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 265 毫秒
1.
Phytate is a potent inhibitor of mineral absorption in humans occurring in plant-based food. Application of lactobacilli that produce phytate-degrading enzymes (phytases) to reduce phytate is an interesting yet a not much explored sector of research. Therefore, phytate dephosphorylation by Lactobacillus plantarum MTCC 1325 was evaluated. Cells at stationary phase showed phytase activity which was maximal at 24 h of growth. Glucose concentration and the type of phosphorous source in the media modulated the enzyme activity. Fermentation of cereal and/or legume flours with the strain resulted in phytate reduction with the highest in sorghum (73%) and the lowest in horse gram (34%). Further, the strain showed tolerance to acid, bile, and simulated gastrointestinal fluid. Significant phytase activity in the presence of simulated gastrointestinal fluids along with the ability to produce phytases post-exposure to simulated gastrointestinal fluids is of interest. To the best of our knowledge, this is the first report on the effect of simulated gastrointestinal fluid on cell-associated phytases of lactobacilli. The results of the investigation indicate that L. plantarum MTCC 1325 could be used as a starter in cereal-legume fermentation and as potential probiotics to achieve phytate hydrolysis in food matrices and also in gastrointestinal tract.  相似文献   

2.
The present study gives an overview on the whole mechanism of phytate degradation in the gut and the enzymes involved. Based on the similarity of the human and pigs gut, the study was carried out in pigs as model for humans. To differentiate between intrinsic feed phytases and endogenous phytases hydrolysing phytate in the gut, two diets, one high (control diet) and the other one very low in intrinsic feed phytases (phytase inactivated diet) were applied. In the chyme of stomach, small intestine and colon inositol phosphate isomers and activities of phytases and alkaline phosphatases were determined. In parallel total tract phytate degradation and apparent phosphorus digestibility were assessed. In the stomach chyme of pigs fed the control diet, comparable high phytase activity and strong phytate degradation were observed. The predominant phytate hydrolysis products were inositol phosphates, typically formed by plant phytases. For the phytase inactivated diet, comparable very low phytase activity and almost no phytate degradation in the stomach were determined. In the small intestine and colon, high activity of alkaline phosphatases and low activity of phytases were observed, irrespective of the diet fed. In the colon, stronger phytate degradation for the phytase inactivated diet than for the control diet was detected. Phytate degradation throughout the whole gut was nearly complete and very similar for both diets while the apparent availability of total phosphorus was significantly higher for the pigs fed the control diet than the phytase inactivated diet. The pathway of inositol phosphate hydrolysis in the gut has been elucidated.  相似文献   

3.
Examples of a new class of phytase related to protein tyrosine phosphatases (PTP) were recently isolated from several anaerobic bacteria from the rumen of cattle. In this study, the diversity of PTP-like phytase gene sequences in the rumen was surveyed by using the polymerase chain reaction (PCR). Two sets of degenerate primers were used to amplify sequences from rumen fluid total community DNA and genomic DNA from nine bacterial isolates. Four novel PTP-like phytase sequences were retrieved from rumen fluid, whereas all nine of the anaerobic bacterial isolates investigated in this work contained PTP-like phytase sequences. One isolate, Selenomonas lacticifex, contained two distinct PTP-like phytase sequences, suggesting that multiple phytate hydrolyzing enzymes are present in this bacterium. The degenerate primer and PCR conditions described here, as well as novel sequences obtained in this study, will provide a valuable resource for future studies on this new class of phytase. The observed diversity of microbial phytases in the rumen may account for the ability of ruminants to derive a significant proportion of their phosphorus requirements from phytate.  相似文献   

4.
Phosphorus (P) deficiency in soil is a major constraint for agricultural production worldwide. Despite this, most soils contain significant amounts of total soil P that occurs in inorganic and organic fractions and accumulates with phosphorus fertilization. A major component of soil organic phosphorus occurs as phytate. We show that when grown in agar under sterile conditions, Arabidopsis thaliana plants are able to obtain phosphorus from a range of organic phosphorus substrates that would be expected to occur in soil, but have only limited ability to obtain phosphorus directly from phytate. In wild-type plants, phytase constituted less than 0.8% of the total acid phosphomonoesterase activity of root extracts and was not detectable as an extracellular enzyme. By comparison, the growth and phosphorus nutrition of Arabidopsis plants supplied with phytate was improved significantly when the phytase gene (phyA) from Aspergillus niger was introduced. The Aspergillus phytase was only effective when secreted as an extracellular enzyme by inclusion of the signal peptide sequence from the carrot extensin (ex) gene. A 20-fold increase in total root phytase activity in transgenic lines expressing ex::phyA resulted in improved phosphorus nutrition, such that the growth and phosphorus content of the plants was equivalent to control plants supplied with inorganic phosphate. These results show that extracellular phytase activity of plant roots is a significant factor in the utilization of phosphorus from phytate and indicate that opportunity exists for using gene technology to improve the ability of plants to utilize accumulated forms of soil organic phosphorus.  相似文献   

5.
Forty crossbred barrows (Camborough 15 Line female x Canabred sire) weighing an average of 79.6 +/- 8.0 kg were used in a factorial design experiment (5 barleys x 2 enzyme levels) conducted to determine the effects of phytase supplementation on nutrient digestibility in low-phytate barleys fed to finishing pigs. The pigs were assigned to one of 10 dietary treatments comprised of a normal 2-rowed, hulled variety of barley (CDC Fleet, 0.26% phytate) or 2 low-phytate hulled genotypes designated as LP422 (0.14% phytate) and LP635 (0.09% phytate). A normal, hulless barley (CDC Dawn, 0.26% phytate) and a hulless genotype designated as LP422H (0.14% phytate) were also included. All barleys were fed with and without phytase (Natuphos 5000 FTU/kg). The diets fed contained 98% barley, 0.5% vitamin premix, 0.5% trace mineral premix, 0.5% NaCl and 0.5% chromic oxide but no supplemental phosphorus. The marked feed was provided for a 7-day acclimatization period, followed by a 3-day faecal collection. In the absence of phytase, phosphorus digestibility increased substantially (P < 0.05) as the level of phytate in the barley declined. For the hulled varieties, phosphorus digestibility increased from 12.9% for the normal barley (0.26% phytate) to 35.3 and 39.8% for the two low-phytate genotypes (0.14 and 0.09% phytate respectively). For the hulless varieties, phosphorus digestibility increased from 9.2% for the normal barley (0.26% phytate) to 34.7% for the hulless variety with 54% of the normal level of phytate (0.14% phytate). In contrast, when phytase was added to the diet, there was little difference in phosphorus digestibility between pigs fed normal barley and those fed the low-phytate genotypes (significant barley x enzyme interaction, P = 0.01). For the hulled varieties, phosphorus digestibility was 50.1% for the barley with the normal level of phytate (0.26% phytate) compared with 51.1 and 52.4% for the varieties with 54 and 35% of the normal level of phytate (0.14 and 0.09% phytate respectively). For the hulless varieties, phosphorus digestibility increased from 47.1% for the normal barley (0.26% phytate) to 54.4% for the hulless variety with 54% of the normal level of phytate (0.14% phytate). In conclusion, both supplementation with phytase and selection for low-phytate genotypes of barley were successful in increasing the digestibility of phosphorus for pigs. Unfortunately, the effects did not appear to be additive. Whether or not swine producers will choose low-phytate barley or supplementation with phytase as a means to improve phosphorus utilization, will likely depend on the yield potential of low-phytate barley and the additional costs associated with supplementation with phytase.  相似文献   

6.
Forty crossbred barrows (Camborough 15 Line female×Canabred sire) weighing an average of 79.6±8.0?kg were used in a factorial design experiment (5 barleys×2 enzyme levels) conducted to determine the effects of phytase supplementation on nutrient digestibility in low-phytate barleys fed to finishing pigs. The pigs were assigned to one of 10 dietary treatments comprised of a normal 2-rowed, hulled variety of barley (CDC Fleet, 0.26% phytate) or 2 low-phytate hulled genotypes designated as LP422 (0.14% phytate) and LP635 (0.09% phytate). A normal, hulless barley (CDC Dawn, 0.26% phytate) and a hulless genotype designated as LP422H (0.14% phytate) were also included. All barleys were fed with and without phytase (Natuphos 5000 FTU/kg). The diets fed contained 98% barley, 0.5% vitamin premix, 0.5% trace mineral premix, 0.5% NaCl and 0.5% chromic oxide but no supplemental phosphorus. The marked feed was provided for a 7-day acclimatization period, followed by a 3-day faecal collection. In the absence of phytase, phosphorus digestibility increased substantially (P<0.05) as the level of phytate in the barley declined. For the hulled varieties, phosphorus digestibility increased from 12.9% for the normal barley (0.26% phytate) to 35.3 and 39.8% for the two low-phytate genotypes (0.14 and 0.09% phytate respectively). For the hulless varieties, phosphorus digestibility increased from 9.2% for the normal barley (0.26% phytate) to 34.7% for the hulless variety with 54% of the normal level of phytate (0.14% phytate). In contrast, when phytase was added to the diet, there was little difference in phosphorus digestibility between pigs fed normal barley and those fed the low-phytate genotypes (significant barley×enzyme interaction, P=0.01). For the hulled varieties, phosphorus digestibility was 50.1% for the barley with the normal level of phytate (0.26% phytate) compared with 51.1 and 52.4% for the varieties with 54 and 35% of the normal level of phytate (0.14 and 0.09% phytate respectively). For the hulless varieties, phosphorus digestibility increased from 47.1% for the normal barley (0.26% phytate) to 54.4% for the hulless variety with 54% of the normal level of phytate (0.14% phytate). In conclusion, both supplementation with phytase and selection for low-phytate genotypes of barley were successful in increasing the digestibility of phosphorus for pigs. Unfortunately, the effects did not appear to be additive. Whether or not swine producers will choose low-phytate barley or supplementation with phytase as a means to improve phosphorus utilization, will likely depend on the yield potential of low-phytate barley and the additional costs associated with supplementation with phytase.  相似文献   

7.
Transgenic Nicotiana tabacum plants expressing a chimeric phytase gene (ex::phyA) from the soil fungus Aspergillus niger were generated. Three independently transformed lines showed increased extracellular phytase activity compared with a vector control and wild-type plants, both of which had no detectable extracellular phytase. Transgenic N. tabacum plants grown in sterile agar supplied with phosphorus (P) as phytate accumulated 3.7-fold more P than vector control plants. Despite this, the expression of ex::phyA in plants did not lead to an improved accumulation of P from two unamended P-deficient soils. However, when soils were amended with either phytate or phosphate and lime, transgenic plants accumulated up to 52% more P than controls. Positive responses by transgenic plants were, in some instances, coincident with a putative increase in soil phytate. We conclude that the development of plants that exude phytase to the soil may not ensure improved plant P nutrition, as the availability of phytate in the soil also appears to be critical. Nevertheless, if plants that express ex::phyA are combined with soil amendments that promote the availability of phytate, there is the potential to enhance the P nutrition of crop plants and to improve the efficiency of P fertilizer use in agricultural systems.  相似文献   

8.
Oh BC  Chang BS  Park KH  Ha NC  Kim HK  Oh BH  Oh TK 《Biochemistry》2001,40(32):9669-9676
The thermostable phytase from Bacillus amyloliquefaciens DS11 hydrolyzes phytate (myo-inositol hexakisphosphate, IP6) to less phosphorylated myo-inositol phosphates in the presence of Ca2+. In this report, we discuss the unique Ca2+-dependent catalytic properties of the phytase and its specific substrate requirement. Initial rate kinetic studies of the phytase indicate that the enzyme activity follows a rapid equilibrium ordered mechanism in which binding of Ca2+ to the active site is necessary for the essential activation of the enzyme. Ca2+ turned out to be also required for the substrate because the phytase is only able to hydrolyze the calcium-phytate complex. In fact, both an excess amount of free Ca2+ and an excess of free phytate, which is not complexed with each other, can act as competitive inhibitors. The Ca2+-dependent catalytic activity of the enzyme was further confirmed, and the critical amino acid residues for the binding of Ca2+ and substrate were identified by site-specific mutagenesis studies. Isothermal titration calorimetry (ITC) was used to understand if the decreased enzymatic activity was related to poor Ca2+ binding. The pH dependence of the Vmax and Vmax/Km consistently supported these observations by demonstrating that the enzyme activity is dependent on the ionization of amino acid residues that are important for the binding of Ca2+ and the substrate. The Ca2+-dependent activation of enzyme and substrate was found to be different from other histidine acid phytases that hydrolyze metal-free phytate.  相似文献   

9.
A novel class of cysteine phytase showing ability to degrade phytate has recently been isolated from rumen bacteria. To expand our knowledge of this enzyme class, a total of 101 distinct cysteine phytase gene fragments were identified from the ruminal genomic DNA of Bore goats and Holstein cows, and most of them shared low identities (< 50%) with known sequences. By phylogenetic analysis, these sequences were separated into three clusters that showed substantial diversity. The two most abundant cysteine phytase genes of goat rumens were cloned and their protein products were characterized. Four findings were revealed based on our results. (i) Compared with soil and water environment, where β‐propeller phytase is the most important phytate‐degrading enzyme, cysteine phytase is the major phytate‐degrading enzyme in the anaerobic ruminal environment. (ii) Cysteine phytase fragments in the rumen contents of goat and cow have the same diversity profile, although most of the sequences and their abundance differ in the two species. (iii) Each species has their respective high‐abundance genes, which may play major roles for phytate degradation. (iv) Compared with previously reported cysteine phytases that have pH optimum at 4.5, the pH optima of the two most abundant secreted goat cysteine phytases are 6.5 and 6.0, which are within the pH range found in the rumens. This study provides valuable information about the diversity, abundance and enzymatic properties of the ruminal cysteine phytases and emphasizes the important role(s) of these cysteine phytases probably in the terrestrial cycle of phosphorus.  相似文献   

10.
AIMS: To screen micro-organisms for the ability to produce phytase enzyme(s) and to use promising strains for the fermentation of pea flour. METHODS AND RESULTS: Two methods using the indirect estimation of phytate degradation were evaluated and both shown to be inadequate. A third method, measuring the inositol phosphate (IP3-IP6) content directly during fermentation, was used instead of the indirect estimations of phytate degradation. In synthetic media, some strains required customized conditions, with no accessible phosphorus sources other than phytate, to express phytase activity. The repression of phytase synthesis by inorganic phosphorus was not detected during fermentation with pea flour as substrate and seemed to be less significant with a higher composition complexity of the substrate. None of the tested lactic acid bacteria strains showed phytase activity. CONCLUSIONS: The methodology for the phytase screening procedure was shown to be critical. Some of the screening methods and media used in previous publications were found to be inadequate. SIGNIFICANCE AND IMPACT OF THE STUDY: This paper highlights the pitfalls and difficulties in the evaluation of phytase production by micro-organisms. The study is of great importance for future studies in this area.  相似文献   

11.
Saccharomyces cerevisiae CY phytase-producing cells were immobilized in calcium alginate beads and used for the degradation of phylate. The maximum activity and immobilization yield of the immobilized phytase reached 280 mU/g-bead and 43%, respectively. The optimal pH of the immobilized cell phytase was not different from that of the free cells. However, the optimum temperature for the immobilized phytase was 50°C, which was 10°C higher than that of the free cells; pH and thermal stability were enhanced as a consequence of immobilization. Using the immobilized phytase, phytate was degraded in a stirred tank bioreactor. Phytate degradation, both in a buffer solution and in soybean-curd whey mixture, showed very similar trends. At an enzyme dosage of 93.9 mU/g-phytate, half of the phytate was degraded after 1 h of hydrolysis. The operational stability of the immobilized beads was examined with repeated batchwise operations. Based on 50% conversion of the phytate and five times of reuse of the immobilized beads, the specific degradation (g phytate/g dry cell weight) for the immobilized phytase increased 170% compared to that of the free phytase.  相似文献   

12.
Culture variables affecting phytase production by a thermophilic mould Sporotrichum thermophile in submerged fermentation were optimized. Soluble starch, peptone, Tween-80 and sodium phytate were identified by Plackett-Burman design as the most significant factors to affect phytase production. The 2(4) full factorial central composite design of response surface methodology was applied for optimizing the concentrations of the significant variables and to delineate their interactions. Starch, Tween-80, peptone and sodium phytate at 0.4%, 1.0%, 0.3% and 0.3% supported maximum enzyme titres, respectively. An overall 3.73-fold improvement in phytase production was achieved due to optimization. When sodium phytate was substituted with wheat bran (3%), the phytase titre in the former was comparable with that in the latter.  相似文献   

13.
Phytases are a special class of phosphatase that catalyze the sequential hydrolysis of phytate to less-phosphorylated myo-inositol derivatives and inorganic phosphate. Phytases are added to animal feedstuff to reduce phosphate pollution in the environment, since monogastric animals such as pigs, poultry, and fish are unable to metabolize phytate. Based on biochemical properties and amino acid sequence alignment, phytases can be categorized into two major classes, the histidine acid phytases and the alkaline phytases. The histidine acid phosphatase class shows broad substrate specificity and hydrolyzes metal-free phytate at the acidic pH range and produces myo-inositol monophosphate as the final product. In contrast, the alkaline phytase class exhibits strict substrate specificity for the calcium–phytate complex and produces myo-inositol trisphosphate as the final product. This review describes recent findings that present novel viewpoints concerning the molecular basis of phytase classification.  相似文献   

14.
PhyA gene products of Aspergillus ficuum (AF) and Peniophora lycii (PL) as expressed in industrial strains of Aspergillus niger and Aspergillus oryzae, respectively, were purified to homogeneity and then characterized for both physical and biochemical properties. The PL phytase is 26 amino acid residues shorter than the AF phytase. Dynamic light scattering studies indicate that the active AF phytase is a monomer while the PL phytase is a dimer. While both of the phytases retained four identical glycosylatable Asn residues, unique glycosylation sites, six for PL and seven for AF phytase, were observed. Global alignment of both the phytases has shown 38% sequence homology between the two proteins. At 58 degrees C and pH 5.0, the PL phytase gave a specific activity of 22,000 nKat/mg as opposed to about 3000 nKat/mg for AF phytase. However, the AF phytase is more thermostable than its counterpart PL phytase at 65 degrees C. Also, AF phytase is more stable at pH 7.5 than the PL phytase. The two phytases differed in K(m) for phytate, K(i) for myo-inositol hexasulfate (MIHS), and pH optima profile. Despite similarities in the active site sequences, the two phytases show remarkable differences in turnover number, pH optima profile, stability at higher temperature, and alkaline pH. These biochemical differences indicate that phytases from ascomycete and basidiomycete fungi may have evolved to degrade phytate in different environments.  相似文献   

15.
The objective of the study was to evaluate the effect of screen size (1, 2 and 3 mm) and microbial phytase (0 and 1000 FTU/kg as-fed) on phytate degradation in maize (100% maize), soybean meal (100% SBM) and maize–SBM (75% maize and 25% SBM) incubated in water for 0, 2, 4, 8 and 24 h at 38°C. Samples were analysed for pH, dry matter and phytate phosphorus (P). Particle size distribution (PSD) and average particle size (APS) of samples were measured by the Laser Diffraction and Bygholm method. PSD differed between the two methods, whereas APS was similar. Decreasing screen size from 3 to 1 mm reduced APS by 48% in maize, 30% in SBM and 26% in maize–SBM. No interaction between screen size and microbial phytase on phytate degradation was observed, but the interaction between microbial phytase and incubation time was significant (P<0.001). This was because microbial phytase reduced phytate P by 88% in maize, 84% in maize–SBM and 75% in SBM after 2 h of incubation (P<0.05), whereas the reduction of phytate P was limited (<50%) in the feeds, even after 24 h when no microbial phytase was added. The exponential decay model was fitted to the feeds with microbial phytase to analyse the effect of screen size and feed on microbial phytase efficacy on phytate degradation. The interaction between screen size and feed affected the relative phytate degradation rate (Rd) of microbial phytase as well as the time to decrease 50% of the phytate P (t1/2) (P<0.001). Thus, changing from 3 to 1 mm screen size increased Rd by 22 and 10%/h and shortened t1/2 by 0.4 and 0.2 h in maize and maize–SBM, respectively (P<0.05), but not in SBM. Moreover, the screen size effect was more pronounced in maize and maize–SBM compared with SBM as a higher phytate degradation rate constant (Kd) and Rd, and a shorter t1/2 was observed in maize compared with SBM in all screen sizes (P<0.05). However, a higher amount of degraded phytate was achieved in SBM than in maize because of the higher initial phytate P content in SBM. In conclusion, reducing screen size from 3 to 1 mm increased Kd and Rd and decreased t1/2 in maize and maize–SBM with microbial phytase. The positive effect of grinding on improving microbial phytase efficacy, which was expressed as Kd, Rd and t1/2, was greater in maize than in SBM.  相似文献   

16.
Experimental data on phytate phosphorus utilisation by ruminants are scarce. The aim of this study was to estimate the phytase activity of rumen micro-organisms when phytate phosphorus supply is high. A semi-continuous culture system fermentor (RUSITEC) was used. The inoculum was obtained from eight goats fed on either high or low forage level diets. Experimental buffers only differed by the nature of phosphorus monosodium phosphate vs. corn sodium phytate. The nylon bags containing 15 g DM of substrate were removed after a 48-hour incubation period. The system was maintained for 15 days: 5 days for adaptation, in order to obtain a steady state, and 10 days for sampling and recording. No significant differences were observed for DM digestibility, gas production, pH, N-NH3, and SCFA for the different treatments. Bacterial efficiency of phytate phosphorus utilisation was significantly higher (p < 0.001) with organic P, but remained lower than the data usually reported in the literature. These results may be explained by the relative saturation of bacterial phytase activity when the buffer contains a high level of phytate phosphorus.  相似文献   

17.
High iron consumption has been proposed to relate to an increase in the risk of colon cancer, whereas high levels of supplemental sodium phytate effectively reduce iron-induced oxidative injury and reverse iron-dependent augmentation of colorectal tumorigenesis. However, the protective role of intrinsic dietary phytate has not been determined. In this study, we examined the impact of removing phytate present in a corn-soy diet by supplemental microbial phytase on susceptibility of pigs to the oxidative stress caused by a moderately high dietary iron intake. Thirty-two weanling pigs were fed the corn-soy diets containing two levels of iron (as ferrous sulfate, 80 or 750 mg/kg diet) and microbial phytase (as Natuphos, BASF, Mt. Olive, NJ, 0 or 1200 units/kg). Pigs fed the phytase-supplemented diets did not receive any inorganic phosphorus to ensure adequate degradation of phytate. After 4 months of feeding, liver, colon, and colon mucosal scrapings were collected from four pigs in each of the four dietary groups. Colonic lipid peroxidation, measured as thiobarbituric acid reacting substances (TBARS), was increased by both the high iron (P< 0.0008) and phytase (P< 0.04) supplementation. Both TBARS and F2-isoprostanes, an in vivo marker of lipid peroxidation, in colonic mucosa were affected by dietary levels of iron (P< 0.03). Mean hepatic TBARS in pigs fed the phytase-supplemented, high iron diet was 43%-65% higher than that of other groups although the differences were nonsignificant. Moderately high dietary iron induced hepatic glutathione peroxidase activity (P= 0.06) and protein expression, but decreased catalase (P< 0.05) in the colonic mucosa. In conclusion, intrinsic phytate in corn and soy was protective against lipid peroxidation in the colon associated with a moderately high level of dietary iron.  相似文献   

18.
In this study 97 soil samples from different soil ecosystems were collected. The initial screening was performed on modified glycerol arginine agar (MGAA) to isolate common actinomycetes and on modified MGA-SE (MMGA-SE) to isolate rare actinomycetes. Sixty-seven isolates potentially producing extracellular phytate-degrading activity were identified. The potential to dephosphorylate phytate was confirmed in liquid culture for 46.3 % of the isolates. 12 strains were selected for a direct determination of their phytate-degrading capacity. The results highlighted that the selected isolates produced extracellular phytate-degrading activity; however their capacity in InsP(6) degradation was different. In addition the fermentation medium had an effect on the extent of phytate degradation. Some enzymatic properties of the phytases from isolate No. 43 and isolate No. 63 were determined after obtaining phytase-enriched samples. The enzymes had maximum phytate-degrading capability at 55 °C and pH 5 (isolate No. 43) and 37 °C and pH 7 (isolates No. 63), respectively. Due to their properties, the phytase of isolate No. 43 behaves like a histidine acid phytase, whereas the phytase of No. 63 showed similar enzymatic properties to the phytase of lily. To our knowledge, the results from this study demonstrated for the first time that actinomycetes produce extracellular phytate-degrading activity. By 16SrRNA sequencing, the more closely studied phytase producers were identified as Streptomyces sp. Isolate No. 43 showed 98 % identity to Streptomyces alboniger and S. venezuelae, while isolate No. 63 exhibited 98 % sequence identity to S. ambofaciens and S. lienomycini.  相似文献   

19.
Two thermostable phytases were identified from Thai isolates of Aspergillus japonicus BCC18313 (TR86) and Aspergillus niger BCC18081 (TR170). Both genes of 1404 bp length, coding for putative phytases of 468 amino acid residues, were cloned and transferred into Pichia pastoris . The recombinant phytases, r-PhyA86 and r-PhyA170, were expressed as active extracellular, glycosylated proteins with activities of 140 and 100 U mL−1, respectively. Both recombinant phytases exhibited high affinity for phytate but not for p -nitrophenyl phosphate. Optimal phytase activity was observed at 50 °C and pH 5.5. High thermostability, which is partly dependent on glycosylation, was demonstrated for both enzymes, as >50% activity was retained after heating at 100 °C for 10 min. The recombinant phytases also exhibited broad pH stability from 2.0 to 8.0 and are resistant to pepsin. In vitro digestibility tests suggested that r-PhyA86 and r-PhyA170 are at least as efficient as commercial phytase for hydrolyzing phytate in corn-based animal feed and are therefore suitable sources of phytase supplement.  相似文献   

20.
Transgenic maize plants expressing a fungal phytase gene   总被引:12,自引:0,他引:12  
Maize seeds are the major ingredient of commercial pig and poultry feed. Phosphorus in maize seeds exists predominantly in the form of phytate. Phytate phosphorus is not available to monogastric animals and phosphate supplementation is required for optimal animal growth. Undigested phytate in animal manure is considered a major source of phosphorus pollution to the environment from agricultural production. Microbial phytase produced by fermentation as a feed additive is widely used to manage the nutritional and environmental problems caused by phytate, but the approach is associated with production costs for the enzyme and requirement of special cares in feed processing and diet formulation. An alternative approach would be to produce plant seeds that contain high phytase activities. We have over-expressed Aspergillus niger phyA2 gene in maize seeds using a construct driven by the maize embryo-specific globulin-1 promoter. Low-copy-number transgenic lines with simple integration patterns were identified. Western-blot analysis showed that the maize-expressed phytase protein was smaller than that expressed in yeast, apparently due to different glycosylation. Phytase activity in transgenic maize seeds reached approximately 2,200 units per kg seed, about a 50-fold increase compared to non-transgenic maize seeds. The phytase expression was stable across four generations. The transgenic seeds germinated normally. Our results show that the phytase expression lines can be used for development of new maize hybrids to improve phosphorus availability and reduce the impact of animal production on the environment.  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