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 531 毫秒
1.
Fan DD  Luo Y  Mi Y  Ma XX  Shang L 《Biotechnology letters》2005,27(12):865-870
Fed-batch cultures of recombinant Escherichia coli BL21 for producing human-like collagen were performed at different specific growth rates (0.1~0.25 h−1) before induction and at a constant value of 0.05 h−1 after induction by the method of pseudo-exponential feeding. Although the final biomass (around 69 g l−1) was almost the same in all fed-batch cultures, the highest product concentration (13.6 g l−1) was achieved at the specific growth rate of 0.15 h−1 and the lowest (9.6 g l−1) at 0.25 h−1. The mean productivity of human-like collagen was the highest at 0.15 h−1 (0.57 g l−1 h−1) and the lowest at 0.1 h−1 (0.35 g l−1 h−1). In the phase before induction, the cell yield coefficient (YX/S) decreased when the specific growth rate increased, while the formation of acetic acid increased upto 2.5 g l−1 at 0.25 h−1. The mean product yield coefficient (YP/S) also decreased with specific growth rate increasing. The respiration quotient (RQ) increased slightly with specific growth rate increasing before induction, and the mean value of RQ was around 72%. The optimum growth rate for human-like collagen production was 0.15~0.2 h−1.  相似文献   

2.
We measured nitrous oxide (N2O), dinitrogen (N2), methane (CH4), and carbon dioxide (CO2) fluxes in horizontal and vertical flow constructed wetlands (CW) and in a riparian alder stand in southern Estonia using the closed chamber method in the period from October 2001 to November 2003. The replicates’ average values of N2O, N2, CH4 and CO2 fluxes from the riparian gray alder stand varied from −0.4 to 58 μg N2O-N m−2 h−1, 0.02–17.4 mg N2-N m−2 h−1, 0.1–265 μg CH4-C m−2 h−1 and 55–61 mg CO2-C m−2 h−1, respectively. In horizontal subsurface flow (HSSF) beds of CWs, the average N2 emission varied from 0.17 to 130 and from 0.33 to 119 mg N2-N m−2 h−1 in the vertical subsurface flow (VSSF) beds. The average N2O-N emission from the microsites above the inflow pipes of the HSSF CWs was 6.4–31 μg N2O-N m−2 h−1, whereas the outflow microsites emitted 2.4–8 μg N2O-N m−2 h−1. In VSSF beds, the same value was 35.6–44.7 μg N2O-N m−2 h−1. The average CH4 emission from the inflow and outflow microsites in the HSSF CWs differed significantly, ranging from 640 to 9715 and from 30 to 770 μg CH4-C m−2 h−1, respectively. The average CO2 emission was somewhat higher in VSSF beds (140–291 mg CO2-C m−2 h−1) and at the inflow microsites of HSSF beds (61–140 mg CO2-C m−2 h−1). The global warming potential (GWP) from N2O and CH4 was comparatively high in both types of CWs (4.8 ± 9.8 and 6.8 ± 16.2 t CO2 eq ha−1 a−1 in the HSSF CW 6.5 ± 13.0 and 5.3 ± 24.7 t CO2 eq ha−1 a−1 in the hybrid CW, respectively). The GWP of the riparian alder forest from both N2O and CH4 was relatively low (0.4 ± 1.0 and 0.1 ± 0.30 t CO2 eq ha−1 a−1, respectively), whereas the CO2-C flux was remarkable (3.5 ± 3.7 t ha−1 a−1). The global influence of CWs is not significant. Even if all global domestic wastewater were treated by wetlands, their share of the trace gas emission budget would be less than 1%.  相似文献   

3.
In many temperate-zone ecosystems, seasonal changes in environmental and biological factors influence the dynamics and magnitude of surface–atmosphere exchange. Research was conducted between July and October 2001 to measure growing season surface-layer fluxes of CO2 in a Deyeuxia angustifolia dominated wetland on the Sanjiang Plain in northeastern China. Seasonal fluctuation and daily change in soil-surface CO2 fluxes were measured as well as the edaphic factors controlling CO2 fluxes. Soil-surface CO2 fluxes were measured with a closed-chamber system. The results revealed that there were both seasonal fluctuations and daily change in CO2 fluxes. The ranges of measured soil-surface CO2 flux were 0.208 – 1.265 g CO2m–2h–1. Soil-surface CO2 fluxes averaged 0.620 g CO2 m–2h–1. An analysis of several edaphic factors including soil temperature and soil moisture of the D. angustifolia wetland showed that there was a significant relationship between flux and temperature (R2 = 0.77).  相似文献   

4.
A 30-l hollow fibre reactor with continuous fermentation for cell recycling of Escherichia coli AS 1.183 was used to remove the inhibitory effects on cell growth and extend the fast growth phase to increase the yield of polynucleotide phosphorylase (PNPase) in E. coli cells. When the dilution rate was 1.5 h−1, the cell concentration of E. coli reached 235 g/l (wet wt, 70% moisture content), with PNPase activity above 90 u/g (wet wt). With the dilution rate is 1.0 h−1, the fermentor volumetric productivity of PNPase in a hollow fiber reactor can reach 974 (u/h * l) compared to 20 (u/h * l) in a conventional batch culture.  相似文献   

5.
Simčič  Tatjana  Brancelj  Anton 《Hydrobiologia》2001,442(1-3):319-328
Seasonal changes of the community composition, oxygen consumption (R) and respiratory electron transport system (ETS)-activity of the Daphnia community living in Lake Bled (Slovenia) were studied between January and December 1998. The ETS activity of ovigerous Daphnia females at in situ temperature ranged from 3.27 l O2 mg dw–1 h–1 in February to 20.91 l O2 mg dw–1 h–1 in July. Respiration rates at in situ temperature varied from 4.04 l O2 mg dw–1h–1 in December to 18.68 l O2 mg dw–1 h–1in July. The influence of four factors (temperature, body size, fecundity, genetic differences) on the metabolism were investigated. Both ETS activity and respiration rate significantly correlated with temperature. The proportion of hybrid D. cucullata× galeata in Daphnia community correlated significantly with respiration rate at in situtemperature and ETS activity at standard temperature also. ETS activity and respiration rate showed no significant correlation with body size and the fecundity of Daphnia, whereas ETS activity in D. hyalina × galeata and D. cucullata× galeataseparately correlated with body size. ETS activity of D. hyalina × galeata also was correlated with fecundity. Hybrid D. hyalina× galeata had up to one third lower ETS activity than D. cucullata× galeata. The mean ETS/R ratio in the Daphnia community was 1.16±0.28 (N= 12). The ETS/R ratio did not correlate significantly with temperature, body size, fecundity or the proportion of D. cucullata× galeatain the Daphnia community. Laboratory experiments showed that both hybrids had similar ETS/R ratios.  相似文献   

6.
Net ecosystem exchange of CO2 (NEE) was measured during 2005 using the eddy covariance (EC) technique over a reed (Phragmites australis (Cav.) Trin. ex Steud.) wetland in Northeast China (121°54′E, 41°08′N). Diurnal NEE patterns varied markedly among months. Outside the growing season, NEE lacked a diurnal pattern and it fluctuated above zero with an average value of 0.07 mg CO2 m−2 s−1 resulting from soil microbial activity. During the growing season, NEE showed a distinct V-like diel course, and the mean daily NEE was −7.48 ± 2.74 g CO2 m−2 day−1, ranging from −13.58 g CO2 m−2 day−1 (July) to −0.10 g CO2 m−2 day−1 (October). An annual cycle was also apparent, with CO2 uptake increasing rapidly in May, peaking in July, and decreasing from August. Monthly cumulative NEE ranged from −115 ± 24 g C m−2 month−1 (the reed wetland was a CO2 sink) in July to 75 ± 16 g C m−2 month−1 (CO2 source) in November. The annual CO2 balance suggests a net uptake of −65 ± 14 g C m−2 year−1, mainly due to the gains in June and July. Cumulative CO2 emission during the non-growing season was 327 g C m−2, much greater than the absolute value of the annual CO2 balance, which proves the importance of the wintertime CO2 efflux at the study site. The ratio of ecosystem respiration (Reco) to gross primary productivity (GPP) for this reed ecosystem was 0.95, indicating that 95% of plant assimilation was consumed by the reed plant or supported the activities of heterotrophs in the soil. Daytime NEE values during the growing season were closely related to photosynthetically active radiation (PAR) (r2 > 0.63, p < 0.01). Both maximum ecosystem photosynthesis rate (Amax) and apparent quantum yield (α) were season-dependent, and reached their peak values in July (1.28 ± 0.11 mg CO2 m−2 s−1, 0.098 ± 0.027 μmol CO2 μmol−1 photon, respectively), corresponding to the observed maximum NEE in July. Ecosystem respiration (Reco) relied on temperature and soil water content, and the mean value of Q10 was about 2.4 with monthly variation ranging from 1.8 to 4.1 during 2005. Annual methane emission from this reed ecosystem was estimated to be about 3 g C m−2 year−1, and about 5% of the net carbon fixed by the reed wetland was released to the atmosphere as CH4.  相似文献   

7.
Sparse Ulmus pumila woodlands play an important role in contributing to ecosystem function in semi-arid grassland of northern China. To understand the key attributes of soil carbon cycling in U. pumila woodland, we studied dynamics of soil respiration in the canopy field (i.e., the projected crown cover area) and the open field at locations differing in distance (i.e., at 1–1.5, 3–4, 10, and >15 m) to tree stems from July through September of 2005, and measured soil biotic factors (e.g., fine root mass, soil microbial biomass, and activity) and abiotic factors [e.g., soil water content (SWC) and organic carbon] in mid-August. Soil respiration was further separated into root component and microbial component at the end of the field measurement in September. Results showed that soil respiration had a significant exponent relationship with soil temperature at 10-cm depth. The temperature sensitivity index of soil respiration, Q 10, was lower than the global average of 2.0, and declined significantly (P < 0.05) with distance. The rate of soil respiration was generally greater in the canopy field than in the open field; monthly mean of soil respiration was 305.5–730.8 mg CO2 m−2 h−1 in the canopy field and 299.6–443.1 mg CO2 m−2 h−1 in the open field from July through September; basal soil respiration at 10°C declined with distance, and varied from ~250 mg CO2 m−2 h−1 near tree stems to <200 mg CO2 m−2 h−1 in the open field. Variations in soil respiration with distance were consistent with patterns of SWC, fine root mass, microbial biomass and activities. Regression analysis indicated that soil respiration was tightly coupled with microbial respiration and only weakly related to root respiration. Overall, variations in SWC, soil nutrients, microbial biomass, and microbial activity are largely responsible for the spatial heterogeneity of soil respiration in this semi-arid U. pumila woodland.  相似文献   

8.
The present study investigates some aspects of the digestive biology and physiological energetics of the Florida lancelet, Branchiostoma floridae. Florida lancelets are able to remove 47.2–56.9% of the energy from a diet of mixed algae. The respiration rate is 0.100 mL O2 (STPD) h− 1 g− 1 (wet), which estimates a metabolic rate of 0.248 J h− 1, at an average body mass of 0.125 g (wet). Published values of the chlorophyll a concentration in its natural habitat indicate that a 125 mg lancelet would need to filter 0.018–0.031 L h− 1 to remove sufficient food to support its resting metabolism. The filtration rate of lancelets has been reported as 0.138 L h− 1, indicating that the actual filtration rate is 4–7 times greater than the filtration rate needed to meet resting metabolic demands. It appears that lancelets have the potential to be raised in aquaculture, because their absorption efficiency and respiration rate are comparable to suspension-feeding invertebrates that have been successfully aquacultured.  相似文献   

9.
Miniature heat balance-sap flow gauges were used to measure water flows in small-diameter roots (3–4 mm) in the undisturbed soil of a mature beech–oak–spruce mixed stand. By relating sap flow to the surface area of all branch fine roots distal to the gauge, we were able to calculate real time water uptake rates per root surface area (Js) for individual fine root systems of 0.5–1.0 m in length. Study aims were (i) to quantify root water uptake of mature trees under field conditions with respect to average rates, and diurnal and seasonal changes of Js, and (ii) to investigate the relationship between uptake and soil moisture θ, atmospheric saturation deficit D, and radiation I. On most days, water uptake followed the diurnal course of D with a mid-day peak and low night flow. Neighbouring roots of the same species differed up to 10-fold in their daily totals of Js (<100–2000 g m−2 d−1) indicating a large spatial heterogeneity in uptake. Beech, oak and spruce roots revealed different seasonal patterns of water uptake although they were extracting water from the same soil volume. Multiple regression analyses on the influence of D, I and θ on root water uptake showed that D was the single most influential environmental factor in beech and oak (variable selection in 77% and 79% of the investigated roots), whereas D was less important in spruce roots (50% variable selection). A comparison of root water uptake with synchronous leaf transpiration (porometer data) indicated that average water fluxes per surface area in the beech and oak trees were about 2.5 and 5.5 times smaller on the uptake side (roots) than on the loss side (leaves) given that all branch roots <2 mm were equally participating in uptake. Beech fine roots showed maximal uptake rates on mid-summer days in the range of 48–205 g m−2 h−1 (i.e. 0.7–3.2 mmol m−2 s−1), oak of 12–160 g m−2 h−1 (0.2–2.5 mmol m−2 s−1). Maximal transpiration rates ranged from 3 to 5 and from 5 to 6 mmol m−2 s−1 for sun canopy leaves of beech and oak, respectively. We conclude that instantaneous rates of root water uptake in beech, oak and spruce trees are above all controlled by atmospheric factors. The effects of different root conductivities, soil moisture, and soil hydraulic properties become increasingly important if time spans longer than a week are considered.  相似文献   

10.
Synechocystis aquatilis SI-2 was grown outdoors in a 12.5cm diam. tubular photobioreactor equipped with static mixers. The static mixers ensured that cells were efficiently circulated between the upper (illuminated) and lower (dark) sections of the tubes. The biomass productivity varied from 22 to 45g m–2d–1, with an average of 35g m–2d–1, etc which corresponded to average CO2 fixation rate of about 57 g CO2 m–2 d–1. The static mixers not only helped in improving the biomass productivities but also have a high potential to lower the photoinhibitory effect of light during the outdoor cultures of algae. Revisions requested 27 July 2004; Revisions received 12 November 2004  相似文献   

11.
CO2 efflux from soil and snow surfaces was measured continuously in a Japanese cedar (Cryptomeria japonica D. Don) forest in central Japan using an open dynamic chamber system. The chamber opens and closes automatically and records measurements based on an open-flow dynamic method. Between May and December, mean soil CO2 efflux ranged from 1,529 mg CO2 m−2 h−1 in September to 255 mg CO2 m−2 h−1 in December. The seasonal change in CO2 efflux from the soil paralleled the seasonal pattern of soil temperature. No marked diurnal trends in soil CO2 efflux were observed on days without rainfall, whereas significant pulses in soil CO2 efflux were observed on days with rainfall. In this plantation, soil CO2 efflux frequently responded to rainfall. Measurements of changes from litter-covered soil to snow-covered surfaces revealed that CO2 efflux decreased from values of ca. 250 mg CO2 m−2 h−1 above soil to less than 33 mg CO2 m−2 h−1 above snow. Soil temperature alone explained 66% of the overall variation in soil CO2 efflux, but explained approximately 85% of the variation when data from two anomalous periods were excluded. Moreover, we found a significant correlation between soil CO2 efflux and soil moisture (which explained 44% of the overall variation) using a second-order polynomial function. Our results suggest that the seasonality of CO2 efflux is affected not only by soil temperature and moisture, but also by drying and rewetting cycles and by litterfall pulses.  相似文献   

12.
Chromosome set manipulation was used to produce rainbow trout, Oncorhynchus mykiss, with identical nuclear backgrounds, but different maternal backgrounds to determine mitochondrial effects on development rate and oxygen consumption. Significant differences in development rate and oxygen consumption were observed between groups from different females. Development rates ranged from a mean of 317.97 degree days (°d) to 335.25 °d in progeny from different females. Mean oxygen consumption rates ranged from 3.31 μmol O2 g− 1 wet mass h− 1 to 9.66 μmol O2 g− 1 wet mass h− 1. Oxygen consumption and development rate analysis revealed the two slowest developing groups had the highest oxygen consumption rates. Development rate differences between second generation clonal females indicate that mitochondrial genomes play a significant role on early development and are comparable to development rate differences between clonal lines of rainbow trout. These results indicate that selection for mitochondrial genomes could increase growth rates and possibly food conversion ratios in aquaculture species.  相似文献   

13.
A family of 10 competing, unstructured models has been developed to model cell growth, substrate consumption, and product formation of the pyruvate producing strain Escherichia coli YYC202 ldhA::Kan strain used in fed-batch processes. The strain is completely blocked in its ability to convert pyruvate into acetyl-CoA or acetate (using glucose as the carbon source) resulting in an acetate auxotrophy during growth in glucose minimal medium. Parameter estimation was carried out using data from fed-batch fermentation performed at constant glucose feed rates of qVG=10 mL h–1. Acetate was fed according to the previously developed feeding strategy. While the model identification was realized by least-square fit, the model discrimination was based on the model selection criterion (MSC). The validation of model parameters was performed applying data from two different fed-batch experiments with glucose feed rate qVG=20 and 30 mL h–1, respectively. Consequently, the most suitable model was identified that reflected the pyruvate and biomass curves adequately by considering a pyruvate inhibited growth (Jerusalimsky approach) and pyruvate inhibited product formation (described by modified Luedeking–Piret/Levenspiel term).List of symbols cA acetate concentration (g L–1) - cA,0 acetate concentration in the feed (g L–1) - cG glucose concentration (g L–1) - cG,0 glucose concentration in the feed (g L–1) - cP pyruvate concentration (g L–1) - cP,max critical pyruvate concentration above which reaction cannot proceed (g L–1) - cX biomass concentration (g L–1) - KI inhibition constant for pyruvate production (g L–1) - KIA inhibition constant for biomass growth on acetate (g L–1) - KP saturation constant for pyruvate production (g L–1) - KP inhibition constant of Jerusalimsky (g L–1) - KSA Monod growth constant for acetate (g L–1) - KSG Monod growth constant for glucose (g L–1) - mA maintenance coefficient for growth on acetate (g g–1 h–1) - mG maintenance coefficient for growth on glucose (g g–1 h–1) - n constant of extended Monod kinetics (Levenspiel) (–) - qV volumetric flow rate (L h–1) - qVA volumetric flow rate of acetate (L h–1) - qVG volumetric flow rate of glucose (L h–1) - rA specific rate of acetate consumption (g g–1 h–1) - rG specific rate of glucose consumption (g g–1 h–1) - rP specific rate of pyruvate production (g g–1 h–1) - rP,max maximum specific rate of pyruvate production (g g–1 h–1) - t time (h) - V reaction (broth) volume (L) - YP/G yield coefficient pyruvate from glucose (g g–1) - YX/A yield coefficient biomass from acetate (g g–1) - YX/A,max maximum yield coefficient biomass from acetate (g g–1) - YX/G yield coefficient biomass from glucose (g g–1) - YX/G,max maximum yield coefficient biomass from glucose (g g–1) - growth associated product formation coefficient (g g–1) - non-growth associated product formation coefficient (g g–1 h–1) - specific growth rate (h–1) - max maximum specific growth rate (h–1)  相似文献   

14.
Anabaena siamensis isolated from rice fields in Thailand is a fast growing cyanobacterium with a high nitrogen-fixing activity. Mutant strains resistant to the l-glutamate analogue, l-methionine sulfoximine (MSX) were isolated by ethyl methanesulfonate mutagenesis. A stable mutant named A. siamensis SS1, which released ammonium to the medium, was studied further. In batch cultures the rate of ammonium production peaked at the early log phase and gradually decreased until the 4th day of growth when the cultures reached a density of 90 μg chl ml−1. To obtain constant release of ammonium by SS1, continuous culture experiments were performed at a cell density of 5 μg chl ml−1 and the following results were obtained: (1) growth rate as the parent (μ:0·123 h−1) in the presence and absence of 500 μm MSX; (2) 48% GS transferase activity when compared with the parent; (3) ammonium excretion at a rate of 8 μmol (mg chl)−1 h−1 as measured up to 20 generations (120 h); (4) depressed nitrogenase activity; and (5) 30% higher nitrogenase activity than that of the parent. SS1 immobilized in alginate beads (5 μg chl ml−1) exhibited values of glutamine synthetase and nitrogenase activity similar to those of free cells. However, ammonium excretion at the rate of 11·61 μmol (mg chl)−1 h−1 was obtained only up to 20 h after loading in bioreactors, due to the fast growth of SS1 as also occurred in batch cultures.  相似文献   

15.
(R)-Phenylacetylcarbinol (PAC), a pharmaceutical precursor, was produced from benzaldehyde and pyruvate by pyruvate decarboxylase (PDC) of Candida utilis in an aqueous/organic two-phase emulsion reactor. When the partially purified enzyme in this previously established in vitro process was replaced with C. utilis cells and the temperature was increased from 4 to 21 °C, a screen of several 1-alcohols (C4–C9) confirmed the suitability of 1-octanol as the organic phase. Benzyl alcohol, the major by-product in the commercial in vivo conversion of benzaldehyde and sugar to PAC by Saccharomyces cerevisiae, was not formed. With a phase volume ratio of 1:1 and 5.6 g C. utilis l−1 (PDC activity 2.5 U ml−1), PAC levels of 103 g l−1 in the octanol phase and 12.8 g l−1 in the aqueous phase were produced in 15 h at 21 °C. In comparison to our previously published process with partially purified PDC in an aqueous/octanol emulsion at 4 °C, PAC was produced at a 4-times increased specific rate (1.54 versus 0.39 mg U−1 h−1) with simplified catalyst production and reduced cooling cost. Compared to traditional in vivo whole cell PAC production, the yield on benzaldehyde was 26% higher, the product concentration increased 3.9-fold (or 6.9-fold based on the organic phase), the productivity improved 3.1-fold (3.9 g l−1 h−1) and the catalyst was 6.9-fold more efficient (PAC/dry cell mass 10.3 g g−1).*Dedicated with gratitude to Prof. Dr. Franz Lingens – “Theo”.  相似文献   

16.
To narrow the differences between the results obtained from radionuclides and heavy metal ecotoxicity investigations in the laboratory and in the abandoned uranium mines, a few standardised plant bioassay procedures were selected from the literature for testing with Lemna gibba L. The bioassay procedures were tested in situ and ex situ. The laboratory culturing was performed in batch and semicontinuous modes. The results revealed that most of the standardised plant bioassay procedures require modification for the L. gibba bioassay to predict the actual effects under field conditions. L. gibba performed relatively better in the field than laboratory batch cultures despite that the batch cultures had many-fold higher nutrient concentrations than in the field. For instance, the phosphorus concentration of the mine tailing water was 0.13 ± 0.09 μg l−1 in the field, while the literature range for phosphorus in the laboratory culture media is 13.6–40 mg l−1. L. gibba growth in the laboratory batch culture was influenced by speciation changes due to consumption of nutrients, CO2 and O2 phase exchanges, and excretion of organic substances by the test plants. Semicontinuous culture modes performed significantly better than batch cultivation even after 10× dilution of the nutrient solution. The growth behaviour revealed that L. gibba exhibited intrapopulation and probiotic interaction for best performance. Growth performance of L. gibba was influenced by the anions that balanced essential cations despite equal cation concentration in the culture media; e.g., the best growth was observed in culture media that had more SO42− than Cl. Water samples from the field had higher SO42− concentrations than Cl. The test vessel material, sterilisation and axenic culturing procedures also influenced the sensitivity of the bioassay. These, for instance, and a few others are neither described nor reported in most standard Lemna tests or the literature. Thus, this work presents results of a series of tests conducted on the selected methods. Common and possible errors and corrective measures in assigning L. gibba bioassay from laboratory population levels to field community levels are discussed.  相似文献   

17.
Two thin layer culture units operated as batch cultures with the algaChlorella kessleri were used in gas exchange experiments. The mass transfer coefficient Kg [g m–2 h–1 kPa–1] of O2 and CO2 desorption from culture surface decreased with increasing culture temperature. Between 60–70% of supplied CO2 was used for algal growth. It was estimated that the length of growth surface may be extended to about 50 m, without additional saturation by CO2. On average 1.35 g CO2 was consumed by the alga per 1 g of produced O2. Net CO2 consumption (RCO2) and O2 production (RO2) were not inhibited by irradiance. RO2 did not decrease (in some cases it even increased) along the culture surface, despite increased accumulation of O2. Measurement of pO2 where the culture leaves the reactor before being pumped back onto the illuminated surface, correlated with O2 production and CO2 consumption and may be used to monitor the reactors growth performance.  相似文献   

18.
Agrobacterium tumefaciens-mediated transformation system for perilla (Perilla frutescens Britt) was developed. Agrobacterium strain EHA105 harboring binary vector pBK I containing bar and γ-tmt cassettes or pIG121Hm containing nptII, hpt, and gusA cassettes were used for transformation. Three different types of explant, hypocotyl, cotyledon and leaf, were evaluated for transformation and hypocotyl explants resulted in the highest transformation efficiency with an average of 3.1 and 2.2%, with pBK I and pIG121Hm, respectively. The Perilla spp. displayed genotype-response for transformation. The effective concentrations of selective agents were 2 mg l−1 phosphinothricin (PPT) and 150 mg l−1 kanamycin, respectively, for shoot induction and 1 mg l−1 PPT and 125 mg l−1 kanamycin, respectively, for shoot elongation. The transformation events were confirmed by herbicide Basta spray or histochemical GUS staining of T0 and T1 plants. The T-DNA integration and transgene inheritance were confirmed by PCR and Southern blot analysis of random samples of T0 and T1 transgenic plants.  相似文献   

19.
Summary From acetylene reduction assays over a 10-month period starting in April 1979, nodule activities averaged 18.78 (se 4.67) moles C2H4 g nodule dw–1 h–1 forAlnus rubra and 59.95 (se 12.14) moles C2H4 g nodule dw–1 h–1 forCytisus scorparius. Plant rates were 1.91 (se. 47) moles C2H4 plant–1 h–1 forA. rubra and 0.55 (se. 17) moles C2H4 plant–1 h–1 forC. Scoparius. Plant activity and total leaf N were strongly correlated with the dw of other plant parts, but nodule activity and percent leaf N were not. Plant and nodule activities were not associated with temperature, moisture stress, precipitation events or percent light for either species over the growing season nor for 54A. rubra sampled in mid-season 1979 on one replication. After 5 to 6 growing seasons, 14A. rubra on the same site ranged from 30 to 332 cm in height and showed strong correlation between nodule dw, leaf dw, plant size and total leaf N. Results from this study and others indicate logistic equations may be modified to predict the effect of adding a N2 fixing plant to a population of non N2 fixing trees.  相似文献   

20.
Net daily budgets of dissolved oxygen (O2), dissolved inorganic carbon (DIC), dissolved inorganic nitrogen (DIN = NH4++NO2+NO3) and soluble reactive phosphorus (SRP) were determined in a pond colonised by Ulva spp. This pond received wastewater from a land-based fish farm and was used as a phytotreatment plant. Three consecutive 24-h cycles of measurements were performed with 8–14 samplings per day. Water samples were collected at the inlet and outlet of the pond and budgets were estimated from differences between inlet and outlet loadings. The first cycle was started when Ulva biomass was 8 kg m−2, as wet weight. The second cycle was performed after the harvest of ~20% of the macroalgal biomass and the third after the harvest of another ~20% of the remaining biomass. Ulva removal was very fast (<1 h) and samplings for cycles 2 and 3 were started two hours after harvesting, so that the whole experiment lasted ~80 h. When Ulva biomass was at its maximum, the aquatic system was heterotrophic with an O2 demand of 519 mol d−1 and a net regeneration of DIC (2686 mol d−1), NH4+ (49 mol d−1) and SRP (2.5 mol d−1). The DIC to O2 ratio was an indicator of persistent anaerobic metabolism. Following the first harvest intervention, this system displayed a prompt response and shifted toward a lower O2 demand (from −519 to −13 mol d−1), with a lesser regeneration degree of NH4+ (11.4 mol d−1) and DIC (1066 mol d−1). After the second Ulva removal the net budget of SRP became negative (−1.0 mol d−1). By integrating these results over the three days cycle we estimated that in order to operate an efficient nutrient control and maintain macroalgal mats in a healthy status the optimal Ulva biomass should be well below ~4 kg m−2 as wet weight. Above this threshold, self-limitation would render most of the algal mat unable to exploit light and nutrients. An efficient removal of nitrogen and phosphorus could be attained through the management of macroalgal biomass only with an optimisation of recipient surface to nutrient loading ratio.  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