with corn oil or beef tallow rather than with starch or sucrose. However, these effects were less marked than those seen with rats fed on the powdered diets. 5. These results are discussed in relation to the mechanism whereby high dietary fat exaggerates the effects of ethanol, fructose and sorbitol in stimulating triacylglycerol synthesis in the liver.  相似文献   

20.
1. The effect of ammonia on purine and pyrimidine nucleotide biosynthesis was studied in rat liver and brain in vitro. The incorporation of NaH14CO3 into acid-soluble uridine nucleotide (UMP) in liver homogenates and minces was increased 2.5–4-fold on incubation with 10mm-NH4Cl plus N-acetyl-l-glutamate, but not with either compound alone. 2. The incorporation of NaH14CO3 into orotic acid was increased 3–4-fold in liver homogenate with NH4Cl plus acetylglutamate. 3. The 5-phosphoribosyl 1-pyrophosphate content of liver homogenate was decreased by 50% after incubation for 10min with 10mm-NH4Cl plus acetylglutamate. 4. Concomitant with this decrease in free phosphoribosyl pyrophosphate was a 40–50% decrease in the rates of purine nucleotide synthesis, both de novo and from the preformed base. 5. Subcellular fractionation of liver indicated that the effects of NH4Cl plus acetylglutamate on pyrimidine and purine biosynthesis required a mitochondrial fraction. This effect of NH4Cl plus acetylglutamate could be duplicated in a mitochondria-free liver fraction with carbamoyl phosphate. 6. A similar series of experiments carried out with rat brain demonstrated a significant, though considerably smaller, effect on UMP synthesis de novo and purine base reutilization. 7. These data indicate that excessive amounts of ammonia may interfere with purine nucleotide biosynthesis by stimulating production of carbamoyl phosphate through the mitochondrial synthetase, with the excess carbamoyl phosphate in turn increasing pyrimidine nucleotide synthesis de novo and diminishing the phosphoribosyl pyrophosphate available for purine biosynthesis.  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