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 46 毫秒
1.
Sediment trap samples collected over a seven-year period (February 1991–October 1997) from Guaymas Basin in the Gulf of California were used to study the oxygen isotope composition of five species of planktonic foraminifera, Globigerinoides ruber (white), Globigerina bulloides, Neogloboquadrina dutertrei, Pulleniatina obliquiloculata, and Globorotalia menardii. The δ18O data were analyzed for temporal and interspecies variability and were compared to local hydrography to evaluate the use of each species in reconstructing past oceanographic applications. The two surface dwelling species, G. ruber and G. bulloides displayed the lowest δ18O values (~ 0.0 to ? 5.0‰), while δ18O values for the thermocline dwelling N. dutertrei, P. obliquiloculata, and G. menardii were higher (~ 0.0 to ? 2.0‰). The δ18O of G. ruber most accurately records measured sea surface temperatures (SSTs) throughout the year. G. bulloides δ18O tracks SSTs during the winter–spring upwelling period but for the remainder of the year records slightly colder, subsurface temperatures. The difference between the δ18O of the surface dwelling species, G. ruber and G. bulloides, and that of the thermocline dwelling species, N. dutertrei, P. obliquiloculata, and G. menardii, was used to estimate the surface to thermocline temperature gradient. During the winter these δ18O differences are small (~ 0.50‰) reflecting a well-mixed water column. These interspecies δ18O differences increase during the summer (~ 1.90‰) in response to the strong thermal stratification that exists at this time of year.  相似文献   

2.
The impact of chronic air pollution such as increased CO2 and NOx emissions on forest ecosystems in the Athabasca oil sands region in Alberta, Canada, was investigated in Pinus banksiana (jack pine) and Populus tremuloides (trembling aspen, aspen) stands in two watersheds (NE7 and SM8) located at different distances from the main emission sources of oil sands mining and upgrading facilities, using δ13C, δ15N, and Ca/Al of soil and tree ring samples as indicators. Watershed NE7 was exposed to greater amounts of acid deposition due to its closeness to the mining and upgrading area. The δ15N in the forest floor was lower (p < 0.05) in NE7 (ranged from −1.42 to −0.87‰) than in SM8 (−0.54 to 1.43‰), implying a greater amount of recent deposition of 15N-depleted N in NE7. Tree ring δ13C gradually decreased over time for both tree species/watersheds, indicating the influence of 13C-depleted CO2 emitted from industrial sources. Tree ring N concentration and δ15N were not different between watersheds and did not significantly change with time. Interestingly, however, the difference between watersheds (NE7–SM8) that is expressed as Diff_N (for N) increased with concomitant decreases in Diff_δ15N over time, implying greater increases in 15N-depleted N input in NE7 than in SM8. Such trends were stronger in aspen stands (R2 = 0.64 and p < 0.001 for Diff_N and R2 = 0.44 and p < 0.01 for Diff_δ15N between 1964 and 2009) than in jack pine stands. We conclude that δ15N in the forest floor and differences in N and δ15N of tree rings between watersheds are useful indicators reflecting the impact of spatial variations of air pollution on forest stands in the Athabasca oil sands region in western Canada.  相似文献   

3.
Prymnesium parvum produces a variety of toxic compounds, which affect other algae, grazers and organisms at higher trophic levels. Here we provide the method for development of a sensitive algal bioassay using a microalgal target, Teleaulax acuta, to measure strain variability in P. parvum toxicity, as well as the temporal stability of both the intracellular and the extracellular lytic compounds of P. parvum. We show high strain variation in toxicities after 3 h incubation with LC50s ranging from 24 to 223 × 103 cells ml−1. Most importantly we prove the necessity of testing physico-chemical properties of P. parvum toxins before attempting to isolate and characterize them. The extracellular toxin in the supernatant is highly unstable, and it loses significant lytic effects after 3 days despite storage at −20 °C and after only 24 h stored at 4 °C. However, when stored at −80 °C, lytic activity is more easily maintained. Reducing oxidation by storing the supernatant with no headspace in the vials significantly slowed loss of activity when stored at 4 °C. We show that the lytic activity of the intracellular toxins, when released by sonication, is not as high as the extracellular toxins, however the stability of the intracellular toxins when kept as a cell pellet at −20 °C is excellent, which proves this is a sufficient storage method for less than 3 months. Our results provide an ecologically appropriate algal bioassay to quantify lytic activity of P. parvum toxins and we have advanced our knowledge of how to handle and store the toxins from P. parvum so as to maintain biologically relevant toxicity.  相似文献   

4.
We investigated the effects of aeration on growth and toxicity of the haptophyte Prymnesium parvum in the presence and absence of the algal prey Rhodomonas salina. Batch monocultures of P-limited P. parvum and N and P sufficient R. salina and mixed cultures of the two microalgae were grown with no, low (20) and high (100) ml min−1 aeration for 18 days. Cell growth of P. parvum and R. salina and cell toxicity of P. parvum were studied over the experimental period. The highest specific growth rates of P. parvum were found at low aeration rates. R. salina in monocultures showed typical growth patterns, while R. salina numbers declined rapidly in the mixed cultures. Of the initial cell densities, 98–100% of the R. salina cells were lysed or ingested within 24 h of mixing with P. parvum cells. The maxima P. parvum biomasses were significantly higher in the mixed cultures than in the monocultures. Cell toxicity of P. parvum increased significantly in response to aeration rates and the highest levels were found in the high aeration condition. Availability of prey and resupply of inorganic nutrients decreased P. parvum cell toxicity. Our study suggests that P. parvum is tolerant and is able to grow over a broad range of aeration and associated turbulence effects though low aeration represents an optimal condition for growth. As P. parvum toxicity was higher in the high aeration treatment we suggest that the higher concentrations of oxygen cause more toxins to be produced, as these are oxygen rich compounds. We suggest that oxygenation and turbulence of surface waters caused by mixing may be involved in promoting high toxic P. parvum blooms in shallow lakes and coastal waters.  相似文献   

5.
Carbon and nitrogen are important elements in biogeochemical studies of tidal wetlands. Three wetland zones in Luoyuan Bay in the Fujian province were chosen for this study; the Spartina alterniflora flat zone with Spartina alterniflora growing, the silt zone with no Spartina alterniflora growing and the Spartina alterniflora-silt flat zone – a transition zone between the two. The spatial and seasonal variations of total organic carbon (TOC), total nitrogen (TN), stable isotopes of organic material (δ13C, δ15N), C/N ratio, average particle size and sediment composition in surface and vertical sediments of different ecological zones were analyzed. Carbon and nitrogen accumulation and particle size effects in the different ecological zones were discussed and the indicators of δ13C and C/N ratios were also compared. TOC, TN, δ13C contents, C/N ratios, and average particle size varied within the ranges of 0.611–1.133%, 0.053–0.090%, ?22.60 to ?18.92‰, 12.3–15.7, and 6.4–8.7 μm, respectively. Sediments were mainly silt-sized. Besides δ15N values, the other parameters, such as TOC, TN, δ13C contents, C/N ratios, and average particle size showed an obvious zonal distribution in surface sediments. The distribution of TOC and TN contents reflected the distribution of Spartina alterniflora within the bay. The profile and seasonal variations of these parameters in different ecological zones indicated that variations in the Spartina alterniflora flat and transition zones were complex because of the effect of Spartina alterniflora. Vertical and seasonal variations were sampled in the silt flat area. The profile and seasonal variations of TOC, TN and δ13C were similar in the transition zone and the Spartina alterniflora flat zone. Seasonal concentrations of TOC, TN and δ13C decreased from autumn > spring > winter > summer. The seasonal variation of carbon and nitrogen in the sediments may be influenced by temperature, particle size, plankton and benthos. The particle size effect was significant in the surface sediments and profile sediments of the transition zone. However, other factors had a greater effect on the distributions of TOC and TN in the Spartina alterniflora flat and silt flat zones. C/N ratios in sediments of the Spartina alterniflora flat, transition zone and silt flat were close to or > 12, indicating that the organic material source was dominated by terrestrial inputs. However, δ13C values decreased from the Spartina alterniflora flat zone > transition zone > silt flat zone indicating that the organic material source was predominantly from marine inputs. Thus the indications from C/N ratios and δ13C were different. There was no clear relationship between C/N ratios and δ13C values and a better relationship between δ13C values and TOC concentrations suggested that δ13C values provided a better indication of the organic source. Limited amounts of organic material came from Spartina alterniflora. This study has provided basic data for researching biogeochemical processes of biogenic elements in tidal wetlands and vegetation restoration, and has also provided a reference for assessing and protecting the environment and ecological systems in wetlands.  相似文献   

6.
Estrogens were prohibited in the food producing animals by European Union (96/22/EC directive) and added to the Report on Carcinogens in United States since 2002. Due to very low concentration in serum or urine (~pg/mL), the method of control its abuse had not been fully developed.The endogenous estrogens were separated from urines of 18 adult men and women. The exogenous estrogens were chemical reference standards and over the counter preparations. Two patients of dysfunctional uterine bleeding (DUB) administered exogenous estradiol and the urines were collected for 72 h. The urinary estrogens were separated by high-performance liquid chromatography (HPLC) and confirmed. The exogenous and exogenous estrogens were analyzed by gas chromatography combustion isotope ratio mass spectrometry (GC–C–IRMS) to determine the 13C/12C ratio (δ13C‰).The δ13C‰ values of reference standard of E1, E2, and E3 were ?29.36 ± 0.72, ?27.98 ± 0.35, ?27.62 ± 0.51, respectively. The δ13C‰ values of the endogenous E1, E2, and E3 were ?21.62 ± 1.07, ?22.14 ± 0.98, and ?21.88 ± 1.16, with P < 0.01 (t-test). Two DUB patients’ urinary estradiol δ13C‰ values was depleted to ?28.02 ± 0.33 after the administration. The progesterone, 17α-hydroxyprogesterone, pregnanediol, as well as desogestrel and ethinylestradiol from contraceptives were also determined.Stable carbon isotope analysis can distinguish the endogenous and exogenous urinary estrogen in human.  相似文献   

7.
Stable isotopes (δ15N and δ13C) are being widely applied in ecological research but there has been a call for ecologists to determine species- and tissue-specific diet discrimination factors (?13C and ?15N) for their study animals. For large sharks stable isotopes may provide an important tool to elucidate aspects of their ecological roles in marine systems, but laboratory based controlled feeding experiments are impractical. By utilizing commercial aquaria, we estimated ?15N and ?13C of muscle, liver, vertebral cartilage and a number of organs of three large sand tiger (Carcharias taurus) and one large lemon shark (Negaprion brevirostris) under a controlled feeding regime. For all sharks mean ± SD for ?15N and ?13C in lipid extracted muscle using lipid extracted prey data were 2.29‰ ± 0.22 and 0.90‰ ± 0.33, respectively. The use of non-lipid extracted muscle and prey resulted in very similar ?15N and ?13C values but mixing of lipid and non-lipid extracted data produced variable estimates. Values of ?15N and ?13C in lipid extracted liver and prey were 1.50‰ ± 0.54 and 0.22‰ ± 1.18, respectively. Non-lipid extracted diet discrimination factors in liver were highly influenced by lipid content and studies that examine stable isotopes in shark liver, and likely any high lipid tissue, should strive to remove lipid effects through standardising C:N ratios, prior to isotope analysis. Mean vertebral cartilage ?15N and ?13C values were 1.45‰ ± 0.61 and 3.75‰ ± 0.44, respectively. Organ ?15N and ?13C values were more variable among individual sharks but heart tissue was consistently enriched by ~ 1–2.5‰. Minimal variability in muscle and liver δ15N and δ13C sampled at different intervals along the length of individual sharks and between liver lobes suggests that stable isotope values are consistent within tissues of individual animals. To our knowledge, these are the first reported diet–tissue discrimination factors for large sharks under semi-controlled conditions, and are lower than those reported for teleost fish.  相似文献   

8.
Stable isotopes (δ15N and δ13C) were used to assess the changes in exposure and assimilation of sewage-derived nutrients in an aquatic food web following changes in effluent quality over an 8 year period at two municipal wastewater treatment plants (WWTPs) that discharge to the Grand River, in southern Ontario. Upgrades at the Kitchener WWTP started in late 2012 to enhance nitrification, while the Waterloo WWTP had a series of construction issues at the plant that resulted in a deterioration of its effluent quality over the study period (2007–2014). Fish (rainbow darter, Etheostoma caeruleum) and primary consumers (benthic invertebrates) were sampled in the receiving waters associated with each outfall. Upgrades at the Kitchener WWTP resulted in improved effluent quality with total annual ammonia output dropping by nearly sixfold (583–100 t), while the Waterloo WWTP increased its total annual ammonia output by nearly fourfold (135–500 t) over the duration of the study. Downstream of the Kitchener WWTP, the reduction in total ammonia output negatively correlated with changes in δ15N of rainbow darter from being depleted (prior to the upgrade) to reflecting signatures similar to those at the upstream reference site. The biota downstream of the Waterloo WWTP showed the opposite trend, going from slightly enriched, to being depleted relative to the upstream reference sites. δ13C was consistently higher downstream of both WWTPs regardless of changing effluent quality, and annual variability in δ13C was associated with annual river discharge. In a laboratory based dietary switch study conducted with rainbow darter, the isotope half-life in muscle (29 days for δ15N and 33 days for δ13C) were determined and these rapid changes were consistent with responses in muscle of wild fish. This is a unique study that was able to contrast two WWTPs in the same watershed as they underwent major changes in treatment processes. Stable isotopes were very effective as a tool to trace the changes in aquatic biota due to changes in wastewater effluent quality, both improvements and deterioration over time.  相似文献   

9.
The unicellular alga Prymnesium parvum has been responsible for toxic incidents with severe ecological impacts in many parts of the world, and causes massive fish kills worldwide. Recently the haptophyte microalgae have caused water-bloom (4.3 × 104 cells ml−1) in 6 fish ponds with high conductivity in Hungary, and caused fish mortality with typical symptoms. Toxicity of P. parvum from water samples was quantified by the assay of the influence of its cell-free filtrates on haemolysis (346 ± 42.2) and in fish and daphnia toxicity tests. High amount of proteases in P. parvum containing waterbloom samples were detected with the help of activity gel electrophoresis. The proteases of investigated P. parvum samples (125–18 kDa) showed high gelatinolytic activity and some of them showed sensitivity to EDTA (inhibitors of metalloproteases) and to PMSF (inhibitors of serine proteases).  相似文献   

10.
Ostreopsis ovata is a benthic dinoflagellate that produces palytoxin and ovatoxins. Blooms of O. ovata causing human health problems and mortality of benthic fauna have been reported from many tropical and temperate marine waters. In the present study we examined the combined effects of temperature and different nutrient conditions on the biochemical composition, growth, toxicity and carbohydrate production of an O. ovata strain originating from the Tyrrhenian Sea. O. ovata cultures with N:P ratios of 1.6, 16 and 160 (N deficient, NP sufficient and P deficient, respectively) were grown at 20 °C and 30 °C. Biomass accumulation, growth rates, cell volumes, biochemical composition, cell toxicity and carbohydrate production in each treatment were studied. Results indicated that under nutrient sufficiency O. ovata biomass accumulation increased significantly compared to N and P deficiency and also that N limitation severely affected growth. The highest growth rates were recorded at 30 °C. Cellular contents and the atomic ratios of C, N and P were higher in the cells grown at 20 °C than in those grown at 30 °C. O. ovata cell volumes increased at 20 °C. N deficiency significantly increased cell toxicity. Toxicity per cell was higher at 20 °C, but per carbon was highest at 30 °C. The highest carbohydrate production was found in conditions of N deficiency and at the lower temperature.Our study suggests that temperature increases due to global warming and nutrient enrichment of coastal waters stimulate the proliferation of O. ovata, particularly for the strains that have become adapted to warm temperate waters.  相似文献   

11.
The intensification of agricultural land use and urbanisation has increased nutrient loads in aquatic ecosystems. Nitrogen loads can alter ecosystem structure and functioning, resulting in increased algal productivity, algal blooms and eutrophication. The principal aim of the present paper is to extend the use of epilithic δ15N signatures to a lake ecosystem in order to evaluate the potential impact of anthropogenic nitrogen discharges (organic and inorganic) that can also reach coastal waters.Epilithic associations were collected from volcanic rocks in different seasons in shallow water along the entire perimeter of Lake Bracciano and analysed for their nitrogen stable isotope signatures. Furthermore, some stones were moved from an unpolluted site to a polluted one in order to verify the effect on the nitrogen signature of the epilithic association. The epilithon’s δ15N signatures provided strong evidence of the space-time variability of N inputs. The differing quality of nitrogen loads was reflected in high isotopic variation within the lake, especially at the beginning of summer (1.7‰  δ15N  13.3‰), while in winter, when anthropogenic pressure was lowest, the δ15N signature variation was less accentuated (3.1‰  δ15N  7.6‰). At all sampling times, spatial variability was found to be related to the various human activities along the lake shore (especially tourism and agriculture), while seasonal variation at all sampling sites was related to the intensity of anthropogenic pressures (higher in summer and lower in winter).Our results showed that epilithic algal associations and the physicochemical properties of the water did not influence the δ15N signature, which in contrast was strongly related to the site-specific effect of human activities around the lake. Thus, the distribution of δ15N across space and time can be used to direct nutrient reduction strategies in the region and can assist in monitoring the effectiveness of environmental protection measures.  相似文献   

12.
The uptake rates of different nitrogen (N) forms (NO3, urea, and the amino acids glycine and glutamic acid) by N-deficient, laboratory-grown cells of the mixotrophic haptophyte, Prymnesium parvum, were measured and the preference by the cells for the different forms determined. Cellular N uptake rates (ρcell, fmol N cell−1 h−1) were measured using 15N-labeled N substrates. P. parvum showed high preference for the tested amino acids, in particular glutamic acid, over urea and NO3 under the culture nutrient conditions. However, extrapolating these rates to Baltic Seawater summer conditions, P. parvum would be expected to show higher uptake rates of NO3 and the amino acids relative to urea because of the difference in average concentrations of these substrates. A high uptake rate of glutamic acid at low substrate concentrations suggests that this substrate is likely used through extracellular enzymes. Nitrate, urea and glycine, on the other hand, showed a non-saturating uptake over the tested substrate concentration (1–40 μM-N for NO3 and urea, 0.5–10 μM-N for glycine), indicating slower membrane-transport rates for these substrates.  相似文献   

13.
The stages of Cryptosporidium parvum asexual exogenous development were investigated at high ultra-structural resolution in cell-free culture using transmission electron microscopy (TEM). Early C. parvum trophozoites were ovoid in shape, 1.07 × 1.47 μm2 in size, and contained a large nucleus and adjacent Golgi complex. Dividing and mature meronts containing four to eight developing merozoites, 2.34 × 2.7 μm2 in size, were observed within the first 24 h of cultivation. An obvious peculiarity was found within the merozoite pellicle, as it was composed of the outer plasma membrane with underlying middle and inner membrane complexes. Further novel findings were vacuolization of the meront's residuum and extension of its outer pellicle, as parasitophorous vacuole-like membranes were also evident. The asexual reproduction of C. parvum was consistent with the developmental pattern of both eimerian coccidia and Arthrogregarinida (formerly Neogregarinida). The unique cell-free development of C. parvum described here, along with the establishment of meronts and merozoite formation, is the first such evidence obtained from in vitro cell-free culture at the ultrastructural level.  相似文献   

14.
We explored the relationships between surface-soil (1–20 cm) organic carbon isotopic signatures and associated climatic factors in central-east Asia in an attempt to develop transfer functions that can be used to retrieve the paleoclimatic information stored in the thick eolian–paleosol sequences within the area. Our analysis shows that the negative correlation between the surface-soil organic δ13C values and the mean annual precipitation is robust (R2 = 0.453; n = 196; p < 0.05) and the negative correlation with the growing-season (April–September) precipitation is more significant (R2 = 0.4966; n = 196; p < 0.05). Our study further shows that the positive correlation between the surface-soil organic δ13C values and mean growing-season aridity is most significant (R2 = 0.5805; n = 196; p < 0.05). We have smoothed both the organic δ13C values and the mean growing-season aridity values using a 3-point moving-window average-filter method in an attempt to remove some of random errors and found that the positive correlation between the two is further increased (R2 =  0.7784; n =  192; p < 0.05). These robust linear relationships demonstrate their value in reconstructing paleoclimate changes in the study area. The documented climatic dependency of the surface-soil carbon isotopic composition in the study area might have resulted both from the humidity-related isotopic enrichment processes of the dominant C3 plants (stomatal conductance and photosynthetic discrimination) and from the aridity-related abundance of C4 plants (mainly Chenopodiaceae species) along the S–N bioclimatic gradient.  相似文献   

15.
《Marine Micropaleontology》2010,77(3-4):92-103
We determined the stable oxygen and carbon isotopic composition of live (Rose Bengal stained) tests belonging to different size classes of two benthic foraminiferal species from the Pakistan continental margin. Samples were taken at two sites, with water depths of about 135 and 275 m, corresponding to the upper boundary and upper part of the core region of the oxygen minimum zone (OMZ). For Uvigerina ex gr. Uvigerina semiornata and Bolivina aff. Bolivina dilatata, δ13C and δ18O values increased significantly with increasing test size. In the case of Uvigerina ex gr. U. semiornata, δ13C increased linearly by about 0.105‰ for each 100-μm increment in test size, whereas δ18O increased by 0.02 to 0.06‰ per 100 μm increment. For Bolivina aff. B. dilatata the relationship between test size and stable isotopic composition is better described by logarithmic equations. A strong positive linear correlation is observed between δ18O and δ13C values of both taxa, with a constant ratio of δ18O and δ13C values close to 2:1. This suggests that the strong ontogenetic effect is mainly caused by kinetic isotope fractionation during CO2 uptake. Our data underline the necessity to base longer δ18O and δ13C isotope records derived from benthic foraminifera on size windows of 100 μm or less. This is already common practice in down-core isotopic studies of planktonic foraminifera.  相似文献   

16.
The 13C/12C ratios of leaves of the conifer morphotype Frenelopsis were measured to decipher the influences of water and salt stress on stomatal density (SD), epidermal cell density (ECD) and stomatal index (SI). Three morphospecies were analyzed: F. ugnaensis from freshwater fluvio-lacustrine deposits (Upper Barremian), F. turolensis and alata from coastal deposits (Lower-Middle Albian and Upper Albian respectively). The cuticle δ13C values show a large variation from ? 28‰ to ? 21‰. Comparison with previously published marine carbonate δ13C records indicate that the difference in cuticle δ13C between the different deposits are mainly due to difference in CO2-plant isotope fractionation rather than to change in isotopic composition of inorganic carbon in the atmosphere and ocean. The less negative δ13C and wide range in δ13C of F. turolensis and alata (? 27.5 to ? 21‰), compared to F. ugnaensis, (? 28 to ? 25‰) are interpreted as a result of salt and/or water stress. The data as a whole yield a good relationship between the 13C/12C ratio and SD (r = 0.67, n = 42, p < 0.001), SI (r = 0.53, n = 41, p < 0.001), hence suggesting that the differences in SD and SI between the three morphospecies are related to freshwater/saline environment. Looking at single morphospecies, the SD of F. ugnaensis decreases with increasing δ13C value (r = ? 0.57, n = 15, p = 0.026) as well as a decrease of SI (r = ? 0.62, n = 15, p = 0.013), possibly reflecting warmer and drier conditions. Average SI of F. alata does not significantly change with δ13C and inferred soil salinity in contrast to SD (p < 0.01).  相似文献   

17.
Planktic and epibenthic foraminiferal δ13C records at Site PS62/015-3 (southwestern Greenland Sea) reveal a series of transient events of extreme 13C depletion down to − 6‰ during the period 90–40 ka. Scanning electron microscope studies of the ultrastructures of foraminiferal tests suggest that 13C depleted specimens are affected by some 10–20% overgrowth by authigenic calcite contributing to the light δ13C signal. Incremental-leaching experiments and census counts of pristine versus overgrowth-affected specimens show that the 13C depleted foraminiferal tests incorporate a primary δ13C signature most likely ranging from + 1‰ to − 1.7‰ and a post-depositional δ13C signature around − 17‰ to − 19.5‰. Extremely low values of productivity and organic carbon in Late Quaternary sediments along the east Greenland margin preclude organic matter as potential source of the isotopically light carbon. In contrast, thermal instability of clathrates and subsequent aerobic oxidation of (highly 12C enriched) methane in pore and ocean water provide a compelling mechanism to account for the negative δ13C excursions of both primary and post-depositional carbonates. Here, pore water methane may have led to a supersaturation of 13C depleted bicarbonate and precipitation of isotopically light authigenic calcite on and in foraminiferal tests, a feature that may serve as a tracer to former sites of clathrate destabilization.  相似文献   

18.
《Aquatic Botany》2007,87(4):292-298
The effect of salinity on leaf area and the relative accumulation of Na+ and K+ in leaves of the mangrove associate Hibiscus tiliaceus were investigated. Photosynthetic gas exchange characteristics were also examined under arid and non-arid leaf conditions at 0, 10, 20 and 30‰ substrate salinity. At salinities  40‰, plants showed complete defoliation followed by 100% mortality within 1 week. Salinities  30‰ were negatively correlated with the total leaf area per plant (r2 = 0.94). The reduction in the total plant leaf area is attributed to the reduction in the area of individual leaves (r2 = 0.94). Selective uptake of K+ over Na+ declined sharply with increasing salinity, where K+/Na+ ratio was reduced from 6.37 to 0.69 in plants treated with 0 and 30‰, respectively. Under non-arid leaf condition, increasing salinity from 0 to 30‰ has significantly reduced the values of the intrinsic components of photosynthesis Vc,max (from 50.4 to 18.4 μmol m−2 s-1), Jmax (from 118.0 to 33.8 μmol photons m−2 s−1), and VTPU (from 6.90 to 2.30 μmol m−2 s−1), while stomatal limitation to gas phase conductance (SL) increased from 14.6 to 38.4%. Water use efficiency (WUE) has subsequently doubled from 3.20 for the control plants to 8.93 for 30‰ treatment. Under arid leaf conditions, the stomatal factor (SL) was more limiting to photosynthesis than its biochemical components (73.4 to 26.6%, respectively, at 30‰). It is concluded that salinity causes a drastic decline in photosynthetic gas exchange in H. tiliaceus leaves through its intrinsic and stomatal components, and that the apparent phenotypic plasticity represented by the leaf area modulation is unlikely to be the mechanism by which H. tiliaceus avoids salt stress.  相似文献   

19.
The δ15N and δ13C signatures of major organic matter (OM) pools were measured across chemical and hydrologic gradients in a large (58,800 ha) subtropical wetland to evaluate whether stable isotopes were useful indicators of environmental change. Once a rainfall-driven wetland, the Loxahatchee National Wildlife Refuge in the Florida Everglades now receives agricultural and urban drainage that has increased phosphorus (P) and mineral loads around the wetland perimeter. Additionally, water impoundment at the southern end has produced a latitudinal hydrologic gradient, with extended hydroperiods in the south and overdrained conditions in the north.Detritus (?4.8‰ to 8.6‰), floc (?1.4‰ to 3.6‰), and metaphyton (?6.6‰ to +7.4‰) δ15N declined southward with changes in hydrology as indicated by water depth. This pattern was attributed to higher mineralization rates under shorter hydroperiods. These signatures were also strongly correlated with increased nutrient and mineral loading. Rooted macrophyte δ15N, by contrast, appeared more responsive to soil nutrient pools. Cattail (?8.9‰ to +7.7‰) was restricted to the wetland perimeter and had the widest δ15N range, which was positively correlated with soil P. Sawgrass (?5.3‰ to +7.7‰) occurred across most of the wetland, but its δ15N was not strongly correlated to any gradient. Patterns for δ13C were more strongly related to chemical gradients caused by canal intrusion than to latitude or hydrology. Again, metaphyton and detrital signatures were more sensitive to water chemistry changes than macrophytes. This pattern is consistent with their locations at the soil–water (detritus-floc), and air–water (metaphyton) interface. Metaphyton δ13C (?36.1‰ to ?21.5‰) which had the broadest range, was affected by DIC source and pool size. In contrast, cattail δ13C (?28.7‰ to ?26.4‰) was more closely related to soil P and sawgrass δ13C (?30.1‰ to ?24.5‰) was not related to any environmental gradient except latitude. There was no correlation between the two isotopes for any OM pool except cattail.These results indicate that isotopic signatures of microbial (metaphyton and detrital) pools are more responsive to changes in wetland hydrology and water chemistry while those of rooted macrophytes respond only to the extent that soil chemistry is altered. Rooted macrophytes also differ in the sensitivity of their isotopic signatures to environmental change. The selection of OM pools for isotopic analysis will, therefore, affect the sensitivity of the analysis and the resulting patterns. Furthermore, δ15N may be more robust and interpretable than δ13C as an indicator of ecosystem change in wetlands exposed to multiple or complex anthropogenic gradients.  相似文献   

20.
To characterize the ecology and physiology of common late Eocene–early Oligocene White River mammals, we analyzed the carbon and oxygen isotope composition of tooth enamel carbonate for six of the most abundant taxa: the perissodactyls Brontops (brontothere), Mesohippus (equid) and Subhyracodon (rhino); and the artiodactyls Merycoidodon (oreodont), Leptomeryx (leptomerycid) and Poebrotherium (camelid).δ13C values of middle and rear molars (M2s and M3s) and premolars range from ? 13.1‰ to ? 7.7‰ (V-PDB), consistent with pure C3 diets. In the late Eocene, Mesohippus, Merycoidodon, and Leptomeryx show indistinguishable average δ13C values (~ ? 10‰). In contrast, Brontops and Subhyracodon exhibit lower (? 11.2‰) and higher (? 8.7‰) values, respectively. Early Oligocene values for Mesohippus and Merycoidodon remain indistinguishable from each other and lower than the value of Leptomeryx, Poebrotherium, and Subhyracodon (~ ? 8.5‰). These results likely indicate niche separation in terms of habitat preference between the investigated sympatric taxa. More specifically, assuming a δ13C value of atmospheric CO2 of ? 5.5‰, our data suggest a preference of Brontops for mesic forested areas, of Mesohippus and Merycoidodon for woodlands, and of Subhyracodon and Poebrotherium for more open habitats (e.g., grasslands). The higher Oligocene versus Eocene average δ13C exhibited by Leptomeryx possibly reflects a preference of the new Oligocene species L. evansi for more open and/or xeric habitats relative to the Eocene species L. speciosus.Late Eocene and early Oligocene average δ18O of Mesohippus, Merycoidodon, Leptomeryx, and Subhyracodon are similar (~ 25‰, V-SMOW), possibly indicating comparable water dependency for these taxa. In contrast, the higher δ18O of Poebrotherium (26.6‰) suggests lower water dependency whereas the lower δ18O of Brontops (23.0‰) may result from a high water dependency or, more likely, from its preference for humid habitats. Because hind-gut fermentation in perissodactyls requires high water intake, whereas fore-gut fermentation does not, our results might indicate the presence of fore-gut fermentation in early Oligocene camelids but its absence or incomplete development in late Eocene–early Oligocene oreodonts and leptomerycids.  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