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 437 毫秒
1.
Mercury (Hg) resistance is widespread among microorganisms and is based on the intracellular transformation of Hg(II) to less toxic elemental Hg(0). The use of microbial consortia to demercurize polluted wastewater streams and environments has been demonstrated. To develop efficient and versatile microbial cleanup strategies requires detailed knowledge of transport and reaction rates. This study focuses on the kinetics of the key enzyme of the microbial transformation, e.g., the mercuric reductase (MerA) under conditions closely resembling the cell interior. To this end, previously constructed and characterized Pseudomonas putida strains expressing MerA from Serratia marcescens were applied. Of the P. putida strains considered in this study P. putida KT2442::mer73 constitutively expressing broad spectrum mercury resistance (merTPAB) yielded the highest mercuric reductase (MerA) activity directly after cell disruption. MerA in the raw extract was further purified (about 100 fold). Reduction rates were measured for various substrates (HgCl2, Hg2SO4, Hg(NO3)2 and phenyl mercury acetate) up to high concentrations dependent on the purification grade. In all cases, a pronounced substrate inhibition was found. The kinetic constants determined for the cell raw extract are in agreement with those measured for intact cells. However, the rate data exhibit reduced affinity and inhibition with rising purification grade (specific activity). Therefore, the findings seemingly point to reactions preceding the catalytic reduction. Based on simplified assumptions, a kinetic model is suggested which reasonably describes the experimental findings and can advantageously be applied to the bioreactor design.  相似文献   

2.
3.
The relationship of mercury resistance to the concentration and chemical speciation of mercurial compounds was evaluated for microbial communities of mercury-polluted and control waters. Methodologies based on the direct viable counting (DVC) method were adapted to enumerate mercury-resistant communities. Elevated tolerance to Hg(II) was observed for the microbial community of one mercury-polluted pond as compared to the community of control waters. These results suggest an in situ acclimation to Hg(II). The results of the methylmercury resistance-DVC assay suggested that minimal acclimation to CH3Hg+ occurred since similar concentrations of CH3HgCl inhibited growth of 50% of organisms in both the control and polluted communities. Analyses of different mercury species in pond waters suggested that total mercury, but not CH3Hg+ concentrations, approached toxic levels in the polluted ponds. Thus, microbial acclimation was specific to the chemical species of mercury present in the water at concentrations high enough to cause toxic effects to nonacclimated bacterial communities.  相似文献   

4.
The kinetics of inorganic Hg [Hg(II)i] association, methylation, and methylmercury (MeHg) demethylation were examined for a group of Desulfovibrio species with and without MeHg production capability. We employed a detailed method for assessing MeHg production in cultures, including careful control of medium chemistry, cell density, and growth phase, plus mass balance of Hg(II)i and MeHg during the assays. We tested the hypothesis that differences in Hg(II)i sorption and/or uptake rates drive observed differences in methylation rates among Desulfovibrio species. Hg(II)i associated rapidly and with high affinity to both methylating and nonmethylating species. MeHg production by Hg-methylating strains was rapid, plateauing after ∼3 h. All MeHg produced was rapidly exported. We also tested the idea that all Desulfovibrio species are capable of Hg(II)i methylation but that rapid demethylation masks its production, but we found this was not the case. Therefore, the underlying reason why MeHg production capability is not universal in the Desulfovibrio is not differences in Hg affinity for cells nor differences in the ability of strains to degrade MeHg. However, Hg methylation rates varied substantially between Hg-methylating Desulfovibrio species even in these controlled experiments and after normalization to cell density. Thus, biological differences may drive cross-species differences in Hg methylation rates. As part of this study, we identified four new Hg methylators (Desulfovibrio aespoeensis, D. alkalitolerans, D. psychrotolerans, and D. sulfodismutans) and four nonmethylating species (Desulfovibrio alcoholivorans, D. tunisiensis, D. carbinoliphilus, and D. piger) in our ongoing effort to generate a library of strains for Hg methylation genomics.  相似文献   

5.
Mercury Adaptation among Bacteria from a Deep-Sea Hydrothermal Vent   总被引:4,自引:1,他引:3       下载免费PDF全文
Since deep-sea hydrothermal vent fluids are enriched with toxic metals, it was hypothesized that (i) the biota in the vicinity of a vent is adapted to life in the presence of toxic metals and (ii) metal toxicity is modulated by the steep physical-chemical gradients that occur when anoxic, hot fluids are mixed with cold oxygenated seawater. We collected bacterial biomass at different distances from a diffuse flow vent at 9°N on the East Pacific Rise and tested these hypotheses by examining the effect of mercuric mercury [Hg(II)] on vent bacteria. Four of six moderate thermophiles, most of which were vent isolates belonging to the genus Alcanivorax, and six of eight mesophiles from the vent plume were resistant to >10 μM Hg(II) and reduced it to elemental mercury [Hg(0)]. However, four psychrophiles that were isolated from a nearby inactive sulfide structure were Hg(II) sensitive. A neighbor-joining tree constructed from the deduced amino acids of a PCR-amplified fragment of merA, the gene encoding the mercuric reductase (MR), showed that sequences obtained from the vent moderate thermophiles formed a unique cluster (bootstrap value, 100) in the MR phylogenetic tree, which expanded the known diversity of this locus. The temperature optimum for Hg(II) reduction by resting cells and MR activity in crude cell extracts of a vent moderate thermophile corresponded to its optimal growth temperature, 45°C. However, the optimal temperature for activity of the MR encoded by transposon Tn501 was found to be 55 to 65°C, suggesting that, in spite of its original isolation from a mesophile, this MR is a thermophilic enzyme that may represent a relic of early evolution in high-temperature environments. Results showing that there is enrichment of Hg(II) resistance among vent bacteria suggest that these bacteria have an ecological role in mercury detoxification in the vent environment and, together with the thermophilicity of MR, point to geothermal environments as a likely niche for the evolution of bacterial mercury resistance.  相似文献   

6.
Herein, we evaluate the binding of Pb(II) and Bi(III) to cysteine-substituted versions of the TRI peptides [AcG-(LKALEEK)4G-NH2] which have previously been shown to bind Hg(II) and Cd(II) in unusual geometries as compared with small-molecule thiol ligands in aqueous solutions. Studies of Pb(II) and Bi(III) with the peptides give rise to complexes consistent with the metal ions bound to three sulfur atoms with M–S distances of 2.63 and 2.54 Å, respectively. Competition experiments between the metal ions Pb(II), Cd(II), Hg(II) and Bi(III) for the peptides show that Hg(II) has the highest affinity, owing to the initial formation of the extremely strong HgS2 bond. Cd(II) and Pb(II) have comparable binding affinities at pH > 8, while Bi(III) displays the weakest affinity, following the model, M(II) + (TRI LXC)3 3? → M(II)(TRI LXC)3 ?. While the relevant equilibria for Hg(II) binding to the TRI peptides corresponds to a strong first step forming Hg(TRI LXC)2(HTRI LXC), followed by a single deprotonation to give Hg(TRI LXC)3 ?, the binding of Cd(II) and Pb(II) is consistent with initial formation of M(II)(TRI LXC)(HTRI LXC)2 + at pH < 5 followed by a two-proton dissociation step (pK a2) yielding M(II)(TRI LXC)3 ?. Pb(II)(TRI LXC)(HTRI LXC)2 + converts to Pb(II)(TRI LXC)3 ? at slightly lower pH values than the corresponding Cd(II)–peptide complexes. In addition, Pb(II) displays a lower pK a of binding to the “d”-substituted peptide, (TRI L12C, pK a2 = 12.0) compared with the “a”-substituted peptide, (TRI L16C, pK a2 = 12.6), the reverse of the order seen for Hg(II) and Cd(II). Pb(II) also showed a stronger binding affinity for TRI L12C (K bind = 3.2 × 107 M?1) compared with that with TRI L16C (K bind = 1.2 × 107 M?1) at pH > 8.  相似文献   

7.
The structure and speciation of the complexes formed between mercury(II) ions and glutathione (GSH = L-glutamyl-L-cysteinyl-glycine) have been studied for a series of alkaline aqueous solutions (\( C_{{{\text{Hg}}^{{2 + }}}}\,{\sim18\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}}\) and C GSH = 40–200 mmol dm?3 at pH ~10.5) by means of extended X-ray absorption fine structure (EXAFS) and 199Hg NMR spectroscopy at ambient temperature. The dominant complexes are [Hg(GS)2]4? and [Hg(GS)3]7?, with mean Hg–S bond distances of 2.32(1) and 2.42(2) Å observed in digonal and trigonal Hg–S coordination, respectively. The proportions of the Hg2+–glutathione complexes were evaluated by fitting linear combinations of model EXAFS oscillations representing each species to the experimental EXAFS spectra. The [Hg(GS)4]10? complex, with four sulfur atoms coordinated at a mean Hg–S bond distance of 2.52(2) Å, is present in minor amounts (<30%) in solutions containing a large excess of glutathione (C GSH ≥ 160 mmol dm?3). Comparable alkaline mercury(II) cysteine (H2Cys) solutions were also investigated and a reduced tendency to form higher complexes was observed, because the deprotonated amino group of Cys2? allows the stable [Hg(S,N-Cys)2]2? chelate to form. The effect of temperature on the distribution of the Hg2+–glutathione complexes was studied by comparing the EXAFS spectra at ambient temperature and at 25 K of a series of glycerol/water (33/67, v/v) frozen glasses with \( C_{{{\text{Hg}}^{{2 + }} }} \,{\sim7\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}} \) and C GSH = 16–81 mmol dm?3. Complexes with high Hg–S coordination numbers, [Hg(GS)3]7? and [Hg(GS)4]10?, became strongly favored when just a moderate excess of glutathione (C GSH ≥28 mmol dm?3) was used in the glassy samples, as expected for a stepwise exothermic bond formation. Addition of glycerol had no effect on the Hg(II)–glutathione speciation, as shown by the similarity of the EXAFS spectra obtained at room temperature for two parallel series of Hg(II)-glutathione solutions with \( C_{{{\text{Hg}}^{{2 + }} }} \,{\sim7\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}},\) with and without 33% glycerol. Also, the 199Hg NMR chemical shifts of a series of ~18 mmol dm?3 mercury(II) glutathione solutions with 33% glycerol were not significantly different from those of the corresponding series in aqueous solution.  相似文献   

8.
Understanding the biogeochemical cycle of the highly toxic element mercury (Hg) is necessary to predict its fate and transport. In this study, we determined that biogenic magnetite isolated from Magnetospirillum gryphiswaldense MSR-1 and Magnetospirillum magnetotacticum MS-1 was capable of reducing inorganic mercury [Hg(II)] to elemental mercury [Hg(0)]. These two magnetotactic bacteria (MTB) lacked mercuric resistance operons in the genomes. However, they revealed high resistance to Hg(II) under atmospheric conditions and an even higher resistance under microaerobic conditions (1% O2 and 99% N2). Neither strain reduced Hg(II) to Hg(0) under atmospheric conditions. However, a slow rate (0.05–0.21 µM·d?1) of Hg(II) loss occurred from late log phase to stationary phase in two MTBs' culture media under microaerobic conditions. Increased Hg(II) entered both cells under microaerobic conditions relative to atmospheric conditions. The majority of Hg(II) was still blocked by the cell membrane. Hg(II) reduction was more effective when biogenic magnetite was extracted out, with or without the magnetosome membrane envelope. When magnetosome membrane was present, 8.55–13.53% of 250 nM Hg(II) was reduced to Hg(0) by 250 mg/L biogenic magnetite suspension within 2 hours. This ratio increased to 55.07–64.70% while magnetosome membrane was removed. We concluded that two MTBs contributed to the reduction of Hg(II) to Hg(0) at a slow rate in vivo. Such reduction was more favorable to occur when biogenic magnetite is released from dead cells. It proposed a new biotic pathway for the formation of Hg(0) in aquatic systems.  相似文献   

9.
The effects of Ag(I) and Hg(II) on membrane potential and integrity of cells of Candida albicans and C. maltosa were determined with a flow cytometric procedure that employed an anionic membrane potential-sensitive dye, bis-(1,3-dibutylbarbituric acid) trimethine oxonol, and a membrane integrity indicator, propidium iodide. The membrane potentials of cells of both species were reduced rapidly within 15 min of exposure to Ag(I). No threshold dose for Hg(II) existed, and cells of both species lost membrane potential gradually in Hg(II) solutions. Cells of both species lost membrane integrity more rapidly in Ag(I) solutions than in Hg(II) solutions. In Ag(I) solutions, the decrease in the numbers of cells recoverable in culture occurred at a rate similar to the rate of cell depolarization and membrane permeabilization. In Hg(II) solutions, loss of cell recoverability preceded the loss of membrane potential and membrane integrity. C. albicans, in contrast to C. maltosa, showed no loss of membrane integrity after exposure to Hg(II) solutions for 1 h. Different rates of binding of Ag(I) and Hg(II) between the two species suggest that the two ions target different primary sites.  相似文献   

10.
The reactivity of N-(2-aminophenyl)-d-glycero-d-gulo-heptonamide (adgha), with the group 12 cations, Zn(II), Cd(II), and Hg(II), was studied in DMSO-d6 solution. The studied system showed a selective coordination to Hg(II), and the products formed were characterized by 1H and 13C NMR in DMSO-d6 solution and fast atom bombardment (FAB+) mass spectra. The expected coordination compounds, [Hg(adgha)](NO3)2 and [Hg(adgha)2](NO3)2, were observed as unstable intermediates that decompose to bis-[2-(d-glycero-d-gulo-hexahydroxyhexyl)-benzimidazole-κN]mercury(II) dinitrate, [Hg(ghbz)2](NO3)2. The chemical transformation of the complexes was followed by NMR experiments, and the nature of the species formed is sustained by a theoretical study done using DFT methodology. From this study, we propose the structure of the complexes formed in solution, the relative stability of the species formed, and the possible role of the solvent in the observed transformations.  相似文献   

11.
The preparation and characterization of mono and binuclear complexes of Ru(II) with a newly synthesized derivate of the terpyridine ligand, 4-(5-bromothiophene)-2,2,6,2″-terpyridine, are communicated. In the binuclear complex, 2,5-bis(2,2,6,2″-terpyridine-4yl)thiophene was used as a bridge between two Ru(II) centers. The new compounds were characterized by H NMR, UV-Vis and IR spectroscopies. Bands at ∼500 nm for the Ru(II) to terpyridine charge transfer transition and absorption bands at λ<400 nm assigned to intraligand transitions, π*←π, centered in the tpy moiety were observed in the UV-Vis spectra of the complexes. Irradiation of the complexes in CH3CN at 337 or 500 nm induced luminescence with maxima at ∼670 nm and lifetimes τ?102 ns. Time-resolved absorption spectroscopy revealed the formation of long-lived species during the decay of the metal to ligand charge transfer excited states. The intermediates were tentatively assigned as unstable products of ligand-substitution or orthometalation excited state reactions.  相似文献   

12.
Acute effects of mercuric chloride (HgCl2) were evaluated on mice. Mice received a single dose of HgCl2 (4.6 mg/kg, subcutaneously) for three consecutive days. Thirty minutes after the last injection with HgCl2, mice received one single injection of 2,3-dimercapto-1-propanesulfonic acid (DMPS) or N-acetylcysteine (NAC) or diphenyl diselenide (PhSe)2. DMPS, NAC and (PhSe)2 were utilized as therapy against mercury exposure. At 24 h after the last HgCl2 injection, blood, liver and kidney samples were collected. δ-Aminolevulinate dehydratase (δ-ALA-D) and Na+, K-+ ATPase activities, thiobarbituric acid-reactive substances (TBARS), non-protein thiols (NPSH) and ascorbic acid concentrations were evaluated. Plasma aspartate (AST) and alanine (ALT) aminotransferase activities, as well as urea and creatinine levels were determined. The group of mice exposed to Hg + (PhSe)2 presented 100% of lethality. Exposure with HgCl2 caused a decrease on the body weight gain and treatments did not modify this parameter. δ-ALA-D, AST and ALT activities, TBARS, ascorbic acid levels and NPSH (hepatic and erythrocytic) levels were not changed after HgCl2 exposure. HgCl2 caused an increase in renal NPSH content and therapies did not modify these levels. Mice treated with (PhSe)2, Hg + NAC and Hg + DMPS presented a reduction in plasma NPSH levels. Creatinine and urea levels were increased in mice exposed to Hg + NAC, while Hg + DMPS group presented an increase only in urea level. Na+, K-+ ATPase activity was inhibited in mice exposed to Hg + DMPS and Hg + NAC. In conclusion, therapies with (PhSe)2, DMPS and NAC following mercury exposure must be better studied because the formation of more toxic complexes with mercury, which can mainly damage renal tissue.  相似文献   

13.
The response of Desulfovibrio vulgaris to Cu(II) and Hg(II) was characterized. Both metals increased the lag phase, and Cu(II) reduced cell yield at concentrations as low as 50 μM. mRNA expression was analyzed using random arbitrarily primed PCR, differential display, and quantitative PCR. Both Cu(II) and Hg(II) (50 μM) caused upregulation of mRNA expression for an ATP binding protein (ORF2004) and an ATPase (ORF856) with four- to sixfold increases for Hg(II) and 1.4- to 3-fold increases with Cu(II). These results suggest that D. vulgaris uses an ATP-dependent mechanism for adapting to toxic metals in the environment.  相似文献   

14.
Although resistance of microorganisms to Hg(II) salts has been widely investigated and resistant strains have been reported from many eubacterial genera, there are few reports of mercuric ion resistance in extremophilic microorganisms. Moderately thermophilic mercury resistant bacteria were selected by growth at 62 °C on Luria agar containing HgCl2. Sequence analysis of 16S rRNA genes of two isolates showed the closest matches to be with Bacillus pallidus and Ureibacillus thermosphaericus. Minimum inhibitory concentration (MIC) values for HgCl2 were 80 μg/ml and 30 μg/ml for these isolates, respectively, compared to 10 μg/ml for B. pallidus H12 DSM3670, a mercury-sensitive control. The best-characterised mercury-resistant Bacillus strain, B. cereus RC607, had an MIC of 60 μg/ml. The new isolates had negligible mercuric reductase activity but removed Hg from the medium by the formation of a black precipitate, identified as HgS by X-ray powder diffraction analysis. No volatile H2S was detected in the headspace of cultures in the absence or presence of Hg2+, and it is suggested that a new mechanism of Hg tolerance, based on the production of non-volatile thiol species, may have potential for decontamination of solutions containing Hg2+ without production of toxic volatile H2S.  相似文献   

15.
Interaction between D-glucuronic acid and Zn(II), Cd(II), and Hg(II) metal ion salts has been studied in solution and solid complexes of the type M(D-glucuronate)X · nH2O and M(D-glucuronate)2·nH2O, where M = Zn(II), Cd(II), and Hg(II), X = Cl or Br, and n = 0–2 were isolated and characterized. Spectroscopic and other evidence indicated that in the metal-halide-sugar complexes the Zn(II) and Cd(II) ions bind to two D-glucuronate moieties via 06, 05 of the carboxyl oxygen atoms of the first and 04, 06' of hydroxyl and carbonyl groups of the second as well as to two H2O molecules, whereas in the corresponding M(D-glucuronate)2 · nH2O salts, the metal ions are bonded to two sugar anions through 06 and 06' of the ionized carboxyl groups and two water molecules, resulting in a six-coordination around each metal cation. The Hg(II) ion binds to 06 and 05 oxygen atoms of a sugar anion and to a halide anion or water molecule, in the Hg(D-glucuronate)X·nH2O compounds, while in the corresponding metal-glucuronate salt mercury is bonded to 06 and 06' of the two glucuronate anions with four-coordination around the Hg(II) ion. The β-anomer sugar conformation is predominant in the free acid and in these series of metal-sugar complexes.  相似文献   

16.
Abstract

Hg(II) and Pb(II) tolerance, removal, bioaccumulation and effects on antioxidant enzymes of thermophilic Exiguobacterium profundum were investigated. The results indicated that Hg(II) was more toxic than Pb(II) to E. profundum. E. profundum was also more tolerant in solid medium than in liquid medium for Pb(II) and Hg(II). The bacterial growth was not significantly influenced at 1.0 and 2.5?mg/L Pb(II) and Hg(II) for 24?h. The highest Hg(II) and Pb(II) bioaccumulation amounts were determined as 37.56 and 54.35?mg metal/dried bacteria, respectively. Bioaccumulation capacities of the cell membrane of E. profundum for Hg(II) and Pb(II) were determined. The different concentrations of Pb(II) and Hg(II) enhanced the SOD and CAT enzymes. In addition, variations of the surface macrostructure and the functionality of E. profundum after the interaction with Hg(II) and Pb(II) were investigated by the scanning electron microscope (SEM) and the Fourier transform infrared spectroscopy (FT-IR), respectively.

This investigation obviously showed that thermophilic E. profundum can also be applied for removal and recovery of toxic metals from industrial wastewater. Clearly, a further investigation should be utilized by thermophilic microorganisms. According to antioxidant enyzme activities, E. profundum can be also used as a bioindicator for the detection of toxic metal pollution in natural water samples.  相似文献   

17.
The rotational spectrum of yeast cells changed after pre-treatment of the cells with HgCl2 or Hg(NO3)2 and became indistinguishable from that of ultrasonically produced cell walls. The spectrum of the affected cells contained a peak which could only be explained by attributing a conductivity to the cell walls that was higher than that of the medium. Theoretical models of the rotational response are fully in accord with the experimental spectra. It is shown that the rotation method is capable of measuring even the low cell wall conductivity of yeast cells (which was found to be 33 μS/cm at 10 μS/cm medium conductivity). Knowledge of the spectra allowed a field frequency to be selected at which untreated cells showed no rotation, but at which cells affected by treatment with Hg(II) identified themselves by rotating in the same direction as the field. Calculation of the percentage of cells showing this co-field rotation gave an index (termed the co-field rotation value) of the proportion of the cells that were affected. Using this technique, effects of 25 nmol/l Hg(II) could be demonstrated. In media of low conductivity (10 μS/cm) the change in the rotational spectrum was usually ‘all-or-none’, whereas at 200 μS/cm a graded Hg(II)-mediated change became apparent. The co-field rotation method showed that the action of small quantities of Hg(II) was still increasing after 3 h of incubation and paralleled the Hg(II)-induced K+ release. A rapid reduction of the effects of Hg(II) was seen when 3–30 mM K+ (or Na+) or when 1 mM Ca2+ were present in the incubation medium, or as the pH was increased. At high incubation cell concentrations the toxic effect of Hg(II) was reduced, apparently due to binding by the cells.  相似文献   

18.
HeLa S3 cells were exposed to varied concentrations of methylmercury over varied periods of time and its binding by the cells was studied using 203Hg-labeled methylmercuric chloride as radioactive marker. Also studied was the effect of cell-bound methylmercury on DNA replication and protein synthesis and on the growth rate of the cells. The results show that methylmercury binding is a rapid process, with much of the organomercurial bound within the the first 60 min of incubation, and that considerable quantities of organic mercury become affixed to the cells. The amounts of bound methylmercury, [CH3Hg(II)]bound, given in mol/cell, range from 2 × 10?16 (at 1 h of incubation and at 1 μM CH3Hg(II) in the medium) to almost 4 × 10?14 (at 24 h of incubation and at 100 μM CH3Hg(II) in the medium). A [CH3Hg(II)]bound value of about 30 × 10?16 mol/cell appears to be the threshold below which cells display a normal growth pattern and below which metabolic events such as DNA replication or protein synthesis are affected only to a minor degree but above which major changes in cell metabolism and cell growth take place. Methylmercury binding by the cells is tight so that only 20% of the bound material is released from the cells over a 3-h incubation period when the cells are placed into fresh, methylmercury-free growth medium. Analysis of the binding data in terms of binding to identical and completely independent sites yields an association constant K of 7.92 × 104 l/mol and for the maximum concentration of cellular binding sites the value 2.40 × 10?14 mol/cell or 1.45 × 1010 sites/cell. Evidence is presented which shows that cellular sulfhydryl groups do not suffice to provide all the sites taken up by methylmercury and that binding, in all likelihood, involves basic nitrogen, too. The levels of cell-bound methylmercury are such that binding to HeLa DNA and HeLa chromatin, for instance, can readily take place. Methylmercury binding data obtained by using the technique of particle-induced X-ray emission (PIXE) are in good agreement with the data obtained via isotope dilution.  相似文献   

19.
The reactions of MCl2 (M = Zn2+, Cd2+, Hg2+) with 2-(α-hydroxy-benzyl)thiamin pyrophosphate (HBTPP) at various pH values (different protonation states) were studied in methanolic solutions. Solid complexes of formulae K[Zn(HBTPP) Cl2 · H2On, K2[Cd(HBTPP)2−Cl2 · 3H2On, K2[Hg(HBTPP)2Cl2 · 3H2O and Zn(HBTPP)20Cl2 were isolated and characterized by elemental analysis and various NMR techniques, namely 13C NMR, 31P NMR, 113Cd NMR, 199Hg NMR and 1H NMR ROESY spectra in D2O. The data provide evidence that Zn(II) in K[Zn(HBTPP) Cl2 · H2On, and Cd(II) in K2[Cd(HBTPP)2Cl2 · 3H2On, are coordinated both to the pyrimidine N(1′) and to the pyrophosphate group. In contrast, Hg(II) in K2[Hg(HBTPP)2Cl2 · 3H2O and Zn(II) in Zn(HBTPP)20Cl2 are bound only to the N(1′) atom or to the pyrophosphate group, respectively.  相似文献   

20.
Several Lupinus species, for example, Lupinus albus, Lupinus luteus, Lupinus angustifolius, and Lupinus hispanicus were used to accumulate Mn(II), Cd(II), Pb(II), Cr(III), Cr(VI), Hg2+, and CH3Hg+ from waste waters. The influence of different species concentrations (50 and 100 mg L-1) and pH on growing behavior as well as the resulting distribution of metals in the plants were investigated. The results obtained showed that lupins were able to germinate and to grow in the presence of the metals mentioned above, even when they were present at levels as high as 50 mg L-1. Accumulation of Pb(II), Cr(III), and Cd(II) was higher in roots than in shoots. As far as mercury is concerned, the highest CH3Hg and Hg2+ accumulation was detected in roots, but fast transport toward the leaves was noticed. In contrast to mercury, the uptake of chromium seems to be influenced by the chemical form of the analyte, remaining Cr(VI) in solution. No differences in growing behavior and accumulation were observed for the four Lupinus species studied. Even though plants were exposed only a relatively short time to the metal solutions, metal concentrations of approximately 2 g/kg of dry matter were detected in the young lupins plants. The feasibility of utilizing Lupinus plants for the removal of heavy metals from wastewater was also investigated. Lupins were able to grow under extreme conditions (wastewater, pH lower than 2) and to remove 98% of the initial amount of toxic metals present in the sample.  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