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 46 毫秒
1.
The osmoprotectant 3-dimethylsulfoniopropionate (DMSP) occurs in Gramineae and Compositae, but its synthesis has been studied only in the latter. The DMSP synthesis pathway was therefore investigated in the salt marsh grass Spartina alterniflora Loisel. Leaf tissue metabolized supplied [35S]methionine (Met) to S-methyl-l-Met (SMM), 3-dimethylsulfoniopropylamine (DMSP-amine), and DMSP. The 35S-labeling kinetics of SMM and DMSP-amine indicated that they were intermediates and, consistent with this, the dimethylsulfonium moiety of SMM was shown by stable isotope labeling to be incorporated as a unit into DMSP. The identity of DMSP-amine, a novel natural product, was confirmed by both chemical and mass-spectral methods. S. alterniflora readily converted supplied [35S]SMM to DMSP-amine and DMSP, and also readily converted supplied [35S]DMSP-amine to DMSP; grasses that lack DMSP did neither. A small amount of label was detected in 3-dimethylsulfoniopropionaldehyde (DMSP-ald) when [35S]SMM or [35S]DMSP-amine was given. These results are consistent with the operation of the pathway Met → SMM → DMSP-amine → DMSP-ald → DMSP, which differs from that found in Compositae by the presence of a free DMSP-amine intermediate. This dissimilarity suggests that DMSP synthesis evolved independently in Gramineae and Compositae.  相似文献   

2.
In the flowering plant Wollastonia biflora (L.) DC. the first step in 3-dimethylsulfoniopropionate (DMSP) synthesis is conversion of methionine to S-methylmethionine (SMM) and the last is oxidation of 3-dimethylsulfoniopropionaldehyde (DMSP-ald) (F. James, L. Paquet, S.A. Sparace, D.A. Gage, A.D. Hanson [1995] Plant Physiol 108: 1439-1448). DMSP-ald was shown to undergo rapid, spontaneous decomposition to dimethylsulfide and acrolein. However, it was stable enough (half-life [greater than or equal to] 1 h) in tertiary amine buffers to use as a substrate for enzyme assays. A dehydrogenase catalyzing DMSP-ald oxidation was detected in extracts of W. biflora mesophyll protoplasts. This enzyme had a high affinity for DMSP-ald (Km = 1.5 [mu]M), was subject to substrate inhibition, preferred NAD to NADP, and was immunologically related to plant betaine aldehyde dehydrogenases. After fractionation of protoplast lysates, [greater than or equal to]90% of DMSP-ald dehydrogenase activity was recovered from the chloroplast stromal fraction, whereas the enzyme that mediates SMM synthesis, S-adenosylmethionine:methionine S-methyltransferase, was found exclusively in the cytosolic fraction. Immunohistochemical analysis confirmed that the S-methyltransferase was cytosolic. Intact W. biflora chloroplasts were able to metabolize supplied [35S]SMM to [35S]DMSP. These findings indicate that SMM is made in the cytosol, imported into the chloroplast, and there converted successively to DMSP-ald and DMSP.  相似文献   

3.
A D Hanson  J Rivoal  L Paquet    D A Gage 《Plant physiology》1994,105(1):103-110
The compatible solute 3-dimethylsulfoniopropionate (DMSP) is accumulated by certain salt-tolerant flowering plants and marine algae. It is the major biogenic precursor of dimethylsulfide, an important sulfur-containing trace gas in the atmosphere. DMSP biosynthesis was investigated in Wollastonia biflora (L.) DC. [= Wedelia biflora (L.) DC., Melanthera biflora (L.) Wild, Asteraceae]. After characterizing DMSP and glycine betaine accumulation in three diverse genotypes, a glycine betaine-free genotype was chosen for radiotracer and stable isotope-labeling studies. In discs from young leaves, label from [U-14C]methionine was readily incorporated into the dimethylsulfide and acrylate moieties of DMSP. This establishes that DMSP is derived from methionine by deamination, decarboxylation, oxidation, and methylation steps, without indicating their order. Five lines of evidence indicated that methylation is the first step in the sequence, not the last. (a) In pulse-chase experiments with [14C]methionine, S-methylmethionine (SMM) had the labeling pattern expected of a pathway intermediate, whereas 3-methylthiopropionate (MTP) did not. (b) [14C]SMM was efficiently converted to DMSP but [14C]MTP was not. (c) The addition of unlabeled SMM, but not of MTP, reduced the synthesis of [14C]DMSP from [14C]methionine. (d) The dimethylsulfide group of [13CH3,C2H3]SMM was incorporated as a unit into DMSP. (e) When [C2H3,C2H3]SMM was given together with [13CH3]methionine, the main product was [C2H3,C2H3]DMSP, not [13CH3,C2H3]DMSP or [13CH3,13CH3]DMSP. The stable isotope labeling results also show that the SMM cycle does not operate at a high level in W. biflora leaves.  相似文献   

4.
Rhodes D  Gage DA  Cooper A  Hanson AD 《Plant physiology》1997,115(4):1541-1548
Leaves of Wollastonia biflora (L.) DC. synthesize the osmoprotectant 3-dimethylsulfoniopropionate (DMSP) from methionine via S-methylmethionine (SMM) and 3-dimethylsulfoniopropionaldehyde (DMSP-ald); no other intermediates have been detected. To test whether the amino group of SMM is lost by transamination or deamination, [methyl-2H3,15N]SMM was supplied to leaf discs, and 15N-labeling of amino acids was monitored, along with synthesis of [2H3]DMSP. After short incubations more 15N was incorporated into glutamate than into other amino acids, and the 15N abundance in glutamate exceeded that in the amide group of glutamine (Gln). This is more consistent with transamination than deamination, because deamination would be predicted to give greater labeling of Gln amide N due to reassimilation, via Gln synthetase, of the 15NH4+ released. This prediction was borne out by control experiments with 15NH4Cl. The transamination product of SMM, 4-dimethylsulfonio-2-oxobutyrate (DMSOB), is expected to be extremely unstable. This was confirmed by attempting to synthesize it enzymatically from SMM using L-amino acid oxidase or Gln transaminase K and from 4-methylthio-2-oxobutyrate using methionine S-methyltransferase. In each case, the reaction product decomposed rapidly, releasing dimethylsulfide. The conversion of SMM to DMSP-ald is therefore unlikely to involve a simple transamination that generates free DMSOB. Plausible alternatives are that DMSOB is channeled within a specialized transaminase-decarboxylase complex or that it exists only as the bound intermediate of a single enzyme catalyzing an unusual transamination-decarboxylation reaction.  相似文献   

5.
Many marine algae produce 3-dimethylsulfoniopropionate (DMSP), a potent osmoprotective compound whose degradation product dimethylsulfide plays a central role in the biogeochemical S cycle. Algae are known to synthesize DMSP via the four-step pathway, l-Met → 4-methylthio-2-oxobutyrate → 4-methylthio-2-hydroxybutyrate → 4-dimethylsulfonio-2-hydroxy-butyrate (DMSHB) → DMSP. Substrate-specific enzymes catalyzing the first three steps in this pathway were detected and partially characterized in cell-free extracts of the chlorophyte alga Enteromorpha intestinalis. The first is a 2-oxoglutarate-dependent aminotransferase, the second an NADPH-linked reductase, and the third an S-adenosylmethionine-dependent methyltransferase. Sensitive radiometric assays were developed for these enzymes, and used to show that their activities are high enough to account for the estimated in vivo flux from Met to DMSP. The activities of these enzymes in other DMSP-rich chlorophyte algae were at least as high as those in E. intestinalis, but were ≥20-fold lower in algae without DMSP. The reductase and methyltransferase were specific for the d-enantiomer of 4-methylthio-2-hydroxybutyrate in vitro, and both the methyltransferase step and the step(s) converting DMSHB to DMSP were shown to prefer d-enantiomers in vivo. The intermediate DMSHB was shown to act as an osmoprotectant, which indicates that the first three steps of the DMSP synthesis pathway may be sufficient to confer osmotolerance.  相似文献   

6.
The fraction of planktonic heterotrophic bacteria capable of incorporating dissolved dimethylsulfoniopropionate (DMSP) and leucine was determined at two coastal sites by microautoradioagraphy (AU). In Gulf of Mexico seawater microcosm experiments, the proportion of prokaryotes that incorporated sulfur from [35S]DMSP ranged between 27 and 51% of 4′,6-diamidino-2-phenylindole (DAPI)-positive cells, similar to or slightly lower than the proportion incorporating [3H]leucine. In the northwest Mediterranean coast, the proportion of cells incorporating sulfur from [35S]DMSP increased from 5 to 42% from January to March, coinciding with the development of a phytoplankton bloom. At the same time, the proportion of cells incorporating [3H]leucine increased from 21 to 40%. The combination of AU and fluorescence in situ hybridization (FISH) revealed that the Roseobacter clade (α-proteobacteria) accounted for 13 to 43% of the microorganisms incorporating [35S]DMSP at both sampling sites. Significant uptake of sulfur from DMSP was also found among members of the γ-proteobacteria and Cytophaga-Flavobacterium groups. Roseobacter and γ-proteobacteria exhibited the highest percentage of DAPI-positive cells incorporating 35S from DMSP (around 50%). Altogether, the application of AU with [35S]DMSP combined with FISH indicated that utilization of S from DMSP is a widespread feature among active marine bacteria, comparable to leucine utilization. These results point toward DMSP as an important substrate for a broad and diverse fraction of marine bacterioplankton.  相似文献   

7.
Kocsis MG  Hanson AD 《Plant physiology》2000,123(3):1153-1161
3-Dimethylsulfoniopropionate (DMSP) is an osmoprotectant accumulated by the cordgrass Spartina alterniflora and other salt-tolerant plants. Previous in vivo isotope tracer and metabolic modeling studies demonstrated that S. alterniflora synthesizes DMSP via the route S-methyl-Met --> 3-dimethylsulfoniopropylamine (DMSP-amine) --> 3-dimethylsulfoniopropionaldehyde --> DMSP and indicated that the first reaction requires a far higher substrate concentration than the second to attain one-half-maximal rate. As neither of these reactions is known from other organisms, two novel enzymes are predicted. Two corresponding activities were identified in S. alterniflora leaf extracts using specific radioassays. The first, S-methyl-Met decarboxylase (SDC), strongly prefers the L-enantiomer of S-methyl-Met, is pyridoxal 5'-phosphate-dependent, generates equimolar amounts of CO(2) and DMSP-amine, and has a high apparent K(m) (approximately 18 mM) for its substrate. The second enzyme, DMSP-amine oxidase (DOX), requires O(2) for activity, shows an apparent K(m) for DMSP-amine of 1.8 mM, and is not accompanied by DMSP-amine dehydrogenase or transaminase activity. Very little SDC or DOX activity was found in grasses lacking DMSP. These data indicate that SDC and DOX are the predicted novel enzymes of DMSP synthesis.  相似文献   

8.
It has been shown that marine algae produce the compatible solute dimethylsulfoniopropionate (DMSP) from methionine (Met) via four enzymatic reactions in which the third step, synthesis of 4‐dimethylsulfonio‐2‐hydroxy‐butyrate (DMSHB) from 4‐methylthio‐2‐hydroxybutyrate (MTHB), is the committing step. However, regulation of the biosynthetic pathways and transport properties of DMSP is largely unknown. Here, the effects of sulfur and sodium concentrations on the uptake and synthesis of DMSHB and DMSP were examined in a sterile mutant of Ulva pertusa Kjellm. Sulfur deficiency increased the activity of the sulfur assimilation enzyme O‐acetyl serine sulfhydrylase but decreased the MTHB S‐methyltransferase activity, suggesting the preferential utilization of sulfur atoms for Met metabolites other than DMSP. Uptake of DMSP and DMSHB was enhanced by S deficiency. High salinity enhanced the MTHB S‐methyltransferase activity as well as the uptake of DMSHB. The MTHB S‐methyltransferase activity was inhibited by its product DMSP. These data demonstrate the importance of MTHB S‐methyltransferase activity and uptake of DMSHB for the regulation of DMSP.  相似文献   

9.
Wollastonia biflora (L.) DC. plants accumulate the osmoprotectant 3-dimethylsulfoniopropionate (DMSP), particularly when salinized. DMSP is known to be synthesized in the chloroplast from S-methylmethionine (SMM) imported from the cytosol, but the sizes of the chloroplastic and extrachloroplastic pools of these compounds are unknown. We therefore determined DMSP and SMM in mesophyll protoplasts and chloroplasts. Salinization with 30% (v/v) artificial seawater increased protoplast DMSP levels from 4.6 to 6.0 μmol mg−1 chlorophyll (Chl), and chloroplast levels from 0.9 to 1.9 μmol mg−1 Chl. The latter are minimum values because intact chloroplasts leaked DMSP during isolation. Correcting for this leakage, it was estimated that in vivo about one-half of the DMSP is chloroplastic and that stromal DMSP concentrations in control and salinized plants are about 60 and 130 mm, respectively. Such concentrations would contribute significantly to chloroplast osmoregulation and could protect photosynthetic processes from stress injury. SMM levels were measured using a novel mass-spectrometric method. About 40% of the SMM was located in the chloroplast in unsalinized W. biflora plants, as was about 80% in salinized plants; the chloroplastic pool in both cases was approximately 0.1 μmol mg−1 Chl. In contrast, ≥85% of the SMM was extrachloroplastic in pea (Pisum sativum L.) and spinach (Spinacia oleracea L.), which lack DMSP. DMSP synthesis may be associated with enhanced accumulation of SMM in the chloroplast.  相似文献   

10.
Soybean (Glycine max L.) plants were grown with sulfate at 2 (S2) or 20 [mu]M (S20) and treated with [35S]sulfate between d 36 and 38. Growth was continued with or without 20 [mu]M sulfate (i.e. S2 -> S0, S2 -> S20, etc.). When the leaves of S20 -> S20 plants were 70% expanded, they exported S and 35S label from the soluble fraction, largely as sulfate, to new expanding leaves. However, 35S label in the insoluble fraction was not remobilized. Very little of the 35S label in the soluble fraction of the leaves of S20 -> S0 plants was redistributed; most was incorporated into the insoluble fraction. The low levels of S remobilization from the insoluble fraction were attributed to the high level of N in the nutrient solution (15 mM). Most of the 35S label in S2 plants at d 38 occurred in the soluble fraction of the roots. In S2 -> S0 plants the 35S label was incorporated into the insoluble fraction of the roots, but in S2 -> S20 plants 35S label was rapidly exported to leaves 3 to 6. It was concluded that the soluble fraction of roots contains a small metabolically active pool of S and another larger pool that is in slow equilibrium with the small pool.  相似文献   

11.
The biosynthetic pathway of sulfoquinovosyldiacylglycerol (SQDG) was investigated using groundnut (Arachis hypogaea) leaf discs and 35S-labeled precursors. [35S]SO4(2-) was actively taken up by the leaf discs and rapidly incorporated into SQDG. After 2 h, 1.5% of the [35S]SO4(2-) added to the incubation medium was taken up, of which 28% was incorporated into SQDG. The methanol-water phases of the lipid extracts of the leaf discs were analyzed for the 35S-labeled intermediates. Up to 2 h of incubation, cysteic acid, 3-sulfopyruvate, 3-sulfolactate, 3-sulfolactaldehyde, and sulfoquinovose (SQ) which have been proposed as intermediates [Davies et al. (1966) Biochem. J. 98, 369-373] were not labeled. Only a negligible amount of radioactivity was observed in these compounds after incubation for 4 h and more. Addition of sodium molybdate inhibited the uptake of [35S]SO4(2-) as well as its incorporation into SQDG by the leaf discs, suggesting that 3'-phosphoadenosine-5'-phosphosulfate may be involved in the biosynthesis of SQDG. Addition of unlabeled cysteic acid to the incubation medium enhanced the uptake of [35S]SO4(2-) but did not affect its incorporation into SQDG. 35S-labeled cysteic acid was taken up by the leaf discs and metabolized to sulfoacetic acid but not incorporated into SQ or SQDG. These results show that cysteic acid is not an intermediate in SQDG biosynthesis. [35S]SQ was taken up by the leaf discs and incorporated into SQDG in a time-dependent manner. [35S]Sulfoquinovosylglycerol was also taken up by the leaf discs but not incorporated into SQDG. It is concluded that SQDG is not biosynthesized by the proposed sulfoglycolytic pathway in higher plants. Though [35S]SQ was converted to SQDG, the rates are much lower compared to [35S]SO4(2-) incorporation, which suggests that a more direct pathway involving sulfonation of a lipid precursor may exist in higher plants.  相似文献   

12.
Adenosine-5′-phosphosulfate (APS) and adenosine-3′-phosphate 5′-phosphosulfate (PAPS) have been used as precursors of sulfoquinovosyldiacylglycerol (SQDG) in intact chloroplasts incubated in the dark. Competition studies demonstrated APS was preferred over PAPS and SO42−. Rates of SQDG synthesis up to 3 nanomoles per milligram of chlorophyll per hour were observed when [35S]APS and appropriate cofactors were supplied to chloroplasts incubated in the dark. The pH optimum for utilization of APS was 7.0. The incorporation was linear for at least 30 minutes. ATP and UTP stimulated the incorporation of sulfur from APS into SQDG, but the most stimulatory additions were DHAP and glycerol-3-P. The concentration curve for APS showed a maximum at 20 micromolar in the absence of DHAP and 30 micromolar in the presence of DHAP. The optimum concentration of DHAP for conversion of APS into SQDG was 2 millimolar. Rates of synthesis up to 4 nanomoles per milligram of chlorophyll per hour were observed when [35S]PAPS was the sulfur donor and appropriate cofactors were supplied to chloroplasts. Optimal rates for conversion of sulfur from PAPS into SQDG occurred with concentrations of DHAP between 5 and 10 millimolar. DHAP was by far the most effective cofactor, although ATP and UTP also stimulated the utilization of PAPS for SQDG biosynthesis. In general, triose phosphates, including glycerol-3-P were not effective cofactors for SQDG biosynthesis.  相似文献   

13.
Organic sulfur compounds are present in all aquatic systems, but their use as sources of sulfur for bacteria is generally not considered important because of the high sulfate concentrations in natural waters. This study investigated whether dimethylsulfoniopropionate (DMSP), an algal osmolyte that is abundant and rapidly cycled in seawater, is used as a source of sulfur by bacterioplankton. Natural populations of bacterioplankton from subtropical and temperate marine waters rapidly incorporated 15 to 40% of the sulfur from tracer-level additions of [(35)S]DMSP into a macromolecule fraction. Tests with proteinase K and chloramphenicol showed that the sulfur from DMSP was incorporated into proteins, and analysis of protein hydrolysis products by high-pressure liquid chromatography showed that methionine was the major labeled amino acid produced from [(35)S]DMSP. Bacterial strains isolated from coastal seawater and belonging to the alpha-subdivision of the division Proteobacteria incorporated DMSP sulfur into protein only if they were capable of degrading DMSP to methanethiol (MeSH), whereas MeSH was rapidly incorporated into macromolecules by all tested strains and by natural bacterioplankton. These findings indicate that the demethylation/demethiolation pathway of DMSP degradation is important for sulfur assimilation and that MeSH is a key intermediate in the pathway leading to protein sulfur. Incorporation of sulfur from DMSP and MeSH by natural populations was inhibited by nanomolar levels of other reduced sulfur compounds including sulfide, methionine, homocysteine, cysteine, and cystathionine. In addition, propargylglycine and vinylglycine were potent inhibitors of incorporation of sulfur from DMSP and MeSH, suggesting involvement of the enzyme cystathionine gamma-synthetase in sulfur assimilation by natural populations. Experiments with [methyl-(3)H]MeSH and [(35)S]MeSH showed that the entire methiol group of MeSH was efficiently incorporated into methionine, a reaction consistent with activity of cystathionine gamma-synthetase. Field data from the Gulf of Mexico indicated that natural turnover of DMSP supplied a major fraction of the sulfur required for bacterial growth in surface waters. Our study highlights a remarkable adaptation by marine bacteria: they exploit nanomolar levels of reduced sulfur in apparent preference to sulfate, which is present at 10(6)- to 10(7)-fold higher concentrations.  相似文献   

14.
Ma JF  Nomoto K 《Plant physiology》1993,102(2):373-378
The biosynthesis of mugineic acids was studied by feeding 2H- or 13C-labeled compounds to water-cultured roots in several gramineous plants. The fate of labeled compounds was monitored by using 2H- and 13C-nuclear magnetic resonance. On investigating the proton changes during biosynthesis by feeding D,L-[3,3,4,4-d4]-methionine (98.6% 2H), 2H-labeled 2[prime]-deoxymugineic, mugineic, and 3-epihydroxymugineic acids were isolated from root washings of wheat (Triticum aestivum L. cv Minori), barley (Hordeum vulgare L. cv Minorimugi), and beer barley (Hordeum vulgare L. cv AM Nijo Tochigi), respectively. The 2H-nuclear magnetic resonance study indicated that 12 deuteriums were incorporated into the labeled 2[prime]-deoxymugineic acid, suggesting that three molecules of L-[3,3,4,4-d4]methionine were combined. In comparison, one of the deuteriums at C-2[prime] position in the mugineic acid, and one each of the deuteriums at C-2[prime] and C-3 positions in the 3-epihydroxymugineic acid, were lost. However, all other deuteriums were incorporated in a manner similar to that of the labeled 2[prime]-deoxymugineic acid. When [1,4[prime],4"-13C3]2[prime]-deoxymugineic acid (20% 13C) was fed to oat roots (Avena sativa L. cv Amuri II), avenic acid A, which was 13C enriched at the corresponding positions, was obtained. These results revealed that L-methionine was the precursor for all these mugineic acids and that cleavage of the azetidine ring or hydroxylation of the 2[prime]-deoxymugineic acid produced two related biosynthetic pathways in different gramineous plant species: L-methionine -> 2[prime]-deoxymugineic acid -> avenic acid A in oat; and L-methionine -> 2[prime]-deoxymugineic acid -> mugineic acid -> 3-epihydroxymugineic acid in barley and beer barley.  相似文献   

15.
Cytokinins of the urea and adenine type induced leaf abscission in young cotton (Gossypium hirsutum) plants in the following order of activity: N-phenyl-N′-1,2,3-thiadiazol-5-ylurea (thidiazuron) » N-phenyl-N′-(2-chloro-4-pyridyl)urea > isopentenyladenine ≥ 6-benzyladenine > zeatin = dihydrozeatin > kinetin. It is suggested that ethylene production is implicated in this response because it was stimulated by the compounds in cotton leaf discs with nearly the same effectiveness. Moreover, similar to thidiazuron (JC Suttle [1985] Plant Physiol 78: 272-276), isopentenyladenine-induced defoliation was inhibited by aminoethoxyvinylglycine, and the effect was restored by 1-aminocyclopropane-1-carboxylic acid.  相似文献   

16.
Sunarpi  Anderson JW 《Plant physiology》1996,110(4):1151-1157
Soybean (Glycine max L.) plants were grown in nutrient solution containing 10 [mu]M sulfate and were treated at various times with [35S]sulfate for 48 h. Growth was then continued in unlabeled solution. The sulfur content of each leaf increased rapidly until it was about 40% expanded; small, additional increases occurred until the leaf was about 70% expanded after which the sulfur content decreased by about 50%. Leaves that were about 60 to 70% expanded during the pulse were strongly labeled but then underwent a significant loss of 35S label. Leaves that were in the early stages of expansion imported little 35S label during the pulse but acquired 35S label during the chase period as they expanded (i.e. redistribution). Most of the redistributed 35S label was derived from other leaves. The rates of both sulfur import and sulfur export by a leaf were greatest at about 70% expansion. Leaves that acquired 35S label during early development retained a much higher proportion of their label than leaves that were more developed, suggesting that the sulfur acquired by leaves during early development is preferentially incorporated into a pool that is less mobile than the sulfur acquired in the later stages of leaf growth.  相似文献   

17.
Organic sulfur compounds are present in all aquatic systems, but their use as sources of sulfur for bacteria is generally not considered important because of the high sulfate concentrations in natural waters. This study investigated whether dimethylsulfoniopropionate (DMSP), an algal osmolyte that is abundant and rapidly cycled in seawater, is used as a source of sulfur by bacterioplankton. Natural populations of bacterioplankton from subtropical and temperate marine waters rapidly incorporated 15 to 40% of the sulfur from tracer-level additions of [35S]DMSP into a macromolecule fraction. Tests with proteinase K and chloramphenicol showed that the sulfur from DMSP was incorporated into proteins, and analysis of protein hydrolysis products by high-pressure liquid chromatography showed that methionine was the major labeled amino acid produced from [35S]DMSP. Bacterial strains isolated from coastal seawater and belonging to the α-subdivision of the division Proteobacteria incorporated DMSP sulfur into protein only if they were capable of degrading DMSP to methanethiol (MeSH), whereas MeSH was rapidly incorporated into macromolecules by all tested strains and by natural bacterioplankton. These findings indicate that the demethylation/demethiolation pathway of DMSP degradation is important for sulfur assimilation and that MeSH is a key intermediate in the pathway leading to protein sulfur. Incorporation of sulfur from DMSP and MeSH by natural populations was inhibited by nanomolar levels of other reduced sulfur compounds including sulfide, methionine, homocysteine, cysteine, and cystathionine. In addition, propargylglycine and vinylglycine were potent inhibitors of incorporation of sulfur from DMSP and MeSH, suggesting involvement of the enzyme cystathionine γ-synthetase in sulfur assimilation by natural populations. Experiments with [methyl-3H]MeSH and [35S]MeSH showed that the entire methiol group of MeSH was efficiently incorporated into methionine, a reaction consistent with activity of cystathionine γ-synthetase. Field data from the Gulf of Mexico indicated that natural turnover of DMSP supplied a major fraction of the sulfur required for bacterial growth in surface waters. Our study highlights a remarkable adaptation by marine bacteria: they exploit nanomolar levels of reduced sulfur in apparent preference to sulfate, which is present at 106- to 107-fold higher concentrations.  相似文献   

18.
1. Protein synthesis has been investigated in different regions of the rat epididymis by measuring incorporation of [35S]methionine in tissue minces incubated in vitro followed by analysis of labelled proteins on polyacrylamide gels containing sodium dodecyl sulphate. Rates of synthesis were highest in the proximal cauda > distal cauda > initial segment > ductuli efferentes > corpus > distal caput > proximal caput. One protein (mol.wt. 23 000) characterized the initial segment, three proteins (mol.wts. 18 500, 19 000 and 32 000) the caput and one protein (mol.wt. 47 000) the cauda. 2. After castration, [35S]methionine incorporation in all regions of the epididymis was reduced to < 10% of that in normal animals but could be restored to control levels within 5 days by testosterone treatment. Other steroids (corticosterone, oestrogen or progesterone) were ineffective. 3. The synthesis of the 18 500, 19 000, and 32 000 mol.wt. proteins in the caput and the 47 000 mol.wt. protein in the cauda were preferentially regulated by androgens, whilst the synthesis of 23 000 and approx. 80 000 mol.wt. proteins in the initial segment was dependent upon factors present in testicular fluid. 4. The androgen-dependent and testicular fluid-dependent proteins were major components of epididymal secretion. Purification and characterization of the 18 500, 19 000, 23 000 and 32 000 mol.wt. proteins showed them to be acidic glycoproteins with a carbohydrate content of 7.6-13.2%. The 47 000 mol.wt. protein, on the other hand, is highly basic. 5. A possible role for these proteins in the acquisition of motility, fertilizing capacity and storage of spermatozoa in the epididymis is discussed.  相似文献   

19.
The effects of glucose on insulin biosynthesis were studied by measuring the incorporation of radiolabelled amino acids into proinsulin/insulin in isolated rat islets. The islets were pulse labelled for 15 min with [3H]leucine (present in rat insulin I and II) or [35S]methionine (unique to rat insulin II) and then incubated for a 165 min post-label (chase) period during which the majority of labelled proinsulin was converted to insulin but under conditions whereby greater than 95% of radiolabelled proinsulin or insulin was retained in the islets. The newly synthesized, labelled, insulin was analyzed by high performance liquid chromatography. Rat I and II insulin biosynthesis was stimulated by 16.7 mM glucose to the same extent.  相似文献   

20.
Metabolic labeling with [35S]cysteine was used to characterize early events in CaSR biosynthesis. [35S]CaSR is relatively stable (half-life ∼8 h), but maturation to the final glycosylated form is slow and incomplete. Incorporation of [35S]cysteine is linear over 60 min, and the rate of [35S]CaSR biosynthesis is significantly increased by the membrane-permeant allosteric agonist NPS R-568, which acts as a cotranslational pharmacochaperone. The [35S]CaSR biosynthetic rate also varies as a function of conformational bias induced by loss- or gain-of-function mutations. In contrast, [35S]CaSR maturation to the plasma membrane was not significantly altered by exposure to the pharmacochaperone NPS R-568, the allosteric agonist neomycin, or the orthosteric agonist Ca2+ (0.5 or 5 mm), suggesting that CaSR does not control its own release from the endoplasmic reticulum. A CaSR chimera containing the mGluR1α carboxyl terminus matures completely (half-time of ∼8 h) and without a lag period, as does the truncation mutant CaSRΔ868 (half-time of ∼16 h). CaSRΔ898 exhibits maturation comparable with full-length CaSR, suggesting that the CaSR carboxyl terminus between residues Thr868 and Arg898 limits maturation. Overall, these results suggest that CaSR is subject to cotranslational quality control, which includes a pharmacochaperone-sensitive conformational checkpoint. The CaSR carboxyl terminus is the chief determinant of intracellular retention of a significant fraction of total CaSR. Intracellular CaSR may reflect a rapidly mobilizable “storage form” of CaSR and/or may subserve distinct intracellular signaling roles that are sensitive to signaling-dependent changes in endoplasmic reticulum Ca2+ and/or glutathione.  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