首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Aerobic denitrification in various heterotrophic nitrifiers   总被引:17,自引:0,他引:17  
Various heterotrophic nitrifiers have been tested and found to also be aerobic denitrifiers. The simultaneous use of two electron acceptors (oxygen and nitrate) permits these organisms to grow more rapidly than on either single electron acceptor, but generally results in a lower yield than is obtained on oxygen, alone. One strain, formerly known as Pseudomonas denitrificans, was grown in the chemostat and shown to achieve nitrification rates of up to 44 nmol NH3 min–1 mg protein–1 and denitrification rates up to 69 nmol NO inf3 sup–1 min–1 mg protein–1.Unlike Thiosphaera pantotropha, this strain needed to induce its nitrate reductase. However, the remainder of the denitrifying pathway was constitutive and, like T. pantotropha, Ps. denitrificans probably possesses the copper nitrite reductase.  相似文献   

2.
The rates of phosphodiesterase-promoted hydrolysis of cGMP and cAMP have been measuted in intact neuroblastoma N1E-115 cells by determining rates of18O incorporation from18O-water into the -phosphoryls of guanine and adenine nucleotides. The basal rate of guanine nucleotide -phosphoryl labeling ranged from 180 to 244 pmol·mg protein–1·min–1. Sodium nitroprusside (SNP) caused a sustained 3,4-fold increase in this18O-labeling rate in conjunction with 28- and 50-fold increases in cellular cGMP concentration at 3 and 6 min, respectively. This18O-labeling rate (795 pmol·mg protein–1·min–1) corresponded with the sum of the low (1.7 M) and high (34 M) Km phosphodiesterase activities assayable in cell lysates which exhibited a combined maximum velocity of 808 pmol·mg protein–1·min–1 to which the highK m species contributed 84%. This information and the characteristics of the profile of18O-labeled molecular species indicate that cGMP metabolism was restricted to a very discrete cellular compartment(s) of approximately 12% of the cell volume. Carbachol (1 mM) produced a transient increase (6-fold) in cellular cGMP concentration and a transient increase (90%) in the rate of18O labeling of -GTP during the first minute of treatment which translates into 30 additional cellular pools of cGMP hydrolyzed in this period. IBMX (1 mM) produced a relatively rapid increase in cellular cGMP (3- to 5-fold) and cAMP (2-fold) concentrations and a delayed inhibition of18O labeling of guanine and adenine nucleotide -phosphoryls without further elevation of cyclic nucleotide levels. These results indicate that besides inhibiting cyclic nucleotide hydrolysis, IBMX also imparts a time-dependent inhibitory influence on the generation of cyclic nucleotides. The data obtained show that measurement of18O labeling of guanine and adenine nucleotide -phosphoryls combined with measurements of cyclic nucleotide steady state levels provides a means to assess the rates of cyclic nucleotide synthesis and hydrolysis within intact cells and to identify the site(s) of action of agents that alter cellular cyclic nucleotide metabolism.Special Issue dedicated to Dr. O. H. Lowry.  相似文献   

3.
Directed evolution of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 previously created the hydroxylase α-subunit (TomA3) V106A variant (TOM-Green) with increased activity for both trichloroethylene degradation (twofold enhancement) and naphthalene oxidation (six-times-higher activity). In the present study, saturation mutagenesis was performed at position A106 with Escherichia coli TG1/pBS(Kan)TOMV106A to improve TOM activity for both chloroform degradation and naphthalene oxidation. Whole cells expressing the A106E variant had two times better naphthalene-to-1-naphthol activity than the wild-type cells (Vmax of 9.3 versus 4.5 nmol·min−1·mg of protein−1 and unchanged Km), and the regiospecificity of the A106E variant was unchanged, with 98% 1-naphthol formed, as was confirmed with high-pressure liquid chromatography. The A106E variant degrades its natural substrate toluene 63% faster than wild-type TOM does (2.12 ± 0.07 versus 1.30 ± 0.06 nmol·min−1·mg of protein−1 [mean ± standard deviation]) at 91 μM and has a substantial decrease in regiospecificity, since o-cresol (50%), m-cresol (25%), and p-cresol (25%) are formed, in contrast to the 98% o-cresol formed by wild-type TOM. The A106E variant also has an elevated expression level compared to that of wild-type TOM, as evidenced by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Another variant, the A106F variant, has 2.8-times-better chloroform degradation activity based on gas chromatography (Vmax of 2.61 versus 0.95 nmol·min−1·mg of protein−1 and unchanged Km) and chloride release (0.034 ± 0.002 versus 0.012 ± 0.001 nmol·min−1·mg of protein−1). The A106F variant also was expressed at levels similar to those of wild-type TOM and 62%-better toluene oxidation activity than wild-type TOM (2.11 ± 0.3 versus 1.30 ± 0.06 nmol·min−1·mg of protein−1). A shift in regiospecificity of toluene hydroxylation was also observed for the A106F variant, with o-cresol (28%), m-cresol (18%), and p-cresol (54%) being formed. Statistical analysis was used to estimate that 292 colonies must be screened for a 99% probability that all 64 codons were sampled during saturation mutagenesis.  相似文献   

4.
Karni  Leah  Moss  Stephen J.  Tel-Or  Elisha 《Archives of microbiology》1984,140(2-3):215-217
Glutathione reductase activity was detected and characterized in heterocysts and vegetative cells of the cyanobacterium Nostoc muscorum. The activity of the enzyme varied between 50 and 150 nmol reduced glutathione· min-1·mg protein-1, and the apparent Km for NADPH was 0.125 and 0.200 mM for heterocysts and vegetative cells, respectively. The enzyme was found to be sensitive to Zn+2 ions, however, preincubation with oxidized glutathione rendered its resistance to Zn+2 inhibition. Nostoc muscorum filaments were found to contain 0.6–0.7mM glutathione, and it is suggested that glutathione reductase can regenerate reduced glutathione in both cell types. The combined activity of glutathione reductase and isocitrate dehydrogenase in heterocysts was as high as 18 nmol reduced glutathione·min-1·mg protein-1. A relatively high superoxide dismutase activity was found in the two cell types; 34.2 and 64.3 enzyme units·min-1·mg protein-1 in heterocysts and vegetative cells, respectively.We suggest that glutathione reductase plays a role in the protection mechanism which removes oxygen radicals in the N2-fixing cyanobacterium Nostoc muscorum.Abbreviations DTNB 5-5-dithiobis-(2-nitrobenzoic acid) - EDTA ethylenediaminetetra-acetic acid - GR glutathione reductase (EC1.6.4.2) - GSH reduced glutathione - GSSG oxidized glutathione - OPT O-phtaldialdehyde - SOD superoxide dismutase (EC 1.15.1.1)  相似文献   

5.
As the only freshwater lake in Israel, Lake Kinneret serves a number of important functions which directly rely upon the viability of the water. The annual outbreak of a dinoflagellate bloom strictly governs the nature of the macro and micro food web and ultimately determines water quality.The freshwater dinoflagellatePeridinium gatunense is subject to a wide range of environmental stresses throughout the spring bloom period. It was confirmed that SOD played an important antioxidative maintainance role throughout the bloom, especially during periods of relatively high photosynthetic activity (820 mg C m–2 day–1), when activity reached 500 Units SOD mg protein–1. In addition, high light stress (>300 mol photons m–2 s–1) induced SOD activity, despite the low dissolved inorganic carbon (DIC) concentrations at the end of the bloom (1.3 mM). Catalase activity was only significant at the end of the bloom, peaking at 120 mol O2 mg protein–1 min–1, when induced by photorespiratory activity.A series of experiments withPeridinium cultures showed that 2 × 10–4 M ascorbate inhibited catalase activity >50% within 15 min incubation, bothin vivo andin vitro. It is suggested that the high concentrations of ascorbate, found previously inPeridinium during early and mid-bloom (0.2–1.6 mM), not only eliminate H2O2 build-up, but also prevent (directly or indirectly) the induction of catalase.  相似文献   

6.
Cell suspension cultures of Taxus chinensis, supplemented with 25 g sucrose l–1, produced 11 mg cephalomanine l–1, 21 g biomass l–1 and 19 nkat geranylgeranyl diphosphate (GGPP) synthase activity g protein–1. Supplementation of the cultures with 100 M methyl jasmonate (MJA) produced 17 mg cephalomanine l–1, 6 g biomass l–1 and 78 nkat GGPP synthase activity g protein–1. Addition of sucrose and MJA together produced 24 mg cephalomanine l–1, 18 g biomass l–1 and 55 nkat GGPP synthase activity g protein–1.  相似文献   

7.
Summary Two new forms of the plasma membrane ATP-ase ofMicrococcus lysodeikticus NCTC 2665 were isolated from a sub-strain of the microorganism by polyacrylamide gel electrophoresis. One of them had a mol.wt of 368,000 and a very low specific activity (0.80 µ mol.min–1.mg protein–1) that could not be stimulated by trypsin. This form has been called BI (strain B, inactive). If the electrophoresis was carried out in the presence of reducing agents (i.e., dithiothreitol) and the pH of the effluent maintained at a value of 8.5 another form of the enzyme was obtained. This had a mol.wt of 385,000 and a specific activity of 2.5–5.0 µ mol.min–1.mg protein–1 that could be stimulated by trypsin to 5–10 µ mol.min–1.mg protein–1. This preparation of the ATPase has been called form BA (strain B, enzyme active). The subunit composition of both forms has been studied by sodium dodecyl sulphate and urea gel electrophoresis and compared to that of the enzyme previously purified from the original strain (form A). The three forms of the enzyme had similar and subunits, with mol.wt of about 50,000 and 30,000 dalton, respectively. They also had in common the component(s) of relative mobility 1.0, whose status as true subunit(s) of the enzyme remains yet to be established. However, subunit, that had a mol.wt of about a 52,500 in form A (Andreu et al. Eur. J. Biochem. (1973) 37, 505–515), had a mol.wt similar to in form BI and about 60,000 in form BA. Furthermore BA usually showed two types of this subunit ( and) and an additional peptide chain () with a mol.wt of about 25,000 dalton. This latter subunit seemed to account for the stimulation by trypsin of form BA.Forms BA could be converted to BI by storage and freezing and thawing. Conventional protease activity could not be detected in any of the purified ATPase forms and addition of protease inhibitors to form BA failed to prevent its conversion to form BI. The low activity form (BI) was more stable than the active forms of the enzyme and also differed in its circular dichroism. These results show thatM. lysodeikticus ATPase can be isolated in several forms. Although these variations may be artifacts caused by the purification procedures, they provide model systems for understanding the structural and functional relationships of the enzyme and for drawing some speculations about its functionin vivo.  相似文献   

8.
Biological oxidation rates of CS2 with a mixed microbial culture obtained from a trickling filter were optimal with 3 mM CS2, pH 7, 30°C and SO4 2– below 25 g l–1. Degradation rates were 3.4 mg CS2/gproteinmin and 13.8 mg H2S/gproteinmin. The concentrations of intermediates (H2S, COS and S°) and the product (SO4 2–) of CS2 oxidation were measured. The biological oxidation was due principally to Gram negative bacteria.  相似文献   

9.
Methyl mercury uptake in free cells and different immobilizates of the cyanobacteriumNostoc calcicola has been examined. The general growth of the immobilized cyanobacterial cells could be negatively correlated with methyl mercury uptake. Alginate spheres proved most efficient in terms of uptake rate (0.48 nmol mg protein–1 min–1, 10 min) and total bioaccumulation (10.71 nmol mg protein–1, 1 h) with a bioconcentration factor of 3.3×103. Alginate biofilms showed a faster methyl mercury accumulation rate (0.83 nmol mg protein–1 min–1, 10 min) with a saturation of 10.28 nmol mg protein–1 reached within only 30 min (bioconcentration factor, 3.1×103). Foam preparations with a slow initial uptake approximated biofilms but were characterized by a lower bioconcentration factor (2.8×103). Free cells, in comparison, maintained the initial slow rate of uptake (0.62 nmol mg protein–1 min–1, 10 min), saturating at 30 min (8.81 nmol mg protein–1), and the resultant lowest bioconcentration factor (2.7×103). Cell ageing (30 days) brought a drastic reduction (3-fold) in organomercury uptake by free cells while alginate spheres maintained the same potential. Foam preparations of the same age showed a significant improvement in methyl mercury uptake followed by only a marginal decline in alginate biofilms. Data are discussed in the light of the physiological efficiency and longevity of immobilized cells.  相似文献   

10.
The initial reactions involved in anaerobic aniline degradation by the sulfate-reducing Desulfobacterium anilini were studied. Experiments for substrate induction indicated the presence of a common pathway for aniline and 4-aminobenzoate, different from that for degradation of 2-aminobenzoate, 2-hydroxybenzoate, 4-hydroxybenzoate, or phenol. Degradation of aniline by dense cell suspensions depended on CO2 whereas 4-aminobenzoate degradation did not. If acetyl-CoA oxidation was inhibited by cyanide, benzoate accumulated during degradation of aniline or 4-aminobenzoate, indicating an initial carboxylation of aniline to 4-aminobenzoate, and further degradation via benzoate of both substrates. Extracts of alinine or 4-aminobenzoategrown cells activated 4-aminobenzoate to 4-aminobenzoyl-CoA in the presence of CoA, ATP and Mg2+. 4-Aminobenzoyl-CoA-synthetase showed a K m for 4-aminobenzoate lower than 10 M and an activity of 15.8 nmol · min-1 · mg-1. 4-Aminobenzoyl-CoA was reductively deaminated to benzoyl-CoA by cell extracts in the presence of low-potential electron donors such as titanium citrate or cobalt sepulchrate (2.1 nmol · min-1 · mg-1). Lower activities for the reductive deamination were measured with NADH or NADPH. Reductive deamination was also indicated by benzoate accumulation during 4-aminobenzoate degradation in cell suspensions under sulfate limitation. The results provide evidence that aniline is degraded via carboxylation to 4-aminobenzoate, which is activated to 4-aminobenzoyl-CoA and further metabolized by reductive deamination to benzoyl-CoA.  相似文献   

11.
During the fermentation of sugars to ethanol relatively high levels of an undesirable coproduct, ethyl acetate, are also produced. With ethanologenic Escherichia coli strain KO11 as the biocatalyst, the level of ethyl acetate in beer containing 4.8% ethanol was 192 mg liter−1. Although the E. coli genome encodes several proteins with esterase activity, neither wild-type strains nor KO11 contained significant ethyl acetate esterase activity. A simple method was developed to rapidly screen bacterial colonies for the presence of esterases which hydrolyze ethyl acetate based on pH change. This method allowed identification of Pseudomonas putida NRRL B-18435 as a source of this activity and the cloning of a new esterase gene, estZ. Recombinant EstZ esterase was purified to near homogeneity and characterized. It belongs to family IV of lipolytic enzymes and contains the conserved catalytic triad of serine, aspartic acid, and histidine. As expected, this serine esterase was inhibited by phenylmethylsulfonyl fluoride and the histidine reagent diethylpyrocarbonate. The native and subunit molecular weights of the recombinant protein were 36,000, indicating that the enzyme exists as a monomer. By using α-naphthyl acetate as a model substrate, optimal activity was observed at pH 7.5 and 40°C. The Km and Vmax for α-naphthyl acetate were 18 μM and 48.1 μmol·min−1·mg of protein−1, respectively. Among the aliphatic esters tested, the highest activity was obtained with propyl acetate (96 μmol·min−1·mg of protein−1), followed by ethyl acetate (66 μmol·min−1·mg of protein−1). Expression of estZ in E. coli KO11 reduced the concentration of ethyl acetate in fermentation broth (4.8% ethanol) to less than 20 mg liter−1.  相似文献   

12.
An intracellular -glucosidase was isolated from the cellobiose-fermenting yeast, Candida wickerhamii. Production of the enzyme was stimulated under aerobic growth, with the highest level of production in a medium containing cellobiose as a carbohydrate source. The molecular mass of the purified protein was approximately 94 kDa. It appeared to exist as a dimeric structure with a native molecular mass of about 180 kDa. The optimal pH ranged from 6.0 to 6.5 with p-nitrophenyl -d-glucopyranoside (NpGlc) as a substrate. The optimal temperature for short-term (15-min) assays was 35°C, while temperature-stability analysis revealed that the enzyme was labile at temperatures of 28° C and above. Using NpGlc as a substrate, the enzyme was estimated to have a K m of 0.28 mM and a V max of 525 mol product min–1 mg protein–1. Similar to the extracellular -glucosidase produced by C. wickerhamii, this enzyme resisted end-product inhibition by glucose, retaining 58% of its activity at 100 mM glucose. The activity of the enzyme was highest against aryl -1,4-glucosides. However, p-nitrophenyl xylopyranoside, lactose, cellobiose, and trehalose also served as substrates for the purified protein. Activity of the enzyme was stimulated by long-chain n-alkanols and inhibited by ethanol, 2-propanol, and 2-butanol. The amino acid sequence, obtained by Edman degradation analysis, suggests that this -glucosidase is related to the family-3 glycosyl hydrolases.  相似文献   

13.
Salmonella enterica forms polyhedral bodies involved in coenzyme-B12-dependent 1,2-propanediol degradation. Prior studies showed that these bodies consist of a proteinaceous shell partly composed of the PduA protein, coenzyme-B12-dependent diol dehydratase, and additional unidentified proteins. In this report, we show that the PduP protein is a polyhedral-body-associated CoA-acylating aldehyde dehydrogenase important for 1,2-propanediol degradation by S. enterica. A PCR-based method was used to construct a precise nonpolar deletion of the gene pduP. The resulting pduP deletion strain grew poorly on 1,2-propanediol minimal medium and expressed 105-fold less propionaldehyde dehydrogenase activity (0.011 mol min–1 mg–1) than did wild-type S. enterica grown under similar conditions (1.15 mol min–1 mg–1). An Escherichia coli strain was constructed for high-level production of His8-PduP, which was purified by nickel-affinity chromatography and shown to have 15.2 mol min–1 mg–1 propionaldehyde dehydrogenase activity. Analysis of assay mixtures by reverse-phase HPLC and mass spectrometry established that propionyl-CoA was the product of the PduP reaction. For subcellular localization, purified His8-PduP was used as antigen for the preparation of polyclonal antiserum. The antiserum obtained was shown to have high specificity for the PduP protein and was used in immunogold electron microscopy studies, which indicated that PduP was associated with the polyhedral bodies involved in 1,2-propanediol degradation. Further evidence for the localization of the PduP enzyme was obtained by showing that propionaldehyde dehydrogenase activity co-purified with the polyhedral bodies. The fact that both Ado-B12-dependent diol dehydratase and propionaldehyde dehydrogenase are associated with the polyhedral bodies is consistent with the proposal that these structures function to minimize propionaldehyde toxicity during the growth of S. enterica on 1,2-propanediol.  相似文献   

14.
An assay is described that allows the direct measurement of the enzyme activity catalyzing the transfer of the methyl group from N 5-methyltetrahydromethanopterin (CH3–H4MPT) to coenzyme M (H–S–CoM) in methanogenic archaebacteria. With this method the topology, the partial purification, and the catalytic properties of the methyltransferase in methanol- and acetate-grown Methanosarcina barkeri and in H2/CO2-grown Methanobacterium thermoautotrophicum were studied. The enzyme activity was found to be associated almost completely with the membrane fraction and to require detergents for solubilization. The transferase activity in methanol-grown M. barkeri was studied in detail. The membrane fraction exhibited a specific activity of CH3–S–CoM formation from CH3–H4MPT (apparent K m=50 M) and H–S–CoM (apparent K m=250 M) of approximately 0.6 mol·min-1·mg protein-1. For activity the presence of Ti(III) citrate (apparent K m=15 M) and of ATP (apparent K m=30 M) were required in catalytic amounts. Ti(III) could be substituted by reduced ferredoxin. ATP could not be substituted by AMP, CTP, GTP, S-adenosylmethionine, or by ATP analogues. The membrane fraction was methylated by CH3–H4MPT in the absence of H–S–CoM. This methylation was dependent on Ti(III) and ATP. The methylated membrane fraction catalyzed the methyltransfer from CH3–H4MPT to H–S–CoM in the absence of ATP and Ti(III). Demethylation in the presence of H–S–CoM also did not require Ti(III) or ATP. Based on these findings a mechanism for the methyltransfer reaction and for the activation of the enzyme is proposed.Abbreviations H4MPT tetrahydromethanopterin - CH3–H4MPT N 5-methyl-H4MPT - H–S–CoM 2-mercaptoethanesulfonate or coenzyme M - CH3–S–CoM 2(methylthio)ethanesulfonate or methylcoenzyme M - SDS-PAGE sodium dodecylsulfate polyacrylamide gel electrophoresis - DTT dithiothreitol - MOPS morpholinopropanesulfonate - CHAPS 3-[(3-cholamidopropyl)-dimethylammonio]-1-propane-sulfonate - 1 U = 1 mol/min  相似文献   

15.
Summary Lactate concentration was measured in the abdominal muscle of the shrimpPalaemon serratus. Rapid and seasonal temperature changes result in an increase of the lactate content of approximately 3–4 fold.Lactate dehydrogenase from the abdominal muscle exhibits a temperature dependent pyruvate inhibition with pyruvate as substrate.The kinetic parameters of lactate dehydrogenase fromPalaemon serratus are found to vary during rapid temperature changes: Vmax increases with temperature from 0.06 mol min–1 (mg protein)–1 at 10°C to 0.28 mol min–1 (mg protein)–1 at 30°C with lactate as substrate, and from 5.5 mol min–1 (mg protein)–1 at 10°C to 26.2 mol min–1 (mg protein)–1 at 30°C, with pyruvate (Table 1). The Hill coefficientn H, decreases with temperature from 2.2 to 1.2 when the pyruvate reduction is examined, but remains near 1.2 when the activity is measured with lactate as substrate (Table 1). The S0.5 values for lactate show a tendency to increase below 30 °C (18.9 mM l–1 at 20 °C) whereas the S0.5 for pyruvate is found to increase greatly with temperature (0.004 mM l–1 at 10 °C and 0.06 mM l–1 at 20 °C).Long term temperature changes involve variations of lactate dehydrogenase activity leading to inverse thermal compensation (Table 2).Activation energy (about 56 kJ both with pyruvate and lactate) does not vary during the year, suggesting that temperature adaptation does not induce important catalytic changes (Table 3).Abbreviation LDH lactate dehydrogenase  相似文献   

16.
The uptake of radioactive ethanolamine has been studied in exclusively neuronal and glial cell cultures from dissociated cerebral hemispheres of chick embryos. Both cell types show saturable kinetics; neurons have an apparentK m of 6.7 M,V max 41.4 pmol mg prot.–1 min–1 and glial cells aK m of 119.6 M,V max 3,917 pmol mg prot–1 min–1. The lower affinity of the transport and the 100 fold increase inV max observed in glial cells correlated with a more important accumulation of free ethanolamine found in glial cells and with a higher degree of phosphorylation of ethanolamine. The uptake appeared to be temperature and Na+ ions dependent but was not affected by CN or ouabain. Monomethyl-, dimethylethanolamine and choline were effective in inhibiting the uptake. Little or no effect was observed with serine, methionine, carnitine, alanine or glutamate.  相似文献   

17.
Two different immobilisation techniques for lipases were investigated: adsorption on to Accurel EP-100 and deposition on to Celite. The specific activities were in the same order of magnitude, 2.9 (mol min–1 mg protein) when Celite was used as support and 2.3 (mol min–1 mg–1 protein) when Accurel EP-100 was used as support, even if the amount of lipase loaded differed by 2 orders of magnitude. Immobilisation on Accurel EP-100 was the preferred technique since 40–100 times more protein can be loaded/per g carrier, thus yielding a more active catalyst. The water activity profiles in lipase catalysed esterification were influenced by the amount of protein adsorbed to Accurel EP-100. Higher protein loading (40 mg g–1) resulted in a bell-shaped water activity profile with highest specific activity (6.1 mol min–1 mg–1 protein) at a w=0.11, while an enzyme preparation with low protein loading (4 mg g–1) showed highest specific activity at a w=0.75.  相似文献   

18.
Summary Acyl-CoA: lysophosphatidylcholine acyltransferase (LPCAT) (EC 2.3.1.23) activity was assayed in liver microsomes from rainbow trout,Salmo gairdneri, acclimated to 5°C and 20°C to assess its contribution to the temperature-induced restructuring of phospholipid acyl chain composition. The synthesis of phosphatidylcholine (PC) (from lyso-PC) was threefold the synthesis of phosphatidylethanolamine (PE) (from lyso-PE) under similar assay conditions. LPCAT activity (i) displayed an absolute requirement for lysophosphatidylcholine (LPC) and was enhanced by the presence of ATP, MgCl2 and CoA (which reduced the impact of endogenous acyl-CoA hydrolase activity by regenerating the acyl-CoA substrate) in the assay medium; (ii) remained linear with time up to 30 min; and (iii) increased linearly with microsomal protein concentration up to 0.2 mg/ml for the 20°C assay and 0.4 mg/ml for the 5°C assay. There was no difference in Km or Vmax values due to the acclimation history of the fish, but there were obvious differences due to assay temperature. The apparent Km values for LPC were 58.54±7.24 M and 12.26±2.14 M when assayed at 5°C and 20°C respectively; values for oleoyl-CoA were 9.11±0.78 M and 1.23±0.25 M under the same assay conditions. Activity was 1.99±0.31 nmol min–1 mg protein–1 when assayed at 5°C, and 3.8±0.45 nmol min–1 mg protein–1 when assayed at 20°C. These findings indicate that adjustments in the activity of LPCAT play no significant role in the temperature-induced restructuring of PC molecular species composition. However, the marked temperature dependence of the Km values for LPC and oleoyl CoA suggest that patterns of fatty acid incorporation (i.e. substrate preference) may vary with assay temperature, and in this way LPCAT could contribute to the restructuring response.Abbreviations PC phosphatidylcholine - PE phosphatidylethanolamine - LPCAT acyl-CoA: lysophosphatidylcholine acyltransferase - LPEAT acyl-CoA: lysophosphatidylethanolamine acyltransferase - LPC 1-palmitoyl,2-lysophosphatidylcholine  相似文献   

19.
The Ca2+-pumping ATPase of erythrocyte plasma membranes of hypertensive humans (HTN) show, in the absence ofcalmodulin, a low Vmax comparable to that of the enzyme of the erythrocyte membranes of normotensive humans (NTN). Although the addition of calmodulin (1.5 gper ml) increased the maximum activity of the calcium pump of membranes of HTN and NTN individuals by at least 2-fold and 4-fold, respectively, the activator protein partially purifed from the erythrocytes of HTN individuals enhanced the activity of the enzyme in a fashion similar to that of the protein obtained from the haemolysate of NTN individuals. A determination of the dependence of the activity of the pump on concentration of ATP revealed that the Km (ATP) of the enzyme of membranes of HTN individuals is 52% higher than that of the enzyme of membranes of NTN individuals, while the Vmax (1.75±0.28 mol ATP mg protein–1 h–1) of the pump is 46% lower in the membranes of HTN humans than that of the enzyme of membranes of normal individuals (3.25 ±0.42 mol ATP mg protein–1 h–1) . It seems likely from these results that elevated erythrocyte Ca2+ concentration associated with essential hypertension may be due to a defective interaction between the Ca2+-pumping ATPase and the calmodulin Ca2+ complex,  相似文献   

20.
Using the impedance cardiography method, heart rate ( c) matched changes on indexed stroke volume (SI) and cardiac output (CI) were compared in subjects engaged in different types of training. The subjects consisted of untrained controls (C), volleyball players (VB) who spent about half of their training time (360 min · week–1) doing anaerobic conditioning exercises and who had a maximal oxygen uptake ( ) 41% higher than the controls, and distance runners (D) who spent all their training time (366 min·week–1) doing aerobic conditioning exercises and who had a 26% higher than VB. The subjects performed progressive submaximal cycle ergometer exercise (10 W·min–1) up to c of 150 beats·min–1. In group C, SI had increased significantly (P<0.05) at c of 90 beats·min–1 ( + 32%) and maintained this difference up to 110 beats·min–1, only to return to resting values on reaching 130 beats·min–1 with no further changes. In group VB, SI peaked (+ 54%) at c of 110 beats·min–1, reaching a value significantly higher than that of group C, but decreased progressively to 22010 of the resting value on reaching 150 beats·min–1. In group D, SI peaked at c of 130 beats·min–1 (+ 54%), reaching a value significantly higher than that of group VB, and showed no significant reduction with respect to this peak value on reaching 150 beats·min–1. As a consequence, the mean CI increase per c unit was progressively higher in VB than in C (+46%) and in D than in VB (+ 105%). It was concluded that thef c value at which SI ceased to increase during incremental exercise was closely related to the endurance component in the training programme.  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