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 46 毫秒
1.
Winter rye (Secale cereale L. cv Musketeer) grown at 5 °C/250 µmol photons m–2 s–1 exhibited a relative reduction state of PS II comparable to that of rye grown at 20 °C but high light (800 µmol photons m–2 s–1) (1-qP = 0.32) whereas winter rye grown at 20 °C/250 µmol photons m–2 s–1 exhibited values of 1-qP ( 0.15) comparable to plants grown at 5 °C but low light (50 µmol photons m–2 s–1). The apparent size of the electron donor pool to PS I, estimated either in vivo or in vitro in the presence of methylviologen by A820 was positively correlated with the relative reduction state of PS II under the steady-state growth conditions. Immunoblotting of rye thylakoid polypeptides indicated that the relative contents of Lhcb1, Lhcb2, D1, Cyt f, PC, PsaA/PsaB heterodimer and the -subunit of ATPase complex exhibited minimal changes on a Chl basis. In contrast, a 2-fold increase in plastoquinone A content was associated with increasing growth irradiance at growth temperatures of either 5 or 20 °C. We suggest that the increases in the apparent size of the electron donor pool to PS I associated with rye grown at either 5 °C/250 µmol photons m–2 s–1or 20 °C/800 µmol photons m–2 s–1 may be explained by an increased thylakoid plastoquinone A content, coupled with possible enhanced PS I cyclic electron transport and/or increased capacity for electron donation from the stroma to the intersystem electron transport chain. The results are discussed with respect to photosynthetic adjustment to changes in PS II excitation pressure in winter rye.  相似文献   

2.
Leaflets of soybean plants which are moderately inorganic nitrogen (N)-limited exhibit either no difference in the rate of net photosynthesis or as much as a 15–23% lower net photosynthesis rate per unit area than leaflets of N-sufficient plants [Robinson JM (1996) Photosynth Res 50: 133–148; Robinson JM (1997a) Int J Plant Sci 158: 32–43]. However, mature leaflets of N-limited soybean plants have a higher CO2photoassimilation rate per unit chlorophyll than leaflets of N-sufficient soybean plants at both moderate light intensity (500 µmol m-2s-1) and saturating light intensity (1200 µmol m-2s-1) [Robinson JM (1996) Photosynth Res 50: 133–148]. This study was undertaken to determine whether chloroplast thylakoids isolated from the leaflets of nitrogen-limited soybean plants displayed similar or higher linear electron transport rates (H2O ferredoxin NADP) per unit chlorophyll than thylakoids isolated from leaflets of N-sufficient plants. Chlorophyll concentration in reaction mixtures containing chloroplast thylakoids prepared from leaflets of N-limited plants was manipulated so that it was similar to the chlorophyll concentration in reaction mixtures of thylakoids prepared from leaflets of N-sufficient plants. Measurements of ferredoxin dependent, NADP dependent, O2photo-evolution in thylakoid isolates were carried out in saturating light (1500 µmol m-2s-1) and with (an uncoupler) in the chloroplast reaction mixtures. Chloroplast thylakoids isolated from N-limited soybean plant leaflets routinely had a 1.5 to 1.7 times higher rate of uncoupled, whole chain electron transport per unit chlorophyll in saturating light than did chloroplast thylakoids isolated from leaflets of N-sufficient plants. The results suggest that the photosystems and photosynthetic electron transport chain components are more active per unit Chl in leaflet chloroplast thylakoids of N-limited soybean plants than in thylakoids of N-sufficient plants.  相似文献   

3.
We have examined the effect of the Ca2+ (Mg2+)-ATPase inhibitors thapsigargin (TG) and vanadate on ATP-dependent 45Ca2+ uptake into IP3-sensitive Ca2+ pools in isolated microsomes from rat pancreatic acinar cells. The inhibitory effect of TG was biphasic. About 40–50% of total Ca2+ uptake was inhibited by TG up to 10 nm (apparent Ki4.2 nm, Ca2+ pool I). An additional increase of inhibition up to 85–90% of total Ca2+ uptake could be achieved at 15 to 20 nm of TG (apparent Ki12.1 nm, Ca2+ pool II). The rest was due to TG-insensitive contaminating plasma membranes and could be inhibited by vanadate (apparent Ki10 m). In the absence of TG, increasing concentrations of vanadate also showed two phases of inhibition of microsomal Ca2+ uptake. About 30–40% of total Ca2+ uptake was inhibited by 100 m of vanadate (apparent Ki18 m, Ca2+ pool II). The remaining 60–70% could be inhibited either by vanadate at concentrations up to 1 mm (apparent Ki300 m) or by TG up to 10 nm (Ca2+ pool I). The amount of IP3-induced Ca2+ release was constant at 25% over a wide range of Ca2+ filling. About 10–20% remained unreleasable by IP3. Reduction of IP3 releasable Ca2+ in the presence of inhibitors showed similar dose-response curves as Ca2+ uptake (apparent Ki 3.0 nm for IP3-induced Ca2+ release as compared to 4.2 nm for Ca2+ uptake at TG up to 10 nm) indicating that the highly TG-sensitive Ca2+ pump fills the IP3-sensitive Ca2+ pool I. At TG concentrations >10 nm which blocked Ca2+ pool II the apparent Ki values were 11.3 and 12.1 nm, respectively. For inhibition by vanadate up to 100 m the apparent Ki values were 18 m for Ca2+ uptake and 7 m for Ca2+ release (Ca2+ pool II). At vanadate concentrations up to 1 mm the apparent Ki values were 300 and 200 m, respectively (Ca2+ pool I). Both Ca2+ pools I and II also showed different sensitivities to IP3. Dose-response curves for IP3 in the absence of inhibitors (control) showed an apparent Km value for IP3 at 0.6 m. In the presence of TG (inhibition of Ca2+ pool I) the curve was shifted to the left with an apparent Km for IP3 at 0.08 m. In the presence of vanadate (inhibition of Ca2+ pool II), the apparent Km for IP3 was 2.1 m. These data allow the conclusion that there are at least three different Ca2+ uptake mechanisms present in pancreatic acinar cells: TG- and IP3 insensitive but highly vanadate-sensitive Ca2+ uptake occurs into membrane vesicles derived from plasma membranes. Two Ca2+ pools with different TG-, vanadate- and IP3-sensitivities are most likely located in the endoplasmic reticulum at different cell sites, which could have functional implications for hormonal stimulation of pancreatic acinar cells.This work was supported by the Deutsche Forschungsgemeinschaft, Sonderforschungsbereich 246. The authors wish to thank Dr. KlausDieter Preuß for valuable discussions and Mrs. Gabriele Mörschbächer for excellent secretarial help.  相似文献   

4.
Summary Two IgE-producing hybridomas were established from spleen cells of Balb/c mice, which had been immunized with mouse mammary tumor virus (MMTV). These IgE monoclonal antibodies (mAbs) reacted specifically with the major envelope glycoprotein (gp36) of MMTV, as established by the immunoblot assay and by passive cutaneous anaphylaxis. The effect of the IgE mAbs (produced by clone A8) on the growth of the MMTV-secreting mammary adenocarcinoma H2712 was investigated in syngeneic C3H/HeJ mice. The mice were inoculated s.c. with either 105 (100 × LD50) or 106 (1000 × LD50) tumor cells and received repeated i.p. injections of 25 µg anti-gp36 IgE mAbs at 4-day intervals for 8 weeks. This treatment prevented the development of subcutaneous tumors in 50% of the animals. Similar protection was observed when the tumor cells (105/animal) were injected i.p. 4 days prior to the beginning of the i.p. treatment consisting of injections of 25 µg mAbs at 4-day intervals for 6 weeks. However, these mAbs did not protect C3H/HeJ mice against the MMTV-negative MA16/c carcinoma cells. Hence, these results support the view that IgE-mediated cytotoxic mechanisms may play an immunologically specific antitumor surveillance role and that laboratory-induced antitumor IgE mAbs have the potential of specific therapeutic agents for in vivo destruction of tumor cells.  相似文献   

5.
In an attempt to assess the influence of tertiary-treated sewage influx on bacterial metabolism in the Danube River, bacterial abundance, ectoenzymatic activity, faecal coliforms, faecal streptococci, chlorophyll a (chl a), concentrations of dissolved (DOC) and total organic carbon (TOC) and inorganic nutrients were measured upstream of the sewage influx and compared with sampling sites downstream. Additional samples were taken near the outlet of the sewage treatment plant. Bacterial abundance as determined by epifluorescence microscopy was compared with plate counts of total heterotrophic bacteria. Significantly higher values were obtained at the stations downstream from the sewage influx only for faecal coliforms and faecal streptococci, for glucuronidase activity and bacterial biovolume. All the other parameters were not significantly different from values obtained at the upstream sampling site. Strong seasonal dependence was detectable for nitrate with high concentrations during the winter ( 250 µM) and low concentrations during summer ( 100 µM). A distinct spring phytoplankton bloom was noticeable in the river reaching chl a concentrations of 70 µg 1–1; during the remaining seasons chl a concentrations were <20 µg 1–1. Highly significant correlations were found between faecal coliform counts and glucuronidase activity. C:N:P ratios of 13.9:10.7:1 (for the upstream station) and 11.7:9.2:1 (for the downstream stations) of dissolved nutrients are rather low indicating severe carbon limitation of bacterioplankton. Based on our results we conclude that the heterotrophic bacterial community is not significantly effected by the input of treated sewage but faecal contamination is readily detectable over a comparatively long reach of 30 km.  相似文献   

6.
One of the most important stages of hormonal signal transduction in cells through the hormone-sensitive adenylyl cyclase signal system (ACS) is functional coupling of receptor of the serpentine type to heterotrimeric GTP-binding protein (G-protein). The main role in realization of such coupling is played by spiralized regions of the receptor cytoplasmic loops proximal in relation to membrane, most of them carrying positive charge. To study molecular mechanisms of interaction of the receptor with G-protein, we compared effects of synthetic cationic peptides containing residues of glutamic acid on the process of regulation of ACS by hormones (biogenic amines) and non-hormonal agents in smooth muscles of the freshwater bivalve mollusc Anodonta cygnea and skeletal muscles of rat. All peptides had the clearly expressed ability to form -helices. Peptides H-(Leu-His-Glu-Lys)4-Leu-NH2 (I), H-(Leu-His-Glu-Lys)3-Lys-His-Glu-Lys-Leu-NH2 (II), H-(Leu-Lys-Glu-Lys)4-Leu-NH2 (III), and H-(Ile-His-Glu-Lys)4-Ala-NH2 (IV) at concentrations of 10–6–10–3 M reduced dose-dependently the value of stimulating effects of serotonin (in mollusc muscles) and isoproterenol (in rat muscles) on the adenylyl cyclase (AC) and protein kinase A (PKA) activities. Values of concentration of these peptide causing a 50% decrease of the hormone-stimulating effect (IC50) vary from 150 to 750 µM. According to the degree of this inhibitory action on stimulating effects of hormones, they may be arranged in the following series: III II > IV I. The peptides I–IV were more effective than the peptide H-(Glu-Lys)8-Ala-NH2 (V) with the charge close to zero, but much less effective than the studied earlier cationic peptides containing only positively charged amino acid residues. The inhibitory effect of the peptides I-IV on stimulation of AC by non-hormonal agents, NaF, Gpp[NH]p, and forskolin, was essentially less pronounced and was marked only at 10–4–10–3 M concentrations. Thus, the inclusion of negatively charged amino acid residues in the primary structure of polycationic peptides leads to a decrease in their ability to inhibit hormonal stimulation of AC and PKA, which indicates importance both of the total positive charge of peptides and of distribution of the charged amino acids in the formed helices for realization of the uncoupling action on the ACS components—the receptor and G-protein.  相似文献   

7.
The tris-chelate formed by biguanide, H2NC(NH)NHC(NH)NH2, and Mn(IV) is one of the rarely encountered examples of water-stable mononuclear complexes of this oxidation state. This cation is reduced in aqueous acid by both V(II) and V(III). In contrast to the kinetically straightforward bimolecular reduction by V(II) in 0.5 M HClO4 (k = 7.4 × 103 M−1 s−1 at 22 °C), reductions by excess V(III) yield profiles which are linear (zero-order in MnIV) until the last few percent reaction. Analyses of these composite curves are consistent with the sequences
  相似文献   

8.
The new acrylamide iron(II)/iron(III) complex [Fe(O-OC(NH2)CHCH2)6][Fe2OCl6] (1) was obtained by the reaction of a mixture of anhydrous FeCl2 and anhydrous FeCl3 with acrylamide (molar ratio 1:2:6) in 98% pure commercial nitromethane under nitrogen atmosphere. According to an X-ray structural analysis, the acrylamide ligands in the cation are coordinated via the amide-oxygen atoms. The formation of the (μ-oxo)bis[trichloroferrate(III)]2− anion presumably resulted from partial hydrolysis of FeCl3 or [FeCl4] by small amounts of water in the nitromethane and/or by the nitromethane itself.  相似文献   

9.
The pharmacological activation of the cystic fibrosis transmembrane conductance regulator (CFTR) chloride channel mutated at glycine 551 (G551D-CFTR) was studied in the presence of the benzimidazolone derivative NS004 and compared to that of wild-type (wt) CFTR. Using iodide (125I) efflux and whole-cell patch-clamp techniques we found dose-dependent stimulation of phosphorylated wt-CFTR channels by NS004 with an EC 50 11 µM. With non-phosphorylated CFTR, the effect of NS004 was apparent only at concentration >100 µM. In G551D-CFTR-expressing CHO cells, neither forskolin (from 0.1 to 10 µM) nor NS004 (from 0.1 to 200 µM) added separately were able to stimulate channel activity. However, in the presence of 10 µM forskolin, NS004 stimulated G551D-CFTR activity in a dose-dependent manner with an EC 50 1.5 µM. We also determined the half-maximal effective concentration of forskolin (EC 50 3.2 µM) required to stimulate G551D channel activity in presence of 1.5 µM NS004. No inhibitory effect was observed at high concentration of NS004 with both wt- and G551D-CFTR. Whole-cell recordings of CFTR chloride currents from cells expressing wild-type or G551D-CFTR in the presence of NS004 were linear, time- and voltage-independent. The inhibitory profile of G551D-CFTR channel activity was similar to that of wild type, i.e., inhibition by glibenclamide (100 µM) and DPC (250 µM) but not by DIDS (200 µM) nor calixarene (100 nM). These results show that NS004 activates wt-CFTR channel and restores G551D-CFTR channel activity, the potency of which depends on both the concentration of NS004 and the phosphorylation status of CFTR.  相似文献   

10.
Summary The objective was to measure the competition for nitrogen among vascular plants, mosses, and soil microbes along a continuum of nitrogen availability, induced by carbon and nitrogen amendments, in a tussock tundra ecosystem.15N was used as a tracer. Vascular plants showed an increasing15N recovery with increasing time and with increasing nitrogen availability; the latter suggests that nitrogen was limiting vascular plant growth. Green mosses took up15N initially, but showed no significant trends with either treatment or time. There was a higher15N recovery in the soil insoluble compartment for the carbon-amended treatment than in the nitrogen-amended treatments; this suggested that carbon as an energy source limited microbial activity. After two months, the relative15N recovery fell in the order: soil microbes (79%)>vascular plants (16%) >green mosses (2%).  相似文献   

11.
This study investigates the occurrence and regulation of serine/threonine protein phosphatases (PPases) in insulin-secreting RINm5F insulinoma cells. PPases types 1 and 2A were identified in crude RINm5F cell homogenates by both enzymatic assay and Western blot analysis. We then characterized and compared the inhibitory actions of several compounds isolated from cyanobacteria, marine dinoflagellates and marine sponges, (viz. okadaic acid, microcystin-LR, calyculin-A and nodularin) cation-independent PPase activities in RINm5F cell homogenates. It was found that okadaic acid was the least potent inhibitor (IC5010–9M, IC10010–6M), while the other compounds exhibited IC50 values of 5·10–10 M and IC100 5·10–9 M. The findings indicate that the inhibitory substances employed in this study may be used pharmacologically to investigate the role of serine/threonine PPases in RINm5F cell insulin secretion, a process that is likely to be regulated to a major extent by protein phosphorylation.  相似文献   

12.
In a computational model of spreading depression (SD), ionic movement through a neuronal syncytium of cells connected by gap junctions is described electrodiffusively. Simulations predict that SD will not occur unless cells are allowed to expand in response to osmotic pressure gradients and K+ is allowed to move through gap junctions. SD waves of [K+]out 25 to 60 mM moving at 2 to 18 mm/min are predicted over the range of parametric values reported in gray matter, with extracellular space decreasing up to 50%. Predicted waveform shape is qualitatively similar to laboratory reports. The delayed-rectifier, NMDA, BK, and Na+ currents are predicted to facilitate SD, while SK and A-type K+ currents and glial activity impede SD. These predictions are consonant with recent findings that gap junction poisons block SD and support the theories that cytosolic diffusion via gap junctions and osmotic forces are important mechanisms underlying SD.  相似文献   

13.
Benthic sulfate reduction and sediment pools of sulfur and iron were examined during January 1992 at 3 stations in the Ao Nam Bor mangrove, Phuket, Thailand. Patterns of sulfate reduction rates (0–53 cm) reflected differences in physical and biological conditions at the 3 stations, and highest rates were found at the vegetated site within the mangrove (Rhizophora apiculata) forest. Due to extended oxidation of mangrove sediments, a large portion of the added35S-label was recovered in the chromium reducible pools (FeS2 and S0) (41–91% of the reduced sulfur). Pyrite was the most important inorganic sulfur component, attaining pool sizes 50–100 times higher than acid volatile pools (FeS). HCl-extractable (0.5 M HCl) iron pools, including Fe(II)HCl and Fe(III)HCl, were generally low and Fe(III)HCl was only present in the upper surface layers (0–5 cm). Maximum concentrations of dissolved Fe2+ (35–285 M) occurred just about the depth where dissolved H2S accumulated. Furthermore Fe2+ and H2S coexisted only where concentrations of both were low. There was an accumulation of organic sulfur in the deep sediment at 2 stations in the inner part of the mangrove. The reoxidation of reduced sulfides was rapid, and storage of sulfur was minor in the upper sediment layers, where factors like bioturbation, the presence of roots, or tidal mixing enhance oxidation processes.Author of correspondence.  相似文献   

14.
Aluminum geochemistry in peatland waters   总被引:4,自引:4,他引:0  
The chemical speciation of aluminum was examined in surface water samples from Sphagnum peatlands in north-central Minnesota, from peatlands along the Canadian east coast, and from bogs in the Pennine Mountain area of England. In highly organic ([DOC] 50 mg L–1 ), low pH waters, 80–90% of total dissolved Al was complexed with organic matter (OM), while in waters with low DOC ([DOC] 5 mg L–1) 54–86% of total dissolved Al existed as Al+3 or other inorganic Al species. Batch titrations of OM with Al revealed a high Al binding capacity, 1.4–2.8 mol (mg DOC)–1, that generally was unsaturated with Al. Titrations of OM with Al in conjunction with a continuous distribution model were used to determine Al-OM conditional stability constants. Binding capacity (mol Al (mg DOC)–1) and strength (formation constant) increased from pH 3 to 5 but decreased above pH 5 due to formation of AI-hydroxy species including A1(OH)3 (s). The high binding capacity of OM in bog waters facilitates metal mobility, especially in low pH (< 5) wetlands where metal solubility is high and OM concentrations are highest. Results showed that the relative degree of organic matter saturation with metal ions was important in modeling AI speciation in bog waters.  相似文献   

15.
Summary Copper(II) complexes CuL1L2 with the ligand pairs 3-phosphoglycerate (PG)/ethylenediamine (en), phosphoserine (PS)/ethylenediamine, phosphoserine/malonate (mal) are shown to be effective in inducing the release of both iron atoms from di-ferric transferrin (Fe2Tf; human serum transferrin) at pH 7.3 in 1 M NaCl at 25°C. Half-times of the reaction with Cu(PG)(en) were less than 1 min at 0.02 M concentration. The iron(III) products are polynuclear hydroxo complexes. There is weaker interaction with Cu(PS) 2 4– and virtually none with Cu(serine)(en) nor Cu(PS)(2,2-bipyridyl), revealing crucial effects of the combined ligand sphere including the phosphomonoester group. The results suggest that the release of iron from Fe2Tf, or from either monoferric transferrins, occurred due to the breakdown of the stability of iron binding in conjunction with the expulsion of the synergistic anion carbonate (or oxalate). The active copper(II) complexes are postulated to be models of membrane components that could liberate iron from transferrin succeeding its uptake at the receptor sites of cells.Abbreviations PG phosphoglycerate - PS phosphoserine - en ethylenediamine - Fe2Tf diferric transferrin - FecTf and FeNTf transferrin with iron bound to the lobe containing the C- or N-terminus, respectively - apoTf apotransferrin - K-3 all-cis-1,3,5-tris(trimethylammonio)-2,4,6-cyclo-hexanetriol - NTA nitrilotriacetic acid; bipy, 2,2-bipyridine; mal, malonate  相似文献   

16.
Sequential fractionation of sediment phosphate   总被引:24,自引:15,他引:9  
By means of sequential extractions with Ca-NTA and EDTA, a separation was performed between Fe(OOH) P and CaC03P in a few sediments; the remaining fraction, considered to be organic phosphate, was quantified as well. We found that with the commonly used method of extraction with NaOH and H2S04, less Fe(OOH) P and much more CaC03 P was found than with the chelating extractants. The organic phosphate pool in live and dead algal material and in some mud samples was partly hydrolysed and therefore recovered as inorganic phosphates with classical extractions. The difference between chelating extractants and the classical ones is discussed.Abbreviations o-P: ortho phosphate (or its concentration) - org-P: organic phosphate - extr-P: extractable sediment bound phosphate - extr-Fe: extractable sediment bound iron - Fe(OOH) P: iron bound, sediment phosphate - CaCO3 P: calcium bound, sediment phosphate - org-C: organic sediment bound carbon  相似文献   

17.
Neutrophilic Fe(II) oxidizing microorganisms are found in many natural environments. It has been hypothesized that, at low oxygen concentrations, microbial iron oxidation is favored over abiotic oxidation. Here, we compare the kinetics of abiotic Fe(II) oxidation to oxidation in the presence of the bacterium Leptothrix cholodnii Appels isolated from a wetland sediment. Rates of Fe(II) oxidation were determined in batch experiments at 20°C, pH 7 and oxygen concentrations between 3 and 120 μmol/l. The reaction progress in experiments with and without cells exhibited two distinct phases. During the initial phase, the oxygen dependency of microbial Fe(II) oxidation followed a Michaelis-Menten rate expression (KM = 24.5 ± 10 μmol O2/l, vmax = 1.8 ± 0.2 μmol Fe(II)/(l min) for 108 cells/ml). In contrast, abiotic rates increased linearly with increasing oxygen concentrations. At similar oxygen concentrations, initial Fe(II) oxidation rates were faster in the experiments with bacteria. During the second phase, the accumulated iron oxides catalyzed further oxidative iron precipitation in both abiotic and microbial reaction systems. That is, abiotic oxidation also dominated the reaction progress in the presence of bacteria. In fact, in some experiments with bacteria, iron oxidation during the second phase proceeded slower than in the absence of bacteria, possibly due to an inhibitory effect of extracellular polymeric substances on the growth of Fe(III) oxides. Thus, our results suggest that the competitive advantage of microbial iron oxidation in low oxygen environments may be limited by the autocatalytic nature of abiotic Fe(III) oxide precipitation, unless the accumulation of Fe(III) oxides is prevented, for example, through a close coupling of Fe(II) oxidation and Fe(III) reduction.  相似文献   

18.
A series of mononuclear iron(III) complexes with containing phenolate donor of substituted-salicylaldimine based ligands [Fe(L1)(TCC)] · CH3OH (1), [Fe(L2)(TCC)] · CH3OH (2), [Fe(L3)(TCC)] (3), and [Fe(L4)(TCC)] (4) have been prepared and studied as functional models for catechol dioxygenases (H2TCC = tetrachlorocatechol, or HL1 = N′-(salicylaldimine)-N,N-diethyldiethylenetriamine, HL2 = N′-(5-Br-salicylaldimine)-N,N-diethyldiethylenetriamine, HL3 = N′-(4,6-dimethoxy-salycyl-aldimine)-N,N-diethyl-diethylenetriamine, HL4 = N′-(4-methoxy-salicylaldimine)-N,N-diethyl-diethylenetriamine). They are structural models for inhibitors of enzyme-substrate adducts from the reactions of catechol 1,2-dioxygenases. Complexes 1-4 were characterized by spectroscopic methods and X-ray crystal structural analysis. The coordination sphere of Fe(III) atom of 1-4 is distorted octahedral with N3O3 donor set from the ligand and the substrate TCC occupying cis position, and Fe(III) is in high-spin (S = 5/2) electronic ground state. The in situ prepared iron(III) complexes without TCC, [Fe(L1)Cl2], [Fe(L2)Cl2], [Fe(L3)Cl2], and [Fe(L4)Cl2] are reactive towards intradiol cleavage of the 3,5-di-tert-butylcatechol (H2DBC) in the presence of O2 or air. The reaction rate of catechol 1,2-dioxygenase depends on the redox potential and acidity of iron(III) ions in complexes as well as the substituent effect of the ligands. We have identified the reaction products and proposed the mechanism of the reactions of these iron(III) complexes with H2DBC with O2.  相似文献   

19.
The iron(II) compound of formula [Fe(NCS)2(dena)2]n (dena = N,N′-diethylnicotinamide) has been prepared by the reaction between iron(III) thiocyanate and dena in ethanol solution. The complex was characterized by elemental analysis, spectral and magnetic measurements. Single-crystal X-ray diffraction methods show that the complex, crystallizing in the triclinic space group, undergoes a phase transition between 220 K and 230 K, connected with the doubling of cell volume. Crystal structures at 230 K (1a; HT phase) and 150 K (1b; LT phase) are described and a transition mechanism is discussed. In both phases the compound has an extended chain structure, in which the neutral molecule of N,N′-diethylnicotinamide acts as a bridging ligand binding through pyridine N atom to one centre and through amide O atom to the neighbouring Fe centre. The Fe2+ ion has a slightly distorted trans-octahedral environment with FeO2N4 chromophore, and all Fe-O and Fe-N bonds in the typical for high-spin iron(II) compounds range. Variable-temperature magnetic susceptibility data in the temperature range 1.8-300 K show that iron(II) is high-spin S = 2(5T2g) and as a result effects due to zero-field splitting are anticipated at low temperatures. The IR spectrum suggested the coordination of N,N′-diethylnicotinamide to the central atom of iron(II) as a bridging ligand and NCS group as a monodentate ligand.  相似文献   

20.
A great number of important chemical reactions that occur in the environment are microbially mediated. In order to understand the kinetics of these reactions it is necessary to develop methods to directly measure in situ reaction rates and to develop models to help elucidate the mechanisms of microbial catalysis. The oxidation of Mn(II) in a zone above the O2/H2S interface in Saanich Inlet, B.C., Canada is one such reaction. We present here a method by which in situ rates of microbial Mn(II) oxidation are measured and a model based on our experimental results to describe the general mechanism of Mn(H) oxidation. We propose a two step process in which Mn(II) is first bound by a site on the bacterial surface and then oxidized. The model is analogous to the Langmuir isotherm model for surface catalyzed gas reactions or the Michaelis-Menten model for enzyme kinetics. In situ Mn(II) oxidation rates were measured during five cruises to Saanich Inlet during the summers of 1983 and 1984. We use the model to calculate the apparent equilibrium binding constant (Ks 0.18 M), the apparent half saturation constant for biological Mn(H) oxidation (Km = 0.22 to 0.89 M), the maximum rate of Mn(II) oxidation (Vmax = 3.5 to 12.1 nM·h-1) and the total microbial surface binding site concentration ( E 51 nM). Vmax for Mn(II) oxidation agrees with the rates calculated from the value of the flux of Mn(II) to the oxidizing zone using the Mn(II) gradient and estimates of the eddy diffusion coefficient. This consistancy verifies our methodology and indicates that the rate of Mn(II) oxidation is nearly equal to the (Vmax for the reaction. We conclude that in this environment the Mn(II) oxidation rate is more a function of the total number of surface binding sites than the Mn(H) concentration.Contribution #1601 from the School of Oceanography, Univ. of Washingtoncorresponding author  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