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.
Addition of osmoprotective compounds has a positive effect on growth and monoclonal antibody production in hyperosmotic hybridoma cell cultures. In order to better understand the processes involved in the osmoprotective response, uptake of the osmoprotective compounds glycine betaine, proline, sarcosine and glycine in mouse hybridoma cell line 6H11 during exposure to hyperosmotic stress was studied. Hyperosmotic stress (510 mOsmol/kg) was introduced through the addition of NaCl (100 mM) to the growth medium, and amino acid transport activity was measured immediately after transfer of the cells to the hyperosmotic medium. The osmoprotective capability of the four osmoprotectants tested was negatively affected if methylaminosobutyric acid (MeAiB), a specific substrate for amino acid transport system A, was simultaneously included in the hyperosmotic medium in equimolar amounts with one of the osmoprotective compounds. This was due to accumulation of MeAiB in the stressed cells, giving a significant reduction in the concentration of the osmoprotective compound inside the cells. Furthermore, addition of excess meAiB gave approx. 905 reduction in the initial rate of uptake of glycine betaine, while 40–50% reduction in the initial rate of uptake of proline, glycine and sarcosine. Similarly, addition of proline, glycine or sarcosine also gave a significant reduction in the initial rate of glycine betaine uptake. These results suggest that the four osmoprotective compounds share, at least in part, a common, MeAiB inhibitable carrier for transport into osmotically stressed hybridoma cells. This carrier is probably equal to amino acid transport system A.  相似文献   

2.
When mouse hybridoma cells were grown in culture media which were made hyperosmotic through the addition of NaCl or sucrose, the specific rate of antibody production increased with medium osmolality, reaching approx. 1.9 times the level obtained at physiological osmolality. However, due to a simultaneous reduction of the maximal cell density in the hyperosmotic media, the effect of the increased production rate did not give significant increases in the maximum antibody titer obtained in the cultures. When the osmoprotective compound, glycine betaine, was included in the NaCl- or sucrose-stressed cultures, the specific antibody production rate wasincreased up to 2.6-fold and maximum antibody titer up to twofold over that obtained in the control culture (physiological osmolality). A similar pattern of response was observed when other osmoprotective compounds (sarcosine, proline, glycine) were added to NaCl-stressed hybridoma cell cultures. For the present experiments, the results suggest that medium osmolality, rather than growth rate, will determine the specific antibody production rate by hybridoma cell line 6H11 growing in hyperosmotic culture media. (c) 1994 John Wiley & Sons, Inc.  相似文献   

3.
4.
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   

5.
Exogenous proline betaine ( stachydrine or N- dimethylproline ) or gamma-butyrobetaine (gamma-trimethylaminobutyrate), at a concentration as low as 1 mM, were found to stimulate the growth rate of Klebsiella pneumoniae, wild type M5A1 , in media of inhibitory osmotic strength (0.8 M NaC1). Simultaneously, nitrogen fixation by whole cells, a process particularly sensitive to osmotic stress, was strongly enhanced by these compounds. However, in the absence of sodium chloride, both the growth and nitrogen fixation were not affected by the addition of the methylammonium derivatives in the medium. The sensitivity of the nitrogen fixation to osmotic stress was used as a bioassay to evaluate the potentiality of osmoprotective compound in relation to the number of methyl groups on the nitrogen atom of glycine, proline, and gamma-aminobutyrate. Experiments with sarcosine ( monomethylglycine ), dimethylglycine, and glycine betaine ( trimethylglycine ), or experiments with mono- and di- methylproline or gamma-mono-, gamma-di, gamma-tri- methylaminobutyrate , indicated that the greatest stress tolerance was always obtained with the more N-methylated compounds.  相似文献   

6.
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   

7.
A total of 15 rhizobial strains representing Rhizobium meliloti, Rhizobium japonicum, Rhizobium trifolii, Rhizobium leguminosarum, Rhizobium sp. (Sesbania rostrata) and Rhizobium sp. (Hedysarum coronarium), were studied with regard to growth rate under salt stress in defined liquid media. In the presence of inhibitory concentrations of NaCl, enhancement of growth resulting from added glycine betaine was observed for R. meliloti strains and Rhizobium sp. (Hedysarum coronarium) but not for other Rhizobium species. The concentration of glycine betaine required for maximal growth stimulation was very low (1 mM) in comparison with the osmolarity of the medium. The stimulation was shown to be independent of any specific solutes. Other related compounds like proline betaine, carnitine, choline, -butyrobetaine and pipecolate betaine were also effective compounds in restoring the growth rate of cells grown in medium of elevated osmolarity. High rate of glycine betaine uptake was demonstrated in R. meliloti cells grown in media of increased osmotic strength. The intracellular concentration of this solute was found to be 308 mM in 0.3 M NaCl-grown cells and 17 times lower in minimal medium-grown cells. Glycine betaine was used for growth under conditions of low osmolarity but could not serve as sole carbon or nitrogen source in medium of increased osmotic strength. Experiments with [14C]glycine betaine showed that this molecule was not metabolized by cells subjected to osmotic stress, whereas it was rapidly converted to dimethylglycine, sarcosine and glycine in minimal medium-grown cells.Abbreviations LAS lactate-aspartate-salts - LGS lactate-glutamate-salts - LS lactate-succinate - MSY mannitol-salts-yeast - YLS yeast-lactate-succinate  相似文献   

8.
Elevated osmolality and pCO(2) have been shown to alter sialylation in a protein-specific manner. In Chinese hamster ovary (CHO)MT2-l-8 cells, tPA sialylation changed only slightly from 40 to 250 mm Hg pCO(2), whereas neural cell adhesion molecule polysialic acid (NCAM PSA) content decreased by up to 70% at 250 mm Hg pCO(2), pH 7.2. NCAM PSA content also decreased with increasing NaCl or NH(4)Cl concentration. This suggests that PSA content is a sensitive indicator of conditions that may alter glycosylation. Amino acids and their derivatives have been used to protect hybridoma and CHO cell growth under hyperosmotic stress. We examined the impact of osmoprotectants on NCAM PSA content in CHO MT2-1-8 cells under hyperosmolality (up to 545 mOsm/kg) and at 195 and 250 mm Hg pCO(2). NCAM PSA content at 545 mOsm/kg was at least two-fold greater in the presence of glycine betaine or L-proline compared to that without osmoprotectant. Surprisingly, in the presence of 20 mM glycine betaine, PSA levels were 50-60% of the control level for osmolalities ranging from 320 to 545 mOsm/kg. Thus, glycine betaine inhibits NCAM polysialylation at osmolalities below 435 mOsm/kg and is beneficial at higher osmolalities. In contrast to glycine betaine, L-proline increased PSA content by 25-120% relative to the unprotected culture at < or =545 mOsm/kg. The decrease in NCAM PSA levels of CHO MT2-1-8 cells cultured at 195 mm Hg pCO(2)-435 mOsm/kg was not mitigated by the presence of 25 mM glycine betaine, glycine, or L-threonine, even though all of these compounds enhanced cell growth. At 250 mm Hg pCO(2), all osmoprotectants tested (20 mM L-threonine, L-proline, glycine, or glycine betaine) increased NCAM polysialylation, with 20 mM glycine betaine restoring NCAM PSA to near control levels. Thus, osmoprotectants may (partially) offset changes in glycosylation, as well as the inhibition of growth, in cells under environmental stress. Supernatant beta-galactosidase levels, which increase upon alkalization of acidic organelles, did not differ significantly under elevated pCO(2) and hyperosmolality from that at control conditions.  相似文献   

9.
Biosynthesis of glycine betaine from simple carbon sources as compatible solute is rare among aerobic heterotrophic eubacteria, and appears to be almost exclusive to the non-halophilic and slightly halophilic phototrophic cyanobacteria. Although Synechococcus sp. WH8102 (CCMP2370), a unicellular marine cyanobacterium, could grow up to additional 2.5% (w/v) NaCl in SN medium, natural abundance 13C nuclear magnetic resonance spectroscopy identified glycine betaine as its major compatible solute. Intracellular glycine betaine concentrations were dependent on the osmolarity of the growth medium over the range up to additional 2% NaCl in SN medium, increasing from 6.8 ± 1.5 to 62.3 ± 5.5 mg/g dw. The ORFs SYNW1914 and SYNW1913 from Synechococcus sp. WH8102 were found as the homologous genes coding for glycine sarcosine N-methyltransferase and sarcosine dimethylglycine N-methyltransferase, heterologously over-expressed respectively as soluble fraction in Escherichia coli BL21(DE3)pLysS and purified by Ni-NTA His•bind resins. Their substrate specificities and the values of the kinetic parameters were determined by TLC and 1H NMR spectroscopy. RT-PCR analysis revealed that the two ORFs were both transcribed in cells of Synechococcus sp. WH8102 growing in SN medium without additional NaCl, which confirmed the pathway of de novo synthesizing betaine from glycine existing in these marine cyanobacteria.  相似文献   

10.
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   

11.
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   

12.
13.
Yersinia enterocolitica is a gram-negative, food-borne pathogen that can grow in 5% NaCl and at refrigerator temperatures. In this report, the compatible solutes (osmolytes) which accumulate intracellularly and confer the observed osmotic tolerance to this pathogen were identified. In minimal medium, glutamate was the only detectable osmolyte that accumulated in osmotically stressed cells. However, when the growth medium was supplemented with glycine betaine, dimethylglycine, or carnitine, the respective osmolyte accumulated intracellularly to high levels and the growth rates of the osmotically stressed cultures improved from 2.4- to 3.5-fold. Chill stress also stimulated the intracellular accumulation of glycine betaine, but the growth rate was only slightly improved by this osmolyte. Both osmotic upshock and temperature downshock stimulated the rate of uptake of [(sup14)C]glycine betaine by more than 30-fold, consistent with other data indicating that the osmolytes are accumulated from the growth medium via transport.  相似文献   

14.
Elevated pCO(2) inhibits cell growth. This growth inhibition is accompanied by a decrease in intracellular pH (pHi), as well as a decrease in glycolysis. Elevated concentrations (mM) of some amino acids have been shown by others to protect cells exposed to two very different environmental stresses: nutrient starvation and hyperosmolality. The fact that many of the amino acids shown to have protective effects against other stresses are transported into the cell through a pHi-sensitive transporter led us to study the possibility of using these amino acids as protective agents under elevated pCO(2). Screening experiments using 5, 15, and 25 mM of each amino acid showed that not all amino acids that protect cells from hyperosmolality protect them from elevated pCO(2). Glycine betaine and glycine were chosen for further characterization in both hybridoma and CHO cells. Asparagine and threonine were also tested in hybridoma and CHO cells, respectively. All amino acids tested under 195 mm Hg pCO(2)/435 mOsm/kg (50% growth inhibition) restored the specific growth rate (mu) in hybridoma cells to that observed under control conditions (40 mm Hg/320 mOsm/kg). Addition of each amino acid resulted in an increase in the consumption rate and intracellular accumulation of that amino acid. In CHO cells, glycine betaine also restored mu to control values, while glycine and threonine partially restored mu. In hybridoma cells, the higher specific antibody productivity obtained at elevated pCO(2) was maintained with the lowest amino acid concentration (5 mM). Productivity decreased toward control values with increasing amino acid concentrations. Elevated pCO(2) decreased the specific tPA productivity in the CHO cell line studied. Only glycine betaine resulted in a 20% increase in productivity at 195 mm Hg/435 mOsm/kg. With the exception of glycine betaine in hybridoma cells, amino acids did not mitigate the associated pHi decrease of at least 0.2 pH units at 195 mm Hg/435 mOsm/kg. pHi in hybridoma cells under elevated pCO(2) in the presence of glycine betaine was about 0.1 pH units below that of control. Amino acids had no effect on the cell size response of hybridoma cells, while they partially offset the increase in CHO cell size at elevated pCO(2). Glycine betaine, asparagine, and glycine increased the specific glucose consumption rate observed at 195 mm Hg/435 mOsm/kg (50% of control) to values greater than 70% of control in hybridoma cells. In CHO cells, only glycine betaine increased q(glc) (by 20%) under elevated pCO(2). All amino acids tested improved the cell yield from glutamine at 195 mm Hg/435 mOsm/kg in both cell lines.  相似文献   

15.
Listeria monocytogenes is a pathogenic bacterium that can grow at low temperatures and elevated osmolarity. The organism survives these stresses by the intracellular accumulation of osmolytes: low-molecular-weight organic compounds which exert a counterbalancing force. The primary osmolyte in L. monocytogenes is glycine betaine, which is accumulated from the environment via two transport systems: glycine betaine porter I, an Na(+)-glycine betaine symporter; and glycine betaine porter II, an ATP-dependent transporter. The biochemical characteristics of glycine betaine porter I were investigated in a mutant strain (LTG59) lacking the ATP-dependent transporter. At 4% NaCl, glycine betaine uptake in LTG59 was about fivefold lower than in strain DP-L1044, which has both transporters, indicating that the ATP-dependent transporter is the primary means by which glycine betaine enters the cell. In the absence of osmotic stress, cold-activated uptake by both transporters was most rapid between 7 and 12 degrees C, but a larger fraction of the total uptake was via the ATP-dependent transporter than was observed under salt-stressed conditions. Twelve glycine betaine analogs were tested for their ability to inhibit glycine betaine uptake and growth of stressed cultures. Carnitine, dimethylglycine, and gamma-butyrobetaine appear to inhibit the ATP-dependent transporter, while trigonelline and triethylglycine primarily inhibit glycine betaine porter I. Triethylglycine was also able to retard the growth of osmotically stressed L. monocytogenes grown in the presence of glycine betaine.  相似文献   

16.
Bacteroids isolated from alfalfa nodules induced by Rhizobium meliloti 102F34 transported glycine betaine at a constant rate for up to 30 min. Addition of sodium salts greatly increased the uptake activity, whereas other salts or non-electrolytes had less effect. The apparent Km for glycine betaine uptake was 8.3 microM and V was about 0.84 nmol min-1 (mg protein)-1 in the presence of 200 mM-NaCl which gave maximum stimulation of the transport. Supplementing bacteroid suspensions with various energy-yielding substrates, or ATP, did not increase glycine betaine uptake rates. The uncoupler carbonyl cyanide m-chlorophenylhydrazone (CCCP), and the respiratory inhibitor potassium cyanide strongly inhibited glycine betaine uptake, but arsenate was totally inactive. Glycine betaine transport showed considerable structural specificity: choline, proline betaine, gamma-butyrobetaine and trigonelline did not competitively inhibit the system, although choline and proline betaine were transported by bacteroids. Both a high-affinity activity and a low-affinity activity were found for choline uptake. These osmoprotective compounds might have a significant role in the maintenance of nitrogenase activity in bacteroids subjected to salt stress.  相似文献   

17.
Glycine betaine transport in Escherichia coli: osmotic modulation.   总被引:58,自引:36,他引:22  
Exogenous glycine betaine highly stimulates the growth rate of various members of the Enterobacteriaceae, including Escherichia coli, in media with high salt concentrations (D. Le Rudulier and L. Bouillard, Appl. Environ. Microbiol. 46:152-159, 1983). In a nitrogen- and carbon-free medium, glycine betaine did not support the growth of E. coli either on low-salt or high-salt media. This molecule was taken up by the cells but was not catabolized. High levels of glycine betaine transport occurred when the cells were grown in media of elevated osmotic strength, whereas relatively low activity was found when the cells were grown in minimal medium. A variety of electrolytes, such as NaCl, KCl, NaH2PO4, K2HPO4, K2SO4, and nonelectrolytes like sucrose, raffinose, and inositol triggered the uptake of glycine betaine. Furthermore, in cells subjected to a sudden osmotic upshock, glycine betaine uptake showed a sixfold stimulation 30 min after the addition of NaCl. Part of this stimulation might be a consequence of protein synthesis. The transport of glycine betaine was energy dependent and occurred against a concentration gradient. 2,4-Dinitrophenol almost totally abolished the glycine betaine uptake. Azide and arsenate exerted only a small inhibition. In addition, N,N'-dicyclohexylcarbodiimide had a very low inhibitory effect at 1 mM. These results indicated that glycine betaine transport is driven by the electrochemical proton gradient. The kinetics of glycine betaine entry followed the Michaelis-Menten relationship, yielding a Km of 35 microM and a Vmax of 42 nmol min-1 mg of protein-1. Glycine betaine transport showed considerable structural specificity. The only potent competitor was proline betaine when added to the assay mixtures at 20-fold the glycine betaine concentration. From these results, it is proposed that E. coli possesses an active and specific glycine betaine transport system which is regulated by the osmotic strength of the growth medium.  相似文献   

18.
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   

19.
Listeria monocytogenes is a pathogenic bacterium that can grow at low temperatures and elevated osmolarity. The organism survives these stresses by the intracellular accumulation of osmolytes: low-molecular-weight organic compounds which exert a counterbalancing force. The primary osmolyte in L. monocytogenes is glycine betaine, which is accumulated from the environment via two transport systems: glycine betaine porter I, an Na+-glycine betaine symporter; and glycine betaine porter II, an ATP-dependent transporter. The biochemical characteristics of glycine betaine porter I were investigated in a mutant strain (LTG59) lacking the ATP-dependent transporter. At 4% NaCl, glycine betaine uptake in LTG59 was about fivefold lower than in strain DP-L1044, which has both transporters, indicating that the ATP-dependent transporter is the primary means by which glycine betaine enters the cell. In the absence of osmotic stress, cold-activated uptake by both transporters was most rapid between 7 and 12°C, but a larger fraction of the total uptake was via the ATP-dependent transporter than was observed under salt-stressed conditions. Twelve glycine betaine analogs were tested for their ability to inhibit glycine betaine uptake and growth of stressed cultures. Carnitine, dimethylglycine, and γ-butyrobetaine appear to inhibit the ATP-dependent transporter, while trigonelline and triethylglycine primarily inhibit glycine betaine porter I. Triethylglycine was also able to retard the growth of osmotically stressed L. monocytogenes grown in the presence of glycine betaine.  相似文献   

20.
In Escherichia coli the osmoprotective compound glycine betaine is produced from choline by two enzymes; choline dehydrogenase (CDH) oxidizes choline to betaine aldehyde and then further on to glycine betaine, while betaine aldehyde dehydrogenase (BADH) facilitates the conversion of betaine aldehyde to glycine betaine. To evaluate the importance of BADH, a BADH/CDH fusion enzyme was constructed and expressed in E. coli and in Nicotiana tabacum. The fusion enzyme displayed both enzyme activities, and a coupled reaction could be measured. The enzyme was characterized regarding molecular weight and the dependence of the enzyme activities on environmental factors (salt, pH, and poly(ethylene glycol) addition). At high choline concentrations, E. coli cells expressing BADH/CDH were able to grow to higher final densities and to accumulate more glycine betaine than cells expressing CDH only. The intracellular glycine betaine levels were almost 5-fold higher for BADH/CDH when product concentration was related to CDH activity. Also, after culturing the cells at high NaCl concentrations, more glycine betaine was accumulated. On medium containing 20 mM choline, transgenic tobacco plants expressing BADH/CDH grew considerably faster than vector-transformed control plants.  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