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.
An isotope dilution procedure for oxalate based upon [1,2-13C2]oxalic acid is described. For routine determinations of urinary concentration, a known quantity of sodium [1,2-13C]oxalate is admixed with the sample, total oxalate precipitated as the calcium salt, and converted by BF3 catalysis to di-n-propyl esters for mass-spectrometric analysis. Selective ion monitoring provides 12C:13C ratios directly, thus precluding the necessity for quantitative recovery at any step of the rapid, single-tube assay. Following a bolus injection of sodium [1,2-13C]oxalate, whole body oxalate pools and their turnover rates can be determined by sequential sampling of urine. Biosynthetic rates calculated from the product of pool size and turnover are in excellent agreement with urinary excretion rates, confirming directly that urinary oxalate is a quantitative index of biosynthesis.  相似文献   

2.
Metabolic control associated with diauxic growth of Pseudomonas oxalaticus in batch cultures on mixtures of formate and oxalate was investigated by measuring intracellular enzyme and coenzyme concentrations and Q O 2values during transition experiments from oxalate to formate and vice versa. In transition from oxalate to formate oxalyl-CoA reductase concentration declined after the exhaustion of oxalate and ribulose-1,5-diphosphate carboxylase and 14CO2 fixation appeared upon addition of formate. In the reciprocal transition, ribulose-1,5-diphosphate carboxylase and 14CO2 fixation rate declined sharply after formate exhaustion, and oxalyl-CoA reductase appeared only after addition of oxalate. The intracellular NAD and NADP concentrations measured in the same experiments are reported. At substrate exhaustion the proportion of NAD in the reduced form fell from 15–20% to 2%. On addition of formate to an oxalate-starved culture there was an immediate increase in the proportion of NADH to 50%; such an increase was not observed in the reverse experiment.Abbreviations RuDP ribulose-1,5-diphosphate - HEPES 2-(N-2 hydroxyethylpiperazin-N-yl) ethane sulphonic acid  相似文献   

3.
Oxalate accumulates in leaves of certain plants such as Rumex species (Polygonaceae). Oxalate plays important roles in defense to predator, detoxification of metallic ions, and in hydrogen peroxide formation upon wounding/senescence. However, biosynthetic pathways of soluble oxalate are largely unknown. In the present study we analysed Rumex obtusifolius L. treated with itaconate (an antagonist to isocitrate). Comprehensive metabolome analysis using capillary electrophoresis-mass spectrometry showed that oxalate content of “new leaves” was notably down-regulated by itaconate, as opposed to the accumulation of citrate. The 13CO2 feeding experiment revealed that oxalate accumulation in new leaves was affected by citrate translocation from stems. The results suggested that excess oxalate in new leaves of R. obtusifolius was synthesized primarily via the isocitrate pathway utilizing citrate delivered from stems.  相似文献   

4.
Oxalate was shown to enter isolated rat hepatocytes and to inhibit gluconeogenesis from lactate, pyruvate, and alanine, but not from glutamine, proline, propionate or dihydroxyacetone. Oxalate apparently acts by inhibiting pyruvate carboxylase (EC 6.4.1.1). It is known to inhibit the isolated enzyme, and inhibition of gluconeogenesis was much greater in a bicarbonate-deficient medium where pyruvate carboxylase activity limits the overall rate of the pathway. A slight inhibition of gluconeogenesis from asparagine was observed, suggesting that oxalate may also inhibit gluconeogenesis at another site. Chelation of extracellular Ca2+ does not contribute to the inhibition of gluconeogenesis. Compared to oxalate, other Ca2+ chelators have little effect upon gluconeogenesis. Also, oxalate inhibits gluconeogenesis effectively both in low Ca2+ medium and in medium containing 2.6 mM Ca2+. Chelation of intracellular Ca2+ also appears to be of little importance, since oxalate does not block the glycogenolytic effects of epinephrine, vasopressin, and angiotensin which are thought to act via Ca2+ as the second messenger. The inhibition of gluconeogenesis could conceivably contribute to the toxic actions of oxalate and to the hypoglycemic action of dichloroacetate, a compound that is metabolized to oxalate. However, oxalate did not cause hypoglycemia in the suckling rat, a model in vivo system very dependent upon gluconeogenesis for maintenance of normal blood glucose levels. Thus, inhibition of gluconeogenesis is probably of little importance in oxalate toxicity and the hypoglycemic effects of dichloroacetate.  相似文献   

5.
Hydroxypyruvate and glycolate inhibited the oxidation of [U-14C]glyoxylate to [14C]oxalate in isolated perfused rat liver, but stimulated total oxalate and glycolate synthesis. [14C]Oxalate synthesis from [14C]glycine similarly inhibited by hydroxypyruvate, but conversion of [14C1]glycolate to [4C]oxalate was increased three-fold. Pyruvate had no effect on the synthesis of [14C]oxalate or total oxalate. The inhibition studies suggest that hydroxypyruvate is a precursor of glycolate and oxalate and that the conversion of glycolate to oxalate does not involve free glyoxylate as an intermediate. [14C3]Hydroxypyruvate, but not [14C1]hydroxypyruvate, was oxidized to [14C]oxalate in isolated perfused rat liver. Isotope dilution studies indicate the major pathway involves the decarboxylation of hydroxypyruvate forming glycolaldehyde which is subsequently oxidized to oxalate via glycolate. The oxidation of serine to oxalate appears to proceed predominantly via hydroxypyruvate rather than glycine or ethanolamine. The hyperoxaluria of L-glyceric aciduria, primary hyperoxaluria type II, is induced by the oxidation of the hydroxypyruvate, which accumulates because of the deficiency of D-glyceric dehydrogenase, to oxalate.  相似文献   

6.
The effect of NaCl stress on molecular and biochemical properties of oxalate oxidase (OXO) was studied in leaves of grain sorghum hybrid (var CSH-14) seedlings. There was no effect on molecular weight and number of subunits of the enzyme but it showed some important changes in its kinetic parameters such as Km for oxalate and Vmax. Optimum pH (5.8), activation energy (5.084 kcal mole?1), time of incubation (6 min) and Km for oxalate (1.21×10t-4M) were increased, while Vmax (0.182 mmole min?1) decreased and no change in optimum temperature was observed. This showed that substrate affinity and maximum activity of the enzyme was adversely affected. The specific activity of oxalate oxidase was increased in seedlings grown in a NaCl containing medium compared to normal, which reveals the increased de novo synthesis of the enzyme to sustain oxalate degradation.  相似文献   

7.
Oxalate metabolism by tobacco leaf discs   总被引:1,自引:0,他引:1       下载免费PDF全文
Havir EA 《Plant physiology》1984,75(2):505-507
The turnover rate of oxalate in leaf discs of Nicotiana tabacum, var Havana Seed, during photosynthesis was estimated to be 1 to 2 micromoles per gram fresh weight per hour. Radioactivity from the enzymic oxidation of [14C]oxalate rapidly appeared in neutral sugars (mainly sucrose), organic acids (mainly malate), and amino acids. Only 5% of the radioactivity was released to the atmosphere as 14CO2, and no formate or formaldehyde could be detected. The metabolism of oxalate was not increased by raising the O2 concentration from 1% to 21% to 60%, nor was the formation of [14C]oxalate from [2-14C]glyoxylate changed under the same conditions as was previously observed in vitro (Havir 1983 Plant Physiol 71: 874-878). While oxalate is not an inert end product of the glycolate pathway, it contributes little to the formation of photorespiratory CO2.  相似文献   

8.
The effect of Ca2+ on morphophysiological parameters of wheat calli (Triticum aestivum L.) infected by the bunt pathogen Tilletia caries, in particular on the level of active oxygen species, activity of oxalate oxidase, peroxidase, and catalase is investigated. The concentration of O2−, H2O2, and activity of oxidoreductases (oxalate oxidase, peroxidase, and catalase) depended on the content of Ca2+ in the culture medium of calli. The increase of the concentration of Ca2+ ions in the culture medium led to forming of calli with high structure, induction of activity of oxalate oxidase and of some isoperoxidase, and to accumulation of active oxygen species. These changes contributed to inhibition of development of the fungus. So this dependence confirm the role of calcium as the intermediant in biochemical reactions related to the formation of the protective response of plant cells to biotic stress.  相似文献   

9.
通过草酸及其与不同抑制剂亚甲基蓝、EGTA、氯丙嗪和Li+组合处理黄瓜叶片,研究了草酸与抑制剂不同处理组合方式对黄瓜叶片POD活性和叶片病情指数的影响,探讨NO、钙信使系统在草酸诱导叶片抗霜霉病中的作用.结果显示,10~70mmol/L草酸均能不同程度诱导黄瓜叶片POD活性的升高,提高叶片对黄瓜霜霉病的抗病性,降低叶片病情指数,并以30mmol/L效果最好.4种抑制剂分别与30mmol/L草酸同时或先于草酸处理,或草酸处理后一定时间再用抑制剂处理,均明显抑制黄瓜叶片POD活性的升高及病情指数的降低.研究表明,NO、Ca2+、钙调素(CaM)和磷酸肌醇均可能参与了草酸诱导黄瓜霜霉病抗性的信号转导过程.  相似文献   

10.
A soluble oxalate oxidase activity has been detected in homogenate of ripened fruits of strawberry (Fragaria ananassa), as confirmed by the stoichiometric relationship between the disappearance of oxalate and utilization of dissolved O2, and generation of H2O2. The enzyme was purified up to apparent homogeneity and had a Mr of 119 kDa with two identical subunits. Km for oxalate was found to be 1.67×10?3 M, and Vmax of 0.741 mmoles ml?1min?1. It retained 76% of its initial activity, when heated at 60°C for 30 min. The enzyme was found to be glycoprotein in nature. The significant increase in the enzyme activity of ripened fruits compared to that in pre-ripened fruit, and decrease in oxalate level (?0.927 correlation with oxalate oxidase) with advancement of ripening indicated the physiological role of enzyme in fruit ripening.  相似文献   

11.
We studied hydrated calcium oxalate and its ions at the restricted Hartree–Fock RHF/6-31G* level of theory. Performing a configurational search seems to improve the fit of the HF/6-31G* level to experimental data. The first solvation shell of calcium oxalate contains 13 water molecules, while the first solvation shell of oxalate ion is formed by 14 water molecules. The first solvation shell of Ca(II) is formed by six water molecules, while the second shell contains five. At 298.15 K, we estimate the asymptotic limits (infinite dilution) of the total standard enthalpies of hydration for Ca(II), oxalate ion and calcium oxalate as ?480.78, –302.78 and –312.73 kcal mol?1, resp. The dissociation of hydrated calcium oxalate is an endothermic process with an asymptotic limit of +470.84 kcal mol?1.
Figure
CaC2O4(H2O)16 and C2O4 2-(H2O)14  相似文献   

12.
1. The labelling patterns of phosphoglycerate obtained from formate-grown or oxalate-grown Pseudomonas oxalaticus after exposure for 15sec. to [14C]formate or [14C]oxalate respectively were determined. 2. The phosphoglycerate obtained from the formate-grown cells contained 78% of the radioactivity in the carboxyl group. This is in accord with that predicted for operation of the ribulose diphosphate cycle of carbon dioxide fixation. 3. The labelling pattern of the phosphoglycerate obtained from the oxalate-grown cells approached uniformity, as predicted for the heterotrophic pathway of oxalate assimilation. The departure from complete uniformity may have been due to concurrent 14CO2 fixation into C4 dicarboxylic acids. 4. The labelling pattern of phosphoglycerate obtained from cells that had just started to grow on oxalate after adaptation from formate was determined after incubation of the cells for 15sec. with [14C]oxalate. This pattern approached uniformity. 5. The pathway of incorporation of 14CO2 into cells that had just started to grow on oxalate after adaptation from formate, in the presence of either formate or oxalate as energy source, was studied by chromatographic and radio-autographic analysis. 6. It is concluded from the isotopic data that a mixed heterotrophic–autotrophic metabolism, with the former mode predominating, operates in the initial stages of growth on oxalate after adaptation from growth on formate.  相似文献   

13.
Serum transferrin (sTf) is a bi-lobal protein. Each lobe of sTf binds one Fe3+ ion in the presence of a synergistic anion. Physiologically, carbonate is the main synergistic anion but other anions such as oxalate, malonate, glycolate, maleate, glycine, etc. can substitute for carbonate in vitro. The present work provides the possible pathways by which the substitution of carbonate with oxalate affects the structural, kinetic, thermodynamic, and functional properties of blood plasma sTf. Analysis of equilibrium experiments measuring iron release and structural unfolding of carbonate and oxalate bound diferric-sTf (Fe2sTf) as a function of pH, urea concentration, and temperature reveal that the structural and iron-centers stability of Fe2sTf increase by substitution of carbonate with oxalate. Analysis of isothermal titration calorimetry (ITC) scans showed that the affinity of Fe3+ with apo-sTf is enhanced by substituting carbonate with oxalate. Analysis of kinetic and thermodynamic parameters measured for the iron release from the carbonate and oxalate bound monoferric-N-lobe of sTf (FeNsTf) and Fe2sTf at pH 7.4 and pH 5.6 reveals that the substitution of carbonate with oxalate inhibits/retards the iron release via increasing the enthalpic barriers.  相似文献   

14.
Pure cadmium oxalate trihydrate (COT) and barium added cadmium oxalate (BCO) single crystals were grown by controlled diffusion of Cd2+ and Ba2+ ions in silica gel at ambient temperature. A single test tube technique coupled with gel aging conferred maximum size crystals by controlling the nucleation rate. It was found that the pH and age of the gel greatly influenced the crystal quality, their size and transparency. Grown crystals CdC2O4 · 3H2O and Ba0.5Cd0.5(C2O4)2 · 5H2O were characterized by X-ray diffraction, Fourier transform infrared spectroscopy and thermal analysis. Effect of barium dopant on the growth and morphology of cadmium oxalate was studied. Pure cadmium oxalate crystallized in triclinic system and the barium-doped cadmium oxalate crystallized in hexagonal system with massive changes in their unit cell parameters. The infrared spectrum revealed the presence of oxalate ligands and water of hydration in both the pure and barium-doped crystals. Thermal analysis showed that the grown crystals were dehydrated thermally even from lower temperatures and the doped crystals were found more stable.  相似文献   

15.
The initial rate of oxalate-facilitated Ca2+ uptake by skeletal microsomes depends on both Ca2+ and oxalate concentrations in the medium. The apparent Km for Ca2+ increases with increasing oxalate concentration, indicating that Ca2+ uptake can involve a carrier-mediated transport system.  相似文献   

16.
The redox cycle of 2,5-dimethoxybenzoquinone (2,5-DMBQ) is proposed as a source of reducing equivalent for the regeneration of Fe2+ and H2O2 in brown rot fungal decay of wood. Oxalate has also been proposed to be the physiological iron reductant. We characterized the effect of pH and oxalate on the 2,5-DMBQ-driven Fenton chemistry and on Fe3+ reduction and oxidation. Hydroxyl radical formation was assessed by lipid peroxidation. We found that hydroquinone (2,5-DMHQ) is very stable in the absence of iron at pH 2 to 4, the pH of degraded wood. 2,5-DMHQ readily reduces Fe3+ at a rate constant of 4.5 × 103 M−1s−1 at pH 4.0. Fe2+ is also very stable at a low pH. H2O2 generation results from the autoxidation of the semiquinone radical and was observed only when 2,5-DMHQ was incubated with Fe3+. Consistent with this conclusion, lipid peroxidation occurred only in incubation mixtures containing both 2,5-DMHQ and Fe3+. Catalase and hydroxyl radical scavengers were effective inhibitors of lipid peroxidation, whereas superoxide dismutase caused no inhibition. At a low concentration of oxalate (50 μM), ferric ion reduction and lipid peroxidation are enhanced. Thus, the enhancement of both ferric ion reduction and lipid peroxidation may be due to oxalate increasing the solubility of the ferric ion. Increasing the oxalate concentration such that the oxalate/ferric ion ratio favored formation of the 2:1 and 3:1 complexes resulted in inhibition of iron reduction and lipid peroxidation. Our results confirm that hydroxyl radical formation occurs via the 2,5-DMBQ redox cycle.  相似文献   

17.
Saprotrophic fungi were investigated for their bioweathering effects on the vanadium‐ and lead‐containing insoluble apatite group mineral, vanadinite [Pb5(VO4)3Cl]. Despite the insolubility of vanadinite, fungi exerted both biochemical and biophysical effects on the mineral including etching, penetration and formation of new biominerals. Lead oxalate was precipitated by Aspergillus niger during bioleaching of natural and synthetic vanadinite. Some calcium oxalate monohydrate (whewellite) was formed with natural vanadinite because of the presence of associated ankerite [Ca(Fe2+,Mg)(CO3)2]. Aspergillus niger also precipitated lead oxalate during growth in the presence of lead carbonate, vanadium(V) oxide and ammonium metavanadate, while abiotic tests confirmed the efficacy of oxalic acid in solubilizing vanadinite and precipitating lead as oxalate. Geochemical modelling confirmed the complexity of vanadium speciation, and the significant effect of oxalate. Oxalate–vanadium complexes markedly reduced the vanadinite stability field, with cationic lead(II) and lead oxalate also occurring. In all treatments and geochemical simulations, no other lead vanadate, or vanadium minerals were detected. This research highlights the importance of oxalate in vanadinite bioweathering and suggests a general fungal transformation of lead‐containing apatite group minerals (e.g. vanadinite, pyromorphite, mimetite) by this mechanism. The findings are also relevant to remedial treatments for lead/vanadium contamination, and novel approaches for vanadium recovery.  相似文献   

18.
A purified and electrophoretically homogeneous blue laccase from the litter-decaying basidiomycete Stropharia rugosoannulata with a molecular mass of approximately 66 kDa oxidized Mn2+ to Mn3+, as assessed in the presence of the Mn chelators oxalate, malonate, and pyrophosphate. At rate-saturating concentrations (100 mM) of these chelators and at pH 5.0, Mn3+ complexes were produced at 0.15, 0.05, and 0.10 μmol/min/mg of protein, respectively. Concomitantly, application of oxalate and malonate, but not pyrophosphate, led to H2O2 formation and tetranitromethane (TNM) reduction indicative for the presence of superoxide anion radical. Employing oxalate, H2O2 production, and TNM reduction significantly exceeded those found for malonate. Evidence is provided that, in the presence of oxalate or malonate, laccase reactions involve enzyme-catalyzed Mn2+ oxidation and abiotic decomposition of these organic chelators by the resulting Mn3+, which leads to formation of superoxide and its subsequent reduction to H2O2. A partially purified manganese peroxidase (MnP) from the same organism did not produce Mn3+ complexes in assays containing 1 mM Mn2+ and 100 mM oxalate or malonate, but omitting an additional H2O2 source. However, addition of laccase initiated MnP reactions. The results are in support of a physiological role of laccase-catalyzed Mn2+ oxidation in providing H2O2 for extracellular oxidation reactions and demonstrate a novel type of laccase-MnP cooperation relevant to biodegradation of lignin and xenobiotics.  相似文献   

19.
The acetogenic bacterium Clostridium thermoaceticum ATCC 39073 grew at the expense of the two-carbon substrates oxalate and glyoxylate. Other two-carbon substrates (acetaldehyde, acetate, ethanol, ethylene glycol, glycolaldehyde, glycolate, and glyoxal) were not growth supportive. Growth increased linearly with increasing substrate concentrations up to 45 mM oxalate and glyoxylate, and supplemental CO2 was not required for growth. Oxalate and glyoxylate yielded 4.9 and 9.4 g, respectively, of cell biomass (dry weight) per mol of substrate utilized. Acetate was the major reduced end product recovered from oxalate and glyoxylate cultures. 14C labeling studies showed that oxalate was subject to decarboxylation, and product analysis indicated that oxalate was utilized by the following reaction: 4-OOC-COO- + 5H2O → CH3COO- + 6HCO3- + OH-. Oxalate- and glyoxylate-dependent growth produced lower acetate concentrations per unit of cell biomass synthesized than did H2-, CO-, methanol-, formate-, O-methyl-, or glucose-dependent growth. Protein profiles of oxalate-grown cells were dissimilar from protein profiles of glyoxylate-, CO-, or formate-grown cells, suggesting induction of new proteins for the utilization of oxalate. C. thermoaceticum DSM 2955 and Clostridium thermoautotrophicum JW 701/3 also grew at the expense of oxalate and glyoxylate. However, oxalate and glyoxylate did not support the growth of C. thermoaceticum OMD (a nonautotrophic strain) or six other species of acetogenic bacteria tested.  相似文献   

20.
Oxalate is quantitated in both urine and plasma samples using reversed-phase ion-pair high-performance liquid chromatography (HPLC) with tris(2,2′-bipyridyl)ruthenium(II) [Ru(bpy)32+]-electrogenerated chemiluminescent (ECL) detection. Underivatized oxalate was separated on a reversed-phase column (Zorbax ODS) using a mobile phase of 10% methanol in 100 mM phsophate buffer at pH 7.0. The eluted compounds were combined with a stream of 2 mM Ru(bpy)32+ at a mixing tee before the ECL flow-cell. In the flow-cell, Ru(bpy)32+ is oxidized to Ru(bpy)32+ at a platinum electrode, and reacts with oxalate to produce chemiluminescence (CL). Urine samples were filtered and diluted prior to injection. Plasma samples were deproteinized before injection. A 25-μl aliquot of sample was injected for analysis. Possible interferants, including amino acids and indole-based compounds, present in biological samples were investigated. Without the separation, amino acids interfere by increasing the total observed CL intensity; this is expected because they give rise to CL emission on their own in reaction with Ru(bpy)33+. Indole compounds exhibit a unique interference by decreasing the CL signal when present with oxalate. Indoles inhibit their own CL emission at high concentration. By use of the indicated HPLC separation, oxalate was adequately separated from both types of interferants, which thus had no effect on the oxalate signal. Urine samples were assayed by both HPLC and enzymatic tests, the two techniques giving similar results, differing only by 1%. Detection limits were determined to be below 1 μM (1 nmol/ml) or 25 pmol injected. The working curve for oxalate was linear throughout the entire clinical range in both urine and plasma.  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