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 62 毫秒
1.
Assimilatory nitrate reductase (NR) was solubilized by acetonetreatment from Plectonema boryanum and was purified 7,700-foldby heat treatment, ammonium sulfate fractionation and chromatographyon DEAE-Sephacel and Sephadex G-150. Purified NR had a specificactivity of 85 µmol NO2 formed min–1 mg–1protein. The enzyme retained both ferredoxin (Fd)- and methylviologen (MV)-linked NR activities throughout the purificationprocedure. Molecular weight was 80,000. The pH optimum was 10.5in the MV-assay and 8.5 when assayed with enzymatically reducedFd as the electron donor. Apparent Km values for nitrate andMV were 700 µM and 2,500µM in the MVassay and 55µM and 75 µM for nitrate and Fd in the Fd-assay.The enzyme was inhibited by thiol reagents and metal-chelatingreagents. (Received October 1, 1982; Accepted March 8, 1983)  相似文献   

2.
During ice-free periods, seston ranges from 10 to 600 and averages29.2 mg dry wt 1–1 in Lake Balaton, the largest shallowlake in Central Europe. Most (80–99%) of the seston consistsof 1–10 µm sized mineral particles. The abiosestoncauses permanent food limitation in the ingestion of the ediblephytoplankton (1–22 µm fraction) by Daphnia cucullataand D.galeata. The postembryonic development time of D.galeataincreases and its fecundity decreases with increasing abioseston.The fecundity and the mortality of D.galeata are balanced atan abioseston concentration of 20.0 mg dry wt 1–1.  相似文献   

3.
During the ANTARES 3 cruise in the Indian sector of the SouthernOcean in October–November 1995, the surface waters ofKerguelen Islands plume, and the surface and deeper waters (30–60m) along a transect on 62°E from 48°36'S to the iceedge (58°50'S), were sampled. The phytoplankton communitywas size-fractionated (2 µm) and cell numbers, chlorophyllbiomass and carbon assimilation, through Rubisco and ß-carboxylaseactivities, were characterized. The highest contribution of<2 µm cells to total biomass and total Rubisco activitywas reported in the waters of the Permanent Open Ocean Zone(POOZ) located between 52°S and 55°S along 62°E.In this zone, the picophytoplankton contributed from 26 to 50%of the total chlorophyll (a + b + c) with an average of 0.09± 0.02 µg Chl l–1 for <2 µm cells.Picophytoplankton also contributed 36 to 64% of the total Rubiscoactivity, with an average of 0.80 ± 0.30 mg C mg Chla–1 h–1 for <2 µm cells. The picophytoplanktoncells had a higher ß-carboxylase activity than largercells >2 µm. The mixotrophic capacity of these smallcells is proposed. From sampling stations of the Kerguelen plume,a relationship was observed between the Rubisco activity perpicophytoplankton cell and apparent cell size, which variedwith the sampled water masses. Moreover, a depth-dependent photoperiodicityof Rubisco activity per cell for <2 µm phytoplanktonwas observed during the day/night cycle in the POOZ. In thenear ice zone, a physiological change in picophytoplankton cellsfavouring phosphoenolpyruvate carboxykinase (PEPCK) activitywas reported. A species succession, or an adaptation to unfavourableenvironmental conditions such as low temperature and/or availableirradiance levels, may have provoked this change. The high contributionof picophytoplankton to the total biomass, and its high CO2fixation capacity via autotrophy and mixotrophy, emphasize thestrong regeneration of organic materials in the euphotic layerin the Southern Ocean.  相似文献   

4.
Size-fractionated chlorophyll a biomass and picophytoplanktoncell number distributions were investigated along a longitudinalaxis of Southampton Water estuary during autumn. Chlorophylla concentration in the >5µm and the 1–5 µmsize fractions was highest midway down the estuary, and decreasedboth in the landward and seaward directions. In contrast, chlorophylla biomass in the 0.2–1 µm size fraction showed nodecline towards the seaward end of the estuary. In agreementwith this observation, phycoerythrin-containing picocyanobacteriacell concentration showed a positive exponential-like relationshipwith salinity and eukaryotic picophytoplankton were also highestat high salinities. Expressed as a percentage of total, chlorophylla standing stock in both the 1–5 µ.m (4.4–28.7%)and the 0.2–1 µm size fractions (1.7–8.6%)was inversely correlated with total chlorophyll a concentration.Both these two fractions made a greater input to the total phaeopigmentconcentration than to the total pool of active chlorophyll a.  相似文献   

5.
Seventeen size-fractionation experiments were carried out duringthe summer of 1979 to compare biomass and productivity in the< 10, <8 and <5 µm size fractions with that ofthe total phytoplankton community in surface waters of NarragansettBay. Flagellates and non-motile ultra-plankton passing 8 µmpolycarbonate filters dominated early summer phytoplankton populations,while diatoms and dinoflagellates retained by 10 µm nylonnetting dominated during the late summer. A significant numberof small diatoms and dinoflagellates were found in the 10–8µm size fraction. The > 10 µm size fraction accountedfor 50% of the chlorophyll a standing crop and 38% of surfaceproduction. The <8 µm fraction accounted for 39 and18% of the surface biomass and production. Production by the< 8 µm fraction exceeded half of the total communityproduction only during a mid-summer bloom of microflagellates.Mean assimilation numbers and calculated carbon doubling ratesin the <8 µm (2.8 g C g Chl a–1 h–1; 0.9day–1)and<5 µm(1.7 g C g Chl a–1h–1; 0.5day–1)size fractions were consistently lower than those of the totalpopulation (4.8 g C g Chl a–1 h–1; 1.3 day–1)and the <10 µm size fraction (5.8 g C g Chl a–1h–1; 1.4 day –1). The results indicate that smalldiatoms and dinoflagellates in fractionated phytoplankton populationscan influence productivity out of proportion to their numbersor biomass. 1Present address: Australian Institute of Marine Science, P.M.B.No. 3, Townsville M.S.O., Qld. 4810, Australia.  相似文献   

6.
Yield stress threshold (Y) and volumetric extensibility () arethe rheological properties that appear to control root growth.In this study they were measured in wheat roots by means ofparallel measurement of the growth rate (r) of intact wheatroots and of the turgor pressures (P) of individual cells withinthe expansion zone. Growth and turgor pressure were manipulatedby immersion in graded osmoticum (mannitol) solutions. Turgorwas measured with a pressure probe and growth rate by visualobservation. The influence of various growth conditions on Yand was investigated; (a) At 27 °C.In 0.5 mol m–3 CaCl2 r, P, Y and were20.7±4.6 µm min–1, 0.77±0.05 MPa,0.07±0.03 MPa and 26±1.9 µm min–1MPa–1 (expressed as increase in length), respectively.Following 24 h growth in 10 mol m–3 KC1 these parametersbecame 12.3±3.5 µm min–1, 0.72±0.04MPa, 0.13±0.01 MPa and 21±0.7 µm min–1MPa–1. After 24 h osmotic adjustment in 150 mol m–3mannitol/0.5 mol m–3 CaCl2 r= 19.6±4.2 µmmin–1, P = 0.68±0.05 MPa and Y and were 0.07±0.04MPa and 30±0.2 µm min–1 MPa–01, respectively.After 24 h growth in 350 mol m–3 mannitol/0.5 mol m–3CaCl2 r= 13.3±4.1 µm min–1, P= 0.58±0.07MPa, Y=0.12±0.01 MPa and ø 32±0.2 tim min–1MPa–1. During osmotic adjustment in 200 mol m–3mannitol/0.5 mol m–3 CaCl2, with or without KCl, the recoveryof growth rate corresponded to turgor pressure recovery (t1/2approximately 3 h). (b) At 15 °C. Lowered temperature dramatically influencedthe growth parameters which became r= 8.3±2.8 um min–1,P=0.78 MPa, r=<0.2 MPa and =15±0.1 µm min–1MPa–1. Therefore, Y and are influenced by 10 mol m–3 K+ ionsand low temperature. In each case the effective pressure forgrowth (P-Y) was large indicating that small fluctuations ofsoil water potential will not stop root elongation. Key words: Yield threshold, cell wall extensibility, wheat root growth, temperature, turgor pressur  相似文献   

7.
Grazing by microzooplankton on autotrophic and heterotrophicpicoplankton as well as >0.7 µm phytoplankton (as measuredby chlorophyll a) was quantified during July, August, October,January and April in the surface layer of Logy Bay, Newfoundland(47°38'14'N, 52°39'36'W). Rates of growth and grazingmortality of bacteria, Synechococcus and >0.7 µm phytoplanktonwere measured using the sea water dilution technique. Microzooplanktoningested 83–184, 96–366 and 64–118% of bacterial,Synechococcus and >0.7 µm phytoplankton daily potentialproduction, respectively and 34–111, 25–30 and 16–131%of bacterial, Synechococcus and >0.7 µm phytoplanktonstanding stocks, respectively. The trends in prey net growthrates followed the seasonal cycles of prey biomass, suggestingthat microzooplankton are important grazers in Newfoundlandcoastal waters. Ingestion was lowest during January and October(~2 µg C l–1 day–1) and highest in August(~20 µg C l–1 day–1). Aside from April when>0.7 µm phytoplankton represented the majority (~80%)of carbon ingested, bacterioplankton and <1 µm phytoplanktonrepresented most of the carbon ingested (~40–100%). Althoughmicrozooplankton have here-to-fore been unrecognized as an importantgrazer population in Newfoundland coastal waters, these resultssuggest that they play an important role in carbon flow withinthe pelagic food web, even at low temperatures in Logy Bay.  相似文献   

8.
The nutritional value of different sized fractions of naturalplankton was investigated for the growth of Eodiaptomus japonicusBurckhardt by comparing the development of its naupliar andcopepodid stages fed on differentially fractionated planktonicassemblages of a eutrophic pond, at 20°C. Water filteredthrough a 0.8 µm Nuclepore filter, containing mainly smallcoccoid bacteria (0.45–0.6 µm in cell diameter),at a concentration of 82.7 µg C 1–1 could not supportthe development of E.japonicus. The 3 µm filtered water,containing bacteria and picoalgae. at a total concentrationof 259 µg C 1–1, supported development but not eggproduction. The 20 µm filtered water, containing bacteria,picoalgae and large algae, at a total concentration of 2600µg C 1–1, supported rapid development of the juvenilesand continuous egg production by the adults. The separated 3–20µm fraction, containing only large algae, could not supportthe development at concentrations of 131 and 196 µg C1–1. However, the same rapid development of the juvenilesand continuous egg production by adults occurred at all of thetested concentrations between 261 and 3920 µg C1–1of the large algae. The results suggest that E.japonicus favoursalgae larger than 3 µm during its complete lifespan, andthat the threshold food concentration for its development variesbetween 200 and 250 µg C 1–1.  相似文献   

9.
The photoregulation of carotenogenesis in Rhodotorula minutawas found to consist of tow phases, a temperature-independentphotochemical reaction (light process) and temperature-dependentbut light-independent biochemical reactions (dark process).These processes were separately examined by regulating the temperatureand were characterized as follows: 1) The quantity of carotenoid produced [C (µg g–1)]and the rate of carotenoid production [Vc (µg g–1hr–1)] in the dark process were regulated by the lightdose [D (erg cm–2)] to which cells were exposed in thelight process. These relationships were expressed by the equations:C=9.1 log D–62.0 and Vc=0.81 log D–5.60. This photoresponsefollowed the Roscoe-Bunsen reciprocity law. 2) The induced state toward carotenogenesis, once acquired inthe light process, was very stable, suggesting that the proposedphotochemical product is stable as an inducer of carotenogenesisand decreases only in conjunction with carotenoid biosynthesis. 3) The photochemical reaction was oxygen-independent, but subsequentdark reactions were completely dependent on oxygen. 4) Postulated compounds related to the photochemical reactionwere not metabolized in vivo. (Received September 12, 1981; Accepted February 20, 1982)  相似文献   

10.
The temporal variability of size-fractioned autotrophic biomassat three depth levels (1, 8 and 25 m) was studied during thewinter-spring transition at two oceanographic stations in ConcepciónBay. Size spectra were obtained on eight occasions by two differentmethods: (i) determining the biomass of seven autotrophic sizefractions by in vivo fluorescence; and (ii) measuring the filamentlength of chain-forming diatoms through direct microscopy. Aclear vertical gradient of biomass was found in all profiles,with maximum values in the surface layer (1 and 8 m levels).Values of chlorophyll were on average 6.2 (range 1.08–25.67)times higher at 1 m than at 25 m, and 7.4 (range 1.15–26.83)times more at 8 m than at 25 m. On a temporal basis, total biomassincreased from low average values in winter (2.5 mg chl-a m–3)to high values in late spring (11.6 mg chl-a m–3). Duringthe whole sampling period (June 8-November 19), the nano- andnet-plankton (1.8–40 µm and 40–335 µmsize fractions respectively) were more abundant near the surface(1 and 8 m depth) than close to the bottom (25 m depth); however,the picoplankton fraction (<1.8 (µm) showed an inverserelationship, with a slight trend to increase near the bottomtoward spring. The highest absolute biomass was concentratedin the net-plankton fraction during the whole period and therelative importance of the picoplankton decreased from winter(6.50 and 15.5% for shallow and bottom levels) to spring (1.5and 10.3% for shallow and bottom levels). This relative effectis caused by the higher absolute values of biomass observedin the net-plankton fraction toward spring. These changing patternsshould have an impact in the size-composition and abundanceof higher trophic levels, mainly through grazing, in particularby modifying food availability to microfJagellates, ciliatesand filter-feeding zooplankton.  相似文献   

11.
Image analyses on the filtering apparatus of Bosmina longirostricshowed that the filter mesh is finer on the gnathobasic filterplates of the second and third trunk limbs (ranges from 0.43to 0.97 µm) and coarser for the outer ones of the thirdlimb (ranges from 0.5 to 1.36 µm), and the intersetulardistances increase with body length. Grazing experiments combinedwith image analysis confirmed the efficient grazing of B.longirostrison natural bacteria with cell lengths equal to or larger thanthe intersetular distances of the gnathobasic filter plates.During the experiments, the animals minimized the average celllength of the bacterioplankton assemblages from 0.77–0.96µm to 0.55–0.68 µm, corresponding to the meanof the filter mesh size on the fine gnathobasic filter platesof the experimental populations. The clearance rates for large,elongated or dividing cells with maximal lengths of 0.88–8.40µm were 2.3–17.7 times higher than those for smallsingle coccoids with a diameter of <0.45 µm. The resultsprovide evidence of a significant differential impact of B.longirostrison the bacterial community structure with respect to the shapeand size of the cells, and demonstrate that the species is amore effective bacterial feeder than considered previously.  相似文献   

12.
Measurements of hydrography, chlorophyll, moulting rates ofjuvenile copepods and egg production rates of adult female copepodswere made at eight stations along a transect across the Skagerrak.The goals of the study were to determine (i) if there were correlationsbetween spatial variations in hydrography, phytoplankton andcopepod production rates, (ii) if copepod egg production rateswere correlated with juvenile growth rates, and (iii) if therewas evidence of food-niche separation among co-occumng femalecopepods The 200 km wide Skagerrak had a stratified water columnin the center and a mixed water column along the margins. Suchspatial variations should lead to a dominance of small phytoplanktoncells in the center and large cells along the margins; however,during our study blooms of Gyrodinium aureolum and Ceratium(three species) masked any locally driven differences in cellsize: 50% of chla was >11 µm, 5% in the 11–50µm fraction and 45% <50 µm. averaged for allstations. Chlorophyll ranged from 0.2 to 2.5 µg l–1at most depths and stations. Specific growth rates of copepodsaveraged 0.10 day–1 for adult females and 0.27 day–1for juveniles The latter is similar to maximum rates known fromlaboratory studies, thus were probably not food-limited. Eggproduction rates were food-limited with the degree of limitationvarying among species: 75% of maximum for Centropages typicus, 50% for Calanus finmarchicus, 30% for Paracalanus parvus and 15% for Acartia longiremis and Temora longicornis. Thedegree of limitation was unrelated to female body size suggestingfood-niche separation among adults. Copepod production, summedover all species, ranged from 3 to 8 mg carbon m–3day–1and averaged 4.6 mg carbon m–1 day–1. Egg productionaccounted for 25% of the total.  相似文献   

13.
The relationships between photosynthesis and photosyntheticphoton flux densities (PPFD, P-l) were studied during a red-tideof Dinophysis norvegica (July-August 1990) in Bedford Basin.Dinophysis norvegica, together with other dinoflagellates suchas Gonyaulax digitate, Ceratium tripos, contributed {small tilde}50%of the phytoplankton biomass that attained a maximum of 16.7µg Chla 1 and 11.93 106 total cells I–1.The atomic ratios of carbon to nitrogen for D.norvegica rangedfrom 8.7 to 10.0. The photosynthetic characteristics of fractionatedphytoplankton (>30 µm) dominated by D.norvegica weresimilar to natural bloom assemblages: o (the initial slope ofthe P-l curves) ranged between 0.013 and 0.047 µg C [µgChla]–1 h–1 [µmol m s–1]–1the maximum photosynthetic rate, pBm, between 0.66 and 1.85µg C [µghla]–1 h–1; lk (the photoadaptationindex) from 14 to 69 µ,mol m–2 s–1. Carbonuptake rates of the isolated cells of D.norvegica (at 780 µmolm–2 s–1) ranged from 16 to 25 pg C cell–1h and were lower than those for C.tripos, G.digitaleand some other dinoflagellates. The variation in carbon uptakerates of isolated cells of D.norvegica corresponded with PBmof the red-tide phytoplankton assemblages in the P-l experiments.Our study showed that D.norvegica, a toxigenic dinoflagellate,was the main contributor to the primary production in the bloom.  相似文献   

14.
Ephyra larvae and small medusae (1.7–95 mm diameter, 0.01–350mg ash-free dry wt, AFDW) of the scyphozoan jellyfish Aureliaaurita were used in predation experiments with phytoplankton(the flagellate Isochrysis galbana, 4 µm diameter, {smalltilde}6 x 10–6 µg AFDW cell–1), ciliates (theoligotrich Strombidium sulcatum, 28 µm diameter, {smalltilde}2 x 10–3 µg AFDW), rotifers (Synchaeta sp.,0.5 µg AFDW individual–1) and mixed zooplankton(mainly copepods and cladocerans, 2.1–3.1 µg AFDWindividual–1). Phytoplankton in natural concentrations(50–200 µg C I–1) were not utilized by largemedusae (44–95 mm diameter). Ciliates in concentrationsfrom 0.5 to 50 individuals ml"1 were consumed by ephyra larvaeand small medusae (3–14 mm diameter) at a maximum predationrate of 171 prey day–1, corresponding to a daily rationof 0.42%. The rotifer Synchaeta sp., offered in concentrationsof 100–600 prey I–1, resulted in daily rations ofephyra larvae (2–5 mm diameter) between 1 and 13%. Mixedzooplankton allowed the highest daily rations, usually in therange 5–40%. Large medusae (>45 mm diameter) consumedbetween 2000 and 3500 prey organisms day"1 in prey concentrationsexceeding 100 I–1. Predation rate and daily ration werepositively correlated with prey abundance. Seen over a broadsize spectrum, the daily ration decreased with increased medusasize. The daily rations observed in high abundance of mixedzooplankton suggest a potential ‘scope for growth’that exceeds the growth rate observed in field populations,and this, in turn, suggests that the natural populations areusually food limited. The predicted predation rate at averageprey concentrations that are characteristic of neritic environmentscannot explain the maximum growth rates observed in field populations.It is therefore suggested that exploitation of patches of preyin high abundance is an important component in the trophodynamicsof this species. 1Present address: University of Bergen, Department of MarineBiology, N-5065 Blomsterdalen, Norway  相似文献   

15.
Using well plates of Phaeocystis pouchetii colonies isolatedfrom experimental mesocosms in western Norway, increases incolony size and division were documented. Median longest lineardimensions increased 0–7 µm h–1; literaturePhaeocystis globosa values are 0.9–4.7 µm h–1.Ten to twelve percent of colonies divided at rates of 0.21–0.28divisions day–1. Daughter colonies were 100 µm smallerthan mother colonies. Colonies delayed 3.5–4.9 days tofirst division, compared with literature values of 4–5days for P. globosa. This study provides the first experimentalevidence for colony division of wild P. pouchetii.  相似文献   

16.
Ritchie, R. J. 1987. The permeability of ammonia, methylamineand ethylamine in the charophyte Chara corallina (C. australis).—J.exp. Bot. 38: 67–76 The permeabilities of the amines, ammonia (NH3), methylamine(CH3NH2) and ethylamine (CH3CH2NH2) in the giant-celled charophyteChara corallina (C. australis) R.Br. have been measured andcompared. The permeabilities were corrected for uptake fluxesof the amine cations. Based on net uptake rates, the permeabilityof ammonia was 6?4?0?93 µm s–1 (n = 38). The permeabilitiesof methylamine and ethylamine were measured in net and exchangeflux experiments. The permeabilities of methylamine were notsignificantly different in net and exchange experiments, norto that of ammonia (Pmethylamine = 6?0?0?49 µm s–1(n = 44)). In net flux experiments the apparent permeabilityof ethylamine was slightly greater than that of ammonia andmethylamine (Pethylamine, net = 8?4?1?2 µm s–1 (n= 40)) but the permeability of ethylamine based on exchangeflux data was significantly higher (Pethylamine, exchange =14?1?2 µm s–1 (n = 20)). Methylamine can be validlyused as an ammonium analogue in permeability studies in Chara. The plasmalemma of Chara has acid and alkaline bands; littlediffusion of uncharged amines would occur across the acid bands.The actual permeability of amines across the alkaline bandsis probably about twice the values quoted above on a whole cellbasis i.e. the permeability of ammonia across the permeablepart of the plasmalemma is probably about 12 µm s–1. Key words: Chara, permeability, ammonia, methylamine  相似文献   

17.
Rhodotorula minuta cells, which have only traces of carotenoidswhen grown in the dark, started carotenoid production with theonset of illumination and the amount increased almost linearlyuntil 70 hr then remained constant thereafter when incubationwas continued under illumination, with the number of cells continuingto increase. The rate of carotenoid production [Vc (µgg–1 hr–1)] depended on the intensity of light [I(ergcm–2 sec–1)], with the relationship of Vc=0.74 logI–1.46. The final carotenoid content [C(µg g–1)]of cells incubated under continuous light was also controlledby the light intensity [I], with the relationship of C=52 logI–81. Control of carotenoid production by light occursas a two-phase process consisting of a temperatureindependentphotochemical reaction and light-independent biochemical reactions. (Received September 12, 1981; Accepted February 20, 1982)  相似文献   

18.
Seedlings of Pharbitis nil, strain Kidachi, were grown undercontinuous light at 20°C in vessels containing 5,000-mlnutrient solution, 24 plants per vessel. NAA (0.005–0.5µM), GA3 (0.1–0.5 µM), kinetin (0.5–5µM), benzyladenine (0.05–5 µM) or abscisicacid (4 µM) added to the nutrient solution induced long-dayflowering, and the flowering was always accompanied by suppressionof root elongation. 3,4-Dichlorobenzoic acid (0.05–10µM) and some other benzoic acid derivatives which arehighly effective for the induction of flowering in Lemna paucicostataalso showed similar effects. Neither NAA, kinetin nor 3,4-dichlorobenzoicacid applied via the apical part of the hypocotyl could causeflowering or suppression of root elongation. Thus, the flower-inducingeffect of the above substances was presumed to be secondaryto the suppression of root elongation. Ethrel (1–50 µM)added to the nutrient solution suppressed root elongation, butdid not induce flowering probably because it has flower-inhibitingactivity. 1 This paper is dedicated to the memory of Dr. Joji Ashida,the first president of the Japanese Society of Plant Physiologists. (Received December 15, 1982; Accepted February 25, 1983)  相似文献   

19.
The chlorophyll a content of nicroparticles which passed throughglass fiber filters Whatman type GF/F but were retained on 0.2µm Nuclepore membranes was analyzed on a weekly basisover the course of 1 year in Kaneohe Bay, Hawaii. Depth profileswere also obtained at four oceanic stations off the islandsof Maui and Molokai, Hawaii. Experimental evidence indicatedthat these microparticles were photosynthetically active. Theproportion of microparticulate chlorophyll a could be up to35% of picoplankton chlorophyll a (2.0–0.2 µm sizerange) retained on a single pass through a 0.2 p.m Nucleporefilter. The filtrate from both GFIF and 0.2µm Nucleporefilters was found to contain chlorophyll a which could be retainedon a subsequent pass through either 0.2 µm Nuclepore orGF/F filters. Only serial filtration can ensure that essentiallyall picoplankton have been filtered from the water when eitherof these types of filters is used.  相似文献   

20.
For decades frequent mass mortalities of Lesser Flamingos (Phoeniconaiasminor Geoffroy) have been observed at alkaline-saline KenyanRift Valley lakes. To estimate the potential influence of toxiccyanobacteria on these mass deaths, the phytoplankton communitieswere investigated in Lakes Bogoria, Nakuru and Elmenteita. Cyanobacterialtoxins were analyzed both in the phytoplankton from the threelakes and in isolated monocyanobacterial strains of Arthrospirafusiformis, Anabaenopsis abijatae, Spirulina subsalsa and Phormidiumterebriformis. Lake Bogoria was dominated by the cyanobacteriumA. fusiformis. In L. Nakuru and L. Elmenteita the phytoplanktonmainly consisted of A. fusiformis, A. abijatae and Anabaenopsisarnoldii, and in L. Nakuru an unknown Anabaena sp. was alsofound. Furthermore, this is the first time A. abijatae and theunknown Anabaena sp. have been found in Kenyan lakes. Phytoplanktonwet weight biomass was found to be high, reaching 777 mg L–1in L. Bogoria, 104 mg L–1 in L. Nakuru and 202 mg L–1in L. Elmenteita. Using HPLC, the cyanobacterial hepatotoxinsmicrocystin-LR, -RR -YR, -LF and -LA and the neurotoxin anatoxin-awere detected in phytoplankton samples from L. Bogoria and L.Nakuru. Total microcystin concentrations amounted to 155 µgmicrocystin-LR equivalents g–1 DW in L. Bogoria, and 4593µg microcystin-LR equivalents g–1 DW in L. Nakuru,with anatoxin-a concentrations at 9 µg g–1 DW inL. Bogoria and 223 µg g–1 DW in L. Nakuru. In L.Elmenteita phytoplankton, no cyanobacterial toxins were found.A. fusiformis was identified as one source of the toxins. Theisolated strain of A. fusiformis from L. Bogoria was found toproduce both microcystin-YR (15.0 µg g–1 DW) andanatoxin-a (10.4 µg g–1 DW), whilst the A. fusiformisstrain from L. Nakuru was found to produce anatoxin-a (0.14µg g–1 DW). Since A. fusiformis mass developmentsare characteristic of alkaline-saline lakes, health risks towildlife, especially the Arthrospira-consuming Lesser Flamingo,may be expected.  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