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.
1. Diel patterns of the frequency of dividing cells (FDC) of the bloom‐forming cyanobacteria Microcystis were investigated using both culture strains and natural populations. 2. In laboratory experiments, diel division cycles were examined twice in a 24‐h light/dark cycle during time‐course batch incubations of six culture conditions using two strains (morphospecies) of Microcystis (M. aeruginosa and M. wesenbergii). While both strains clearly showed phased cell division in the light period during the logarithmic growth phase, the peaks of FDC became unclear towards the stationary phases. Some dividing cells were always found in the dark period regardless of whether or not division had paused at the same time. 3. This result implied the inadequacy of applying the model of McDuff & Chisholm [Limnology and Oceanography (1982) vol. 27 , pp. 783–787] directly to calculate the duration of cell division. Modified equations are proposed to calculate the duration of cytokinesis as a terminal event, in which the FDC values at night are regarded as 0% and all FDC values are subtracted by the minimum FDC value. 4. The diel FDC in natural populations of M. aeruginosa and M. wesenbergii were examined at five sites from a harbour to several distances offshore in Lake Biwa. While both species showed phased cell division patterns in the daytime at the harbour, no peaks in FDC were discernible in the samples taken from the offshore sites. These results strongly suggested that Microcystis growth was higher inshore than offshore. The in situ growth rates were estimated using the new equations.  相似文献   

2.
Natural populations of planktonic Synechococcus spp. exhibit diel variations in abundance and frequency of dividing cells (FDC) during the warm seasons (>25°C) in coastal water of the western subtropical Pacific ocean. We hypothesized that differences in grazing rates during the day/dark cycle were a major cause of the observed diel variations in Synechococcus spp. abundance. We used fluorescently labeled particles (FLP) as tracers of feeding on Synechococcus spp. in June and August 2007. Our results showed that FDC of Synechococcus spp. were highest at dusk (50%) and lowest (<10%) between midnight and early morning. Synechococcus spp. had three abundance phases, accrual (I), peak (II), and diminished (III). Moreover, FDC values gradually increased in phase I, declined to the lowest values during phase II, and remained at low levels throughout phase III. Our results strongly indicated that pigmented nanoflagellate (PNF) grazing was the underlying biological factor regulating diel variations in Synechococcus spp. abundance. And, the rate of PNF ingestion peaked during the night on smaller non-dividing cells following a peak of Synechococcus spp. abundance in phase II. These findings provide evidence that diel variations in ingestion rates are affected by non-dividing cells of Synechococcus spp. and imply that the impacts of PNF grazing on Synechococcus spp. is based on food selectivity by size.  相似文献   

3.
The diel variability in picophytoplankton cell death was analyzed by quantifying the proportion of dead cyanobacteria Prochlorococcus and Synechococcus cells along several in situ diel cycles in the open Mediterranean Sea. During the diel cycle, total cell abundance varied on average 2.8 ± 0.6 and 2.6 ± 0.4 times for Synechococcus and Prochlorococcus populations, respectively. Increasing percentages of dead cells of Prochlorococcus and Synechococcus were observed during the course of the day reaching the highest values around dusk and decreasing as the night progressed, indicating a clear pattern of diel variation in the cell mortality of both cyanobacteria. Diel cycles of cell division were also monitored. The maximum percentage of dead cells (Max % DC) and the G2 + M phase of the cell division occurred within a period of 2 h for Synechoccoccus and 4.5 h for Prochlorococcus, and the lowest fraction of dead cells occurred at early morning, when the maximum number of cells in G1 phase were also observed. The G1 maximum corresponded with the maximal increase in newly divided cells (minimum % dead cells), and the subsequent exposure of healthy daughter cells to environmental stresses during the day resulted in the progressive increase in dying cells, with the loss of these cells from the population when cell division takes place. The discovery of diel patterns in cell death observed revealed the intense dynamics of picocyanobacterial populations in nature.  相似文献   

4.
Eugene gracilis Klebs (Z) was grown in a cyclostat (continuous culture on a light/dark cycle) at growth limiting levels of phosphate. Cell division was restricted to the dark period regardless of the proportion of the cells dividing during each 24 h period. Growth rate, as reflected by the amplitude of the cell density oscillation, was correlated with dilution rate. The width of the division gate was analyzed using a phasing index and found to be narrowest at dilution rates where the mean generation time of the cell population was an even multiple of 24 h. The effect was attributed to enhanced phasing of the cell division process by the biological clock of Euglena. Residual phosphate levels in the cyclostat were less than 0.3 μM PO4 at all submaximal growth rates. Cellular phosphorus concentration increased with dilution rate as described by a hyperbola saturating at Dmax= 0.74 day−1 with 8 × 10−8μM P/cell as the minimum intracellular phosphorus concentration for growth. The results are discussed, in terms of the inherent similarities and differences between a cyclostat and a steady state chemostat, and the advantages of the cyclostat for studies in phytoplankton ecology.  相似文献   

5.
Diel changes in the frequency of dividing cells (FDC) of three Microcystis species were investigated in a small eutrophic pond from July to October 2005. The representative species was M. aeruginosa (Kütz.) Kütz., constituting 57%–86% of the Microcystis population throughout the study period, and the remainder were M. viridis (A. Braun) Lemmerm. and M. wesenbergii (Komárek) Komárek. The FDC of M. aeruginosa and M. wesenbergii increased in the daytime and fell in the nighttime in July and August, but this regular variation was not observed in September or October. The in situ specific growth rates of Microcystis species were estimated based on the assumption that the specific growth rate can be given as an absolute value of the derivative of FDC with respect to time. The calculated values were similar among species—0.15–0.38 · d?1 for M. aeruginosa, 0.14–0.63 · d?1 for M. viridis, and 0.18–0.61 · d?1 for M. wesenbergii. The specific growth rates in July and August slightly exceeded those in September and October. The analysis of the in situ specific growth rate of Microcystis indicated that recruitment of the benthic population or morphological change, rather than massive growth, was at least partly responsible for the dominance of M. aeruginosa in the study pond.  相似文献   

6.
A pico sized Synechococcus species isolated from Lake Balaton was studied in batch and continuous cultures. This picocyanobacterium had a pH optimum at 8.5 and a temperature optimum at 28-30°C. The Ik value for growth was 52 μEinstein m−2 S−1, the maximum growth rate 2.27 d−1, the half saturation Constant of growth 1.2 μg PO4-P I−1 and the minimal cell quota 1.74 nig P g dry weight−1. The dry weight of cells showed a minimum, the chlorophyll-a/biomass ratio a maximum as a function of growth rate. Above the quota of 3.4 fg P Cell−1 significant amounts of non-reactive dissolved Phosphorous were released.  相似文献   

7.
In situ incubations of natural autotrophic picoplankton populations during a 15 month study were used to test the frequency of dividing cells proceduresin estimating phototrophic picoplankton growth rates. These rates were estimated using dilution experiments and compared to the average frequency of dividing cells over the same time interval. The regression equation of µ = 2.85 × 10–3 (FDC) + 0.022 was calculated to relate autotrophic picoplankton growth rate and the frequency of dividing cells in this study. The resulting relationship was compared to 14C-bicarbonate derived growth rates. Productivity estimates using frequency of dividing cells correlated closely to sodium 14C-bicarbonate results and indicated a range of productivity by autotrophic picoplankton of 55.6% the total phytoplankton primary productivity in July to a January rate of 2.3%. Annual autotrophic picoplankton abundance varied seasonally in the lower Chesapeake Bay ranging from 7.26 × 106 cells 1–1 in winter to 9.28 × 108 cells 1–1 during late summer.  相似文献   

8.
In the North Atlantic over a wide geographic region that includes various oceanic regimes and a temperature range from 10 to 22° C, an increase in the number of nondividing Synechococcus cells (X) was generally accompanied by a greater-than-proportional increase in the number of dividing cells (Y). As a result, the fraction of dividing cells (FDC = Y · (Y + X)?1) was positively related to population size (Y + X). Recognizing that FDC is generally greater in a rapidly growing population than in a slowly growing one, our empirical finding implies a positive correlation between specific growth rate and standing stock for Synechococcus. One notable exception occurred during winter (T < 5°C) in a eutrophic coastal embayment when a decrease in cell abundance was not matched by a decrease in FDC.  相似文献   

9.
A mathematical model of predator-prey interactions was used to predict the relationship between population size and cellular growth rate in a two-tiered trophic system consisting of Synechococcus PCC 6301 and Tetrahymena pyriformis. As predicted, axenic chemostat cultures of Synechococcus responded to increased nutrient availability by expanding the equilibrium population size without a concurrent change in growth rate. Likewise, the addition of the predator Tetrahymena pyriformis decreased the Synechococcus population size by 85% and increased the Synechococcus growth rate. Synechococcus populations in the surface waters of the Gulf of Mexico were sampled to ascertain whether the relationship between population size and cellular 16S rRNA concentration conformed to that predicted by the model. Direct counts of autofluorescent cells in size-fractionated seawater samples provided an estimate of Synechococcus population size. The growth rate of in situ populations was estimated by measuring the extent of hybridization of an oligonucleotide probes complementary to Synechococcus 16S rRNA, based on evidence that ribosomal RNA content increases concurrently with growth rate. The comparison of in situ population sizes and specific growth rates revealed that relatively large Synechococcus populations were growing slowly, indicative of nutrient limitation, and that quickly growing populations were relatively small, as predicted for predator-limited populations.  相似文献   

10.
The present study reports on effects of different light:dark periods, light intensities, N:P ratios and temperature on the specific growth rate of flagellated cells of Phaeocystis pouchetii in culture. The specific growth rate was estimated by diel changes in cellular DNA content. The cellular DNA content and cell cycle of flagellated cells of P. pouchetii are shown, and the importance of light:dark period in cell division is demonstrated. Diel patterns of the cellular DNA content showed that cell division was confined to the dark period. The cells dealt with more than one division per day by rapid divisions shortly after each other.The specific growth rates (μDNA) based on the DNA cell cycle model were in close agreement with specific growth rates (μCell) determined from cell counts. The temperature affected the specific growth rates (multiple regression, p < 0.01) and were higher at 5 °C (μ ≤ 2.2 d−1) than at 10 °C (μ ≤1.6 d−1). Increasing the light:dark period from 12:12 h to 20:4 h affected the specific growth rate of P. pouchetii at the lower temperature tested (5 °C) (multiple regression, p < 0.01), resulting in higher specific growth rates than at 10 °C. At 10 °C, the effect of light:dark period was severely reduced. Neither light nor nutrients could compensate the reduction in specific growth rates caused by elevated temperature. The specific growth rates was not affected by the N:P ratios tested (multiple regression, p = 0.21). The experiments strongly suggest that the flagellated cells have a great growth potential and could play a dominating role in northern areas at increased day length.  相似文献   

11.
Measurements of bacterial secondary production were carried out during 13 diel studies at one coastal marine station and in five lakes differing with respect to nutrient concentration and primary production. Bacterial secondary production was measured in situ every 3 to 5 h by [3H]thymidine incorporation into DNA. In some of the diel studies, these results were compared with results obtained from dark 14CO2 uptake and frequency of dividing cells. Only minor diel changes were observed. The rate of [3H]thymidine incorporation into DNA and the frequency of dividing cells varied from 23 to 194% of the diel mean. The dark CO2 uptake rate varied from 12 to 259% of the diel mean. An analysis of variance demonstrated that no specific time periods during 24 h showed significantly different production rates, supporting the idea that bacterial activities in natural assemblages are controlled by a variety of events. The best correction (r2 = 0.74) was obtained between the [3H]thymidine incorporation and frequency of dividing cells procedures from the lake water samples. The actual production rates calculated by [3H]thymidine incorporation into DNA were appreciably lower than those obtained by the frequency of dividing cells and the dark CO2 uptake techniques. Diel rates of bacterial production are discussed in relation to sampling frequency, statistical errors, and choice of method.  相似文献   

12.
The characteristics of phosphate uptake and photosynthetic capacity were studied in P-limited populations of Euglena gracilis Klebs (Z), using both P-limited batch cultures in stationary phase and cyclostat cultures grown on 14:10 LD. P uptake obeyed Michaelis-Menten kinetics between 0 and 150 μM PO4 under both growth conditions. The value of Vmax was 35% lower in the dark than in the light in the stationary phase cells. The value of K8 was not affected by light conditions, and uptake was completely inhibited in the presence of 1 mm KCN. P uptake (at 2.0 μM PO4) and photosynthetic capacity showed diel periodicity with peak rates occurring just before the beginning of the dark period for P uptake, and 8 h into the light period for photosynthetic capacity. Vmax for P uptake increased by a factor of 1.5 over the light period, whereas K8 remained constant at 1.4 μM PO4. These patterns were displayed by both nondividing stationary phase cells and populations in which less than a third of the cells divided each day, indicating that the rhythmicity is not coupled to cell division.  相似文献   

13.
Heterotrophic nanoflagellates are ubiquitous and known to be major predators of bacteria. The feeding of free-living heterotrophic nanoflagellates on phytoplankton is poorly understood, although these two components usually co-exist. To investigate the feeding and ecological roles of major heterotrophic nanoflagellates Katablepharis spp., the feeding ability of Katablepharis japonica on bacteria and phytoplankton species and the type of the prey that K. japonica can feed on were explored. Furthermore, the growth and ingestion rates of K. japonica on the dinoflagellate Akashiwo sanguinea—a suitable algal prey item—heterotrophic bacteria, and the cyanobacteria Synechococcus sp., as a function of prey concentration were determined. Among the prey tested, K. japonica ingested heterotrophic bacteria, Synechococcus sp., the prasinophyte Pyramimonas sp., the cryptophytes Rhodomonas salina and Teleaulax sp., the raphidophytes Heterosigma akashiwo and Chattonella ovata, the dinoflagellates Heterocapsa rotundata, Amphidinium carterae, Prorocentrum donghaiense, Alexandrium minutum, Cochlodinium polykrikoides, Gymnodinium catenatum, A. sanguinea, Coolia malayensis, and the ciliate Mesodinium rubrum, however, it did not feed on the dinoflagellates Alexandrium catenella, Gambierdiscus caribaeus, Heterocapsa triquetra, Lingulodinium polyedra, Prorocentrum cordatum, P. micans, and Scrippsiella acuminata and the diatom Skeletonema costatum. Many K. japonica cells attacked and ingested a prey cell together after pecking and rupturing the surface of the prey cell and then uptaking the materials that emerged from the ruptured cell surface. Cells of A. sanguinea supported positive growth of K. japonica, but neither heterotrophic bacteria nor Synechococcus sp. supported growth. The maximum specific growth rate of K. japonica on A. sanguinea was 1.01 d−1. In addition, the maximum ingestion rate of K. japonica for A. sanguinea was 0.13 ng C predator−1d−1 (0.06 cells predator−1d−1). The maximum ingestion rate of K. japonica for heterotrophic bacteria was 0.019 ng C predator−1d−1 (266 bacteria predator−1d−1), and the highest ingestion rate of K. japonica for Synechococcus sp. at the given prey concentrations of up to ca. 107 cells ml−1 was 0.01 ng C predator−1d−1 (48 Synechococcus predator−1d−1). The maximum daily carbon acquisition from A. sanguinea, heterotrophic bacteria, and Synechococcus sp. were 307, 43, and 22%, respectively, of the body carbon of the predator. Thus, low ingestion rates of K. japonica on heterotrophic bacteria and Synechococcus sp. may be responsible for the lack of growth. The results of the present study clearly show that K. japonica is a predator of diverse phytoplankton, including toxic or harmful algae, and may also affect the dynamics of red tides caused by these prey species.  相似文献   

14.
Estimates of bacterial production based on total trichloroacetic acid (TCA)-precipitable [methyl-3H]thymidine incorporation and frequency of dividing cell (FDC) techniques were compared to sediment respiration rates in Lake George, New York. Bacterial growth rates based on thymidine incorporation ranged from 0.024 to 0.41 day–1, while rates based on FDC ranged from 1.78 to 2.48 day–1. Respiration rates ranged from 0.11 to 1.8mol O2·hour–1·g dry weight sediment–1. Thymidine incorporation yielded production estimates which were in reasonable agreement with respiration rates. Production estimates based on FDC were 4- to 190-fold higher than those predicted from respiration rates.  相似文献   

15.
The productivity of the final larval and adult populations of a large sized diplopod, Japonaria laminata armigera were studied at a warm temperate ever-green broad leaf forest in Chiba Japan. The population density was about 200/m2 in May 1962, and 7/m2 in May 1966, the former being about 30 times as much as the latter. A remarkable growth of animals was seen from June to November, and the growth pattern of individuals in both populations was very similar. The growth and mortality were calculated by the following formulae, where G is growth, M mortality, Nt number at the time t,Wt−1 mean body weight of animals at the time t−1, ΔN decrease in number during the period of time from t−1 to t, ΔW growth in mean body weight during the time from t−1 to t. The constant b in the formula R=aWb representing the relation between body weight and oxygen consumption of Japonaria laminata was determined at 0.81, and Q10 was estimated at about 2. The amounts of growth and assimilation were calculated. The pattern of dynamics in terms of productivity of the population in the two periods were very similar. But, the amounts of assimilation were computed at 57.8 kcal/m2/from May 1962 to April 1963, and 3.01 kcal from May 1966 to April 1967. The former is 20 times as much as the latter. The growth or ecological efficiencies for the populations were similar.  相似文献   

16.
Wolinella succinogenes was found to grow on H2S plus fumarate with the formation of elemental sulfur and succinate. The growth rate was 0.18 h-1 (t d=3.8 h) and the growth yield was estimated to be 6.0 g per mol fumarate used. Growth also occurred on formate plus elemental sulfur; the products formed were H2S and CO2. The growth rate and estimated growth yield were 0.58 h-1 (t d=1.2 h) and 3.5 g per mol formate used, respectively. These results suggest that certain chemotrophic anaerobes may be involved in both the formation and reduction of sulfur.  相似文献   

17.
The cyanobacteria Synechococcus and Prochlorococcus are important primary producers in marine ecosystems. Because currently available approaches for estimating microbial growth rates can be difficult to apply in the field, we have been exploring the feasibility of using quantitative rRNA measurements as the basis for making such estimates. In this study we examined the relationship between rRNA and growth rate in several Synechococcus and Prochlorococcus strains over a range of light‐regulated growth rates. Whole‐cell hybridization with fluorescently labeled peptide nucleic acid (PNA) probes was used in conjunction with flow cytometry to quantify rRNA on a per cell basis. This PNA probing technique allowed rRNA analysis in a phycoerythrin‐containing Synechococcus strain (WH7803) and in a non–phycoerythrin‐containing strain and in Prochlorococcus. All the strains showed a qualitatively similar tri‐phasic relationship between rRNA·cell?1 and growth rate, involving relatively little change in rRNA·cell?1 at low growth rates, linear increase at intermediate growth rates, and a plateau and/or decrease at the highest growth rates. The onset of each phase was associated with the relative, rather than absolute, growth rate of each strain. In the Synechococcus strains, rRNA normalized to flow cytometrically measured forward angle light scatter (an indicator of size) was well‐correlated with growth rate across strains. These findings support the idea that cellular rRNA may be useful as an indicator of in situ growth rate in natural Synechococcus and Prochlorococcus populations.  相似文献   

18.
Gyrodinium corsicum Paulmier produced massive blooms in Alfacs Bay (Ebro Delta, NW Mediterranean Sea) from December 1994 to April 1995, a period characterized by an initial period of water stability and low outward flux of water during subsequent months. The highest cell densities of G.corsicum were found at the bottom of the bay during the population maintenance period (from January to March). During the day, no differences in the vertical distribution of G.corsicum were observed and no diel in situ vertical migration of the population was recorded. The in situ growth rate of G.corsicum was estimated using the cell cycle method since the cell division cycle was well phased with the daily light cycle. The S phase fraction of G.corsicum reached a maximum near the middle of the dark period. The G2M peak usually occurred in the dark period. The duration of the S phase was 2.1-8.7 h and the duration of the G2M phase was 2.6-7.7 h. The estimated specific in situ growth rate of G.corsicum ranged from 0.94 day-1 during the initial phases of the bloom to 0.3 day-1 when the bloom was well developed. Growth rates, measured in different locations at the same time, varied by a factor of 1.5, suggesting that different parts of the same blooming population were at different stages of growth. The persistence of the G.corsicum bloom with high cell densities (105-106 cells l-1) at the bottom of the bay is discussed in relation to the low cell losses by physical dispersion in this area and the lack of upward vertical migration of G.corsicum.   相似文献   

19.
Kelp forests provide vital ecosystem services such as carbon storage and cycling, and understanding primary production dynamics regarding seasonal and spatial variations is essential. We conducted surveys at three sites in southeast Tasmania, Australia, that had different levels of water motion, across four seasons to determine seasonal primary production and carbon storage as living biomass for kelp beds of Lessonia corrugata (Order Laminariales). We quantified blade growth, erosion rates, and the variation in population density and estimated both the net biomass accumulation (NBA) per square meter and the carbon standing stock. We observed a significant difference in blade growth and erosion rates between seasons and sites. Spring had the highest growth rate (0.02 g C · blade−1 · d−1) and NBA (1.62 g C · m−2 · d−1), while summer had the highest blade erosion (0.01 g C · blade−1 · d−1), with a negative NBA (−1.18 g C · m−2 · d−1). Sites exhibiting lower blade erosion rates demonstrated notably greater NBA than sites with elevated erosion rates. The sites with the highest water motion had the slowest erosion rates. Moreover, the most wave-exposed site had the densest populations, resulting in the highest NBA and a greater standing stock. Our results reveal a strong seasonal and water motion influence on carbon dynamics in L. corrugata populations. This knowledge is important for understanding the dynamics of the carbon cycle in coastal regions.  相似文献   

20.
Cryptophytes are ubiquitous and one of the major phototrophic components in marine plankton communities. They often cause red tides in the waters of many countries. Understanding the bloom dynamics of cryptophytes is, therefore, of great importance. A critical step in this understanding is unveiling their trophic modes. Prior to this study, several freshwater cryptophyte species and marine Cryptomonas sp. and Geminifera cryophila were revealed to be mixotrophic. The trophic mode of the common marine cryptophyte species, Teleaulax amphioxeia has not been investigated yet. Thus, to explore the mixotrophic ability of T. amphioxeia by assessing the types of prey species that this species is able to feed on, the protoplasms of T. amphioxeia cells were carefully examined under an epifluorescence microscope and a transmission electron microscope after adding each of the diverse prey species. Furthermore, T. amphioxeia ingestion rates heterotrophic bacteria and the cyanobacterium Synechococcus sp. were measured as a function of prey concentration. Moreover, the feeding of natural populations of cryptophytes on natural populations of heterotrophic bacteria was assessed in Masan Bay in April 2006. This study reported for the first time, to our knowledge, that T. amphioxeia is a mixotrophic species. Among the prey organisms offered, T. amphioxeia fed only on heterotrophic bacteria and Synechococcus sp. The ingestion rates of T. amphioxeia on heterotrophic bacteria or Synechococcus sp. rapidly increased with increasing prey concentrations up to 8.6 × 106 cells ml−1, but slowly at higher prey concentrations. The maximum ingestion rates of T. amphioxeia on heterotrophic bacteria and Synechococcus sp. reached 0.7 and 0.3 cells predator−1 h−1, respectively. During the field experiments, the ingestion rates and grazing coefficients of cryptophytes on natural populations of heterotrophic bacteria were 0.3–8.3 cells predator−1 h−1 and 0.012–0.033 d−1, respectively. Marine cryptophytes, including T. amphioxeia, are known to be favorite prey species for many mixotrophic and heterotrophic dinoflagellates and ciliates. Cryptophytes, therefore, likely play important roles in marine food webs and may exert a considerable potential grazing impact on the populations of marine bacteria.  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