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.
Reaction of (PhMe2P)2PtMe2 or [(κ2-P,N)-Ph2PC2H4NMe2]PtMe2 with an excess of H2SnBu2 or H2SnPh2 resulted in the catalytic formation of cyclo-, oligo- and/or polystannanes. In the reaction of (PhMe2P)2PtMe2 with H2SnBu2, linear oligomeric species H(SnBu2)nH were observed in the initial stage of the reaction, which eventually converted into cyclostannanes. Only polystannanes were observed in the reaction of [(κ2-P,N)-Ph2PC2H4NMe2]PtMe2 with H2SnBu2. The reactions of H2SnPh2 were similar, but more difficult to analyze due to redistribution reactions and the formation of insoluble products. The mechanism of the reactions is clearly different to that previously observed for HSnR3 because metal complexes indicative of oxidative addition/reductive elimination reactions were only observed as minor products.  相似文献   

2.
The chemical decontamination of infected dental implants is essential for the successful treatment of peri-implantitis. The aim of this study was to assess the antibacterial effect of a hydrogen peroxide-titanium dioxide (H2O2–TiO2) suspension against Staphylococcus epidermidis biofilms. Titanium (Ti) coins were inoculated with a bioluminescent S. epidermidis strain for 8 h and subsequently exposed to H2O2 with and without TiO2 nanoparticles or chlorhexidine (CHX). Bacterial regrowth, bacterial load and viability after decontamination were analyzed by continuous luminescence monitoring, live/dead staining and scanning electron microscopy. Bacterial regrowth was delayed on surfaces treated with H2O2–TiO2 compared to H2O2. H2O2-based treatments resulted in a lower bacterial load compared to CHX. Few viable bacteria were found on surfaces treated with H2O2 and H2O2–TiO2, which contrasted with a uniform layer of dead bacteria for surfaces treated with CHX. H2O2–TiO2 suspensions could therefore be considered an alternative approach in the decontamination of dental implants.  相似文献   

3.
《Inorganica chimica acta》1986,113(2):131-135
Treatment of Al2iBu4 with THF and Et2O results in partial decomposition to afford Al metal while reaction with γ-picoline products a bis-adduct of limited stability. Reactions with AlMe3 and BCl3, separately, involves ligand exchange with accompanying disproportionation to yield Al metal. Dimethylamine induces disproportionation to afford AliBu3·HNMe2 and an intemediate trialuminum species. The latter undergoes AlAl bond cleavage with formation of H2, iBu2AlNMe2, and [Me2N- (iBu)AlAliBu2]2· Al2iBu4 eliminates Me2CCH2 in solution at 80 °C, and the catenated AlH intermediate reacts with ethylene to afford AlEt and AlCH2CH2CH2CH3 moieties.  相似文献   

4.
Summary Six early successional plant species with differing photosynthetic pathways (3 C3 species and 3 C4 species) were grown at either 300, 600, or 1,200 ppm CO2 and at either 0.0 or 0.25 ppm SO2. Total plant growth increased with CO2 concentration for the C3 species and varied only slightly with CO2 for the C4 species. Fumigation with SO2 caused reduced growth of the C3 species at 300 ppm CO2 but not at the higher concentrations of CO2. Fumigation with SO2 reduced growth of the C4 species at high CO2 and increased growth at 300 ppm CO2. Leaf area increased with increasing CO2 for all plant species. Fumigation with SO2 reduced leaf area of C3 plants more at low CO2 than at high CO2 while leaf area of C4 plants was reduced more at high CO2 than at low CO2. These results support the notion that C3 species are more sensitive to SO2 fumigation than are C4 species at concentrations of CO2 equal to that found in normal ambient air. However, the difference in sensitivity to SO2 between C3 and C4 species was found to be reversed at higher concentrations of CO2. A possible explanation for this reversal based upon differences in stomatal response to elevated CO2 between C3 and C4 species is discussed.  相似文献   

5.
《Inorganica chimica acta》1986,122(2):207-211
Treatment of [M(CO)4Ph2PCHPPh2] with CH3- OCH2Cl at 20 °C gave the methoxymethyl derivations [M(CO)4{Ph2PCH(CH2OCH3)PPh2}] (MCr or W), but a similar treatment at 80 °C gave derivatives of a vinylidene diphosphine [M(CO)4(Ph2P)2C CH2]. Treatment of [M(CO)4Ph2PCHPPh2]with CH3CHClOCH3 at 20 or 80 °C gave only [M(CO)4- (Ph2P)2CHCH(CH3)OCH3] (MCr or W). The vinylidene diphosphine complexes [M(CO)4(Ph2P)2- CCH2] (MCr, Mo or W) were even more easily prepared by treating [M(CO)6] with (Ph2P)2CCH2 (vdpp) in hot solvents such as CH3OCH2CH2OCH2- CH2OCH3.Treatment of [W(CO)4vdpp] with LiBun followed by methanol gave [W(CO)4(Ph2P)2CHCH2Bun] (1c), i.e. conjugate addition to the CCH2 occurs. 1c was also made by treating [W(CO)4(Ph2P)2CH] with n-pentyl-iodide. Similarly LiMe was added to [W(CO)4(Ph2P)2CCH2]. Treatment of [M(CO)4- vdpp] with NaCH(COOEt)2 gave [M(CO)4(Ph2- P)2CHCH2CH(COOEt)2] (MW or Mo). Pyrrolidine added to the CCH2 bonds of [M(CO)4vddp] to give [M(CO)4(Ph2P)2CHCH2NC4H8]. 31p and 1H NMR and IR data are given.  相似文献   

6.
A rabbit antiserum directed toward the prostaglandin E2 metabolite 13,14-dihydro-15-keto-prostaglandin E2 (KH2PGE2) was produced by immunization with a human albumin-KH2PGE2 conjugate. The antiserum recognized the 15-keto-group (it cross reacts with 13,14-dihydro-prostaglandin E2 0.2%); the saturated 13,14-bond (it cross reacts with 15-keto-prostaglandin E2 7%); the 9-keto group (it cross reacts with 13,14-dihydro-15-keto-prostaglandin F 5%); and the 11-hydroxy group (it cross reacts with 13,14-dihydro-15-keto-PGA2 0.4%).By subjecting the antiserum to preparative isoelectric electrofocusing, populations of antibodies that varied in their cross reaction with 13,14-dihydro-15-keto-prostaglandin F (KH2PGF) from 20% to 1% were obtained. The levels of KH2PGE2 in plasma of rat and mouse as measured by radioimmunoassay of the unfractionated plasma were 0.39 ± 0.07 ng/ml and 0.41 ± 0.13 ng/ml, respectively. Recovery of exogenously added KH2PGE2 from human plasma was 100%. Radioimmunoassay with two antisera; an antiserum directed toward KH2PGF that cross reacts with KH2PGE2 1% and the antiserum to KH2PGE2, demonstrated that KH2PGE2, not KH2PGF, was being measured with the anti-KH2PGE2. The levels of KH2PGE2 in rat plasma did not vary with sex. In rats, the levels of KH2PGE2 markedly increased after exercise stress.In mice carrying a spindle-cell sarcoma (SAI) and a fibrosarcoma (SaD2), the levels of KH2PGE2 in the plasma increased with time after transplantation. The increase was not observed in the plasma of mice carrying a transplantable anaplastic carcinoma (15091AK), a lymphatic leukemia (AW5147), two mammary adenocarcinomas (CADI, CAD2), a myeloid leukemia (C1498), and a hepatoma (BW7756).  相似文献   

7.
We investigated the reaction mechanism and thermochemical property of conjugated dienes or mono-olefins with nickel dithiolenes (Ni(S2C2R2)2) using density functional theory. The reactions between conjugated dienes and nickel dithiolenes are concerted reactions. The thermochemical property study shows that the introduction of electron-withdrawing groups (–CF3 or –CN) to nickel dithiolene (Ni(S2C2H2)2) not only significantly lowers the activation energy barrier but also strongly stabilises the products. The introduction of electron-donating group (–CH3) to butadiene has the same effect. So, we conclude that the reactions between nickel dithiolenes and conjugated dienes are electrophilic cycloaddition. Mono-olefins can add to nickel dithiolenes through interligand pathway, which is a two-step process or through intraligand pathway, which is a one-step process. The thermochemical property study shows that the activation enthalpy for the reaction of butadiene with Ni(S2C2(CF3)2)2 is much lower than those of C4 mono-olefins with Ni(S2C2(CF3)2)2 for both interligand addition and intraligand addition. The Gibbs free energy for the reaction of butadiene with Ni(S2C2(CF3)2)2 is also more favourable than those of C4 mono-olefins with Ni(S2C2(CF3)2)2. It is the very preferential pathway for Ni(S2C2(CF3)2)2 to bind butadiene than C4 mono-olefins.  相似文献   

8.
Compounds 1-6 of the type MoO2X2L2 (X=F, Cl, Br; L=OPMePh2, OPPh3) have been prepared in order to investigate the variation in catalytic activity with changes in electronic and steric properties. All six complexes catalyze the epoxidation of cyclohexene with tert-butylhydroperoxide, and the species with X=Cl and L=OPMePh2 (2) displays the best activity with 83% conversion and 90% selectivity in one hour at ambient atmosphere. These inexpensive and easily prepared dioxo catalysts are stable to air and water. Reactions of the dioxo compounds with H2O2 and t-BuOOH have also been carried out. The structures of MoO2F2(OPMePh2)2 (1) and the product of its reaction with H2O2, MoO(O2)2(OPMePh2)2 (7) have been solved by single crystal X-ray diffraction.  相似文献   

9.
New linear and tripodal tetradentate ligands, LH2, are reported and their syntheses are described. The new linear ligands L = HSCH2CH2SCH2CH2NRCH2CR2SH, R = H, CH3) and the new tripodal ligands N(CH2CH2SH)2CH2Z, Z = CH2NH2, CH2N(CH3)2, CH2N(C2H5)2, CH2SCH3 and CO2- were synthesized. The known linear ligands HSCH2CH2NCH3(CH2)nNCH3CH2CH2SH (n = 2, 3) and HSCR2CH2NHCH2CH2NHCH2CR2SH (R = H, CH3) were also utilized. These ligands react with MoO2(acac)2 in CH3OH to yield MoO2L complexes in high yield. Infra-red and 1H nmr spectra provide evidence to supplement X-ray crystallographic results reported elsewhere for selected numbers of the series. Octahedral structures with cis MoO22+ groupings are assigned. Solution 1H nmr studies are consistent with a trans placement of the two thiolate donors in agreement with the X-ray studies.  相似文献   

10.
The effects of C2H2 metabolism on N2O production were examined in soil slurries. Enrichment of C2H2 consumption activity occurred only in aerobic incubations. Rapid disappearance of subsequent C2H2 additions, stimulation of CO2 production, and most-probable-number enumerations of C2H2 utilizers indicated enrichment of the population responsible. During C2H2 consumption in slurries incubated statically under air, maximal rates of N2O evolution were 19 times higher than those in anaerobic incubations. After 20 days of enrichment with C2H2, the production of N2O by slurries supplemented with C2H2 and nitrate was 10 times higher than that in the unenriched controls. A Nocardia- or Arthrobacter-like bacterium was isolated that grew on C2H2 but did not denitrify. The behavior of soil inoculated with this bacterium became similar to that of C2H2-enriched soil incubated aerobically. Ethanol, acetate, and acetaldehyde were identified in enrichment experiments, and denitrification in soil slurries was stimulated by addition of the supernatant from a pure culture grown on mineral medium with C2H2. These results indicate that denitrification can be stimulated by the actions of an aerobic, nondenitrifying C2H2-metabolizing population. Utilization of intermediate metabolites by denitrifiers and enhanced O2 consumption are two possible mechanisms for this stimulation.  相似文献   

11.
Classical benzodiazepines, such as diazepam, interact with αxβ2γ2 GABAA receptors, x = 1, 2, 3, 5 and modulate their function. Modulation of different receptor isoforms probably results in selective behavioural effects as sedation and anxiolysis. Knowledge of differences in the structure of the binding pocket in different receptor isoforms is of interest for the generation of isoform-specific ligands. We studied here the interaction of the covalently reacting diazepam analogue 3-NCS with α1S204Cβ2γ2, α1S205Cβ2γ2 and α1T206Cβ2γ2 and with receptors containing the homologous mutations in α2β2γ2, α3β2γ2, α5β1/2γ2 and α6β2γ2. The interaction was studied using radioactive ligand binding and at the functional level using electrophysiological techniques. Both strategies gave overlapping results. Our data allow conclusions about the relative apposition of α1S204Cβ2γ2, α1S205Cβ2γ2 and α1T206Cβ2γ2 and homologous positions in α2, α3, α5 and α6 with C-atom adjacent to the keto-group in diazepam. Together with similar data on the C-atom carrying Cl in diazepam, they indicate that the architecture of the binding site for benzodiazepines differs in each GABAA receptor isoform α1β2γ2, α2β2γ2, α3β2γ2, α5β1/2γ2 and α6β2γ2.  相似文献   

12.
The new bis(pyrazolyl)amine ligand NH2CH2CH(pz)2 (1) was prepared from the reaction of N-[2,2-bis(pyrazolyl)ethyl]-1,8-naphthalimide with hydrazine monohydrate. A substituted derivative, C6H5CH2NHCH2CH(pz)2 (2), was prepared by the reaction of 1 with benzaldehyde followed by reduction with NaBH4. Ligand 1 was also converted by two methods to the new bitopic, para-linked bis(pyrazolyl)amine ligand p-C6H4(CH2NHCH2CH(pz)2)2, (3). The reactions of the ligands 1-3 with [Cu(PPh3)2]NO3 yields {(PPh3)Cu[(pz)2CHCH2NH2]}NO3, {(PPh3)Cu[(pz)2CHCH2NHCH2C6H5]}NO3 and {[(PPh3)Cu]2[p-((pz)2CHCH2NHCH2)2C6H4]}(NO3)2·solvate, respectively. Complex {(N3)2Cu[(pz)2CHCH2NHCH2C6H5]} was obtained from a methanol solution of 2, copper(II) acetate monohydrate and sodium azide. The complex {Cd[(pz)2CHCH2NHCH2C6H5]2}(PF6)2·3C3H6O was synthesized by reaction of the protonated form of ligand 2, [(pz)2CHCH2NH2CH2C6H5]PF6, with Cd(acac)2. In all of the structures the ligands are tridentate, bonding to the metal through the lone pair on the amine group as well as through the pyrazolyl rings - they act as true scorpionates. The solid state structures all have extensive non-covalent interactions, with the N-H functional groups of the amines participating in both N-H?π and N-H?O or N-H?N hydrogen bonding interactions.  相似文献   

13.
Three mono- and dinuclear nickel complexes with dichalcogenolate o-carboranyl ligands were synthesized and characterized by X-ray crystallography. The reactions of Ni(COD)2(COD=1,5-octadiene) with [(THF)3LiE2C2B10H10Li(THF)]2 (E=S, Se) in THF in the presence of air in different ratios afforded the mono- and dinuclear nickel complexes of formulae Li(THF)4]2[Ni(E2C2B10H10)2] (E=S, 1a; E=Se, 1b) and [Li(THF)4]2[Ni2(E2C2B10H10)3] (E=S, 2a; E=Se, 2b). In 2a, two nickel atoms are connected by one chalcogen (η12-S2C2B10H10) bridging ligand with strong metal-metal interaction. Complex of formula (PPh3)2Ni(S2C2B10H10) · 0.5THF (3a) was also obtained from the reaction of (PPh3)2NiCl2 and [(THF)3LiS2C2B10H10Li(THF)]2.  相似文献   

14.
《Inorganica chimica acta》2004,357(2):571-580
Treatment of the ligand N-(2-mercaptoethyl)-3,5-dimethylpyrazole with [Pd(CH3COO)2]3 and reaction of [PdCl(μ-med)]2 with pyridine (py) or triphenylphosphine (PPh3) in the presence of AgBF4 produced the following complexes: [Pd(CH3COO)(μ-med)]2, [Pd(μ-med)(py)]2(BF4)2 and [Pd(μ-med)(PPh3)]2(BF4)2. Similar reactions carried out with 2,2-bipyridine (bpy) or 1,3-bis(diphenylphosphino)propane (dppp) produced [Pd(μ-med)(bpy)]x(BF4)x (x=1 or 2) and [Pd(μ-med)(dppp)]x(BF4)x (x=1 or 2). Treatment of [Pd(μ-med)(bpy)]x(BF4)x with [PdCl2(CH3CN)2] produced [Pd3Cl2(μ-med)2(bpy)2](BF4)2. Treatment of [Pd(μ-med)(dppp)]x(BF4)x with [PdCl2(CH3CN)2] produced a mixture of [Pd(μ-Cl)(dppp)]2(BF4)2 and [Pd(μ-med)2(dppp)]2+. X-ray crystal structures of [Pd(μ-med)(PPh3)]2(BF4)2 · 2CH3CN and [Pd(μ-med)(bpy)]2(BF4)2 · 0.5CH3OH are presented.  相似文献   

15.
The effect of environmental factors on the post-illumination burst of CO2 (PIB) and O2 inhibition of apparent photosynthesis (APS) in wheat (Triticum aestivum L.) was studied in an open gas exchange system utilizing the mathematics of non-steady-state systems. Two components of inhibition by O2 are suggested: one is caused by photorespiration as measured from the maximum rate of the PIB, and the second is direct inhibition as taken as APS2%O2— (APSx%O2+ PIBx%O2) where X is the oxygen concentration. A primary PIB which occurred from 16–28 s after the darkening of the foliage was attributed to photorespiration. No primary PIB was observed at 2% O2. At a CO2 concentration of 100 μ/1 in the atmosphere (about 2.5 μM based on leaf intercellular concentration) and at 30°C and 145 nE/cm2 nE/cm2·s, APS decreased curve-linearly with increasing O2 and reached an O2 compensation point of 560 μM (48% by volume), above which there was a net loss of CO2 in the light. The PIB increased with increasing O2 and became saturated at about 500 μM O2 but decreased above 900 μM O2. Direct inhibition of photosynthesis by O2 increased with increasing O2 concentration. Decreasing CO2 concentration had an effect on the magnitude of the PIB similar to that of increasing O2. At 30°C and 21% O2, the PIB increased with decreasing CO2 down to the CO2 compensation point (I) of 1.4 μM (47 μM/l). Below Γ, both PIB and CO2 evolution into the air in the light (at 21% O2) increased and then decreased at CO2 below 0.8 μM. The ratio of the PIB to APS2% o O2 increased linearly with increasing O2/CO2 ratio where O2 was held constant at 21% and CO2 was varied from 1.4 to 8.5 μM, while direct inhibition of photosynthesis expressed as a proportion of APS2%O2 remained constant over this range. At low CO2 concentration photorespiration as estimated by the PIB is the major part of O2 photosynthesis, while at atmospheric CO2 levels, direct inhibition is the major component. The PIB and APS at 2% and 21% O2 increased hyperbolically with increasing irradiance and all became light-saturated at about 65 nE/cm2 s. The percentage total O2 inhibition of photosynthesis remained constant with increasing irradiance as did the relative contribution of direct O2 inhibition or photorespiration (PIB) to total O2 inhibition. The PIB and APS at 21% O2 had similar temperature optima of 30°C when experimental conditions were adjusted to provide a constant internal O2/CO2 solubility ratio at varying temperatures. However, with a constant external CO2 concentration, the temperature optimum for the PIB shifted upward to 35°C while that for APS at 21% O2 remained at 30°C, which may be due to an increased O2/CO2 concentration in the leaf with increasing temperature.  相似文献   

16.
[Rh2(μ-Cl)2(cod)2] reacts with Ph2PCH2CH2OMe (PC2O), Ph2P(CH2)3NMe2 (PC3N), PBunPh2 or PPh3 to give [Rh(cod)L2]Cl (L = PC2O, PC3N, PBunPh2, PPh3). In the presence of hydrogen, [Rh(cod)L2]Cl is converted to [RhClH2L3]. In contrast, [Rh(cod)(PC2O)2]BPh4 reacts with H2 to give [RhH2(PC2O)2S2]BPh4 (S = solvent). With Ph2PCH2CH2NMe2 (PC2N) or Ph2PCH2CH2SMe (PC2S), [Rh2(μ-Cl)2(cod)2] reacts to form the chelate complexes cis- [Rh(PC2N)2]+ or cis-[Rh(PC2S)2]+, neither of which reacts with hydrogen under ambient conditions. The products of the reactions are characterized in situ by 31P1H NMR spectroscopy.  相似文献   

17.
Fusicoccin (FC) treatment prevents dark‐induced stomatal closure, the mechanism of which is still obscure. By using pharmacological approaches and laser‐scanning confocal microscopy, the relationship between FC inhibition of dark‐induced stomatal closure and the hydrogen peroxide (H2O2) levels in guard cells in broad bean was studied. Like ascorbic acid (ASA), a scavenger of H2O2 and diphenylene iodonium (DPI), an inhibitor of H2O2‐generating enzyme NADPH oxidase, FC was found to inhibit stomatal closure and reduce H2O2 levels in guard cells in darkness, indicating that FC‐caused inhibition of dark‐induced stomatal closure is related to the reduction of H2O2 levels in guard cells. Furthermore, like ASA, FC not only suppressed H2O2‐induced stomatal closure and H2O2 levels in guard cells treated with H2O2 in light, but also reopened the stomata which had been closed by darkness and reduced the level of H2O2 that had been generated by darkness, showing that FC causes H2O2 removal in guard cells. The butyric acid treatment simulated the effects of FC on the stomata treated with H2O2 and had been closed by dark, and on H2O2 levels in guard cells of stomata treated with H2O2 and had been closed by dark, and both FC and butyric acid reduced cytosol pH in guard cells of stomata treated with H2O2 and had been closed by dark, which demonstrates that cytosolic acidification mediates FC‐induced H2O2 removal. Taken together, our results provide evidence that FC causes cytosolic acidification, consequently induces H2O2 removal, and finally prevents dark‐induced stomatal closure.  相似文献   

18.
The reactivity of [Pt2(μ-S)2(PPh3)4] towards a range of nickel(II) complexes has been probed using electrospray ionisation mass spectrometry coupled with synthesis and characterisation in selected systems. Reaction of [Pt2(μ-S)2(PPh3)4] with [Ni(NCS)2(PPh3)2] gives [Pt2(μ-S)2(PPh3)4Ni(NCS)(PPh3)]+, isolated as its BPh4 − salt; the same product is obtained in the reaction of [Pt2(μ-S)2(PPh3)4] with [NiBr2(PPh3)2] and KNCS. An X-ray structure determination reveals the expected sulfide-bridged structure, with an N-bonded thiocyanate ligand and a square-planar coordination geometry about nickel. A range of nickel(II) complexes NiL2, containing β-diketonate, 8-hydroxyquinolinate, or salicylaldehyde oximate ligands react similarly, giving [Pt2(μ-S)2(PPh3)4NiL]+ cations.  相似文献   

19.
《Inorganica chimica acta》1988,154(2):177-182
TiCl4 reacts with t-butylamine in benzene to give [Ti(NCMe3)Cl2(NH2CMe3)2]x and t-butylamine hydrochloride. The IR spectrum indicates both c/s and trans metal dichlorides (300; and 308, 208 cm−1). In the 13C NMR spectrum the t-butylimido quaternary carbon resonance occurs at 72.1 ppm. A dimeric structure incorporating symmetric t-butylimido bridges is proposed. TiCl4 in benzene react under reflux with two equivalents of Me3SiNHCMe3 to give [Ti(NCMe3)Cl2(NH2CMe3)]x and with iso-propylamine and ethylamine to give complexes of the form [Ti(NR)Cl2(NH2R)2]x. Broad bands below 800 cm−1 in the IR spectra suggest polymeric MNM bridges. For [Ti(NCHMe2)Cl2(NH2CHMe2)]x the iso-propylimido CH resonance in the 13C NMR spectrum occurs at 67 ppm. [Ti(NCMe3)Cl2(NH2CMe3)2]2 reacts with L=bipy or tmed to give [Ti(NCMe3)Cl2(L)]2, and TiCl4 reacts with two equivalents of Me3SiNHCMe3 in benzene and then tmed to give [Ti(NCMe3)Cl2(tmed)]2. The 13C NMR spectrum shows the t-butylimido quaternary carbon resonance at 73.5 ppm and the tmed resonances are chemically equivalent. A dimeric μ-NCMe3 bridging structure is proposed for the complex.  相似文献   

20.
The effects of several metals on microbial methane, carbon dioxide, and sulfide production and microbial ATP were examined in sediments from Spartina alterniflora communities. Anaerobically homogenized sediments were amended with 1,000 ppm (ratio of weight of metal to dry weight of sediment) of various metals. Time courses in controls were similar for CH4, H2S, and CO2, with short initial lags (0 to 4 h) followed by periods of constant gas production (1 to 2 days) and declining rates thereafter. Comparisons were made between control and experimental assays with respect to initial rates of production (after lag) and overall production. Methane evolution was inhibited both initially and overall by CH3HgCl, HgS, and NaAsO2. A period of initial inhibition was followed by a period of overall stimulation with Hg, Pb, Ni, Cd, and Cu, all as chlorides, and with ZnSO4, K2CrO4, and K2Cr2O7. Production of CO2 was generally less affected by the addition of metals. Inhibition was noted with NaAsO2, CH3HgCl, and Na2MoO4. Minor stimulation of CO2 production occurred over the long term with chlorides of Hg, Pb, and Fe. Sulfate reduction was inhibited in the short term by all metals tested and over the long term by all but FeCl2 and NiCl2. Microbial biomass was decreased by FeCl2, K2Cr2O7, ZnSO4, CdCl2, and CuCl2 but remained generally unaffected by PbCl2, HgCl2, and NiCl2. Although the majority of metals produced an immediate inhibition of methanogenesis, for several metals this was only a transient phenomenon followed by an overall stimulation. The initial suppression of methanogenesis may be relieved by precipitation, complexation, or transformation of the metal (possibly by methylation), with the subsequent stimulation resulting from a sustained inhibition of competing organisms (e.g., sulfate-reducing bacteria). For several environmentally significant metals, severe metal pollution may substantially alter the flow of carbon in sediments.  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