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.
Low concentrations of urea and GuHCl (2 M) enhanced the activity of endoglucanase (EC 3.1.2.4) from Aspergillus aculeatus by 2.3- and 1.9-fold, respectively. The Km values for controls, in the presence of 2 M urea and GuHCl, were found to be 2.4 ± 0.2 × 10−8 mol L−1, 1.4 ± 0.2 × 10−8 mol L−1, and 1.6 ± 0.2 × 10−8 mol L−1, respectively. The dissociation constant (Kd) showed changes in the affinity of the enzyme for the substrate with increases in the Kcat suggesting an increased turnover number in the presence of urea and GuHCl. Fluorescence studies showed changes in the microenvironment of the protein. The increase in the activity of this intermediate state was due to conformational changes accompanied by increased flexibility at the active site.  相似文献   

2.
Our study aimed to test the ability of aquatic plants to use bicarbonate when acclimated to three different bicarbonate concentrations. To this end, we performed experiments with the three species Ceratophyllum demersum, Egeria densa, Lagarosiphon major to determine photosynthetic rates under varying bicarbonate concentrations. We measured bicarbonate use efficiency, photosynthetic performance and respiration. For all species, our results revealed that photosynthetic rates were highest in replicates grown at low alkalinity. Thus, E. densa had approx. five times higher rates at low (264 ± 15 μmol O2 g−1 DW h−1) than at high alkalinity (50 ± 27 μmol O2 g−1 DW h−1), C. demersum had three times higher rates (336 ± 95 and 120 ± 31 μmol O2 g−1 DW h−1), and L. major doubled its rates at low alkalinity (634 ± 114 and 322 ± 119 μmol O2 g−1 DW h−1). Similar results were obtained for bicarbonate use efficiency by E. densa (136 ± 44 and 43 ± 10 μmol O2 mequiv. L−1 g−1 DW h−1) and L. major (244 ± 29 and 82 ± 24 μmol O2 mequiv. L−1 g−1 DW h−1). As to C. demersum, efficiency was high but unaffected by alkalinity, indicating high adaptation ability to varied alkalinities. A pH drift experiment supported these results. Overall, our results suggest that the three globally widespread worldwide species of our study adapt to low inorganic carbon availability by increasing their efficiency of bicarbonate use.  相似文献   

3.
The uptake kinetics of phosphate (Pi) by Myriophyllum spicatum was determined from adsorption and absorption under light and dark conditions. Pi uptake was light dependent and showed saturation following the Michaelis-Menten relation (in light: V = 16.91 × [Pi](1.335 + [Pi]), R2 = 0.90, p < 0.001; in the dark: V = 5.13 × [Pi](0.351 + [Pi]), R2 = 0.77, p < 0.001). Around 77% of the loss of Pi in the water column was absorbed into the tissue of M. spicatum, and only 23% was adsorbed on the surface of the plant shoots. Our study shows that M. spicatum shoots have a much higher affinity (in light: 3.9 μmol g−1 dw h−1 μM−1; in the dark: 3.7 μmol g−1 dw h−1 μM−1) and Vmax (maximum uptake rate, shoot light) for Pi uptake than many other aquatic macrophytes (in light: 0.002-0.23 μmol g−1 dw h−1 μM−1; in the dark: 0.002-0.19 μmol g−1 dw h−1 μM−1), which may provide a competitive advantage over other macrophytes across a wide range of Pi concentrations.  相似文献   

4.
The microalgae, Chlorella sp., were cultivated in various culture modes to assess biomass and lipid productivity in this study. In the batch mode, the biomass concentrations and lipid content of Chlorella sp. cultivated in a medium containing 0.025–0.200 g L−1 urea were 0.464–2.027 g L−1 and 0.661–0.326 g g−1, respectively. The maximum lipid productivity of 0.124 g d−1 L−1 occurred in a medium containing 0.100 g L−1 urea. In the fed-batch cultivation, the highest lipid content was obtained by feeding 0.025 g L−1 of urea during the stationary phase, but the lipid productivity was not significantly increased. However, a semi-continuous process was carried out by harvesting the culture and renewing urea at 0.025 g L−1 each time when the cultivation achieved the early stationary phase. The maximum lipid productivity of 0.139 g d−1 L−1 in the semi-continuous culture was highest in comparison with those in the batch and fed-batch cultivations.  相似文献   

5.
A novel nanocomposite material of multiwalled carbon nanotubes (MWCNTs) and room temperature ionic liquid (RTIL) N-butylpyridinium hexafluorophosphate (BPPF6) was explored and used to construct a novel microperoxidase-11 (MP-11) biosensor for the determination of hydrogen peroxide (H2O2). Cyclic voltammetry (CV) and differential pulse voltammetry (DPV) were used to characterize the performance of the biosensor. Under the optimized experimental conditions, H2O2 could be detected in a linear calibration range of 0.5 to 7.0 × 10−7 mol L−1 with a correlation coefficient of 0.9949 (n = 9) and a detection limit of 3.8 × 10−9 mol L−1 at 3σ. The modified electrodes displayed excellent electrochemical response, high sensitivity, long-term stability, and good bioactivity and selectivity.  相似文献   

6.
Previous work demonstrated that a mixture of NH4Cl and KNO3 as nitrogen source was beneficial to fed-batch Arthrospira (Spirulina) platensis cultivation, in terms of either lower costs or higher cell concentration. On the basis of those results, this study focused on the use of a cheaper nitrogen source mixture, namely (NH4)2SO4 plus NaNO3, varying the ammonium feeding time (T = 7-15 days), either controlling the pH by CO2 addition or not. A. platensis was cultivated in mini-tanks at 30 °C, 156 μmol photons m−2 s−1, and starting cell concentration of 400 mg L−1, on a modified Schlösser medium. T = 13 days under pH control were selected as optimum conditions, ensuring the best results in terms of biomass production (maximum cell concentration of 2911 mg L−1, cell productivity of 179 mg L−1 d−1 and specific growth rate of 0.77 d−1) and satisfactory protein and lipid contents (around 30% each).  相似文献   

7.
Calcification and primary production responses to irradiance in the temperate coralline alga Lithothamnion corallioides were measured in summer 2004 and winter 2005 in the Bay of Brest. Coralline algae were incubated in dark and clear bottles exposed to different irradiances. Net primary production reached 1.5 μmol C g−1 dry wt h−1 in August and was twice as high as in January–February. Dark respiration showed significant seasonal variations, being three-fold higher in summer. Maximum calcification varied from 0.6 μmol g−1 dry wt h−1 in summer 2004 to 0.4 μmol g−1 dry wt h−1 in winter 2005. According to PE curves and the daily course of irradiance, estimated daily net production and calcification reached 131 μg C g−1 dry wt and 970 μg CaCO3 g−1 dry wt in summer 2004, and 36 μg C g−1 dry wt and 336 μg CaCO3 g−1 dry wt in winter 2005. The net primary production of natural L. corallioides populations in shallow waters was estimated at 10–600 g C m−2 y−1, depending on depth and algal biomass. The mean annual calcification of L. corallioides populations varied from 300 to 3000 g CaCO3 m−2. These results are similar to those reported for tropical coralline algae in terms of carbon and carbonate productivity. Therefore, L. corallioides can be considered as a key element of carbon and carbonate cycles in the shallow coastal waters where they live.  相似文献   

8.
Gametophyte strains originating from indigenous sporophytes of Undaria pinnatifida (Harvey) Suringar in Iwate Prefecture, Northeast Japan, were maintained for 9–10 months at 45 μmol photons m−2 s−1. Before cryopreservation in liquid nitrogen for more than 12 h (1–14 days) using a two-step cooling method with a mixture of cryoprotectants (10% l-proline and 10% glycerol), these were pre-incubated for 2, 4 and 8 months at 15 μmol photons m−2 s−1. After 1 week of thawing, no surviving gametophytes were detected in the strains without pre-incubation, but both the female and male gametophytes, pre-incubated for more than 4 months, showed high survival rates (43–60% for females and 64–100% for males). This revealed the induction of freezing tolerance by incubation at low irradiance. Thereafter, sporophytes derived from cryopreserved gametophytes and subcultured gametophytes, stored under pre-incubation conditions, were formed from the strain, and a morphological comparison was conducted with 10 characters (stipe length, stipe wet weight, blade length, blade wet weight, blade width, incision depth, blade thickness, sporophyll length, sporophyll wet weight, and sporophyll width). The morphology of the sporophytes formed from the cryopreserved gametophytes corresponded well with that of the subcultured gametophytes from the same strain. The results suggest that the cryopreservation method is applicable for preserving culture stocks of U. pinnatifida to be used in mariculture.  相似文献   

9.
The acid biocoagulants produced from non-sterile lactic acid fermentation by Lactobacillus casei TISTR 1500 were used to settle colloidal protein, mainly casein, at the isoelectric point in dairy effluent prior to secondary treatment. High concentration of azo dye (Ponceau 4R) in the dairy wastewater and the stress of starvation decreased the efficiencies of the micro-aerobic SBR. Consequently, low casein recovery obtained and organic removal suffered a decline. The number of lactic acid bacteria (LAB) also declined from log 7.4 to log 5.30 in the system fed with 400 mg L−1 of the dye containing wastewater. The recovery of the system, however, showed that 25,000 mg COD L−1 influent with 200 mg L−1 of the dye maintained the growth of LAB in the range of log 7.74–8.12, with lactic and acetic production (2597 and 197 mg L−1) and 83% protein removal. The results in this study suggested that the inhibitory effects were compensated with high organic content feeding.  相似文献   

10.
The folding mechanism and stability of dimeric formate dehydrogenase from Candida methylica was analysed by exposure to denaturing agents and to heat. Equilibrium denaturation data yielded a dissociation constant of about 10−13 M for assembly of the protein from unfolded chains and the kinetics of refolding and unfolding revealed that the overall process comprises two steps. In the first step a marginally stable folded monomeric state is formed at a rate (k1) of about 2 × 10−3 s−1 (by deduction k−1 is about10−4 s−1) and assembles into the active dimeric state with a bimolecular rate constant (k2) of about 2 × 104 M−1 s−1. The rate of dissociation of the dimeric state in physiological conditions is extremely slow (k−2 ∼ 3 × 10−7 s−1).  相似文献   

11.
We tested the effects of UV radiation (UVR) and nitrate limitation on the production of dimethylsulfide (DMS), particulate dimethylsulfoniopropionate (DMSPp), and particulate dimethylsulfoxide (DMSOp) in natural seawater from the Gulf of Mexico and in phytoplankton cultures. DMS/Chl a ratios in PAR-only and PAR + UV-exposed seawater were 0.44–2.0 and 0.46–1.9 nmol DMS μg−1 Chl a, respectively, whereas the ratios in cultures of Amphidinium carterae were 1.0–2.2 nmol μg−1 in PAR-exposed samples and 0.91–2.1 nmol μg−1 in PAR + UV-exposed samples. These results suggested that UVR did not substantially affect DMS/Chl a ratios in seawater and A. carterae culture samples. Similarly, UVR had no significant effect on DMSOp/Chl a in seawater samples (0.83–1.6 nmol DMSO μg−1 Chl a for PAR + UV vs. 0.70–1.5 nmol μg−1 for PAR-exposed seawater samples, respectively) or in A. carterae cultures (0.20–1.3 and 0.19–0.88 nmol DMSO μg−1 Chl a in PAR + UV- and PAR-exposed cultures, respectively). In an experiment with the diatom, Thalassiosira oceanica, the culture was grown in high nitrate (30 μM) or low nitrate (6 μM) media and exposed to PAR-only or PAR + UV. The low nitrate, PAR-only samples showed an increase of intracellular dimethylsulfoniopropionate (DMSP) concentration from 2.1 to 15 mmol L−1 in 60 h, but the increase occurred only after cultures reached the stationary phase. Cultures of T. oceanica grown under UVR had lower growth rates than those under PAR-only (μ′ = 0.17 and 0.32 d−1, respectively) and perhaps did not experience severe nitrate limitation even in the low nitrate treatment. These results suggest that the elevated UVR in low nitrate environments could result in reduction of DMSP in some species, whereas DMSP concentrations would not be affected in eutrophic areas.  相似文献   

12.
The Iberian Peninsula encompasses more than 80% of the species richness of European aquatic ranunculi. The floristic diversity of the phytocoenosis characterised by aquatic Ranunculus and the main physical–chemical factors of the water were studied in 43 localities of the central Iberian Peninsula. Four aquatic Ranunculus communities are found in most of the aquatic environments. These are species-poor and have an uneven distribution: three species of Batrachium are heterophyllous and their communities are distributed in different aquatic ecosystems on silicated substrates; one species is homophyllous and its community occurs in various aquatic ecosystems with carbonated waters. In the Mediterranean climate, Ranunculus species are present in different habitats, as shown by the results of all the statistical analyses. Ranunculus trichophyllus communities occur in base-rich waters with a high buffering capacity (2273.44 ± 794.57 mg CaCO3 L−1) and a high concentration of cations (Ca2+, 121 ± 33.12 mg L−1; Mg2+, 71.64 ± 82.77 mg L−1), nitrates (2.89 ± 4.80 mg L−1), ammonium (2.19 ± 1.36 mg L−1) and sulphates (216.25 ± 218.54 mg L−1). Ranunculus penicillatus communities are found in flowing waters with a high concentration of phosphates (0.48 ± 0.6 mg L−1) and intermediate buffering capacity (683.66 ± 446.76 mg CaCO3 L−1). Both Ranunculus pseudofluitans and Ranunculus peltatus communities grow in waters with low buffering capacity (R. pseudofluitans, 385.91 ± 209.2 mg CaCO3 L−1; R. peltatus, 263.3 ± 180.36 mg CaCO3 L−1), and a low concentration of cations (R. pseudofluitans: Ca2+, 12.57 ± 9.42 mg L−1; Mg2+, 3.42 ± 1.67 mg L−1; R. peltatus: Ca2+, 15 ± 18.26 mg L−1; Mg2+, 6.26 ± 8.89 mg L−1) and nutrients (R. pseudofluitans: nitrates, 0.23 ± 0.2 mg L−1; phosphates, 0.09 ± 0.1 mg L−1; R. peltatus: nitrates, 0.19 ± 0.21 mg L−1; phosphates, 0.09 ± 0.12 mg L−1); the first in flowing waters, the latter in still waters.  相似文献   

13.
Two extracellular chitinases (designated as Chi-56 and Chi-64) produced by Massilia timonae were purified by ion-exchange chromatography, ammonium sulfate precipitation, and gel-filtration chromatography. The molecular mass of Chi-56 was 56 kDa as determined by both SDS-PAGE and gel-filtration chromatography. On the other hand, Chi-64 showed a molecular mass of 64 kDa by SDS-PAGE and 28 kDa by gel-filtration chromatography suggesting that its properties may be different from those of Chi-56. The optimum temperature, optimum pH, pI, Km, and Vmax of Chi-56 were 55 °C, pH 5.0, pH 8.5, 1.1 mg mL−1, and 0.59 μmol μg−1 h−1, respectively. For Chi-64, these values were 60 °C, pH 5.0, pH 8.5, 1.3 mg mL−1, and 1.36 μmol μg−1 h−1, respectively. Both enzymes were stimulated by Mn2+ and inhibited by Hg2+, and neither showed exochitinase activity. The N-terminal sequences of Chi-56 and Chi-64 were determined to be Q-T-P-T-Y-T-A-T-L and Q-A-D-F-P-A-P-A-E, respectively.  相似文献   

14.
Gamete production after exposure to hypoxia or sulphide was studied in the marine macroalga Ulva sp. collected in the Sacca di Goro, Italy. Experiments were carried out on discs (12 mm diameter) of thalli cultured in artificial sea water in laboratory at 20 ± 1 °C, 152 μmol m−2 s−1, 16 h photoperiod and 30‰ salinity. Dehydration of thallus was used as inducer of gametogenesis and growth and gamete release during recovery after 10, 20, 30 or 40 min dehydration (20 ± 1 °C, 25% humidity) were analysed. Unlike non-dehydrated thalli the dehydrated ones produced gametes. Thallus discs, non-dehydrated or subjected to 30 min dehydration, were exposed to hypoxia (1.78–4.02 μmol O2 L−1) or sulphide (1 mM) for 3, 5, or 7 days at 20 °C in the dark. Non-dehydrated and dehydrated thalli maintained in normoxic conditions in the dark were the controls. Gamete density was checked by counting at the end of the incubation period and during the subsequent 7 days of recovery under 16 h photoperiod in normoxic conditions. Non-dehydrated thalli maintained in normoxic conditions in the dark released gametes when returned to light suggesting that dark constitutes a stimulus to gamete production. The presence of gametes at the end of 3 days incubation of dehydrated thalli in normoxia demonstrated that gametogenesis can occur even in the dark. However, gametes were not present at the end of incubation in hypoxic and sulphidic conditions. Actually, during hypoxic incubation oxygen consumption in D-thalli was very low, only 0.117 × 10−3 μmol O2 mg−1 h−1 compared to 5.93 × 10−3 μmol O2 mg−1 h−1 in normoxia, denoting a reduction of the metabolic rate that could not sustain gametogenesis. During recovery after incubation in normoxic, hypoxic or sulphidic conditions densities of gametes from dehydrated thalli showed significant differences and resulted after hypoxia > after normoxia > after sulphide. Differences in non-dehydrated thalli were not significant. Dehydrated thalli, still green at the end of the incubation period, underwent blanching in the course of recovery in parallel to gamete production, while non-dehydrated thalli maintained their green colour even after exposure to sulphide. Our findings suggest that macroalga Ulva sp. can survive exposure to darkness, severe hypoxia and high sulphide levels and can maintain gamete production even when the exposure to these stress conditions is joined to dehydration.  相似文献   

15.
The sydnone SYD-1 (3-[4-chloro-3-nitrophenyl]-1,2,3-oxadiazolium-5-olate] possesses important antitumor activity against Sarcoma 180 and Ehrlich tumors. We previously showed that SYD-1 depresses mitochondrial phosphorylation efficiency, which could be involved in its antitumoral activity. Considering the important role of mitochondria in the generation of reactive oxygen species (ROS) and the involvement of ROS in cell death mechanisms, we evaluated the effects of SYD-1 on oxidative stress parameters in rat liver mitochondria. SYD-1 (0.5 and 0.75 μmol mg−1 protein) inhibited the lipoperoxidation induced by Fe3+/ADP-oxoglutarate by approximately 75% and promoted total inhibition at the highest concentration tested (1.0 μmol mg−1 protein). However, SYD-1 did not affect lipoperoxidation started by peroxyl radicals generated by α-α′-azodiisobutyramidine dihydrochloride. The mesoionic compound (0.25–1.0 μmol mg−1 protein) demonstrated an ability to scavenge superoxide radicals, decreasing their levels by 9–19%. The activities of catalase and superoxide dismutase did not change in the presence of SYD-1 (0.25–1.0 μmol mg−1 protein). SYD-1 inhibited mitochondrial swelling dependent on the formation/opening of the permeability transition pore induced by Ca2+/phosphate by approximately 30% (1.0 μmol mg−1 protein). When Ca2+/H2O2 were used as inducers, SYD-1 inhibited swelling only by approximately 12% at the same concentration. NADPH oxidation was also inhibited by SYD-1 (1.0 μmol mg−1 of protein) by approximately 48%. These results show that SYD-1 is able to prevent oxidative stress in isolated mitochondria and suggest that the antitumoral activity of SYD-1 is not mediated by the increasing generation of ROS.  相似文献   

16.
Studies on Caulerpa prolifera rhizoids were carried out to determine the possibility of mass culture, because the rhizoids produce a bio-adhesive. Rhizoids can be induced by cutting the base of a blade and floating it in a media or planting it in sand. Measurement of rhizoid production included determination of number, length, and the weight of attached sand grains. The growth experiments were for 1–2 weeks and fronds growth was compared to rhizoid production. Optimal conditions for rhizoid growth included low levels of nitrogen and phosphate (less than 5 and 2 μM, respectively), low irradiance (30 μmol photon m−2 s−1), moderate temperature (22–28°), continuous shaking, addition of microelements and auxin (1 ppm) and initially detached fronds followed by attachment. Under these optimal conditions maximal weekly growth reached 70–170 rhizoids per blade, 7–11 mm length and maximal attachment to sand grains. Blade growth of C. prolifera responded similarly to rhizoid production and reached a weekly growth rate of 30–130%.  相似文献   

17.
The effects of inorganic nitrogen (N) source (NH4+, NO3 or both) on growth, biomass allocation, photosynthesis, N uptake rate, nitrate reductase activity and mineral composition of Canna indica were studied in hydroponic culture. The relative growth rates (0.05-0.06 g g−1 d−1), biomass allocation and plant morphology of C. indica were indifferent to N nutrition. However, NH4+ fed plants had higher concentrations of N in the tissues, lower concentrations of mineral cations and higher contents of chlorophylls in the leaves compared to NO3 fed plants suggesting a slight advantage of NH4+ nutrition. The NO3 fed plants had lower light-saturated rates of photosynthesis (22.5 μmol m−2 s−1) than NH4+ and NH4+/NO3 fed plants (24.4-25.6 μmol m−2 s−1) when expressed per unit leaf area, but similar rates when expressed on a chlorophyll basis. Maximum uptake rates (Vmax) of NO3 did not differ between treatments (24-35 μmol N g−1 root DW h−1), but Vmax for NH4+ was highest in NH4+ fed plants (81 μmol N g−1 root DW h−1), intermediate in the NH4NO3 fed plants (52 μmol N g−1 root DW h−1), and lowest in the NO3 fed plants (28 μmol N g−1 root DW h−1). Nitrate reductase activity (NRA) was highest in leaves and was induced by NO3 in the culture solutions corresponding to the pattern seen in fast growing terrestrial species. Plants fed with only NO3 had high NRA (22 and 8 μmol NO2 g−1 DW h−1 in leaves and roots, respectively) whereas NRA in NH4+ fed plants was close to zero. Plants supplied with both forms of N had intermediate NRA suggesting that C. indica takes up and assimilate NO3 in the presence of NH4+. Our results show that C. indica is relatively indifferent to inorganic N source, which together with its high growth rate contributes to explain the occurrence of this species in flooded wetland soils as well as on terrestrial soils. Furthermore, it is concluded that C. indica is suitable for use in different types of constructed wetlands.  相似文献   

18.
We studied the decolorization of malachite green (MG) by the fungus Cunninghamella elegans. The mitochondrial activity for MG reduction was increased with a simultaneous increase of a 9-kDa protein, called CeCyt. The presence of cytochrome c in CeCyt protein was determined by optical absorbance spectroscopy with an extinction coefficient (E550-535) of 19.7 ± 6.3 mM−1 cm−1 and reduction potential of + 261 mV. When purified CeCyt was added into the mitochondria, the specific activity of CeCyt reached 440 ± 122 μmol min−1 mg−1 protein. The inhibition of MG reduction by stigmatellin, but not by antimycin A, indicated a possible linkage of CeCyt activity to the Qo site of the bc1 complex. The RT-PCR results showed tight regulation of the cecyt gene expression by reactive oxygen species. We suggest that CeCyt acts as a protein reductant for MG under oxidative stress in a stationary or secondary growth stage of this fungus.  相似文献   

19.
A case study on Centaurea gymnocarpa Moris & De Not., a narrow endemic species, was carried out by analyzing its morphological, anatomical, and physiological traits in response to natural habitat stress factors under Mediterranean climate conditions. The results underline that the species is particularly adapted to the environment where it naturally grows. At the plant level, the above-ground/below-ground dry mass (1.73 ± 0.60) shows its investment predominately in the above-ground structure with a resulting total leaf area per plant of 1399 ± 94 cm2. The senescent attached leaves at the base of the plant contribute to limit leaf transpiration by shading soil around the plant. Moreover, the dense C. gymnocarpa leaf pubescence, leaf rolling, the relatively high leaf mass area (LMA = 12.3 ± 1.3 mg cm−2) and leaf tissue density (LTD = 427 ± 44 mg cm−3) contribute to limit leaf transpiration, also postponing leaf death under dry conditions. At the physiological level, a relatively low respiration/photosynthesis ratio (R/PN) in spring results from high R [2.26 ± 0.59 μmol (CO2) m−2 s−1] and PN [12.3 ± 1.5 μmol (CO2) m−2 s−1]. The high photosynthetic nitrogen use efficiency [PNUE = 15.5 ± 0.4 μmol (CO2) g−1 (N) s−1] shows the large amount of nitrogen (N) invested in the photosynthetic machinery of new leaves, associated to a high chlorophyll content (Chl = 35 ± 5 SPAD units). On the contrary, the highest R/PN ratio (1.75 ± 0.19) in summer is due to a significant PN decrease and increase of R in response to drought. The low PNUE [1.5 ± 0.2 μmol (CO2) g−1 (N) s−1] in this season is indicative of a greater N investment in leaf cell walls which may contribute to limit transpiration. On the contrary, the low R/PN ratio (0.05 ± 0.02) in winter is resulting from the limited enzyme activity of the respiratory apparatus [R = 0.23 ± 0.08 μmol (CO2) m−2 s−1] while the low PNUE [3.5 ± 0.2 μmol (CO2) g−1 (N) s−1] suggests that low temperatures additionally limit plant production. The experiment of the imposed water stress confirms that the C. gymnocarpa growth capability is in conformity with the severe conditions of its natural habitat, likewise as it may be the case with others narrow endemic species that have occupied niches with similar extreme conditions.  相似文献   

20.
The response of rapid light–response curves (RLCs) of variable fluorescence to changes in short- and long-term photoacclimation status was studied in an estuarine benthic diatom. The diatom Nitzschia palea was grown under low- (LL, 20 μmol m−2 s−1) and high-light (HL, 400 μmol m−2 s−1) conditions, with the purpose of characterising the effects of long-term photoacclimation on (i) steady-state light–response curves (LC) of relative electron transport rate, rETR, (ii) the response of RLCs to changes in ambient irradiance (E, the irradiance to which the sample is acclimated to immediately before the RLCs), (iii) the relationship of RLCs to LC parameters and non-photochemical quenching (NPQ). Photoacclimation to LL and HL conditions induced distinct light–response patterns of rETR and NPQ. Higher growth light resulted in rETR vs. E curves with lower initial slopes (α, 0.591 μmol−1 m2 s vs. 0.661 μmol−1 m2 s, for HL and LL, respectively) and markedly higher maximum rates (rETRm, 95.9 vs. 29.3), reached under higher E levels (higher light-saturation coefficient, Ek: 162.4 μmol m−2 s−1 vs. 44.3 μmol m−2 s−1). Acclimation to HL induced bi-phasic NPQ vs. E curves, with minimum values reached under low E levels (15–25 μmol m−2 s−1) and not on dark-acclimated samples. The response of RLCs to changes in ambient irradiance varied with the long-term photoacclimation status of the samples. The initial slope, αRLC, decreased monotonically with E in LL cultures, from 0.68 to 0.25 μmol−1 m2 s, while varied bi-phasically in HL-acclimated samples. Typically, αRLC of HL cultures increased under low E, reaching a maximum of 0.61 μmol−1 m2 s under 25–55 μmol m−2 s−1, and decreased gradually under higher E levels to 0.25 μmol−1 m2 s. RLC maximum rETR, rETRm,RLC, and saturation coefficient Ek,RLC, increased with E following a saturation-like pattern, with the HL cultures presenting markedly higher values for all the E range (maximum rETRm,RLC values were 108.6 and 33.4 for HL and LL cultures, respectively). An inverse relationship was consistently found between αRLC and NPQ, both on LL and HL cultures, causing strong correlations (P < 0.001 in all cases) between NPQ and the high light-induced decrease of αRLC, ΔαRLC. RLCs were confirmed to also provide information on the long-term photoacclimation status, as significant correlations (P < 0.001 both for HL and LL cultures) were verified between Ek and an index based on RLC parameters, Êk, both for LL and HL cultures. These results reinforce the usefulness of RLCs as a tool for inferring on the short- and long-term photoacclimation status of samples with different long-term light histories, through the estimation of LC parameters and the monitoring of NPQ levels.  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