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.
The interaction of phytic acid with Ca(II) has been studied by potentiometric titration and by measurement of free Ca(II) concentrations using an ion Selective electrode. With increasing Ca(II) concentration, the titration curve of phytic acid is displaced to regions of lower pH. In the binding of calcium ions to phytic acid, there is no evidence that significant binding occurs below approximately pH 5. Above this pH, the extent of binding is dependent upon both pH and the calcium to phytic acid ratios. Maximum binding obtains at a Ca(II):phytate ratio of 6 with 4.8 mol of Ca(II) bound per mol of phytate above pH ca. 8. Binding constants are apparently very large since binding isotherms at any Ca(II):phytate ratio are a linear function of the total calcium ion concentration. In all cases, binding occurs only when one or more phosphate groups have been converted to the oxo dianion form. The apparent pK' values (curve-fit parameters) that describe the potentiometric titration data are in good agreement with the constants evaluated from the binding of Ca(II) to phytate as a function of pH. Using CPK space-filling models, structures containing six metal ions in coordinate linkage to pairs of oxo dianions have been constructed and discussed within the framework of the axial conformation of phytic acid and the order of proton removal with an increase in pH based upon NMR studies.  相似文献   

2.
Using an isoperibolic titration microcalorimeter, the ionization characteristics and associated heat changes of phytic acid (myo-inositol hexaphosphate) and phytic acid in the presence of varying Zn(II) concentrations have been examined over the pH range 2.5–11 at 25°C in 0.2 M KCl. In the absence of Zn(II), ca. 7 of the 12 ionizable protons in phytic acid are titrated in this pH range with ionization heats varying from ca. 2 to −3 kcal-mol-1. At Zn(II): phytate mol ratios of 4:1 and greater, the dissociation of all protons and complex formation of phytate with Zn(II) occurs below pH 6. From the difference titration curves of phytic acid plus Zn(II) versus Zn(II) alone, ca. 3.5 mol Zn(II) bind per mol phytate. Since Zn(II):phytate complexes are insoluble, the observed heat changes contain contributions not only from heats of precipitation but also from binding, ionization, neutralization, and hydration effects. From the heat change for the titration of (a) phytic acid, pH 2.6–10.4; (b) phytic acid + Zn(II), pH 2.6–6.1; and (c) Zn(II), pH 2.6–6.1 at Zn(II): phytate ratios of 4 to 10, the value of 24.7 ± 0.5 kcal mol−1 phytate has been obtained for the binding of 3.5 mols Zn(II). This figure also includes the heat of precipitation of the complex. In pH-drop experiments, with the initial pH at 8.65, the value of 23.9 kcal mol-1 was obtained for ΔH°. Hysteresis effects are prevalent in these reaction solutions. Time-dependent changes in pH occur with a change in pH. For the phytate-Zn(II) reactions, the time-course curves are biphasic and fit a rate equation for two simultaneous first order reactions. Hysteresis effects seen in the titration of Zn(II) fit simple first-order kinetics. These effects most probably arise from the ejection of a proton from the aqua ion or aqua ion ligand complex(es).  相似文献   

3.
The interaction of the cupric ion with phytic acid as a function of pH has been studied by potentiometric and thermal titration and by the determination of ligand binding. As has been found for the reaction of zinc and calcium cations with phytate, the presence of the Cu(II) ion results in a displacement of the titration curves to more acid values. Evaluation of the parameters that describe such changes in ionization behavior by curve-fit analysis showed that as the Cu(II):phytate mol ratio was increased from one to eight, the pK' values of the ionizable group sets of phytic acid (ranging from 1.59 to 9.79) were consolidated into just two sets with curve-fit (CP) values ca. 1.5 and 3.7. Marked pH hysteresis effects are seen in such systems because of the pronounced acid strength of the Cu(II):aqua ion and the Cu(II) ligand aqua ion complex. The combined heat of binding and precipitation (plus solvation changes, etc.) of Cu(II) to phytate is endothermic (21.8–22.2 kcal mol−1). This is similar in magnitude to that reported for the binding of either Zn(II) or Ca(II) to phytate. In the titration of Cu(NO3)2 with KOH, presumably to form Cu(OH)2, ΔH° was exothermic (−12.5 kcal mol−1). From measurements of free Cu(II) cation concentration in the presence of phytate the binding reaction was found to be stoichiometric with 6 mols Cu(II) bound at pH 6. Binding occurs within the pH range 2–6. An apparent necessary requirement for binding is the availability of the oxo dianion structure formed from the second dissociation step of a phosphoryl group. Curve-fit analysis of the binding data as a function of pH showed that a group or group set with CP value ca. 4 governs the binding reaction(s) at all mol ratios of Cu(II) to phytate examined. It is suggested that the binding of cupric ions to phytate may occur to the equatorial rather than the axial configuration as suggested for Ca(II) binding. A space-filling molecular model to illustrate this has been constructed. Soluble Cu(II):phytate complexes are formed within the pH range from 2 to ca. 3.4. This is supported by the results of difference absorption spectrometry.  相似文献   

4.
Phytic acid is a major determinant of zinc bioavailability. Little is known about phytic acid intakes or indices of zinc bioavailability in type 2 diabetes mellitus (DM), a condition that predisposes to zinc deficiency. The aim of this cross-sectional study was to measure and explore the relationships among phytic acid intake, zinc bioavailability, and molecular markers of zinc homeostasis in 20 women with DM compared to 20 healthy women. The phytate/zinc, (calcium)(phytate)/zinc, and (calcium + magnesium)(phytate)/zinc molar ratios were used to indicate zinc bioavailability. Plasma zinc concentrations and zinc transporter (ZnT1, ZnT8, and Zip1) gene expression in mononuclear cells were measured. Participants with DM consumed 1,194?±?824?mg/day (mean?±?SD) phytic acid, an amount similar to the intake of healthy women (1,316?±?708?mg/day). Bread products and breakfast cereals contributed more than 40?% of the phytic acid intake in each group. A positive relationship was observed in all participants between phytic acid and dietary fiber (r?=?0.6, P?<?0.001) and between dietary fiber and the (calcium)(phytate)/zinc ratio (r?=?0.5, P?<?0.001). Compared to the healthy group, the messenger RNA ratio of ZnT1 (zinc export) to Zip1 (zinc import) was lower in participants with DM, which may indicate perturbed zinc homeostasis in the disorder. The plasma zinc concentration was not predicted by age, body mass index, health status, zinc bioavailability, or zinc transporter expression. Healthy and diabetic women consume phytic acid in amounts that are likely to decrease the bioavailability of dietary zinc. Recommendations to consume greater amounts of dietary fiber, much of which is associated with phytate, increase the risk of zinc deficiency.  相似文献   

5.
For the past 50 years laboratory animals have been used to ascertain the metabolic bases for signs of zinc deficiency such as sharply reduced food intake, severe dermatitis, slow wound healing, delayed sexual development and function, reduced immunocompetence, severe teratogenic abnormalities, and abnormal metabolism of carbohydrate, lipid, and protein. Current evidence indicates that many of these symptoms may be consequences of inhibition of early steps in nucleic acid metabolism that lead to problems with cellular replication and growth and also that zinc plays an important role in membrane structure and function. Bioavailability of zinc to experimental animals was early shown to be reduced by plant protein diets and to be further reduced by feeding excess calcium. Current evidence indicates phytic acid in plant proteins to be a major inhibitor of zinc absorption, although food-processing methods can either increase or decrease zinc bioavailability. The inhibitory effect of phytic acid is very dependent on dietary calcium in association with phytate and zinc. Usual calcium intakes by humans are much below those demonstrated in animals to cause phytate inhibition of dietary zinc availability.  相似文献   

6.
BACKGROUND: Phytases hydrolyze phytic acid (myo-inositol-hexakisphosphate) to less-phosphorylated myo-inositol derivatives and inorganic phosphate. Phytases are used in animal feed to reduce phosphate pollution in the environment. Recently, a thermostable, calcium-dependent Bacillus phytase was identified that represents the first example of the beta propeller fold exhibiting phosphatase activity. We sought to delineate the catalytic mechanism and property of this enzyme. RESULTS: The crystal structure of the enzyme in complex with inorganic phosphate reveals that two phosphates and four calcium ions are tightly bound at the active site. Mutation of the residues involved in the calcium chelation results in severe defects in the enzyme's activity. One phosphate ion, chelating all of the four calcium ions, is close to a water molecule bridging two of the bound calcium ions. Fluoride ion, which is expected to replace this water molecule, is an uncompetitive inhibitor of the enzyme. The enzyme is able to hydrolyze any of the six phosphate groups of phytate. CONCLUSIONS: The enzyme reaction is likely to proceed through a direct attack of the metal-bridging water molecule on the phosphorous atom of a substrate and the subsequent stabilization of the pentavalent transition state by the bound calcium ions. The enzyme has two phosphate binding sites, the "cleavage site", which is responsible for the hydrolysis of a substrate, and the "affinity site", which increases the binding affinity for substrates containing adjacent phosphate groups. The existence of the two nonequivalent phosphate binding sites explains the puzzling formation of the alternately dephosphorylated myo-inositol triphosphates from phytate and the hydrolysis of myo-inositol monophosphates.  相似文献   

7.
This review describes the present state of knowledge about phytic acid (phytate), which is often present in legume seeds. The antinutritional effects of phytic acid primarily relate to the strong chelating associated with its six reactive phosphate groups. Its ability to complex with proteins and particularly with minerals has been a subject of investigation from chemical and nutritional viewpoints. The hydrolysis of phytate into inositol and phosphates or phosphoric acid occurs as a result of phytase or nonenzymatic cleavage. Enzymes capable of hydrolysing phytates are widely distributed in micro-organisms, plants and animals. Phytases act in a stepwise manner to catalyse the hydrolysis of phytic acid. To reduce or eliminate the chelating ability of phytate, dephosphorylation of hexa- and penta-phosphate forms is essential since a high degree of phosphorylation is necessary to bind minerals. There are several methods of decreasing the inhibitory effect of phytic acid on mineral absorption (cooking, germination, fermentation, soaking, autolysis). Nevertheless, inositol hexaphosphate is receiving increased attention owing to its role in cancer prevention and/or therapy and its hypocholesterolaemic effect.  相似文献   

8.
9.
Phytate hydrolysis by germfree and conventional rats.   总被引:1,自引:1,他引:0       下载免费PDF全文
Phytic acid is naturally occurring compound that reduces intestinal absorption of many metals. Early work suggests that some dietary phytate may be hydrolyzed in the large intestines by bacteria, but more recently nutritionists have suggested that a mucosal enzyme is responsible. This paper reports a study intended to resolve this controversy. The hydrolysis of dietary phytic acid was measured in germfree and conventional rats fed either of two diets that differed in their calcium content. Negligible phytate hydrolysis occurred in the germfree rats, whereas 22 and 56% of the phytic acid was hydrolyzed by conventional rats fed high- and low-calcium diets, respectively. We concluded that bacteria were responsible for the hydrolysis of phytate in these diets and that any activity of endogenous enzyme was negligible.  相似文献   

10.
Phytases hydrolyze phytic acid to less phosphorylated myo-inositol derivatives and inorganic phosphate. A thermostable phytase is of great value in applications for improving phosphate and metal ion availability in animal feed, and thereby reducing phosphate pollution to the environment. Here, we report a new folding architecture of a six-bladed propeller for phosphatase activity revealed by the 2.1 A crystal structures of a novel, thermostable phytase determined in both the partially and fully Ca2+-loaded states. Binding of two calcium ions to high-affinity calcium binding sites results in a dramatic increase in thermostability (by as much as approximately 30 degrees C in melting temperature) by joining loop segments remote in the amino acid sequence. Binding of three additional calcium ions to low-affinity calcium binding sites at the top of the molecule turns on the catalytic activity of the enzyme by converting the highly negatively charged cleft into a favorable environment for the binding of phytate.  相似文献   

11.
Ability of some strains of lactic acid bacteria to degrade phytic acid   总被引:2,自引:0,他引:2  
Twelve strains of lactic acid bacteria were examined for their ability to degrade phytate. In media in which phytic acid was the source of phosphate, phytate degradation was observed. Phytate disappearance may however not only be due to phytase, as phytic acid coprecipitated with protein as a consequence of a fall in pH during fermentation.  相似文献   

12.
The interactions of phytic acid with Cu(II) and Zn(II) ions were examined as functions of metal ion concentrations and pH. Cu(II) ion-selective potentiometric and electron spin resonance (ESR) experiments provide strong evidence for the binding of Cu(II) ions to the phytic acid molecule at low pH (2.4–3.4) values. The relative stabilities of the copper and zinc phytates at low pH values were found to be very similar. For systems with metal ion:phytic acid molar ratios of 1:1–4:1 and 5:1–6:1 and pH values in the 3.4–5.9 and 3.4–5.0 ranges, respectively, Zn(II) ions were found to form complexes with phytic acid that were more stable than those of Cu(II) ions with phytic acid. The phytic acid molecule, however, was found to accommodate Cu(II) ions more readily than Zn(II) ions. For example, in systems containing equal amounts of Cu(II) and Zn(II) ions, 2 Zn(II) ions and 2, 3, 4, or 4.5 Cu(II) ions were found per phytic acid molecule depending upon metal ion:phytic acid molar ratios in the systems and pH. Total metal ion:phytic acid molar ratios and pH affected resultant metal ion solubilities and were factors influencing the effects of Zn(II) and Cu(II) ions on the binding of each other by phytic acid. Zn(II) and Cu(II) ions were observed to potentiate the binding of each other by phytic acid in some systems and compete with each other for phytate binding sites in others.  相似文献   

13.
The heats of complex formation of Cu(II) with phytic acid to form soluble complexes in the absence of precipitation at acid pH have been measured. The reaction was examined over a wide range of mol ratios of Cu(II):phytate. In all cases the heats of reaction were endothermic. Measurements of the uncombined copper by use of a copper electrode allowed calculation of the combined copper and hence the enthalpies. These latter values varied to some extent, depending upon the Cu(II):phytate ratio and the pH region where the reaction was examined. Factors which could contribute to the variation in the enthalpy terms include changes in the heats of ionization and possible structural changes with Cu(II) bound.  相似文献   

14.
Phytic acid. A natural antioxidant   总被引:14,自引:0,他引:14  
The catalysis by iron of radical formation and subsequent oxidative damage has been well documented. Although many iron-chelating agents potentiate reactive oxygen formation and lipid peroxidation, phytic acid (abundant in edible legumes, cereals, and seeds) forms an iron chelate which greatly accelerates Fe2+-mediated oxygen reduction yet blocks iron-driven hydroxyl radical generation and suppresses lipid peroxidation. Furthermore, high concentrations of phytic acid prevent browning and putrefaction of various fruits and vegetables by inhibiting polyphenol oxidase. These observations indicate an important antioxidant function for phytate in seeds during dormancy and suggest that phytate may be a substitute for presently employed preservatives, many of which pose potential health hazards.  相似文献   

15.
王忠华 《植物学报》2005,22(4):463-470
植酸是玉米(Zea mays)、小麦(Triticum aestivum)、大麦(Hordeum vulgare)、水稻(Oryza sativa)和大豆(Glvcine max)等籽粒中广泛存在的一种有机酸(6-肌醇磷酸), 其与K+、Ca2+、Mg2+和Fe3+等金属离子形成的植酸盐是微量营养元素的重要贮存形式。植酸及植酸盐不能被人和非反刍动物所吸收利用; 植酸摄入体内后还会和其他来源的微量营养元素结合形成植酸盐, 造成这些营养元素的生物有效性下降, 从而造成微量元素缺乏症。此外, 大量的植酸及植酸盐随粪便排出, 造成严重的环境污染, 尤其是水体富营养化。由于土壤中缺乏分解微生物, 即使畜禽粪便作有机肥还田仍不能被作物吸收利用。近年来, 利用理化诱变与转基因技术已成功地获得了玉米、大麦、水稻和大豆等作物的低植酸突变体。本文对植酸的生物合成过程、低植酸突变体的诱发与研究、低植酸突变体的遗传特征与可能机理及营养评价进行了综述, 并对低植酸作物的应用前景进行了简要分析。  相似文献   

16.
磷是植物生长发育所必需的大量营养元素。在种子发育过程中,植酸是磷的贮存库,对维持植物体内磷平衡有重要的作用。在种子萌发过程中,植酸酶分解植酸盐,释放磷、矿质营养和肌醇供幼苗生长。本文综述了近年来植物(作物)种子中植酸的生物合成途径、种子植酸含量的遗传、低植酸作物的育种等研究进展。首先,植酸生物合成途径中最初的反应底物为葡萄糖-6-磷酸,形成肌醇后,以肌醇为底物合成植酸共有两条路径:依赖脂类与不依赖脂类,目前,已分离鉴定若干植酸合成所需的关键酶及其编码基因,包括肌醇-3-磷酸合成酶、肌醇激酶、肌醇多磷酸盐激酶,以及参与植酸运输的ATP结合盒转运子。其次,利用作图群体及关联分析群体,分别在水稻(Oryza sativa L.)、白菜(Brassica rapa L.)、菜豆(Phaseolus vulgaris L.)等植物中鉴定出多个与种子植酸磷含量相关的遗传位点。第三,筛选获得有价值的低植酸突变体是培育低植酸作物的主要途径。当把低植酸作为育种目标时,可能会忽略种子植酸含量的降低给植物带来的不利影响,如何消除低植酸造成的不利影响,成为科学家们亟需解决的问题。  相似文献   

17.
Phytate is the primary form of phosphorus found in mature cereal grain. This form of phosphorus is not available to monogastric animals due to a lack of the enzyme phytase in their digestive tract. Several barley low phytic acid (lpa) mutants have been identified that contain substantial decreases in seed phytate accompanied by concomitant increases in inorganic phosphorus. Seed homozygous for low phytic acid 1-1 (lpa1-1) or low phytic acid 2-1 (lpa2-1) has a 50% and 70% decrease in seed phytate respectively. These mutations were previously mapped to chromosomes 2HL and 7HL respectively. The RFLP marker ABC153 located in the same region of 2H was converted to a sequence-characterized-amplified-region (SCAR) marker. Segregation analysis of the CDC McGwire × Lp422 doubled haploid population confirmed linkage between the SCAR marker and the lpa1-1 locus with 15% recombination. A third low phytic acid mutant, M635, has a 75% decrease in phytate. This mutation was located to chromosome 1HL by linkage with an inter-simple sequence repeat (ISSR) based marker (LP75) identified through bulked-segregant analysis, and has been designated lpa3-1. Based on analysis of recombination between marker LP75 and low phytic acid in an additional mutant line M955 (95% phytate decrease), lpa3-1 and the mutation in M955 are in the same region on chromosome 1HL, and may be allelic.  相似文献   

18.
低植酸作物突变体研究进展   总被引:3,自引:0,他引:3  
王忠华 《植物学通报》2005,22(4):463-470
植酸是玉米(Zea mays)、小麦(Triticum aestivum)、大麦(Hordeum vulgare)、水稻(Oryza sativa)和大豆(Glvcine max)等籽粒中广泛存在的一种有机酸(6-肌醇磷酸),其与K 、Ca2 、Mg2 和Fe3 等金属离子形成的植酸盐是微量营养元素的重要贮存形式.植酸及植酸盐不能被人和非反刍动物所吸收利用;植酸摄入体内后还会和其他来源的微量营养元素结合形成植酸盐,造成这些营养元素的生物有效性下降,从而造成微量元素缺乏症.此外,大量的植酸及植酸盐随粪便排出,造成严重的环境污染,尤其是水体富营养化.由于土壤中缺乏分解微生物,即使畜禽粪便作有机肥还田仍不能被作物吸收利用.近年来,利用理化诱变与转基因技术已成功地获得了玉米、大麦、水稻和大豆等作物的低植酸突变体.本文对植酸的生物合成过程、低植酸突变体的诱发与研究、低植酸突变体的遗传特征与可能机理及营养评价进行了综述,并对低植酸作物的应用前景进行了简要分析.  相似文献   

19.
Development of effective polymerase chain reaction (PCR)-based diagnostic tests using ruminant fecal specimens has been thwarted by excessive inhibition. A PCR system based on amplification of 1000 copies of bacteriophage lambda-DNA was used as a model to evaluate inhibition levels in bovine feces. Dilution experiments using a bovine fecal specimen suggested that as little as 40 microg of feces (in a 100-microl PCR) affected the efficiency of amplification. It was discovered that phytic acid (the hexaphosphoric ester of inositol) is a powerful inhibitor of PCR. Above 0.3 mM phytate, the PCR is completely inhibited. In a very narrow range around 0.2 mM target-specific amplification proceeds efficiently. At concentrations between 10 and 100 microM, phytate nonspecific amplification (e.g., primer-dimer formation) is dominant. Below 10 microM, phytate target-specific amplification proceeds efficiently. A simple processing procedure using 50 units/ml of Aspergillus niger 3-phytase [E.C. 3.1.3.8] was developed that reduced PCR inhibition levels in bovine fecal specimens by approximately 500-fold.  相似文献   

20.
Evidence suggests that rats can tolerate a dietary phytate/Zn molar ratio greater than 15 if the dietary Zn concentration is high. High dietary Ca exacerbates the effect of phytic acid on Zn utilization by rats. In a short term (15 d) balance trial with adult men, we observed slightly greater Zn balance when whole compared to dephytinized wheat bran was consumed (molar ratios 12 and 1.2, respectively). There was, however, greater fecal excretion of Zn during the first 5 d whole bran was consumed. In a second study, Na phytate was the major source of phytic acid and Zn balance was less when the phytate/Zn molar ratio was greater than 16 compared to 4. The difference was not significant, however, and there was evidence of physiological adjustments to maintain homeostasis when the high ratio diet was consumed. Mean Zn intake averaged 17 mg (0.26 mmole) and 11 mg (0.17 mmole) daily for the bran and Na phytate studies, respectively. The level of Zn intake may influence the response of humans to varying phytate/Zn ratios. Comparison of isotope retention studies and the balance data is discussed. Some information on the relationship of dietary Ca to the phytate/Zn effect in human diets is gathered from current literature. The phytate/Zn molar ratio is a useful index of Zn bioavailability.  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