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.
The effect of salinity and growth temperature on the accumulation of intracellular organic solutes was examined in the hyperthermophilic archaeon Palaeococcus ferrophilus. The genus Palaeococcus represents a deep-branching lineage of the order Thermococcales, which diverged before Thermococcus and Pyrococcus. Palaeococcus ferrophilus accumulated mannosylglycerate, glutamate, and aspartate as major compatible solutes. Unlike members of the genera Pyrococcus and Thermococcus, Palaeococcus ferrophilus did not accumulate di-myo-inositol phosphate, a canonical solute of hyperthermophiles. The level of mannosylglycerate increased in response to both heat and salt stress; glutamate increased at supraoptimal growth temperatures, whereas aspartate increased at supraoptimal salt concentration. Proline, alanine, and trehalose were also found in lesser amounts, but their levels did not respond significantly to any of the stresses imposed. Additionally, the genes involved in the synthesis of mannosylglycerate in Palaeococcus ferrophilus and Thermococcus litoralis were identified. In both organisms the synthesis proceeds via the two-step pathway comprising mannosyl-3-phosphoglycerate synthase (MPGS) (EC 2.4.1.217) and mannosyl-3-phosphoglycerate phosphatase (MPGP) (EC 3.1.3.70). The mpgS and mpgP genes of Palaeococcus ferrophilus were expressed in Escherichia coli and the proteins were characterized. MPGS had maximal activity at 90°C and pH near 7.0, and was strictly dependent on Mg2+. MPGP had optimal activity at 90°C and pH 6.0 and was barely dependent on Mg2+. The half-life values for inactivation of MPGS and MPGP at 83°C were 18 and 25 min, respectively. A comparative discussion of the osmo- and thermoadaptation strategies in these three genera of the Thermococcales is presented.  相似文献   

2.
The accumulation of compatible solutes was studied in the hyperthermophilic bacterium Aquifex pyrophilus as a function of the temperature and the NaCl concentration of the growth medium. Nuclear magnetic resonance analysis of cell extracts revealed the presence of α- and β-glutamate, di-mannosyl-di-myo-inositol phosphate, di-myo-inositol phosphate, and an additional compound here identified as 1-glyceryl-1-myo-inosityl phosphate. All solutes accumulated by A. pyrophilus are negatively charged at physiological pH. The intracellular levels of di-myo-inositol phosphate increased in response to supraoptimal growth temperature, while α- and β-glutamate accumulated in response to osmotic stress, especially at growth temperatures below the optimum. The newly discovered compound, 1-glyceryl-1-myo-inosityl phosphate, appears to play a double role in osmo- and thermoprotection, since its intracellular pool increased primarily in response to a combination of osmotic and heat stresses. This work also uncovered the nature of the unknown compound, previously detected in Archaeoglobus fulgidus (L. O. Martins et al., Appl. Environ. Microbiol. 63:896-902, 1997). The curious structural relationship between diglycerol phosphate (found only in Archaeoglobus species), di-myo-inositol phosphate (a canonical solute of hyperthermophiles), and the newly identified solute is highlighted. This is the first report on the occurrence of 1-glyceryl-1-myo-inosityl phosphate in living systems.  相似文献   

3.
Many Archaea and Bacteria isolated from hot, marine environments accumulate di-myo-inositol-phosphate (DIP), primarily in response to heat stress. The biosynthesis of this compatible solute involves the activation of inositol to CDP-inositol via the action of a recently discovered CTP:inositol-1-phosphate cytidylyltransferase (IPCT) activity. In most cases, IPCT is part of a bifunctional enzyme comprising two domains: a cytoplasmic domain with IPCT activity and a membrane domain catalyzing the synthesis of di-myo-inositol-1,3′-phosphate-1′-phosphate from CDP-inositol and l-myo-inositol phosphate. Herein, we describe the first X-ray structure of the IPCT domain of the bifunctional enzyme from the hyperthermophilic archaeon Archaeoglobus fulgidus DSMZ 7324. The structure of the enzyme in the apo form was solved to a 1.9-Å resolution. The enzyme exhibited apparent Km values of 0.9 and 0.6 mM for inositol-1-phosphate and CTP, respectively. The optimal temperature for catalysis was in the range 90 to 95°C, and the Vmax determined at 90°C was 62.9 μmol · min−1 · mg of protein−1. The structure of IPCT is composed of a central seven-stranded mixed β-sheet, of which six β-strands are parallel, surrounded by six α-helices, a fold reminiscent of the dinucleotide-binding Rossmann fold. The enzyme shares structural homology with other pyrophosphorylases showing the canonical motif G-X-G-T-(R/S)-X4-P-K. CTP, l-myo-inositol-1-phosphate, and CDP-inositol were docked into the catalytic site, which provided insights into the binding mode and high specificity of the enzyme for CTP. This work is an important step toward the final goal of understanding the full catalytic route for DIP synthesis in the native, bifunctional enzyme.  相似文献   

4.
The accumulation of organic solutes was investigated in the thermophilic bacteria Persephonella marina and Marinitoga piezophila, two representatives of the deepest lineages in the domain Bacteria. These organisms grow optimally at around 70 °C in medium containing 3 % NaCl. A new disaccharide, accumulating in Persephonella marina, was identified as ??(1?C6)glucosyl-??(1?C2)glucosylglycerate (GGG), by nuclear magnetic resonance. This identification was validated by comparison with the spectra of the compound obtained by chemical synthesis. Besides GGG, the solute pool of Persephonella marina comprised ??-glutamate, di-myo-inositol-1,3??-phosphate and 2-O-??-glucosylglycerate. In contrast, amino acids such as ??-glutamate, proline and alanine were the dominant components of the solute pool of Marinitoga piezophila and sugar derivatives were absent. The ability of GGG to protect protein structure against heat denaturation was assessed using model proteins. A genomic search for the biosynthetic pathways of known ionic solutes in Aquificales and Thermotogales shows the inability of this analysis to predict the nature of compatible solutes and underlines the need for efficient cultivation techniques.  相似文献   

5.
Protein misfolding and deposition in the brain are implicated in the etiology of numerous neurodegenerative disorders. Here, organic solutes characteristic of microorganisms adapted to hot environments, were tested on experimental cell models of Huntington’s and Parkinson’s diseases. Diglycerol phosphate, di-myo-inositol phosphate, mannosylglycerate, and mannosylglyceramide were not toxic to the cells, at 10 mM concentration, but caused a decrease in cell density, which suggested an effect on proliferation. In contrast, mannosyl-lactate, an artificial analogue of mannosylglycerate, had a negative impact on cell viability. Concerning protein aggregation, inclusions of mutant huntingtin were reduced in the presence of diglycerol phosphate and di-myo-inositol phosphate, increased with mannosylglycerate, while mannosyl-lactate and mannosylglyceramide had no significant effect. α-Synuclein aggregation was not affected by the solutes tested, except for di-myo-inositol phosphate that led to a slight increased percentage of cells displaying visible aggregates. These solutes might be useful in the development of therapies for protein misfolding diseases.  相似文献   

6.
In addition to di-myo-inositol-1,3′-phosphate (DIP), a compatible solute widespread in hyperthermophiles, the organic solute pool of Thermotoga maritima comprises 2-(O-β-d-mannosyl)-di-myo-inositol-1,3′-phosphate (MDIP) and 2-(O-β-d-mannosyl-1,2-O-β-d-mannosyl)-di-myo-inositol-1,3′-phosphate (MMDIP), two newly identified β-1,2-mannosides. In cells grown under heat stress, MDIP was the major solute, accounting for 43% of the total pool; MMDIP and DIP accumulated to similar levels, each corresponding to 11.5% of the total pool. The synthesis of MDIP involved the transfer of the mannosyl group from GDP-mannose to DIP in a single-step reaction catalyzed by MDIP synthase. This enzyme used MDIP as an acceptor of a second mannose residue, yielding the di-mannosylated compound. Minor amounts of the tri-mannosylated form were also detected. With a genomic approach, putative genes for MDIP synthase were identified in the genome of T. maritima, and the assignment was confirmed by functional expression in Escherichia coli. Genes with significant sequence identity were found only in the genomes of Thermotoga spp., Aquifex aeolicus, and Archaeoglobus profundus. MDIP synthase of T. maritima had maximal activity at 95°C and apparent Km values of 16 mM and 0.7 mM for DIP and GDP-mannose, respectively. The stereochemistry of MDIP was characterized by isotopic labeling and nuclear magnetic resonance (NMR): DIP selectively labeled with carbon 13 at position C1 of the l-inositol moiety was synthesized and used as a substrate for MDIP synthase. This β-1,2-mannosyltransferase is unrelated to known glycosyltransferases, and within the domain Bacteria, it is restricted to members of the two deepest lineages, i.e., the Thermotogales and the Aquificales. To our knowledge, this is the first β-1,2-mannosyltransferase characterized thus far.Thermotoga maritima was first isolated from hot marine sediments on Vulcano Island, Italy, being able to grow between 55°C and 90°C (14). This strictly anaerobic bacterium ferments a variety of simple and complex carbohydrates to acetate, hydrogen, and CO2 (10). In line with these metabolic traits, a substantial percentage of the genes annotated in the genome of this hyperthermophile are allocated to the metabolism of mono- and polysaccharides (8, 23). Therefore, T. maritima has been pointed out as a source of glycoside hydrolases with potential industrial relevance, namely, in processes of conversion of biomass into biofuels (3, 34).Like many other hyperthermophiles isolated from marine environments, Thermotoga maritima is slightly halophilic (optimum NaCl concentration of 2.7%, wt/vol) and has developed biochemical strategies to counterbalance the external osmotic pressure. The accumulation of low-molecular-mass organic compounds in the cytoplasm is the most common osmoadaptation mechanism, which enables a rapid response to fluctuations in the salinity of the external medium. Interestingly, the organic solutes encountered in organisms adapted to thrive in hot environments are clearly different from those used by mesophiles, leading to the view that osmolytes of (hyper)thermophiles could play an additional role as protectors of macromolecules and other cellular components against heat damage. This notion is further fuelled by the finding that the total pool of organic solutes of (hyper)thermophiles increases notably not only at supraoptimal salinity but also in response to heat stress conditions (30).Over the last decade, our team has directed considerable effort to assess the role of osmolytes in the thermo-adaptation strategies of hyperthermophiles. Despite the scarcity of genetic tools for manipulation of marine hyperthermophiles, a number of novel organic solutes were identified and the corresponding biosynthetic pathways characterized at the genetic and biochemical levels (15, 17, 30), providing critical knowledge for engaging in elucidation of the molecular basis of the whole process, from the sensing of stress to the synthesis of specific osmolytes. In this context, we recently reported the characterization of the pathway for synthesis of di-myo-inositol-1,3′-phosphate, the most common solute within hyperthermophiles (5). Additionally, the genes and enzymes involved in the relevant reaction steps were disclosed. The synthesis proceeds via a phosphorylated form of DIP, and the respective synthase is a membrane-associated enzyme that catalyzes the condensation of CDP-inositol with inositol-1-phosphate (26, 27).The solute pool in members of the order Thermotogales was investigated a few years ago (19). Thermotoga neapolitana responded to heat stress with a strong accumulation of DIP and DIP derivatives. One of the solutes was assigned to a mannosylated form of DIP, at that time designated di-mannosyl-di-myo-inositol phosphate; moreover, the presence of a second DIP derivative was proposed, but its structure remained elusive. Therefore, we set out to fully characterize the solute pool of Thermotoga spp. and to identify the genes and the enzyme(s) involved in the synthesis of the DIP derivatives. Members of the genus Thermotoga accumulated DIP and two mannosylated forms of this compound, herein fully characterized using isotopic labeling, NMR, and mass spectrometry. Moreover, the pathway for the synthesis of these novel solutes was identified, leading to the discovery of a unique β-1,2-mannosyltransferase that catalyzes the transfer of the mannosyl group from GDP-mannose to DIP.  相似文献   

7.
Many of the marine microorganisms which are adapted to grow at temperatures above 80°C accumulate di-myo-inositol phosphate (DIP) in response to heat stress. This led to the hypothesis that the solute plays a role in thermoprotection, but there is a lack of definitive experimental evidence. Mutant strains of Thermococcus kodakarensis (formerly Thermococcus kodakaraensis), manipulated in their ability to synthesize DIP, were constructed and used to investigate the involvement of DIP in thermoadaptation of this archaeon. The solute pool of the parental strain comprised DIP, aspartate, and α-glutamate. Under heat stress the level of DIP increased 20-fold compared to optimal conditions, whereas the pool of aspartate increased 4.3-fold in response to osmotic stress. Deleting the gene encoding the key enzyme in DIP synthesis, CTP:inositol-1-phosphate cytidylyltransferase/CDP-inositol:inositol-1-phosphate transferase, abolished DIP synthesis. Conversely, overexpression of the same gene resulted in a mutant with restored ability to synthesize DIP. Despite the absence of DIP in the deletion mutant, this strain exhibited growth parameters similar to those of the parental strain, both at optimal (85°C) and supraoptimal (93.7°C) temperatures for growth. Analysis of the respective solute pools showed that DIP was replaced by aspartate. We conclude that DIP is part of the strategy used by T. kodakarensis to cope with heat stress, and aspartate can be used as an alternative solute of similar efficacy. This is the first study using mutants to demonstrate the involvement of compatible solutes in the thermoadaptation of (hyper)thermophilic organisms.Hyperthermophilic bacteria and archaea isolated from saline environments accumulate unusual organic solutes in response to osmotic as well as heat stress. Mannosylglycerate, mannosylglyceramide, di-myo-inositol phosphate, mannosyl-di-myo-inositol phosphate (DIP), diglycerol phosphate, and glycero-phospho-myo-inositol are examples of compatible solutes highly restricted to thermophiles and hyperthermophiles (27, 31). Our team has, over several years, examined the compatible solute composition in a large number of hyperthermophiles and their accumulation under stressful conditions. The data reveal a trend toward specialization of roles in thermoadaptation and osmoadaptation. Indeed, mannosylglycerate and diglycerol phosphate typically accumulate in response to increased NaCl concentration in the growth medium, whereas the levels of DIP and derivatives consistently increase at supraoptimal growth temperatures (11, 16, 17, 27, 31).DIP is widespread among extreme archaeal hyperthermophiles, such as Methanotorris igneus, Aeropyrum pernix, Stetteria hydrogenophila, Pyrodictium occultum, Pyrolobus fumarii, Archaeoglobus spp., and all the members of the Thermococcales examined thus far, except Palaeococcus ferrophilus (5, 7, 11, 13, 16, 18, 31). This organic solute has also been found in representatives of the two hyperthermophilic bacterial genera, Aquifex and Thermotoga (14, 17, 22).The specific chemical nature of solutes encountered in hyperthermophiles, together with their accumulation in response to elevated temperatures, led to the hypothesis that they play a role in thermoprotection of cellular components in vivo. However, there is a lack of convincing experimental evidence, such as that obtained with suitable mutants. Progress toward understanding the physiological functions of these solutes critically depends on two conditions: the availability of genetic tools to manipulate hyperthermophilic organisms and knowledge about the genes and enzymes implicated in the synthesis of these unusual solutes.Thermococcus kodakarensis (formerly Thermococcus kodakaraensis) is a member of the order Thermococcales with an optimal growth temperature of 85°C and is able to grow at temperatures up to 94°C in batch cultures. The NaCl concentration for optimal growth matches that of seawater (1). T. kodakarensis is the only marine hyperthermophile for which a number of genetic tools have been developed, including Escherichia coli-T. kodakarensis shuttle vectors and a reliable gene disruption system (19, 29, 32, 34). The genome of T. kodakarensis possesses a gene encoding CTP:inositol-1-phosphate cytidylyltransferase/CDP-inositol:inositol-1-phosphate transferase (IPCT/DIPPS), a key enzyme in DIP synthesis (2, 25, 26). This enzyme catalyzes the synthesis of CDP-inositol from CTP and inositol-1-phosphate as well as the transfer of the inositol group from CDP-inositol to a second molecule of inositol-1-phosphate to yield a phosphorylated form of DIP (2). Therefore, we set out to investigate whether DIP was involved in thermoadaptation of T. kodakarensis. A DIP-deficient mutant was constructed by deleting the IPCT/DIPPS gene; subsequently, this strain was complemented in this activity by inserting the gene under the control of a constitutive promoter, resulting in a construct with restored ability to synthesize DIP. The effects of heat and osmotic stress on the pattern of solute accumulation and on the growth profiles of the two mutants provided evidence for the involvement of DIP in thermoprotection.  相似文献   

8.
We describe a new species, Thermococcus litoralis, which is different from the type species Thermococcus celer in molecular, morphological and physiological characteristics.Abbreviations 3 x SSC (standard saline citrate) - 0.45 M NaCl 0.045 M Na3-citrate  相似文献   

9.
Extracellular pullulanases were purified from cell-free culture supernatants of the marine thermophilic archaea Thermococcus litoralis (optimal growth temperature, 90°C) and Pyrococcus furiosus (optimal growth temperature, 98°C). The molecular mass of the T. litoralis enzyme was estimated at 119,000 Da by electrophoresis, while the P. furiosus enzyme exhibited a molecular mass of 110,000 Da under the same conditions. Both enzymes tested positive for bound sugar by the periodic acid-Schiff technique and are therefore glycoproteins. The thermoactivity and thermostability of both enzymes were enhanced in the presence of 5 mM Ca2+, and under these conditions, enzyme activity could be measured at temperatures of up to 130 to 140°C. The addition of Ca2+ also affected substrate binding, as evidenced by a decrease in Km for both enzymes when assayed in the presence of this metal. Each of these enzymes was able to hydrolyze, in addition to the α-1,6 linkages in pullulan, α-1,4 linkages in amylose and soluble starch. Neither enzyme possessed activity against maltohexaose or other smaller α-1,4-linked oligosaccharides. The enzymes from T. litoralis and P. furiosus appear to represent highly thermostable amylopullulanases, versions of which have been isolated from less-thermophilic organisms. The identification of these enzymes further defines the saccharide-metabolizing systems possessed by these two organisms.  相似文献   

10.
The product of myo-inositol-1-phosphate synthase, EC 5.5.1.4, from mature pollen of Lilium longiflorum Thunb., cv Ace (Easter lily) and that of myo-inositol kinase, EC 2.7.1.64, from wheat germ has been identified as 1l-myo-inositol-1-phosphate by gas chromatography of its trimethylsilyl-methyl phosphate derivative on a glass capillary column bearing a chiral phase.  相似文献   

11.
Fourteen strains of hyperthermophilic organotrophic anaerobic marine Archaea were isolated from shallow water and deep-sea hot vents, and four of them were characterized. These isolates, eight previously published strains, and six type strains of species of the order Thermococcales were selected for the study of cell wall components by means of thin sectioning or freeze-etching electron microscopy. The cell envelopes of most isolates were shown to consist of regularly arrayed surface protein layers, either single or double, with hexagonal lattice (p6) symmetry, as the exclusive constituents outside the cytoplasmic membrane. The S-layers studied differed in center-to-center spacing and molecular mass of the constituent protein subunits. Polyclonal antisera raised against the cells of 10 species were found to be species-specific and allowed 12 new isolates from shallow water hot vents to be identified as representatives of the species Thermococcus litoralis, Thermococcus stetteri, Thermococcus chitonophagus, and Thermococcus pacificus. Of the 7 deep-sea isolates, only 1 was identified as a T. litoralis strain. Thus, hyperthermophilic marine organotrophic isolates obtained from deep-sea hot vents showed greater diversity with regard to their S-layer proteins than shallow water isolates. Received: February 5, 1999 / Accepted: May 11, 1999  相似文献   

12.
β-Alanine is a precursor for coenzyme A (CoA) biosynthesis and is a substrate for the bacterial/eukaryotic pantothenate synthetase and archaeal phosphopantothenate synthetase. β-Alanine is synthesized through various enzymes/pathways in bacteria and eukaryotes, including the direct decarboxylation of Asp by aspartate 1-decarboxylase (ADC), the degradation of pyrimidine, or the oxidation of polyamines. However, in most archaea, homologs of these enzymes are not present; thus, the mechanisms of β-alanine biosynthesis remain unclear. Here, we performed a biochemical and genetic study on a glutamate decarboxylase (GAD) homolog encoded by TK1814 from the hyperthermophilic archaeon Thermococcus kodakarensis. GADs are distributed in all three domains of life, generally catalyzing the decarboxylation of Glu to γ-aminobutyrate (GABA). The recombinant TK1814 protein displayed not only GAD activity but also ADC activity using pyridoxal 5′-phosphate as a cofactor. Kinetic studies revealed that the TK1814 protein prefers Asp as its substrate rather than Glu, with nearly a 20-fold difference in catalytic efficiency. Gene disruption of TK1814 resulted in a strain that could not grow in standard medium. Addition of β-alanine, 4′-phosphopantothenate, or CoA complemented the growth defect, whereas GABA could not. Our results provide genetic evidence that TK1814 functions as an ADC in T. kodakarensis, providing the β-alanine necessary for CoA biosynthesis. The results also suggest that the GAD activity of TK1814 is not necessary for growth, at least under the conditions applied in this study. TK1814 homologs are distributed in a wide range of archaea and may be responsible for β-alanine biosynthesis in these organisms.  相似文献   

13.
The intracellular accumulation of small organic solutes was described in the marine bacterium Rhodopirellula baltica, which belongs to the globally distributed phylum Planctomycetes whose members exhibit an intriguing lifestyle and cell morphology. Sucrose, α-glutamate, trehalose and mannosylglucosylglycerate (MGG) are the main solutes involved in the osmoadaptation of R. baltica. The ratio and total intracellular organic solutes varied significantly in response to an increase in salinity, temperature and nitrogen content. R. baltica displayed an initial response to both osmotic and thermal stresses that includes α-glutamate accumulation. This trend was followed by a rather unique and complex osmoadaptation mechanism characterized by a dual response to sub-optimal and supra-optimal salinities. A reduction in the salinity to sub-optimal conditions led primarily to the accumulation of trehalose. In contrast, R. baltica responded to salt stress mostly by increasing the intracellular levels of sucrose. The switch between the accumulation of trehalose and sucrose was by far the most significant effect caused by increasing the salt levels of the medium. Additionally, MGG accumulation was found to be salt- as well as nitrogen-dependent. MGG accumulation was regulated by nitrogen levels replacing α-glutamate as a K+ counterion in nitrogen-poor environments. This is the first report of the accumulation of compatible solutes in the phylum Planctomycetes and of the MGG accumulation in a mesophilic organism.  相似文献   

14.
1. A number of compounds known to inhibit polyamine biosynthesis at various steps in the biosynthetic pathway were tested for their ability to inhibit growth and decrease polyamine concentrations in virally transformed mouse fibroblasts (SV-3T3 cells). 2. Virtually complete inhibition of growth was produced by the inhibitors of ornithine decarboxylase α-methylornithine and α-difluoromethylornithine and by the inhibitors of S-adenosylmethionine decarboxylase 1,1′-[(methylethanediylidene)dinitrilo]diguanidine and 1,1′-[(methylethanediylidene)dinitrilo]bis-(3-aminoguanidine). The former inhibitors decreased putrescine and spermidine contents in the cells to very low values, whereas the latter substantially increased putrescine but decreased spermidine concentrations. The inhibitory effects of all of these inhibitors on cell growth could be prevented by the addition of spermidine, suggesting that spermidine depletion is the underlying cause of their inhibition of growth. 3. α-Difluoromethylornithine, which is an irreversible inhibitor of ornithine decarboxylase, was a more potent inhibitor of growth and polyamine production (depleting spermidine almost completely and spermine significantly) than α-methylornithine, which is a competitive inhibitor. This was not the case with the inhibitors of S-adenosylmethionine decarboxylase where 1,1′-[(methylethanediylidene)dinitrilo]diguanidine, a reversible inhibitor, was more active than 1,1′-[(methylethanediylidene)dinitrilo]bis-(3-aminoguanidine), an irreversible inhibitor. It is suggested that this effect may be due to the lesser uptake and/or greater chemical reactivity of the latter compound. 4. Various nucleoside derivatives of S-adenosylhomocysteine that inhibited spermidine synthase in vitro did not have significant inhibitory action against polyamine accumulation in the cell. These compounds, which included S-adenosylhomocysteine sulphone, decarboxylated S-adenosylhomocysteine sulphone, decarboxylated S-adenosylhomocysteine sulphoxide and S-adenosyl-4-thio-butyric acid sulphone did not inhibit cell growth or polyamine content until cytotoxic concentrations were added. 5. 5′-Methylthioadenosine, 5′-isobutylthioadenosine and 5′-methylthiotubercidin, which inhibit aminopropyltransferase activity in vitro, all inhibited cell growth and decreased spermidine content. Although these compounds were most active against spermine synthase in vitro, they acted in the cell primarily to decrease spermidine content. Cell growth could not be restored to normal values by addition of spermidine, suggesting that these nucleosides have another inhibitory action towards cellular proliferation. 6. 5′-Methylthioadenosine and 5′-isobutylthioadenosine are degraded by a phosphorylase present in SV3T3 cells, yielding 5-methylthioribose-1-phosphate and 5-isobutylthioribose-1-phosphate respectively, and adenine. This degradation appears to decrease the inhibitory action towards cell growth, suggesting that the nucleosides themselves are exerting the inhibitory action. 5′-Methylthiotubercidin, which is not a substrate for the phosphorylase and is a competitive inhibitor of it, was the most active of these nucleosides in inhibiting cell growth and spermidine content. 5′-Methylthiotubercidin and α-difluoromethylornithine had additive effects on retarding cell growth, but not on cellular spermine accumulation, also suggesting that the primary growth-inhibiting action of the nucleoside was not on polyamine production. 7. These results support the concept that 5′-methylthioadenosine phosphorylase plays an important role in permitting cell growth to continue by preventing the build-up of inhibitory intracellular concentrations of 5′-methylthioadenosine.  相似文献   

15.
Marine hyperthermophiles accumulate small organic compounds, known as compatible solutes, in response to supraoptimal temperatures or salinities. Pyrococcus furiosus is a hyperthermophilic archaeon that grows optimally at temperatures near 100°C. This organism accumulates mannosylglycerate (MG) and di-myo-inositol phosphate (DIP) in response to osmotic and heat stress, respectively. It has been assumed that MG and DIP are involved in cell protection; however, firm evidence for the roles of these solutes in stress adaptation is still missing, largely due to the lack of genetic tools to produce suitable mutants of hyperthermophiles. Recently, such tools were developed for P. furiosus, making this organism an ideal target for that purpose. In this work, genes coding for the synthases in the biosynthetic pathways of MG and DIP were deleted by double-crossover homologous recombination. The growth profiles and solute patterns of the two mutants and the parent strain were investigated under optimal growth conditions and also at supraoptimal temperatures and NaCl concentrations. DIP was a suitable replacement for MG during heat stress, but substitution of MG for DIP and aspartate led to less efficient growth under conditions of osmotic stress. The results suggest that the cascade of molecular events leading to MG synthesis is tuned for osmotic adjustment, while the machinery for induction of DIP synthesis responds to either stress agent. MG protects cells against heat as effectively as DIP, despite the finding that the amount of DIP consistently increases in response to heat stress in the nine (hyper)thermophiles examined thus far.  相似文献   

16.
Summary Hyperthermostable proteases were characterized from five archaeobacterial species (Thermococcus celer, T. stetteri, Thermococcus strain AN 1, T. litoralis, Staphylothermus marinus) and the hyperthermophilic eubacterium Thermobacteroides proteolyticus. These proteases, which were found to be of the serine type, exhibited a preference for phenylalanine in the carboxylic side of the peptide. The enzymes from Thermococcus stetteri and T. litoralis hydrolysed most substrates (peptides) tested. All proteases were extremely thermostable and demonstrated optimal activities between 80 and 95°C. The pH optimum was either neutral (T. celer, Thermococcus strain AN 1) or alkaline. The protease of Thermobacteroides proteolyticus was optimally active at pH 9.5. Zymogram staining showed the presence of multiple protease bands for all strains investigated.Offprint requests to: G. Antranikian  相似文献   

17.
RNA terminal phosphate cyclase catalyzes the ATP-dependent conversion of a 3′-phosphate RNA end to a 2′,3′-cyclic phosphate via covalent enzyme-(histidinyl-Nϵ)-AMP and RNA(3′)pp(5′)A intermediates. Here, we report that Escherichia coli RtcA (and its human homolog Rtc1) are capable of cyclizing a 2′-phosphate RNA end in high yield. The rate of 2′-phosphate cyclization by RtcA is five orders of magnitude slower than 3′-phosphate cyclization, notwithstanding that RtcA binds with similar affinity to RNA3′p and RNA2′p substrates. These findings expand the functional repertoire of RNA cyclase and suggest that phosphate geometry during adenylate transfer to RNA is a major factor in the kinetics of cyclization. RtcA is coregulated in an operon with an RNA ligase, RtcB, that splices RNA 5′-OH ends to either 3′-phosphate or 2′,3′-cyclic phosphate ends. Our results suggest that RtcA might serve an end healing function in an RNA repair pathway, by converting RNA 2′-phosphates, which cannot be spliced by RtcB, to 2′,3′-cyclic phosphates that can be sealed. The rtcBA operon is controlled by the σ54 coactivator RtcR encoded by an adjacent gene. This operon arrangement is conserved in diverse bacterial taxa, many of which have also incorporated the RNA-binding protein Ro (which is implicated in RNA quality control under stress conditions) as a coregulated component of the operon.  相似文献   

18.
We previously found that mice fed lutein accumulated its oxidative metabolites (3′-hydroxy-ε,ε-caroten-3-one and ε,ε-carotene-3,3′-dione) as major carotenoids, suggesting that mammals can convert xanthophylls to keto-carotenoids by the oxidation of hydroxyl groups. Here we elucidated the metabolic activities of mouse liver for several xanthophylls. When lutein was incubated with liver postmitochondrial fraction in the presence of NAD+, (3′R,6′R)-3′-hydroxy-β,ε-caroten-3-one and (6RS,3′R,6′R)-3′-hydroxy-ε,ε-caroten-3-one were produced as major oxidation products. The former accumulated only at the early stage and was assumed to be an intermediate, followed by isomerization to the latter. The configuration at the C3′ and C6′ of the ε-end group in lutein was retained in the two oxidation products. These results indicate that the 3-hydroxy β-end group in lutein was preferentially oxidized to a 3-oxo ε-end group via a 3-oxo β-end group. Other xanthophylls such as β-cryptoxanthin and zeaxanthin, which have a 3-hydroxy β-end group, were also oxidized in the same manner as lutein. These keto-carotenoids, derived from dietary xanthophylls, were confirmed to be present in plasma of normal human subjects, and β,ε-caroten-3′-one was significantly increased by the ingestion of β-cryptoxanthin. Thus, humans as well as mice have oxidative activity to convert the 3-hydroxy β-end group of xanthophylls to a 3-oxo ε-end group.  相似文献   

19.
Biosynthesis of di-myo-inositol-1,1′-phosphate (DIP) is proposed to occur with myo-inositol and myo-inositol 1-phosphate (I-1-P) used as precursors. Activation of the I-1-P with CTP and condensation of the resultant CDP-inositol (CDP-I) with myo-inositol then generates DIP. The sole known biosynthetic pathway of inositol in all organisms is the conversion of d-glucose-6-phosphate to myo-inositol. This conversion requires two key enzymes: l-I-1-P synthase and I-1-P phosphatase. Enzymatic assays using 31P nuclear magnetic resonance spectroscopy as well as a colorimetric assay for inorganic phosphate have confirmed the occurrence of l-I-1-P synthase and a moderately specific I-1-P phosphatase. The enzymatic reaction that couples CDP-I with myo-inositol to generate DIP has also been detected in Methanococcus igneus. 13C labeling studies with [2,3-13C]pyruvate and [3-13C]pyruvate were used to examine this pathway in M. igneus. Label distribution in DIP was consistent with inositol units formed from glucose-6-phosphate, but the label in the glucose moiety was scrambled via transketolase and transaldolase activities of the pentose phosphate pathway.Di-myo-inositol-1,1′-phosphate (DIP) is an unusual inositol derivative that has been identified as a major solute in hyperthermophilic archaea including Pyrococcus woesei (22), Pyrococcus furiosus (16), Methanococcus igneus (5), and several eubacteria of the order Thermotogales (15). Intracellular DIP increases with increasing extracellular concentrations of NaCl in both M. igneus (5) and P. furiosus (16). DIP also increases dramatically at supraoptimal growth temperatures (>80°C for M. igneus and 98 to 101°C for P. furiosus). The unusual intracellular high concentration of K+ ions and the extreme optimal growth temperatures (100 to 104°C) of P. woesei (30) suggested the role of DIP as a main counterion of K+ with a possible thermostabilizing action. Scholz et al. (22) demonstrated that among several salts, the potassium salt of DIP provided optimum enzyme stabilization when the activity of glyceraldehyde-3-phosphate dehydrogenase of P. woesei was tested at 105°C under anaerobic conditions.Since de novo synthesis of DIP occurs in response to external levels of NaCl and temperature, there must be regulatory biosynthetic mechanisms linked to osmotic pressure and temperature. To study the regulation, the enzymes and/or other proteins responsible for synthesis of this compatible solute must be isolated. This requires knowledge of the biosynthetic pathways involved in the synthesis of DIP. The sole known pathway for inositol biosynthesis in all other organisms is the conversion of d-glucose-6-phosphate to l-myo-inositol 1-phosphate (l-I-1-P) via l-myo-inositol 1-monophosphate (I-1-P) synthase and hydrolysis of I-1-P to myo-inositol via a specific phosphatase, I-1-P phosphatase (13, 14). Similar enzymes are likely to exist in methanogens. A logical pathway for the biosynthesis of DIP would then use myo-inositol and I-1-P as precursors. Activation of the I-1-P with CTP and condensation of the resultant CDP-inositol (CDP-I) with myo-inositol would generate DIP. As summarized in Fig. Fig.1,1, DIP biosynthesis requires four key enzymes: I-1-P synthase (step 1), I-1-P phosphatase (step 2), CTP:I-1-P cytidylyltransferase (step 3), and DIP synthase (step 4). The enzymes that catalyze steps 1 and 2 have been well studied in plants, yeasts, and mammalian tissues. However, the enzymes invoked for steps 3 and 4 are novel activities, although based on similar chemical transformations in cells. Open in a separate windowFIG. 1Proposed biosynthetic pathway for DIP showing the four key enzymatic activities. Based on similar transformations in other organisms, cofactors are indicated for several of the steps.This work describes the use of 31P nuclear magnetic resonance (NMR) and colorimetric assays to verify the existence of three of these activities in cell extracts of M. igneus. Specific labeling of DIP with [13C]pyruvate was also used to probe the DIP biosynthetic pathway. The pattern of 13C label incorporation from [3-13C]pyruvate and [2,3-13C]pyruvate coupled with the known stereochemistry of DIP provided evidence that M. igneus also has enzymes of the pentose phosphate pathway (transaldolase and transketolase) that scramble label in glucose-6-phosphate.  相似文献   

20.
A Mg2+-dependent, alkaline phosphatase has been isolated from mature pollen of Lilium longiflorum Thunb., cv. Ace and partially purified. It hydrolyzes 1l- and 1d-myo-inositol 1-phosphate, myo-inositol 2-phosphate, and β-glycerophosphate at rates decreasing in the order named. The affinity of the enzyme for 1l- and 1d-myo-inositol 1-phosphate is approximately 10-fold greater than its affinity for myo-inositol 2-phosphate. Little or no activity is found with phytate, d-glucose 6-phosphate, d-glucose 1-phosphate, d-fructose 1-phosphate, d-fructose 6-phosphate, d-mannose 6-phosphate, or p-nitrophenyl phosphate. 3-Phosphosphoglycerate is a weak competitive inhibitor. myo-Inositol does not inhibit the reaction. Optimal activity is obtained at pH 8.5 and requires the presence of Mg2+. At 4 millimolar, Co2+, Fe2+ or Mn2+ are less effective. Substantial inhibition is obtained with 0.25 molar Li+. With β-glycerophosphate as substrate the Km is 0.06 millimolar and the reaction remains linear at least 2 hours. In 0.1 molar Tris, β-glycerophosphate yields equivalent amounts of glycerol and inorganic phosphate, evidence that transphosphorylation does not occur.  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