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.
In a previous study, the essential role of 3-sulfinopropionyl coenzyme A (3SP-CoA) desulfinase acyl-CoA dehydrogenase (Acd) in Advenella mimigardefordensis strain DPN7T (AcdDPN7) during degradation of 3,3′-dithiodipropionic acid (DTDP) was elucidated. DTDP is a sulfur-containing precursor substrate for biosynthesis of polythioesters (PTEs). AcdDPN7 showed high amino acid sequence similarity to acyl-CoA dehydrogenases but was unable to catalyze a dehydrogenation reaction. Hence, it was investigated in the present study whether 3SP-CoA desulfinase activity is an uncommon or a widespread property within the acyl-CoA dehydrogenase superfamily. Therefore, proteins of the acyl-CoA dehydrogenase superfamily from Advenella kashmirensis WT001, Bacillus cereus DSM31, Cupriavidus necator N-1, Escherichia coli BL21, Pseudomonas putida KT2440, Burkholderia xenovorans LB400, Ralstonia eutropha H16, Variovorax paradoxus B4, Variovorax paradoxus S110, and Variovorax paradoxus TBEA6 were expressed in E. coli strains. All purified acyl-CoA dehydrogenases appeared as homotetramers, as revealed by size exclusion chromatography. AcdS110, AcdB4, AcdH16, and AcdKT2440 were able to dehydrogenate isobutyryl-CoA. AcdKT2440 additionally dehydrogenated butyryl-CoA and valeryl-CoA, whereas AcdDSM31 dehydrogenated only butyryl-CoA and valeryl-CoA. No dehydrogenation reactions were observed with propionyl-CoA, isovaleryl-CoA, succinyl-CoA, and glutaryl-CoA for any of the investigated acyl-CoA dehydrogenases. Only AcdTBEA6, AcdN-1, and AcdLB400 desulfinated 3SP-CoA and were thus identified as 3SP-CoA desulfinases within the acyl-CoA dehydrogenase family, although none of these three Acds dehydrogenated any of the tested acyl-CoA thioesters. No appropriate substrates were identified for AcdBL21 and AcdWT001. Spectrophotometric assays provided apparent Km and Vmax values for active substrates and indicated the applicability of phylogenetic analyses to predict the substrate range of uncharacterized acyl-CoA dehydrogenases. Furthermore, C. necator N-1 was found to utilize 3SP as the sole source of carbon and energy.  相似文献   

2.
The act gene of Variovorax paradoxus TBEA6 encodes a succinyl-CoA:3-sulfinopropionate coenzyme A (CoA)-transferase, ActTBEA6 (2.8.3.x), which catalyzes the activation of 3-sulfinopropionate (3SP), an intermediate during 3,3′-thiodipropionate (TDP) degradation. In a previous study, accumulation of 3SP was observed in a Tn5::mob-induced mutant defective in growth on TDP. In contrast to the wild type and all other obtained mutants, this mutant showed no growth when 3SP was applied as the sole source of carbon and energy. The transposon Tn5::mob was inserted in a gene showing high homology to class III CoA-transferases. In the present study, analyses of the translation product clearly allocated ActTBEA6 to this protein family. The predicted secondary structure indicates the lack of a C-terminal α-helix. ActTBEA6 was heterologously expressed in Escherichia coli Lemo21(DE3) and was then purified by Ni-nitrilotriacetic acid (NTA) affinity chromatography. Analytical size exclusion chromatography revealed a homodimeric structure with a molecular mass of 96 ± 3 kDa. Enzyme assays identified succinyl-CoA, itaconyl-CoA, and glutaryl-CoA as potential CoA donors and unequivocally verified the conversion of 3SP to 3SP-CoA. Kinetic studies revealed an apparent Vmax of 44.6 μmol min−1 mg−1 for succinyl-CoA, which corresponds to a turnover number of 36.0 s−1 per subunit of ActTBEA6. For 3SP, the apparent Vmax was determined as 46.8 μmol min−1 mg−1, which corresponds to a turnover number of 37.7 s−1 per subunit of ActTBEA6. The apparent Km values were 0.08 mM for succinyl-CoA and 5.9 mM for 3SP. Nonetheless, the V. paradoxus Δact mutant did not reproduce the phenotype of the Tn5::mob-induced mutant. This defined deletion mutant was able to utilize TDP or 3SP as the sole carbon source, like the wild type. Complementation of the Tn5::mob-induced mutant with pBBR1MCS5::acdDPN7 partially restored growth on 3SP, which indicated a polar effect of the Tn5::mob transposon on acdTBEA6, located downstream of actTBEA6.  相似文献   

3.
The catabolism of the disulfide 3,3′-dithiodipropionic acid (DTDP) is initiated by the reduction of its disulfide bond. Three independent Tn5::mob-induced mutants of Advenella mimigardefordensis strain DPN7T were isolated that had lost the ability to utilize DTDP as the sole source of carbon and energy and that harbored the transposon insertions in three different sites of the same dihydrolipoamide dehydrogenase gene encoding the E3 subunit of the pyruvate dehydrogenase multi-enzyme complex of this bacterium (LpdAAm). LpdAAm was analyzed in silico and compared to homologous proteins, thereby revealing high similarities to the orthologue in Ralstonia eutropha H16 (PdhLRe). Both bacteria are able to cleave DTDP into two molecules of 3-mercaptopropionic acid (3MP). A. mimigardefordensis DPN7T converted 3MP to 3-sulfinopropionic acid, whereas R. eutropha H16 showed no growth with DTDP as the sole carbon source but was instead capable of synthesizing heteropolythioesters using the resulting cleavage product 3MP. Subsequently, the genes lpdAAm and pdhLRe were cloned, heterologously expressed in Escherichia coli applying the pET23a expression system, purified, and assayed by monitoring the oxidation of NADH. The physiological substrate lipoamide was reduced to dihydrolipoamide with specific activities of 1,833 mkat/kg of protein (LpdAAm) or 1,667 mkat/kg of protein (PdhLRe). Reduction of DTDP was also unequivocally detected with the purified enzymes, although the specific enzyme activities were much lower: 0.7 and 0.5 mkat/kg protein, respectively.In Advenella mimigardefordensis strain DPN7T (15, 42), three independent mutants with an insertion of Tn5::mob in the lpdA gene coding for the E3 component of the pyruvate dehydrogenase multi-enzyme complex revealed an interesting phenotype: these mutants were fully impaired in utilizing 3,3′-dithiodipropionic acid (DTDP) as the sole carbon and energy source, whereas the growth on no other tested carbon sources was affected (41). Our main interest in the catabolism of DTDP is to unravel the pathway and to identify the involved enzymes. Furthermore, the application of this disulfide as precursor substrate for biotechnological production of polythioesters (PTE) (22) is of interest. Since poly(3-mercaptopropionate) (PMP) biosynthesis depends hitherto on supplying the harmful thiol 3-mercaptopropionic acid (3MP) (35), an improvement of the recombinant Escherichia coli system by heterologous expression of enzymes capable of cleaving the less toxic DTDP symmetrically into two molecules of 3MP, which are then polymerized, could be an important achievement toward large-scale biotechnological production of PMP.Two different enzyme systems catalyzing the conversion of disulfides into the corresponding thiols are already known and have been described in detail. (i) Enzymes belonging to the well-characterized family of pyridine-nucleotide disulfide oxidoreductases (25) contain a redox center formed by a disulfide bridge coupled to a flavin ring. They catalyze a simultaneous two-electron transfer via the enzymatic active disulfides associated with the pyridine nucleotides and flavin, toward the substrate (39, 40). (ii) An alternative disulfide reduction is catalyzed by enzymes using iron-sulfur clusters to cleave of disulfide substrates in two one-electron reduction steps (37). The disrupted gene in A. mimigardefordensis was designated lpdAAm (EC 1.8.1.4), and it encodes a homodimeric flavoprotein, the dihydrolipoamide dehydrogenase LpdAAm (i.e., the E3 component of the pyruvate dehydrogenase multi-enzyme complex of A. mimigardefordensis strain DPN7T) belonging to the above-mentioned family of pyridine nucleotide-disulfide oxidoreductases. Enzymes of this class share high sequence and structural similarities and catalyze reduction of compounds which are linked by disulfide bonds (38). Alkylhydroperoxide reductases, coenzyme A disulfide reductases, glutathione reductases, mycothione reductases, thioredoxin reductases, and trypanothione reductases also, in addition to dihydrolipoamide dehydrogenases, belong to this family (3, 38). The physiological function of LpdAAm is most probably the conversion of lipoamide to dihydrolipoamide, but the reduction of DTDP into two molecules of 3MP (Fig. (Fig.1)1) is also predicted, enabling the first step in DTDP catabolism in A. mimigardefordensis strain DPN7T (41).Open in a separate windowFIG. 1.Reactions catalyzed by LpdAAm and PdhLRe. Presented are the enzymatic conversions of DTDP into two molecules of 3MP (A), lipoamide into dihydrolipoamide (B), and DTNB into two molecules of NTB (C). Abbreviations: DTDP, 3,3′-dithiodipropionic acid; 3MP, 3-mercaptopropionic acid; DTNB, 5,5′-dithiobis-(2-nitrobenzoic acid); NTB, 2-nitro-5-thiobenzoic acid.Ralstonia eutropha H16 synthesizes copolymers of 3-hydroxybutyrate and 3MP, if 3MP (23) or DTDP (22) is supplied as a precursor in addition to a second utilizable carbon source. Although R. eutropha is not able to grow with DTDP as the sole carbon source, it must be capable of cleaving this organic disulfide symmetrically, because it synthesizes from it heteropolymers containing the resulting 3MP. Thus, R. eutropha must possess at least one gene encoding a DTDP-cleaving enzyme. Five genes coding for homologues of a dihydrolipoamide dehydrogenase (DHLDH), which in A. mimigardefordensis DPN7T is obviously involved in DTDP degradation, are known to exist in the genome of R. eutropha H16 (27; M. Raberg, J. Bechmann, U. Brandt, J. Schlüter, B. Uischner, and A. Steinbüchel, unpublished data). Therefore, LpdAAm and the five DHLDH paralogues of R. eutropha H16 were aligned and compared (Fig. (Fig.2).2). Subsequently, lpdAAm and the gene encoding the DHLDH belonging to the pyruvate dehydrogenase complex of R. eutropha H16 (pdhLRe) were cloned, heterologously expressed in Escherichia coli, purified, and assayed.Open in a separate windowFIG. 2.Phylogenetic relationships of the A. mimigardefordensis strain DPN7T LpdA (boldface), R. eutropha H16 PdhL (boldface), and homologues. The neighbor-joining plot was derived from a CLUSTAL X alignment of amino acid sequences closely related to LpdAAm. The amino acid sequence of the outer membrane protein P64K from Neisseria meningitidis was used as the outgroup. GenBank accession numbers are given in parentheses. Scale bar, 10% sequence divergence.  相似文献   

4.
Steudle E 《Plant physiology》1980,66(6):1155-1163
Water-relation parameters of leaf mesophyll cells of the CAM plant Kalanchoë daigremontiana have been determined directly in cells of tissue slices using the pressure-probe technique. Turgor pressures measured in cells of the second to fourth layer from the cut surface showed an average of 1.82 ± 0.62 bar (mean ± sd; n = 157 cells). This was lower than expected from measurements of the osmotic pressure of the cell sap. The half-time (T1/2) for water-flux equilibration of individual cells was 2.5 to 8.8 seconds. This is the fastest T1/2 found so far for higher-plant cells. The calculated values of the hydraulic conductivity were in the range of 0.20 to 1.6 × 10−5 centimeters second−1 bar−1, with an average of (0.69 ± 0.46) × 10−5 centimeters second−1 bar−1 (mean ± sd; n = 8 cells). The T1/2 values of water exchange of individual cells are consistent with the overall rates of water-flux equilibration measured for tissue slices.The volumetric elastic moduli (∈) of individual cells were in the range 13 to 128 bar for turgor pressures between 0.0 and 3.4 bar; the average ∈ value was 42.4 ± 27.7 bar (mean ± sd; n = 21 cells). This ∈ value is similar to that observed for other higher-plant cells.The water-storage capacity of individual cells, calculated as Cc = V/(∈ + πi) (where V = cell volume and πi = internal osmotic pressure) was 9.1 × 10−9 cubic centimeters bar−1 per cell, and the capacity for the tissue was 2.2 × 10−2 cubic centimeters bar−1 gram−1 fresh weight. The significance of the water-relation parameters determined at the cellular level is discussed in terms of the water relations of whole leaves and the high water-use efficiency characteristic of CAM plants.  相似文献   

5.
The hitherto unstudied microbial degradation of the organic disulfide 3,3'-dithiodipropionic acid (DTDP) was investigated with the recently described bacterium Tetrathiobacter mimigardefordensis strain DPN7(T) (DSM 17166(T); LMG 22922(T)), which is able to use DTDP as the sole carbon source for growth. 3-Mercaptopropionic acid (3MP) and 3-sulfinopropionic acid (3SP) were detected in the growth medium and occurred as intermediates during DTDP degradation. To identify genes coding for enzymes of DTDP catabolism, Tn5::mob-induced mutants of T. mimigardefordensis were generated. Screening of transposon mutant libraries yielded many mutants fully or partially impaired in utilizing DTDP as a carbon source. Mapping of the insertion loci in some mutants identified four disrupted open reading frames (ORFs) with putative metabolic functions. The ORFs were assigned function on the basis of homologies with lpdA (EC 1.8.1.4), cdo (EC 1.13.11.20), sucCD (EC 6.2.1.5), and acnB (EC 4.2.1.3). Tn5::mob insertions occurred additionally in the vicinity of heat shock protein-encoding genes. The predicted function of the LpdA homologue in T. mimigardefordensis is cleavage of the disulfide bond of DTDP to form two molecules of 3MP. Cdo catalyzes the conversion of the sulfhydryl group of 3MP, yielding the corresponding sulfinic acid, 3SP. SucCD exhibits thiokinase activity, ligating coenzyme A (CoA) with 3SP to form 3SP-CoA. Afterwards, an elimination of sulfite via a putative desulfinase is expected. acnB encodes a putative 2-methylisocitrate dehydratase. Therefore, a new pathway is proposed for the catabolism of DTDP via 3MP, 3SP, and 3SP-CoA toward propionyl-CoA, which is then further catabolized via the 2-methylcitric acid cycle in T. mimigardefordensis.  相似文献   

6.
Lü Z  Lu Y 《PloS one》2012,7(4):e35279

Background

Methanocellales contributes significantly to anthropogenic methane emissions that cause global warming, but few pure cultures for Methanocellales are available to permit subsequent laboratory studies (physiology, biochemistry, etc.).

Methodology/Principal Findings

By combining anaerobic culture and molecular techniques, a novel thermophilic methanogen, strain HZ254T was isolated from a Chinese rice field soil located in Hangzhou, China. The phylogenetic analyses of both the 16S rRNA gene and mcrA gene (encoding the α subunit of methyl-coenzyme M reductase) confirmed its affiliation with Methanocellales, and Methanocella paludicola SANAET was the most closely related species. Cells were non-motile rods, albeit with a flagellum, 1.4–2.8 µm long and by 0.2–0.3 µm in width. They grew at 37–60°C (optimally at 55°C) and salinity of 0–5 g NaCl l−1 (optimally at 0–1 g NaCl l−1). The pH range for growth was 6.4–7.2 (optimum 6.8). Under the optimum growth condition, the doubling time was 6.5–7.8 h, which is the shortest ever observed in Methanocellales. Strain HZ254T utilized H2/CO2 but not formate for growth and methane production. The DNA G+C content of this organism was 52.7 mol%. The sequence identities of 16S rRNA gene and mcrA gene between strain HZ254T and SANAET were 95.0 and 87.5% respectively, and the genome based Average Nucleotide Identity value between them was 74.8%. These two strains differed in phenotypic features with regard to substrate utilization, possession of a flagellum, doubling time (under optimal conditions), NaCl and temperature ranges. Taking account of the phenotypic and phylogenetic characteristics, we propose strain HZ254T as a representative of a novel species, Methanocella conradii sp. nov. The type strain is HZ254T ( = CGMCC 1.5162T = JCM 17849T = DSM 24694T).

Conclusions/Significance

Strain HZ254T could potentially serve as an excellent laboratory model for studying Methanocellales due to its fast growth and consistent cultivability.  相似文献   

7.
Heavy-metal-tolerant bacteria, GIMN1.004T, was isolated from mine soils of Dabaoshan in South China, which were acidic (pH 2–4) and polluted with heavy metals. The isolation was Gram-negative, aerobic, non-spore-forming, and rod-shaped bacteria having a cellular width of 0.5−0.6 µm and a length of 1.3−1.8 µm. They showed a normal growth pattern at pH 4.0–9.0 in a temperature ranging from 5°C to 40°C.The organism contained ubiquinone Q-8 as the predominant isoprenoid quinine, and C16∶0, summed feature 8 (C18∶1 ω7c and C18∶1 ω6c), C18∶0, summed feature 3 (C16∶1 ω7c or iso-C15∶0 2-OH), C17∶0 cyclo, C18∶1 ω9c, C19∶0 cyclo ω8c, C14∶0 as major fatty acid. These profiles were similar to those reported for Burkholderia species. The DNA G+C % of this strain was 61.6%. Based on the similarity to 16S rRNA gene sequence, GIMN1.004T was considered to be in the genus Burkholderia. The similarities of 16S rRNA gene sequence between strain GIMN1.004T and members of the genus Burkholderia were 96−99.4%, indicating that this novel strain was phylogenetically related to members of that genus. The novel strain showed the highest sequence similarities to Burkholderia soli DSM 18235T (99.4%); Levels of DNA-DNA hybridization with DSM 18235T was 25%. Physiological and biochemical tests including cell wall composition analysis, differentiated phenotype of this strain from that closely related Burkholderia species. The isolation had great tolerance to cadmium with MIC of 22 mmol/L, and adsorbability of 144.94 mg/g cadmium,and it was found to exhibit antibiotic resistance characteristics. The adsorptive mechanism of GIMN1.004T for cadmium depended on the action of the amide,carboxy and phosphate of cell surface and producing low-molecular-weight (LMW ) organic acids to complex or chelated Cd2+.Therefore, the strain GIMN1.004T represented a new cadmium resistance species, which was tentatively named as Burkholderia dabaoshanensis sp. nov. The strain type is GIMN1.004T ( = CCTCC M 209109T =  NRRL B-59553T ).  相似文献   

8.
The relative bioavailability of chlorothiazide from mucoadhesive polymeric compacts is compared to commercial oral suspension in pigs. A single-dose randomized study was conducted in 12 healthy pigs that are 9–10 weeks old. After overnight fasting, pigs were divided into two groups of six animals. To the first group, a reference product containing 50 mg of chlorothiazide suspension, and in the second group, test product (mucoadhesive compacts) chlorothiazide (50 mg) was administered with 75 mL of water via gastric tubes. Blood samples were collected between 0 to 24 h using catheters inserted into the jugular vein. Plasma was separated by protein precipitation, and chlorothiazide concentrations were determined using a high-performance liquid chromatography method. The mean Tmax and the Cmax of chlorothiazide following the administration of oral suspension and mucoadhesive compacts were 0.58 ± 0.20 h and 682.97 ± 415.69 ng/mL and 2.17 ± 0.98 h and 99.42 ± 124.08 ng/mL, respectively. The Kel and T1/2 of chlorothiazide were found to be 1.06 ± 0.28 h−1 and 0.70 ± 0.21 h from suspension and 0.95 ± 1.11 h−1 and 2.05 ± 1.90 h from the compacts, respectively. The Tmax of mucoadhesive compacts were significantly longer (p < 0.05; 2.17 h) than the reference products (0.58 h), whereas the Cmax of compacts were significantly lower (99 ng/mL) than the reference product (683 ng/mL; p < 0.05). The area under the curve (AUC) of compacts accounts only 50.15% (404.32 ± 449.93 ng h/mL) of the reference product’s AUC (806.27 ± 395.97 ng h/mL). The relative bioavailability of the compacts was lower than that of the suspension, and this may be due to the narrow window of absorption for chlorothiazide.Key words: bioavailability, chlorothiazide, mucoadhesive compacts, pigs  相似文献   

9.

Background

South Asians are more insulin resistant than Europeans, which cannot be fully explained by differences in adiposity. We investigated whether differences in oxidative capacity and capacity for fatty acid utilisation in South Asians might contribute, using a range of whole-body and skeletal muscle measures.

Methodology/Principal Findings

Twenty men of South Asian ethnic origin and 20 age and BMI-matched men of white European descent underwent exercise and metabolic testing and provided a muscle biopsy to determine expression of oxidative and lipid metabolism genes and of insulin signalling proteins. In analyses adjusted for age, BMI, fat mass and physical activity, South Asians, compared to Europeans, exhibited; reduced insulin sensitivity by 26% (p = 0.010); lower VO2max (40.6±6.6 vs 52.4±5.7 ml.kg−1.min−1, p = 0.001); and reduced fat oxidation during submaximal exercise at the same relative (3.77±2.02 vs 6.55±2.60 mg.kg−1.min−1 at 55% VO2max, p = 0.013), and absolute (3.46±2.20 vs 6.00±1.93 mg.kg−1.min−1 at 25 ml O2.kg−1.min−1, p = 0.021), exercise intensities. South Asians exhibited significantly higher skeletal muscle gene expression of CPT1A and FASN and significantly lower skeletal muscle protein expression of PI3K and PKB Ser473 phosphorylation. Fat oxidation during submaximal exercise and VO2max both correlated significantly with insulin sensitivity index and PKB Ser473 phosphorylation, with VO2max or fat oxidation during exercise explaining 10–13% of the variance in insulin sensitivity index, independent of age, body composition and physical activity.

Conclusions/Significance

These data indicate that reduced oxidative capacity and capacity for fatty acid utilisation at the whole body level are key features of the insulin resistant phenotype observed in South Asians, but that this is not the consequence of reduced skeletal muscle expression of oxidative and lipid metabolism genes.  相似文献   

10.
Rev is an essential HIV-1 regulatory protein which binds to the Rev responsive element (RRE) present within the env gene of HIV-1 RNA genome. This binding facilitates the transport of the RNA to the cytoplasm, which in turn triggers the switch between viral latency and active viral replication. Essential components of this complex have been localized to a minimal arginine rich Rev peptide and stem IIB region of RRE. A synthetic peptide known as RSG-1.2 binds with high binding affinity and specificity to the RRE-IIB than the Rev peptide, however the thermodynamic basis of this specificity has not yet been addressed. The present study aims to probe the thermodynamic origin of this specificity of RSG-1.2 over Rev Peptide for RRE-IIB. The temperature dependent melting studies show that RSG-1.2 binding stabilizes the RRE structure significantly (ΔT m = 4.3°C), in contrast to Rev binding. Interestingly the thermodynamic signatures of the binding have also been found to be different for both the peptides. At pH 7.5, RSG-1.2 binds RRE-IIB with a Ka = 16.2±0.6×107 M−1 where enthalpic change ΔH = −13.9±0.1 kcal/mol is the main driving force with limited unfavorable contribution from entropic change TΔS = −2.8±0.1 kcal/mol. A large part of ΔH may be due to specific stacking between U72 and Arg15. In contrast binding of Rev (Ka = 3.1±0.4×107 M−1) is driven mainly by entropy (ΔH = 0 kcal/mol and TΔS = 10.2±0.2 kcal/mol) which arises from major conformational changes in the RNA upon binding.  相似文献   

11.
In this study, we investigated an SBP (DctPAm) of a tripartite ATP‐independent periplasmic transport system (TRAP) in Advenella mimigardefordensis strain DPN7T. Deletion of dctPAm as well as of the two transmembrane compounds of the tripartite transporter, dctQ and dctM, impaired growth of A. mimigardefordensis strain DPN7T, if cultivated on mineral salt medium supplemented with d ‐glucose, d ‐galactose, l ‐arabinose, d ‐fucose, d ‐xylose or d ‐gluconic acid, respectively. The wild type phenotype was restored during complementation studies of A. mimigardefordensis ΔdctPAm using the broad host vector pBBR1MCS‐5::dctPAm. Furthermore, an uptake assay with radiolabeled [14C(U)]‐d ‐glucose clearly showed that the deletion of dctPAm, dctQ and dctM, respectively, disabled the uptake of this aldoses in cells of either mutant strain. Determination of KD performing thermal shift assays showed a shift in the melting temperature of DctPAm in the presence of d ‐gluconic acid (KD 11.76 ± 1.3 µM) and the corresponding aldonic acids to the above‐mentioned carbohydrates d ‐galactonate (KD 10.72 ± 1.4 µM), d ‐fuconic acid (KD 13.50 ± 1.6 µM) and d ‐xylonic acid (KD 8.44 ± 1.0 µM). The sugar (glucose) dehydrogenase activity (E.C.1.1.5.2) in the membrane fraction was shown for all relevant sugars, proving oxidation of the molecules in the periplasm, prior to transport.  相似文献   

12.

Background and Aims

Floral thermogenesis occurs in at least 12 families of ancient seed plants. Some species show very high rates of respiration through the alternative pathway, and some are thermoregulatory, with increasing respiration at decreasing ambient temperature. This study assesses the intensity and regulation of respiration in three species of African Hydnora that represent the Hydnoraceae, an unusual family of holoparasitic plants from arid environments.

Methods

Long-term respirometry (CO2 production) and thermometry were carried out on intact flowers of H. africana, H. abyssinica and H. esculenta in the field, and short-term measurements were made on floral parts during the protogynous flowering sequence.

Key Results

For H. africana, there was no temperature elevation in either the osmophores or the gynoecial chamber in any phase, and mass-specific respiration rates of the flower parts were low (maximum 8·3 nmol CO2 g−1 s−1 in osmophore tissue). Respiration tracked ambient and floral temperatures, eliminating the possibility of the inverse relationship expected in thermoregulatory flowers. Hydnora abyssinica flowers had higher respiration (maximum 27·5 nmol g−1 s−1 in the osmophores) and a slight elevation of osmophore temperature (maximum 2·8 °C) in the female stage. Respiration by gynoecial tissue was similar to that of osmophores in both species, but there was no measurable elevation of gynoecial chamber temperature. Gynoecial chamber temperature of H. esculenta could reach 3·8 °C above ambient, but there are no respiration data available. Antheral tissue respiration was maximal in the male phase (4·8 nmol g−1 s−1 in H. africana and 10·3 nmol g−1 s−1 in H. abyssinica), but it did not raise the antheral ring temperature, which showed that thermogenesis is not a by-product of pollen maturation or release.

Conclusions

The exceptionally low thermogenesis in Hydnora appears to be associated with scent production and possibly gynoecial development, but has little direct benefit to beetle pollinators.Key words: Pollination biology, Hydnora, thermogenesis, respiration rate, temperature, flowers, insects  相似文献   

13.
14.
Soil physic-chemical properties differ at different depths; however, differences in afforestation-induced temporal changes at different soil depths are seldom reported. By examining 19 parameters, the temporal changes and their interactions with soil depth in a large chronosequence dataset (159 plots; 636 profiles; 2544 samples) of larch plantations were checked by multivariate analysis of covariance (MANCOVA). No linear temporal changes were found in 9 parameters (N, K, N:P, available forms of N, P, K and ratios of N: available N, P: available P and K: available K), while marked linear changes were found in the rest 10 parameters. Four of them showed divergent temporal changes between surface and deep soils. At surface soils, changing rates were 262.1 g·kg−1·year−1 for SOM, 438.9 mg·g−1·year−1 for C:P, 5.3 mg·g−1·year−1 for C:K, and −3.23 mg·cm−3·year−1 for bulk density, while contrary tendencies were found in deeper soils. These divergences resulted in much moderated or no changes in the overall 80-cm soil profile. The other six parameters showed significant temporal changes for overall 0–80-cm soil profile (P: −4.10 mg·kg−1·year−1; pH: −0.0061 unit·year−1; C:N: 167.1 mg·g−1·year−1; K:P: 371.5 mg·g−1 year−1; N:K: −0.242 mg·g−1·year−1; EC: 0.169 μS·cm−1·year−1), but without significant differences at different soil depths (> 0.05). Our findings highlight the importance of deep soils in studying physic-chemical changes of soil properties, and the temporal changes occurred in both surface and deep soils should be fully considered for forest management and soil nutrient balance.  相似文献   

15.

Background

Zebrafish has been largely accepted as a vertebrate multidisciplinary model but its usefulness as a model for exercise physiology has been hampered by the scarce knowledge on its swimming economy, optimal swimming speeds and cost of transport. Therefore, we have performed individual and group-wise swimming experiments to quantify swimming economy and to demonstrate the exercise effects on growth in adult zebrafish.

Methodology/Principal Findings

Individual zebrafish (n = 10) were able to swim at a critical swimming speed (Ucrit) of 0.548±0.007 m s−1 or 18.0 standard body lengths (BL) s−1. The optimal swimming speed (Uopt) at which energetic efficiency is highest was 0.396±0.019 m s−1 (13.0 BL s−1) corresponding to 72.26±0.29% of Ucrit. The cost of transport at optimal swimming speed (COTopt) was 25.23±4.03 µmol g−1 m−1. A group-wise experiment was conducted with zebrafish (n = 83) swimming at Uopt for 6 h day−1 for 5 days week−1 for 4 weeks vs. zebrafish (n = 84) that rested during this period. Swimming zebrafish increased their total body length by 5.6% and body weight by 41.1% as compared to resting fish. For the first time, a highly significant exercise-induced growth is demonstrated in adult zebrafish. Expression analysis of a set of muscle growth marker genes revealed clear regulatory roles in relation to swimming-enhanced growth for genes such as growth hormone receptor b (ghrb), insulin-like growth factor 1 receptor a (igf1ra), troponin C (stnnc), slow myosin heavy chain 1 (smyhc1), troponin I2 (tnni2), myosin heavy polypeptide 2 (myhz2) and myostatin (mstnb).

Conclusions/Significance

From the results of our study we can conclude that zebrafish can be used as an exercise model for enhanced growth, with implications in basic, biomedical and applied sciences, such as aquaculture.  相似文献   

16.
Ephedra foliata Boiss. & Kotschy ex Boiss., (family – Ephedraceae), is an ecologically and economically important threatened Gymnosperm of the Indian Thar Desert. A method for micropropagation of E. foliata using nodal explant of mature female plant has been developed. Maximum bud-break (90 %) of the explant was obtained on MS medium supplemented with 1.5 mg l−1 of benzyl adenine (BA) + additives. Explant produces 5.3 ± 0.40 shoots from single node with 3.25 ± 0.29 cm length. The multiplication of shoots in culture was affected by salt composition of media, types and concentrations of plant growth regulators (PGR’s) and their interactions, time of transfer of the cultures. Maximum number of shoots (26.3 ± 0.82 per culture vessel) were regenerated on MS medium modified by reducing the concentration of nitrates to half supplemented with 200 mg l−1 ammonium sulphate {(NH4) 2SO4} (MMS3) + BA (0.25 mg l−1), Kinetin (Kin; 0.25 mg l−1), Indole-3-acetic acid (IAA; 0.1 mg l−1) and additives. The in vitro produced shoots rooted under ex vitro on soilrite moistened with one-fourth strength of MS macro salts in screw cap bottles by treating the shoot base (s) with 500 mg l−1 of Indole-3-butyric acid (IBA) for 5 min. The micropropagated plants were hardened in the green house. The described protocol can be applicable for (i) large scale plant production (ii) establishment of plants in natural habitat and (iii) germplasm conservation of this endemic Gymnosperm of arid regions.  相似文献   

17.
As a stable analog for ADP-sensitive phosphorylated intermediate of sarcoplasmic reticulum Ca2+-ATPase E1PCa2·Mg, a complex of E1Ca2·BeFx, was successfully developed by addition of beryllium fluoride and Mg2+ to the Ca2+-bound state, E1Ca2. In E1Ca2·BeFx, most probably E1Ca2·BeF3, two Ca2+ are occluded at high affinity transport sites, its formation required Mg2+ binding at the catalytic site, and ADP decomposed it to E1Ca2, as in E1PCa2·Mg. Organization of cytoplasmic domains in E1Ca2·BeFx was revealed to be intermediate between those in E1Ca2·AlF4 ADP (transition state of E1PCa2 formation) and E2·BeF3·(ADP-insensitive phosphorylated intermediate E2P·Mg). Trinitrophenyl-AMP (TNP-AMP) formed a very fluorescent (superfluorescent) complex with E1Ca2·BeFx in contrast to no superfluorescence of TNP-AMP bound to E1Ca2·AlFx. E1Ca2·BeFx with bound TNP-AMP slowly decayed to E1Ca2, being distinct from the superfluorescent complex of TNP-AMP with E2·BeF3, which was stable. Tryptophan fluorescence revealed that the transmembrane structure of E1Ca2·BeFx mimics E1PCa2·Mg, and between those of E1Ca2·AlF4·ADP and E2·BeF3. E1Ca2·BeFx at low 50–100 μm Ca2+ was converted slowly to E2·BeF3 releasing Ca2+, mimicking E1PCa2·Mg → E2P·Mg + 2Ca2+. Ca2+ replacement of Mg2+ at the catalytic site at approximately millimolar high Ca2+ decomposed E1Ca2·BeFx to E1Ca2. Notably, E1Ca2·BeFx was perfectly stabilized for at least 12 days by 0.7 mm lumenal Ca2+ with 15 mm Mg2+. Also, stable E1Ca2·BeFx was produced from E2·BeF3 at 0.7 mm lumenal Ca2+ by binding two Ca2+ to lumenally oriented low affinity transport sites, as mimicking the reverse conversion E2P· Mg + 2Ca2+E1PCa2·Mg.Sarcoplasmic reticulum Ca2+-ATPase (SERCA1a),2 a representative member of the P-type ion transporting ATPases, catalyze Ca2+ transport coupled with ATP hydrolysis (Fig. 1) (19). The enzyme forms phosphorylated intermediates from ATP or Pi in the presence of Mg2+ (1013). In the transport cycle, the enzyme is first activated by cooperative binding of two Ca2+ ions at high affinity transport sites (E2 to E1Ca2, steps 1–2) (14) and autophosphorylated at Asp351 with MgATP to form the ADP-sensitive phosphoenzyme (E1P, step 3), which reacts with ADP to regenerate ATP in the reverse reaction. Upon this E1P formation, the two bound Ca2+ are occluded in the transport sites (E1PCa2). Subsequent isomeric transition to the ADP-insensitive form (E2PCa2), i.e. loss of ADP sensitivity at the catalytic site, results in rearrangement of the Ca2+ binding sites to deocclude Ca2+, reduce the affinity, and open the lumenal gate, thus releasing Ca2+ into the lumen (E2P, steps 4–5). Finally Asp351-acylphosphate in E2P is hydrolyzed to form the Ca2+-unbound inactive E2 state (steps 6 and 7). Mg2+ bound at the catalytic site is required as a physiological catalytic cofactor in phosphorylation and dephosphorylation and thus for the transport cycle. The cycle is totally reversible, e.g. E2P can be formed from Pi in the presence of Mg2+ and absence of Ca2+, and subsequent Ca2+ binding at lumenally oriented low affinity transport sites of E2P reverses the Ca2+-releasing step and produces E1PCa2, which is then decomposed to E1Ca2 by ADP.Open in a separate windowFIGURE 1.Ca2+ transport cycle of Ca2+-ATPase.Various intermediate structural states in the transport cycle were fixed as their structural analogs produced by appropriate ligands such as AMP-PCP (non-hydrolyzable ATP analog) or metal fluoride compounds (phosphate analogs), and their crystal structures were solved at the atomic level (1522). The three cytoplasmic domains, N, P, and A, largely move and change their organization state during the transport cycle, and the changes are coupled with changes in the transport sites. Most remarkably, in the change from E1Ca2·AlF4·ADP (the transition state for E1PCa2 formation, E1PCa2·ADP·Mg) to E2·BeF3 (the ground state E2P·Mg) (2325), the A domain largely rotates by more than 90° approximately parallel to the membrane plane and associates with the P domain, thereby destroying the Ca2+ binding sites, and opening the lumenal gate, thus releasing Ca2+ into the lumen (see Fig. 2). E1PCa2·Ca·AMP-PN formed by CaAMP-PNP without Mg2+ is nearly the same as E1Ca2·AlF4·ADP and E1Ca2·CaAMP-PCP in their crystal structures (17, 18, 22).Open in a separate windowFIGURE 2.Structure of SERCA1a and its change during processing of phosphorylated intermediate. E1Ca2·AlF4·ADP (the transition state analog for phosphorylation E1PCa2·ADP·Mg) and E2·BeF3 (the ground state E2P analog (25)) were obtained from the Protein Data Bank (PDB accession code 1T5T (17) and 2ZBE (21), respectively). Cytoplasmic domains N (nucleotide binding), P (phosphorylation), and A (actuator), and 10 transmembrane helices (M1–M10) are indicated. The arrows on the domains, M1′ and M2 (Tyr122) in E1Ca2·AlF4·ADP, indicate their approximate motions predicted for E1PCa2·ADP·MgE2P·Mg. The phosphorylation site Asp351, TGES184 of the A domain, Arg198 (tryptic T2 site) on the Val200 loop (DPR198AV200NQD) of the A domain, and Thr242 (proteinase K site) on the A/M3-linker are shown. Seven hydrophobic residues gather in the E2P state to form the Tyr122-hydrophobic cluster (Y122-HC); Tyr122/Leu119 on the top part of M2, Ile179/Leu180/Ile232 of the A domain, and Val705/Val726 of the P domain. The overall structure of E1Ca2·AlF4·ADP is virtually the same as those of E1Ca2·CaAMP-PCP and E1PCa2·Ca·AMP-PN (17, 18, 22).Despite these atomic structures, yet unsolved is the structure of E1PCa2·Mg, the genuine physiological intermediate E1PCa2 with bound Mg2+ at the catalytic site without the nucleotide. Its stable structural analog has yet to be developed. E1PCa2·Mg is the major intermediate accumulating almost exclusively at steady state under physiological conditions. Its rate-limiting isomerization results in Ca2+ deocclusion/release producing E2P·Mg as a key event for Ca2+ transport. In E1Ca2·CaAMP-PCP, E1Ca2·AlF4·ADP, and E1PCa2·Ca·AMP-PN, the N and P domains are cross-linked and strongly stabilized by the bound nucleotide and/or Ca2+ at the catalytic site, thus they are crystallized (17, 18, 22). Kinetically, E1PCa2·Ca formed with CaATP is markedly stabilized due to Ca2+ binding at the catalytic Mg2+ site, and its isomerization to E2P is strongly retarded in contrast to E1PCa2·Mg (26, 27). Thus, the bound Ca2+ at the catalytic Mg2+ site likely produces a significantly different structural state from that with bound Mg2+.Therefore, it is now essential to develop a genuine E1PCa2·Mg analog without bound nucleotide and thereby gain further insight into the structural mechanism in the Ca2+ transport process. It is also crucial to further clarify the structural importance of Mg2+ as the physiological catalytic cation. In this study, we successfully developed the complex E1Ca2·BeFx, most probably E1Ca2·BeF3, as the E1PCa2·Mg analog by adding beryllium fluoride (BeFx) to the E1Ca2 state without any nucleotides. For its formation, Mg2+ binding at the catalytic site was required and Ca2+ substitution for Mg2+ was absolutely unfavorable, revealing a likely structural reason for its preference as the physiological cofactor. In E1Ca2·BeF3, two Ca2+ ions bound at the high affinity transport sites are occluded. It was also produced from E2·BeF3 by lumenal Ca2+ binding at the lumenally oriented low affinity transport sites, mimicking E2P·Mg + 2Ca2+E1PCa2·Mg. All properties of the newly developed E1Ca2·BeF3 fulfilled the requirements as the E1PCa2·Mg analog, and hence we were able to uncover the hitherto unknown nature of E1PCa2·Mg as well as structural events occurring in the phosphorylation and isomerization processes. Also, we successfully found the conditions that perfectly stabilize the E1Ca2·BeF3 complex.  相似文献   

18.
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   

19.
Naturally occurring tufts of the mixotroph Thiothrix nivea blanketed the East Everglades (Dade County, Fla.) Chekika artesian well and runoff areas. The rate of HCO3 fixation by these Thiothrix tufts was determined to be 14.0 ± 5.4 nmol of HCO3 per min per mg of dry weight, which reflected a growth rate of 5.0%/h. The addition of 10 mM glucose, ribose, acetate, or pyruvate or 0.05% Casamino Acids (Difco Laboratories, Detroit, Mich.) did not appear to alter the HCO3 fixation rate. Whereas 1 mM acetate or 10 mM lactate, ethanol, glycerol, α-ketoglutarate, succinate, fumarate, or citrate slightly stimulated HCO3 fixation, 5 to 10 mM malate inhibited HCO3 fixation by 90%. Pure Thiothrix cultures isolated from Chekika fixed HCO3 at rates as high as 29.9 ± 2.8 nmol of HCO3 per min per mg of dry weight in the presence of growth medium. Malate did not have a suppressive effect but rather slightly stimulated in vivo HCO3 fixation.  相似文献   

20.
Extracellular glucoamylase of Colletotrichum sp. KCP1 produced through solid state fermentation was purified by two steps purification process comprising ammonium sulphate precipitation followed by gel permeation chromatography (GPC). The Recovery of glucoamylase after GPC was 50.40 % with 19.3-fold increase in specific activity. The molecular weight of enzyme was found to be 162.18 kDa by native-PAGE and was dimeric protein of two sub-units with molecular weight of 94.62 and 67.60 kDa as determined by SDS-PAGE. Activation energy for starch hydrolysis was 26.45 kJ mol−1 while temperature quotient (Q10) was found to be 1.9. The enzyme was found to be stable over wide pH range and thermally stable at 40–50 °C up to 120 min while exhibited maximum activity at 50 °C with pH 5.0. The pKa1 and pKa2 of ionisable groups of active site controlling Vmax were 3.5 and 6.8, respectively. Vmax, Km and Kcat for starch hydrolysis were found to be 58.82 U ml−1, 1.17 mg (starch) ml−1 and 449 s−1, respectively. Activation energy for irreversible inactivation (Ea(d)) of glucoamylase was 74.85 kJ mol−1. Thermodynamic parameters of irreversible inactivation of glucoamylase and starch hydrolysis were also determined.  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