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.
Vertical distribution of organic constituents, i.e. total organic carbon (TOC), hydrocarbons, fatty acids and hydroxy acids in water and sediment samples from Lake Fryxell (77° 35 S, 163° 15 E) of southern Victoria Land, Antarctica were studied to elucidate their features in relation to stratification of the lake waters and likely distribution of microorganisms. The TOC content of the surface water (5.0 m; just below the ice cover of 4.50 m thickness) was 1.4 mg l–1. It increased markedly with depth and attained a maximum value of 21.7 mg C l–1 at a depth of 17.5 m, but decreased to the bottom (13.3 mg C l–1). The high TOC content of the anoxic bottom layers (> 15 m) is attributable to the concentration of refractory organic substances over long periods following the degradation of labile organic constituents. Hydrocarbons were not found in the water column, but the major constituent of the bottom sediment was n-C29 : 2 alkene. Total concentrations of fatty acids in the oxic layers ( 10 m) were highest at 10.0 m and much higher than those in the anoxic layers (> 10 m), probably reflecting the phytoplankton population. The content of branched (iso and anteiso) fatty acids and 3-hydroxy acids in the anoxic layers were much greater than those in the oxic layers which would seem to reflect the distribution of bacterial abundance. The differences of organic composition between the water column and sediments imply that sinking dead organisms were quickly degraded in the lake bottom. Also, the composition of microorganisms in the water column must be very different from that in the sediments.  相似文献   

2.
A laboratory experiment was conducted to determine the effect of tubificid worms on the flux of zinc into lake sediments. Forty-six cores of Lake Erie sediment, with and without (control) tubificid worm populations, were exposed to aquarium water with a zinc concentration of about 5 mg 1–1 for 139 days. Pore water and exchangeable particulate zinc concentrations in the top 12 cm of sediment were periodically determined in pairs of cores — one with worms and one without worms — at 1 cm depth increments. After 139 days, pore water zinc concentrations in sediments with and without worms were nearly identical in the 0–1 cm interval (4.1 and 4.3 mg 1–1 respectively), but were significantly greater in the sediments with worms in the 1–2 cm (4.4 vs. 0.3 mg1–1) and the 2–3 cm (1.3 vs. 0.3 mg 1–1) intervals. Exchangeable particulate zinc concentrations in the 0–1, 1–2, and 2–3 cm intervals in sediments with worms were 612.3, 750.7, and 191.5 µg g–1 dry sediment respectively, whereas in sediments without worms, concentrations were 375.4, 5.9, and 3.2 µg g–1 dry sediment. The increased flux of zinc into tubificid-inhabited sediments was caused by the conveyor belt feeding activity of the worms, which continuously exposed sedimentary particles to the overlying water. Movement of zinc into sediments with worms was dominated by adsorption and by particle movement, whereas movement of zinc into control sediments was by adsorption at the sediment-water interface and diffusion. The increased concentration of zinc in tubificid-inhabited sediments has important implications with respect to the trophic transfer of zinc through the aquatic food chain.  相似文献   

3.
The effects of bioventing, nutrient addition and inoculation with an oil-degrading bacterium on biodegradation of diesel oil in unsaturated soil were investigated. A mesocosm system was constructed consisting of six soil compartments each containing 6 m3 of naturally contaminated soil mixed 11 with silica sand, resulting in a diesel oil content of approximately 2000 mg kg–1. Biodegradation was monitored over 112 days by determining the actual diesel oil content of the soil and by respirometric tests. The best agreement between calculations of degradation rates based upon the two methods was in July, when venting in combination with nutrient addition resulted in degradation rates of 23 mg kg–1 day–1 based on actual oil concentration in the soil and 33 mg kg–1 day–1 calculated from respirometric data. In September, these rates decreased to 9 and 1.4 mg kg–1 day–1, and in October the degradation rates were 5 and 0.7 mg kg–1 day–1 based upon the two methods. The average ambient temperature during the respirometric tests was 14,10 and 2°C in July, September and October, respectively. The combination of venting and nutrient addition resulted in an average residual oil content of the soil of 380 mg kg–1. Neither venting alone nor inoculation enhanced oil degradation. The respiratory quotient averaged 0.40. The oil composition changed following degradation resulting in the unresolved complex mixture constituting up to 96% of the total oil content at the end of the experimental period.  相似文献   

4.
Arsenic transport between water and sediments   总被引:1,自引:1,他引:0  
Cornett  Jack  Chant  Lorna  Risto  Bert 《Hydrobiologia》1992,(1):533-544
Arsenic discharged into the Moira River has accumulated in the sediments of Moira Lake during the past century. The chronology of arsenic concentrations in the sediments, established using Pb-210 dating, has a subsurface concentration maximum (> 1000 g g–1) that reflects higher inputs to the lake 15 to 45 years ago. The distribution coefficient (Kd) of arsenic in the surficial sediments was low (4000–6000 L kg–1) and decreased below the sediment water interface. Higher concentrations of exchangeable As also were extracted deeper in the sediments. As a result, arsenic is mobile in the sediment column and the flux of arsenic via diffusion and particle resuspension from the sediments into the water is greater than current external loading from the Moira River. Less than 20% of the external input of arsenic is buried in the lake sediments. Using these flux measurements and a one dimensional model of arsenic transport in the sediment column, we constructed the history of arsenic exchange between water and sediments throughout the past century. The simulations predict that arsenic input into the water from the sediments has been > 20 % of external loading for the past 25 years and will continue to be important in the future as diffusion and resuspension regenerate arsenic from the mixed layer of the sediments into the overlying water.  相似文献   

5.
Frignani  M.  Bellucci  L. G.  Favotto  M.  Albertazzi  S. 《Hydrobiologia》2003,494(1-3):283-290
Three sediment cores were collected in the Venice Lagoon: two from mud flats (E, F) and one from the San Giuliano Canal (I1), which borders the industrial district. Samples were analysed for the 15 polycyclic aromatic hydrocarbons (PAHs) listed as priority pollutants by the U.S. EPA. Sediment chronologies were established using both 137Cs and 210Pb activity-depth profiles, and confirmed by independent information. The highest levels of PAHs, up to 16,474 g kg–1, characterise the sediment from the industrial canal. In lagoon sediments maximums were 618–1,531 g kg–1, while surficial values were 315 and 810 g kg–1. Dated concentration-depth profiles suggest that highest inputs occurred in the first half of last century and were followed by significant decreases. The industrial activities played a major role in the PAH contamination of lagoon sediments, as suggested by the high concentration gradients in the study area. The main source, based on the information provided by the relative abundance of congeners, is represented by high temperature combustion processes. Petrogenic sources may have influenced some samples, whereas the effects of selective transport and diagenesis are difficult to assess. The sediment of the industrial canal has the potential to occasionally cause adverse effects in sensitive species.  相似文献   

6.
Amano  Koji  Fukushima  Takehiko  Nakasugi  Osami 《Hydrobiologia》1992,235(1):491-499
Linear alkylbenzenesulfonate (LAS) was detected in a 0–30 cm deep sediment column collected in Lake Teganuma (one of the most polluted lakes in Japan). The range of the LAS concentration in sediments was between 0.1 and 500 µg g–1 (C11-C14 homologs per dry solid) and its vertical profile showed a seasonal variation. A mathematical model, which includes a diffusion term and a biodegradation term, was used to simulate the temporal variation of LAS in the sediment column and to calculate the diffusive flux rate of LAS across the sediment/water interface. An averaged diffusion coefficient of 2.4 × 10–5 cm2 s–1 for the sediment interstitial water was obtained from sediment core samples located in Lake Teganuma. The biodegradation rate constant (0.002 d–1) of LAS in the sediment obtained from the model analysis was considerably less than that reported for LAS in anaerobic waters. These results confirm that a model describing diffusive transport and biodegradation of LAS in the sediments can simulate the temporal variation of LAS in near surface sediments. The diffusive flux rate from overlying water to bottom sediment was calculated to be between –0.20 and 0.52 (C11-C14 LAS) mg m–2 h–1 and the annual net flux rate was 0.7 g m–2 y–1.  相似文献   

7.
From 2000 to 2002, sediment contamination by Cd, Cu, Hg, Pb and Zn was analysed in the Pialassa Baiona salt marsh, which receives petrochemical wastewaters from the industrial district of Ravenna (Italy). The recent contamination levels were compared with data of previous studies carried out in 1982, in order to assess whether environmental policies and remedial measures reduced sediment pollution. Sedimentary profiles of Cu and Pb were homogeneous along the uppermost 0–10 cm horizon, which corresponded to the sedimentation in the last 30 years. Concentrations of Zn attained a peak (up to 800 mg kg−1 dry weight) in the 0–4 cm sediment horizon, which was assumed to correspond to the last 10–15 years. A wide-spread contamination by Hg was detected in the salt marsh as well as in the main channel with peaks up to 20–40 mg kg−1 dry weight. Nonetheless, recent sediments resulted less contaminated, since Hg discharge from industrial plants ceased about 20 years ago. Contamination levels by Hg values were two orders of magnitude higher than the international sediment quality standards. Cadmium, which was analysed for the first time in 2000–2002, attained a peak in the surface layers (1–2.5 mg kg−1 d.w.), with a progressive decline along the sediment column. Through comparison with pre-industrial values detected in the deeper sediment horizons (before 1920), Hg showed the highest enrichment factor, up to 300 times. Cd and Zn concentrations in recent sediments were from 2 to 10 times higher than background values. In terms of possible adverse effects, Hg posed the highest risk, and Cd and Zn were frequently above the recommended thresholds.  相似文献   

8.
The effect of redox potential and pH on the phosphate mobility in two sediments were investigated using both consolidated and suspended sediments from the area where the Parana Medio long reservoir (Atgentina) is to be built (Smirnov, 1984). In addition to direct chemical sediment analysis, extraction techniques were carried out with a stepwise NH4Cl-NaOH-HCl shaking method, the latter supposedly separating the weakly bound, the Fe- and Al- bound and the Ca- bound phosphates in the sediments.Phosphate released into water depends upon redox potential and pH, which both were modified in an experimental setup. The source of the phosphate was the fraction of Fe and/or Al bound phosphate present both in the sediment and in the suspended solids.Abbreviations cm centimeter - km kilometer - gg gram - l liter - ¬m micrometer - °C grade centigrades - km2 square kilometer - m.s–1 meter per second - m3.s–1 cubic meter per second - mg.11 miligram per liter  相似文献   

9.
Behavioral and physiological responses to hypoxia were examined in three sympatric species of sharks: bonnethead shark Sphyrna tiburo, blacknose shark, Carcharhinus acronotus, and Florida smoothhound shark, Mustelus norrisi, using closed system respirometry. Sharks were exposed to normoxic and three levels of hypoxic conditions. Under normoxic conditions (5.5–6.4mg l–1), shark routine swimming speed averaged 25.5 and 31.0cm s–1 for obligate ram-ventilating S. tiburo and C. acronotus respectively, and 25.0cm s–1 for buccal-ventilating M. norrisi. Routine oxygen consumption averaged about 234.6 mg O2kg–1h–1 for S. tiburo, 437.2mg O2kg–1h–1 for C. acronotus, and 161.4mg O2 kg–1 h–1 for M. norrisi. For ram-ventilating sharks, mouth gape averaged 1.0cm whereas M. norrisi gillbeats averaged 56.0 beats min–1. Swimming speeds, mouth gape, and oxygen consumption rate of S. tiburo and C. acronotus increased to a maximum of 37–39cm s–1, 2.5–3.0cm and 496 and 599mg O2 kg–1 h–1 under hypoxic conditions (2.5–3.4mg l–1), respectively. M. norrisi decreased swimming speeds to 16cm s–1 and oxygen consumption rate remained similar. Results support the hypothesis that obligate ram-ventilating sharks respond to hypoxia by increasing swimming speed and mouth gape while buccal-ventilating smoothhound sharks reduce activity.  相似文献   

10.
Serrano  L.  Calzada-Bujak  I.  Toja  J. 《Hydrobiologia》2003,492(1-3):159-169
This study reports on the spatial and temporal variability of the phosphate composition in the sediment of a temporary pond over a period of 3 years using the EDTA-method for P-fractionation. Sediment samples were collected at three different sites (open-water, littoral and flood plain) to compare the effect of the length of the wet/dry phase on the sediment phosphate composition, with special emphasis on the potential bioavailability of the P-fractions.Fine sediments (<0.1 mm) were rich in organic matter (9–25%) and contained high mean concentrations of Tot-P (182–655 mg kg–1 d.w.), especially in the flood plain sediment. The sediment P composition was dominated by P-organic fractions at all sites (64–94%). The average C/N ratios were 8.8, 6.0 and 5.9 for sediments of the flood plain, littoral and open-water sites, respectively. The flood plain sediment was significantly poorer in iron-bound P (FeOOHP), but richer in the P-organic fractions extracted by EDTA than the sediment of the open-water site (P<0.01). The percentage of organic matter increased significantly in the sediment of the open-water site at the end of each dry season (P<0.05), while it decreased in the sediment of the flood plain site (P<0.01). In all sediments, the fraction of Fe(OOH)P decreased at the end of each dry season and some of these changes were significant (P<0.05). The decrease in the fraction of Fe(OOH)P was not related to changes in the sediment redox potential. Although the flood plain site was dry longer than the open-water site during the study period, the differences between the sediment composition of both sites were probably due to the effect of plant growth on the dry sediments of the flood plain site rather than to a direct effect of desiccation.  相似文献   

11.
Long-term monitoring (May 1990 – November 1994) of benthic macrofauna and sediment composition was undertaken to examine changes in community structure following the construction of a tidal barrage at the entrance to Sutton Harbour, Plymouth (South West England). The harbour has permanently anoxic sediments, below a depth of 3 cm, consisting mainly of silt with relatively high total metal concentrations in the m<63 m fraction for Cu (96 – 222 g g–1), Hg (0.7 – 2.0 g g–1), Pb (93 – 297 g g–1) and Zn (114 – 460 g g–1). Polychaete worms, particularly of the family Cirratulidae, dominated the 93 taxa that form the macrofauna of the fine sediments. Multi-dimensional scaling (MDS) identified significant changes, surpassing all seasonal variability, in benthic community structure between pre-construction and construction phases and the similarities percentages procedure (SIMPER) isolated the species differentiating pre- and post-construction samples. Principal components analysis (PCA) revealed changes in heavy metal concentrations and sediment size distribution, primarily between surveys encompassing the start of construction. Rank correlations between the biotic and abiotic similarity matrices (BIOENV) were highest for sediment Pb and Zn concentrations, indicating that these variables offered a potential explanation of the changes in community structure. The results demonstrate the sensitivity of the benthic community to man-induced change and the need for ecological factors to be accounted for in harbour development at other locations.  相似文献   

12.
We studied five 20-m transects onthe lower slope under tropical lower montanerain forest at 1900–2200 m above sea level. We collectedsamples of soil and of weekly rainfall,throughfall, litter leachate, and stream waterbetween 14 March 1998 and 30 April 1999 anddetermined the concentrations of Al, totalorganic C (TOC), Ca, Cl, Cu, K, Mg, Mn,NH4 +-N, NO3 -N, total N (TN), Na, P, S, and Zn. The soils were shallowInceptisols; pH ranged 4.4–6.3 in the Ohorizons and 3.9–5.3 in the A horizons, totalCa (6.3–19.3 mg kg–1) and Mgconcentrations (1.4–5.4) in the O horizon weresignificantly different between the transects.Annual rainfall was 2193 mm; throughfall variedbetween 43 and 91% of rainfall, cloud waterinputs were 3.3 mm a–1 except forone transect (203). The volume-weighted mean pHwas 5.3 in rainfall and 6.1–6.7 in throughfall.The median of the pH of litter leachate andstream water was 4.8–6.8 and 6.8, respectively.The concentrations of Ca and Mg in litterleachate and throughfall correlatedsignificantly with those in the soil (r =0.76–0.95). Element concentrations inthroughfall were larger than in rainfallbecause of leaching from the leaves (Al, TOC,Ca, K, Mg), particulate dry deposition (TOC,Cu, Cl, NH4 +-N), and gaseousdry deposition (NO3 -N, total N, S).Net throughfall (= throughfall-rainfalldeposition) was positive for most elementsexcept for Mn, Na, and Zn. High-flow eventswere associated with elevated Al, TOC, Cu, Mn,and Zn concentrations.  相似文献   

13.
Water, suspended particulate matter (SPM) and sediments were collected from the Esteiro de Estarreja (Ria de Aveiro, Portugal), which receives considerable quantities of waste mercury from a chlor-alkali plant. Dissolved and particulate Hg concentrations in the effluent ranged between 4 –167 g I–1 and 141–3144 g g–1, respectively, at pH values of >10. The effluent plume undergoes significant chemical changes during advection downestuary. The evidence suggested that adsorption of dissolved Hg onto organic-rich SPM was an important process. A maximum sediment Hg concentration of 500 g g–1 was found about 1.5 km from the discharge, as a result of the settling of Hg-rich SPM. Downestuary Hg concentrations in sediments decline to about 100 g g–1 at the mouth of the Esteiro. The particle-water interactions are discussed in terms of the transport of dissolved and particulate Hg into the Ria de Aveiro.  相似文献   

14.
Effects of picolinic acid (2-pyridinecarboxylic acid) and chromium(III) picolinate was studied on the chromium (Cr) accumulation of fodder radish (Raphanus sativus L. convar. oleiformis Pers., cv. Leveles olajretek) and komatsuna (Brassica campestris L. subsp. napus f. et Thoms. var. komatsuna Makino, cv. Kuromaru ) grown in a pot experiment. Control cultures, grown in an uncontaminated soil (UCS; humous sand with pHKCl 7.48, sand texture with 12.4% clay+silt content, organic carbon 0.56%, CaCO3 2.2%, CEC 6.2 cmolc kg–1, Cr 10.6 mg kg–1), accumulated low amounts of chromium (less than 5.4 g g–1) in their roots or shoots. When this UCS was artificially contaminated with 100 mg kg–1 Cr (CrCl3) later picolinic acid treatment promoted the translocation of chromium into the shoots of both species. In fodder radish shoots Cr concentration reached 30.4 g g–1 and in komatsuna shoots 44.5 g g–1. Application of ethylene diamine tetra-acetic acid (EDTA) to this Cr contaminated soil had similar effect to picolinic acid. When the UCS was amended with leather factory sewage sediment (which resulted in 853 mg kg–1 Cr in soil), Cr mobilization was observed only after repeated soil picolinic acid applications. From a galvanic mud contaminated soil (brown forest soil with pHKCl 6.77, loamy sand texture with 26.6% clay+silt content, organic carbon 1.23%, CaCO3 0.7%, CEC 24.5 cmolc kg–1, Cd 5.0 mg kg–1, Cr 135 mg kg–1, and Zn 360 mg kg–1) the rate of Cr mobilization was negligible, only a slight increase was observed in Cr concentration of fodder radish shoots after repeated picolinic acid treatments of soil. Presumably picolinic acid forms a water soluble complex (chromium(III) picolinate) with Cr in the soil, which promotes translocation of this element (and also Cu) into the shoots of plants. The rate of complex formation may be related to the binding forms and/or concentration of Cr in soil and also to soil characteristics (i.e. pH, CEC), since the rate of Cr translocation was the following: artificially contaminated soil > leather factory sewage sediment amended soil > galvanic mud contaminated soil. Four times repeated 10 mg kg–1 chromium(III) picolinate application to UCS multiplied the transport of chromium to shoots, as compared to single 10 mg kg–1 CrCl3 treatment. This also suggests that chromium(III) picolinate is forming in the picolinic acid treated Cr-contaminated soils, and plants more readily accumulates and translocates organically bound Cr than ionic Cr. Picolinic acid promotes Cr translocation in soil-plant system. This could be useful in phytoextraction (phytoremediation) of Cr contaminated soils or in the production of Cr enriched foodstuffs.  相似文献   

15.
Inorganic phosphate in exposed sediments of the River Garonne   总被引:6,自引:0,他引:6  
Fractionation of inorganic phosphates in sediments of the River Garonne was carried out during a period of low water discharge. Sediments were collected under 5 cm of water (LS), in the dried river bank (MS) and in the riparian forest (RS). Sediments were sampled at two dates, during a period in which the water level fell gradually, causing sediment LS to be air exposed.Sediments were analysed for total phosphate, iron bound phosphate (Fe(OOH)P) using Ca-NTA and Ca bound phosphate (CaCO3P) using Na-EDTA.Total-P varies from 552 (RS at date 1) to 2072 µg g–1 (LS at date 2). There are significant differences between sediments and a significant increase from date 1 to date 2 in sediment LS only (1825 to 2072 µg g–1). Fe(OOH)P varies from 186 (RS at date 1) to 874 µg g–1 (LS at date 2). The highest values correspond to sediment LS. Moreover, Fe(OOH)P increased significantly between date 1 and 2 in LS (560 to 874 µg g–1) as well as in sediment MS (248 to 432 µg g–1). Ca bound P concentrations differed significantly between sediments (75, 112, 235 µg g–1 for sediments RS, MS and LS respectively) but not between sampling dates.These differences are attributed to the conditions of deposition of the sediments (such as morphology and hydrology of the river) and to the changes in chemical composition during the drying out of the sediments.  相似文献   

16.
An experimental approach was taken to examine the processes of detritus decomposition in river sediments.Addition of macrophyte detritus (Alternanthera philoxeroides (Mart.) Griseb) to river sediment resultedin an increase in carbon mineralization from a sbasal rate of 200 up to 500 mg C m–2 d–1.Carbon mineralization after addition of oak detritus was only slightly higher than mineralization in sediments thatreceived no addition ( 200 mg C m–2 d–1). Bacterial biomass and production in sediments \s+ alligator-weedwere higher than in sediments + oak or zero-addition. In these experiments the major fate of addedalligatorweed was mineralization. For alligatorweed detritus, microbial metabolism depletes organic carbonrather than leading to increases in food quality. Therefore, a pulse input of alligatorweed detritus would notbe available as a long-term source of organic carbon. Oak detritus was not rapidly decomposed, and sopersists in these sediments for a longer period.  相似文献   

17.
Summary The effect of soy sauce oil and various other oils on protease production by Aspergillus oryzae NISL 1913 was studied in chemostat cultures (dilution rate=0.02 h–1). Soy sauce oil was consumed as a carbon source by the cells and also accelerated protease production. When soy sauce oil was used as sole carbon source, the specific protease production rate was 2.89 protease units·(mg dry weight of mycelium)–1·h–1, which was threefold higher than that with starch. The specific protease production rate with linoleic acid, oleic acid, Tween 80 and soybean oil exhibited similar values to that with soy sauce oil but the fatty acids with carbon chains shorter than six, such as caproic acid and acetic acid, did not stimulate protease production. The oils did not cause an increase in other exocellular enzymes such as -amylase, indicating that the protease production was selectively stimulated by the oils. Offprint requests to: Y. Fukushima  相似文献   

18.
Sulfate reduction rates were measured in waters and sediments from four antarctic lakes and an antarctic fjord basin by a radiometric technique. There was generally a linear correlation between the period of incubation and sulfate reduced; the average of the correlation coefficients was 0.76 ± 0.1. The rates at 6 °C were very low (0.0–1.1 µmol kg–1 d–1) when compared to most other marine and non-marine environments for which sulfate reduction rates have been reported. Lactate and acetate did not stimulate sulfate reduction. Temperatures of the sediments selected from the different sites varied from –0.4 to 4.5 °C and the chloride and sulfate concentrations of the sediments varied from 0.19 to 0.83 mol kg–1 and 0.04 to 41.01 mmol kg–1 respectively. Sulfate reduction rates did not correlate with the chlorosity of sediment porewaters.  相似文献   

19.
Rates of bacterial production were measured in the water column, on the surface of plant detritus, and in the surface sediments of a freshwater marsh in the Okefenokee Swamp, Georgia, USA. Bacterioplankton production rates were not correlated with several measures of quantity and quality of dissolved organic matter, including an index of the relative importance of vascular plant derivatives. Bacterioplankton productivity was high (mean: 63 g C liter–1 day–1) compared with rates reported for other aquatic ecosystems. Somewhat paradoxically, bacterial productivity on plant detritus (mean: 13 g C g–1 day–1) and sediments (mean: 15 g C g–1 day–1) was low relative to other locations. On an a real basis, total bacterial productivity in this marsh ecosystem averaged 22 mg C m–2 day–1, based on sample dates in May 1990 and February 1991. Marsh sediments supported the bulk of the production, accounting for 46% (May) and 88% (February) of the total. The remainder was contributed approximately equally by bacteria in the water column and on accumulated stores of plant detritus. Send offprint requests to: M. A. Moran.  相似文献   

20.
A sequential five-step extraction scheme for phosphorus pools in freshwater sediment was modified for use in marine sediments. In the second step phosphate bound to reducible forms of iron and manganese (iron-bound P) is extracted by a bicarbonate buffered dithionite solution (BD-reagent). The extraction scheme was tested on sediment from 16 m water depth in Aarhus Bay, DK and used in two other marine sediments: Kattegat at 56 m and Skagerrak at 695 m depth. By comparing the BD-extractable P-pool with both the pool of iron in the BD-fraction and the pool of oxidized, amorphous or poorly crystalline iron (am.FeOOH), highly significant correlations (p < 0.001) were observed in all three sediments. Thus, we conclude that the BD-reagent was very specific for iron-bound P. Further evidence for this came from two experiments: 1) Enhanced BD treatment did not result in additional phosphate extraction and 2) by sequential extraction of phosphorus pools in pure cultures of diatoms and cyanobacteria no phosphate was recovered in the BD-fraction. The pool of am.FeOOH was very important for controlling porewater phosphate concentration which was inferred from the significant inverse relationships between the two parameters (p < 0.001) in all sediments studied. Further, an isotopic exchange experiment with 32POf4/p3– revealed that BD-extractable P was by far the most exchangeable P-pool even deep in the sediment where the pool size was small. Iron-bound P made up 33–45% of total P in the surface sediments. The ratio between iron-bound phosphate and am.FeOOH was 8–11 in Aarhus Bay and Kattegat. In Skagerrak the ratio was 17, which may indicate that the iron mineral extracted from this sediment is less capable of adsorbing phosphate or less saturated with phosphate.  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