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 740 毫秒
1.
A decrease in citric acid and increases in acetic acid, acetoin and diacetyl were found in the test red wine after inoculation of intact cells of Leuconostoc mesenteroides subsp. lactosum ATCC 27307. a malo-lactic bacterium, grown on the malate plus citrate-medium. Citric acid in the buffer solution was transformed to acetic acid, acetoin and diacetyl in the pH range of 2 to 6 after inoculation with intact cells of this bacterial species. It was concluded that citric acid in wine making involving malolactic fermentation, at first, was converted by citrate lyase to acetic and oxaloacetic acids, and the latter was successively transformed by decarboxylation to pyruvic acid which was subsequently converted to acetoin, diacetyl and acetic acid.

Both the activities of citrate lyase and acetoin formation from pyruvic acid in the dialyzed cell-free extract were optimal at pH 6.0. Divalent cations such as Mn2+, Mg2+, Co2+ and Zn2+ activated the citrate lyase. The citrate lyase was completely inhibited by EDTA, Hg2+ and Ag2+ . The acetoin formation from pyruvic acid was significantly stimulated by thiamine pyrophosphate and CoCl2, and inhibited by oxaloacetic acid. Specific activities of the citrate lyase and acetoin formation were considerably variable among the six strains of malo-lactic bacteria examined. Some activities of irreversible reduction of diacetyl to acetoin were found in the cell-free extracts of four of the malo-lactic bacteria strains and the optimal pH was 6.0 for this activity of Leu. mesenteroides.  相似文献   

2.
The isolation of a soluble brain fraction which behaves as an endogenous ouabain-like substance, termed endobain E, has been described. Endobain E contains two Na+, K+-ATPase inhibitors, one of them identical to ascorbic acid. Neurotransmitter release in the presence of endobain E and ascorbic acid was studied in non-depolarizing (0 mM KCl) and depolarizing (40 mM KCl) conditions. Synaptosomes were isolated from cerebral cortex of male Wistar rats by differential centrifugation and Percoll gradient. Synaptosomes were preincubated in HEPES-saline buffer with 1 mM d-[3H]aspartate (15 min at 37°C), centrifuged, washed, incubated in the presence of additions (60 s at 37°C) and spun down; radioactivity in the supernatants was quantified. In the presence of 0.5–5.0 mM ascorbic acid, d-[3H]aspartate release was roughly 135–215% or 110–150%, with or without 40 mM KCl, respectively. The endogenous Na+, K+-ATPase inhibitor endobain E dose-dependently increased neurotransmitter release, with values even higher in the presence of KCl, reaching 11-times control values. In the absence of KCl, addition of 0.5–10.0 mM commercial ouabain enhanced roughly 100% d-[3H]aspartate release; with 40 mM KCl a trend to increase was recorded with the lowest ouabain concentrations to achieve statistically significant difference vs. KCl above 4 mM ouabain. Experiments were performed in the presence of glutamate receptor antagonists. It was observed that MPEP (selective for mGluR5 subtype), failed to decrease endobain E response but reduced 50–60% ouabain effect; LY-367385 (selective for mGluR1 subtype) and dizocilpine (for ionotropic NMDA glutamate receptor) did not reduce endobain E or ouabain effects. These findings lead to suggest that endobain E effect on release is independent of metabotropic or ionotropic glutamate receptors, whereas that of ouabain involves mGluR5 but not mGluR1 receptor subtype. Assays performed at different temperatures indicated that in endobain E effect both exocytosis and transporter reversion are involved. It is concluded that endobain E and ascorbic acid, one of its components, due to their ability to inhibit Na+, K+-ATPase, may well modulate neurotransmitter release at synapses.  相似文献   

3.
Single crystals of KCl doped with Ce3+,Tb3+ were grown using the Bridgeman–Stockbarger technique. Thermoluminescence (TL), optical absorption, photoluminescence (PL), photo‐stimulated luminescence (PSL), and thermal‐stimulated luminescence (TSL) properties were studied after γ‐ray irradiation at room temperature. The glow curve of the γ‐ray‐irradiated crystal exhibits three peaks at 420, 470 and 525 K. F‐Light bleaching (560 nm) leads to a drastic change in the TL glow curve. The optical absorption measurements indicate that F‐ and V‐centres are formed in the crystal during γ‐ray irradiation. It was attempted to incorporate a broad band of cerium activator into the narrow band of terbium in the KCl host without a reduction in the emission intensity. Cerium co‐doped KCl:Tb crystals showed broad band emission due to the d–f transition of cerium and a reduction in the intensity of the emission peak due to 5D37Fj (j = 3, 4) transition of terbium, when excited at 330 nm. These results support that energy transfer occurs from cerium to terbium in the KCl host. Co‐doping Ce3+ ions greatly intensified the excitation peak at 339 nm for the emission at 400 nm of Tb3+. The emission due to Tb3+ ions was confirmed by PSL and TSL spectra. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

4.
A procedure for preparing highly purified brush border membranes from rabbit kidney cortex using differential and density gradient centrifugation is described. Brush border membranes prepared by this procedure were substantially free of basal-lateral membranes, mitochondria, endoplasmic reticulum and nuclear material as evidenced by an enrichment factor of less than 0.3 for (Na+ + K+)-ATPase, succinate dehydrogenase, NADPH-cytochrome c reductase and DNA. Alkaline phosphatase was enriched ten fold indicating that the membranes were enriched at least 30 fold with respect to other cellular organelles. The yield of brush border membranes was 20%.Transport of d-glucose by the membranes was identical to that previously reported except that the Arrhenius plot for temperature dependence of transport was curvilinear (EA = 11.3–37.6 kcal/mol) rather than biphasic. Transport of p-aminohippuric acid and uric acid were increased by the presence of NaCl, either gradient or preequilibrated. However, no overshoot was obtained in the presence of a NaCl gradient, and KCl and LiCl also produced equivalent stimulation of transport suggesting a nonspecific ionic strength effect. Uptakes of p-aminohippuric acid and uric acid were not saturable, and were increased markedly by reducing the pH from 7.5 to 5.6. Probenecid (1 mM) reduced p-aminohippuric acid and uric acid (50 μM) uptake by 49% and 21%, respectively. We conclude that the uptake of uric acid and p-aminohippuric acid by renal brush border membranes of the rabbit occurs primarily by a simple solubility-diffusion mechanism.  相似文献   

5.
Membranes of corn (Zea mays, cv Trojan 929) coleoptiles were fractionated by sucrose density gradient centrifugation and the locations of organelles were determined using marker enzymes and electron microscopy. Latent IDPase (or UDPase) was selected as the Golgi marker and UDPG-sterol glucosyl transferase was selected as the plasma membrane (PM) marker, because they were clearly separable from markers for the other organelles. Golgi-rich and PM-rich fractions were studied in relation to their ATPase activities. The pH optimum of the KCl, Mg2+-ATPase of the PM-rich fraction from a step gradient was 6.0 to 6.5, while the Golgi-rich fraction had peaks at pH 6.0 to 6.5 and pH 7.5. It is hypothesized that the peak at pH 6.0 to 6.5 for the Golgi-rich fraction is due to PM-contamination, while the peak at pH 7.5 represents the activity of a Golgi ATPase. To reduce PM contamination, Golgi-rich fractions obtained from step or rate-zonal gradients were recentrifuged isopycnically on linear sucrose gradients. The distribution of KCl, Mg2+-ATPase activity was measured at pH 6.5 and 7.5. The pH 6.5 ATPase was coincident with UDPG-sterol glucosyl transferase, a PM marker, while the pH 7.5 ATPase overlapped with latent UDPase, a Golgi marker. These results provide strong evidence for a KCl, Mg2+-ATPase, active at pH 7.5, associated with the Golgi membranes of corn coleoptiles.  相似文献   

6.
Barnea caridida oocytes release acid (1.35 pmole H+/oocyte) upon fertilization. After artificial activation by an excess of KCl, germinal vesicle breakdown (GVBD) occurs normally and a quite similar, but not identical, acid release is recorded (1.10 pmole H+/oocyte). KCl activation of Barnea oocytes is completely inhibited in 100 mM sodium-acetate sea water at pH 6.5 and fertilization does not result in activation when the oocytes are transferred after one minute into 100 mM sodium-acetate sea water at pH 6.3. When D–600, a calcium transmembrane fluxes inhibitor, is added 20 seconds after fertilization, GVBD is inhibited but a normal acid release is recorded. The presence of at least 10 mM sodium ions in the external medium is required for 100% activation of these oocytes by an excess of KCl. These results suggest that while an intracellular pH increase may be a requisite for GVBD, this can not be a sufficient condition to trigger it unless a calcium influx is allowed to occur. Moreover, the acid release does not result from a Ca++-H+ exchange transport but appears more likely to be due to a Na+-H* exchange as it has been demontrated in sea urchin eggs.  相似文献   

7.
The influence of NO3 uptake and reduction on ionic balance in barley seedlings (Hordeum vulgare, cv. Compana) was studied. KNO3 and KCl treatment solutions were used for comparison of cation and anion uptake. The rate of Cl uptake was more rapid than the rate of NO3 uptake during the first 2 to 4 hours of treatment. There was an acceleration in rate of NO3 uptake after 4 hours resulting in a sustained rate of NO3 uptake which exceeded the rate of Cl uptake. The initial (2 to 4 hours) rate of K+ uptake appeared to be independent of the rate of anion uptake. After 4 hours the rate of K+ uptake was greater with the KNO3 treatment than with the KCl treatment, and the solution pH, cell sap pH, and organic acid levels with KNO3 increased, relative to those with the KCl treatment. When absorption experiments were conducted in darkness, K+ uptake from KNO3 did not exceed K+ uptake from KCl. We suggest that the greater uptake and accumulation of K+ in NO3-treated plants resulted from (a) a more rapid, sustained uptake and transport of NO3 providing a mobile counteranion for K+ transport, and (b) the synthesis of organic acids in response to NO3 reduction increasing the capacity for K+ accumulation by providing a source of nondiffusible organic anions.  相似文献   

8.
Total phospholipids were extracted from cells of temperature sensitive unsaturated fatty acid auxotrophs of Escherichia coli (K-12 UFAts) grown at 28°C (PL28), and at 42°C in the presence of 2% KCl as an osmotic stabilizer (PL42 (KCl)). From the analysis of fatty acids, it was shown that the content of unsaturated fatty acids of PL42 (KCl) is only 9% of the total fatty acids, while that of PL28 is 54%. The thermal phase transitions of the bilayers prepared from the phospholipid fractions were studied by proton magnetic resonance. The line widths of the methylene signals and the sums of the methylene and methyl signal intensities were plotted against reciprocal values of absolute temperature 1/T or temperature itself. From the plots phase transitions were detected at about 19°C for PL28 and at 43°C for PL42 (KCl). In spite of its complex composition of fatty acids a highly cooperative transition was observed in the case of PL42 (KCl). It was also suggested that the phospholipids bilayers in the biomembranes of this strain at the growth temperature (42°C) are in the state where the gel and liquid crystalline phases coexist.  相似文献   

9.
Factors behind the small-scale variaton of pH were examined in O horizons (humus layers) developed under two stands of Picea abies (L.) Karst. (Site F and K) by combining data on the composition of the cation exchange complex with data from titrations of corresponding H+-saturated samples. Cations extractable in 0.5 M CuCl2 (S=cmolc kg–1 [2Ca+2Mg+2Mn+K+Na]), aluminium extractable in 1.0 M KCl (Ale=cmolc kg–1 [3Al]) and in 0.5 M CuCl2 (Alorg=cmolc kg–1 [3Al]), as well as pH measured in 0.01 M CaCl2 (pHCa) were analysed in one-cm-layers of 13 O horizon cores at each site. Composite samples representing each of the one-cm-layers at each site, as well as samples with two different levels of Al saturation at Site K, were H+ saturated and titrated with NaOH to chosen end points of pHCa=4.0 and 5.5 in a 0.01 M CaCl2 ionic medium. The Acid Neutralisation Capacity (ANC) was estimated as the amount of base needed to increase pHCa of the composite H+-saturated samples to the mean pHCa of the corresponding natural samples. The ANC was found to be similar in magnitude or higher than the amount of sites binding S+Ale, which suggests that 1.0 M KCl exchangeable Al ions are nonacidic in acid O horizons. The relative contribution from i) the capacity of acidic functional groups, ii) their acid strength and iii) their degree of neutralisation to differences in pHCa between sites, among cm-layers and between samples with different levels of Al saturation were estimated from titration curves adjusted to hold two out of three factors (i, ii and iii) to be constant. The degree of neutralisation explained most of the differences in pHCa between the two sites, as well as between samples with different levels of Al saturation at Site K. The pHCa decrease by depth at site F was, however, partly explained by an increasing acid strength. The study emphasizes the importance of examining not only changes in the degree of neutralisation, but also changes in the acid strength and the capacity of buffering functional groups before conclusions about causes behind acidification processes can be made. Difficulties of accurately estimating the degree of neutralisation (base saturation) of acidic functional groups from the composition of adsorbed cations, owing to the unknown acidity of adsorbed Al, was also demonstrated.  相似文献   

10.
Abstract: To see the effect of a γ-aminobutyric acid GABA uptake inhibitor on the efflux and content of endogenous and labeled GABA, rat cortical slices were first labeled with [3H]GABA and then superfused in the absence or presence of 1 mM nipecotic acid. Endogenous GABA released or remaining in the slices was measured with high performance liquid chromatography, which was also used to separate [3H]GABA from its metabolites. In the presence of 3 mM K+, nipecotic acid released both endogenous and [3H]GABA, with a specific activity four to five times as high as that present in the slices. The release of labeled metabolite(s) of [3H]GABA was also increased by nipecotic acid. The release of endogenous GABA evoked by 50 mM K+ was enhanced fourfold by nipecotic acid but that of [3H]GABA was only doubled when expressed as fractional release. In a medium containing no Ca2+ and 10 mM Mg2+, the release evoked by 50 mMK+ was nearly suppressed in either the absence or the presence of nipecotic acid. In the absence of nipecotic acid electrical stimulation (bursts of 64 Hz) was ineffective in evoking release of either endogenous or [3H]GABA, but in the presence of nipecotic acid it increased the efflux of endogenous GABA threefold, while having much less effect on that of [3H]GABA. Tetrodotoxin (TTX) abolished the effect of electrical stimulation. Both high K+ and electrical stimulation increased the amount of endogenous GABA remaining in the slices, and this increase was reduced by omission of Ca2+ or by TTX. The results suggest that uptake of GABA released through depolarization is of major importance in removing GABA from extracellular spaces, but the enhancement of spontaneous release by nipecotic acid may involve intracellular heteroexchange. Depolarization in the presence of Ca2+ leads to an increased synthesis of GABA, in excess of its release, but the role of this excess GABA remains to be established.  相似文献   

11.
Plasma membranes were prepared from guinea pig ventricle by a procedure which involved differential centrifugation at low gravitational forces, extraction with KCl, and centrifugation in a discontinuous sucrose gradient. Adenylate cyclase was purified 10–15-fold over the starting homogenate with a yield of 75%. The membranes contained an active Ca2+ binding and uptake system as well as Ca2+-activated adenosine triphosphatase; protein kinase and phosphoprotein phosphatase activities were also present. The membranes could be phosphorylated by either intrinsic or exogenous protein kinase, and phosphorylation was stimulated by cyclic AMP and was reversible. Phosphorylated membranes accumulated twice as much Ca2+ as control preparations.  相似文献   

12.
Depolymerization of hyaluronic acid obtained from Streptococcus zooepidemicus by D-fructose 6-phosphate was investigated for characterization of reducing sugar-mediated degradation of biopolymers under physiological conditions. The extent of depolymerization was monitored by the decrease of viscosity of a reaction mixture containing 1.0% hyaluronic acid, D-fructose 6-phosphate, and 1.0 × 10?2 mM of Cu2+ in phosphate buffer, pH 7.4. It was found that the depolymerization of hyaluronic acid was dependent on the concentration of the reducing sugar and was specifically accelerated by the presence of Cu2+. The reaction was found to be significantly inhibited by catalase, superoxide dismutase (SOD), 1,2-dihy­ droxybenzene 3,5-disulfonic acid (Tiron), and chelating agents such as EDTA and diethylene triamine penta­ acetic acid (DETAPAC), although the inhibition by SOD was low. Almost the same depolymerization rates were observed in hyaluronic acid preparations of different molecular weight (1.1 × 106, 8.8 × 105, and 6.8 × 105). The rates, however, were different for hyaluronic acids obtained from S. zooepidemicus, rooster comb, and umbilical cord. It was concluded that depolymerization of the polysaccharide was caused by active oxygen species generated by the autoxidation of D-fructose 6-phosphate in the presence of Cu2+, in a mechanism similar to that previously reported for the degradation of DNA and inactivation of virus in vitro.  相似文献   

13.
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   

14.
The effects of abscisic acid, salicylic acid and trans-cinnamic acid were tested on the light-induced phosphorylating reactions and oxygen evolution of the unicellular green alga Scenedesmus. It was found that abscisic acid and cinnamic acid had practically no influence on the total inorganic phosphate uptake, while salicylic acid in the concentration range of 10-6 to 10-3M gave a small decrease in the total inorganic phosphate uptake. The ATP level in the cells is in most cases increased when these three acids are given to the algae. The oxygen output is not significantly changed by abscisic acid or salicylic acid. Trans- cinnamic acid inhibits the oxygen evolution at concentrations of 10-4–10-3M None of the substances investigated caused such effects on photophosphorylation and oxygen evolution in Scenedesmus as those caused by the inhibitor β-complex from potatoes according to earlier reports. It is suggested that these effects are due to other components in the inhibitor β-complex.  相似文献   

15.
  • 1.1. In the contents of the oesophagus and stomach, one form of acid phosphatase is found. Its electrophoretic mobility is identical to that of the multiple form 3 of acid phosphatase from the hepatopancreas.
  • 2.2. The enzyme is not stimulated by divalent cations. It is inhibited by molybdate, Cu2+, Hg2+. F and tartrate L+.
  • 3.3. The optimum pH of the enzyme is 4.5. The Km for paranitrophenylphosphate as substrate amounts to 0.25 mM. The enzyme is stable at a temperature of up to 55°C.
  相似文献   

16.
A preliminary analysis of Fatty Acid synthesis in pea roots   总被引:3,自引:3,他引:0       下载免费PDF全文
Subcellular fractions from pea (Pisum sativum L.) roots have been prepared by differential centrifugation techniques. Greater than 50% of the recovered plastids can be isolated by centrifugation at 500g for 5 minutes. Plastids of this fraction are largely free from mitochondrial and microsomal contamination as judged by marker enzyme analysis. De novo fatty acid biosynthesis in pea roots occurs in the plastids. Isolated pea root plastids are capable of fatty acid synthesis from acetate at rates up to 4.3 nanomoles per hour per milligram protein. ATP, bicarbonate, and either Mg2+ or Mn2+ are all absolutely required for activity. Coenzyme A at 0.5 millimolar improved activity by 60%. Reduced nucleotides were not essential but activity was greatest in the presence of 0.5 millimolar of both NADH and NADPH. The addition of 0.5 millimolar glycerol-3-phosphate increased activity by 25%. The in vitro and in vivo products of fatty acid synthesis from acetate were primarily palmitate, stearate, and oleate, the proportions of which were dependent on experimental treatments. Fatty acids synthesized by pea root plastids were recovered in primarily phosphatidic acid and diacylglycerol or as water soluble derivatives and the free acids. Lesser amounts were found in phosphatidylcholine, phosphatidylethanolamine, phosphatidylglycerol, and monogalactosyldiacylglycerol.  相似文献   

17.
Schwartz A 《Plant physiology》1985,79(4):1003-1005
Ca2+ (0.1-1.0 millimolar) accelerated dark-induced stomatal closure and reduced stomatal apertures in the light in epidermal peels of Commelina communis L. In contrast, ethyleneglycol-bis-(β-aminoethyl ether) N,N′tetraacetic acid (EGTA) (2 millimolar), a Ca2+ chelator, prevented closure in the dark and accelerated opening in the light. EGTA did not promote significant opening in the dark. It is therefore concluded that EGTA does not increase ion uptake into guard cells, but rather prevents ion efflux. Addition of EGTA to incubating solutions with 10 millimolar KCl resulted in steady state apertures of 15.6 micrometers, whereas in the absence of EGTA similar apertures required 55 millimolar KCl and 150 millimolar KCl was needed in the presence of 1 millimolar CaCl2. The results demonstrate the importance of Ca2+ in the regulation of stomatal closure and point to a role of Ca2+ in the regulation of K+ efflux from stomatal guard cells.  相似文献   

18.
Primary cell walls, free from cytoplasmic contamination were prepared from corn (Zea mays L.) roots and potato (Solanum tuberosum) tubers. After EDTA treatment, the bound acid phosphatase activities were measured in the presence of various multivalent cations. Under the conditions of minimized Donnan effect and at pH 4.2, the bound enzyme activity of potato tuber cell walls (PCW) was stimulated by Cu2+, Mg2+, Zn2+, and Mn2+; unaffected by Ba2+, Cd2+, and Pb2+; and inhibited by Al3+. The bound acid phosphatase of PCW was stimulated by a low concentration but inhibited by a higher concentration of Hg2+. On the other hand, in the case of corn root cell walls (CCW), only inhibition of the bound acid phosphatase by Al3+ and Hg2+ was observed. Kinetic analyses revealed that PCW acid phosphatase exhibited a negative cooperativity under all employed experimental conditions except in the presence of Mg2+. In contrast, CCW acid phosphatase showed no cooperative behavior. The presence of Ca2+ significantly reduced the effects of Hg2+ or Al3+, but not Mg2+, to the bound cell wall acid phosphatases. The salt solubilized (free) acid phosphatases from both PCW and CCW were not affected by the presence of tested cations except for Hg2+ or Al3+ which caused a Ca2+-insensitive inhibition of the enzymes. The induced stimulation or inhibition of bound acid phosphatases was quantitatively related to cation binding in the cell wall structure.  相似文献   

19.
We have studied the relationship between acid release, cytoplasmic alkalinization, and the extent of chromosome condensation during parthenogenetic activation of sea urchin eggs. The relative rate of acid release in Strongylocentrotus purpuratus eggs was determined from pH measurements of egg suspensions. Acid release in inseminated eggs began after a lag of 0.4 min and the relative rate increased 108-fold, declined, and release was essentially complete by 8-min postinsemination. An average of 3.8 ± 0.23 × 10?12moles H+ cell? was released as determined by backtitration with NaOH. Acid release characteristics of eggs parthenogenetically activated with either NH4C1, methylamine ethylamine, n-propylamine, n-butylamine, or benzylamine were qualitatively similar. There was no detectable lag peroid and the increase in relative rate of acid release was directly proportional to the carbon number of the amine used, eg, from 8.3-fold methylamine to 470-fold with benzylamine. The total equivalents of acid released ranged from 0.50–8.2 × 10?12 moles H+·cell? in direct proportion to the concentration of amine used. The degree fo cytoplasmic alkalinization induced as a function of methylamine and benzylamine concentration was determined by pH measurements fo egg homogenates; egg cultures were also prepared for microscopic examination of chromosome condensation. None of the eggs had condensed chromosomes at 0.5-mM methylamine whereas a cytoplasmic alkalinization of 0.6 pH units was observed. Increased methylamine levels up to 10mM resulted in chromiosome condensation in only 20% of the eggs. A similar result was found with benzylamine. We conclude that acid release and cytoplasmic alkalinization during chemical parthenogenesis are insufficient to mimic sperm induction of chromiosome condensation and suggest that an additional factor(s) is required for chromosome condensation by low concentration of amines.  相似文献   

20.
An isozyme of acid phosphatase-1, acid phosphatase-11, was purified from the leaves of tomato (Lycopersicon esculentum) to homogeneity and characterized. The purified enzyme was homogeneous on polyacrylamide gel electrophoresis with or without sodium dodecyl sulfate. The gel filtration analysis showed that the native molecule had a relative molecular mass of about 61 kilodaltons (kDa). The relative molecular mass of the subunit on gel electrophoresis with sodium dodecyl sulfate was about 32 kDa, indicating that the native form of the enzyme was a homodimer. It was suggested by periodic acid-Schiff staining on the gel that the enzyme was a glycoprotein. The Km for p-nitrophenylphosphate was 2.9 × 10?3 m. The enzyme had a pH optimum of 4.5 in 0.15 m potassium acetate buffer with p-nitrophenylphosphate as a substrate. This enzyme was activated by divalent metal ions, such as Zn2+, Mg2+, and Mn2+. The N-terminal amino acids were sequenced after the purified enzyme was treated with pyroglutamylpeptidase. It was suggested that the N-terminal amino acid was pyroglutamate.  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