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.
When methylmercury was incubated in the presence of selenite and reduced glutathione (GSH), the mercury which was extracted into benzene under acidic condition decreased gradually with the elapse of time. This decrease was due to the cleavage of mercury-carbon bond of methylmercury. The reaction did not proceed when selenite or GSH was singly added to the reaction mixture. L-Cysteine, 2-mercaptoethanol and sodium sulfide in place of GSH also were effective for decomposition of methylmercury in combination with selenite, but oxidized glutathione (GSSG) and L-cystine were not. This suggests that reduction of selenite is needed for the degradation of methylmercury. Thus, the effect of reduced metabolites of selenite produced by GSH was investigated. Glutathione selenotrisulfide (GSSeSG) requierd GSH for the degradation of methylmercury, whereas H2Se possessed a strong activity even in the absence of GSH. This may indicate that H2Se is involved directly in the conversion of methylmercury to inorganic mercury. This phenomenon found in in vitro experiments is discussed in relation to the biotransformation of methylmercury.  相似文献   

2.
Luminescent lanthanide (III) ions have been exploited for circularly polarized luminescence (CPL) for decades. However, very few of these studies have involved chiral samarium (III) complexes. Complexes are prepared by mixing axial chiral ligands (R/S))‐2,2’‐bis(diphenylphosphoryl)‐1,1′‐binaphthyl (BINAPO) with europium and samarium Tris (trifluoromethane sulfonate) (Eu (OTf)3 and Sm (OTf)3). Luminescence‐based titration shows that the complex formed is Ln((R/S)‐BINAPO)2(OTf)3, where Ln = Eu or Sm. The CPL spectra are reported for Eu((R/S)‐BINAPO)2(OTf)3 and Sm((R/S)‐BINAPO)2(OTf)3. The sign of the dissymmetry factors, gem, was dependent upon the chirality of the BINAPO ligand, and the magnitudes were relatively large. Of all of the complexes in this study, Sm((S)‐BINAPO)2(OTf)3 has the largest gem = 0.272, which is one of the largest recorded for a chiral Sm3+ complex. A theoretical three‐dimensional structural model of the complex that is consistent with the experimental observations is developed and refined. This report also shows that (R/S)‐BINAPO are the only reported ligands where gem (Sm3+) > gem (Eu3+).  相似文献   

3.
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   

4.
Accurate measurement of sapwood depth (D S) is essential for calculating volumetric water use of individual trees and stands. Various methods are available to measure D S but their accuracy is rarely cross-validated. We sampled 15 Eucalyptus and 1 Corymbia species along a gradient of aridity and obtained reference values of D S in fresh wood cores using light microscopy, which represents our reference method. We compared this method to the simpler and widely used macroscopic method: visual assessment of natural or induced colour change from sapwood to heartwood. In a third method, estimation of D S was based on species-specific models that rely on wood properties measured using near infrared spectroscopy (NIR). Microscopy allowed clear identification of D S based on the presence of blocked vessels. Measurement of D S using microscopic methods was possible for 78 of a total of 80 cores and ranged from 3.6 mm (E. loxophleba) to 43.8 mm (E. viminalis). Macroscopic assessment clearly differentiated sapwood and heartwood in 60 cores. Results from microscopic and macroscopic methods agreed closely (<10% deviation between estimates) in 35 of 78 cores. After elimination of clearly erroneous measurements (>50% deviation between estimates), macroscopic measurement across all species agreed well with microscopic assessment of D S (R 2 = 0.92). Models developed for differentiation between sapwood and heartwood using NIR spectroscopy were very robust (high coefficient of determination) for four species, but D S could only be predicted well for one (E. obliqua) of the four species. Even after elimination of apparent false estimates, prediction of D S by NIR across species was not as strong as for macroscopic assessment (R 2 = 0.88). D S can accurately be measured using microscopy if vessel occlusion is clearly visible. Although slightly overestimated, D S from macroscopic assessment was generally similar to that measured by microscopy. NIR spectroscopy was unable to predict D S with acceptable accuracy for the majority of species. Further improvements in the prediction of D S using NIR will require more intensive model calibration and validation, and may not be applicable to all species.  相似文献   

5.
A series of twelve novel diamminetetrakis(carboxylato)platinum(IV) and 18 novel bis(carboxylato)dichlorido(ethane‐1,2‐diamine)platinum(IV) complexes with mixed axial carboxylato ligands was synthesized and characterized by multinuclear 1H‐, 13C‐, 15N‐, and 195Pt‐NMR spectroscopy. Their cytotoxic potential was evaluated (by MTT assay) against three human cancer cell lines derived from ovarian teratocarcinoma (CH1/PA‐1), lung (A549), and colon carcinoma (SW480). In the cisplatin‐sensitive CH1/PA‐1 cancer cell line, diamminetetrakis(carboxylato)platinum(IV) complexes showed IC50 values in the low micromolar range, whereas, for the most lipophilic compounds of the bis(carboxylato)dichlorido(ethane‐1,2‐diamine)platinum(IV) series, IC50 values in the nanomolar range were found.  相似文献   

6.
The interaction of selenium with methylmercury was investigated in brain of animals labeled with 75SeO32? and CH3203Hg+. Brains were fractionated into subcellular components and the cytosol was further fractionated by chromatography on Sephadex G-150 and G-200. The main result of these studies was evidence suggesting a shift of 75Se from the cytosol to the mitochondrial fraction in brain when CH3Hg+ was given. Concurrent equimolar (10 μmoles/kg) selenite injections increased the uptake of Hg but did not alter 203Hg distribution in brain. Changing the dose of CH3Hg+ from 1 to 38 μmoles/kg had little effect on Hg uptake (% of dose per g). Gel filtrations on Sephadex G-150 and G-200 revealed that 203Hg in cytosol followed a pattern more closely related to protein (A280) than to 75Se, although a considerable portion of both isotopes eluted with proteins in the void volume. Assays of whole brain homogenates revealed a slight reduction in glutathione peroxidase activity in CH3Hg+-treated rats which was not seen when equimolar selenite was injected with the CH3Hg+.  相似文献   

7.
Besides the open circuit voltage (VOC) deficit, fill factor (FF) is the second most significant parameter deficit for earth‐abundant kesterite solar cell technology. Here, various pathways for FF loss are discussed, with focus on the series resistance issue and its various contributing factors. Electrical and physical characterizations of the full range of bandgap (Eg = 1.0–1.5 eV) Cu2ZnSn(SxSe1?x)4 (CZTSSe) devices, as well as bare and exfoliated films with various S/(S + Se) ratios, are performed. High intensity Suns‐VOC measurement indicates a nonohmic junction developing in high bandgap CZTSSe. Grazing incidence X‐ray diffraction, Raman mapping, field emission scanning electron microscopy, and X‐ray photoelectron spectroscopy indicate the formation of Sn(S,Se)2, Mo(S,Se)2, and Zn(S,Se) at the high bandgap CZTSSe/Mo interface, contributing to the increased series resistance (RS) and nonohmic back contact characteristics. This study offers some clues as to why the record‐CZTSSe solar cells occur within a bandgap range centered around 1.15 eV and offers some direction for further optimization.  相似文献   

8.
31P NMR has been employed to study the interaction between zinc(II) bis(O,O′-di-iso-butyldithiophosphate), Zn[S2P(OiBu)2]2, and four multidentate amines (diethylenetriamine, triethylenetetramine, tetraethylenepentamine and pentaethylenehexamine) in chloroform at 294 K. The major interaction of Zn[S2P(OiBu)2]2 and these polyamines involves displacement of the {S2P(OiBu)2} ligands from the zinc giving [Zn(amine)]2+ and [S2P(OiBu)2] ions in solution. The magnitudes of the equilibrium constants, K1 (=[{Zn(amine)}2+][{DDP}]2/[Zn(DDP)2][amine]), have been evaluated in the cases of triethylenetetramine (20.0 l mol−1), tetraethylenepentamine (19.1 l mol−1) and pentaethylenehexamine (1.58 l mol−1). Crystalline 1:1 ionic complexes have also been isolated from these systems and characterised.  相似文献   

9.
To develop the potential of plant for phytoremediation of methylmercury pollution, a genetically engineered tobacco plant that coexpresses organomercurial lyase (MerB) with the ppk-specified polyphosphate (polyP) and merT-encoding mercury transporter was constructed by integrating a bacterial merB gene into ppk/merT-transgenic tobacco. A large number of independent transgenic tobaccos was obtained, in some of which the merB gene was stably integrated in the plant genome and substantially translated to the expected MerB enzyme in the transgenic tobacco. The ppk/merT/merB-transgenic tobacco callus showed more resistance to methylmercury (CH3Hg+) and accumulated more mercury from CH3Hg+-containing medium than the ppk/merT-transgenic and wild-type progenitors. These results suggest that the MerB enzyme encoded by merB degraded the incorporated CH3Hg+ to Hg2+, which then accumulated as a less toxic Hg-polyP complex in the tobacco cells. Phytoremediation of CH3Hg+ and Hg2+ in the environment with this engineered ppk/merT/merB-transgenic plant, which prevents the release mercury vapor (Hg0) into the atmosphere in addition to generating potentially recyclable mercury-rich plant residues, is believed to be more acceptable to the public than other competing technologies, including phytovolatilization.  相似文献   

10.
The accumulation of methylmercury (MeHg) to an oligochaete worm Lumbriculus variegatus (Müller) was measured in two different lake sediments in the laboratory. 14C-labelled MeHg was added to sediments at the nominal concentration of 95 ng/g dw sediment. Groups of six oligochaete worms were exposed in glass beakers to 35 g of spiked sediment for 14 days. The two sediments had organic carbon concentrations of 3.4% and 9.9% and natural selenium concentrations of 1.45 and 0.28 mg/kg (dw), respectively. After two weeks exposure, both the accumulation rate of MeHg and the body residue in the worms were much lower in the sediment having a high organic carbon content. The effect of selenium concentration in the sediment on bioaccumulation of MeHg in Lumbriculus variegatus was measured in one sediment (organic carbon 3.4% and Se 1.45 mg/kg) by adding sodiumselenite (Na2SeO3) at different concentrations. The added amounts of selenium were 0, 0.1, 0.5, 2.5, 15.0, and 50.0 mg Se/kg dry sediment. In this exposure the nominal concentration of MeHg was 102 ng/g dw sediment. The two lowest selenium concentrations did not affect the bioaccumulation of MeHg. But, the dose of 2.5 mg Se/kg resulted in a 25% reduction in the body residue after two weeks exposure. When 15 and 50 mg Se/kg were added to the sediment the accumulation of MeHg in the organisms was decreased by 75% and 86%, respectively, as compared to the reference.  相似文献   

11.
The effectiveness of eight thiol ligands for removing methylmercury (CH3Hg(II)) from its glutathione and hemoglobin complexes in hemolyzed erythrocytes has been studied by 1H nuclear magnetic resonance spectroscopy. These complexes are the predominant methylmercury species in human erythrocytes. The effectiveness was determined from the exchange-averaged chemical shift of the resonance for the proton on the α-carbon of the cysteinyl residue and from the intensity of the resonance for the methylene protons of the glycine residue of reduced glutathione (GSH), both of which provide a measure of the amount of glutathione in the CH3Hg(II)-complexed form. The thiol ligands were found to release GSH from its CH3Hg(II) complex in the order 2, 3-dimercap-tosuccinic acid > mercaptosuccinic acid > cysteine > mercaptoacetic acid > D-penicillamine > 2, 3-dimercaptopropanesulfonic acid > N-acetyl-D,L-penicillamine > D.L-homocysteine.  相似文献   

12.
Copper and other transition metal ions and their complexes are catalysts for the decomposition of nitrosothiols. In this way they catalyze the biological functions of nitrosothiols. The kinetics and mechanism of the reaction of two nitrosothiols, S-nitrosothiolactic acid and S-nitrosoglutathione (GSNO), with copper(I) are reported. The kinetics of the reaction of Cu(MeCN) n + (n=0–3) with the nitrosothiols were studied. The results indicate that Cu+ aq is the active species in the GSNO system, with k(Cu+ aq+GSNO)=(9.4 ±2.0)×107 dm3 mol−1 s−1 . The results also indicate that the Cu(MeCN) n + (n=0–3) complexes react with S-nitrosothiolactic acid. Transient species are formed in these processes. The results suggest that these species contain copper(I) and thiol. The results shed light on the catalytic role of copper complexes in the decomposition of S-nitrosothiols. Received 10 April 1999 / Accepted 17 December 1999  相似文献   

13.
The optically active mixed‐ligand fac(S)‐tris(thiolato)rhodium(III) complexes, ΔLfac(S)‐[Rh(aet)2(L‐cys‐N,S)]? (aet = 2‐aminoethanethiolate, L‐cys = L‐cysteinate) ( 1 ) and ΔLLfac(S)‐[Rh(aet)(L‐cys‐N,S)2]2? were newly prepared by the equatorial preference of the carboxyl group in the coordinated L‐cys ligand. The amide formation reaction of 1 with 1,10‐diaminodecane and polyallylamine gave the diamine‐bridged dinuclear Rh(III) complex and the single‐chain polymer‐supported Rh(III) complex with retention of the ΔL configuration of 1 , respectively. These Rh(III) complexes reacted with Co(III) or Co(II) to give the linear‐type trinuclear structure with the S‐bridged Co(III) center and the two Δ‐Rh(III) terminal moieties. The polymer‐supported Rh(III) complex was applied not only to the CD spectropolarimetric detection and determination of a trace of precious metal ions such as Au(III), Pt(II), and Pd(II) but also to concentration and extraction of these metal ions into the solid polymer phase. Chirality 28:85–91, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

14.
Strain MPA‐C3 was isolated by incubating arsenic‐bearing sediments under anaerobic, mesophilic conditions in minimal media with acetate as the sole source of energy and carbon, and As(V) as the sole electron acceptor. Following growth and the respiratory reduction of As(V) to As(III), a yellow precipitate formed in active cultures, while no precipitate was observed in autoclaved controls, or in uninoculated media supplemented with As(III). The precipitate was identified by X‐ray diffraction as alacranite, As8S9, a mineral previously only identified in hydrothermal environments. Sequencing of the 16S rRNA gene indicated that strain MPA‐C3 is a member of the Deferribacteres family, with relatively low (90%) identity to Denitrovibrio acetiphilus DSM 12809. The arsenate respiratory reductase gene, arrA, was sequenced, showing high homology to the arrA gene of Desulfitobacterium halfniense. In addition to As(V), strain MPA‐C3 utilizes NO3?, Se(VI), Se(IV), fumarate and Fe(III) as electron acceptors, and acetate, pyruvate, fructose and benzoate as sources of carbon and energy. Analysis of a draft genome sequence revealed multiple pathways for respiration and carbon utilization. The results of this work demonstrate that alacranite, a mineral previously thought to be formed only chemically under hydrothermal conditions, is precipitated under mesophilic conditions by the metabolically versatile strain MPA‐C3.  相似文献   

15.
Resistance to coffee leaf rust is conferred by SH3, a major dominant gene that has been introgressed from a wild coffee species Coffea liberica (genome L) into the allotetraploid cultivated species, Coffea arabica (genome CaEa). As the first step toward the map-based cloning of the SH3 gene, using a bacterial artificial chromosome (BAC) library, we describe the construction of a physical map in C. arabica spanning the resistance locus. This physical map consists in two homeologous BAC-contigs of 1,170 and 1,208 kb corresponding to the subgenomes Ca and Ea, respectively. Genetic analysis was performed using a single nucleotide polymorphism detection assay based on Sanger sequencing of amplicons. The C. liberica-derived chromosome segment that carries the SH3 resistance gene appeared to be introgressed on the sub-genome Ca. The position of the SH3 locus was delimited within an interval of 550 kb on the physical map. In addition, our results indicated a sixfold reduction in recombination frequency in the introgressed SH3 region compared to the orthologous region in Coffea canephora.  相似文献   

16.
Conformations of disulfide and diselenide were compared in (Boc‐Cys/Sec‐NHMe)2 and (Boc‐Cys/Sec‐OMe)2 using X‐ray crystallography, nuclear magnetic resonance (NMR) spectroscopy, density functional theory (DFT), and circular dichroism (CD) spectroscopy. Conformations of disulfide/diselenide in polypeptides are defined based on the sign of side chain torsion angle χ3 (–CH2–S/Se–S/Se–CH2–); negative indicates left‐handed and positive indicates right‐handed orientation. In the crystals of (Boc‐Cys‐OMe)2 and (Boc‐Sec‐OMe)2, the disulfide exhibits a left‐handed and the diselenide a right‐handed orientation. Characterization of cystine and selenocystine derivatives in solution using 1H‐NMR, natural abundant 77Se NMR, 2D‐ROESY, and chemical shift analysis coupled to DMSO titration has indicated the symmetrical nature and antiparallel orientation of Cys/Sec residues about the disulfide/diselenide bridges. Structural calculations of cystine and selenocystine derivatives using DFT further support the antiparallel orientation of Cys/Sec residues about disulfide/diselenide. The far‐ultraviolet (UV) region CD spectra of cystine and selenocystine derivatives have exhibited the negative Cotton effect (CE) for disulfide and positive for diselenide confirming the difference in the conformational preference of disulfide and diselenide. In the previously reported polymorphic structure of (Boc‐Sec‐OMe)2, the diselenide has right‐handed orientation. In the X‐ray structures of disulfide and diselenide analogues of Escherichia coli protein encoded by curli specific gene C (CgsC) retrieved from Protein Databank (PDB), disulfide has left‐handed and the diselenide right‐handed orientation. The current report provides the evidence for the local conformational difference between a disulfide and a diselenide group under unconstrained conditions, which may be useful for the rational replacement of disulfide by diselenide in polypeptide chains.  相似文献   

17.
The binding of the methylmercury cation CH3Hg+ by poly(L -glutamic acid) (PGA) and by poly(L -lysine) (PLL) has been investigated by Raman spectroscopy. Coordination on the side-chain COO? and NH groups of these polypeptides gave characteristic ligand–Hg stretching modes at ca. 505 and 450 cm?1, respectively. Precipitation generally occurred upon formation of the complexes and changes of conformation were common. The solid complex obtained from PGA at pH 4.6 was found to have a mostly disordered conformation, which differed from the respective α-helical and β-sheet structures of the dissolved and precipitated uncomplexed polypeptide in the same conditions. An α-helical structure was generally adopted by the complex formed with PLL, even in pH and temperature conditions where the free polypeptide normally exists in another conformation. The addition of a stronger complexing agent, glutathione, to the PLL/CH3Hg+ complex caused a migration of the bound cations and a restoration of the polypeptide to its original state.  相似文献   

18.
《Inorganica chimica acta》1986,111(2):171-178
The structures of solvated methylmercury(II) halides in pyridine solution were determined by a large angle X-ray scattering technique. Near-linear CH3HgX (X = Cl, Br and I) species solvated by two weakly-coordinated pyridine molecules are indirectly interpreted. Additional mercury-pyridine interactions, through van der Waals forces, are found at the sum of the van der Waals radii. The HgX bond distances in the methylmercury(II) halides are found to be 2.325(8), 2.480(3) and 2.649(3) Å for chloride, bromide and iodide, respectively. The HgC bond distances are assumed to be ∼2.08 Å. This interaction is indicated in the radial distribution functions. The bond distance between mercury and the two solvating pyridine molecules is ∼2.8 Å, e.g., 2.84(2) Å in methylmercury(II) bromide. The additional mercury interactions with roughly two pyridine molecules at the sum of van der Waals radii are revealed at around 3.15 Å. Comparison between Raman stretching vibrations and the solvated structures of methylmercury(II) complexes found in various solvents indicates a lower limit in solvent donor property for the formation of solvate bonds to mercury for the methylmercury(II) halides.  相似文献   

19.
Scheuhammer  A. M.  Blancher  P. J. 《Hydrobiologia》1994,279(1):445-455
Piscivorous birds and mammals in areas remote from point sources of Hg contamination may be exposed to dietary methylmercury concentrations that are sufficiently high to cause reproductive impairment. Common loons (Gavia immer) were observed to show aberrant nesting behavior and low overall reproductive success when Hg concentrations in prey (small fish and crayfish) averaged > 0.3 µg g–1 wet weight (Barr, 1986), levels known to occur in fish from many lakes in central Ontario. We used data on Hg in Ontario fish to estimate the proportion of lakes where fish small enough for loons to eat (< 250 g) had Hg concentrations that exceeded estimated thresholds for reproductive impairment. Up to 30 % of lakes exceeded thresholds for reproductive impairment, depending on the species of fish and the threshold Hg concentrations chosen. There was a significant negative correlation between fish-Hg concentration and lake pH in most fish species examined. For these species, reductions in sulfate deposition rates are predicted to result in a corresponding reduction of lakes in Ontario having fish with potentially toxic concentrations of Hg.  相似文献   

20.
The interest for lanthanide circularly polarized luminescence (CPL) has been quickly growing for 10 years. However, very few of these studies have involved correlation between the dissymmetry factor (glum) and the chemical modifications in a series of chiral ligands. Four polymeric compounds of Eu(III) were prepared by using a series of binaphtyl derivatives for which the size of the π system as well as the number of stereogenic elements (i.e., the binaphtyl moiety) are modulated. The resulting {[Eu(hfac)3((S)/(R)-Lx)]}n (x = 1 and 3) and {[Eu(hfac)3((S,S,S)/(R,R,R)-Lx)]}n (x = 2 and 4) have been characterized by powder X-ray diffraction by comparison with the X-ray structures on single crystal of the Dy(III) analogs. In solution, the structure of the complexes is deeply modified and becomes monomeric. The nature of the ligand induces change in the shape of the CPL spectra in CH2Cl2 solution. Furthermore, a large |glum| = 0.12 of the magnetic-dipole transition for the [Eu(hfac)3((S,S,S)/(R,R,R)-L2)] complex involving the ligand with three stereogenic elements and an extended ?? system has been measured. This report also shows CPL measurements in solid state for the series of {[Eu(hfac)3((S)/(R)-Lx)]}n (x = 1 and 3) and {[Eu(hfac)3((S,S,S)/(R,R,R)-Lx)]}n (x = 2 and 4) polymers.  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