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.
Origin of the sarcosine molecules of actinomycins   总被引:1,自引:0,他引:1       下载免费PDF全文
1. Streptomyces V–187 produces on minimal medium a mixture composed mainly of actinomycin C1 (actinomycin D) and actinomycin A1 (actinomycin I). If sarcosine is added to the medium, the micro-organism produces, in addition to actinomycins C1 and A1, actinomycin F8 (actinomycin II) and actinomycin F9 (actinomycin (III), characterized by the substitution by sarcosine of one or both the proline molecules present in actinomycin C1. 2. Exogenous sarcosine seems to be incorporated as such by Streptomyces V–187 only in the sarcosine molecule(s) that replace proline in the actinomycins of the F group, whereas, for the synthesis of the other sarcosine molecules, the amino acid is first demethylated to glycine. 3. The incorporation of sarcosine and glycine into actinomycin by Streptomyces antibioticus appears to follow a similar pattern, except that a portion of the methyl group produced in the degradation of sarcosine is utilized as a source of the methyl groups of the antibiotic. This explains the previously reported lack of cross-dilution between glycine and sarcosine observed in the incorporation of these amino acids into actinomycin.  相似文献   

2.
3.
Protoplasts actively synthesizing actinomycins have been prepared from Streptomyces, antibioticus. They showed an absolute requirement for the presence of oxygen, galactose, and alkaline earth ions. Sucrose was most efficient as an osmotic stabilizer. However, in air-saturated buffer the protoplasts seemed to be slightly inhibited in their metabolism. This is expressed by the appearance of 4-methyl-3-hydroxyanthranilic acid and the inability to utilize [1?14C]sarcosine for actinomycin synthesis. Evidence has been obtained that sarcosine and N-methyl-l-valine are not free precursors of the peptide-bound N-methyl-amino acids. It is further demonstrated that synthesis of actinomycin IV and actinomycin V differ from each other with respect to their different susceptibilities against the changings in the physiological environment of the protoplasts. Actinomycin synthesis is severely reduced when protoplasts are incubated in the presence of 10?3, m methionine.  相似文献   

4.
Actinomycin synthesis by washed mycelia of Streptomyces antibioticus has been conducted in the presence of 3-hydroxy-4-methylanthranilate-(carboxyl-14C). Incorporation of this compound into actinomycins has been observed, which constitutes further evidence that 3-hydroxy-4-methylanthranilate is an intermediate in actinomycin biosynthesis. The position of the incorporated label has been determined to be within the actinomycin chromophore, and the label appears to be equally distributed between both halves of the chromophore. Incidental to these findings was the observation that the 14C-labeled actinomycins were subject to rapid reabsorption by the organism with actinomycin V taken up preferentially to actinomycin IV.  相似文献   

5.
In addition to actinomycins D, X2 and X, Streptomyces antibioticus No. B-1625 produces minor acidic actinomycin congeners (FA-components). To increase the production of the FA-components, improvement of medium constituents was attempted for both chemically defined and complex media. Addition of trace metals, especially FeSO4, increased FA-components production and, moreover, the addition of sarcosine was found to increase the production of a selected component, B-1625 FA. Finally, a complex medium, consisting of starch 3.0, Polypepton 0.1, meat extract 0.1, corn steep liquor 3.0, NaCl 0.3, CaCO3 0.3, sarcosine 0.1 and FeSO4 · 7H2O 0.05%, was developed for the increased production of FA-components, in particular, the selected component of B-1625 FA.  相似文献   

6.
A new actinomycin was isolated from a mixture of actinomycins formed by Actinomyces sp. No. 2, an organism producing auranthin, an actinomycetous antibiotic. The peptide chains of the new actinomycin contain such amino acids as threonine, valine, proline and sarcosine in a ratio of 2 : 4 : 2 : 2. N-Methyl-valine characteristic of all actinomycins is replaced in position 5 of both pentapeptide chains of the new actinomycin by valine. The new actinomycin is actinomycin D undermethylated in position 5 by valine. When the growing culture of Actinomyces olivobrunneus producing actinomycin D was exposed to sulfadimesine, an inhibitor of biological methylation, production of actinomycin D0 (sarcosine replaced by glycine in one of the pentapeptide chains) markedly increased, which indicated impairment of the glycine residue methylation. Still, no impairment of the valine residue methylation in position 5 of the pentapeptide chains was observed an no actinomycin with N-methyl-valine replaced by valine was formed.  相似文献   

7.
Colwellia is a genus of mostly psychrophilic halophilic Gammaproteobacteria frequently isolated from polar marine sediments and sea ice. In exploring the capacity of Colwellia psychrerythraea 34H to survive and grow in the liquid brines of sea ice, we detected a duplicated 37 kbp genomic island in its genome based on the abnormally high G + C content. This island contains an operon encoding for heterotetrameric sarcosine oxidase and is located adjacent to several genes used in the serial demethylation of glycine betaine, a compatible solute commonly used for osmoregulation, to dimethylglycine, sarcosine, and glycine. Molecular clock inferences of important events in the adaptation of C. psychrerythraea 34H to compatible solute utilization reflect the geological evolution of the polar regions. Validating genomic predictions, C. psychrerythraea 34H was shown to grow on defined media containing either choline or glycine betaine, and on a medium with sarcosine as the sole organic source of carbon and nitrogen. Growth by 8 of 9 tested Colwellia species on a newly developed sarcosine-based defined medium suggested that the ability to catabolize glycine betaine (the catabolic precursor of sarcosine) is likely widespread in the genus Colwellia. This capacity likely provides a selective advantage to Colwellia species in cold, salty environments like sea ice, and may have contributed to the ability of Colwellia to invade these extreme niches.  相似文献   

8.
Relaxases act as DNA selection sieves in conjugative plasmid transfer. Most plasmid relaxases belong to the HUH endonuclease family. TrwC, the relaxase of plasmid R388, is the prototype of the HUH relaxase family, which also includes TraI of plasmid F. In this article we demonstrate that TrwC processes its target nic-site by means of a highly secure double lock and key mechanism. It is controlled both by TrwC–DNA intermolecular interactions and by intramolecular DNA interactions between several nic nucleotides. The sequence specificity map of the interaction between TrwC and DNA was determined by systematic mutagenesis using degenerate oligonucleotide libraries. The specificity map reveals the minimal nic sequence requirements for R388-based conjugation. Some nic-site sequence variants were still able to form the U-turn shape at the nic-site necessary for TrwC processing, as observed by X-ray crystallography. Moreover, purified TrwC relaxase effectively cleaved ssDNA as well as dsDNA substrates containing these mutant sequences. Since TrwC is able to catalyze DNA integration in a nic-site-containing DNA molecule, characterization of nic-site functionally active sequence variants should improve the search quality of potential target sequences for relaxase-mediated integration in any target genome.  相似文献   

9.
The effects of sarcosine on the processes driving prostate cancer (PCa) development remain still unclear. Herein, we show that a supplementation of metastatic PCa cells (androgen independent PC-3 and androgen dependent LNCaP) with sarcosine stimulates cells proliferation in vitro. Similar stimulatory effects were observed also in PCa murine xenografts, in which sarcosine treatment induced a tumor growth and significantly reduced weight of treated mice (p < 0.05). Determination of sarcosine metabolism-related amino acids and enzymes within tumor mass revealed significantly increased glycine, serine and sarcosine concentrations after treatment accompanied with the increased amount of sarcosine dehydrogenase. In both tumor types, dimethylglycine and glycine-N-methyltransferase were affected slightly, only. To identify the effects of sarcosine treatment on the expression of genes involved in any aspect of cancer development, we further investigated expression profiles of excised tumors using cDNA electrochemical microarray followed by validation using the semi-quantitative PCR. We found 25 differentially expressed genes in PC-3, 32 in LNCaP tumors and 18 overlapping genes. Bioinformatical processing revealed strong sarcosine-related induction of genes involved particularly in a cell cycle progression. Our exploratory study demonstrates that sarcosine stimulates PCa metastatic cells irrespectively of androgen dependence. Overall, the obtained data provides valuable information towards understanding the role of sarcosine in PCa progression and adds another piece of puzzle into a picture of sarcosine oncometabolic potential.  相似文献   

10.
The gene encoding sarcosine oxidase from Arthrobacter sp. TE1826 (soxA) was cloned in Escherichia coli by a convenient plate assay. It was located within a 1.7-kbp PstI-EcoRI fragment of the recombinant plasmid pSAOEP3. The purified sarcosine oxidase from the recombinant strain was found to be the same as that from the parental strain. The DNA sequence of soxA was determined, and an open reading frame composed of 389 amino acid residues was found. By Edman degradation of the enzyme, it was revealed that the amino-terminal amino acid (methionine) was eliminated in the parental strain and E. coli. The molecular weight (43,249) of the enzyme was consistent with the result from SDS-polyacrylamide gel electrophoresis. The FAD-binding site was found in the amino-terminal region of sarcosine oxidase by a homology search. The soxA gene was subcloned on a shuttle vector, pHY300PLK, and was expressed in both E. coli and Bacillus subtillis in the absence of an inducer, although the enzyme was induced with sarcosine in the parental strain.  相似文献   

11.
Sarcosinemia is an autosomal recessive metabolic trait manifested by relatively high concentrations of sarcosine in blood and urine. Sarcosine is a key intermediate in 1-carbon metabolism and under normal circumstances is converted to glycine by the enzyme sarcosine dehydrogenase. We encountered six families from two different descents (French and Arab), each with at least one individual with elevated levels of sarcosine in blood and urine. Using the “candidate gene approach” we sequenced the gene encoding sarcosine dehydrogenase (SARDH), which plays an important role in the conversion of sarcosine to glycine, and found four different mutations (P287L, V71F, R723X, R514X) in three patients. In an additional patient, we found a uniparental disomy in the region of SARDH gene. In two other patients, we did not find any mutations in this gene. We have shown for the first time that mutations in the SARDH gene are associated with sarcosinemia. In addition, our results indicate that other genes are most probably involved in the pathogenesis of this condition.  相似文献   

12.
《Insect Biochemistry》1987,17(1):17-20
Sarcosine and methionine sulfoxide were investigated in several wild or laboratory-reared symbiotic and aposymbiotic strains of Sitophilus oryzae and S. zeamais. The amino acid composition of fourth-instar larvae indicated that a high level in sarcosine found together with a low level of methionine sulfoxide were biochemical characteristics of the aposymbiotic state in this genus. Nutritional experiments demonstrated that the synthesis of these two amino acids depended on dietary precursors. Since sarcosine and methionine sulfoxide are both methionine derivatives, it is therefore suggested that methionine metabolism in Sitophilus larvae might differ according to the presence or the absence of the symbiotic bacteria.  相似文献   

13.
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   

14.
The obligate anaerobe Eubacterium acidaminophilum metabolized the glycine derivatives sarcosine (N-monomethyl glycine) and betaine (N-trimethyl glycine) only by reduction in a reaction analogous to glycine reductase. Using formate as electron donor, sarcosine and betaine were stoichiometrically reduced to acetate and methylamine or trimethylamine, respectively. The N-methyl groups of the cosubstrates or of the amines produced were not transformed to CO2 or acetate. Under optimum conditions (formate/acceptor ratio of 1 to 1.2, 34°C, pH 7.3) the doubling times were 4.2 h on formate/sarcosine and 3.6 h on formate/betaine. The molar growth yields were 8.15 and 8.5 g dry cell mass per mol sarcosine and betaine, respectively. The assays for sarcosine reductase and betaine reductase were optimized in cell extracts; NADPH was preferred as physiological electron donor compared to NADH, dithioerythritol was used as artificial donor; no requirements for AMP and ADP could be detected. Growth experiments mostly revealed diauxic substrate utilization pattern using different combinations of glycine, sarcosine, and betaine (plus formate) and inocula from different precultures. Glycine was always utilized first, what coincided with the presence of glycine reductase activity under all growth conditions except for serine as substrate. Sarcosine reductase and betaine reductase were only induced when E. acidaminophilum was grown on sarcosine and betaine, respectively. Creatine was metabolized via sarcosine. [75Se]-selenite labeling revealed about the same pattern of predominant labeled proteins in glycine-, sarcosine-, and betaine-grown cells.Abbreviations DTE dithioerythritol - TES N-Tris (hydroxymethyl) methyl-2-amino-ethane sulfonic acid  相似文献   

15.
The jejunal mechanisms for the electrogenic transfer of four neutral amino acids (alanine, leucine, methionine, valine) and for sarcosine were characterised by an electrical method in vitro. The values for apparent Km obtained electrically agree well with those assessed by conventional chemical techniques. Hypothyroidism and/or fasting rats for 3 days induced differential changes in the apparent Km and p.d.max for the various amino acids. These alterations were interpreted as indicating the presence of at least three mechanisms for neutral amino acid transfer and one for sarcosine.In euthyroid rats, only alanine showed changes in apparent Km (decrease) and p.d.max (decrease) after fasting for 3 days. With hypothyroidism the kinetic parameters of electrogenic transfer for alanine, valine and sarcosine were significantly altered while those for leucine and methionine were unaffected.  相似文献   

16.
Astrid R. Klingen  Carola Hunte 《BBA》2007,1767(3):204-221
Cytochrome bc1 is a major component of biological energy conversion that exploits an energetically favourable redox reaction to generate a transmembrane proton gradient. Since the mechanistic details of the coupling of redox and protonation reactions in the active sites are largely unresolved, we have identified residues that undergo redox-linked protonation state changes. Structure-based Poisson-Boltzmann/Monte Carlo titration calculations have been performed for completely reduced and completely oxidised cytochrome bc1. Different crystallographically observed conformations of Glu272 and surrounding residues of the cytochrome b subunit in cytochrome bc1 from Saccharomyces cerevisiae have been considered in the calculations. Coenzyme Q (CoQ) has been modelled into the CoQ oxidation site (Qo-site). Our results indicate that both conformational and protonation state changes of Glu272 of cytochrome b may contribute to the postulated gating of CoQ oxidation. The Rieske iron-sulphur cluster could be shown to undergo redox-linked protonation state changes of its histidine ligands in the structural context of the CoQ-bound Qo-site. The proton acceptor role of the CoQ ligands in the CoQ reduction site (Qi-site) is supported by our results. A modified path for proton uptake towards the Qi-site features a cluster of conserved lysine residues in the cytochrome b (Lys228) and cytochrome c1 subunits (Lys288, Lys289, Lys296). The cardiolipin molecule bound close to the Qi-site stabilises protons in this cluster of lysine residues.  相似文献   

17.
ABSTRACT

Pyoverdines, a group of peptide siderophores produced by Pseudomonas species, function not only in iron acquisition, but also in their virulence in hosts. Thus, chemical inhibition of pyoverdine production may be an effective strategy to control Pseudomonas virulence. In the plant pathogen Pseudomonas cichorii SPC9018 (SPC9018), pyoverdine production is required for virulence on eggplant. We screened microbial culture extracts in a pyoverdine-production inhibition assay of SPC9018 and found Streptomyces sp. RM-32 as a candidate-producer. We isolated two active compounds from RM-32 cultures, and elucidated their structures to be actinomycins X2 and D. Actinomycins X2 and D inhibited pyoverdine production by SPC9018 with IC50 values of 17.6 and 29.6 μM, respectively. Furthermore, pyoverdine production in other Pseudomonas bacteria, such as the mushroom pathogen P. tolaasii, was inhibited by the actinomycins. Therefore, these actinomycins may be useful as chemical tools to examine pyoverdine functions and as seed compounds for anti-Pseudomonas virulence agents.  相似文献   

18.
a fluorometric-colorimetric fluorescamine analyzer was successfully utilized for the analysis of a series of actinomycin hydrolysates containing N-methyl-amino acids, proline, and proline derivatives. An explanation is proposed for the low threonine values consistently observed in the analysis of actinomycin hydrolysates and alternate conditions for improved hydrolysis of actinomycins are proposed.  相似文献   

19.
Our previous work described a clear loss of Escherichia coli (E. coli) membrane integrity after incubation with glycine or its N-methylated derivatives N-methylglycine (sarcosine) and N,N-dimethylglycine (DMG), but not N,N,N-trimethylglycine (betaine), under alkaline stress conditions. The current study offers a thorough viability analysis, based on a combination of real-time physiological techniques, of E. coli exposed to glycine and its N-methylated derivatives at alkaline pH. Flow cytometry was applied to assess various physiological parameters such as membrane permeability, esterase activity, respiratory activity and membrane potential. ATP and inorganic phosphate concentrations were also determined. Membrane damage was confirmed through the measurement of nucleic acid leakage. Results further showed no loss of esterase or respiratory activity, while an instant and significant decrease in the ATP concentration occurred upon exposure to either glycine, sarcosine or DMG, but not betaine. There was a clear membrane hyperpolarization as well as a significant increase in cellular inorganic phosphate concentration. Based on these results, we suggest that the inability to sustain an adequate level of ATP combined with a decrease in membrane functionality leads to the loss of bacterial viability when exposed to the proton scavengers glycine, sarcosine and DMG at alkaline pH.  相似文献   

20.
The binding of substrates to the A-site half (A′) and the P-site half (P′) of the peptidyltransferase center was measured by means of equilibrium dialysis. The tRNA fragments C-A-C-C-A-Leu and C-A-C-C-A-(N-acetyl)Leu were used as A′-site and P′-site substrates, respectively. The A′- and P′-substrates bound well to 50 S in contrast to 30 S subunits; significant binding to 23 S and 16 S RNA was also found. The binding of the P′-site substrate to 23 S RNA and 50 S subunits was very similar at various Mg2+ and K+ concentrations, indicating that the 23 S RNA is probably directly involved in the binding of the 3′-end of the peptidyl-tRNA. Cooperative effects at the peptidyltransferase center were found using chloramphenicol and deacylated tRNA as competitors, which completely inhibited the substrate binding to one site whilst drastically stimulating binding to the other. Chloramphenicol inhibited the binding of the A′-site substrate C-A-C-C-A-Leu, whereas the binding of the corresponding P′-site substrate was stimulated. In contrast, deacylated tRNA blocked the binding of the P′-site substrate, but stimulated the corresponding A′-site binding. Similarly, the trinucleotide Cp,CpA inhibited binding of the P′-site substrate (showing complete inhibition at 70 μm) whereas binding of the A′-site substrate was slightly stimulated at concentrations below 70 μm.  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