首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Ambient sea-water nitrate and tissue nitrogen (ethanol soluble nitrate and amino acids, as well as total nitrogen) of Macrocystis integrifolia Bory were monitored over a 2-yr period in Bamfield, Vancouver Island, British Columbia. Sea-water nitrate varied from a high of 12 μmol · 1?1 (individual values as high as 23 μmol · 1?1 were recorded) in late winter to below detection limits for most of the summer. Tissue nitrate and total nitrogen paralleled the ambient nitrate levels and showed summer minima and winter maxima (from 0 to 70 μmol · g fresh wt?1 for nitrate and from 0.8 to 2.9% of dry wt for total N). The nitrate uptake capacity was inversely proportional to tissue nitrate concentration and, furthermore, was much higher for subapical surface blades (60–70 nmol · cm?2 · h?1) than for older, deeper blades (5–10 nmol · cm?2 · h?1). Nitrate uptake by subapical blade disks in summer is apparently higher in dark (1.0–1.7 μmol · g fresh wt?1 · h?1) than in light (0.6–1.3 μmol · g fresh wt?1 · h?1) and the data obtained in 36–108 h experiments indicate nitrate pool sizes of 30–90 μmol · g fresh wt?1. These pools are 23 to nearly full in winter. Ammonium does not inhibit nitrate uptake. It is taken up and apparently utilized much faster than nitrate and it may well be an important source of nitrogen for marine macrophytes.  相似文献   

2.
W. Kaiser  W. Urbach 《BBA》1976,423(1):91-102
1. Dihydroxyacetone phosphate in concentrations ? 2.5 mM completely inhibits CO2-dependent O2 evolution in isolated intact spinach chloroplasts. This inhibition is reversed by the addition of equimolar concentrations of Pi, but not by addition of 3-phosphoglycerate. In the absence of Pi, 3-phosphoglycerate and dihydroxyacetone phosphate, only about 20% of the 14C-labelled intermediates are found in the supernatant, whereas in the presence of each of these substances the percentage of labelled intermediates in the supernatant is increased up to 70–95%. Based on these results the mechanism of the inhibition of O2 evolution by dihydroxyacetone phosphate is discussed with respect to the function of the known phosphate translocator in the envelope of intact chloroplasts.2. Although O2 evolution is completely suppressed by dihydroxyacetone phosphate, CO2 fixation takes place in air with rates of up to 65μ mol · mg?1 chlorophyll · h?1. As non-cyclic electron transport apparently does not occur under these conditions, these rates must be due to endogenous pseudocyclic and/or cyclic photophosphorylation.3. Under anaerobic conditions, the rates of CO2 fixation in presence of dihydroxyacetone phosphate are low (2.5–7 μmol · mg?1 chlorophyll · h?1), but they are strongly stimulated by addition of dichlorophenyl-dimethylurea (e.g. 2 · 10?7 M) reaching values of up to 60 μmol · mg?1 chlorophyll · h?1. As under these conditions the ATP necessary for CO2 fixation can be formed by an endogenous cyclic photophosphorylation, the capacity of this process seems to be relatively high, so it might contribute significantly to the energy supply of the chloroplast. As dichlorophenyl-dimethylurea stimulates CO2 fixation in presence of dihydroxyacetone phosphate under anaerobic but not under aerobic conditions, it is concluded that only under anaerobic conditions an “overreduction” of the cyclic electron transport system takes place, which is removed by dichlorophenyl-dimethylurea in suitable concentrations. At concentrations above 5 · 10?7 M dichlorophenyl-dimethylurea inhibits dihydroxyacetone phosphate-dependent CO2 fixation under anaerobic as well as under aerobic conditions in a similar way as normal CO2 fixation. Therefore, we assume that a properly poised redox state of the electron transport chain is necessary for an optimal occurrence of endogenous cyclic photophosphorylation.4. The inhibition of dichlorophenyl-dimethylurea-stimulated CO2 fixation in presence of dihydroxyacetone phosphate by dibromothymoquinone under anaerobic conditions indicates that plastoquinone is an indispensible component of the endogenous cyclic electron pathway.  相似文献   

3.
Light intensity and temperature interactions have a complex effect on the physiological process rates of the filamentous bluegreen alga Anabaena variabilis Kütz. The optimum temperature for photosynthesis increased with increasing light intensity from 10°C at 42 μE·m?2·s?1 to 35°C at 562 μE·m?2·s?1. The light saturation parameter, IK, increased with increasing temperatures. The maximum photosynthetic rate (2.0 g C·g dry wt.?1·d?1) occurred at 35°C and 564 μE·m?2·s?1. At 15°C, the maximum rate was 1.25 g C·g dry wt.?1·d?1 at 332 μE·m?2·s?1. The dark respiration rate increased exponentially with temperature. Under favorable conditions of light intensity and temperature the percent of extracellular release of dissolved organic carbon was less than 5% of the total C fixed. This release increased to nearly 40% under combinations of low light intensity and high temperature. A mathematical model was developed to simulate the interaction of light intensity and temperature on photosynthetic rate. The interactive effects were represented by making the light-saturation parameters a function of temperature.  相似文献   

4.
Methods were developed for obtaining highly viable mouse hepatocytes in single cell suspension and for maintaining the hepatocytes in adherent static culture. The characteristics of transferrin binding and iron uptake into these hepatocytes was investigated. (1) After attachment to culture dishes for 18–24 h hepatocytes displayed an accelerating rate of iron uptake with time. Immediately after isolation mouse hepatocytes in suspension exhibited a linear iron uptake rate of 1.14·105molecules/cell per min in 5 μM transferrin. Iron uptake also increased with increasing transferrin concentration both in suspension and adherent culture. Pinocytosis measured in isolated hepatocytes could account only for 10–20% of the total iron uptake. Iron uptake was completely inhibited at 4°C. (2) A transferrin binding component which saturated at 0.5 μM diferric transferrin was detected. The number of specific, saturable diferric transferrin binding sites on mouse hepatocytes was 4.4·104±1.9·104 for cells in suspension and 6.6·104±2.3·104 for adherent cultured cells. The apparent association constants were 1.23·107 1·mol?1 and 3.4·106 1·mol?1 for suspension and cultured cells respectively. (3) Mouse hepatocytes also displayed a large component of non-saturable transferrin binding sites. This binding increased linearly with transferrin concentration and appeared to contribute to iron uptake in mouse hepatocytes. Assuming that only saturable transferrin binding sites donate iron, the rate of iron uptake is about 2.5 molecules iron/receptor per min at 5 μM transferrin in both suspension and adherent cells and increases to 4 molecules iron/receptor per min at 10 μM transferrin in adherent cultured cells. These rates are considerably greater than the 0.5 molcules/receptor per min observed at 0.5 μM transferrin, the concentration at which the specific transferrin binding sites are fully occupied. The data suggest that either the non-saturable binding component donates some iron or that this component stimulates the saturable component to increase the rate of iron uptake. (4) During incubations at 4°C the majority of the transferrin bound to both saturable and nonsaturable binding sites lost one or more iron atoms. Incubations including 2 mM α,α′-dipyridyl (an Fe11 chelator) decreased the cell associated 59Fe at both 4 and 37°C while completely inhibiting iron uptake within 2–3 min of exposure at 37°C. These observations suggest that most if not all iron is loosened from transferrin upon interaction of transferrin with the hepatocyte membrane. There is also greater sensitivity of 59Fe uptake compared to transferrin binding to pronase digestion, suggesting that an iron acceptor moiety on the cell surface is available to proteolysis.  相似文献   

5.
Transport of l-proline into Saccharomyces cerevisiae K is mediated by two systems, one with a KT of 31 μM and Jmax of 40 nmol · s?1 · (g dry wt.)?1, the other with KT > 2.5 mM and Jmax of 150–165 nmol · s?1 · (g dry wt.)?1, The kinetic properties of the high-affinity system were studied in detail. It proved to be highly specific, the only potent competitive inhibitors being (i) l-proline and its analogs l-azetidine-2-carboxylic acid, sarcosine, d-proline and 3,4-dehydro-dl-proline, and (ii) l-alanine. The other amino acids tested behaved as noncompetitive inhibitors. The high-affinity system is active, has a sharp pH optimum at 5.8–5.9 and, in an Arrhenius plot, exhibits two inflection points at 15°C and 20–21°C. It is trans-inhibited by most amino acids (but probably only the natural substrates act in a trans-noncompetitive manner) and its activity depends to a considerable extent on growth conditions. In cells grown in a rich medium with yeast extract maximum activity is attained during the stationary phase, on a poor medium it is maximal during the early exponential phase. Some 50–60% of accumulated l-proline can leave cells in 90 min (and more if washing is done repeatedly), the efflux being insensitive to 0.5 mM 2,4-dinitrophenol and uranyl ions, to pH between 3 and 7.3, as well as to the presence of 10–100 mM unlabeled l-proline in the outside medium. Its rate and extent are increased by 1% d-glucose and by 10 μg nystatin per ml.  相似文献   

6.
Uptake of 14C-labelled sucrose and glucose by isolated seed coat halves of pea (Pisum sativum L. cv. Marzia) seeds was measured in the concentration range <0.1 μM to 100 mM. The initial influx of sucrose was strictly proportional to the external concentration, with a coefficient of proportionality (k) of 6.2 μmol·(g FW)?1·min?1·M?1. Sucrose influx was not affected by 10 μM carbonylcyanide m-chlorophenylhydrazone (CCCP), but it was inhibited by 40% in the presence of 2.5 mM p-chloromercuribenzenesulfonic acid (PCMBS). Influx with diffusional kinetics was also observed for glucose (k = 4.8 μmol·(g FW)?1·min ?1·M ?1) and mannitol (k = 5.1 μmol·(g FW)?1·min?1·M?1). For glucose an additional saturable system was found (Km = 0.26 mM, V max = 4.2 nmol·(g FW)?1·min?1), which appeared to be completely inhibited by CCCP and partly by PCMBS. In contrast to the diffusional pathway, uptake by this saturable system was slightly pH-dependent, with an optimum at pH 5.5. The influx of sucrose appears to be by the same pathway as the efflux of endogenous sucrose, which was inhibited by 36% in the presence of 2.5 mM PCMBS (De Jong A, Wolswinkel P, 1995, Physiol Plant 94: 78–86). It is argued that passive transport may be the only mechanism for sucrose transport through the plasma membrane of seed coat parenchyma cells. The estimated permeability coefficient of the plasma membrane for sucrose (P = 3.5·10?7 cm·s?1) is more than 1 × 106-fold higher than that reported for artificial lipid membranes. This relatively high permeability is hypothesized to result from pore-forming proteins that allow the diffusion of sucrose. Furthermore, it is shown that a sucrose gradient across the plasma membrane of the seed coat parenchyma of only 22 mM will suffice to result in the net efflux of sucrose which is required to feed the embryo.  相似文献   

7.
Activity of the enzyme glutamine synthetase (GS, EC 6.3.1.2) was determined in vitro for roots of the marine angiosperm Zostera marina L. (eelgrass) collected from a population in Great Harbor, Woods Hole, Massachusetts, U.S.A. The GS synthetase activity was lowest in roots of plants collected from the shallow region of the eelgrass bed (12.0 μmol·g−1 (fresh wt)· h−1) and increased in the mid (3.0 m, 40.3 μmol·g−1 (fresh wt)·h−1) and deep (5.0 m, 72.3 μmol·g−1 (fresh wt)·h−1) plant collection depths. GS transferase activity increased with collection depth in a similar manner: shallow, 28.6 μmol·g−1 (fresh wt)·h−1; mid, 52.0 μmol·g−1 (fresh wt)·h−1; deep, 92.8 μmol·g−1 (fresh wt)·h−1. When sediment-embedded plants were held in continuous darkness for 2 days to create extended root anoxia, root GS activities nearly doubled. In contrast, in vivo incorporation of 14C-glutamate into glutamine and protein residue remained constant or declined under short-term hypoxia and anoxia. During aerobic recovery from anoxia, root labelling of glutamine and protein increased markedly. Free amino acid patterns of eelgrass roots growing in situ were determined over a diurnal cycle. Total free amino acid content was maximal at dawn and decreased 50% by noon. In contrast, the proportion of glutamine was lowest at dawn and maximal at noon for both shallow and deep-growing plants. Despite differences in depth-specific plant sizes, root/rhizome/shoot ratios, and relative growth rates, the daily whole plant nitrogen demand of shallow and deep growing plants were equivalent. When corrected for assay temperature response, the enzyme synthetase activities measured in vitro suggest that all of the plant nitrogen assimilation requirements can be met within daylight hours during the period of peak summer biomass.  相似文献   

8.
The electrophoretic mobility of L5178Y cells in 0.0145 M NaCl, 4.5% sorbitol, 0.6 mM NaHCO3, pH 7.2, at 25°C was — 1.78 μ·s?1·V?1·cm?1 while that of an L-asparaginase resistant subline, L5178Y/ASN, was — 1.11 μm·s?1·V?1·cm?1. Both cell lines were characterized by terminal sialic acid residues on their surfaces. Treatment of L5178Y cells for 90 min with 10 units of L-asparaginase per ml in saline decreased the electrophoretic mobility of the cells to — 1.65 μm·s?1·V?1·cm?1 while treatment in Fischer's medium decreased the mobility to — 1.25 μm·s?1·V?1·cm?1; neither treatment had a significant effect on the L5178Y/ASN electrophoretic mobility. The results suggest that L-asparaginase has an immediate and specific effect on synthesis of cell surface asparaginyl glycoproteins.  相似文献   

9.
125I-labelled α2-macroglobulin-typrin complex (125I-labelled α2-macroglobulin·trypsin) was associated to isolated rat adipocytes and hepatocytes with a half-time of about 60 min at 37°C. The association of 0.5 μg/ml 125I-labelled α2-macroglobulin·trypsin was inhibited by unlabelled α2-macroglobulin·trypsin with a half-inhibition constant of about 8 μg/ml (11 nM). 125I-Labelled α2-macrioglubulin became cell-associated to a smaller extent (10–40% of that of α2-macroglobulin·trypsin) and the half-inhibition constant was about 35 μg/ml in adipocytes. The cell associated of 125I-labelled α-macroglobulin·trypsin was markedly inhibited by dansylcadaverin, bacitracin, omission of Ca2+ from the medium or pretreatment of the cell with trypsin. After incubation for 180 min more than 60% of the cell-associated 125-Ilabelled α2-macroglobulin·trypsin was not removed by treatment of the cells with trypsin-EDTA and represented probably internalized marterial. 125I-Labelled α2-macroglobulin·trypsin was degraded to trichloroacetic acid-soluble fragments by suspensions of both cell types but only to a negligible extent by incubation media preincubated with these cells. The rate of degradation of 0.5 μg/ml 125I-labelled α2-macroglobulin was approx. 40% of that of 125I-labelled α2-macroglobulin·trypsin. Degradation of 125I-labelled α2-macroglobulin·trypsin was abolished by a high concentration (0.5 mg/ml) and α2-macroglobulin·trypsin. It is concluded that α2-macroglobulin·trypsin by a specific and saturable mechanism is bound to, internalized and degraded by isolated rat adipocytes and hepatocytes.  相似文献   

10.
The transport of 2-methyl-4-amino-5-hydroxymethylpyrimidine (MAHMP) by Salmonella typhimurium was studied using synthetic [methyl-3H3]MAHMP. It was found that an active transport system existed for MAHMP, having Km of 0.07 μM and Vmax 45 nmol·min?1·(g dry wt. cells)?1, that required glucose as a source of energy and was pH and temperature dependent. Uptake was inhibited by cyanide, azide, N-ethylmaleimide, 2,4-dinitrophenol and carbonyl cyanide m-chlorophenylhydrazone. Uptake was also weakly inhibited by oxythiamine, but not by thiamine, 2-methyl-4-amino-5-aminomethylpyrimidine, or 4-amino-5-hydroxymethylpyrimidine, indicating that the transport system is specific for MAHMP.  相似文献   

11.
The presence of muscarinic receptors in islets of Langerhans was assessed by measurement of specific binding of [3H]methylscopolamine. Specific binding was defined as total binding minus binding obtained in the presence of 1000-fold or higher excess of unlabeled methylscopolamine. At 37°C specific binding was significant after 1 min and plateaued after 10 min of incubation. Displacement of label by increasing concentrations of unlabeled methylscopolamine indicated a dissociation constant of 1.5·10?12 M. Effects of methylscopolamine on insulin release were evaluated from the inhibitions of cholinergic-induced insulin release. 4·10?10 M methylscopolamine inhibited acetylcholine (20 μM)-induced insuliln release more than 60%. Binding was not influenced by the following variations during binding incubations: changing the glucose concentration from 0 to 83 mM, adding rotenon (1 μM) or omitting calcium from the incubation medium. Islets kept in tissue culture exhibited higher binding when cultured at 11.1 than at 3.3 mM glucose for 96 h. It is concluded that islets contain muscarinic receptors, the binding to which can be subject to alteration by the long-term glucose environment.  相似文献   

12.
Nef is a multifunctional gene of HIV which can increase virus replication either directly or by modulating the target cell's metabolism. Nevertheless the role of the exogenous Nef protein is not yet well understood. To investigate it, we studied the effects of the recombinant Nef protein on the expression of some antigens of lymphoid T‐cells permissive to HIV‐1 replication, and on their proliferation and on apoptosis induction. For this purpose, we utilised MT‐4 and H9 T‐cell lines. We evaluated FN (fibronectin), CD4 and CD71 expression in uninfected and acutely or chronically infected cells, both untreated and treated with Nef. Our studies showed a significant up‐regulation of FN especially in uninfected cells, with a dose of 2·5 μg ml−1; in contrast, a significant down‐modulation of CD4 and CD71 both in uninfected and in acutely or chronically infected cells, was detected. The proliferation of H9 uninfected cells was significantly reduced 24 h after treatment with Nef protein in a dose‐dependent manner (ranging from 0·02 to 2·5 μg ml−1); likewise a significant inhibition of proliferation of acutely and chronically infected cells was evident with 2·5 μg ml−1. Moreover, we demonstrated a dose‐dependent activity of Nef on inducing apoptosis in H9 uninfected cells and no effects of this protein on modulation of INF α and γ production in peripheral blood mononucleated cells of health donors. Nef appeared to be able to increase the effect of apoptotic stimuli. In conclusion, our data suggest that in our experimental system, the exogenous Nef protein can inhibit cellular synthesis facilitating the metabolic pathway involved in virus replication. In addition it modulates the susceptibility to the HIV‐1 infection and finally, that apoptosis induction or enhancement can facilitate the release of neoformed virions. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

13.
Novel techniques were devised to explore the mechanisms mediating the adverse effects of compacted soil on plants. These included growing plants in: (i) profiles containing horizons differing in their degree of compaction and; (ii) split-pots in which the roots were divided between compartments containing moderately (1·4 g cm ? 3) and severely compacted (1·7 g cm ? 3) soil. Wild-type and ABA-deficient genotypes of barley were used to examine the role of abscisic acid (ABA) as a root-to-shoot signal. Shoot dry weight and leaf area were reduced and root : shoot ratio was increased relative to 1·4 g cm ? 3 control plants whenever plants of both genotypes encountered severely compacted horizons. In bartey cultivar Steptoe, stomatal conductance decreased within 4 d of the first roots encountering 1·7 g cm ? 3 soil and increased over a similar period when roots penetrated from 1·7 g cm ? 3 into 1·4 g cm ? 3 soil. Conductance was again reduced by a second 1·7 g cm ? 3 horizon. These responses were inversely correlated with xylem sap ABA concentration. No equivalent stomatal responses occurred in Az34 (ABA deficient genotype), in which the changes in xylem sap ABA were much smaller. When plants were grown in 1·7 : 1·4 g cm ? 3 split-pots, shoot growth was unaffected relative to 1·4 g cm ? 3 control plants in Steptoe, but was significantly reduced in Az34. Excision of the roots in compacted soil restored growth to the 1·4 g cm ? 3 control level in Az34. Stomatal conductance was reduced in the split-pot treatment of Steptoe, but returned to the 1·4 g cm ? 3 control level when the roots in compacted soil were excised. Xylem sap ABA concentration was initially higher than in 1·4 g cm ? 3 control plants but subsequently returned to the control level; no recovery occurred if the roots in compacted soil were left intact. Xylem sap ABA concentration in the split-pot treatment of Az34 was initially similar to plants grown in uniform 1·7 g cm ? 3 soil, but returned to the 1·4 g cm ? 3 control level when the roots in the compacted compartment were excised. These results clearly demonstrate the involvement of a root-sourced signal in mediating responses to compacted soil; the role of ABA in providing this signal and future applications of the compaction procedures reported here are discussed.  相似文献   

14.
Adult rat heart muscle cells obtained by perfusion of the heart with collagenase have been used to characterize the insulin receptors by equilibrium binding and kinetic measurements. Binding of 125I-labelled insulin to heart cells exhibited a high degree of specificity; it was dependent on pH and temperature, binding at steady increased with decreasing temperatures. About 70% of the radioactivity bound at equilibrium at 25°C could be dissociated by addition of an excess of unlabelled insulin. 54 and 40% of 125I-labelled insulin was degraded by isolated heart cells after 2 h at 37°C and 4 h at 25°C, respectively. This degrading activity was effectively inhibited by high concentration of albumin.Equilibrium binding studies were conducted at 25°C using insulin concentrations ranging from 2.5 · 10?11 mol/l to 10?6 mol/l. Scatchard analysis of the binding data resulted in a curvilinear plot (concave upward), which was further analyzed using the average affinity profile. The empty site affinity constant was calculated to be 9.5 · 107 l/mol with a total receptor concentration of 3.4 · 106 sites per cell.The presence of site-site interactions of the negative cooperative type among the insulin receptors has been confirmed by kinetic experiments. The rate of dilution induced dissociation was enhanced in the presence of native insulin (5 · 10?9 mol/l), both, under conditions of low and high fractional saturation of receptors.  相似文献   

15.
Cellular nutrient concentrations and nutrient uptake rates of Cladophora glomerata (L.) Kuetzing were determined during summer and fall in 1989–1990 at a site on the upper Clark Fork of the Columbia River, Montana. Both physiological tests indicated that Cladophora growth is likely to be limited by nitrogen during late summer-early fall. Maximum uptake rates of ammonia-N and nitrate-N were 5935–6991 and 507–984 μg · g DW?1· h?1, respectively, during July–October when dissolved inorganic nitrogen (DIN) concentrations in the river were less than 10 μg · L?1. During November-December, when DIN was 72–376 μg · L?1, maximum ammonia-N uptake was 1137–1633 μg · g DW?1· h?1 and maximum nitrate-N uptake was 0–196 μg · g DW?1· h?1. Cellular nitrogen during summer–early fall was 0.78–1.80% of Cladophora dry weight, frequently at or below 1.1%, a level suggested as a critical minimum N concentration for maximum growth. In contrast, cellular P was 0.18–0.36% of dry weight, 3–6 times the suggested critical P concentration of 0.06%. Molar ratios of cellular N:P (< 16:1) and DIN: SRP (< 4:1) during late summer-early fall also indicated potential N limitation. Cellular N and P from Cladophora collected from a second site influenced by a municipal wastewater discharge in 1990 displayed similar seasonal trends. At both sites, seasonal fluctuations in DIN were closely tracked by changes in cellular N, Cellular P, however, increased through the growing season despite declining levels of SRP in the river.  相似文献   

16.
Combined gas chromatography-mass spectrometry (GCMS) was used to identify and quantify specific cytokinins from Porphyra perforate J. Ag. and Sargassum muticum (Yendo) Fensh. The level of isopentenyladenosine was estimated to be 0.6 μ·kg?1 fresh weight in Porphyra and 0.9 μ·kg?1 fresh weight in Sargassum. The level of cis-zeatin riboside was estimated to be 0.2 μ·kg?1 fresh weight in Sargassum. This is the first definitive identification of a cytokinin from a red alga, and the second report from a brown alga.  相似文献   

17.
CPTA and cycocel cause accumulation of lycopene and γ-carotene, simultaneously inhibiting the formation of β-carotene and β-zeacarotene in Phycomyces blakesleeanus mutant strain C115. Phytoene synthesis is enhanced. CPTA is more effective than cycocel. Kinetic studies show that with increasing concentrations of CPTA, lycopene and γ-carotene increase with the concomitant decrease in β-carotene, the total of these three carotenes being almost equal to β-carotene present in the control. When CPTA-treated mycelium is washed free of the chemical and resuspended in phosphate buffer solution containing 2·5% glucose (pH 5·6), β-carotene is formed at the expense of both γ-carotene and lycopene. β-Zeacarotene, which is not present in the mycelium, reappears upon resuspension. These results indicate that CPTA is inhibiting the enzymes causing cyclization both at neurosporene and lycopene levels. Studies on the effect of CPTA on the high lycopene mutant strain C9 reveal that with increasing concentrations of the compound, lycopene increases slightly and both β-carotene and γ-carotene decrease. Phytoene synthesis is stimulated up to a certain level of CPTA and then becomes steady. In the albino mutant strain C5, there is a slight increase in phytoene formation on the addition of CPTA to the medium. No other carotenoid is formed, suggesting that CPTA cannot remove the block caused by genetic mutation and exerts its influence in an already existing biosynthetic pathway.  相似文献   

18.
Aims: To investigate the effect of lactic acid (LA), copper (II), and monolaurin as natural antimicrobials against Cronobacter in infant formula. Methods and Results: The effect of LA (0·1, 0·2 and 0·3% v/v), copper (II) (10, 50 and 100 μg ml?1) and monolaurin (1000, 2000, and 3000 μg ml?1) suspended into tween‐80? or dissolved in ethanol against Cronobacter in infant formula was investigated. Reconstituted infant formula and powdered infant formula were inoculated with five strains of Cronobacter spp. at the levels of c. 1 × 106 CFU ml?1 and 1 × 103 CFU g?1, respectively. LA at 0·2% v/v had a bacteriostatic effect on Cronobacter growth, whereas 0·3% v/v LA resulted in c. 3 log10 reduction. Copper (II) at the levels of 50 μg ml?1 and 100 μg ml?1 elicited c. 1 and 2 log10 reductions, respectively. The combination of 0·2% LA and 50 μg ml?1 copper (II) resulted in a complete elimination of the organism. Monolaurin exhibited a slight inhibitory activity against Cronobacter (c. 1·5 log10 difference) compared to the control when ethanol was used to deliver monolaurin. Conclusions: A complete elimination of Cronobacter was obtained when a combination of sublethal concentrations of LA (0·2%) and copper (II) (50 μg ml?1) was used. Significance and Impact of the Study: The use of the synergistic interactive combination of LA and copper (II) could be beneficial to control Cronobacter in the infant formula industry.  相似文献   

19.
The steady-state rate of ATP synthesis in the isolated, Langendorff-perfused rat heart was determined using a 31P NMR saturation transfer method. At 37°C and a perfusion pressure of 70 cm H2O the value is 2.8 ± 0.3 (n=5 ± S.E.M.) μmol.s?1. (g. dry wt.)?1. The activity of creatine phosphokinase measured in the same experiments was 14.6 ± 1.0 μ mol.s?1 .(g. dry wt.)?1. From the rate of ATP synthesis and the separately measured oxygen consumption we calculated an apparent mitochondrial ADP:O ratio of 3.5 ± 0.8 in the intact tissue.  相似文献   

20.
Leakage of the entrapped anionic fluorophore carboxyfluorescein was used as a measure of the permeability of liposomes to several different acids. Carboxyfluorescein leakage increased with increasing buffer concentration at a given pH and depended on its chemical nature: apolar weak acids such as acetic or pyruvic acids induced fast leakage at relatively high pH (4 to 5), while glycine, aspartic, citric and hydrochloric acids induced leakage only at lower pH. Fluorescence leakage measurements reflected the acidification of the liposomes' aqueous spaces, which was primarily caused by the diffusion of undissociated acid molecules across the lipid bilayer. A simple mathematical model in accord with this hypothesis and assuming that carboxyfluorescein leakage was directly related to the proportion of its neutral lactone form, described satisfactorily the carboxyfluorescein leakage kinetics and allowed rough estimation of permeability coefficients for carboxyfluorescein (neutral lactone form; 9 · 10?9 cm · s?1), acetic acid (>1 · 10?7cm · s?1) and glycine (cation: 6 · 10?9 cm · s?1). These results are consistent with low effective proton permeability of liposomes (<5 · 10?12cm · s?1) and with the permeability coefficient of HCl (3 · 10?3 cm · s?1) reported by Nozaki and Tanford ((1981) Proc. Natl. Acad. Sci. U.S.A. 78, 4324–4328). Diffusion of weak acid molecules across lipid membranes has implications for drug encapsulation and delivery, and may be of biological significance.  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