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.
AIMS: To isolate and characterize an oxalate-degrading Pandoraea sp. OXJ-11. METHODS AND RESULTS: A new bacterium Pandoraea sp. OXJ-11 was isolated from soil samples, which can grow in the medium with oxalate as the sole carbon and energy source. The isolate OXJ-11 is Gram-negative straight rod. It occurs singly and is motile by means of a double polar flagellum. Catalase is positive and nitrate is not reduced. It grows aerobically and the optimum growth temperature and the optimum pH are at 30 degrees C and pH 6.0, respectively. The polyphasic taxonomic data along with 16S rRNA sequence comparison demonstrate that the isolate OXJ-11 should belong to the genus Pandoraea and represent a new member in this family. CONCLUSIONS: Oxalate could be degraded and the oxalate-degrading enzyme activity was detected when the isolate OXJ-11 grew in the medium with oxalate as carbon source. SIGNIFICANCE AND IMPACT OF THE STUDY: Oxalate-degrading Pandoraea sp. OXJ-11 would be beneficial to the potential application in the control of sclerotinia stem rot in economically important plants caused by fungus Sclerotinia sclerotiorum, and in making plants resistant to the white mold disease by oxalate-degrading enzyme transgene.  相似文献   

2.
In order to establish a standard procedure for the spectrophotometric determination of urinary and plasma oxalate with oxalate oxidase (Laker, M.F., et al. (1980) Clin. Chem. 26, 827-830; Sugiura, M., et al. (1980) Clin. Chim. Acta 105, 393-399) and to define the limitations of the method, the procedures and reactions involved in the assay have been examined. Among the chromogenic hydrogen donors for peroxidase tested, a combination of 3-methyl-2-benzothiazolinone hydrazone (MBTH) and sodium N-sulfopropylaniline (HALPS) was found to be best for the oxalate determination under the conditions used. Urine contained substance(s) which were inhibitory to the measurement of hydrogen peroxide by the peroxidase-catalyzed oxidative condensation of MBTH and HALPS, but they were largely removed by charcoal treatment at pH 5.6 without significant loss of oxalate. Deproteinization of plasma was carried out by ultrafiltration through a membrane cone (Centriflo CF-25) at neutral pH. The plasma oxalate ultrafiltrability under the conditions employed was calculated to be approximately 95%. A standard assay system for oxalate in these urine and plasma samples was then set up based on a series of studies on the reactions involved in the assay. In the case of normal plasma, however, the absorbance change was very small due to the low concentration of oxalate, and in addition, pretreatment of plasma with excess oxalate decarboxylase followed by the ultrafiltration and oxalate determination did not abolish completely the oxalate oxidase-dependent absorbance increase. It was concluded that the enzymic method was useful for the assay of urinary oxalate and in detecting elevated levels of plasma oxalate such as those in hemodialysis patients but was not sensitive enough to determine accurately the normal or decreased level of oxalate in plasma. The apparent concentration of oxalate in normal human plasma was measured in this work as 3.5 +/- 0.8 microM (mean +/- S.D., n = 8), and this result was interpreted to mean that the concentration of plasma oxalate was less than approximately 3.5 microM, as estimated by the present method.  相似文献   

3.
Oxalate decarboxylase is a manganese-dependent enzyme that catalyzes the conversion of oxalate to formate and carbon dioxide. We have determined the structure of oxalate decarboxylase from Bacillus subtilis at 1.75 A resolution in the presence of formate. The structure reveals a hexamer with 32-point symmetry in which each monomer belongs to the cupin family of proteins. Oxalate decarboxylase is further classified as a bicupin because it contains two cupin folds, possibly resulting from gene duplication. Each oxalate decarboxylase cupin domain contains one manganese binding site. Each of the oxalate decarboxylase domains is structurally similar to oxalate oxidase, which catalyzes the manganese-dependent oxidative decarboxylation of oxalate to carbon dioxide and hydrogen peroxide. Amino acid side chains in the two metal binding sites of oxalate decarboxylase and the metal binding site of oxalate oxidase are very similar. Four manganese binding residues (three histidines and one glutamate) are conserved as well as a number of hydrophobic residues. The most notable difference is the presence of Glu333 in the metal binding site of the second cupin domain of oxalate decarboxylase. We postulate that this domain is responsible for the decarboxylase activity and that Glu333 serves as a proton donor in the production of formate. Mutation of Glu333 to alanine reduces the catalytic activity by a factor of 25. The function of the other domain in oxalate decarboxylase is not yet known.  相似文献   

4.
Bacillus subtilis YvrK is an acid-induced oxalate decarboxylase   总被引:5,自引:0,他引:5       下载免费PDF全文
Bacillus subtilis has been shown to express a cytosolic oxalate decarboxylase (EC 4.1.1.2). The enzyme was induced in acidic growth media, particularly at pH 5.0, but not by oxalate. The enzyme was purified, and N-terminal sequencing identified the protein to be encoded by yvrK. The role of the first oxalate decarboxylase to be identified in a prokaryote is discussed.  相似文献   

5.
The metabolism of hydroxypyruvate to oxalate was studied in isolated rat hepatocytes. [14C]Oxalate was produced from [2-14C]- and [3-14C]- but not [1-14C]hydroxypyruvate. No oxalate was produced from similarly labeled pyruvate. The mechanism by which hydroxypyruvate is metabolized to oxalate involves decarboxylation at the carbon 1 position as the initial step. This activity was distinct from that which produced CO2 from the carbon 1 position of pyruvate. Hydroxypyruvate decarboxylase activity was found mainly in the mitochondria, with the remainder (25%) in the cytosol. No activity was present in the peroxisomes, the probable site of oxalate production from glycolate and glyoxylate. Hydroxypyruvate, but not pyruvate stimulated [14C]oxalate production from [U-14C]fructose, suggesting that hydroxypyruvate is either an intermediate in the fructose-oxalate pathway, or that it prevents carbon from leaving that pathway. The lack of effect of pyruvate in this regard is evidence against redox being the primary effect of hydroxypyruvate and focuses attention on hydroxypyruvate and its precursors as important sources of carbon for oxalate synthesis from both carbohydrate and protein.  相似文献   

6.
The equilibrium parameters for calcium oxalate solubility in tissue culture media were investigated because of the current interest in oxalate toxicity. The calcium selective ion electrode methodology was evaluated and calcium concentrations from potentiometric calculations were verified by d-c argon plasma emission spectroscopy. The experimental K(sp)'s at 25 degrees C for Dulbecco's modified Eagle media and McCoys 5A media are equivalent to the literature K(sp) of 2.3 x 10(-9) for low ionic strength. The equilibrium concentration products, [Ca2+] [C2O2-(4)], are ten times higher than the K(sp)'s due to the high ionic strengths of tissue culture media. At 37 degrees C, addition of soluble oxalate at the 10(-3) to 10(-4) M level causes >50% precipitation of the oxalate resulting in equilibrium oxalate concentrations of less than 6 x 10(-5) M. This relatively inexpensive selective ion technique allows the determination of oxalate concentrations in equilibrium-saturated media which are substantially less than those calculated by the amount of soluble oxalate added to the media.  相似文献   

7.
Oxalate decarboxylase converts oxalate to formate and carbon dioxide and uses dioxygen as a cofactor despite the reaction involving no net redox change. We have successfully used Fourier transform infrared spectroscopy to monitor in real time both substrate consumption and product formation for the first time. The assignment of the peaks was confirmed using [(13)C]oxalate as the substrate. The K(m) for oxalate determined using this assay was 3.8-fold lower than that estimated from a stopped assay. The infrared assay was also capable of distinguishing between oxalate decarboxylase and oxalate oxidase activity by the lack of formate being produced by the latter. In D(2)O, the product with oxalate decarboxylase was C-deuterio formate rather than formate, showing that the source of the hydron was solvent as expected. Large solvent deuterium kinetic isotope effects were observed on V(max) (7.1 +/- 0.3), K(m) for oxalate (3.9 +/- 0.9), and k(cat)/K(m) (1.8 +/- 0.4) indicative of a proton transfer event during a rate-limiting step. Semiempirical quantum mechanical calculations on the stability of formate-derived species gave an indication of the stability and nature of a likely enzyme-bound formyl radical catalytic intermediate. The capability of the enzyme to bind formate under conditions in which the enzyme is known to be active was determined by electron paramagnetic resonance. However, no enzyme-catalyzed exchange of the C-hydron of formate was observed using the infrared assay, suggesting that a formyl radical intermediate is not accessible in the reverse reaction. This restricts the formation of potentially harmful radical intermediates to the forward reaction.  相似文献   

8.
Carbon-14 was incorporated from citrate-1,5-14C, glyoxylate-14C(U), or glyoxylate-1-14C into oxalate by cultures of Aspergillus niger pregrown on a medium with glucose as the sole source of carbon. Glyoxylate-14C(U) was superior to glyoxylate-1-14C and citrate-1,5-14C as a source of incorporation. By addition of a great amount of citrate the accumulation of oxalate was accelerated and its maximum yield increased. In a cell-free extract from mycelium forming oxalate from citrate the enzyme oxaloacetate hydrolase (EC3.7.1.1) was identified. Its in vitro activity per flask exceeded the rate of in vivo accumulation of oxalate. Glyoxylate oxidizing enzymes (glycolate oxidase, EC1.1.3.1; glyoxylate oxidase, EC1.2.3.5;NAD(P)-dependent glyoxylate dehydrogenase; glyoxylate dehydrogenase, CoA-oxalylating, EC1.2.1.7) could not be detected in cell-free extracts. It is concluded that in cultures accumulating oxalate from citrate after pregrowth on glucose, oxalate arises by hydrolytic cleavage of oxaloacetate but not by oxidation of glyoxylate.  相似文献   

9.
Calcium oxalate formation in Lemna minor L. occurs in structurally specialized cells called crystal idioblasts. Cytochemical and immunocytochemical protocols were employed to study the distribution of peroxisomes and the enzymes glycolate oxidase, glycine decarboxylase and ribulose 1,5-bisphosphate carboxylase-oxygenase (RuBisCO) in relation to synthesis of oxalate used for Ca oxalate formation. These enzymes are necessary for photorespiratory glycolate synthesis and metabolism. Using catalase cytochemistry, microbodies were found to exist in crystal idioblasts but were smaller and fewer than those found in mesophyll cells. Glycolate oxidase, which can oxidize glycolate to oxalate via glyoxylate, could not be found in microbodies of crystal idioblasts at any stage of development. This enzyme increased in amount in microbodies of mesophyll cells as they matured and could even be found in dense amorphous inclusions of mature cell peroxisomes. Glycine decarboxylase and RuBisCO could also be detected in increasing amount in mesophyll cells as they matured but could not be detected in idioblasts or were just detectable. Thus, Lemna idioblasts lack the machinery for synthesis of oxalate from glycolate. Based on these results and other available information, two general models for the generation and accumulation of oxalate used for Ca oxalate formation in crystal idioblasts are proposed. The biochemical specialization of crystal idioblasts indicated by this study is also discussed with respect to differentiation of cellular structure and function.  相似文献   

10.
Oxalate decarboxylases and oxalate oxidases are members of the cupin superfamily of proteins that have many common features: a manganese ion with a common ligand set, the substrate oxalate, and dioxygen (as either a unique cofactor or a substrate). We have hypothesized that these enzymes share common catalytic steps that diverge when a carboxylate radical intermediate becomes protonated. The Bacillus subtilis decarboxylase has two manganese binding sites, and we proposed that Glu162 on a flexible lid is the site 1 general acid. We now demonstrate that a decarboxylase can be converted into an oxidase by mutating amino acids of the lid that include Glu162 with specificity switches of 282,000 (SEN161-3DAS), 275,000 (SENS161-4DSSN), and 225,000 (SENS161-4DASN). The structure of the SENS161-4DSSN mutant showed that site 2 was not affected. The requirement for substitutions other than of Glu162 was, at least in part, due to the need to decrease the Km for dioxygen for the oxidase reaction. Reversion of decarboxylase activity could be achieved by reintroducing Glu162 to the SENS161-4DASN mutant to give a relative specificity switch of 25,600. This provides compelling evidence for the crucial role of Glu162 in the decarboxylase reaction consistent with it being the general acid, for the role of the lid in controlling the Km for dioxygen, and for site 1 being the sole catalytically active site. We also report the trapping of carboxylate radicals produced during turnover of the mutant with the highest oxidase activity. Such radicals were also observed with the wild-type decarboxylase.  相似文献   

11.
We used ethylenediaminetetraacetic acid dianhydride (EDTAD) to modify oxalate decarboxylase (OXDC) to improve its adsorption on calcium oxalate stones. The modified sites were identified by Ultra performance liquid chromatography-mass spectrometry (UPLC-MS) and the adsorption mechanism of the EDTAD-modified OXDC on calcium oxalate (CaOx) was investigated. We investigated adsorption time, initial enzyme concentration, temperature and solution pH on the adsorption process. Data were analyzed using kinetics, thermodynamics and isotherm adsorption models. UPLC-MS showed that EDTAD was attached to OXDC covalently and suggested that the chemical modification occurred at both the free amino of the side chain and the α-NH2 of the peptide. The adsorption capacity of the EDTAD-OXDC on calcium oxalate was 53.37% greater than that of OXDC at the initial enzyme concentration of 5 mg/ml, pH = 7.0, at 37° C. The modified enzyme (EDTAD-OXDC) demonstrated improved oxalate degradation activity at pH 4.5?6.0. Kinetic data fitting analysis suggested a pseudo second order kinetic model. Estimates of the thermodynamic parameters including ΔG0, ΔH0 and ΔS0 of the adsorption process showed it to be feasible, spontaneous and endothermic. Isotherm data fitting analysis indicated that the adsorption process is reduced to monolayer adsorption at a low enzyme concentration and to multilayer adsorption at a high enzyme concentration. It may be possible to apply OXDC to degradation of calcium oxalate stones.  相似文献   

12.
EPR spin trapping experiments on bacterial oxalate decarboxylase from Bacillus subtilis under turn-over conditions are described. The use of doubly (13)C-labeled oxalate leads to a characteristic splitting of the observed radical adducts using the spin trap N-tert-butyl-α-phenylnitrone linking them directly to the substrate. The radical was identified as the carbon dioxide radical anion which is a key intermediate in the hypothetical reaction mechanism of both decarboxylase and oxidase activities. X-ray crystallography had identified a flexible loop, SENS161-4, which acts as a lid to the putative active site. Site directed mutagenesis of the hinge amino acids, S161 and T165 was explored and showed increased radical trapping yields compared to the wild type. In particular, T165V shows approximately ten times higher radical yields while at the same time its decarboxylase activity was reduced by about a factor of ten. This mutant lacks a critical H-bond between T165 and R92 resulting in compromised control over its radical chemistry allowing the radical intermediate to leak into the surrounding solution.  相似文献   

13.
Sclerotinia sclerotiorum D-E7 was studied to determine the impact of nutritional supplements and monosaccharides on growth, oxalate accumulation, and culture pH in broth media (initial pH c. 5). Cultures with 0.1% nutritional supplement (tryptone, yeast extract, or soytone) yielded minimal growth, 2-3 mM oxalate, and a final culture pH of 4.2-4.8. In contrast, cultures with 0.1% nutritional supplement and 25 mM glucose yielded significant growth, minimal oxalate (<1 mM), and a final culture pH of 2.8-3.7. Similar trends were observed when glucose in 0.1% soytone cultures was replaced with 25 mM d-mannose, l-arabinose, or d-xylose. With 1% soytone-25 mM glucose cultures, growth and oxalate accumulation ( approximately 21 mM) occurred with little change in initial pH. This was not the case with 1% soytone-250 mM glucose cultures; increased glucose levels resulted in a decrease in oxalate accumulation ( approximately 7 mM) and in final culture pH (3.4). Time-course studies with these cultures revealed that oxalate accumulation was suppressed during growth when the culture pH dropped to <4. Overall, these results indicate that (1) the decrease in external pH (i.e. acidification) was independent of oxalate accumulation and (2) acidification coupled to glucose-dependent growth regulated oxalate accumulation by Sclerotinia sclerotiorum.  相似文献   

14.
Enzymatic oxalate decarboxylation in isolates of Sclerotinia sclerotiorum   总被引:4,自引:0,他引:4  
Abstract Sclerotinia sclerotiorum isolates B24 (virulent) and SS41 (hypovirulent) possess oxalate decarboxylase. Production was regulated by composition and pH of culture medium and required the presence of oxalate or its precursor, succinic acid, as inducers. Mycelia of both isolates contain equivalent amounts of enzyme.  相似文献   

15.
Oxalate oxidase is thought to be involved in the production of hydrogen peroxide for lignin degradation by the dikaryotic white rot fungus Ceriporiopsis subvermispora. This enzyme was purified, and after digestion with trypsin, peptide fragments of the enzyme were sequenced using quadrupole time-of-flight mass spectrometry. Starting with degenerate primers based on the peptide sequences, two genes encoding isoforms of the enzyme were cloned, sequenced, and shown to be allelic. Both genes contained 14 introns. The sequences of the isoforms revealed that they were both bicupins that unexpectedly shared the greatest similarity to microbial bicupin oxalate decarboxylases rather than monocupin plant oxalate oxidases (also known as germins). We have shown that both fungal isoforms, one of which was heterologously expressed in Escherichia coli, are indeed oxalate oxidases that possess < or =0.2% oxalate decarboxylase activity and that the organism is capable of rapidly degrading exogenously supplied oxalate. They are therefore the first bicupin oxalate oxidases to have been described. Heterologous expression of active enzyme was dependent on the addition of manganese salts to the growth medium. Molecular modeling provides new and independent evidence for the identity of the catalytic site and the key amino acid involved in defining the reaction specificities of oxalate oxidases and oxalate decarboxylases.  相似文献   

16.
Many plants accumulate crystals of calcium oxalate. Just how these crystals form remains unknown. To gain insight into the mechanisms regulating calcium oxalate crystal formation, a crystal engineering approach was initiated utilizing the non-crystal-accumulating plant, Arabidopsis. The success of this approach hinged on the ability to transform Arabidopsis genetically into a calcium oxalate crystal-accumulating plant. To accomplish this transformation, two oxalic acid biosynthetic genes, obcA and obcB, from the oxalate-secreting phytopathogen, Burkholderia glumae were inserted into the Arabidopsis genome. The co-expression of these two bacterial genes in Arabidopsis conferred the ability not only to produce a measurable amount of oxalate but also to form crystals of calcium oxalate. Biochemical and cellular studies of crystal accumulation in Arabidopsis revealed features that are similar to those observed in the cells of crystal-forming plants. Thus, it appears that at least some of the basic components that comprise the calcium oxalate crystal formation machinery are conserved even in non-crystal-accumulating plants.  相似文献   

17.
Intravenously injected 14C labelled oxalate was rapidly removed from the blood stream via the kidney in 2 sheep, 75% being cleared within 8 h. Mean daily urinary oxalate excretions over 5 days were 21-2 and 27-5 mg and the derived plasma oxalate concentrations were 52-6 and 74-4 mug/100 ml, respectively. Oxalate was both filtered and secreted by the renal tubule with oxalate/inulin ratios varying from 1-11 to 1-57 in 6 normal sheep. A large increase in calcium excretion induced by calcium borogluconate infusion over 5 days was accompanied by a small but consistent increase in urinary oxalate excretion relative to calcium. Oxalate in blood was to be found mainly in the plasma, there being a small (8%) proporation within erythrocytes. This is lower than that reported for man, and yet in its excretion of oxalate via the kidney the sheep appears to closely resemble man and dog.  相似文献   

18.
This contribution describes the trapping of the hydroperoxyl radical at a pH of 4 during turnover of wild-type oxalate decarboxylase and its T165V mutant using the spin-trap BMPO. Radicals were detected and identified by a combination of EPR and mass spectrometry. Superoxide, or its conjugate acid, the hydroperoxyl radical, is expected as an intermediate in the decarboxylation and oxidation reactions of the oxalate monoanion, both of which are promoted by oxalate decarboxylase. Another intermediate, the carbon dioxide radical anion was also observed. The quantitative yields of superoxide trapping are similar in the wild type and the mutant while it is significantly different for the trapping of the carbon dioxide radical anion. This suggests that the two radicals are released from different sites of the protein.  相似文献   

19.
Tian H  Jiang L  Liu E  Zhang J  Liu F  Peng X 《Physiologia plantarum》2008,133(2):180-189
Oxalate, a common constituent in many plants, is known to play important functional roles in plants. However, excess levels of oxalate in edible parts of plants adversely affect their quality as food. Understanding the regulatory mechanism in plants, particularly in food crops, is of both scientific and practical significance. While a number of studies have shown that nitrate can efficiently induce oxalate accumulation in plants, how it elicits such an effect is not well understood. This study aimed to gain a further insight into the mechanism underlying the nitrate-induced oxalate accumulation. Nitrate-N efficiently caused oxalate accumulation in rice leaves, depending on the nitrate concentrations and treatment time. In contrast, same nitrogen molar levels of the other N forms such as nitrite, ammonium, glutamate and urea either had no effect on the accumulation or even reduced the oxalate level. When glutamate, glutamine, asparate and asparagine were added into the nutrient solution that already contained saturating concentration of nitrate, both oxalate levels and NR activity were correspondingly decreased. In all of these modes of treatment, the change in NR activity was positively paralleled to that in oxalate levels. For a further confirmation, we generated the transgenic rice plants with a NR interference gene introduced. The result further demonstrated that in the transgenic plants, unlike in wild-type plants, oxalate was no longer able to accumulate in response to the nitrate treatment even though the endogenous nitrate levels were substantially elevated. Taken together, our results suggest that the nitrate-induced oxalate accumulation in rice leaves is dependent on the NR-catalyzed nitrate reduction, rather than on nitrate itself or nitrite reduction or its downstream metabolites.  相似文献   

20.
A new method for rapid determination of oxalic acid was developed using oxalate oxidase and a biosensor based on SIRE (sensors based on injection of the recognition element) technology. The method was selective, simple, fast, and cheap compared with other present detection systems for oxalate. The total analysis time for each assay was 2-9 min. A linear range was observed between 0 and 5 mM when the reaction conditions were 30 degrees C and 60 s. The linear range and upper limit for concentration determination could be increased to 25 mM by shortening the reaction time. The lower limit of detection in standard solutions, 20 microM, could be achieved by means of modification of the reaction conditions, namely increasing the temperature and the reaction time. The biosensor method was compared with a conventional commercially available colorimetric method with respect to the determination of oxalic acid in urine samples. The urine oxalic acid concentrations determined with the biosensor method correlated well (R=0.952) with the colorimetric method.  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