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 375 毫秒
1.
The quaternary ammonium compounds accumulated in saline conditions by five salt-tolerant species of Limonium (Plumbaginaceae) were analyzed by fast atom bombardment mass spectrometry. Three species accumulated β-alanine betaine and choline-O-sulfate; the others accumulated glycine betaine and choline-O-sulfate. Three lines of evidence indicated that β-alanine betaine and choline-O-sulfate replace glycine betaine as osmo-regulatory solutes. First, tests with bacteria showed that β-alanine betaine and choline-O-sulfate have osmoprotective properties comparable to glycine betaine. Second, when β-alanine betaine and glycine betaine accumulators were salinized, the levels of their respective betaines, plus that of choline-O-sulfate, were closely correlated with leaf solute potential. Third, substitution of sulfate for chloride salinity caused an increase in the level of choline-O-sulfate and a matching decrease in glycine betaine level. Experiments with 14C-labeled precursors established that β-alanine betaine accumulators did not synthesize glycine betaine and vice versa. These experiments also showed that β-alanine betaine synthesis occurs in roots as well as leaves of β-alanine betaine accumulators and that choline-O-sulfate and glycine betaine share choline as a precursor. Unlike glycine betaine, β-alanine betaine synthesis cannot interfere with conjugation of sulfate to choline by competing for choline and does not require oxygen. These features of β-alanine betaine may be advantageous in sulfate-rich salt marsh environments.  相似文献   

2.
We report here that the naturally occurring choline ester choline-O-sulfate serves as an effective compatible solute for Bacillus subtilis, and we have identified a high-affinity ATP-binding cassette (ABC) transport system responsible for its uptake. The osmoprotective effect of this trimethylammonium compound closely matches that of the potent and widely employed osmoprotectant glycine betaine. Growth experiments with a set of B. subtilis strains carrying defined mutations in the glycine betaine uptake systems OpuA, OpuC, and OpuD and in the high-affinity choline transporter OpuB revealed that choline-O-sulfate was specifically acquired from the environment via OpuC. Competition experiments demonstrated that choline-O-sulfate functioned as an effective competitive inhibitor for OpuC-mediated glycine betaine uptake, with a Ki of approximately 4 μM. Uptake studies with [1,2-dimethyl-14C]choline-O-sulfate showed that its transport was stimulated by high osmolality, and kinetic analysis revealed that OpuC has high affinity for choline-O-sulfate, with a Km value of 4 ± 1 μM and a maximum rate of transport (Vmax) of 54 ± 3 nmol/min · mg of protein in cells grown in minimal medium with 0.4 M NaCl. Growth studies utilizing a B. subtilis mutant defective in the choline to glycine betaine synthesis pathway and natural abundance 13C nuclear magnetic resonance spectroscopy of whole-cell extracts from the wild-type strain demonstrated that choline-O-sulfate was accumulated in the cytoplasm and was not hydrolyzed to choline by B. subtilis. In contrast, the osmoprotective effect of acetylcholine for B. subtilis is dependent on its biotransformation into glycine betaine. Choline-O-sulfate was not used as the sole carbon, nitrogen, or sulfur source, and our findings thus characterize this choline ester as an effective compatible solute and metabolically inert stress compound for B. subtilis. OpuC mediates the efficient transport not only of glycine betaine and choline-O-sulfate but also of carnitine, crotonobetaine, and γ-butyrobetaine (R. Kappes and E. Bremer, Microbiology 144:83–90, 1998). Thus, our data underscore its crucial role in the acquisition of a variety of osmoprotectants from the environment by B. subtilis.  相似文献   

3.
A search was undertaken for osmoprotective compounds for mouse hybridoma cell line 6H11 grown in culture. When the osmolality of the growth medium was increased above the normal osmolality of 330 mOsmol/kg, growth rates were decreased in a dose-dependent fashion, reaching zero when the osmolality of the medium reached approx. 435 mOsmol/kg through the addition of KCl (60 mM), or 510 mOsmol/kg through the addition of NaCl (100 mM), or sucrose (175 mM). For NaCl or sucrose-stressed cultures, the inclusion of glycine betaine, sarcosine, proline, glycine, or asparagine in the growth medium gave a moderate to strong osmoprotective effect, measured as the ability of these compounds to enhance cell growth rates under hyperosmotic conditions. Inclusion of dimethylglycine may also give a strong osmoprotective effect under these stress conditions.In KCl-stressed cell cultures, addition of glycine betaine, sarcosine, or dimethylglycine gave strong osmoprotective effects. Of 38 compounds tested during NaCl stress, 7 gave weak osmoprotective effects and 25 gave no osmoprotective effect. The osmoprotective compounds accumulated inside the stressed cells. Accumulation was completed after 4 to 8 h, reaching intracellular concentrations of approx. 0.27 pmol/cell, or 0.15 M, in NaCl stressed cells (100 mM NaCl added).Glycine betaine, dimethylglycine, and sarcosine accumulation was observed only when these protectants were included in the medium. For all osmoprotectants, a growth medium concentration between 5 and 30 mM gave the maximal protective effect, with the exception of dimethylglycine, for which the optimum concentration was approx. 65 mM. Osmoprotective effects obtained with glycine, sarcosine, dimethylglycine, and glycine betaine, indicate that the more methylated compounds are the most effective protectants.The cellular content of glycine betaine and the glycine betaine uptake rate increased with medium osmolality in a linear fashion. Glycine betaine uptake was described by a model comprising a saturable component obeying Michaelis-Menten kinetics and a nonsaturable component. K(m) and V(max) for glycine betaine uptake were determined at 420 mOsmol/kg (50 mM NaCl added) and 510 mOsmol/kg (100 mM NaCl added). A K(m) value of approx. 2.5 mM was obtained at both medium osmolalities, while V(max) increased from 0.010 pmol/cell . h to 0.018 pmol/cell . h as the osmolality of the growth medium was increased, indicating an effect of medium osmolality on the maximal rate of transport rather than on the affinity of the transporters for glycine betaine. Hybridoma cells were not able to utilize the glycine betaine precursors choline or glycine betaine aldehyde for osmoprotection, suggesting that the cells lack part, or all, of the choline-glycine betaine pathway or the appropriate uptake mechanism.The uptake rate for glycine in NaCl-stressed hybridoma cells was approx. four times higher than the uptake rate for glycine betaine. Furthermore, if equimolar amounts of glycine betaine, glycine, sarcosine, and proline were simultaneously added to NaCl-stressed cell cultures, the intracellular concentrations of glycine, proline, and sarcosine were significantly higher than the concentration of glycine betaine.A 40% increase in hybridoma cell volume was observed when the growth medium osmolality was increased from 300 to 520 mOsmol/kg. (c) 1994 John Wiley & Sons, Inc.  相似文献   

4.
The role of phosphatidylcholine turnover during hypersaline stress is investigated in Saccharomyces cerevisiae. In the wild-type strain, 2180-1A hypersaline stress induced the rapid turnover of phosphatidylcholine, a major membrane lipid. Yeast cells were grown in the presence of [14C]-choline to label phosphatidylcholine. Upon shifting the cells to medium with 0.8 M NaCl, phosphatidylcholine levels were diminished by c. 30% within 20 min to yield glycerophosphocholine, a methylamine osmoprotectant that has been previously identified in renal cells. High-performance liquid chromatography studies showed that osmotically mediated glycerophosphocholine production was enhanced if 10 mM choline was added as a supplement to synthetic dextrose medium with 1.6 M NaCl, but glycine betaine was not detected. Enhanced glycerophosphocholine production also correlated with improved growth in media containing 1.6 M NaCl and choline. Enhanced growth is specific to methylamines: salt-stressed cells supplemented with 10 mM choline or glycine betaine showed enhanced growth relative to unsupplemented control cultures, but other additives had no effect on growth or adversely affected it. Nutritional effects are ruled out because yeast cannot use choline or glycine betaine as carbon or nitrogen sources in normal or high-salt medium. Finally, enhanced growth in hypersaline media with choline or glycine betaine is dependent on the choline permease Hnm1. These results in yeast highlight a similarity with mammalian renal cells, namely that phosphatidylcholine turnover contributes to osmotic adaptation via synthesis of the osmoprotectant glycerophosphocholine.  相似文献   

5.
Among the Rhizobiaceae, Bradyrhizobium japonicum strain USDA110 appears to be extremely salt sensitive, and the presence of glycine betaine cannot restore its growth in medium with an increased osmolarity (E. Boncompagni, M. Østerås, M. C. Poggi, and D. Le Rudulier, Appl. Environ. Microbiol. 65:2072-2077, 1999). In order to improve the salt tolerance of B. japonicum, cells were transformed with the betS gene of Sinorhizobium meliloti. This gene encodes a major glycine betaine/proline betaine transporter from the betaine choline carnitine transporter family and is required for early osmotic adjustment. Whereas betaine transport was absent in the USDA110 strain, such transformation induced glycine betaine and proline betaine uptake in an osmotically dependent manner. Salt-treated transformed cells accumulated large amounts of glycine betaine, which was not catabolized. However, the accumulation was reversed through rapid efflux during osmotic downshock. An increased tolerance of transformant cells to a moderate NaCl concentration (80 mM) was also observed in the presence of glycine betaine or proline betaine, whereas the growth of the wild-type strain was totally abolished at 80 mM NaCl. Surprisingly, the deleterious effect due to a higher salt concentration (100 mM) could not be overcome by glycine betaine, despite a significant accumulation of this compound. Cell viability was not significantly affected in the presence of 100 mM NaCl, whereas 75% cell death occurred at 150 mM NaCl. The absence of a potential gene encoding Na+/H+ antiporters in B. japonicum could explain its very high Na+ sensitivity.  相似文献   

6.
Unlike most Lactobacillus acidophilus strains, a specific strain, L. acidophilus IFO 3532, was found to grow in rich medium containing 1 M sodium acetate, KCl, or NaCl. This strain could also grow with up to 1.8 M NaCl or 3 M nonelectrolytes (fructose, xylose, or sorbitol) added. Thus, this strain was tolerant to osmotic pressures up to 2.8 osM. A search for an intracellular solute which conferred osmoprotection led to the identification of glycine betaine (betaine). Betaine was accumulated to high concentrations in cells growing in MRS medium supplemented with 1 M KCl or NaCl. Uptake of [14C]betaine by L. acidophilus 3532 cells suspended in buffer was stimulated by increasing the medium osmotic pressure with 1 M KCl or NaCl. The accumulated betaine was not metabolized further; transport was relatively specific for betaine and was dependent on an energy source. Other lactobacilli, more osmosensitive than strain 3532, including L. acidophilus strain E4356, L. bulgaricus 8144, and L. delbrueckii 9649, showed lower betaine transport rates in response to an osmotic challenge than L. acidophilus 3532. Experiments with chloramphenicol-treated L. acidophilus 3532 cells indicated that the transport system was not induced but appeared to be activated by an increase in osmotic pressure.  相似文献   

7.
Glycine betaine relieved sodium chloride-mediated inhibition of growth in Azospirillum lipoferum ATCC 29708. 35S-methionine labelling of proteins after salinity up-shock revealed strong induction of a 30 kDa protein which cross-reacted with the anti-glycine betaine binding protein antibody from Escherichia coli. This suggested that A. lipoferum had a salinity-induced ProU-like high-affinity glycine betaine transport system. A genomic library of A. lipoferum ATCC 29708 was screened for the proU-like gene by complementation of a proU mutant of E. coli. Four recombinant cosmids, capable of restoring growth of the proU mutant on plates containing 600 mM NaCl and 1 mM glycine betaine were selected. Selected recombinant cosmids hybridized with a proU gene probe from E. coli. Complementation of E. coli proU mutant with the A. lipoferum genomic DNA was evident by the ability of proU mutant (containing selected recombinant cosmids) to grow on minimal medium supplemented with 600 mM NaCl and 1 mM glycine betaine.  相似文献   

8.
The present study demonstrates the regeneration of plantlets of guava (Psidium guajava L.) from somatic embryos developed under salt-stress conditions. With increasing concentrations of NaCl in induction medium (MS + 4.52 μM 2,4-d + 5% sucrose) from 0 to 200 mM, the number of somatic embryos per responsive explant decreased. Somatic embryos induced on 0–100 mM NaCl containing medium developed into torpedo stages, whereas, the development of somatic embryos that differentiated on 150 and 200 mM NaCl-supplemented medium was arrested prior to torpedo stage and did not undergo maturation phase. Somatic embryos that developed on NaCl-containing medium, showed better germination in the presence of NaCl as compared with those developed on medium without NaCl. The effect of increasing salt-stress was also investigated on plant growth, chlorophyll and carotenoids, Na+ and K+, and proline and glycine betaine accumulation in in vitro grown plantlets. The level of Na+ in leaves increased with increasing concentrations of NaCl in the medium. Accumulation of free proline and glycine betaine in leaves significantly increased with increasing salinity. The results suggest that accumulation of proline and glycine betaine may be important for osmotic adjustment in guava under salinity stress.  相似文献   

9.
Ensifer (Sinorhizobium) meliloti is a nitrogen-fixing α-proteobacterium able to biosynthesize the osmoprotectant glycine betaine from choline sulfate through a metabolic pathway that starts with the enzyme choline-O-sulfatase. This protein seems to be widely distributed in microorganisms and thought to play an important role in their sulfur metabolism. However, only crude extracts with choline sulfatase activity have been studied. In this work, Ensifer (Sinorhizobium) meliloti choline-O-sulfatase was obtained in a high degree of purity after expression in Escherichia coli. Gel filtration and dynamic light scattering experiments showed that the recombinant enzyme exists as a dimer in solution. Using calorimetry, its catalytic activity against its natural substrate, choline-O-sulfate, gave a kcat=2.7×10?1 s?1 and a KM=11.1 mM. For the synthetic substrates p-nitrophenyl sulfate and methylumbelliferyl sulfate, the kcat values were 3.5×10?2 s?1 and 4.3×10?2 s?1, with KM values of 75.8 and 11.8 mM respectively. The low catalytic activity of the recombinant sulfatase was due to the absence of the formylglycine post-translational modification in its active-site cysteine 54. Nevertheless, unmodified Ensifer (Sinorhizobium) meliloti choline-O-sulfatase is a multiple-turnover enzyme with remarkable catalytic efficiency.  相似文献   

10.
The sensitivity of industrial strains Acetobacter aceti, Gluconobacter frateurii, and Propionibacterium acidipropionici to osmotic stress was studied. Growth of A. aceti and G. frateurii was totally inhibited at 0.4 M NaCl concentration, but P. acidipropionici was able to grow on a medium containing 1.2 M NaCl. Addition of glycine betaine to the medium had no detectable osmoprotective effect on A. aceti and G. frateurii cultivations in elevated NaCl concentrations, but it enabled cells of P. acidipropionici to achieve faster the maximum specific growth rate after the prolonged lag phase and therefore to gain faster the final biomass and product concentrations. The final concentrations of biomass and product of P. acidipropionici were the same as for the cultivations of the bacterium without NaCl and glycine betaine present in the medium. Intracellular accumulation of glycine betaine was detected in P. acidipropionici cells cultivated in the medium containing glycine betaine. The amount accumulated increased with NaCl concentration, suggesting that glycine betaine plays an important role in the osmoadaptation. Received: 31 March 2000 / Received revision: 22 May 2000 / Accepted: 3 June 2000  相似文献   

11.
Natural-abundance 13C-nuclear magnetic resonance was used to probe the intracellular organic solute content of the moderately halophilic bacterium Tetragenococcus halophila. When grown in complex growth media supplemented or not with NaCl, T. halophila accumulates glycine betaine and carnitine. Unlike other moderate halophiles, T. halophila was not able to produce potent osmoprotectants (such as ectoines and glycine betaine) through de novo synthesis when cultured in defined medium under hyperosmotic constraint. Addition of 2 mM carnitine, glycine betaine, or choline to defined medium improved growth parameters, not only at high salinity (up to 2.5 M NaCl) but also in media lacking NaCl. These compounds were taken up when available in the surrounding medium. The transport activity occurred at low and high salinities and seems to be constitutive. Glycine betaine and carnitine were accumulated by T. halophila in an unmodified form, while exogenously provided choline led to an intracellular accumulation of glycine betaine. This is the first evidence of the existence of a choline-glycine betaine pathway in a lactic acid bacterium. An assay showed that the compatible solutes strikingly repressed the accumulation of glutamate and slightly increased the intracellular potassium level only at high salinity. Interestingly, osmoprotectant-treated cells were able to maintain the intracellular sodium concentration at a relatively constant level (200 to 300 nmol/mg [dry weight]), independent of the NaCl concentration of the medium. In contrast, in the absence of osmoprotectant, the intracellular sodium content increased sharply from 200 to 2,060 nmol/mg (dry weight) when the salinity of the medium was raised from 1 to 2 M. Indeed, the imported compatible solutes play an actual role in regulating the intracellular Na+ content and confer a much higher salt tolerance to T. halophila.  相似文献   

12.
Abstract The addition of 1 mM glycine betaine to the growth medium of Chromatium sp. NCIMB 8379 relieved growth inhibition caused by exposure to supra-optimal Nad concentrations. Intracellular glycine betaine concentrations were dependent upon the NaCl concentration of the growth medium up to 3 M exogenous Nad. Kinetic data for the accumulation of [methyl-14C]-glycine betaine demonstrated that Chromatium sp. NCIMB 8379 possesses a constitutively expressed active transport system for glycine betaine. The transport system was saturable with respect to glycine betaine concentration and exhibited typical Michaelis-Menten type kinetics: K m= 24 μ M, V max= 306 nmol min−1 mg protein−1 at an external NaCl concentration of 1 M. The rate of glycine betaine transport decreased progressively with increasing growth medium NaCl concentration. This transport system may represent an adaptive response to growth in high osmolarity environments in this halotolerant isolate, allowing accumulation of glycine betaine from the external cell environment or recycling synthesised glycine betaine which has passively diffused from the cell.  相似文献   

13.
In this study, we have investigated intrinsic salt tolerance of Astragalus cicer microsymbionts (USDA3350, ACMP18) and the role of exogenous glycine betaine in osmoprotection in these bacteria. Salt stress was imposed by NaCl concentrations ranging from 0.5 to 2 %. A. cicer mesorhizobia were capable of tolerating up to 2 % sodium chloride with a population count that was inversely proportional to the salt content. When the extracellular concentration of NaCl was raised to 2 %, the generation time of the UDSA3350 strain in the mid-exponential phase of growth was 3.9-times greater than that in the no-salt control medium, whereas the ACMP18 strain survived under the same conditions but did not multiply. Application of 1 mM glycine betaine into the salt-stressed rhizobium cultures increased the number of culturable bacteria, pointing out that this molecule was involved in restoration of osmotic balance. The decline in A. cicer symbiont viability in the medium with sodium chloride and the osmoprotective role of glycine betaine for these bacteria was confirmed in the experiment using the live/dead Bac Light Bacterial Vibility Kit. Data presented in this study showed the presence of proU-like genes in the genomes of A. cicer rhizobia with high sequence similarity to the genes of the ProU-like system in Sinorhizobium meliloti and the proU operon of Escherichia coli.  相似文献   

14.
The halophilic methanoarchaeon Methanohalophilus portucalensis can synthesize de novo and accumulate β-glutamine, N-acetyl-β-lysine, and glycine betaine (betaine) as compatible solutes (osmolytes) when grown at elevated salt concentrations. Both in vivo and in vitro betaine formation assays in this study confirmed previous nuclear magnetic resonance 13C-labelling studies showing that the de novo synthesis of betaine proceeded from glycine, sarcosine, and dimethylglycine to form betaine through threefold methylation. Exogenous sarcosine (1 mM) effectively suppressed the intracellular accumulation of betaine, and a higher level of sarcosine accumulation was accompanied by a lower level of betaine synthesis. Exogenous dimethylglycine has an effect similar to that of betaine addition, which increased the intracellular pool of betaine and suppressed the levels of N-acetyl-β-lysine and β-glutamine. Both in vivo and in vitro betaine formation assays with glycine as the substrate showed only sarcosine and betaine, but no dimethylglycine. Dimethylglycine was detected only when it was added as a substrate in in vitro assays. A high level of potassium (400 mM and above) was necessary for betaine formation in vitro. Interestingly, no methylamines were detected without the addition of KCl. Also, high levels of NaCl and LiCl (800 mM) favored sarcosine accumulation, while a lower level (400 mM) favored betaine synthesis. The above observations indicate that a high sarcosine level suppressed multiple methylation while dimethylglycine was rapidly converted to betaine. Also, high levels of potassium led to greater amounts of betaine, while lower levels of potassium led to greater amounts of sarcosine. This finding suggests that the intracellular levels of both sarcosine and potassium are associated with the regulation of betaine synthesis in M. portucalensis.  相似文献   

15.
Carthamus tinctorius L., rich in antioxidant compounds, is a herbal medicine. Biochemical mechanisms of adaptation to salinity stress in safflower are still poorly understood at the cellular level. For this purpose, callus cultures of four different genotypes of safflower were used in this study to evaluate changes in their biochemical (ionic content, proline, and glycine betaine), total phenolics content (TPC), total flavonoids content (TFD), antioxidant responses (2,2-diphenyl-1-picrylhydrazyl: DPPH assay and carotenoid content), and lipid peroxidation (malon dialdehyde content: MDA) under salinity stress. The calluses derived from hypocotyls were exposed to in vitro salt stress at different concentrations of sodium chloride (0, 100, 200, and 300 mM). A reducing trend was observed in K+ and carotenoid reserves of the calluses with increasing NaCl concentration while an increasing trend was observed in Na+ content, proline, MDA, TPC, TFD, and DPPH activity under the same conditions. Callus glycine betaine content was found to decrease in the medium containing 100 mM NaCl but increased beyond this concentration up to 300 mM NaCl. Positive and significant correlations were recognized between DPPH and total phenolics as well as DPPH and total flavonoid contents, demonstrating that phenolics are the main contributors to the potential antioxidant activity of safflower at the cellular level. Overall, the salt-tolerant genotypes of Mex.2-137 and Mex.2-138 were found capable of being processed for the production of secondary metabolites via NaCl elicitation.  相似文献   

16.
When 23 recombinant Chinese hamster ovary (rCHO) cell clones were cultivated in hyperosmolar medium resulting from NaCl addition (533 mOsm/kg), their specific thrombopoietin (TPO) productivity (q(TPO)) was increased. However, due to depressed cell growth at elevated osmolality, no enhancement in the maximum TPO titer was made in batch cultures of all 23 clones. To test the feasibility of using glycine betaine, known as a strong osmoprotective compound, for improved TPO production in hyperosmotic rCHO cell cultures, hyperosmotic batch cultures of 23 clones were performed in the presence of 15 mM glycine betaine. Glycine betaine was found to have a strong osmoprotective effect on all 23 clones. Inclusion of 15 mM glycine betaine in hyperosmolar medium enabled 22 clones to grow at 542 mOsm/kg, where most clones could not grow in the absence of glycine betaine, but at a cost of reduced q(TPO). However, the relative decrease in q(TPO) varied significantly among clones. Thus, efficacy of the simultaneous use of hyperosmotic pressure and glycine betaine as a means to improve foreign protein production was variable among clones. Six out of 23 clones displayed more than a 40% increase in the maximum TPO titer in the hyperosmolar medium containing glycine betaine, compared with that in the standard medium with a physiological osmolality. Taken together, the results obtained here emphasize the importance of selection of clones for the successful use of hyperosmotic pressure and glycine betaine as an economical means to improve TPO production.  相似文献   

17.
The concentrations of intracellular solutes in Listeria monocytogenes were examined in cells grown at various concentrations of NaCl. At 5% NaCl, cells contained elevated concentrations of potassium and glycine betaine compared with concentrations in cells grown without NaCl. At 7.5% NaCl, cells contained increased concentrations of K+, glycine betaine, glycine, alanine, and proline. Only glycine betaine, choline, or glycine promoted growth on a solidified defined medium containing 4% NaCl; there was no growth at higher concentrations of NaCl in the defined medium.  相似文献   

18.
The concentrations of intracellular solutes in Listeria monocytogenes were examined in cells grown at various concentrations of NaCl. At 5% NaCl, cells contained elevated concentrations of potassium and glycine betaine compared with concentrations in cells grown without NaCl. At 7.5% NaCl, cells contained increased concentrations of K+, glycine betaine, glycine, alanine, and proline. Only glycine betaine, choline, or glycine promoted growth on a solidified defined medium containing 4% NaCl; there was no growth at higher concentrations of NaCl in the defined medium.  相似文献   

19.
Among the Rhizobiaceae, Bradyrhizobium japonicum strain USDA110 appears to be extremely salt sensitive, and the presence of glycine betaine cannot restore its growth in medium with an increased osmolarity (E. Boncompagni, M. Osteras, M. C. Poggi, and D. Le Rudulier, Appl. Environ. Microbiol. 65:2072-2077, 1999). In order to improve the salt tolerance of B. japonicum, cells were transformed with the betS gene of Sinorhizobium meliloti. This gene encodes a major glycine betaine/proline betaine transporter from the betaine choline carnitine transporter family and is required for early osmotic adjustment. Whereas betaine transport was absent in the USDA110 strain, such transformation induced glycine betaine and proline betaine uptake in an osmotically dependent manner. Salt-treated transformed cells accumulated large amounts of glycine betaine, which was not catabolized. However, the accumulation was reversed through rapid efflux during osmotic downshock. An increased tolerance of transformant cells to a moderate NaCl concentration (80 mM) was also observed in the presence of glycine betaine or proline betaine, whereas the growth of the wild-type strain was totally abolished at 80 mM NaCl. Surprisingly, the deleterious effect due to a higher salt concentration (100 mM) could not be overcome by glycine betaine, despite a significant accumulation of this compound. Cell viability was not significantly affected in the presence of 100 mM NaCl, whereas 75% cell death occurred at 150 mM NaCl. The absence of a potential gene encoding Na(+)/H(+) antiporters in B. japonicum could explain its very high Na(+) sensitivity.  相似文献   

20.
On transferring three-week-old plants of Aster tripolium L. growing in a half strength Hoagland's medium to the same medium containing 333 m M NaCl a very quick uptake of salt and, after a lag phase of 3 to 5 h, an increase in free proline level was observed. During the time course of imino acid storage, the accumulation rates were higher in the light than in the dark, thereby suggesting some kind of photocontrol on solute metabolism. At zero time, high levels of glycine betaine were present in young plants grown without salt. However, after the application of saline shock, the betaine level also increased significantly. The highest rate of betaine accumulation was detected during the third day of treatment when the rate of proline storage decreased. Glycine betaine storage could also be linked to light dependent processes; whatever its importance in response to salt shock was, the levels observed were lower than those of plants directly grown on 333 m M NaCl for three weeks. When saline stressed plants were transferred to a medium without NaCl, the proline level quickly decreaed while that of glycine betaine remained stable.  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