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.
Denitrification beds are a simple approach for removing nitrate (NO3) from a range of point sources prior to discharge into receiving waters. These beds are large containers filled with woodchips that act as an energy source for microorganisms to convert NO3 to nitrogen (N) gases (N2O, N2) through denitrification. This study investigated the biological mechanism of NO3 removal, its controlling factors and its adverse effects in a large denitrification bed (176 m × 5 m × 1.5 m) receiving effluent with a high NO3 concentration (>100 g N m−3) from a hydroponic glasshouse (Karaka, Auckland, New Zealand). Samples of woodchips and water were collected from 12 sites along the bed every two months for one year, along with measurements of gas fluxes from the bed surface. Denitrifying enzyme activity (DEA), factors limiting denitrification (availability of carbon, dissolved organic carbon (DOC), dissolved oxygen (DO), temperature, pH, and concentrations of NO3, nitrite (NO2) and sulfide (S2−)), greenhouse gas (GHG) production - as nitrous oxide (N2O), methane (CH4), carbon dioxide (CO2) - and carbon (C) loss were determined. NO3-N concentration declined along the bed with total NO3-N removal rates of 10.1 kg N d−1 for the whole bed or 7.6 g N m−3 d−1. NO3-N removal rates increased with temperature (Q10 = 2.0). In laboratory incubations, denitrification was always limited by C availability rather than by NO3. DO levels were above 0.5 mg L−1 at the inlet but did not limit NO3-N removal. pH increased steadily from about 6 to 7 along the length of the bed. Dissolved inorganic carbon (C-CO2) increased in average about 27.8 mg L−1, whereas DOC decreased slightly by about 0.2 mg L−1 along the length of the bed. The bed surface emitted on average 78.58 μg m−2 min−1 N2O-N (reflecting 1% of the removed NO3-N), 0.238 μg m−2 min−1 CH4 and 12.6 mg m−2 min−1 CO2. Dissolved N2O-N increased along the length of the bed and the bed released on average 362 g dissolved N2O-N per day coupled with N2O emission at the surface about 4.3% of the removed NO3-N as N2O. Mechanisms to reduce the production of this GHG need to be investigated if denitrification beds are commonly used. Dissolved CH4 concentrations showed no trends along the length of the bed, ranging from 5.28 μg L−1 to 34.24 μg L−1. Sulfate (SO42−) concentrations declined along the length of the bed on three of six samplings; however, declines in SO42− did not appear to be due to SO42− reduction because S2− concentrations were generally undetectable. Ammonium (NH4+) (range: <0.0007 mg L−1 to 2.12 mg L−1) and NO2 concentrations (range: 0.0018 mg L−1 to 0.95 mg L−1) were always very low suggesting that anammox was an unlikely mechanism for NO3 removal in the bed. C longevity was calculated from surface emission rates of CO2 and release of dissolved carbon (DC) and suggested that there would be ample C available to support denitrification for up to 39 years.This study showed that denitrification beds can be an efficient tool for reducing high NO3 concentrations in effluents but did produce some GHGs. Over the course of a year NO3 removal rates were always limited by C and temperature and not by NO3 or DO concentration.  相似文献   

2.
Syntheses and room-temperature single crystal X-ray structural characterizations are recorded for a variety of silver(I) oxyanion (perchlorate, nitrate and trifluoroacetate (‘tfa’) (increasing basicity)) adducts, AgX, with a number of pyridine (‘py’) bases, L, functionalized in the 2-position with N- or O-donor groups, namely 2-amino-, 2-amino-6-methyl-, 2-aminomethyl-, 2-hydroxy-, 2-methoxy- and 2-acetyl- pyridines, ‘2np’, ‘nmp’, ‘amp’, ‘ohp’, ‘mop’, and ‘acp’. A variety of stoichiometries and associated structural types are defined: [Ag(chelate)2]X, L/X = amp,acp/ClO4, [XAg(chelate)2], L/X = acp/tfa, of 1:2 AgX:L stoichiometry; for 1:1 stoichiometry, although a discrete mononuclear complex [(chelate)Ag(O2NO)] is defined for AgNO3: acp (1:1), all others are polymers, successive silver atoms being linked by N,N′-bridging ligands singly (L/X = 2np/ClO4 (?HAgHTAgTHAgH?), amp/ClO4, NO3 (?HTAgHTAg?) (‘H’ ≡ head, ‘T’ = tail)) or pairwise, ?L2AgX2AgL2Ag? (L/X = 2np/tfa, nmp/NO3). More complex polymeric arrays are found with L/X = ohp/NO3, tfa, where interaction with the metal takes place via the O-donor only, the py functionality being protonated, and in adducts of more complex stoichiometry AgNO3:mop (2:3) and AgNO3:2np (3:4).  相似文献   

3.
Synthetic, single crystal X-ray structural characterizations and vibrational spectroscopic studies are recorded for a number of adducts of 1:2 stoichiometry of silver(I) oxyanion salts for oxyanions of differing basicity (perchlorate, nitrate, carboxylate (as trifluoroacetate (≡‘tfa’))), with a variety of pyridine (≡‘py’) or piperidine (≡‘pip’) bases hindered in the 2- (and, sometimes, 6-) position(s) by methyl or non-coordinating functionalities of other types, the ligands employed being 2-methylpyridine (‘2mp’), 2,6-dimethylpyridine (‘lut’), 2,4,6-trimethylpyridine (‘coll’), quinoline (‘quin’), 2,2,6,6-tetramethylpiperidine (‘tmp’), 2-amino-,6-methylpyridine (‘nmp’), 2-methoxypyridine (‘mop’) and 2-cyanomethylpyridine (‘pcn’); studies are also recorded of adducts with the parent, ‘py’, base and with 4-cyanopyridine (‘cnp’). In the majority of the complexes, the NAgN motif predominates, as might be expected, variously distorted from linearity in response to changes in (competing) basicities of the nitrogen base and any nearby anion or solvent molecule; an unusual variation is found in the highly hindered tmp/tfa adduct which is a monohydrate with interacting water displacing the rather basic anion, the converse being the case in the corresponding nitrate, also a monohydrate. With the less-hindered base mpy, both nitrate and trifluoroacetate are binuclear, with O and OCO bridges corresponding to centrosymmetric four- and eight-membered rings, respectively; the quin/nitrate adduct is more complex, also binuclear but with bis(chelating) nitrate. AgNO3:py (1:3) is found to be binuclear, while with Agtfa/py, a 3:2 adduct [Ag(py)2][Ag2(tfa)3](∞|∞) is found with a novel, polymeric, strongly interacting anion. A further pair of 1:3 adducts, AgNO3:2np (2np = 2-aminopyridine) and Agtfa:nmp, both mononuclear [AgL3]+X are described, differing in the modes of interaction of silver with the three N-bases. In all simple NAgN systems with aromatic ligands, the pair of ligand ‘planes’ is disposed quasi-parallel.The far-IR spectra of [AgL2]Y (L = lut, coll; Y = ClO4, NO3, tfa) and of [Ag(py)n](ClO4) (n = 2,4) have been recorded and the ν(AgN) bands assigned in the range 80-240 cm−1. For the L = lut, coll complexes, there is a clear trend of decreasing ν(AgN) following increasing r(AgN) as the interaction with the counterion increases along the series Y = ClO4, NO3, tfa.  相似文献   

4.
The metabolic capability of denitrifying sludge to oxidize ammonium and p-cresol was evaluated in batch cultures. Ammonium oxidation was studied in presence of nitrite and/or p-cresol by 55 h. At 50 mg/L NH4+-N and 76 mg/L NO2-N, the substrates were consumed at 100% and 95%, respectively, being N2 the product. At 50 mg/L NH4+-N and 133 mg/L NO2-N, the consumption efficiencies decreased to 96% and 70%, respectively. The increase in nitrite concentration affected the ammonium oxidation rate. Nonetheless, the N2 production rate did not change. In organotrophic denitrification, the p-cresol oxidation rate was slower than ammonium oxidation. In litho-organotrophic cultures, the p-cresol and ammonium oxidation rates were affected at 133 mg/L NO2-N. Nonetheless, at 76 mg/L NO2-N the denitrifying sludge oxidized ammonium and p-cresol, but at different rate. Finally, this is the first work reporting the simultaneous oxidation of ammonium and p-cresol with the production of N2 from denitrifying sludge.  相似文献   

5.
We studied microbial N2 production via anammox and denitrification in the anoxic water column of a restored mining pit lake in Germany over an annual cycle. We obtained high-resolution hydrochemical profiles using a continuous pumping sampler. Lake Rassnitzer is permanently stratified at ca. 29 m depth, entraining anoxic water below a saline density gradient. Mixed-layer nitrate concentrations averaged ca. 200 μmol L−1, but decreased to zero in the anoxic bottom waters. In contrast, ammonium was <5 μmol L−1 in the mixed layer but increased in the anoxic waters to ca. 600 μmol L−1 near the sediments. In January and October, 15N tracer measurements detected anammox activity (maximum 504 nmol N2 L−1 d−1 in 15NH4+-amended incubations), but no denitrification. In contrast, in May, N2 production was dominated by denitrification (maximum 74 nmol N2 L−1 d−1). Anammox activity in May was significantly lower than in October, as characterized by anammox rates (maximum 6 vs. 16 nmol N2 L−1 d−1 in incubations with 15NO3), as well as relative and absolute anammox bacterial cell abundances (0.56% vs. 0.98% of all bacteria, and 2.7×104 vs. 5.2×104 anammox cells mL−1, respectively) (quantified by catalyzed reporter deposition-fluorescence in situ hybridization (CARD-FISH) with anammox bacteria-specific probes). Anammox bacterial diversity was investigated with anammox bacteria-specific 16S rRNA gene clone libraries. The majority of anammox bacterial sequences were related to the widespread Candidatus Scalindua sorokinii/brodae cluster. However, we also found sequences related to Candidatus S. wagneri and Candidatus Brocadia fulgida, which suggests a high anammox bacterial diversity in this lake comparable with estuarine sediments.  相似文献   

6.
The heavy use of fertilizers in agricultural lands can result in significant nitrate (NO3) loadings to the aquatic environment. We hypothesized that biological denitrification in agricultural ditches and streams could be enhanced by adding elemental sulfur (So) to the sediment layer, where it could act as a biofilm support and electron donor. Using a bench-scale stream mesocosm with a bed of So granules, we explored NO3 removal fluxes as a function of the effluent NO3 concentrations. With effluent NO3 ranging from 0.5 mg N L−1 to 4.1 mg N L−1, NO3 removal fluxes ranged from 228 mg N m−2 d−1 to 708 mg N m−2 d−1. This is as much as 100 times higher than for agricultural drainage streams. Sulfate (SO42−) production was high due to aerobic sulfur oxidation. Molecular studies demonstrated that the So amendment selected for Thiobacillus species, and that no special inoculum was required for establishing a So-based autotrophic denitrifying community. Modeling studies suggested that denitrification was diffusion limited, and advective flow through the bed would greatly enhance NO3 removal fluxes. Our results indicate that amendment with So is an effective means to stimulate denitrification in a stream environment. To minimize SO42− production, it may be better to place So deeper in the sediment layer.  相似文献   

7.
Climatic influence (global warming and decreased rainfall) could lead to an increase in the ecological and toxicological effects of the pollution in aquatic ecosystems, especially contamination from agricultural nitrate (NO3) fertilizers. Physicochemical properties of the surface waters and sediments of four selected sites varying in NO3 concentration along La Rocina Stream, which feeds Marisma del Rocio in Doñana National Park (South West, Spain), were studied. Electrical conductivity, pH, content in macro and microelements, total organic carbon and nitrogen, and dissolved carbon and nitrogen were affected by each sampling site and sampling time. Contaminant NO3 in surface water at the site with the highest NO3 concentration (ranged in 61.6-106.6 mg L−1) was of inorganic origin, most probably from chemical fertilizers, as determined chemically (90% of the total dissolved nitrogen from NO3) and by isotopic analysis of δ15N-NO3. Changes in seasonal weather conditions and hydrological effects at the sampling sites were also responsible for variations in some biological activities (dehydrogenase, β-glucosidase, arylsulphatase, acid phosphatase and urease) in sediments, as well as in the production of the greenhouse gases CO2, CH4 and N2O. Both organic matter and NO3 contents influenced rates of gas production. Increased NO3 concentration also resulted in enhanced levels of potential denitrification measured as N2O production. The denitrification process was affected by NO3 contamination and the rainfall regimen, increasing the greenhouse gases emissions (CO2, CH4 and especially N2O) during the driest season in all sampling sites studied.  相似文献   

8.
This study focused on effects from Monoporeia affinis reworking and ventilation activities on benthic fluxes and mineralization processes during a simulated bloom event. The importance of M. affinis density for benthic solute (O2, ΣNO2 + NO3, NH4+ and HPO42−) fluxes and sediment reactivity (mobilization of NH4+ and HPO42−) following additions of organic material to the sediment surface was experimentally investigated using sediment-water and closed sediment (jar) incubations. Three different densities of M. affinis were used to resemble a low, medium and high density situation (1300, 2500 and 6400 ind. m− 2, respectively) of a natural amphipod community. The degradation of phytodetritus (Tetraselmis sp., 5 g C m− 2) added to the sediment surface was followed over a period of 20 days. Benthic solute fluxes of O2, ΣNO2 + NO3 and NH4+ were generally progressively stimulated with increasing number of M. affinis, while no such correlation was found for HPO42−. Solute fluxes were initially enhanced 1 to 2 days after the addition of phytodetritius, caused by mineralization of the most labile organic material and a food-stimulated irrigation by the amphipods. There was no effect from the activity of M. affinis on total denitrification (Dtot = Dn + Dw) or denitrification utilizing nitrate from coupled nitrification/denitrification (Dn) for any of the densities examined. Denitrification utilizing overlying water nitrate (Dw) was only about 10% of Dtot. Dw was significantly enhanced for the highest M. affinis density investigated. The reactivity of the sediment decreased progressively with increasing density of M. affinis and with time of the experiment. However, enhanced ammonium production at least 6 days after the organic addition indicated excretion of N-containing organic compounds by M. affinis. In conclusion, large spatial and temporal variations in density of M. affinis may be of significant importance for benthic solute fluxes and overall mineralization of organic material in Baltic Sea sediments.  相似文献   

9.
The meagre (structurally defined) array of 1:2 silver(I) (pseudo-)halide:unidentate nitrogen base adducts is augmented by the single-crystal X-ray structural characterization of the 1:2 silver(I) thiocyanate:piperidine (‘pip’) adduct. It is of the one-dimensional ‘castellated polymer’ type previously recorded for the chloride: ?Ag(pip)2(μ-SCN)Ag(pip)2? a single bridging atom (S) linking successive silver atoms. By contrast, in its copper(I) counterpart, also a one-dimensional polymer, the thiocyanate bridges as end-bound SN-ambidentate: ?CuSCNCuSCN? A study of the 1:1 silver(I) bromide:quinoline (‘quin’) adduct is recorded, as the 0.25 quin solvate, isomorphous with its previous reported ‘saddle polymer’ chloride counterpart.Recrystallization of 1:1 silver(I) iodide:tris(2,4,6-trimethoxyphenyl)phosphine (‘tmpp’) mixtures from py and quinoline (‘quin’)/acetonitrile solutions has yielded crystalline materials which have also been characterized by X-ray studies. In both cases the products are salts, the cation in each being the linearly coordinated silver(I) species [Ag(tmpp)2]+, while the anions are, respectively, the discrete [Ag5I7(py)2]2− species, based on the already known but unsolvated [Cu5I7]2− discrete, and the polymeric, arrays, and polymeric . The detailed stereochemistry of the [Ag(tmpp)2]+ cation is a remarkably constant feature of all structures, as is its tendency to close-pack in sheets normal to their P-Ag-P axes.The far-IR spectra of the above species and of several related complexes have been recorded and assigned. The vibrational modes of the single stranded polymeric AgX chains in [XAg(pip)2](∞|∞) (X = Cl, SCN) are discussed, and the assignments ν(AgX) = 155, 190 cm−1 (X = Cl) and 208 cm−1 (X = SCN) are made. The ν(AgX) and ν(AgN) modes in the cubane tetramers [XAg(pip)]4 (X = Br, I) are assigned and discussed in relation to the assignments for the polymeric AgX:pip (1:2) complexes, and those for the polymeric [XAg(quin)](∞|∞) (X = Cl, Br) compounds. The far-IR spectra of [Ag(tmpp)2]2[Ag5I7(py)2] and its corresponding 2-methylpyridine complex show a single strong band at about 420 cm−1 which is assigned to the coordinated tmpp ligand in [Ag(tmpp)2]+, and a partially resolved triplet at about 90, 110 and 140 cm−1 which is assigned to the ν(AgI) modes of the [Ag5I7L2]2− anion. An analysis of this pattern is given using a model which has been used previously to account for unexpectedly simple ν(CuI) spectra for oligomeric iodocuprate(I) species.  相似文献   

10.
Ten novel adducts of the form AgClO4:PR3:L (1:1:1) (R = Ph, cy, o-tolyl; L = 2,2′-bipyridyl (‘bpy’), 2,2′-biquinoline (‘bq’), bis(2-pyridyl)amine (‘dpa’), bis(2-picolyl)amine (‘dpca’)) have been synthesized and characterized by analytical, spectroscopic (IR, far-IR, 1H and 31P NMR) and single crystal X-ray diffraction studies. The solid state molecular structures show that the complexes predominantly take the form [(R3P)AgL]+X, with a trigonal PAgN2 coordination environment, where the approach of the anion or the solvent may perturb the planarity of the silver environment. The ClO4 anion shows uni- or semi-bidentate coordination, except in the complexes AgClO4:PR3:dpca (1:1:1) (R = Ph and o-tolyl), where the anion remains uncoordinated and the dpca donor is a three-coordinate pincer-like ligand.  相似文献   

11.
Fluxes of oxygen, inorganic nitrogen (DIN) and denitrification (isotope pairing) were measured from January 1997 to February 1998 via intact cores incubation in a shallow brackish area within the eutrophic Valli di Comacchio (northern Adriatic coast, Italy). Rates were measured in the light and in the dark in sediments colonized by the rooted macrophyte Ruppia cirrhosa and in adjacent sediments with benthic microalgae. Ruppia biomass (25-414 g DW m− 2) exhibited a seasonal evolution whilst that of microphytobenthos (12-66 mg chl a m− 2) was more erratic. Net (NP) and gross (GP) primary productivity was 1.15 and 6.89 mol C m− 2y− 1 for bare and 25.4 and 51.7 mol C m− 2y− 1 for Ruppia vegetated sediments. Nitrogen pools in Ruppia standing stock varied from 43.6 to 631.4 (annual average 201.2) mmol N m− 2; the macrophyte N content was correlated with DIN concentration in the water column. Estimated N pool in microphytobenthos was one order of magnitude lower (from 2.4 to 14.5 mmol N m− 2, annual average 7.2). Theoretical DIN assimilation calculated from NP was 127.8 and 1112.6 mmol N m− 2y− 1 whilst that calculated from GP was 765 and 2282 mmol N m− 2y− 1 for microphytobenthos and Ruppia respectively. Measured annual fluxes of DIN were 974.6 and − 577 mmol N m− 2y− 1 in bare and Ruppia vegetated sediments meaning that the two sites were a source and sink for DIN and that from 25 to 50% of Ruppia annual DIN requirements came from the water column. During the period of this study total denitrification was lower in the macrophyte colonized (92.3 mmol N m− 2y− 1) compared to bare sediments (163.3 mmol N m− 2y− 1) as a probable consequence of higher competition between denitrifiers and phanerogams. At both sites the ratio between denitrification of water column nitrate (DW) and denitrification coupled to nitrification (DN) was >1.6 due to little oxygen penetration in reducing sediments (< 1.2 mm) and scarce nitrification activity. DW (0-35 µmol N m− 2h− 1) was significantly correlated with water column NO3−  (2-16 µM). Theoretical DIN assimilation to denitrification ratio varied from 12.0 to 24.8 for Ruppia vegetated and from 0.8 to 4.7 for unvegetated sediments.At Valle Smarlacca, Ruppia may influence nitrogen cycling by incorporating large DIN pools in biomass which is scattered in surrounding areas and fuels intense bacterial activity. With increasing anthropogenic nutrient input and insignificant organic matter export in the open sea the already severe eutrophic conditions are enhanced and may accelerate the decline of the macrophyte meadow.  相似文献   

12.
The effects of inorganic nitrogen (N) source (NH4+, NO3 or both) on growth, biomass allocation, photosynthesis, N uptake rate, nitrate reductase activity and mineral composition of Canna indica were studied in hydroponic culture. The relative growth rates (0.05-0.06 g g−1 d−1), biomass allocation and plant morphology of C. indica were indifferent to N nutrition. However, NH4+ fed plants had higher concentrations of N in the tissues, lower concentrations of mineral cations and higher contents of chlorophylls in the leaves compared to NO3 fed plants suggesting a slight advantage of NH4+ nutrition. The NO3 fed plants had lower light-saturated rates of photosynthesis (22.5 μmol m−2 s−1) than NH4+ and NH4+/NO3 fed plants (24.4-25.6 μmol m−2 s−1) when expressed per unit leaf area, but similar rates when expressed on a chlorophyll basis. Maximum uptake rates (Vmax) of NO3 did not differ between treatments (24-35 μmol N g−1 root DW h−1), but Vmax for NH4+ was highest in NH4+ fed plants (81 μmol N g−1 root DW h−1), intermediate in the NH4NO3 fed plants (52 μmol N g−1 root DW h−1), and lowest in the NO3 fed plants (28 μmol N g−1 root DW h−1). Nitrate reductase activity (NRA) was highest in leaves and was induced by NO3 in the culture solutions corresponding to the pattern seen in fast growing terrestrial species. Plants fed with only NO3 had high NRA (22 and 8 μmol NO2 g−1 DW h−1 in leaves and roots, respectively) whereas NRA in NH4+ fed plants was close to zero. Plants supplied with both forms of N had intermediate NRA suggesting that C. indica takes up and assimilate NO3 in the presence of NH4+. Our results show that C. indica is relatively indifferent to inorganic N source, which together with its high growth rate contributes to explain the occurrence of this species in flooded wetland soils as well as on terrestrial soils. Furthermore, it is concluded that C. indica is suitable for use in different types of constructed wetlands.  相似文献   

13.
Twenty-one adducts of the form AgX:ER3:L (1:1:1) (X = CF3COO (‘tfa’), CH3COO (‘ac’), E = P, As; R = Ph, cy, o-tolyl; L = 2,2′-bipyridyl (‘bpy’)-based ligand) have been synthesized and characterized by analytical, spectroscopic (IR, far-IR, 1H, 19F and 31P NMR) and single crystal X-ray diffraction studies. The resulting complexes are predominantly of the form [(R3E)AgL]+X, with a trigonal EAgN2 coordination environment, the planarity of which may be perturbed by the approach of anion or solvent. The carboxylate anions have been found to be uni-, or semi-bidentate, or also completely ionic, as in the complexes [Ag(PPh3)(bpy)(H2O)](tfa) and [Ag(PPh3)(dpk · H2O)](tfa) (‘dpk · H2O’ = bis(2-pyridyl)ketone (hydrated)). The complexes Agac:PPh3:dpa (1:1:1) and Agac:P(o-tol)3:dpa:MeCN (1:1:1:1) are dinuclear, with bridging unidentate acetate and terminal unidentate dpa (‘dpa’ = bis(2-pyridyl)amine).  相似文献   

14.
Cytidine (cyt) and adenosine (ado) react with cis-[L2Pt(μ-OH)]2(NO3)2 (L = PMe3, PPh3) in various solvents to give the nucleoside complexes cis-[L2Pt{cyt(− H),N3N4}]3(NO3)3 (L = PMe3, 1),cis-[L2Pt{cyt(− H),N4}(cyt,N3)]NO3 (L = PPh3, 2), cis-[L2Pt{ado(− H),N1N6}]2(NO3)2 (L = PMe3, 3) and cis-[L2Pt{ado(− H),N6N7}]NO3 (L = PPh3, 4). When the condensation reaction is carried out in solution of nitriles (RCN, R = Me, Ph) the amidine derivatives cis-[(PPh3)2PtNH=C(R){cyt(− 2H)}]NO3 (R = Me, 5a; R = Ph, 5b) and cis-[(PPh3)2PtNH=C(R){ado(− 2H)}]NO3 (R = Me, 6a: R = Ph, 6b) are quantitatively formed. The coordination mode of these nucleosides, characterized in solution by multinuclear NMR spectroscopy and mass spectrometry, is similar to that previously observed for the nucleobases 1-methylcytosine (1-MeCy) and 9-methyladenine (9-MeAd). The cytotoxic properties of the new complexes, and those of the nucleobase analogs, cis-[(PPh3)2PtNH=C(R){1-MeCy(− 2H)}]NO3 (R = Me, 7a: R = Ph, 7b), cis-[(PPh3)2PtNH=C(R){9-MeAd(− 2H)}]NO3 (R = Me, 8a: R = Ph, 8b) have been investigated in a wide panel of human cancer cells. Interestingly, whereas the Pt(II) nucleoside complexes (1-4) did not show appreciable cytotoxicity, the corresponding amidine derivatives (7a, 7b, 8a, 8b, 5b, and 6b) exhibited a significant in vitro antitumor activity.  相似文献   

15.
In this study we assessed the growth, morphological responses, and N uptake kinetics of Salvinia natans when supplied with nitrogen as NO3, NH4+, or both at equimolar concentrations (500 μM). Plants supplied with only NO3 had lower growth rates (0.17 ± 0.01 g g−1 d−1), shorter roots, smaller leaves with less chlorophyll than plants supplied with NH4+ alone or in combination with NO3 (RGR = 0.28 ± 0.01 g g−1 d−1). Ammonium was the preferred form of N taken up. The maximal rate of NH4+ uptake (Vmax) was 6–14 times higher than the maximal uptake rate of NO3 and the minimum concentration for uptake (Cmin) was lower for NH4+ than for NO3. Plants supplied with NO3 had elevated nitrate reductase activity (NRA) particularly in the roots showing that NO3 was primarily reduced in the roots, but NRA levels were generally low (<4 μmol NO2 g−1 DW h−1). Under natural growth conditions NH4+ is probably the main N source for S. natans, but plants probably also exploit NO3 when NH4+ concentrations are low. This is suggested based on the observation that the plants maintain high NRA in the roots at relatively high NH4+ levels in the water, even though the uptake capacity for NO3 is reduced under these conditions.  相似文献   

16.
The new mononuclear bis(oxamato) complex [n-Bu4N]2[Cu(obbo)] (1) (obbo=o-benzyl-bis(oxamato)) has been synthesized as a precursor for trinuclear oxamato-bridged transition metal complexes. Starting from 1 the homotrinuclear complexes [Cu3(obbo)(pmdta)2(NO3)](NO3)·CH2Cl2·H2O (2) and [Cu3(obbo)(tmeda)2(NO3)2(dmf)] (3) have been prepared, where pmdta = N,N,N′,N″,N″-pentamethyldiethylenetriamine, tmeda = N,N,N′,N′-tetramethylethylenediamine and dmf = dimethylformamide. The crystal structures of 1-3 were solved. The magnetic properties of 2 and 3 were studied by susceptibility measurements versus temperature. For the intramolecular J parameter values of −111 cm−1 (2) and −363 cm−1 (3) were obtained.  相似文献   

17.
Wong BT  Lee DJ 《Bioresource technology》2011,102(12):6673-6679
The inhibitory effects of 90-189 mg l−1 of sulfide and 25-75 mg-N l−1 of nitrate on methanogenesis were investigated in a mixed methanogenic culture using butyrate as carbon source. In the initial phase of 90 mg l−1 S2− test, autotrophic denitrification of nitrate occurred with sulfide as the electron donor. Then the sulfate-reducing strains converted the produced sulfur back to sulfide via heterotrophic oxidation pathway. Methanogenesis was not markedly inhibited when 90 mg l−1 of sulfide was dosed alone. When 25-75 mg-N l−1 of nitrate was presented, initiation of methanogenesis was seriously delayed. Nitrogen oxides (NOx), the intermediates for nitrate reduction via denitrification pathway, inhibited methanogenesis. The 90 mg l−1 of sulfide favored heterotrophic dissimilatory nitrate reduction to ammonia (DNRA) pathway for nitrate reduction. Possible ways of maximizing methane production from an organic carbon-rich wastewater with high levels of sulfide and nitrate were discussed.  相似文献   

18.
The effects of short term hypoxia on bioturbation activity and inherent solute fluxes are scarcely investigated even if increasing number of coastal areas are subjected to transient oxygen deficits. In this work dark fluxes of oxygen (O2), dissolved inorganic carbon (TCO2) and nutrients across the sediment-water interface, as well as rates of denitrification (isotope pairing), were measured in intact sediment cores collected from the dystrophic pond of Sali e Pauli (Sardinia, Italy). Sediments were incubated at 100, 70, 40 and 10% of O2 saturation in the overlying water, with both natural benthic communities, dominated by the polychaete Polydora ciliata (11.100 ± 2.500  ind. m− 2), and after the addition of individuals of the deep-burrower polychaete Hediste diversicolor. Below an uppermost oxic layer of ~ 1 mm, sediments were highly reduced, with up to 6 mM of S2− in the 5 mm layer. Flux of S2− and O2 calculated from pore water gradients were 8.61 ± 1.12 and − 2.27 ± 0.56 mmol m− 2 h− 1, respectively. However, sediment oxygen demand (SOD) calculated from core incubation was − 10.52 ± 0.33 mmol m− 2 h− 1, suggesting a major contribution of P. ciliata to O2-mediated sulphide oxidation. P. ciliata also strongly stimulated NH4+ and PO43− fluxes, with rates ~ 15 and ~ 30 folds higher, respectively, than those estimated from pore water gradients. P. ciliata activity was significantly reduced at 10% O2 saturation, coupled to decreased rates of solutes transfer. The addition of H. diversicolor further stimulated SOD, NH4+ efflux and SiO2 mobilisation. Similarly to P. ciliata, the degree of stimulation of SOD and NH4+ flux by H. diversicolor depended on the level of oxygen saturation. TCO2 regeneration, respiratory quotients, PO43− fluxes and denitrification of added 15NO3 were not affected by the addition of H. diversicolor, but depended upon the O2 levels in the water column. Denitrification rates supported by water column 14NO3 and sedimentary nitrification were both negligible (< 0.5 µmol m− 2 h− 1). They were not significantly affected by oxygen saturation nor by bioturbation, probably due to the limited availability of NO3 in the water column (< 3 µM) and O2 in the sediments. This study demonstrates for the first time the integrated short term effect of transient hypoxia and bioturbation on solute fluxes across the sediment-water interface within a simplified lagoonal benthic community.  相似文献   

19.
This study assesses the growth and morphological responses, nitrogen uptake and nutrient allocation in four aquatic macrophytes when supplied with different inorganic nitrogen treatments (1) NH4+, (2) NO3, or (3) both NH4+ and NO3. Two free-floating species (Salvinia cucullata Roxb. ex Bory and Ipomoea aquatica Forssk.) and two emergent species (Cyperus involucratus Rottb. and Vetiveria zizanioides (L.) Nash ex Small) were grown with these N treatments at equimolar concentrations (500 μM). Overall, the plants responded well to NH4+. Growth as RGR was highest in S. cucullata (0.12 ± 0.003 d−1) followed by I. aquatica (0.035 ± 0.002 d−1), C. involucratus (0.03 ± 0.002 d−1) and V. zizanioides (0.02 ± 0.003 d−1). The NH4+ uptake rate was significantly higher than the NO3 uptake rate. The free-floating species had higher nitrogen uptake rates than the emergent species. The N-uptake rate differed between plant species and seemed to be correlated to growth rate. All species had a high NO3 uptake rate when supplied with only NO3. It seems that the NO3 transporters in the plasma membrane of the root cells and nitrate reductase activity were induced by external NO3. Tissue mineral contents varied with species and tissue, but differences between treatments were generally small. We conclude, that the free-floating S. cucullata and I. aquatica are good candidate species for use in constructed wetland systems to remove N from polluted water. The rooted emergent plants can be used in subsurface flow constructed wetland systems as they grow well on any form of nitrogen and as they can develop a deep and dense root system.  相似文献   

20.
Competitive abilities of Lagarosiphon major (Ridley) Moss (invasive in Belgium) and native Ceratophyllum demersum L. were assessed experimentally in relation to sediment dredging. We mimicked these conditions by taking undisturbed sediment (‘before dredging’ treatment) and by using restored sediment where the uppermost nutrient rich top layer was removed (‘after dredging’ treatment). Both the species were allowed to grow for seven weeks in monocultures and mixed cultures at different planting densities. Overall, invasive L. major performed better than native C. demersum independent of the characteristics of the growth environment. L. major achieved a higher relative growth rate (RGR) in both treatments based on total length (0.17-0.21 week−1) and weight (0.10-0.19 week−1) compared to C. demersum (length: 0.04-0.07 week−1; weight: 0.03-0.17 week−1). The better performance of L. major was due to a high plasticity under stressful conditions of low free CO2 and high pH. Intraspecific competition and niche partitioning were observed between the two species indicating that species coexistence is favoured instead of competitive exclusion. L. major performed better in the ‘after dredging’ treatment. Consequently, we deduce that sediment dredging will not lead to a decline of the invasive L. major.  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