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 93 毫秒
1.
All animals, except for the placental mammals, conjugate their bile acids exclusively with taurine. However, in certain of the placental mammals, glycine conjugates are also found. The basis for the appearance of glycine conjugation among the placental mammals was investigated. The reaction of choloyl-CoA with glycine and taurine, as catalysed by the soluble fraction from guinea-pig liver, had a high affinity for taurine and a poor affinity for glycine. The predominant synthesis of glycine conjugates in the guinea pig can be related to the fact that guinea-pig liver contains an unusually low concentration of taurine and a high concentration of glycine. Rabbits make exclusively glycine conjugates and their livers also contain low concentrations of taurine. However, the biochemical basis for their glycine conjugation is more straightforward than in the guinea pig in that the soluble fraction from rabbit liver has a high affinity for glycine and a poor affinity for taurine. Alternative-substrate-inhibition studies with glycine and taurine in soluble fractions from guinea-pig and rabbit liver revealed that glycine and taurine were mutually inhibitory. This suggests that there is only one enzyme for glycine and taurine conjugation in these tissues. The soluble fractions from bovine liver and human liver also made both glycine and taurine conjugates and evidence is presented that suggests that there is only one enzyme in these tissues too. Even the rat, which excretes mostly taurine conjugates, could make both glycine and taurine conjugates in vitro. However, in contrast with all of the placental mammals studied, the supernatant fraction from liver of the chicken, and other non-mammals, could not make glycine conjugates even in the presence of very high concentrations of glycine.  相似文献   

2.
The genes from the extreme halophile Ecto-thiorhodospira halochloris encoding the biosynthesis of glycine betaine from glycine were cloned into Escherichia coli. The accumulation of glycine betaine and its effect on osmotolerance of the cells were studied. In mineral medium with NaCl concentrations from 0.15 to 0.5 M, the accumulation of both endogenously synthesized and exogenously provided glycine betaine stimulated the growth of E. coli. The intracellular levels of glycine betaine and the cellular yields were clearly higher for cells receiving glycine betaine exogenously than for cells synthesizing it. The lower level of glycine betaine accumulation in cells synthesizing it is most likely a consequence of the limited availability of precursors (e.g. S-adenosylmethionine) rather than the result of a low expression level of the genes. Glycine betaine also stimulated the growth of E. coli and decreased acetate formation in mineral medium with high sucrose concentrations (up to 200 g.l(-1)).  相似文献   

3.
Experiments at 298.15 K have been performed to measure the density, velocity of sound and refractive index in three water+glycine+electrolyte systems. The electrolytes studied were KCl, KNO3 and NaNO3. The values of apparent molar volume and isentropic compressibility of glycine in aqueous electrolyte solutions were calculated from the measured data. The results obtained in this study and those reported previously for water+glycine+NaCl system have been comparatively studied. The results show that the nature of both the cation and the anion of an electrolyte influence the behaviour of glycine in aqueous solutions. For all four electrolytes studied, the comparison shows a positive volume transfer for glycine from an electrolyte solution to a more concentrated solution of the same electrolyte. The results also show a negative apparent isentropic compressibility for glycine in the presence of the electrolytes studied. These effects indicate that the volume of a glycine molecule is larger in solutions with higher electrolyte concentration and the water molecules around the glycine molecules are less compressible than the water molecules in the bulk solution. These effects were attributed to the doubly charged behaviour of glycine and to the formation of physically bonded ion-pairs between the charged groups of glycine and the cation and the anion of the electrolyte.  相似文献   

4.
Transport Is the Primary Determinant of Glycine Content in Retinal Neurons   总被引:1,自引:0,他引:1  
Abstract: This study demonstrates that in mammalian and nonmammalian species it is possible to deplete selectively and reversibly retinal glycinergic neurons of their content of glycine by exposure to sarcosine, a competitive inhibitor of glycine transporter 1 (glyt-1). This observation was used as a tool to test the hypothesis that uptake of glycine rather than de novo synthesis is the main determinant of glycine content in retinal neurons. Isolated retinae were depleted of immunocytochemically detectable pools of glycine. Thereafter retinae were exposed either to physiological medium containing glycine or to medium lacking glycine but containing precursors for the synthesis of glycine. Retinae exposed to glycine-containing medium rapidly recovered their content of glycine, whereas retinae exposed to medium lacking glycine but containing serine, a substrate for synthesis of glycine, showed only a slow recovery of immunoreactivity for glycine in a few amacrine cells. These data indicate that uptake of glycine is the primary determinant of glycine content in most retinal glycinergic neurons. The origins of the extracellular pools of glycine remain to be identified; however, it is suggested that such glycine may be derived from the vitreous humor and that in turn this glycine may be derived from the peripheral circulation.  相似文献   

5.
R M Kappes  B Kempf    E Bremer 《Journal of bacteriology》1996,178(17):5071-5079
The accumulation of the osmoprotectant glycine betaine from exogenous sources provides a high degree of osmotic tolerance to Bacillus subtilis. We have identified, through functional complementation of an Escherichia coli mutant defective in glycine betaine uptake, a new glycine betaine transport system from B. subtilis. The DNA sequence of a 2,310-bp segment of the cloned region revealed a single gene (opuD) whose product (OpuD) was essential for glycine betaine uptake and osmoprotection in E. coli. The opuD gene encodes a hydrophobic 56.13-kDa protein (512 amino acid residues). OpuD shows a significant degree of sequence identity to the choline transporter BetT and the carnitine transporter CaiT from E. coli and a BetT-like protein from Haemophilus influenzae. These membrane proteins form a family of transporters involved in the uptake of trimethylammonium compounds. The OpuD-mediated glycine betaine transport activity in B. subtilis is controlled by the environmental osmolarity. High osmolarity stimulates de novo synthesis of OpuD and activates preexisting OpuD proteins to achieve maximal glycine betaine uptake activity. An opuD mutant was constructed by marker replacement, and the OpuD-mediated glycine betaine uptake activity was compared with that of the previously identified multicomponent OpuA and OpuC (ProU) glycine betaine uptake systems. In addition, a set of mutants was constructed, each of which synthesized only one of the three glycine betaine uptake systems. These mutants were used to determine the kinetic parameters for glycine betaine transport through OpuA, OpuC, and OpuD. Each of these uptake systems shows high substrate affinity, with Km values in the low micromolar range, which should allow B. subtilis to efficiently acquire the osmoprotectant from the environment. The systems differed in their contribution to the overall glycine betaine accumulation and osmoprotection. A triple opuA, opuC, and opuD mutant strain was isolated, and it showed no glycine betaine uptake activity, demonstrating that three transport systems for this osmoprotectant operate in B. subtilis.  相似文献   

6.
In a previous paper, we pointed out that the capability to synthesize glycine from serine is constrained by the stoichiometry of the glycine hydroxymethyltransferase reaction, which limits the amount of glycine produced to be no more than equimolar with the amount of C 1 units produced. This constraint predicts a shortage of available glycine if there are no adequate compensating processes. Here, we test this prediction by comparing all reported fluxes for the production and consumption of glycine in a human adult. Detailed assessment of all possible sources of glycine shows that synthesis from serine accounts for more than 85% of the total, and that the amount of glycine available from synthesis, about 3 g/day, together with that available from the diet, in the range 1.5–3.0 g/day, may fall significantly short of the amount needed for all metabolic uses, including collagen synthesis by about 10 g per day for a 70 kg human. This result supports earlier suggestions in the literature that glycine is a semi-essential amino acid and that it should be taken as a nutritional supplement to guarantee a healthy metabolism.  相似文献   

7.
The synthesis and characterization of a new photolabile precursor of glycine (caged glycine) is described. The alpha-carboxyl group of glycine is covalently coupled to the alpha-carboxy-2-nitrobenzyl (alphaCNB) protecting group. Photolysis of the caged glycine with UV light produces free glycine. At 308 nm, the compound photolyzes with a quantum yield of 0.38. The absorption spectrum and the pH dependence of a transient absorption produced after laser-flash illumination are typical for aci-nitro intermediates of alphaCNB-protected compounds. The time constant for the major component of the aci-nitro intermediate decay ( approximately 84% of the total aci-nitro absorbance) was determined to be 7 micros at physiological pH. A minor component ( approximately 16%) decays with a rate constant of 170 micros. The compound does not activate or inhibit the alpha(1)-homomeric glycine receptor transiently expressed in HEK293 cells. After photolysis with a 10 ns pulse of 325 nm laser light, the glycine released from the caged compound activates glycine-mediated whole-cell currents in the same cells. The rise of these currents can be measured in a time-resolved fashion and occurs on a millisecond to sub-millisecond time scale. It can be described with a single-exponential function over >85% of the total current. The rate constant of the current rise is about 2 orders of magnitude slower than the rate constant of caged glycine photolysis. Thermal hydrolysis of the alphaCNB-caged glycine takes place with a half-life of 15.6 h at physiological pH. The new caged glycine is the first in a series of photoprotected glycine derivatives that has the required properties for use with chemical kinetic methods for investigation of glycine-activated cell surface receptors. Photolysis is rapid and efficient with respect to the receptor reactions to be studied; hydrolysis in aqueous solution is sufficiently slow, and the compound is biologically inert. It will, therefore, be a useful tool for investigation of the processes leading to channel opening of glycine receptor channels and the effects of mutations of the glycine receptor and of inhibitors on these processes.  相似文献   

8.
Glycine betaine is an osmoprotectant found in many organisms, including bacteria and higher plants. The bacterium Escherichia coli produces glycine betaine by a two-step pathway where choline dehydrogenase (CDH), encoded by betA, oxidizes choline to betaine aldehyde which is further oxidized to glycine betaine by the same enzyme. The second step, conversion of betaine aldehyde into glycine betaine, can also be performed by the second enzyme in the pathway, betaine aldehyde dehydrogenase (BADH), encoded by betB. Transformation of tobacco (Nicotiana tabacum), a species not accumulating glycine betaine, with the E. coli genes for glycine betaine biosynthesis, resulted in transgenic plants accumulating glycine betaine. Plants producing CDH were found to accumulate glycine betaine as did F1 progeny from crosses between CDH- and BADH-producing lines. Plants producing both CDH and BADH generally accumulated higher amounts of glycine betaine than plants producing CDH alone, as determined by 1H NMR analysis. Transgenic tobacco lines accumulating glycine betaine exhibited increased tolerance to salt stress as measured by biomass production of greenhouse-grown intact plants. Furthermore, experiments conducted with leaf discs from glycine betaine-accumulating plants indicated enhanced recovery from photoinhibition caused by high light and salt stress as well as improved tolerance to photoinhibition under low temperature conditions. In conclusion, introduction of glycine betaine production into tobacco is associated with increased stress tolerance probably partly due to improved protection of the photosynthetic apparatus.  相似文献   

9.
Experiments have been performed at 298.15 K to measure the density, sound velocity and refractive index of glycine in aqueous solutions of NaCl over a wide range of both glycine and NaCl concentrations. The values of apparent molar volume and isentropic compressibility of glycine were calculated from the measured data. The results show a positive transfer volume of glycine from an NaCl solution to a more concentrated NaCl solution. This indicates that the size of a glycine molecule is larger in a solution with higher NaCl concentration. The negative values of apparent isentropic compressibility imply that the water molecules around the glycine molecules are less compressible than the water molecules in the bulk solution. These effects are attributed to the doubly charged behaviour of glycine and to the formation of physically bonded ion-pairs between the charged groups of glycine and sodium and chloride ions. The formation of ion-pairs, whose extents of binding reactions depend on the concentrations of both NaCl and glycine, alter the hydration number of glycine. This also explains the reason for the increase in the size of glycine with an increase in the NaCl concentration. A model based on the Pitzer formalism has been developed to correlate the activity coefficient, apparent molar volume and isentropic compressibility of glycine in aqueous solutions of NaCl. The results show that the model can accurately correlate the interactions in aqueous solutions of glycine and NaCl.  相似文献   

10.
D. A. Sinclair  I. W. Dawes 《Genetics》1995,140(4):1213-1222
Saccharomyces cerevisiae can grow on glycine as sole nitrogen source and can convert glycine to serine via the reaction catalyzed by the glycine decarboxylase multienzyme complex (GDC). Yeast strains with mutations in the single gene for lipoamide dehydrogenase (lpd1) lack GDC activity, as well as the other three 2-oxoacid dehydrogenases dependent on this enzyme. The LPD1 gene product is also required for cells to utilize glycine as sole nitrogen source. The effect of mutations in LPD1 (L-subunit of GDC), SER1 (synthesis of serine from 3-phosphoglycerate), ADE3 (cytoplasmic synthesis of one-carbon units for the serine synthesis from glycine), and all combinations of each has been determined. The results were used to devise methods for isolating mutants affected either in the generation of one-carbon units from glycine (via GDC) or subsequent steps in serine biosynthesis. The mutants fell into six complementation groups (gsd1-6 for defects in conversion of glycine to serine). Representatives from three complementation groups were also unable to grow on glycine as sole nitrogen source (gsd1-3). Assays of the rate of glycine uptake and decarboxylation have provided insights into the nature of the mutations.  相似文献   

11.
The specific radioactivity of urinary hippurate glycine was determined after injecting guinea pigs with benzoate and either dl-[2-(14)C]glutamate or dl-[5-(14)C]glutamate. The isotope dilution factor for the formation of [(14)C]glycine was significantly greater (30%) with C-2 labelled glutamate. With either form of labelled glutamate the hippurate glycine was largely carboxyl-group labelled. The observations suggest a route for the incorporation of glutamate carbon into glycine that involves C-5 but not C-2. A hypothesis for glycine biosynthesis from l-glutamate is advanced, consistent with these findings, that includes conversion of l-glutamate to 4-hydroxy-2-oxoglutarate, the scission of the latter to glyoxylate and pyruvate, and the formation of glycine by transamination.  相似文献   

12.
The aim of the study was to determine the relative bioavailability of zinc gluconate stabilized with glycine in a Petit Suisse cheese from an infant dessert. Weight gain and bone zinc content were the nutritional responses evaluated for the diets of different zinc content: 2 ppm (basal) and 5, 10, and 30 ppm from zinc gluconate stabilized with glycine and zinc sulfate. Nonlinear regression analysis of the fitted curves for weight gain determined a relative zinc bioavailability of 100% for the Y max ratio and 96% for Y max/t 1/2 ratio for zinc gluconate stabilized with glycine (R 2=0.7996 for zinc sulfate and 0.8665 for zinc gluconate stabilized with glycine). The slope ratio analysis from linear regression of femur zinc determined a relative zinc bioavailability of 93% for zinc gluconate stabilized with glycine (R 2=0.8693 for zinc sulfate and 0.8307 for zinc gluconate stabilized with glycine). Zinc gluconate stabilized with glycine has similar bioavailability as zinc sulfate in a Petit Suisse cheese nutritional matrix, with the advantage that the stabilized compound does not modify the sensorial characteristics of the fortified cheese.  相似文献   

13.
Mitochondria isolated from pea leaves possess a glycine transporter that is capable of moving glycine from the cytosol into the matrix, the site of glycine decar?ylase. The carrier was inhibited by mersalyl, p-chloromercuribenzoate, and the glycine analogues, glycine hydroxamate and aminoacetonitrile. Glycine uptake was dependent on the transmembrane pH gradient and was inhibited by uncouplers and electron transport inhibitors. Glycine transport was not, however, inhibited by the glycine decar?ylase inhibitor, arsenite. This transporter is responsible for the movement of glycine into the mitochondria and provides an important step in photorespiration.  相似文献   

14.
Two different aminotransferases, that have glyoxylate as the amino acceptor, have specific activities of 1 to 2 mumol . min-1 . mg of protein-1 in the isolated peroxisomal fraction from spinach leaves. Their properties were evaluated after separation on a hydroxylapatite column. Both enzymes had a Km for glyoxylate of 0.15 mM and an amino acid Km of 2 to 3 mM. Reactions proceeded by a Ping Pong Bi Bi mechanism. Serine:glyoxylate aminotransferase was relatively specific for both substrates and could only be slightly reversed with 100 mM glycine, although the Ki of glycine was 33 mM. The glutamate:glyoxylate amino-transferase protein was equally active in catalyzing an alanine:glyoxylate aminotransferase reaction, but the reverse reactions with 100 mM glycine were hardly measureable, although the Ki (glycine) was 8.7 mM. Protection against hydroxylamine inhibition from reaction with pyridoxal phosphate was used to investigate the specificity of amino acid binding. Substrate amino acids protected at about the same concentration as their Km, while glycine protected at its Ki concentration. Thus, the nearly irreversible catalysis with glycine is not due to a failure to bind glycine. The significance of a peroxisomal alanine:glyoxylate aminotransferase activity has not been incorporated into schemes for the oxidative photosynthetic carbon cycle.  相似文献   

15.
[3H]Glycine binding and glycine modulation of [3H]MK-801 binding have been used to study the glycine allosteric site associated with the N-methyl-D-aspartate receptor complex in postmortem human brain. The effect of glycine on [3H]MK-801 binding appeared sensitive to duration of terminal coma, and possibly postmortem delay. Thirty percent of the binding occurred in a subfraction of brain tissue and did not show enhancement by glycine and glutamic acid. [3H]Glycine binding to a subfraction free from this component was studied and showed high specific binding. KD and Bmax values showed considerable intersubject variability which did not appear to be due to demographic features or to tissue content of amino acids with an affinity for this site. The pharmacological characteristics of binding in this subfraction and a correlation between Bmax values and the maximal enhancement of [3H]MK-801 binding by glycine are consistent with [3H]glycine binding occurring to an N-methyl-D-aspartate receptor complex associated site. Further support for this is provided by a significantly lower Bmax value for [3H]glycine binding in subjects with Alzheimer's disease and reduced glycine enhancement of [3H]MK-801 binding. However, the effect of perimortem factors makes it difficult to confidently attribute this solely to a disease-related change in the receptor. The possible role of the glycine allosteric site in the treatment of neuropsychiatric disorders is discussed.  相似文献   

16.
Glucagon stimulates 14CO2 production from [1-14C] glycine by isolated rat hepatocytes. Maximal stimulation (70%) of decarboxylation of glycine by hepatocytes was achieved when the concentration of glucagon in the medium reached 10 nM; half-maximal stimulation occurred at a concentration of about 2 nM. A lag period of 10 min was observed before the stimulation could be measured. Inclusion of beta-hydroxybutyrate (10 mM) or acetoacetate (10 mM) did not affect the magnitude of stimulation suggesting that the effects of glucagon were independent of mitochondrial redox state. Glucagon did not affect either the concentration or specific activity of intracellular glycine, thus excluding the possibilities that altered concentration or specific activity of intracellular glycine contributes to the observed stimulation. The stimulation of decarboxylation of glycine by glucagon was further studied by monitoring 14CO2 production from [1-14C]glycine by mitochondria isolated from rats previously injected with glucagon. Glycine decarboxylation was significantly stimulated in the mitochondria isolated from the glucagon-injected rats. We suggest that glucagon is a major regulator of hepatic glycine metabolism through the glycine cleavage enzyme system and may be responsible for the increased hepatic glycine removal observed in animals fed high-protein diets.  相似文献   

17.
The efflux and exchange of glycine were studied in plasma membrane vesicles isolated from cultured glioblastoma cells. The mechanism of glycine translocation has been probed by comparing the ion dependence of net efflux to that of exchange. Dilution-induced efflux requires the simultaneous presence of internal sodium and chloride, while influx is dependent on the presence of these two ions on the outside (Zafra, F. and Giménez, C. (1986) Brain Res. 397, 108-116). Glycine efflux from the membrane vesicles is stimulated by external glycine, this exchange being dependent on external sodium, but not on external chloride. The parallelism observed in influx and efflux processes suggests that glycine is translocated in both directions across the membrane, probably by interacting with the carrier. To account for all the observed effects of external ions, glycine concentrations and membrane potential on glycine influx and efflux, a kinetic model of the Na+/Cl-/glycine cotransport system is discussed.  相似文献   

18.
Mortality associated with endotoxin shock is likely mediated by Kupffer cells, alveolar macrophages, and circulating neutrophils. Acute dietary glycine prevents mortality and blunts increases in serum tumor necrosis factor-alpha (TNF-alpha) following endotoxin in rats. Furthermore, acute glycine blunts activation of Kupffer cells, alveolar macrophages, and neutrophils by activating a glycine-gated chloride channel. However, in neuronal tissue, glycine rapidly downregulates chloride channel function. Therefore, the long-term effects of a glycine-containing diet on survival following endotoxin shock were investigated. Dietary glycine for 4 wk improved survival after endotoxin but did not improve liver pathology, decrease serum alanine transaminase, or effect TNF-alpha levels compared with animals fed control diet. Interestingly, dietary glycine largely prevented inflammation and injury in the lung following endotoxin. Surprisingly, Kupffer cells from animals fed glycine for 4 wk were no longer inactivated by glycine in vitro; however, isolated alveolar macrophages and neutrophils from the same animals were sensitive to glycine. These data are consistent with the hypothesis that glycine downregulates chloride channels on Kupffer cells but not on alveolar macrophages or neutrophils. Importantly, glycine diet for 4 wk protected against lung inflammation due to endotoxin. Chronic glycine improves survival by unknown mechanisms, but reduction of lung inflammation is likely involved.  相似文献   

19.
High doses of glycine have been reported to improve negative schizophrenic symptoms, suggesting that ingested glycine activates glutamatergic transmission via N-methyl-d-aspartate (NMDA) receptors. However, the pharmacokinetics of administered glycine in the brain has not been evaluated. In the present study, the time- and dose-dependent distributions of administered glycine were investigated from a pharmacokinetic viewpoint. Whole-body autoradiography of radiolabeled glycine was performed, and time–concentration curves for glycine and serine in plasma, cerebrospinal fluid (CSF), and brain tissues were obtained. Furthermore, pharmacokinetic parameters were calculated. For a more detailed analysis, the amount of glycine uptake in the brain was evaluated using the brain uptake index method. Radiolabeled glycine was distributed among periventricular organs in the brain. Oral administration of 2?g/kg of glycine significantly elevated the CSF glycine concentration above the ED50 value for NMDA receptors. The glycine levels in CSF were 100 times lower than those in plasma. Glycine levels were elevated in brain tissue, but with a slower time-course than in CSF. Serine, a major metabolite of glycine, was elevated in plasma, CSF, and brain tissue. Glycine uptake in brain tissue increased in a dose-dependent manner. Time–concentration curves revealed that glycine was most likely transported via the blood–CSF barrier and activated NMDA receptors adjacent to the ventricles. The pharmacokinetic analysis and the brain uptake index for glycine suggested that glycine was transported into brain tissue by passive diffusion. These results provide further insight into the potential therapeutic applications of glycine.  相似文献   

20.
The kinetic properties of glycine oxidase from Bacillus subtilis were investigated using glycine, sarcosine, and d-proline as substrate. The turnover numbers at saturating substrate and oxygen concentrations were 4.0 s(-1), 4.2 s(-1), and 3.5 s(-1), respectively, with glycine, sarcosine, and D-proline as substrate. Glycine oxidase was converted to a two-electron reduced form upon anaerobic reduction with the individual substrates and its reductive half-reaction was demonstrated to be reversible. The rates of flavin reduction extrapolated to saturating substrate concentration, and under anaerobic conditions, were 166 s(-1), 170 s(-1), and 26 s(-1), respectively, with glycine, sarcosine, and D-proline as substrate. The rate of reoxidation of reduced glycine oxidase with oxygen in the absence of product (extrapolated rate approximately 3 x 10(4) M(-1) x s(-1)) was too slow to account for catalysis and thus reoxidation started from the reduced enzyme:imino acid complex. The kinetic data are compatible with a ternary complex sequential mechanism in which the rate of product dissociation from the reoxidized enzyme form represents the rate-limiting step. Although glycine oxidase and D-amino acid oxidase differ in substrate specificity and amino acid sequence, the kinetic mechanism of glycine oxidase is similar to that determined for mammalian D-amino acid oxidase on neutral D-amino acids, further supporting a close similarity between these two amine oxidases.  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