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 31 毫秒
1.
SYNOPSIS. Continuous growth of one cell line (UCI variant) of Leishmania tarentolae was achieved in the absence of organic sulfur. These cells were able to use sodium sulfate, and, to a limited extent, sodium sulfite as their sole sulfur source and could utilize methionine sulfoxide in place of L-methionine. A related cell line (RU variant) was unable to grow in organic sulfate-free media nor could these cells utilize methionine sulfoxide. UCI promastigotes incorporated significant amounts of 35S sodium sulfate; killed cells did not take up the label. 35S incorporation was inhibited by sodium molybdate (5 × 10?4 M), sodium arsenite (5 × 10?4 M), 2,4-dinitrophenol (1 × 10?4 M), or KCN (5 × 10?4 M). RU promastigotes did not incorporate significant amounts of 35S sodium sulfate. Thin layer chromatographs of protein hydrolysates from UCI cells incubated in 35S sodium sulfate revealed several radio opaque spots, one of which had chromatographic properties of cystine. UCI variants of L. tarentolae were therefore capable of assimilatory sulfate reduction whereas RU cells lacked this ability.  相似文献   

2.
We have studied the regulation of sulfate assimilation by the gliding bacterium Cytophaga johnsonae in which 20% of the total sulfur is in the sulfornate moiety of sulfonolipid. Added cystine inhibited sulfate uptake and growth with cystine as sulfur source resulted in a repression of sulfate uptake. However, low concentrations of cystine preferentially repressed the terminal reactions of sulfate assimilation responsible for cysteine synthesis while allowing the transport and activation of sulfate for sulfonolipid synthesis. The significance of this novel pattern of regulation in bacteria is discussed.  相似文献   

3.
Brassica oleracea L. was rather insensitive to atmospheric H2S: growth was only negatively affected at ≥0.4 μl I?1. Shoots formed a sink for H2S and the uptake rate showed saturation kinetics with respect to the atmospheric concentration. The H2S uptake rate was high in comparison with other species, which may reflect the high sulfur need of Brassica. The net uptake of sulfate by roots of hydroponically grown plants was substantially reduced after one week of exposure to 0.25 μl l?1 H2S, indicating that plants switched in part from sulfate to H2S as sulfur source for plant growth. Plants were sulfur deficient after two weeks of sulfur deprivation, illustrated by reduced growth, which was more pronounced for shoots than for roots, and in enhanced shoot dry matter content. The latter could for the greater part be attributed to enhanced levels of soluble sugars and starch. Sulfur deficiency was further characterized by a low pigment content, extremely low levels of sulfate and water-soluble non-protein thiols, and by enhanced levels of nitrate and free amino acids, particularly in the shoots. Furthermore, sulfur deficient plants contained a lower total lipid content in shoots, whereas its content in roots was unaffected. The level of sulfolipids was decreased in both roots and shoots. When sulfur deprived plants were exposed to 0.25 μl I?1 H2S for one week, all sulfur deficiency symptoms were abolished and growth was restored. Furthermore, plants were able to grow with 0.4 μl I?1 H2S as the sole sulfur source. Water-soluble non-protein thiol content was enhanced in both shoots and roots of H2S exposed plants, irrespective of the sulfate supply to the roots, whereas plants grown with H2S as sole sulfur source contained very low sulfate levels. The interaction between atmospheric and pedospheric sulfur nutrition in plants is discussed.  相似文献   

4.
Metabolism of various sulfur compounds in Bacillus subtilis during growth and sporulation was investigated by use of tracer techniques, in an attempt to clarify the mechanism involved in the formation of cystine rich protein of the spore coat.

Methionine, homocysteine, cystathionine, cysteine and some inorganic sulfur compounds (sulfate, sulfite and thiosulfate) were utilized by this organism as sulfur sources for its growth and sporulation. Biosynthesis of methionine from sulfate during growth was more or less inhibited by the addition of cysteine, homocysteine or cystathionine to the culture.

It is suggested from these results that in Bacillus subtilis methionine is synthesized from sulfate through cysteine, cystathionine and homocysteine as is the case in Salmonella or Neurospora. The results also suggest that the metabolism of sulfur-containing amino acids in Bacillus subtilis is strongly regulated by methionine and homocysteine.  相似文献   

5.
The qualitative and quantitative aspects of elemental sulfur metabolization in wheat leaves and its effect upon photosynthetic metabolism were studied through the application of micronized sulfur upon the third leaf. Energy-dispersive x-ray analysis combined with scanning electron microscopy emphasized the existence of a sulfur peak associated with a strong potassium peak in the spectra of different tissue regions for treated leaves only, supplying an original evidence of sulfur uptake. Experiments with35S-labeled micronized sulfur showed that about 2% of the labeled S was absorbed and metabolized into cystine, methionine, glutathione, and sulfate. The close correlation between the excess of oxygen uptake and oxygen needs for sulfur oxidation in conjunction with the absence of hydrogen sulfide released by treated leaves support direct and fast oxidation of sulfur into sulfate according to a pathway still unclear but independent of photosynthetic CO2 metabolism in treated leaf. The mechanisms involved in the primary metabolism of element sulfur in wheat therefore appear to be different from those in fungi.  相似文献   

6.
The uptake and distribution of sulfate in BRASSICA OLERACEA, a species characterised by its high sulfate content in root and shoot, are coordinated and adjusted to the sulfur requirement for growth, even at external sulfate concentrations close to the K (m) value of the high-affinity sulfate transporters. Plants were able to grow normally and maintain a high sulfur content when grown at 5 or 10 microM sulfate in the root environment. Abundance of mRNAs for the high affinity sulfate transporters, BolSultr1;1 and BolSultr1;2, were enhanced at 相似文献   

7.
Previous studies of hyaluronan uptake and catabolism by lymph nodes indicated that the nodes might also add some HA of low molecular weight to the unabsorbed fraction that passes through from afferent to efferent lymph vessels.The ability of lymph nodes to synthesise HA and proteoglycans was therefore examined (i) by perfusion of [3H] acetate through an afferent lymph vessel in vivo, and recovery of labeled products from the efferent lymph vessel and from the node after perfusion; and (ii) by tissue culture of lymph nodes with [3H] acetate.Perfusion of lymph nodes with [3H] acetate in situ yielded: (a), in outflowing lymph, small amounts of chondroitin/dermatan sulfate within the first hour which continued to be produced for up to 24[emsp4 ]h; heparin in the second hour and HA in the third. In the nodes removed 17 to 19[emsp4 ]h later, equal amounts of hyaluronan and chondroitin/dermatan sulfate and heparan sulfate proteoglycans were detected. In the tissue culture of lymph nodes: (1) HA, heparin and proteoglycans of heparan sulfate and chondroitin/dermatan sulfate were released into the medium but in the cell extract only heparan sulfate proteoglycan was detected; and (ii) molecular weight of the released hyaluronan ranged widely but was mostly less than 4–5×105[emsp4 ]D; heparan sulfate proteoglycan was 2.8×104 to 9.4×105[emsp4 ]D; heparin 7.9×104[emsp4 ]D and chondroitin sulfate 1.3×104[emsp4 ]D, suggesting that the chondrotin sulfate were released from their proteoglycans core by enzymic degradation.It is concluded that lymph nodes can release HA, heparin, heparan sulfate and chondroitin/dermatan sulfate proteoglycans into efferent lymph but the amount of hyaluronan is likely to be small without immune or other stimulation and its molecular weight is lower than in other tissues.  相似文献   

8.
Photoheterotrophic and heterotrophic suspension cultures of tobacco (Nicotiana tabacum L.) were grown with 1 mM glutathione (reduced; GSH) as sole source of sulfur. Addition of sulfate to both cultures did not alter the rate of exponential growth, but affected the removal of GSH and sulfate in different ways. In photoheterotrophic suspensions, addition of sulfate caused a decline in the net uptake of GSH, whereas sulfate was taken up by the green cells immediately. In heterotrophic suspensions, however, addition of sulfate did not affect the net uptake of GSH and sulfate was only taken up by the cells after the GSH supply in the medium had been exhausted. Apparently, GSH uptake in photoheterotrophic cells is inhibited by sulfate, whereas sulfate uptake is inhibited by GSH in heterotrophic cells. The differences in the effect of GSH on sulfate uptake in photoheterotrophic and heterotrophic tobacco suspensions cannot be attributed to differences in the kinetic properties of sulfate carriers. In short-time transport experiments, both cultures took up sulfate almost entirely by an active-transport system as shown by experiments with metabolic inhibitors; sulfate transport of both cultures obeyed monophasic Michaelis-Menten kinetics with similar app. Km (photoheterotrophic cells: 16.0±2.0 M; heterotrophic cells: 11.8±1.8 M) and Vmax (photoheterotrophic cells: 323±50 nmol·min-1·g-1 dry weight; heterotrophic cells: 233±3 nmol·min-1·g-1 dry weight). Temperature- and pH-dependence of sulfate transport showed almost identical patterns. However, the cultures exhibited remarkable differences in the inhibition of sulfur influx by GSH in short-time transport experiments. Whereas 1 mM GSH inhibited sulfate transport into heterotrophic tobacco cells completely, sulfate transport into photoheterotrophic cells proceeded at more than two-thirds of its maximum velocity at this GSH concentration. The mode of action of GSH on sulfate transport in chloroplast-free tobacco cell does not appear to be direct: a 14-h exposure to 1 mM GSH was found to be necessary to completely block sulfate transport; a 4-h time of exposure did not affect this process. Consequently, glutathione does not seem to be a product of sulfur metabolism acting on sulfate-carrier entities by negative feedback control. When transferred to the whole plant, the observed differences in sulfate and glutathione influx into green and chloroplast-free cells may be interpreted as a regulatory device to prevent the uptake of excess sulfate by plants.Abbreviations DCCD N,N-dicyclohexylcarbodiimide - DNP dinitrophenol - DW dry weight - FW fresh weight - GSH reduced glutathione  相似文献   

9.
10.
Twenty-six different species of keratinophilic fungi were examined to determine their ability to utilize free cystine. Of the fungi tested, the majority metabolized free L-cystine in a glucose-peptone culture medium. Cystine was used as source of sulfur, and carbon and nitrogen as well. Excess sulfur was excreted into the culture fluid, as thiosulfate and sulfate, following oxidation. The rate of cystine oxidation varied with the different fungal strains, but was maximal for Graphium penicilloideus (88.5%). Low quantities of thiols were found in the medium. Cystine oxidation and inorganic thiosulfate excretion were found to correlate significantly (r = 0.94).  相似文献   

11.
Chlorella pyrenoidosa Chick (Emerson strain 3) utilizes thiosulfate for growth as effectively as sulfate, and more effectively than a variety of organic sulfur compounds containing sulfur in various oxidation states. Thiosulfates, differentially labeled with 35S in either the SH— or SO3 — sulfur moieties, were used to follow the incorporation of thiosulfate-sulfur into constituents of the insoluble fraction and of the soluble pools. Labeled sulfate was also used for purposes of comparison. Label from both sulfur atoms of thiosulfate and from sulfate is incorporated into the cysteine, homocysteine, and glutathione of the soluble pools, and into the methionine and cystine of protein in the insoluble fraction. Label from SO3-sulfur of thiosulfate is incorporated more slowly into protein methionine and cystine than label from the SH-sulfur. Moreover, the SO3-sulfur of thiosulfate is recovered largely as sulfate in both the soluble pools and the insoluble fraction, while only a trace of SH-sulfur is recovered as sulfate in either case. Consistent with this, the metabolism of the SO3-sulfur of thiosulfate more closely resembles the metabolism of sulfate. Thus it would appear that exogenous thiosulfate undergoes early dismutation in which the SO3-sulfur is preferentially oxidized, and the SH-sulfur is preferentially incorporated in a reduced state. These results are discussed in relation to the conversion of sulfate to thiosulfate by cell-free extracts of Chlorella previously described.  相似文献   

12.
Mature leaves of Ricinus communis fed with 35SO 4 2- in the light export labeled sulfate and reduced sulfur compounds by phloem transport. Only 1–2% of the absorbed radiosulfur is exported to the stem within 2–3 h, roughly 12% of 35S recovered was in reduced form. The composition of phloem translocate moving down the stem toward the root was determined from phloem exudate: 20–40% of the 35S moved in the form of organic sulfur compounds, however, the bulk of sulfur was transported as inorganic sulfate. The most important organic sulfur compound translocated was glutathione, carrying about 70% of the label present in the organic fraction. In addition, methionine and cysteine were involved in phloem sulfur transport and accounted for roughly 10%. Primarily, the reduced forms of both, glutathione and cysteine are prsent in the siever tubes.Abbreviations CySH cysteine - GSH glutathione - GSSG glutathione disulfide - NEM N-ethylmaleimide - CyS-SCy cystine  相似文献   

13.
Studies of microbial sulfate reduction have suggested that the magnitude of sulfur isotope fractionation varies with sulfate concentration. Small apparent sulfur isotope fractionations preserved in Archean rocks have been interpreted as suggesting Archean sulfate concentrations of <200 μm , while larger fractionations thereafter have been interpreted to require higher concentrations. In this work, we demonstrate that fractionation imposed by sulfate reduction can be a function of concentration over a millimolar range, but that nature of this relationship depends on the organism studied. Two sulfate‐reducing bacteria grown in continuous culture with sulfate concentrations ranging from 0.1 to 6 mm showed markedly different relationships between sulfate concentration and isotope fractionation. Desulfovibrio vulgaris str. Hildenborough showed a large and relatively constant isotope fractionation (34εSO4‐H2S ? 25‰), while fractionation by Desulfovibrio alaskensis G20 strongly correlated with sulfate concentration over the same range. Both data sets can be modeled as Michaelis–Menten (MM)‐type relationships but with very different MM constants, suggesting that the fractionations imposed by these organisms are highly dependent on strain‐specific factors. These data reveal complexity in the sulfate concentration–fractionation relationship. Fractionation during MSR relates to sulfate concentration but also to strain‐specific physiological parameters such as the affinity for sulfate and electron donors. Previous studies have suggested that the sulfate concentration–fractionation relationship is best described with a MM fit. We present a simple model in which the MM fit with sulfate concentration and hyperbolic fit with growth rate emerge from simple physiological assumptions. As both environmental and biological factors influence the fractionation recorded in geological samples, understanding their relationship is critical to interpreting the sulfur isotope record. As the uptake machinery for both sulfate and electrons has been subject to selective pressure over Earth history, its evolution may complicate efforts to uniquely reconstruct ambient sulfate concentrations from a single sulfur isotopic composition.  相似文献   

14.
Abdin MZ  Akmal M  Ram M  Nafis T  Alam P  Nadeem M  Khan MA  Ahmad A 《Protoplasma》2011,248(3):591-600
Lycopersicon esculantum sulfate transporter gene (LeST 1.1) encodes a high-affinity sulfate transporter (HAST) located in root epidermis. In this study, the LeST 1.1 gene was constitutively expressed in Indian mustard (Brassica juncea cv. Pusa Jai Kisan). Transgenic as well as untransformed plants were grown in sulfur-insufficient (25 and 50 μM) and sulfur-sufficient (1,000 μM) conditions for 30 days. Two-fold increase was noticed in the sulfate uptake rate of transgenic plants grown in both sulfur-insufficient and -sufficient conditions as compared to untransformed plants. The transgenic B. juncea plants were able to accumulate higher biomass and showed improved sulfur status even in sulfur-insufficient conditions when compared with untransformed plants. Chlorophyll content, ATP sulfurylase activity and protein content were also higher in transgenic plants than untranformed plants under sulfur-insufficient conditions. Our results, thus, clearly indicate that constitutive expression of LeST 1.1 gene in B. juncea had led to enhanced capacity of sulfur uptake and assimilation even in sulfur-insufficient conditions. This approach can also be used in other crops to enhance their sulfate uptake and assimilation potential under S-insufficient conditions.  相似文献   

15.
Cystine catabolism in mycelia of Microsporum gypseum,a dermatophytic fungus   总被引:3,自引:0,他引:3  
The fate of 35S label was studied during cystine degradation by mycelia of the dermatophytic fungus Microsporum gypseum. Excess free cystine in the medium was readily taken up and its sulfur moiety excreted as inorganic sulfate and sulfite. At intervals after 3–60 min of incubation with 35S cystine the products of cystine catabolism were extracted from the mycelia by boiling water and separated by thin layer chromatography and electrophoresis. A total of 10 sulfur-containing compounds were identified, and their relative radioactivity was assessed. After 3 min the mycelia contained, in addition to cystine, labeled cysteine and particularly cysteine sulfinic acid which was accompanied by a smaller amount of cysteic acid. Later on, oxidized and reduced glutathione, inorganic sulfate and taurine appeared consecutively. In all extracts, small amounts of labeled S-sulfocysteine were found, not, however, sulfite.The results suggest that the intermediates of cysteine degradation in the fungal mycelia are cysteine, cysteine sulfinate, unstable sulfinylpyruvate, sulfite and sulfate, i.e., that the catabolic pattern is similar to that of higher organisms.The formation and the role of S-sulfocysteine, cysteic acid, and of taurine is not yet completely understood, although certainly autoxidative processes are involved in the formation of the latter two compounds, and sulfitolysis in that of the former compound.  相似文献   

16.

In the present study, coupled stable sulfur and oxygen isotope fractionation during elemental sulfur disproportionation according to the overall reaction: 4H2O + 4S? → 3H2S + SO4 2 ? + 2H+, was experimentally investigated for the first time using a pure culture of the sulfate reducer Desulfobulbus propionicus at 35?C. Bacterial disproportionation of elemental sulfur is an important process in the sulfur cycle of natural surface sediments and leads to the simultaneous formation of sulfide and sulfate. A dual-isotope approach considering both sulfur and oxygen isotope discrimination has been shown to be most effective in evaluating specific microbial reactions. The influence of iron- and manganese bearing-solids (Fe(II)CO3, Fe(III)OOH, Mn(IV)O2) acting in natural sediments as scavengers for hydrogen sulfide, was considered, too. Disproportionation of elemental sulfur was observed in the presence of iron solids at a cell-specific sulfur disproportionation rate of about 10? 9.5± 0.4 μ mol S? cell? 1 h? 1. No disproportionation, however, was observed with MnO2. In the presence of iron solids, newly formed sulfate was enriched in 18 O compared to water by about +21‰ (≡ ? H2O ), in agreement with a suggested oxygen isotope exchange via traces of intra- or extracellular sulfite that is formed as a disproportionation intermediate. Dissolved sulfate was also enriched in 34S compared to elemental sulfur by up to +35%. Isotope fractionation by Desulfobulbus propionicusis highest for all disproportionating bacteria investigated, so far, and may impact on the development of isotope signals at the redox boundary of surface sediments.  相似文献   

17.
Sulfate transport capacity was not regulated by cysteine, methionine, or glutathione in Pseudomonas halodurans, but growth on sulfate or thiosulfate suppressed transport. Subsequent sulfur starvation of cultures grown on all sulfur sources except glutathione stimulated uptake. Only methionine failed to regulate sulfate transport in Alteromonas luteo-violaceus, and sulfur starvation of all cultures enhanced transport capacity. During sulfur starvation of sulfate-grown cultures of both bacteria, the increase in transport capacity was mirrored by a decrease in the low-molecular-weight organic sulfur pool. Little metabolism of endogenous inorganic sulfate occurred. Cysteine was probably the major regulatory compound in A. luteo-violaceus, but an intermediate in sulfate reduction, between sulfate and cysteine, controlled sulfate transport in P. halodurans. Kinetic characteristics of sulfate transport in the marine bacteria were similar to those of previously reported nonmarine systems in spite of significant regulatory differences. Sulfate and thiosulfate uptake in P. halodurans responded identically to inhibitors, were coordinately regulated by growth on various sulfur compounds and sulfur starvation, and were mutually competitive inhibitors of transport, suggesting that they were transported by the same mechanism. The affinity of P. halodurans for thiosulfate was much greater than for sulfate.  相似文献   

18.
J. Edelman  A. D. Hanson 《Planta》1971,101(2):122-132
Summary Free space invertase activities were determined in carrot callus strains CRT1 and CRT2 grown under conditions in which sucrose suppression of chlorophyll synthesis occurred in CRT1 but not CRT2. CRT2 possessed a high free space acid invertase activity (pH optimum 5.0 Km for sucrose 3.1×10-3M) while CRT1 lacked this enzyme. [U-14C] sucrose introduced into the free space of calluses was rapidly inverted by CRT2, but not by CRT1.Despite their different invertase levels, CRT1 and CRT2 showed similar sucrose uptake rates and took up [U-14C-glucosyl] sucrose and [5-T-glucosyl] sucrose from external bathing media essentially without prior inversion.It is concluded that acid invertase in callus tissue relieves the suppression of chlorophyll synthesis caused by sucrose in the free space. The invertase may in some circumstances hydrolyse sucrose before uptake, but is not an essential part of the sucrose uptake mechanism in carrot tissue cultures.  相似文献   

19.
Soil management practices that involve additions of organic materials may influence plant sulfur availability in highly-weathered, acid soils. This study evaluated the effects of organic additions on sulfate adsorption and sulfur availability in a limed (3,4 t ha-1) and unlimed Typic Haplustox soil of the Cerrado Region of Brazil. In unlimed soil, the proportion of applied sulfate (600 kg S ha-1 as gypsum) that was adsorbed temporarily decreased over two cropping seasons by incorporation of 10 t dry matter ha-1 crop-1 of guinea grass (Panicum maximum Jacq.) but not when a similar quantity of a tropical legume, feijâo de porco (Canavalia ensiformis L.), was added. Liming reduced sulfate adsorption and resulted in sulfate leaching to a depth of 30 to 45 cm. Both plant materials temporarily reduced sulfate adsorption in laboratory studies when added to an unlimed soil at a rate equivalent to 40 t ha-1. Analysis of soil properties affected by organic additions and liming showed significant correlations between sulfate adsorption and soil pH, extractable aluminum, calcium and magnesium, and surface charge. Maize dry matter yields increased by 1.3 to 3.5 t ha-1 with addition of both organic materials. However, only the feijâo de porco treatment resulted in increases in sulfur uptake for the years in which organic materials were applied. Determining the effects of organic material additions on plant sulfur availability is complicated by the combined effects of sulfur mineralization, sulfate adsorption, and the plant's ability to utilize adsorbed subsoil sulfate.Joint contribution of Cornell University and CPAC-EM- BRAPA. This research was supported by USAID through the Title XII CRSP subgrant SM-CRSP-10 from North Carolina State University  相似文献   

20.
All 16 strains of dermatophytes investigated utilized cystine (added to the gelatin medium) as a source of sulfur and also of carbon and nitrogen. Excess sulfur oxidized and excreted to the medium, primarily as inorganic sulfate. Six strains used up all cystine and excreted more than 90% stoichiometric amount of sulfur. Cystine utilization proceeded in parallel with the development of the culture and was terminated during the stationary phase or as late as in the autolytic phase. Other strains did not use up cystine completely and excreted 17-70% sulfur in the oxidized form. In addition to sulfate, sulfite was always produced during the initial growth phases and in poorly growing strains. Free sulfite was only rarely detected; it usually reacted with the residual cystine yielding S-sulfocysteine that was also used up later. Specific features of cystine metabolism (known from Microsporum gypseum) are generally valid in dermatophytes.  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