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 140 毫秒
1.
2.
Resistance to the non‐selective herbicide dl ‐phosphinothricin (PPT) was introduced into commercial Lotus corniculatus cv. Bokor by co‐cultivation of cotyledons with Agrobacterium tumefaciens AGL1 harbouring the binary vector pDM805 which contains the bialaphos resistance gene (bar) from Streptomyces hygroscopicus encoding phosphinothricin acetyltransferase (PAT) and the uidA gene encoding β‐glucuronidase. The half‐cotyledon explants were precultured on regeneration Murashige and Skoog's (MS) medium supplemented with 6‐benzyladenine (BA) and 1‐naphthaleneacetic acid (NAA) at 0.5 mg L?1 each, 3 days prior to infection. Upon co‐cultivation, the explants were cultured on PPT‐free regeneration medium for 10 days, and then subcultured on regeneration/selection media with increasing PPT concentrations (5–7 mg L?1) for about 18 weeks. Out of 480 initially co‐cultivated explants, 272 regenerated shoots survived the entire PPT selection procedure. Resistant shoots were grown further, multiplied by tillering that was additionally promoted by PPT and rooted on hormone‐free MS medium containing 5 mg L?1 PPT. Established shoot cultures, continuously maintained on the same medium, have preserved PPT resistance up to now (more than 2 years). Transformed plants assessed in vitro and in a greenhouse were tolerant to the herbicide PPT at 300 mg L?1 equivalent to more than twofold the recommended field dosage for weed eradication. Applied PPT treatment did not affect the activities of glutamine synthetase (GS; EC 6.3.1.2) and NADH‐dependent glutamate dehydrogenase (NADH‐GDH; EC 1.4.1.2) in transformed plants. However, PPT did increase the mobility of glutamine synthetase isoforms GS1 and GS2 as well as the inhibition of an additional high mobility GS (hmGS) activity. In untransformed plants, PPT treatment reduced total GS activity by 4.4‐fold while contrary the activity of NADH‐GDH was increased by ninefold. All transformed herbicide‐resistant plants were phenotypically normal and exhibited genomic stability, as were the untransformed plants analysed by flow cytometry. Under greenhouse conditions, they grew to maturity, flowered and set seeds. Stable integration and expression of the bar gene in T0 and T1 plants were confirmed by Southern and Western blot analysis, while integration of the reporter uidA gene did not occur. The bar gene was inherited in a Mendelian fashion by the progeny, as detected by PPT resistance. The production of PPT‐resistant plants may have significant practical applications in weed control in fields of L. corniculatus.  相似文献   

3.
D‐alanine is widely used in medicine, food, additives, cosmetics, and other consumer items. Esterase derived from Bacillus cereus WZZ001 exhibits high hydrolytic activity and stereoselectivity. In this study, we expressed the esterase gene in Escherichia coli BL21 (DE3). We analyzed the biocatalytic resolution of N‐acetyl‐DL‐alanine methyl ester by immobilized whole Ecoli BL21 (DE3) cells, which were prepared through embedding and cross‐linking. We analyzed biocatalytic resolution under the optimal conditions of pH of 7.0, temperature of 40°C and substrate concentration of at 700 mM with an enantiomeric excess of 99.99% and e.e.p of 99.50%. The immobilized recombinant Bcereus esterase Ecoli BL21 (DE3) cells exhibited excellent reusability and retained 86.04% of their initial activity after 15 cycles of repeated reactions. The immobilized cells are efficient and stable biocatalysts for the preparation of N‐acetyl‐D‐alanine methyl esters.  相似文献   

4.
The continuously growing interest in the understanding of peptide folding led to the conformational investigation of methylamides of N‐acetyl‐amino acids as diamide models. Here we report the results of detailed conformational analysis on Ac‐Pro‐NHMe and Ac‐β‐HPro‐NHMe diamides. These compounds were analyzed by experimental and computational methods, the conformational distributions obtained by Density Functional Theory (DFT) calculations for isolated and solvated diamide compounds are discussed. The conformational preference of proline‐containing diamide compounds as a function of the ambience was observed by a number of chiroptical spectroscopic techniques, such as vibrational circular dichroism (VCD), electronic circular dichroism (ECD), Raman optical activity (ROA) spectroscopy, and additionally by single crystal X‐ray diffraction analyses. Based on a comparison between Ac‐Pro‐NHMe and Ac‐β‐HPro‐NHMe, one can conclude that due to the greater conformational freedom of the β‐HPro derivative, Ac‐β‐HPro‐NHMe shows different behavior in solid‐ and solution‐phase, as well. Ac‐β‐HPro‐NHMe tends to form cis Ac‐β‐HPro amide conformation in water, dichloromethane, and acetonitrile in contrast to its α‐Pro analog. On the other hand, the crystal structure of the β‐HPro compound cannot be related to any of the conformers obtained in vacuum and solution while the X‐ray structure of Ac‐Pro‐NHMe was identified as tαL–, which is a trans Ac‐Pro amide containing conformer also predominant in polar solvents. Chirality 26:228–242, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

5.
The B subunit of Escherichia coli heat-labile enterotoxin (LTB) has been transformed to plants for use as an edible vaccine. We have developed a simple and reliable Agrobacterium-mediated transformation method to express synthetic LTB gene in N. tabacum using a phosphinothricin acetyltransferase (bar) gene as a selectable marker. The synthetic LTB gene adapted to the coding sequence of tobacco plants was cloned to a plant expression vector under the control of the ubiquitin promoter and transformed to tobacco by Agrobacterium-mediated transformation. Transgenic plants were selected in the medium supplemented with 5 mg l-1 phosphinothricin (PPT). The amount of LTB protein detected in the transgenic tobacco was approximately 3.3% of the total soluble protein, approximately 300-fold higher than in the plants generated using the native LTB gene under the control of the CaMV 35S promoter. The transgenic plants that were transferred to a greenhouse had harvested seeds that proved to be resistant to herbicide. Thus, the described protocol could provide a useful tool for the transformation of tobacco plants.  相似文献   

6.
Alanine oligomers provide a key structure for silk fibers from spider and wild silkworms.We report on structural analysis of l ‐alanyl‐l ‐alanyl‐l ‐alanyl‐l ‐alanine (Ala)4 with anti‐parallel (AP) β‐structures using X‐ray and solid‐state NMR. All of the Ala residues in the (Ala)4 are in equivalent positions, whereas for alanine trimer (Ala)3 there are two alternative locations in a unit cell as reported previously (Fawcett and Camerman, Acta Cryst., 1975, 31, 658–665). (Ala)4 with AP β‐structure is more stable than AP‐(Ala)3 due to formation of the stronger hydrogen bonds. The intermolecular structure of (Ala)4 is also different from polyalanine fiber structure, indicating that the interchain arrangement of AP β‐structure changes with increasing alanine sequencelength. Furthermore the precise 1H positions, which are usually inaccesible by X‐ray diffraction method, are determined by high resolution 1H solid state NMR combined with the chemical shift calculations by the gauge‐including projector augmented wave method. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 13–20, 2014.  相似文献   

7.
A peptide, N‐Ac‐Phe‐Tyr‐NH2, with angiotensin I‐converting enzyme (ACE) inhibitor activity was synthesized by an α‐chymotrypsin‐catalyzed condensation reaction of N‐acetyl phenylalanine ethyl ester (N‐Ac‐Phe‐OEt) and tyrosinamide (Tyr‐NH2). Three kinds of solvents: a Tris–HCl buffer (80 mM, pH 9.0), dimethylsulfoxide (DMSO), and acetonitrile were employed in this study. The optimum reaction solvent component was determined by simplex centroid mixture design. The synthesis efficiency was enhanced in an organic‐aqueous solvent (Tris‐HCl buffer: DMSO: acetonitrile = 2:1:1) in which 73.55% of the yield of N‐Ac‐Phe‐Tyr‐NH2 could be achieved. Furthermore, the effect of reaction parameters on the yield was evaluated by response surface methodology (RSM) using a central composite rotatable design (CCRD). Based on a ridge max analysis, the optimum condition for this peptide synthesis included a reaction time of 7.4 min, a reaction temperature of 28.1°C, an enzyme activity of 98.9 U, and a substrate molar ratio (Phe:Tyr) of 1:2.8. The predicted and the actual (experimental) yields were 87.6 and 85.5%, respectively. The experimental design and RSM performed well in the optimization of synthesis of N‐Ac‐Phe‐Tyr‐NH2, so it is expected to be an effective method for obtaining a good yield of enzymatic peptide. © 2012 American Institute of Chemical Engineers Biotechnol. Prog., 2012  相似文献   

8.
We present evidence, for the first time, of the occurrence of a transport system common for amino acid methionine, and methionine/glutamate analogues l-methionine-dl-sulfoximine (MSX) and phosphinothricin (PPT) in cyanobacterium Nostoc muscorum. Methionine, which is toxic to cyanobacterium, enhanced its nitrogenase activity at lower concentrations. The cyanobacterium showed a biphasic pattern of methionine uptake activity that was competitively inhibited by the amino acids alanine, isoleucine, leucine, phenylalanine, proline, valine, glutamine, and asparagine. The methionine/glutamate analogue-resistant N. muscorum strains (MSX-R and PPT-R strains) also showed methionine-resistant phenotype accompanied by a drastic decrease in 35S methionine uptake activity. Treatment of protein extracts from these mutant strains with MSX and PPT reduced biosynthetic glutamine synthetase (GS) activity only in vitro and not in vivo. This finding implicated that MSX- and PPT-R phenotypes may have arisen due to a defect in their MSX and PPT transport activity. The simultaneous decrease in methionine uptake activity and in vitro sensitivity toward MSX and PPT of GS protein in MSX- and PPT-R strains indicated that methionine, MSX, and PPT have a common transport system that is shared by other amino acids as well in N. muscorum. Such information can become useful for isolation of methionine-producing cyanobacterial strains.  相似文献   

9.
The microprojectile bombardment method was used to transfer DNA into embryogenic callus of asparagus (Asparagus officcinalis L.) and to produce stably transformed asparagus plants. Embryogenic callus, derived from UC 157 and UC72 asparagus cultivars, was bombarded with tungsten particles coated with plasmid DNA that contained genes encoding hygromycin phosphotransferase, phosphinothricin acetyl transferase and -glucuronidase. Putatively transformed calli were identified from the bombarded tissue after 4 months selection on 25 mg/L hygromycin B plus 4 mg/L phosphinothricin (PPT). By selecting embryogenic callus on hygromycin plus PPT the overall transformation and selection efficiencies were substantially improved over selection with hygromycin or PPT alone, where no transgenic clones were recovered. The transgenic nature of the selected material was demonstrated by GUS histochemical assays and Southern blot hybridization analysis. Transgenic asparagus plants were found to withstand the prescribed levels of the PPT-based herbicide BASTATM for weed control.Abbreviations GUS -glucuronidase - HPT hygromycin phosphotransferase - bar phosphinothricin acetyl transferase gene - PPT phosphophinothricin - NAA naphthalene acetic acid - 2iP 2-isopentenyl adenine  相似文献   

10.
A system for the production of transgenic papaya (Carica papaya L.) plants using zygotic embryos and embryogenic callus as target cells for particle bombardment is described. Phosphinothricin (bar ) and kanamycin (npt II) resistance genes were used as selectable markers, and the gus gene (uidA) as a reporter gene. Selection with 100 mg/l kanamycin and 4 mg/l phosphinothricin (PPT) yielded a total of over 90 resistant embryogenic colonies from three independent experiments using embryogenic callus as a target tissue. This represents an efficiency of 60 transgenic clones per gram of fresh weight callus bombarded. The efficiency of genetic transformation using zygotic embryos was lower, as only 8 independent resistant clones were recovered out of 645 bombarded zygotic embryos, giving a efficiency of 1.24%. Subsequent subculture of transgenic somatic embryos both from zygotic embryos and embryogenic callus led to the development of plants with apparently normal morphology. Histological, fluorimetric assay for GUS, NPT II assay and DNA analysis (Southern hybridization) showed that kanamycin /PPT resistant plants carried and expressed the transgenes.Abbreviations Gus -glucuronidase - NPTII neomycin phophotransferase II - bar phophinothricin acetyl transferase gene - Pat phosphinothricin acetyl transferase - PPT phosphinothricin - Km kanamycin - 2,4-D 2,4-dichlorophenoxyacetic acid - K kinetin - BAP benzylaminopurine - IBA indolbutyric acid  相似文献   

11.
Dehydrobutyrine is the most naturally occurring dehydroamino acid. It is also the simplest dehydroamino acid having the geometrical isomers E/Z. To investigate its conformational properties, a theoretical analysis was performed on N‐acetyl‐α,β‐dehydrobutyrine N′‐methylamides, Ac‐(E)‐ΔAbu‐NHMe and Ac‐(Z)‐ΔAbu‐NHMe, as well as the dehydrovaline derivative Ac‐ΔVal‐NHMe. The ?, ψ potential energy surfaces and the localised conformers were calculated at the B3LYP/6‐311 + + G(d,p) level of theory both in vacuo and with inclusion of the solvent (chloroform, water) effect (SCRF method). The X‐ray crystal structures of Ac‐(Z)‐ΔAbu‐NHMe and Ac‐ΔVal‐NHMe were determined at 85 and 100 K, respectively. The solid‐state conformational preferences for the studied residues have been analysed and compared with the other related structures. Despite the limitations imposed by the Cα = Cβ double bond on the topography of the side chains, the main chains of the studied dehydroamino acids are more flexible than in standard alanine. The studied dehydroamino acids differ in their conformational preferences, which depend on the polarity of the environment. This might be a reason why the nature quite precisely differentiates between ΔVal and each of the ΔAbu isomers, and why, particularly so with the latter, they are used as a conformational tool to influence the biological action of usually small, cyclic dehydropeptides. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Cleavage reactions at backbone loci are one of the consequences of oxidation of proteins and peptides. During α‐amidation, the Cα–N bond in the backbone is cleaved under formation of an N‐terminal peptide amide and a C‐terminal keto acyl peptide. On the basis of earlier works, a facilitation of α‐amidation by the thioether group of adjacent methionine side chains was proposed. This reaction was characterized by using benzoyl methionine and benzoyl alanyl methionine as peptide models. The decomposition of benzoylated amino acids (benzoyl‐methionine, benzoyl‐alanine, and benzoyl‐methionine sulfoxide) to benzamide in the presence of different carbohydrate compounds (reducing sugars, Amadori products, and reductones) was studied during incubation for up to 48 h at 80 °C in acetate‐buffered solution (pH 6.0). Small amounts of benzamide (0.3–1.5 mol%) were formed in the presence of all sugars and from all benzoylated species. However, benzamide formation was strongly enhanced, when benzoyl methionine was incubated in the presence of reductones and Amadori compounds (3.5–4.2 mol%). The reaction was found to be intramolecular, because α‐amidation of a similar 4‐methylbenzoylated amino acid was not enhanced in the presence of benzoyl‐methionine and carbohydrate compounds. In the peptide benzoyl‐alanyl‐methionine, α‐amidation at the methionine residue is preferred over α‐amidation at the benzoyl peptide bond. We propose here a mechanism for the enhancement of α‐amidation at methionine residues. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Disaccharide containing unit of peptidoglycan from bacterial cell wall, N‐acetyl‐d ‐glucosaminyl‐N‐acetylmuramyl‐l ‐alanyl‐d ‐glutaminamide (gluсosaminyl‐muramyl‐dipeptide) registered in Russia as an immunomodulatory drug, is shown to participate in slow equilibrium of α and β anomeric forms. Data of NMR spectra and molecular dynamics indicate that the α‐anomer predominantly acquires a folded conformation stabilized by intramolecular hydrogen bond between the alanyl carbonyl and muramyl NH proton. The β‐form displays a considerable fraction of extended, non‐hydrogen bonded structures. In the standard immunoadjuvant test system, the α‐form is practically inactive, and the activity of the equilibrium mixture with α : β = 68 : 32 ratio is due to the presence of β‐anomer. Such unique α–β selectivity of biological action must be considered at the design of related immunoactive glycopeptides. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
The folding and assembly of Rubisco large and small subunits into L8S8 holoenzyme in chloroplasts involves many auxiliary factors, including the chaperone BSD2. Here we identify apparent intermediary Rubisco‐BSD2 assembly complexes in the model C3 plant tobacco. We show BSD2 and Rubisco content decrease in tandem with leaf age with approximately half of the BSD2 in young leaves (~70 nmol BSD2 protomer.m2) stably integrated in putative intermediary Rubisco complexes that account for <0.2% of the L8S8 pool. RNAisilencing BSD2 production in transplastomic tobacco producing bacterial L2 Rubisco had no effect on leaf photosynthesis, cell ultrastructure, or plant growth. Genetic crossing the same RNAi‐bsd2 alleles into wild‐type tobacco however impaired L8S8 Rubisco production and plant growth, indicating the only critical function of BSD2 is in Rubisco biogenesis. Agrobacterium mediated transient expression of tobacco, Arabidopsis, or maize BSD2 reinstated Rubisco biogenesis in BSD2‐silenced tobacco. Overexpressing BSD2 in tobacco chloroplasts however did not alter Rubisco content, activation status, leaf photosynthesis rate, or plant growth in the field or in the glasshouse at 20°C or 35°C. Our findings indicate BSD2 functions exclusively in Rubisco biogenesis, can efficiently facilitate heterologous plant Rubisco assembly, and is produced in amounts nonlimiting to tobacco growth.  相似文献   

15.
Summary Selectable marker genes play an important role in plant transformation. The level of selection pressure is generally established by generating a kill curve for the selectable marker. In most cases, the lowest concentration which kills all explants is used. This study examined two selectable marker genes, phosphinothricin acetyl transferase (PAT) and hygromycin phosphotransferase (HPT), in transformation of tobacco leaf disks. Experiments to determine the lethal level of the herbicide, glufosinate-ammonium (phosphinothricin) (PPT) using a leaf-disk regeneration assay established that no shoots regenerated at 2 to 4 mg PPT per 1. Likewise with the antibiotic, hygromycin (HYG), no plants regenerated at 50 mg hygromycin per 1. In contrast, after cocultivation of the leaf disks withAgrobacterium tumefaciens containing either the PAT or HPT gene in combination with a Bt gene for insect resistance, plants were successfully regenerated from leaf disks at 2 to 4 mg PPT per 1 and 50 mg hygromycin per 1. However, most plants regenerated at 2 and 3 mg PPT per 1 were found to be nontransformed (95–100% escapes) by i) Southern-blot analysis, ii) herbicide application test, and iii) insect feeding bioassay. On the other hand, plants that regenerated on 50 mg hygromycin per 1 and 4 mg PPT per 1 were transgenic as determined by Southern analysis, leaf assay for PPT or HYG resistance, and death of tobacco budworms feeding on these leaves. This study showed a significant level of cross-protection and/or transient expression of the PAT selectable marker gene allowing escapes (95–100%) at selection levels of 2 and 3 mg PPT per 1 which completely kill controls. On the other hand, the HPT gene at 50 mg is efficient in selecting for T-DNA integration.  相似文献   

16.
Efficient Agrobacterium tumefaciens-mediated transformation and a higher recovery of transformed plants of cucumber cv. Poinsett76 were achieved via direct organogenesis from cotyledon explants. Stable transformants were obtained by inoculating explants with A. tumefaciens strains EHA105 or LBA4404, both harboring the binary vector pME508, which contains the neomycin phosphotransferase II (nptII) and phosphinothricin resistance genes (bar) conferring resistance to kanamycin and PPT, respectively, as selectable markers and the sgfp-tyg gene for the green fluorescent protein (GFP) as a visual marker driven by the constitutive CaMV35S promoter in the presence of acetosyringone (50 μM). Transformed shoots were obtained on MS Murashige and Skoog (Plant Physiol. 15: 473–497, 1962) medium supplemented with 1 mg L−1 benzyladenine (BA), 20 mg L−1 l-glutamine and 2 mg L−1 phosphinothricin (PPT) or 100 mg L−1 kanamycin. The regenerated shoots were examined in vivo using a hand-held long wave UV lamp for GFP expression. The GFP screening helped identify escapes and chimeric shoots at regular intervals to increase the growth of transformed shoots on cotyledon explants. Elongation and rooting of putative transformants were achieved on PPT (2 mg L−1) containing MS media with 0.5 mg L−1 gibberellic acid (GA3) and 0.6 mg L−1 indole butyric acid (IBA), respectively. PCR and Southern analyses confirmed the integration of the sgfp gene into the genome of T0 and the progenies. T1 segregation of transgenic progeny exhibited Mendelian inheritance of the transgene. The use of EHA105 resulted in 21% transformation efficiency compared to 8.5% when LBA4404 was used. This higher rate was greatly facilitated by PPT selection coupled with effective screening of transformants for GFP expression, thus making the protocol highly useful for the recovery of a higher number of transgenic cucumber plants.  相似文献   

17.
Summary The generation of transgenic Cucumis sativus cv. Greenlong plants resistant to phosphinothricin (PPT) was obtained using Agrobacterium tumefaciens-mediated gene transfer. The protocol relied on the regeneration of shoots from cotyledon explants. Transformed shoots were obtained on Murashige and Skoog medium supplemented with 4.4 μM 6-benzylaminopurine 3.8 μM abscisic acid, 108.5 μM adenine sulfate, and 2 mg l−1 phosphinothricin. Cotyledons were inoculated with the strain EHA105 harboring the neomycin phosphotransferase II (npt II), and phosphinothricin resistance (bar) genes conferring resistance to kanamycin and PPT. Transformants were selected by using increasing concentrations of PPT (2–6 mg l−1). Elongation and rooting of putative transformants were performed on PPT-containing (2 mg l−1) medium with 1.4 μM gibberellic acid and 4.9 μM indolebutyric acid, respectively. Putative transformants were confirmed for transgene insertion through PCR and Southern analysis. Expression of the bar gene in transformed plants was demonstrated using a leaf painting test with the herbicide Basta. Pre-culture of explants followed by pricking, addition of 50 μM acetosyringone during infection, and selection using PPT rather than kanamycin were found to enhance transformation frequency as evidenced by transient β-glucuronidase assay. Out of 431 co-cultivated explants, 7.2% produced shoots that rooted and grew on PPT, and five different plants (1.1%) were demonstrated to be transgenic following Southern hybridization.  相似文献   

18.
The influence of aqueous environment on the main‐chain conformation (ω0, ?, and ψ dihedral angles) of two model peptoids: N‐acetyl‐N‐methylglycine N’‐methylamide (Ac‐N(Me)‐Gly‐NHMe) ( 1 ) and N‐acetyl‐N‐methylglycine N’,N’‐dimethylamide (Ac‐N(Me)‐Gly‐NMe2) ( 2 ) was investigated by MP2/6‐311++G(d,p) method. The Ramachandran maps of both studied molecules with cis and trans configuration of the N‐terminal amide bond in the gas phase and in water environment were obtained and all energy minima localized. The polarizable continuum model was applied to estimate the solvation effect on conformation. Energy minima of the Ac‐N(Me)‐Gly‐NHMe and Ac‐N(Me)‐Gly‐NMe2 have been analyzed in terms of the possible hydrogen bonds and C = O dipole attraction. To validate the theoretical results obtained, conformations of the similar structures gathered in the Cambridge Crystallographic Data Centre were analyzed. Obtained results indicate that aqueous environment in model peptoids 1 and 2 favors the conformation F (? and ψ = ?70º, 180º), and additionally significantly increases the percentage of structures with cis configuration of N‐terminal amide bond in studied compounds. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Summary Suspension cells of Oryza sativa L. (rice) were transformed, by microprojectile bombardment, with plasmids carrying the coding region of the Streptomyces hygroscopicus phosphinothricin acetyl transferase (PAT) gene (bar) under the control of either the 5 region of the rice actin 1 gene (Act1) or the cauliflower mosaic virus (CaMV) 35S promoter. Subsequently regenerated plants display detectable PAT activity and are resistant to BASTATM, a phosphinothricin (PPT)-based herbicide. DNA gel blot analyses showed that PPT resistant rice plants contain a bar-hybridizing restriction fragment of the expected size. This report shows that expression of the bar gene in transgenic rice plants confers resistance to PPT-based herbicide by suppressing an increase of ammonia in plants after spraying with the herbicide.  相似文献   

20.
We investigated the effects of genetic modification of nitrogen metabolism via the bacterial glutamate dehydrogenase (GDH) on plant growth and metabolism. The gdhA gene from Escherichia coli encoding a NADPH-GDH was expressed in tobacco plants under the control of the 35 S promoter. The specific activity of GDH in gdhA plants was 8-fold of that in E. coli. Damage caused by spray application of 1.35 mM of phosphinothricin (PPT) herbicide, a glutamine synthetase (GS) inhibitor, was less pronounced in gdhA plants as compared with the control plants which suggests that the introduced GDH can assimilate some of the excess ammonium, at least during GS inhibition. However, gdhA plants were susceptible to 2.7 mM PPT. Biomass production was consistently increased in gdhA transgenic plants grown under controlled conditions and in the field. Total free amino acids and total carbohydrates were increased in gdhA plants grown in the greenhouse suggesting that both nitrogen and carbon metabolism were altered. We conclude that the modifications in transgenic plants may result from both increased nitrogen efficiency and altered gene expression and metabolism. This revised version was published online in June 2006 with corrections to the Cover Date.  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