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.
Plants that can take up amino acids directly from the soil solution may have a competitive advantage in ecosystems where inorganic nitrogen sources are scarce. We hypothesized that diverse mosses in cold, N-stressed ecosystems share this ability. We experimentally tested 11 sub-arctic Swedish moss species of wide-ranging taxa and growth form for their ability to take up double labelled (15N and 13C) glycine and aspartic acid in a laboratory setup as well as in a realistic field setting. All species were able to take up amino acids injected into the soil solution to some extent, although field uptake was marginal to absent for the endohydric Polytrichum commune. The 11 moss species on average took up 36 ± 5% of the injected glycine and 18 ± 2% of the aspartic acid in the lab experiment. Field uptake of both glycine (24 ± 5%) and aspartic acid (10 ± 2%) was lower than in the lab. Overall differences in uptake amongst species appeared to be positively associated with habitat wetness and/or turf density among different Sphagnum species and among non-Sphagnum species, respectively. Species from habitats of lower inorganic N availability, as indicated tentatively by lower tissue N concentrations, showed relatively strong amino acid uptake, but this was only significant for the field uptake among non-Sphagnum mosses. Further experiments are needed to test for consistent differences in amino acid uptake capacity among species and functional groups as determined by their functional traits, and to test how the affinity of cold-biome mosses for amino acids compares to that for ammonium or nitrate. Still, our results support the view that widespread moss species in cold, N-stressed ecosystems may derive a significant proportion of their nitrogen demand from free amino acids. This might give them a competitive advantage over plants that depend strongly on inorganic N sources. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

2.
Summary As steroid hormones are known to influence amino acid metabolism we tested the hypothesis that ovariectomy should lead to significant changes in this system.We found that after ovariectomy serum alanine was significantly decreased (p = 0.0006) in contrast to serum glycine and branched chain amino acids (BCAA). The ratio of glycine/BCAA, a parameter for anabolism or catabolism was not changed after ovariectomy. If, however, the amino acid alanine as the link to carbohydrate and lipid metabolism was introduced the alanine/BCAA ratio was significantly altered (p = 0.01).Although serum cholesterol was altered as well (increased,p = 0.03), no significant correlation with alanine was found. We can therefore assume that there are two independent mechanisms for lipid and amino acid changes after ovariectomy.The most prominent finding was that estradiol replacement after ovariectomy restored increased cholesterol levels but did not restore alanine levels. Other ovarial hormones must be incriminated for the regulation of alanine metabolism. The anabolic effects of estradiol as decreasing glycine and BCAA were noticed which rules out insufficient estradiol replacement.  相似文献   

3.
It was found that the production of amylase and proteinase by washed cells of Bacillus subtilis var. amyloliquefaciens was inhibited by glycine and its peptides but not by glycine derivatives, in which the free amino group was protected with various groups. Incorporation experiments of glycine-C14 revealed that about 60 per cent of the radioactivity which had been incorporated into the cells was found in the free amino acid fraction of the bacteria. The inhibitory effect of glycine was easily reversed by the addition of amino acid such as alanine, methionine and glutamic acid. Spermine also caused the reversal of inhibition of the enzyme production by glycine.  相似文献   

4.
Aims: It is well established that the bile salt sodium taurocholate acts as a germinant for Clostridium difficile spores and the amino acid glycine acts as a co‐germinant. The aim of this study was to determine whether any other amino acids act as co‐germinants. Methods and Results: Clostridium difficile spore suspensions were exposed to different germinant solutions comprising taurocholate, glycine and an additional amino acid for 1 h before heating shocking (to kill germinating cells) or chilling on ice. Samples were then re‐germinated and cultured to recover remaining viable cells. Only five amino acids out of the 19 common amino acids tested (valine, aspartic acid, arginine, histidine and serine) demonstrated co‐germination activity with taurocholate and glycine. Of these, only histidine produced high levels of germination (97·9–99·9%) consistently in four strains of Cl. difficile spores. Some variation in the level of germination produced was observed between different PCR ribotypes, and the optimum concentration of amino acids with taurocholate for the germination of Cl. difficile NCTC 11204 spores was 10–100 mmol l?1. Conclusions: Histidine was found to be a co‐germinant for Cl. difficile spores when combined with glycine and taurocholate. Significance and Impact of the Study: The findings of this study enhance current knowledge regarding agents required for germination of Cl. difficile spores which may be utilized in the development of novel applications to prevent the spread of Cl. difficile infection.  相似文献   

5.
Influx of glycine from dilute solution in the medium into Phaeodactylum tricornutum (Bohlin) was studied using radiochemical techniques. Comparison of these data with simultaneous measurements of the disappearance of primary amines from the medium by fluorometry indicates that influx of 14C-labeled glycine accurately reflects net entry of substrate. High performance liquid chromatography (HPLC) was employed to demonstrate net removal of fourteen different amino acids from dilute solution by P. tricornutum. HPLC was also used to demonstrate net removal of free amino acids naturally present in sea water.  相似文献   

6.
WhenStaphylococcus aureus cells were labeled with a single radioactive amino acid for 20 minutes, the highest activity, except for alanine, leucine, and glycine, was found in the free pool. Significant amounts of the above amino acids and also valine and methionine were incorporated into the protein — cell wall fraction.Cells previously labeled with a single amino acid underwent a net loss of radioactivity when transferred to buffer, glucose, or complete medium. An exception was glycine. The greatest loss in activity occurred in the free pool.While some amino acids (alanine, cystine) were transferred from the free pool to the protein — cell wall fraction under all conditions tested, others (glutamic acid, proline) were transferred only under conditions of growth.Cells labeled with certain single amino acids and then transferred to a complete medium lost a significant portion of the label. The most extreme case noted was proline, but other amino acids also effluxed from the cell under these conditions.  相似文献   

7.
To disclose the addition of some strong promotional amino acids (namely glycine, glutamate, lysine and aspartic acid) is how to improve the glycerol productivity of Candida glycerinogenes. An amino acid addition strategy based on dynamic enzyme activity was applied to improve glycerol productivity and decrease the byproducts formation in a fermentation of C. glycerinogenes in a 7-1 bioreactor. Compared with the control, after feeding glycine, glutamate, lysine and aspartic acid, glycerol productivity obtained an increase of 22.3, 25.6, 23.5 and 28.6%, respectively; meanwhile, the amounts of ethanol, acetic acid and pyruvate decreased largely. Whichever glycine, lysine, glutamate or aspartic acid was fed could elevate the activities of glucose-6-phosphate dehydrogenase (G6PDH), citrate synthase (CIT), triosephosphate isomerase (TPI) and cytoplasmic NAD+AEAAKw-dependent glycerol-3-phosphate dehydrogenase (ctGPD), and reduce the activities of pyruvate kinase (PYK), phosphofructokinase (PFK) and alcohol dehydrogenase (ADH). The reason of glycerol overproduction by the yeasts after feeding glycine, glutamate, lysine or aspartic acid is that the anaplerosis of intermediate metabolites in TCA cycle for amino acid degradation can decrease the flux from Embden-Meyerhof-Parnas (EMP) pathway to TCA cycle and enhance the flux through glycerol biosynthesis pathway. Above all, not only the high active hexose monophosphate (HMP) pathway but also the high dihydroxyacetone phosphate (DHAP) level plays an important role in the high glycerol productivity of C. glycerinogenes. The strategy of amino acid supplement is significant and can be economically implemented by an online process control strategy for higher yield of glycerol in industrial scale. Published in Russian in Prikladnaya Biokhimiya i Mikrobiologiya, 2009, Vol. 45, No. 3, pp. 338–343. The article is published in the original.  相似文献   

8.
Amino acid polyoxometalate nanoparticles-embedded microcapsules were in situ fabricated by layer-by-layer (LbL) self-assembly method [polyoxometalate, H3PMo12O40·nH2O (PMo12); amino acid, glycine (Gly)]. The morphology of the obtained microcapsules was characterized by transmission electron microscopy and scanning electron microscopy. The electrochemical behavior of the amino acid polyoxometalate nanoparticles-embedded microcapsules was studied by cyclic voltammetry. The microcapsules show the pH-dependent properties, indicating that the pH of solution plays an important role in the electrochemical behavior of heteropolyanions.  相似文献   

9.
The characteristics of amino acid uptake by frog spinal cord slices was studied by in vitro incubations in appropriate media. The uptake mechanisms exhibited saturation; kinetic analysis demonstrated 2 distinct systems for the influx of the possible neurotransmitters: GABA, glycine, L-glutamic acid and L-aspartic acid. One system showed a comparatively high substrate affinity (Km values, 10-26 μM) while the other system had a lower affinity (Km, 0.4-1.6 mM).-Leucine, an amino acid presumably not a transmitter, was accumulated only by a low affinity mechanism (Km 1.6 mM). The process responsible for high affinity uptake had many of the properties of an active transport mechanism. These included temperature sensitivity, energy dependence, requirement for Na+ ions and inhibition by ouabain. GABA and glycine uptake was inhibited only by closely related amino acids or structural analogues. The influx of L-glutamic acid was competitively inhibited by the presence of L-aspartic acid in the medium; the converse was also demonstrated. Thus, the high affinity uptake system for possible transmitter amino acids in the frog spinal cord closely resembles that described for mammalian CNS tissue. These results are compatible with the assumption that GABA, glycine, L-glutamic acid and L-aspartic acid are neurotransmitters in the amphibian spinal cord.  相似文献   

10.
This study analyzes the relationship between amino acids and pain perception during active labor. Cerebrospinal fluid (CSF) levels of the excitatory amino acids (EAAs)—glutamate, aspartate and their amide forms, inhibitory amino acids (IAAs)—glycine, γ-amino butyric acid (GABA) and taurine and nitric oxide (NO) related compounds—arginine and citrulline (by-product of NO synthesis) were compared between pregnant women at term pregnancy with labor pain (n = 38) and without labor pain (Caesarian section; n = 30). The levels of aspartate, glycine, GABA and citrulline were significantly higher; whilst taurine was significantly lower in the labor pain group. These findings suggest that aspartate and NO are associated with labor pain. An inhibitory role for the IAA taurine and a pronociceptive role for glycine in labor pain are proposed.  相似文献   

11.
An evolutionary scheme is postulated in which a primitive code, involving only guanine and cytosine, would code for glycine (GG), alanine (GC), arginine (CG) and proline (CC). From each of these amino acids and their codons, there evolves a family of related amino acids as the code expands. The four families are: (1)alanine valine, leucine, isoleucine, phenylalanine, tyrosine, methionine and tryptophane; (2)proline, threonine and serine; (3)arginine, lysine, and histidine; (4)glycine, serine, cysteine, glutamic acid, glutamine, aspartic acid and asparagine. Except for the glycine relation to glutamic acid and aspartic acid, all amino acids are related by chemical similarities in their side chains. Glycine not having a side chain would permit a more complex set of substitutions.  相似文献   

12.
The in vitro digestibility of rice glutelin and wheat glutenin was investigated with a view to assessing their nutritional qualities, using casein and bovine serum albumin (BSA) as references. The following hydrolytic processes were adopted: pepsin-pancreation digestion (a model system before intestinal absorption) and aminopeptidase-prolidase hydrolysis [a model system for the intestinal mucosa (membrane digestion) and after intestinal absorption (intracellular hydrolysis)]. The pepsin-pancreatin digests were first examined. The degree of amino acid released from the proteins was 30% (glutelin), 23% (glutenin), 24% (casein) and 30% (BSA). A similar release pattern of individual amino acids was observed for all the proteins. The amounts of large peptide fractions increased in the order: glutelin < glutenin < casein < BSA. Glutelin was highly digestible. Apart from containing high amounts of glutamic acid (glutamine), cystine and proline, the large peptide fractions of glutelin were also rich in threonine, glycine and isoleucine while those of glutenin were only rich in glycine. The aminopeptidase-prolidase digests were examined next. Glutelin was almost completely hydrolyzed to amino acid, except for a low release of cystine, suggesting that the amino acid residues constituting glutelin could be easily utilized as nutrients in the living tissues. The degree of amino acid released from the proteins was 97% (glutelin), 93% (glutenin), 90% (casein) and 79% (BSA).

The convenient application of these model systems for the assessment of the in vitro digestibility of food proteins have been discussed.  相似文献   

13.
Threonine content of brain decreases in young rats fed a threonine-limiting, low protein diet containing a supplement of small neutral amino acids (serine, glycine and alanine), which are competitors of threonine transport in other systems (Tews et al., 1977). Threonine transport by brain slices was inhibited more by a complex amino acid mixture resembling plasma from rats fed the small neutral amino acid supplement than by mixtures resembling plasma from control rats or from rats fed a supplement of large neutral amino acids. Greater inhibition was seen with mixtures containing only the small neutral amino acids than with mixtures containing only large neutral amino acids. On an equimolar basis, serine and alanine were the most inhibitory; large neutrals were moderately so; and glycine and lysine were without effect. Threonine transport was also strongly inhibited by α-amino-n-butyric acid and homoserine, less so by α-aminoisobutyric acid, and not at all by GABA. The complex amino acid mixtures strongly inhibited α-aminoisobutyric acid transport by brain or liver slices but, in contrast to effects in brain, the extent of the inhibition in liver was not much affected by altering the composition of the mixture. Tryptophan accumulation by brain slices was effectively inhibited by other large neutral amino acids in physiologically occurring concentrations. Threonine, or a mixture of serine, glycine and alanine only slightly inhibited tryptophan uptake; basic amino acids were without effect and histidine stimulated tryptophan transport slightly. These results support the conclusion that a diet-induced decrease in the concentration in brain of a specific amino acid may be related to increased inhibition of its transport into brain by increases in the concentrations of transport-related, plasma amino acids.  相似文献   

14.
This work was devoted to the study of the structure-affinity relationships in neutral amino acid transport by intestinal brush border of marine fish (Dicentrarchus labrax). The effects of the length of the side chain on kinetics of glycine, alanine, methionine and amino isobutyric acid were investigated. In the presence of K+ two components were characterized: one is saturable by increased substrate concentrations, whereas the other can be described by simple diffusion mechanism. Simple diffusion, a passive, non-saturable, Na+-independent route, contributes largely to the transport of methionine and to a much lesser extend to alanine, glycine or alphaaminoisobutyric acid uptakes. If a branched chain is present, as in the case of amino isobutyric acid, diffusion is low. A Na+-independent, saturable system has been fully characterized for methionine, but not for branched amino acids such as amino isobutyric acid. In the presence of Na+ saturable components were shown. Two distinct Na+-dependent pathways have been characterized for glycine uptake, with low and high affinities. For alanine and methionine only one Na+-dependent high affinity system exists with the same half-saturation concentration and the same maximum uptake at saturable concentrations. Glycine high affinity system has the same half-saturation concentration as methionine or alanine uptake, whereas maximum uptake is lower. The substitution of the hydrogen by a methyl group results in a severe decrease of uptake (aminoisobutyric acid). Mutual inhibition experiments indicate that the same carriers could be responsible for methionine and alanine uptakes and probably glycine Na+-dependent uptake. The influence of Na+ concentrations (100-1 mol·l-1) on amino acid uptake was examined. Glycine, alanine, methionine and amino isobutyric acid transport can be described by a hyperbolic function, with a saturation uptake which is highly increased for methionine. However, the half-saturation concentration does not seem to be strongly affected by the amino acid structure. The effect of Na+ concentration (25 and 100 mmol·l-1) on the kinetics of methionine uptake have been also examined. The maximum uptake of the saturable system clearly shows a typical relationship with concentration.Abbreviations [AA] amino acid concentration - AIB aminoisobutyric acid - [I] Inhibitor amino acid concentration - J i uptake in the presence of inhibitor - J o uptake without inhibitor - K d passive diffusion constant - K i inhibitor constant - K t concentration of test amino acid for half-maximal flux - MES 2[N-morpholino]ethanesulphonic acid - V max maximum uptake at saturable amino acid concentrations - V tot total amino acid uptake  相似文献   

15.
The effect of amino acids on nitrate transport was studied in Zea mays cell suspension cultures and in Zea mays excised roots. The inclusion of aspartic acid, arginine, glutamine and glycine (15mM total amino acids) in a complete cell-culture media containing 1.0 mM NO3 - strongly inhibited nitrate uptake and the induction of accelerated uptake rates. The nitrate uptake rate increased sharply once solution amino acid levels fell below detection limits. Glutamine alone inhibited induction in the cell suspension culture. Maize seedlings germinated and grown for 7 days in a 15 mM mixture of amino acids also had lower nitrate uptake rates than seedlings grown in 0.5 mM Ca(NO3)2 or 1 mM CaCl2. As amino acids are the end product of nitrate assimilation, the results suggest an end-product feed-back mechanism for the regulation of nitrate uptake.  相似文献   

16.
The dinitrophenyl-derivative of N-glycylglucosamine was isolated from partially hydrolyzed dinitrophenylated cell walls of Halococcus morrhuae CCM 859. To increase the yield of amino-terminal glycine residues, halococcal cell walls were treated with alkali or acid prior to dinitrophenylation. Authentic N-glycyl-glucosamine was used as a reference substance. A substitution of the amino group of glucosamine by an amino acid has so far not been found in any other wall of a pro- or eucaryotic cell. Since only 5% of the glycine residues reveal an unsubstituted carboxyl group within intact cell walls, glycine may play a role in connecting glycan strands through peptidic linkages between the amino group of glucosamine and the carboxyl group of an uronic acid or gulosaminuronic acid.Abbreviations DNP dinitrophenyl  相似文献   

17.
Oenococcus oeni has numerous amino acid requirements for growth and dipeptides could be important for its nutrition. In this paper the individual or combined effect of dipeptides on growth of O. oeni X2L in synthetic media deficient in one or more amino acids with L-malic acid was investigated. Utilization of dipeptides, glucose, and L-malic acid was also analyzed. Dipeptides were constituted by at least one essential amino acid for growth. Dipeptides containing two essential amino acids, except leucine, had a more favorable effect than free amino acids on the growth rate. Gly-Gly was consumed to a greater extent than Leu-Leu and a rapid exodus of glycine to the extracellular medium accompanied it. The microorganism could use glycine in exchange for other essential amino acids outside the cell, favoring growth. In the presence of Leu-Leu, the increase in glucose consumption rate could be related to the additional energy required for dipeptide uptake.  相似文献   

18.
Effect of Sugars and Amino Acids on Androgenesis of Cucumis sativus   总被引:3,自引:1,他引:2  
The effects of sugars (sucrose, maltose, glucose and fructose) and amino acids (glutamine, glycine, arginine, asparagine and cysteine) on embryogenesis and plantlet regeneration from cultured anthers of Cucumis sativus L. cv. Calypso and Green Long were studied. Type and concentration of sugar and amino acid influenced embryogenesis. Among the different sugars tested, sucrose was the best for embryo induction with an optimal concentration of 0.25 M. Maximum of 72 and 80 embryos per 60 anthers of Calypso and Green Long, respectively, were induced on embryo induction medium [B5 (Gamborg, Miller and Ojima (1968) Exp. Cell Res. 50: 151–158) supplemented with 2.0 μM 2,4-dichlorophenoxyacetic acid (2,4-D), 1.0 μM 6-benzyladenine (BA)] containing 0.25 M sucrose. The addition of amino acids to the embryo induction medium improved embryo yield with a combination of amino acids (glutamine, glycine, arginine, asparagine and cysteine of 1.0 mM each) giving the best response. Embryo differentiation was achieved on B5 medium supplemented with 0.25 μM of α-naphthaleneacetic acid (NAA), 0.25 μM kinetin (KN) and 0.09 M sucrose. Embryos were converted on B5 medium supplemented with abscisic acid (ABA) (10 μM) and 0.09 M sucrose. Embryos that developed on B5 medium supplemented with a combination of amino acids (glutamine, glycine, arginine, asparagine and cysteine of 1.0 mM each) exhibited the highest plantlet regeneration frequency.  相似文献   

19.
Clostridium histolyticum grew on glycine, arginine, or threonine as sole substrate. Arginine degradation preceded that of glycine and partially inhibited that of threonine when two amino acids were present. Each amino acid seemed to be individually catabolized, not by a Stickland type of reaction. Glycine fermentation required the presence of complex ingredients. Therefore, an effect of selenite on glycine catabolism could only be demonstrated after scavenging selenium contamination by preculturing Peptostreptococcus glycinophilus in that medium. C. acidiurici was not suited as selenium accumulating organism as C. histolyticum was inhibited by the residual uric acid. Arginine catabolism was unaffected by seleniuum depriviation. The labelling pattern obtained in acetate after incubation of C. histolyticum with [1-14C]- or [2-14C]glycine strongly indicated the metabolism of glycine via the glycine reductase pathway.  相似文献   

20.
Intestinal and hepatic ornithine decarboxylase (ODC) activities increased to a peak 4 h after administration of a diet containing casein or an amino acid mixture simulating that of casein to rats starved for 12 h. All amino acids except cysteine with a two or three carbon skeleton, including those with a D-configuration, and alpha-amino-isobutyric acid (AIB) strongly induced intestinal ODC when given in the diet or administered intragastrically. Amino acids with a four carbon skeleton were far less effective as inducers and other amino acids did not induce intestinal ODC at all. The amino acids that induced hepatic ODC showed no particular structural characteristics: glycine and cysteine were very effective, threonine, tryptophan, methionine, and phenylalanine were less effective, and serine, valine, isoleucine, and histidine were only slightly effective. Elevation of ODC activity after amino acid administration was not due to stabilization of the enzyme protein with the amino acids. Intestinal ODC was induced by intragastric but not intraperitoneal injection of glycine, although these treatments resulted in similar increases in the tissue concentration of glycine. On the contrary, hepatic ODC was induced by glycine regardless of the administration route. Intestinal ODC was also induced only in the segment of the intestine perfused with a solution of an amino acid with which the activity increased in the feeding experiment. These results suggest that the accumulation of an amino acid per se is not a trigger for induction of intestinal ODC and that an amino acid must act on the mucosal surface to induce the enzyme.  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