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 65 毫秒
1.
Evaluation of DNA damage and mutagenicity induced by lead in tobacco plants   总被引:1,自引:0,他引:1  
Tobacco (Nicotiana tabacum L. var. xanthi) seedlings were treated with aqueous solutions of lead nitrate (Pb2+) at concentrations ranging from 0.4 mM to 2.4 mM for 24 h and from 25 microM to 200 microM for 7 days. The DNA damage measured by the comet assay was high in the root nuclei, but in the leaf nuclei a slight but significant increase in DNA damage could be demonstrated only after a 7-day treatment with 200 microM Pb2+. In tobacco plants growing for 6 weeks in soil polluted with Pb2+ severe toxic effects, expressed by the decrease in leaf area, and a slight but significant increase in DNA damage were observed. The tobacco plants with increased levels of DNA damage were severely injured and showed stunted growth, distorted leaves and brown root tips. The frequency of somatic mutations in tobacco plants growing in the Pb2+-polluted soil did not significantly increase. Analytical studies by inductively coupled plasma optical emission spectrometry demonstrate that after a 24-h treatment of tobacco with 2.4 mM Pb2+, the accumulation of the heavy metal is 40-fold higher in the roots than in the above-ground biomass. Low Pb2+ accumulation in the above-ground parts may explain the lower levels or the absence of Pb2+-induced DNA damage in leaves.  相似文献   

2.
Direct evidence for the occurrence of an ADP-sensitive phosphoenzyme of (K+ + H+)-ATPase, the proton-pumping system of the gastric parietal cell is presented. The enzyme is phosphorylated with 5 μM [γ-32P]ATP in 50 mM imidazole-HCl (pH 7.0) and in the presence of 7–15 μM Mg2+. Addition of 5 mM ADP to this preparation greatly accelerates its hydrolysis. We have been able to establish this by stopping the phosphorylation with radioactive ATP, by adding 1 mM non-radioactive ATP, which leads to a slow monoexponential process of dephosphorylation of 32P-labeled enzyme. The relative proportion of the ADP-sensitive phosphoenzyme is 22% of the total phosphoenzyme. Values for the rate constants of breakdown and interconversion of the two phosphoenzyme forms have been determined.  相似文献   

3.
The presence of five P2X-like genes (p2xA–E) in Dictyostelium suggests that nucleotides other than cAMP may act as extracellular signalling molecules in this model eukaryote. However, p2xA was found to have an exclusively intracellular localisation making it unclear whether Dictyostelium utilise P2 receptors in a manner analogous to vertebrates. Using an apoaequorin expressing strain we show here that Dictyostelium do possess cell surface P2 receptors that facilitate Ca2+ influx in response to extracellular ATP and ADP (EC50 = 7.5 μM and 6.1 μM, respectively). Indicative of P2X receptor activation, responses were rapid reaching peak within 2.91 ± 0.04 s, required extracellular Ca2+, were inhibited by Gd3+, modified by extracellular pH and were not affected by deletion of either the single or iplA genes. Responses also remained unaffected by disruption of p2xA or p2xE showing that these genes are not involved. Cu2+ and Zn2+ inhibited purine-evoked Ca2+ influx with IC50 values of 0.9 and 6.3 μM, respectively. 300 μM Zn2+ completely abolished the initial large rapid rise in intracellular Ca2+ revealing the presence of an additional smaller, slower P2Y-like response. The existence of P2 receptors in Dictyostelium makes this organism a valuable model to explore fundamental aspects of purinergic signalling.  相似文献   

4.
Tobacco seedlings (Nicotiana tabacum var. xanthi) were treated for 24 h with mono-(2- and 3-CBA), di-(2,5- and 3,4-CBA), and tri-(2,4,6- and 2,3,5-CBA)-chlorobenzoic acids (CBAs) and with the mixture of polychlorinated biphenyls – Delor 103, or cultivated for 1 or 2 weeks in soil polluted with the CBAs. DNA damage in nuclei of leaves and roots was evaluated by the comet assay. A significant increase in DNA damage was observed only at concentrations of CBAs that caused withering of leaves or had lethal effects within 2–4 weeks after the treatments. As the application of CBAs did not induce somatic mutations, the induced DNA migration is probably caused by necrotic DNA fragmentation and not by DNA damage resulting in genetic alteration. In contrast, the application of the monofunctional alkylating agent ethyl methanesulphonate as a positive control resulted in a dose–response increase of DNA damage and an increase of somatic mutations. Thus, the EMS-produced DNA migration is probably associated with genotoxin-induced DNA fragmentation. The data demonstrate that the comet assay in plants should be conducted together with toxicity studies to distinguish between necrotic and genotoxin-induced DNA fragmentation. The content of 2,5-CBA in tobacco seedlings was measured by reverse-phase high pressure liquid chromatography.  相似文献   

5.
The effects of NaCl salinity on growth, morphology and photosynthesis of Salvinia natans (L.) All. were investigated by growing plants in a growth chamber at NaCl concentrations of 0, 50, 100 and 150 mM. The relative growth rates were high (ca. 0.3 d−1) at salinities up to 50 mM and decreased to less than 0.2 d−1 at higher salinities, but plants produced smaller and thicker leaves and had shorter stems and roots, probably imposed by the osmotic stress and lowered turgor pressure restricting cell expansion. Na+ concentrations in the plant tissue only increased three-fold, but uptake of K+ was reduced, resulting in very high Na+/K+ ratios at high salinities, indicating that S. natans lacks mechanisms to maintain ionic homeostasis in the cells. The contents of proline in the plant tissue increased at high salinity, but concentrations were very low (<0.1 μmol g−1 FW), indicating a limited capacity of S. natans to synthesize proline as a compatible compound. The potential photochemical efficiency of PSII (Fv/Fm) of S. natans remained unchanged at 50 mM NaCl but was reduced at higher salinities, and the photosynthetic capacity (ETRmax) was significantly reduced at 50 mM NaCl and higher. It is concluded that S. natans is a salt-sensitive species lacking physiological measures to cope with exposure to high NaCl salinity. At low salinities salts are taken up and accumulate in old leaves, and high growth rates are maintained because new leaves are produced at a higher rate than for plants not exposed to salt.  相似文献   

6.
Recent novel mixed blooms of several species of toxic raphidophytes have caused fish kills and raised health concerns in the highly eutrophic Inland Bays of Delaware, USA. The factors that control their growth and dominance are not clear, including how these multi-species HAB events can persist without competitive exclusion occurring. We compared and contrasted the relative environmental niches of sympatric Chattonella subsalsa and Heterosigma akashiwo isolates from the bays using classic Monod-type experiments. C. subsalsa grew over a temperature range from 10 to 30 °C and a salinity range of 5–30 psu, with optimal growth occurring from 20 to 30 °C and 15 to 25 psu. H. akashiwo had similar upper temperature and salinity tolerances but also lower limits, with growth occurring from 4 to 30 °C and 5 to 30 psu and optimal growth between 16 and 30 °C and 10 and 30 psu. These culture results were confirmed by field observations of bloom occurrences in the Inland Bays. Maximum nutrient-saturated growth rates (μmax) for C. subsalsa were 0.6 d−1 and half-saturation concentrations for growth (Ks) were 9 μM for nitrate, 1.5 μM for ammonium, and 0.8 μM for phosphate. μmax of H. akashiwo (0.7 d−1) was slightly higher than C. subsalsa, but Ks values were nearly an order of magnitude lower at 0.3 μM for nitrate, 0.3 μM for ammonium, and 0.2 μM for phosphate. H. akashiwo is able to grow on urea but C. subsalsa cannot, while both can use glutamic acid. Cell yield experiments at environmentally relevant levels suggested an apparent preference by C. subsalsa for ammonium as a nitrogen source, while H. akashiwo produced more biomass on nitrate. Light intensity affected both species similarly, with the same growth responses for each over a range from 100 to 600 μmol photons m−2 s−1. Factors not examined here may allow C. subsalsa to persist during multi-species blooms in the bays, despite being competitively inferior to H. akashiwo under most conditions of nutrient availability, temperature, and salinity.  相似文献   

7.
Jingmei Zhang  Jiaxi Liu  Zukeng Chen  Jinxing Lin   《Flora》2007,202(7):581-588
The calcium inhibitors A23187, EGTA and La3+ inhibit pollen grain germination and growth of pollen tubes of Lilium davidii var. unicolor at different concentrations. Treatment with 10−4 or 10−5 M ionophores A23187 reduced germination rate and resulted in distortion of pollen tube. Addition of 2 or 10 mM of the chelator EGTA disturbed the direction of pollen tube growth and extended the diameter of pollen tube as observed by light and confocal microscopy. The Ca2+-channel blocker lanthanum chloride (La3+) restrained germination or markedly caused transformation of pollen tube. Furthermore, all treatments led to disappearance of any calcium gradient. Calcium distribution in pollen grain and pollen tube was altered as shown by confocal microscopy for each treatment. This indicates that the inhibitors influence pollen development by affecting the calcium gradient which may play a critical role in germination and tube growth. Fourier transform infrared (FTIR) spectra indicated slight increases in contents of amide I and a substantial decrease in the content of aliphatic esters and saturated esters in treated pollen tubes compared with normal pollen tubes. The FTIR analysis confirmed that EGTA and La3+ weakened the accumulation of ester in pollen tubes, which may be associated with an increased content of amide I.  相似文献   

8.
Summary A gram negative, motile, short rod-shaped, and nickel resistant (tolerating 6.5 mM Ni2+) bacterium, strain BB1A, was isolated from the waters of the River Torsa in Hashimara, Jalpaiguri district, West Bengal, India. The isolate BB1A was identified as a strain of Acinetobacter junii following detailed analysis of morphological, physio-biochemical and 16S rRNA gene sequence. The expression of nickel resistance in BB1A was inducible by exposure to nickel chloride at a concentration as low as 50 μM Ni2+. The other metal ions, Cu2+, Zn2+, or Pb2+ at a concentration range of 20–30 μM, also induced the nickel resistance system in this bacterium. Southern hybridizations of BB1A genomic DNA with digoxigenin-dUTP labeled DNA probes specific for well known nickel resistance determinants, cnr, ncc or nre, resulted in no detectable signal, but nir specific probe yielded weak hybridization signal with restricted genomic DNA of BB1A. The isolate BB1A, therefore, carries out a novel induction phenomenon of nickel resistance and presumably with a nickel resistance genetic system different from that previously characterized in other bacteria.  相似文献   

9.
In vitro selection of Zn2+-dependent RNA-cleaving DNAzymes with activity at 90°C has yielded a diverse spool of selected sequences. The RNA cleavage efficiency was found in all cases to be specific for Zn2+ over Pb2+, Ca2+, Cd2+, Co2+, Hg2+, and Mg2+. The Zn2+-dependent activity assay of the most active sequence showed that the DNAzyme possesses an apparent Zn2+-binding dissociation constant of 234 μM and that its activity increases with increasing temperatures from 50–90°C. A fit of the Arrhenius plot data gave Ea = 15.3 kcal mol−1. Surprisingly, the selected Zn2+-dependent DNAzymes showed only a modest (∼3-fold) activity enhancement over the background rate of cleavage of random sequences containing a single embedded ribonucleotide within an otherwise DNA oligonucleotide. The result is attributable to the ability of DNA to sustain cleavage activity at high temperature with minimal secondary structure when Zn2+ is present. Since this effect is highly specific for Zn2+, this metal ion may play a special role in molecular evolution of nucleic acids at high temperature.  相似文献   

10.
The physiological effects of thallium(I)-acetate on the duckweed Lemna minor after 1-, 4-, 7- and 14-d exposure were analyzed. High bioaccumulation of Tl (221 mg kg−1 dry wt at 2.0 μM Tl-acetate) caused an inhibition of plant growth. After 14-d exposure, 0.2, 0.5, 1.0 and 2.0 μM Tl-acetate reduced the frond-number growth rate by 21.1%, 39.4%, 66% and 83.1%, respectively. Tl-acetate also induced a modulation of the antioxidative response by depleting the ascorbate content and affecting the antioxidative enzymes activities. Superoxide dismutase showed a continuous increase of activity (31–67%) after Tl-acetate exposure. Other antioxidative enzymes displayed a biphasic response to both the concentration and the exposure period. Exposure up to 7 d decreased the catalase activity (up to 40%) in plants treated with higher Tl-acetate concentrations. In contrast, 14-d exposure increased the activity of the enzyme (≥90%). Short-term exposure increased ascorbate peroxidase activity (13–41%), except in plants exposed to the highest Tl-acetate concentration. However, 14-d exposure decreased the enzyme activity at all concentrations tested (38–60%). Although pyrogallol peroxidase activity increased (up to 26%) during 4-d exposure, longer exposures to the highest two concentrations decreased the activity of the enzyme (25–48%). In general, short-term exposure to Tl-acetate activated the antioxidant capacity, which resulted in recovery of the frond-number growth rates in Tl-treated plants. In spite of the activation of the antioxidative response during short-term exposure, higher Tl-acetate concentrations increased the hydrogen peroxide level (up to 45%) and induced marked oxidative damage to lipids, proteins and DNA. Longer exposure induced a decline of the antioxidative response, and plants showed the symptoms of oxidative damage even at lower Tl-acetate concentrations. The genotoxic effect was evaluated by an alkaline version of the cellular and acellular Comet assay, which revealed an indirect genotoxic effect of Tl-acetate, suggesting oxidatively induced damage to DNA.  相似文献   

11.
The NAD-dependent glutamate dehydrogenase (GDH) (EC 1.4.1.2) fromLaccaria bicolorwas purified 410-fold to apparent electrophoretic homogeneity with a 40% recovery through a three-step procedure involving ammonium sulfate precipitation, anion-exchange chromatography on DEAE–Trisacryl, and gel filtration. The molecular weight of the native enzyme determined by gel filtration was 470 kDa, whereas sodium dodecyl sulfate–polyacrylamide gel electrophoresis gave rise to a single band of 116 kDa, suggesting that the enzyme is composed of four identical subunits. The enzyme was specific for NAD(H). The pH optima were 7.4 and 8.8 for the amination and deamination reactions, respectively. The enzyme was found to be highly unstable, with virtually no activity after 20 days at −75°C, 4 days at 4°C, and 1 h at 50°C. The addition of ammonium sulfate improved greatly the stability of the enzyme and full activity was still observed after several months at −75°C. NAD-GDH activity was stimulated by Ca2+and Mg2+but strongly inhibited by Cu2+and slightly by the nucleotides AMP, ADP, and ATP. The Michaelis constants for NAD, NADH, 2-oxoglutarate, and ammonium were 282 μM, 89 μM, 1.35 mM, and 37 mM, respectively. The enzyme had a negative cooperativity for glutamate (Hill number of 0.3), and itsKmvalue increased from 0.24 to 3.6 mM when the glutamate concentration exceeded 1 mM. These affinity constants of the substrates, compared with those of the NADP-GDH of the fungus, suggest that the NAD-GDH is mainly involved in the catabolism of glutamate, while the NADP-GDH is involved in the catalysis of this amino acid.  相似文献   

12.
ATP and adenylylimidodiphosphate (AdoPP[NH]P) bind to (Na+ + K+)-ATPase in the absence of Mg2+ (EDTA present) with a homogeneous but 15-fold different affinity, the Kd values being 0.13 μM and 1.9 μM, respectively. The binding capacities of the two nucleotides are nearly equal and amount to 3.9 and 4 nmol/mg protein or 1.7 and 1.8 mol/mol (Na+ + K+)-ATPase, respectively. The Kd value for ATP is equal to the Km for phosphorylation by ATP (0.05–0.25 μM) and the binding capacity is equivalent to the phosphorylation capacity of 1.8 mol/mol (Na+ + K+)-ATPase. Hence, the enzyme contains two high-affinity nucleotide binding and phosphorylating sites per molecule, or one per α-subunit. Additional low-affinity nucleotide binding sites are elicited in the presence of Mg2+, as shown by binding studies with the non-phosphorylating (AdoPP[NH]P). The Kd and binding capacity for AdoPP[NH]P at these sites is dependent on the Mg2+ concentration. The Kd increases from 0.06 mM at 0.5 mM Mg2+ to a maximum of 0.26 mM at 2 mM Mg2+ and the binding capacity from 1.5 nmol/mg protein at 0.5 mM Mg2+ to 3.3 nmol/mg protein at 4 mM Mg2+. Extrapolation of a double reciprocal plot of binding capacity vs. total Mg2+ concentration yields a maximal binding capacity at infinite Mg2+ concentration of 3.8 nmol/mg protein or 1.7 mol/mol (Na+ + K+)-ATPase. The Kd for Mg2+ at the sites, where it exerts this effect, is 0.8 mM. The Kd for the high-affinity sites increases from 1.5–1.9 μM in the absence of Mg2+ to a maximum of 4.2 μM at 2 mM Mg2+ concentration. The binding capacity of these sites (1.8 mol/mol enzyme) is independent of the Mg2+ concentration. Hence, Mg2+ induces two low-affinity non-phosphorylating nucleotide binding sites per molecule (Na+ + K+)-ATPase in addition to the two high-affinity, phosphorylating nucleotide binding sites.  相似文献   

13.
Field and laboratory experiments were designed to determine the differential growth and toxin response to inorganic and organic nitrogen additions in Pseudo-nitzschia spp. Nitrogen enrichments of 50 μM nitrate (KNO3), 10 μM ammonium (NH4Cl), 20 μM urea and a control (no addition) were carried out in separate carboys with seawater collected from the mouth of the San Francisco Bay (Bolinas Bay), an area characterized by high concentrations of macronutrients and iron. All treatments showed significant increases in biomass, with chlorophyll a peaking on days 4–5 for all treatments except urea, which maintained exponential growth through the termination of the experiment. Pseudo-nitzschia australis Frenguelli abundance was 103 cells l−1 at the start of the experiment and increased by an order of magnitude by day 2. Particulate domoic acid (pDA) was initially low but detectable (0.15 μg l−1), and increased throughout exponential and stationary phases across all treatments. At the termination of the experiment, the urea treatment produced more than double the amount of pDA (9.39 μg l−1) than that produced by the nitrate treatment (4.26 μg l−1) and triple that of the control and ammonium treatments (1.36 μg l−1 and 2.64 μg l−1, respectively). The mean specific growth rates, calculated from increases in chlorophyll a and from cellular abundance of P. australis, were statistically similar across all treatments.These field results confirmed laboratory experiments conducted with a P. australis strain isolated from Monterey Bay, CA (isolate AU221-a) grown in artificial seawater enriched with 50 μM nitrate, 50 μM ammonium or 25 μM of urea as the sole nitrogen source. The exponential growth rate of P. australis was significantly slower for cells grown on urea (ca. 0.5 day−1) compared to the cells grown on either nitrate or ammonium (ca. 0.9 day−1). However the urea-grown cells produced more particulate and dissolved domoic acid (DA) than the ammonium- or nitrate-grown cells. The field and laboratory experiments demonstrate that P. australis is able to grow effectively on urea as the primary source of nitrogen and produced more pDA when grown on urea in both natural assemblages and unialgal cultures. These results suggest that the influence of urea from coastal runoff may prove to be more important in the development or maintenance of toxic blooms than previously thought, and that the source of nitrogen may be a determining factor in the relative toxicity of west coast blooms of P. australis.  相似文献   

14.
Furan is found in various food items and is cytotoxic and carcinogenic in the liver of rats and mice. Metabolism of furan includes the formation of an unsaturated dialdehyde, cis-2-butene-1,4-dial (BDA). In view of the multifunctional electrophilic reactivity of BDA, adduct formation with protein and DNA may explain some of the toxic effects. Short-term tests for genotoxicity of furan in mammalian cells are inconclusive, little is known for BDA. We investigated BDA generated by hydrolysis of 2,5-diacetoxy-2,5-dihydrofuran for genotoxicity in L5178Y tk+/− mouse lymphoma cells using standard procedures for the comet assay, the micronucleus test, and the mouse lymphoma thymidine kinase gene mutation assay, using 4-h incubation periods. Cytotoxicity was remarkable: cell viability at concentrations ≥50 μM was reduced to <50%. In the dose range up to 25 μM, viability was >90%. Measures of comet-tail length and thymidine–kinase mutant frequency were increased 1.6- and 2.4-fold above control, respectively. Analysis of three fully independent replicates with a linear mixed-effects model showed a highly significant increase with concentration for both endpoints. Compared to methyl methanesulfonate used as a positive control, BDA was of similar potency with respect to genotoxicity, but it was much more cytotoxic. Furan added to cell cultures at doses that resulted in time-averaged effective concentrations of up to 3100 μM was neither cytotoxic nor genotoxic. A potential cross-linking activity of BDA was investigated by checking whether gamma radiation-induced DNA migration in the comet assay could be reduced by pre-treatment with BDA. In contrast to the effect of the positive control glutaraldehyde, BDA treatment did not reduce the comet tail length. On the contrary, an increase was observed at ≥100 μM BDA, which was attributable to early apoptotic cells. Although BDA was found to be a relatively potent genotoxic agent in terms of the concentration necessary to double the background measures, cytotoxicity strongly limited the concentration range that produced interpretable results. This may explain some of the inconclusive results and indicates that non-genotoxic effects must be taken into account in the discussion of the modes of toxic and carcinogenic action of furan.  相似文献   

15.
Nitraria retusa and Atriplex halimus (xero-halophytes) plants were grown in the range 0–800 mM NaCl while Medicago arborea (glycophyte) in 0–300 mM NaCl. Salt stress caused a marked decrease in osmotic potential and a significant accumulation of Na+ and Cl in leaves of both species. Moderate salinity had a stimulating effect on growth rate, net CO2 assimilation, transpiration and stomatal conductance for the xero-halophytic species. At higher salinities, these physiological parameters decreased significantly, and their percentages of reduction were higher in A. halimus than in N. retusa whereas, in M. arborea they decreased linearly with salinity. Nitraria retusa PSII photochemistry and carotenoid content were unaffected by salinity, but a reduction in chlorophyll content was observed at 800 mM NaCl. Similar results were found in A. halimus, but with a decrease in the efficiency of PSII (F′v/F′m) occurred at 800 mM. Conversely, in M. arborea plants we observed a significant reduction in pigment concentrations and chlorophyll fluorescence parameters. The marked toxic effect of Na+ and/or Cl observed in M. arborea indicates that salt damage effect could be attributed to ions’ toxicity, and that the reduction in photosynthesis is most probably due to damages in the photosynthetic apparatus rather than factors affecting stomatal closure. For the two halophyte species, it appears that there is occurrence of co-limitation of photosynthesis by stomatal and non-stomatal factors. Our results suggest that both N. retusa and A. halimus show high tolerance to both high salinity and photoinhibition while M. arborea was considered as a slightly salt tolerant species.  相似文献   

16.
A batch sorption technique was used to study the biosorption of Pb2+, Cd2+ and Zn2+ ions onto the vastly abundant water hyacinth weed, Eichhornia crassipes biomass in binary and ternary systems at a temperature of 30 °C and pH 4.84. Mutual interference effects were probed using equilibrium adsorption capacity ratios, , where the prime indicates the presence of one or two other metal ions. The combined action of the metals was found to be antagonistic, and the metal sorption followed the order Pb2+  Cd2+  Zn2+. The behaviour of competitive biosorption for Pb–Cd and Pb–Zn combinations were successfully described by the Langmuir Competitive Model (CLM), whilst the model showed poor fitting to the Cd–Zn data. In conclusion, Pb2+ ions could still be effectively removed from aqueous solution in the presence of both Cd2+ and Zn2+ ions, but removal of the Cd2+ and Zn2+ ions would be suppressed in the presence of Pb2+.  相似文献   

17.
The Delaware Inland Bays (DIB) have experienced harmful algal blooms of dinoflagellates and raphidophytes in recent years. We used quantitative polymerase chain reaction (QPCR) techniques to investigate the community dynamics of three DIB dinoflagellates (Karlodinium veneficum, Gyrodinium instriatum, and Prorocentrum minimum) and one raphidophyte (Heterosigma akashiwo) at a single site in the DIB (IR-32) in summer 2006 relative to salinity, temperature and nutrient concentrations. We also carried out complementary laboratory culture studies. New primers and probes were developed and validated for the 18S rRNA genes in the three dinoflagellates. K. veneficum, H. akashiwo, and G. instriatum were present in almost all samples throughout the summer of 2006. In contrast, P. minimum was undetectable in late June through September, when temperatures ranged from 20 to 30 °C (average 25.7 °C). Dissolved nutrients ranged from 0.1 to 2.8 μM PO43− (median = 0.3 μM), 0.7–30.2 μM NOx (median = 12.9 μM), and 0–19.4 μM NH4+ (median = 0.7 μM). Dissolved N:P ratios covered a wide range from 2.6 to 177, with a median of 40. There was considerable variability in occurrence of the four species versus nutrients, but in general P. minimum and H. akashiwo were most abundant at higher (>40) N:P ratios and dissolved nitrogen concentrations, while K. veneficum and G. instriatum were most abundant at low dissolved N:P ratios (<20) and dissolved nitrogen concentrations < 10 μM. The semi-continuous laboratory competition experiment used mixed cultures of K. veneficum, P. minimum, and H. akashiwo grown at dissolved N:P ratios of 5, 16, and 25. At an N:P of 16 and 25 P. minimum was the dominant alga at the end of the experiment, even at a temperature that was much higher than that at which this alga was found to bloom in the field (27 °C). P. minimum and H. akashiwo had highest densities in the N:P of 25. K. veneficum grew equally well at all three N:P ratios, and was co-dominant at times at an N:P of 5. H. akashiwo had the lowest densities of the three algae in the laboratory experiment. Laboratory and field results showed both interesting similarities and significant differences in the influences of important environmental factors on competition between these harmful algal species, suggesting the need for more work to fully understand HAB dynamics in the DIB.  相似文献   

18.
Two extracellular chitinases were purified from Paecilomyces variotii DG-3, a chitinase producer and a nematode egg-parasitic fungus, to homogeneity by DEAE Sephadex A-50 and Sephadex G-100 chromatography. The purified enzymes were a monomer with an apparent molecular mass of 32 kDa (Chi32) and 46 kDa (Chi46), respectively, and showed chitinase activity bands with 0.01% glycol chitin as a substrate after SDS-PAGE. The first 20 and 15 N-terminal amino acid sequences of Chi32 and Chi46 were determined to be Asp-Pro-Typ-Gln-Thr-Asn-Val-Val-Tyr-Thr-Gly-Gln-Asp-Phe-Val-Ser-Pro-Asp-Leu-Phe and Asp-Ala-X-X-Tyr-Arg-Ser-Val-Ala-Tyr-Phe-Val-Asn-Trp-Ala, respectively. Optimal temperature and pH of the Chi32 and Chi46 were found to be both 60°C, and 2.5 and 3.0, respectively. Chi32 was almost inhibited by metal ions Ag+ and Hg2+ while Chi46 by Hg2+ and Pb2+ at a 10 mM concentration but both enzymes were enhanced by 1 mM concentration of Co2+. On analyzing the hydrolyzates of chitin oligomers [(GlcNAc) n , n = 2–6)], it was considered that Chi32 degraded chitin oligomers as an exo-type chitinase while Chi46 as an endo-type chitinase.  相似文献   

19.
Cell suspension cultures of Agave amaniensis and Costus speciosus were grown in media containing Cd2 + up to 25 and 20 mg l–1, respectively, and Pb2+ up to 40 mg l–1. The cultures hyper-accumulated Cd2+ up to 900 and 530 g g–1 and Pb2+ up to 1390 and 1170 g g–1 dry wt. in their respective biomasses. Increasing Pb2+ up to 30 mg l–1 increased the biomass production and total sitosterol content of Costus speciosus by up to 1.7- and 1.3-fold, respectively.  相似文献   

20.
Store-operated Ca2+ entry (SOCE) is an important mechanism for Ca2+ influx in smooth muscle cells; however the activation and regulation of this influx pathway are incompletely understood. In the present study we have examined the effect of several protein kinases in regulating SOCE in pulmonary artery smooth muscle cells (PASMCs) of the rat. Inhibition of protein kinase C with chelerythrine (3 μM) potentiated SOCE by 47 ± 2%, while the tyrosine kinase inhibitors genistein (100 μM) and tyrphostin 23 (100 μM) caused a significant reduction in SOCE of 55 ± 9% and 43 ± 7%, respectively. It has been proposed that Ca2+-insensitive phospholipase A2 (iPLA2) is involved in the activation of SOCE in many different cell types. The iPLA2 inhibitor, bromoenol lactone had no effect on SOCE, suggesting that this mechanism was not involved in the activation of the pathway. The calmodulin antagonists, calmidazolium (CMZ) (10 μM) and W-7 (10 μM) appeared to potentiate SOCE in PASMCs. Further investigation established that CMZ was actually activating a Ca2+ influx pathway that was independent of the filling state of the sarcoplasmic reticulum. The CMZ-activated Ca2+ influx was blocked by Gd3+ (10 μM), but unaffected by 2-APB (75 μM), indicating a pharmacological profile distinct from the classical SOCE pathway.  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