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 156 毫秒
1.
Summary The seed of peach fruits develop the capacity to produce ethylene with a lag phase of about 1 h after excision. The site of ethylene synthesis is in the seed coat and rates as high as 6,000 l kg–1 h–1 were recorded. Ethylene production was reduced to less than 1% of the control by 10 g/ml cycloheximide. Although the tissue had only a small methionine pool, supplying the seed with exogenous methionine did not influence ethylene production at any stage of seed development. Label from [U-14C]methionine was readily incorporated into ethylene.  相似文献   

2.
Dormancy in seeds of Manihot glaziovii is overcome at 25°C by application of ethrel at effective ethylene concentrations equal to and greater than 10 ll–1. Imbibition of seeds in ethrel broadens the temperature optimum for germination but does not prevent the development of secondary dormancy at temperatures of 35°C and greater and 15°C and lower. Secondary dormancy must be due to factors other than reduced ethylene production.  相似文献   

3.
The object of this work was to determine, using a full-factorial experiment, the influence of temperature, irradiance and salinity on growth and hepatotoxin production by Nodularia spumigena, isolated from Lake Alexandrina in the south-east of South Australia. Higher levels of biomass (determined as particulate organic carbon, POC), toxin production and intracellular toxin concentration per mg POC were produced under light limited conditions (30 mol m–2 s–1) and at salinities equal to or greater than those experienced in Lake Alexandrina. Both highest biomass and total toxin production rates were recorded at temperatures equal to or greater than those of the lake (20 and 30°C). The temperature at which maximum biomass and toxin production was recorded decreased from 30°C for cultures grown at 30 mol m–2 s–1 to 20°C when grown at 80 mol m–2 s–1. In contrast, intracellular toxin per mg POC was highest at the lowest growth temperature, 10°C, at both 30 and 80 mol m–2 s–1. It appears that the optimum temperature for biosynthetic pathways used in the production of toxin is lower than the optimum temperature for those pathways associated with growth. Intracellular toxin levels were higher in cells cultured at 10°C/30 mol m–2 s–1 whereas the majority of the toxin was extracellular in cells grown at 30°C/30 mol m–2 s–1. This implies that the highest concentration of toxin in lake water would occur under high temperature and high irradiance conditions. Individual environmental parameters of salinity, irradiance and temperature were all shown to influence growth and toxin production. Notwithstanding, the overall influence of these three parameters on toxin production was mediated through their effect upon growth rate.  相似文献   

4.
Summary Respiration of an undescribed species of soil nematode of the genus Chiloplacus from the Canadian High Arctic was measured at 2°, 5°, 10°, 15°, 20° and 25°C. The corresponding metabolic rates were 0.2697×10-3 l, 0.3406×10-3 l, 0.8408×10-3 l, 0.8539×10-3 l, 1.8420×10-3 l and 2.9360×10-3 l O2 ind-1 h-1, respectively, for a nematode of 1.0 g dry weight. The relationship between respiration and dry weight for Chiloplacus sp. at 10°C is described by the function log R=-3.0693+0.8844 log W. Q10 values for the 2°–5°, 5°–10°, 10°–15°, 15°–20° and 20°–25°C temperature intervals were 2.18, 6.09, 1.03, 4.65 and 2.54, respectively. Chiloplacus sp. showed raised metabolic rates at low tempetatures compared with species from warmer environments. Metabolic rates of representative samples of the soil, nematode fauna (dominated by individuals of the genus Plectus) from the same location were 0.1593×10-3 l, 0.3603×10-3 l and 0.5332×10-3 l O2 ind-1 h-1 at 5°, 10° and 15°C for an average nematode of 0.4297 g dry weight.  相似文献   

5.
Dormant Amaranthus retroflexus seeds do not germinate in the dark at temperatures below 35°C. Fully dormant seeds germinate only at 35–40°C whereas non-dormant ones germinate within a wider range of temperatures (15 to 40°C). Germination of non-dormant seeds requires at least 10% oxygen, but the sensitivity of seeds to oxygen deprivation increases with increasing depth of dormancy. 10–6 to 10–4 M ethephon, 10–3 M 1-aminocyclopropane 1-carboxylic acid (ACC) and 10–3 M gibberellic acid (GA3) break this dormancy. In the presence of 10–3 M GA3 dormant seeds are able to germinate in the same range of temperatures as non-dormant seeds. The stimulatory effect of GA3 is less dependent on temperature than that of ethephon, while ACC stimulates germination only at relatively high temperatures (25–30°C). The results obtained are discussed in relation to the possible involvement of endogenous ethylene in the regulation of germination of A. retroflexus seeds.Abbreviations ACC 1-aminocyclopropane 1-carboxylic acid - GA3 gibberellic acid - SD standard deviation  相似文献   

6.
Summary In order to obtain a better understanding of the behaviour ofPediococcus pentosaceus in food products as well to facilitate the designing of industrial production processes for the organism, the growth and lactic acid production ofPediococcus pentosaceus in a complex glucose medium was followed in batch cultures at different gas environments (CO2, air, N2 and static cultures without gasflow), temperatures (10–50°C), pH (4.3–7.3) and nitrite concentrations (0–700 ppm). Optimal growth was obtained in CO2 at 40°C and pH 6.3 and resulted in a maximum specific growth rate ( max) of 1.27 h–1. In static culture at 40°C and pH 6.3 the max was 1.21 h–1. The max was, compared with static culture, reduced in air (12%) and nitrogen (26%). At 10°C the max was reduced by 99% and at 50°C by 88%. The reduction at pH 4.3 and 7.3 was 65% and 57%, respectively. Nitrite did not affect the max at any pH but increased the lag phase at pH 4.3 by a factor of 12. The lactic acid production was linked to the growth. The total amount of lactic acid produced was the same in all the tested gases and nitrite concentrations and also within the wide temperature range (15–45°C) and pH range (5.3–7.3). Mainly L(+)-lactic acid was produced during the exponential growth phase, but after this growth declined about 30% of the L(+)-lactic acid was converted to D(–)-lactic acid. The lactic acid product yield and the cellmass yied were both affected by the temperature but not by the pH.  相似文献   

7.
Plants of Solanum tuberosum L. potato do not cold acclimate when exposed to low temperature such as 5°C, day/night. When ABA (45 M) was added to the culture medium, stem-cultured plantlets of S. tuberosum, cv. Red Pontiac, either grown at 20°C/15°C, day/night, or at 5°C, increased in cold hardiness from –2°C (killing temperature) to –4.5°C. The increase in cold hardiness could be inhibited in both temperature regimes if cycloheximide (70 M) was added to the culture medium at the inception of ABA treatment. Cycloheximide did not inhibit cold hardiness development, however, when it was added to the culture medium 3 days after ABA treatment.When pot-grown plants were foliar sprayed with mefluidide (50 M), ABA content increased from 10 nmol to 30 nmol g–1 dry weight and plants increased in cold hardiness from –2°C to about –3.5°C. The increases in free ABA and cold hardiness occurred only in plants grown at 20°C/15°C; neither ABA nor cold hardiness increased in plants grown at 5°C.The results suggest that an increase in ABA and a subsequent de novo synthesis of proteins are required for the development of cold hardiness in S. tuberosum regardless of temperature regime, and that the inability to synthesize ABA at low temperature, rather than protein synthesis, appears to be the reason why S. tuberosum does not cold acclimate.  相似文献   

8.
Summary Gossypium hirsutum L. var. Delta Pine 61 was cultivated in controlled-environment chambers at 1000–1100 mol photosynthetically active photons m-2 s-1 (medium photon flux density) and at 1800–2000 mol photons m-2 s-1 (high photon flux density), respectively. Air temperatures ranged from 20° to 34°C during 12-h light periods, whereas during dark periods temperature was 25° C in all experiments. As the leaf temperature decreased from about 33° to 27° C, marked reductions in dry matter production, leaf chlorophyll content and photosynthetic capacity occurred in plants growing under high light conditions, to values far below those in plants growing at 27° C and medium photon flux densities. The results show that slightly suboptimum temperatures, well above the so-called chilling range (0–12° C), greatly reduce dry matter production in cotton when combined with high photon flux densities equivalent to full sunlight.Abbreviations DW dry weight - F v variable fluorescence yield - F M maximum fluorescence yield - PFD photon flux density (400–700 nm)  相似文献   

9.
The kinetics of CNProto- and CNDeutero-hemin binding to apohemoglobin A2 was investigated in a stopped-flow device in 0.05 M potassium phosphate buffer, pH 7, at 10°C. The overall kinetic profile exhibited multiple phases: Phases I–IV corresponding with heme insertion (8.5–13 × 107 M–1 s–1), local structural rearrangement (0.21–0.23 s–1), global structural event (0.071–0.098 s–1), and formation of the Fe–His bond (0.009–0.012 s–1), respectively. Kinetic differences observed between apohemoglobin A2 and apohemoglobin A (previously studied) prompted an analysis of the structures of and chains through molecular modeling. This revealed a structural repositioning of the residues not only at, but also distant from the site of the amino acid substitutions, specifically those involved in the heme contact and subunit interface. A significant global change was observed in the structure of the exon-coded 3 region and provided additional evidence for the designation of this as the subunit assembly domain.  相似文献   

10.
The effects of daminozide (butanedioic acid-2,2-dimethylhydrazide) on ethylene synthesis by apple fruits were investigated. The objective was to determine the effects of postharvest applications as compared to the standard application of diaminozide in the orchard. Immersion in a solution containing 4.25 g L–1 active ingredient for 5 min delayed the rise in ethylene production in individual Cox apples at 15°C by about 2 days, whereas orchard application of 0.85 g L–1 caused delays of about 3 days. Both modes of application depressed the maximal rate of ethylene production attained by ripe apples by about 30%. Daminozide did not affect the stimulation of respiration by ethylene treatment of Gloster apples, but it delayed the increase in ethylene synthesis. Daminozide applied immediately after harvest delayed the rise in ethylene synthesis in Golden Delicious held at 15°C, but it was less effective when applied 48 h after harvest or when apples were held at 5°C. Exposure to 1–2 l L–1 ethylene for 48 h was less effective in promoting the rise in ethylene in daminozidetreated Cox and Gloster apples than in untreated fruit. High (100–1000 l L–1) concentrations of ethylene more or less overcame the daminozide effect. Apples absorbed about 40% of surface-applied [14C]daminozide in 48 h, but more than 90% of the radioactivity in the fruit was recovered from the peel and outer 1 cm of the cortex. Daminozide was partly converted to carbon dioxide and other metabolites.  相似文献   

11.
Summary Four 1,3--glucanases GI, GII, GIV and GVIII from a culture filtrate ofStreptomyces sp. 1228 were purified by anion exchange chromatography using DEAE-Sepharose Cl-6B or DEAE-Cellulose, gel filtration on Bio-Gel P-200 or Sephacryl S-200, Amicon ultrafiltration and preparative PAGE. The Mr of these enzymes were 19000, 74000, 78000 and 56000 respectively. The glucanase GVIII consisted of two subunits. The optimal catalytic activity of the purified preparations was at 50–55°C and pH 5.5–6.0. The enzymes were also most stable at this pH. Both glucanases GI and GVIII were characterized by high thermostability. The glucanases showed different affinities towards laminarin with Km values of 6.65 x 10–5 mol/l for GI, 2.35 x 10–4 mol/l for GII, 8.1 x 10–5mol/l for GIV and 8.1 x 10–4mol/l for GVIII. The presence of metal ions was not required for activity of these enzymes but thiol groups increased their activity. D-glucono--lactone did not inhibit the enzymes.  相似文献   

12.
In vitro cultures of Nephrolepis exaltata and Cordyline fruticosa were stored at 5°, 9° or 13°C, at a low irradiance (3–5 mol m–2 s–1) or in darkness. Prior to storage the cultures were subjected to 18°, 21°, 24° or 27°C and 15, 30 or 45 mol m–2 s–1 in a factorial combination.The optimal storage conditions for Nephrolepis were 9°C in complete darkness. These cultures were still transferable to a peat/perlite mixture at the end of the experimental period of 36 months.The optimal storage conditions for Cordyline were 13°C and a low light level (±3–5 mol m-2 s-1). When the pre-storage conditions were normal growth room conditions (24°C and 30 mol m-2 s-1), in vitro cultures could be stored for 18 months. With the most favourable pre-storage treatment (18°C and 15 mol m-2 s-1) some cultures still had green shoots after 36 months of storage, but did not survive transfer to peat/perlite.Pre-conditioning before storage was most favourable for Nephrolepis, and not that important, but still favourable, for Cordyline. There was an interaction between pre-storage temperature and pre-storage irradiance. For both species a high irradiance level was less favourable than a low irradiance level when combined with high growth room temperatures.Abbreviations BA 6-benzyladenine - IAA indole-3-acetic acid - NOA 2-naphthoxyacetic acid  相似文献   

13.
Ornithine transcarbamylase (OTCase) (E.C.2.1.3.3) was partially purified fromLactobacillus buchneri NCDO110. It was stabilized by the presence of glycerol. The optimal pH for enzyme activity is 8.5. The positive cooperativeness was observed among ornithine molecules at pH values different from the optimum. The Mr of the enzyme was calculated to be 162,000 by gel filtration on Ultrogel ACA-34. Maximum activity occurred at 35°C. G* of the reaction was calculated from Arrhenius plot. The values were 9100 cal mol–1 below 35°C, and 4300 cal mol–1 above 35°C. The Km value for carbamylphosphate was 7.1×10–4 M, and the Km for ornithine was 1.6×10–3 M, under the conditions described here. Dead-end inhibition analysis was performed with norvaline, which is a structural analogue of ornithine. Norvaline acted as a noncompetitive inhibitor when carbamylphosphate was the variable substrate, and as a competitive inhibitor when ornithine was the variable substrate. The results are consistent with a ping-pong mechanism.  相似文献   

14.
Summary A new, sensitive and continuous assay for -glucosidase is described exploiting the different angles of rotation for the substrate maltose and the product glucose. Kinetic experiments revealed a very pronounced product inhibition of -glucosidase fromSaccharomyces carlsbergensis with a Ki of 4.85·10–3 M for glucose.The KM of maltose was found to be 37.8·10–3 M. Taking these values, an integral kinetic curve for the enzymatic hydrolysis of maltose was calculated, which is shown to fit the experimental data.Symbols used k1 (min–1) pseudo first-order rate constant (for enzymatic cleavage) - k2 (min–1) rate constant (for mutarotation reaction) - I, P (mol/1) inhibitor (product) concentration - ki (mmol/1) inhibitor constant - KM (mmol/l) Michaelis constant - [M] 589 30 (degree/m · l/mol) molecular rotation at 30°C and 589 nm - s (mmol/l) substrate concentration - R (mmol/mg · min) reaction rate - Vmax (mmol/mg · min) maximal rate - U (mol/min) activity unit (here at 30°C and pH=6.8) Indices O initial value - max maximal value  相似文献   

15.
The role of ethylene in the formation of adventitious roots in vitro was studied in tomato (Lycopersicon esculentum Mill. cv. UC 105) cotyledons and lavandin (Lavandula officinalis Chaix × Lavandula latifolia microshoots. Both systems were able to form roots on hormone-free medium evolving low amounts of ethylene. The addition of 20–50 M indole-3-acetic acid (IAA) inhibited root formation in tomato cotyledons while increasing ethylene production. Naphthaleneacetic acid (NAA, 3 M) stimulated root number in lavandin explants and induced a transient rise in ethylene evolution. Enhanced ethylene levels via the endogenous precursors 1-aminocyclopropane-1-carboxylic acid (ACC, 25–50 M) drastically impaired root regeneration and growth in tomato. In lavandin, 10 M ACC stimulated ethylene production and significantly inhibited the rooting percentage and root growth. Conversely, ACC enhanced the root number in the presence of NAA only. Severe inhibition of rooting was also caused by ethylene reduction via biosynthetic inhibitors, aminoethoxyvinylglycine (AVG, 5–10 M) in tomato, and salicylic acid (SA, 100 M) in lavandin. A strict requirement of endogenous ethylene for adventitious root induction and growth is thus suggested.Abbreviations LS Linsmaier and Skoog medium - BA N6-benzyladenine - NAA 1-naphthaleneacetic acid - IAA Indole-3-acetic acid - AVG Aminoethoxyvinylglycine - SA Salicylic acid - ACC 1-aminocyclopropane-1-carboxylic acid  相似文献   

16.
Summary Respiratory gas exchange and blood respiratory properties have been studied in the East-African tree frogChiromantis petersi. This frog is unusually xerophilous, occupies dry habitats and prefers body temperatures near 40°C and direct solar exposure. Total O2 uptake was low at 81 l O2·g–1·h–1±19.0 (SD) at 25°C increasing to 253.5 l O2·g–1·h–1±94.8 (SD) at 40°C giving aQ 10 value of 2.1. Skin O2 uptake at 25°C was 38.5% of total. The gas exchange ratio was 0.71 for whole body gas exchange, 0.61 for the lungs and 1.02 for the skin at 25°C.Blood O2 affinity was low with aP 50 of 47.5 mmHg at 25°C and pH 7.65. Then H-value at 25°C increased from 2.7 aroundP 50 to 5.0 at O2 saturations exceeding 70–80%. Surprisingly, blood O2 affinity was nearly insensitive to temperature expressed by a H value of ±1.0 kcal·mole between 25 and 40°C.The adaptive significance of the low O2 affinity, the increase ofn H with O2 saturation and the temperature insensitive O2-Hb binding is discussed in relation to the high and fluctuating body temperatures ofChiromantis.  相似文献   

17.
Influence of soil temperature on methane emission from rice paddy fields   总被引:18,自引:2,他引:16  
Methane emission rates from an Italian rice paddy field showed diel and seasonal variations. The seasonal variations were not closely related to soil temperatures. However, the dieL changes of CH4 fluxes were significantly correlated with the diel changes of the temperature in a particular soil depth. The soil depths with the best correlations between CH4 flux and temperature were shallow (1–5cm) in May and June, deep (10–15cm) in June and July, and again shallow (1–5 cm) in August. Apparent activation energies (Ea) calculated from these correlations using the Arrhenius model were relatively low (50–150 kJ mol–1) in May and June, but increased to higher values (80–450 kJ mol–1) in August. In the laboratory, CH4 emission from two rice cultures incubated at temperatures between 20 and 38°C showed E . values of 41 and 53 kJ mol–1) Methane production in anoxic paddy soil suspensions incubated between 7 and 43°C showed E values between 53 and 132 kJ mol–1 with an average value of 85 kJ mol–1) and in pure cultures of hydrogenotrophic methanogenic bacteria E a values between 77 and 173 (average 126) kJ mol–1. It is suggested that diel changes of soil properties other than temperature affect CH4 emission rates, e.g. diel changes in root exudation or in efficiency of CH4 oxidation in the rhizosphere.  相似文献   

18.
Anaerobic enrichment cultures inoculated with neutral and alkaline (pH 7.0–9.0) sediment and biomat samples from hot-springs in Hveragerdi and Fluir, Iceland, were screened for growth on beech xylan from pH 8.0 to 10.0 at 68° C: no growth occured in cultures above pH 8.4. Five anaerobic xylanolytic bacteria were isolated from enrichment cultures at pH 8.4; all five microbes were Gram-positive rods with terminal spores, and produced CO2, H2, acetate, lactate and ethanol from xylan and xylose. One of the isolates, strain A2, grew from 50 to 75° C, with optimum growth near 68° C, and from pH 5.2 to 9.0 with an optimum between 6.8 and 7.4. Taxonomically, strain A2 was most similar to Clostridium thermohydrosulfuricum. At pH 7.0, the supernatant xylanases of strain A2 had a temperature range from 50 to 78° C with an optimum between 68 and 78° C. At 68° C, xylanase activity occurred from pH 4.9 to 9.1, with an optimum from pH 5.0 to 6.6. At pH 7.0 and 68° C, the K m of the supernatant xylanases was 2.75 g xylan/l and the V max was 2.65 × 10–6 kat/l culture supernatant. When grown on xylose, xylanase production was as high as when grown on xylan. Correspondence to: B. K. Ahring  相似文献   

19.
Summary The following equations represent the influence of the ethanol concentration (E) on the specific growth rate of the yeast cells () and on the specific production rate of ethanol () during the reactor filling phase in fed-batch fermentation of sugar-cane blackstrap molasses: = 0 - k · E and v = v 0 · K/(K +E) Nomenclature E ethanol concentration in the aqueous phase of the fermenting medium (g.L–1) - Em value of E when = 0 or = 0 (g.L–1) - F medium feeding rate (L.h–1) - k empirical constant (L.g–1.h–1) - K empirical constant (g.L–1) - Mas mass of TRS added to the, reactor (g) - Mcs mass of consumed TRS (g) - Me mass of ethanol in the aqueous phase of the fermenting medium (g) - Ms mass of TRS in the aqueous phase of the fermenting medium (g) - Mx mass of yeast cells (dry matter) in the fermenting medium (g) - r correlation coefficient - S TRS concentration in the aqueous phase of the fermenting medium (g.L–1) - Sm TRS concentration of the feeding medium (g.L–1) - t time (h) - T temperature (° C) - TRS total reducing sugars calculated as glucose - V volume of the fermenting medium (L) - V0 volume of the inoculum (L) - X yeast cells concentration (dry matter) in the fermenting medium (g.L–1) - filling-up time (h) - specific growth rate of the yeast cells (h–1) - 0 value of when E=0 - specific production rate of ethanol (h–1) - 0 value of when E=0 - density of the yeast cells (g.L–1) - dry matter content of the yeast cells  相似文献   

20.
Young sporophytes of short-stipe ecotype ofEcklonia cavafrom a warmer locality (Tei, Kochi Pref., southern Japan) and those of long-stipe ecotype from a cooler locality (Nabeta, Shizuoka Pref., central Japan) were transplanted in 1995 to artificial reefs immersed at the habitat of long-stipe ecotype in Nabeta Bay, Shizuoka Pref., central Japan. The characteristics of photosynthesis and respiration of bladelets of the transplanted sporophytes of the two ecotypes were compared in winter and summer 1997; the results were assessed per unit area, per unit chlorophyllacontent and per unit dry weight. In photosynthesis-light curves at 10–29 °C, light saturation occurred at 200–400 mol photon m–2s–1in sporophytes from both Tei and Nabeta. The maximum photosynthetic rate (P max) at 10–29 °C and the light-saturation index (I k) at 25–29 °C in sporophytes from both localities were generally higher in winter than in summer.P maxat 25–29 °C (per unit area and chlorophylla) were higher in sporophytes from Tei than those from Nabeta in both seasons. The optimum temperature for photosynthesis was 25 °C in winter and 27 °C in summer at high light intensities of 100–400 mol photon m–2s–1. However, at lower light intensities of 12.5–50 mol photon m–2s–1, it was 20 °C in winter and 25–27 °C in summer for sporophytes from both locations. Dark respiration increased with temperature rise in the range of 10–29 °C in sporophytes from both locations in summer and winter. The sporophytes transplanted from Tei (warmer area) showed higher photosynthetic activities than those from Nabeta (cooler area) at warmer temperatures even under the same environmental conditions. This indicates that these physiological ecotypes have arisen from genetic differentiation.  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