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 15 毫秒
1.
The possibility of low but nontrivial atmospheric oxygen (O2) levels during the mid‐Proterozoic (between 1.8 and 0.8 billion years ago, Ga) has important ramifications for understanding Earth's O2 cycle, the evolution of complex life and evolving climate stability. However, the regulatory mechanisms and redox fluxes required to stabilize these O2 levels in the face of continued biological oxygen production remain uncertain. Here, we develop a biogeochemical model of the C‐N‐P‐O2‐S cycles and use it to constrain global redox balance in the mid‐Proterozoic ocean–atmosphere system. By employing a Monte Carlo approach bounded by observations from the geologic record, we infer that the rate of net biospheric O2 production was Tmol year?1 (1σ), or ~25% of today's value, owing largely to phosphorus scarcity in the ocean interior. Pyrite burial in marine sediments would have represented a comparable or more significant O2 source than organic carbon burial, implying a potentially important role for Earth's sulphur cycle in balancing the oxygen cycle and regulating atmospheric O2 levels. Our statistical approach provides a uniquely comprehensive view of Earth system biogeochemistry and global O2 cycling during mid‐Proterozoic time and implicates severe P biolimitation as the backdrop for Precambrian geochemical and biological evolution.  相似文献   

2.
Abstract: The earliest O2--evolvers were marine cyanobacteria (3.5 billion years ago) with marine eukaryotic phototrophs from 2.0 billion years ago. These organisms were, and are, poikilo-hydric, i.e., cannot remain hydrated when exposed to a desiccating atmosphere (as can occur for intertidal benthic algae and cy anobacteria at low tide). The smallest marine primarily poikilo-hydric O2--evolvers are close to the lower size limit imposed by non-scaleable components such as minimum genome size and constant membrane thickness, with cyanobacterial unicells 0.65 μn in diameter and eukaryotic unicells 0.95 μm in diameter. The largest (multicellular) marine primarily aquatic poikilohydric O2--evolvers are brown algae at least 60 m long and over 100 kg fresh mass; there are no obvious constraints on the max imum size of such organisms. In freshwaters the size range for primarily poikilohydric O2--evolving organisms is smaller, due to the absence of very large organisms. An even smaller size range characterizes terrestrial algae and cyanobacteria which have occurred for about 1 billion years. Desiccation-tolerant cyanobacterium and algae (intertidal, freshwater, terrestrial) are at the lower end of the size ranges. Embryophytic terrestrial O2--evolvers arose some 450 million years ago and were than all poikilohydric and (probably) desiccation-tolerant. Embryophytic defining structural features re quire organisms of at least 100 μm equivalent spherical diameter for both gametophyte and sporophyte phases. Primarily poi kilohydric embryophytes are not more than 1 m tall as a result of a mechanistically mysterious size limit for desiccation-tolerant organisms. Homoiohydric embryophytes evolved some 420 mil lion years ago in the sporophyte phase (later to become the dominant terrestrial vegetation) and possibly in the gameto phyte phase (although no such homoiohydric gametophytes are known today). The homoiohydric features of gas spaces, stomata, cuticle, endohydric water conducting system and water and nutrient uptake structures require an organism at least 5 mm high; this has implications for the minimum size of mega-spores and seeds. The tallest homoiohydric plants are (or were within historic times) 130 m high, with height constrained by re source costs of the synthesis and maintenance of the mechanical and water conduction systems, andbr of xylem water trans port. Secondarily poikilohydric embryophytes in aquatic, or very damp terrestrial, habitats are derived from homoiohydric plants; they retain most homoiohydric features but are not functionally homoiohydric. The smaller secondarily poikilohydric plants are less than one tenth of the size of the smallest functionally homoiohydric plants.  相似文献   

3.
The evolution of vascular plants and their spread across the land surface, beginning ~420 Ma, progressively increased the rate of weathering of phosphorus from rocks. This phosphorus supply promoted terrestrial and marine productivity and the burial of organic carbon, which has been the major source of O2 over geological timescales. Hence, it is predicted that the rise of plants led to an increase in the O2 content of the atmosphere from ~12 vol %, 570–400 Ma to its present level of ~21 vol % by ~340 Ma. Previous modelling studies suggest that O2 then rose to ~35 vol % ~300 Ma. Such high concentrations are difficult to reconcile with the known persistence of forests, because rising O2 increases the frequency and intensity of vegetation fires, tending to decrease biomass and cause ecological shifts toward faster regenerating ecosystems. Rising O2 also directly inhibits C3 photosynthetic carbon assimilation and increases the production of toxic reactive oxygen species in cells. These effects suppress plant‐induced phosphorus weathering and hence organic carbon burial, providing a sensitive negative feedback on O2. A revised model predicts that this mechanism could have regulated atmospheric O2 within the range 15–25 vol % for the last 350 million years.  相似文献   

4.
Summary Photon absorption and photosynthesis under conditions of light limitation were determined in six temperate marine macroalgae and eight submerged angiosperms. Photon absorption and photosynthetic efficiency based on incident light increased in proportion to chlorophyll density per area and approached saturation at the highest densities (300 mg chlorophyll m–2) encountered. Absorption and photosynthetic efficiency were higher in brown and red algae than in green algae and angiosperms for the same chlorophyll density because of absorption by accessory pigments. Among thin macroalgae and submerged angiosperms chlorophyll variations directly influence light absorption and photosynthesis, whereas terrestrial leaves have chlorophyll in excess and thus there is only a minor influence of pigment variability on light-limited photosynthesis. The quantum efficiency of photosynthesis averaged 0.062±0.019 (±SD) mol O2 mol–1 photons absorbed for macroalgae and, significantly less, 0.049±0.016 mol O2 mol–1 photons for submerged angiosperms. Of the measurements 80% were between 0.037 and 0.079 mol O2 mol–1 photons. The results are lower than values given in the literature for unicellular algae and terrestrial C3 species at around 0.1 mol O2 mol–1 photons, but resemble values for other marine macroalgae and terrestrial C4 species. The reason for these differences remains unknown, but may be sought for in differential operation of cyclic photophosphorylation and photorespiration.  相似文献   

5.
The induction of astaxanthin formation by reactive oxygen species in mixotrophic culture of Chlorococcum sp. was investigated. H2O2 (0.1 mM) enhanced the total astaxanthin formation from 5.8 to 6.5 mg g–1 cell dry wt. Fe2+ (0.5 mM) added to the medium with H2O2 (0.1 mM) further promoted astaxanthin formation to 7.1 mg g–1 cell dry wt. Similarly, Fe2+ (0.5 mM) together with methyl viologen (0.01 mM) promoted astaxanthin formation to 6.3 mg g–1 cell dry wt. In contrast, an addition of KI (1 mM), a specific scavenger for hydroxyl radicals (OH), together with H2O2 (0.1 mM) and Fe2+ (0.5 mM), to the medium decreased astaxanthin formation to 1.8 mg g–1 cell dry wt. KI (1 mM) also inhibited the enhancement of carotenogenesis by superoxide anion radicals (O2 ), with a decrease of astaxanthin formation to 1.7 mg g–1 cell dry wt. This suggested that O2 might be transformed to OH before promoting carotenogenesis in Chlorococcum sp.  相似文献   

6.
The induction of hydroxyl radical (OH) production via quinone redox cycling in white-rot fungi was investigated to improve pollutant degradation. In particular, we examined the influence of 4-methoxybenzaldehyde (anisaldehyde), Mn2+, and oxalate on Pleurotus eryngii OH generation. Our standard quinone redox cycling conditions combined mycelium from laccase-producing cultures with 2,6-dimethoxy-1,4-benzoquinone (DBQ) and Fe3+-EDTA. The main reactions involved in OH production under these conditions have been shown to be (i) DBQ reduction to hydroquinone (DBQH2) by cell-bound dehydrogenase activities; (ii) DBQH2 oxidation to semiquinone (DBQ) by laccase; (iii) DBQ autoxidation, catalyzed by Fe3+-EDTA, producing superoxide (O2) and Fe2+-EDTA; (iv) O2 dismutation, generating H2O2; and (v) the Fenton reaction. Compared to standard quinone redox cycling conditions, OH production was increased 1.2- and 3.0-fold by the presence of anisaldehyde and Mn2+, respectively, and 3.1-fold by substituting Fe3+-EDTA with Fe3+-oxalate. A 6.3-fold increase was obtained by combining Mn2+ and Fe3+-oxalate. These increases were due to enhanced production of H2O2 via anisaldehyde redox cycling and O2 reduction by Mn2+. They were also caused by the acceleration of the DBQ redox cycle as a consequence of DBQH2 oxidation by both Fe3+-oxalate and the Mn3+ generated during O2 reduction. Finally, induction of OH production through quinone redox cycling enabled P. eryngii to oxidize phenol and the dye reactive black 5, obtaining a high correlation between the rates of OH production and pollutant oxidation.The degradation of lignin and pollutants by white-rot fungi is an oxidative and rather nonspecific process based on the production of substrate free radicals (36). These radicals are produced by ligninolytic enzymes, including laccase and three kinds of peroxidases: lignin peroxidase, manganese peroxidase, and versatile peroxidase (VP) (23). The H2O2 required for peroxidase activities is provided by several oxidases, such as glyoxal oxidase and aryl-alcohol oxidase (AAO) (9, 18). This free-radical-based degradative mechanism leads to the production of a broad variety of oxidized compounds. Common lignin depolymerization products are aromatic aldehydes and acids, and quinones (34). In addition to their high extracellular oxidation potential, white-rot fungi show strong ability to reduce these lignin depolymerization products, using different intracellular and membrane-bound systems (4, 25, 39). Since reduced electron acceptors of oxidized compounds are donor substrates for the above-mentioned oxidative enzymes, the simultaneous actions of both systems lead to the establishment of redox cycles (35). Although the function of these redox cycles is not fully understood, they have been hypothesized to be related to further metabolism of lignin depolymerization products that require reduction to be converted in substrates of the ligninolytic enzymes (34). A second function attributed to these redox cycles is the production of reactive oxygen species, i.e., superoxide anion radicals (O2), H2O2, and hydroxyl radicals (OH), where lignin depolymerization products and fungal metabolites act as electron carriers between intracellular reducing equivalents and extracellular oxygen. This function has been studied in Pleurotus eryngii, whose ligninolytic system is composed of laccase (26), VP (24), and AAO (9). Incubation of this fungus with different aromatic aldehydes has been shown to provide extracellular H2O2 on a constant basis, due to the establishment of a redox cycle catalyzed by an intracellular aryl-alcohol dehydrogenase (AAD) and the extracellular AAO (7, 10). The process was termed aromatic aldehyde redox cycling, and 4-methoxybenzaldehyde (anisaldehyde) serves as the main Pleurotus metabolite acting as a cycle electron carrier (13). A second cyclic system, involving a cell-bound quinone reductase activity (QR) and laccase, was found to produce O2 and H2O2 during incubation of P. eryngii with different quinones (11). The process was described as the cell-bound divalent reduction of quinones (Q) by QR, followed by extracellular laccase oxidation of hydroquinones (QH2) into semiquinones (Q), which autoxidized to some extent, producing O2 (Q + O2 ⇆ Q + O2). H2O2 was formed by O2 dismutation (O2 + HO2 + H+ → O2 + H2O2). In an accompanying paper, we describe the extension of this O2 and H2O2 generation mechanism to OH radical production by the addition of Fe3+-EDTA to incubation mixtures of several white-rot fungi with different quinones (6). Among them, those derived from 4-hydroxyphenyl, guaiacyl, and syringyl lignin units were used: 1,4-benzoquinone (BQ), 2-methoxy-1,4-benzoquinone (MBQ), and 2,6-dimethoxy-1,4-benzoquinone (DBQ), respectively. Semiquinone autoxidation under these conditions was catalyzed by Fe3+-EDTA instead of being a direct electron transfer to O2. The intermediate Fe2+-EDTA reduced not only O2, but also H2O2, leading to OH radical production by the Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+).Although OH radicals are the strongest oxidants produced by white-rot fungi (2, 14), studies of their involvement in pollutant degradation are quite scarce. In this context, the objectives of this study were to (i) determine possible factors enhancing the production of OH radicals by P. eryngii via quinone redox cycling and (ii) test the validity of this inducible OH production mechanism as a strategy for pollutant degradation. Our selection of possible OH production promoters was guided by two observations (6). First, the redox cycle of benzoquinones working with washed P. eryngii mycelium is rate limited by hydroquinone oxidation, since the amounts of the ligninolytic enzymes that remained bound to the fungus under these conditions were not large. Second, H2O2 is the limiting reagent for OH production by the Fenton reaction.With these considerations in mind, anisaldehyde and Mn2+ were selected to increase H2O2 production. As mentioned above, anisaldehyde induces H2O2 production in P. eryngii via aromatic aldehyde redox cycling (7). Mn2+ has been shown to enhance H2O2 production during the oxidation of QH2 by P. eryngii laccase by reducing the O2 produced in the semiquinone autoxidation reaction (Mn2+ + O2 → Mn3+ + H2O2 + 2 H+) (26). Mn2+ was also selected to increase the hydroquinone oxidation rate, since this reaction has been shown to be propagated by the Mn3+ generated in the latter reaction (QH2 + Mn3+ → Q + Mn2+ + 2 H+). The replacement of Fe3+-EDTA by Fe3+-oxalate was also planned in order to increase the QH2 oxidation rate above that resulting from the action of laccase. Oxalate is a common extracellular metabolite of wood-rotting fungi to which the function of chelating iron and manganese has been attributed (16, 45). The use of Fe3+-oxalate and nonchelated Fe3+, both QH2 oxidants, has been proven to enable quinone redox cycling in fungi that do not produce ligninolytic enzymes, such as the brown-rot fungus Gloeophyllum trabeum (17, 40, 41). Finally, phenol and the azo dye reactive black 5 (RB5) were selected as model pollutants.  相似文献   

7.
A simple strategy for the induction of extracellular hydroxyl radical (OH) production by white-rot fungi is presented. It involves the incubation of mycelium with quinones and Fe3+-EDTA. Succinctly, it is based on the establishment of a quinone redox cycle catalyzed by cell-bound dehydrogenase activities and the ligninolytic enzymes (laccase and peroxidases). The semiquinone intermediate produced by the ligninolytic enzymes drives OH production by a Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+). H2O2 production, Fe3+ reduction, and OH generation were initially demonstrated with two Pleurotus eryngii mycelia (one producing laccase and versatile peroxidase and the other producing just laccase) and four quinones, 1,4-benzoquinone (BQ), 2-methoxy-1,4-benzoquinone (MBQ), 2,6-dimethoxy-1,4-benzoquinone (DBQ), and 2-methyl-1,4-naphthoquinone (menadione [MD]). In all cases, OH radicals were linearly produced, with the highest rate obtained with MD, followed by DBQ, MBQ, and BQ. These rates correlated with both H2O2 levels and Fe3+ reduction rates observed with the four quinones. Between the two P. eryngii mycelia used, the best results were obtained with the one producing only laccase, showing higher OH production rates with added purified enzyme. The strategy was then validated in Bjerkandera adusta, Phanerochaete chrysosporium, Phlebia radiata, Pycnoporus cinnabarinus, and Trametes versicolor, also showing good correlation between OH production rates and the kinds and levels of the ligninolytic enzymes expressed by these fungi. We propose this strategy as a useful tool to study the effects of OH radicals on lignin and organopollutant degradation, as well as to improve the bioremediation potential of white-rot fungi.White-rot fungi are unique in their ability to degrade a wide variety of organopollutants (36, 47), mainly due to the secretion of a low-specificity enzyme system whose natural function is the degradation of lignin (11). Components of this system include laccase and/or one or two types of peroxidase, such as lignin peroxidase (LiP), manganese peroxidase (MnP), and versatile peroxidase (VP) (31). Besides acting directly, the ligninolytic enzymes can bring about lignin and pollutant degradation through the generation of low-molecular-weight extracellular oxidants, including (i) Mn3+, (ii) free radicals from some fungal metabolites and lignin depolymerization products (7, 22), and (iii) oxygen free radicals, mainly hydroxyl radicals (OH) and lipid peroxidation radicals (21). Although OH radicals are the strongest oxidants found in cultures of white-rot fungi (1), studies of their involvement in pollutant degradation are scarce. One of the reasons is that the mechanisms proposed for OH production still await in vivo validation.Several potential sources of extracellular OH based on the Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+) have been postulated for white-rot fungi. In one case, an extracellular fungal glycopeptide has been shown to reduce O2 and Fe3+ to H2O2 and Fe2+ (45). Enzymatic sources include cellobiose dehydrogenase, LiP, and laccase. Among these, only cellobiose dehydrogenase is able to directly catalyze the formation of Fenton''s reagent (33). The ligninolytic enzymes, however, act as an indirect source of OH through the generation of Fe3+ and O2 reductants, such as formate (CO2) and semiquinone (Q) radicals. The first time evidence was provided that a ligninolytic enzyme was involved in OH production, oxalate was used to generate CO2 in a LiP reaction mediated by veratryl alcohol (4). The proposed mechanism consisted of the following cascade of reactions: production of veratryl alcohol cation radical (Valc+) by LiP, oxidation of oxalate to CO2 by Valc+, reduction of O2 to O2 by CO2, and a superoxide-driven Fenton reaction (Haber-Weiss reaction) in which Fe3+ was reduced by O2. The OH production mechanism assisted by Q was inferred from the oxidation of 2-methoxy-1,4-benzohydroquinone (MBQH2) and 2,6-dimethoxy-1,4-benzohydroquinone (DBQH2) by Pleurotus eryngii laccase in the presence of Fe3+-EDTA. The ability of Q radicals to reduce both Fe3+ to Fe2+ and O2 to O2, which dismutated to H2O2, was demonstrated (14). In this case, OH radicals were generated by a semiquinone-driven Fenton reaction, as Q radicals were the main agents accomplishing Fe3+ reduction. The first evidence of the likelihood of this OH production mechanism being operative in vivo had been obtained from incubations of P. eryngii with 2-methyl-1,4-naphthoquinone (menadione [MD]) and Fe3+-EDTA (15). Extracellular OH radicals were produced on a constant basis through quinone redox cycling, consisting of the reduction of MD by a cell-bound quinone reductase (QR) system, followed by the extracellular oxidation of the resulting hydroquinone (MDH2) to its semiquinone radical (MD). The production of extracellular O2 and H2O2 by P. eryngii via redox cycling involving laccase was subsequently confirmed using 1,4-benzoquinone (BQ), 2-methyl-1,4-benzoquinone, and 2,3,5,6-tetramethyl-1,4-benzoquinone (duroquinone), in addition to MD (16). However, the demonstration of OH production based on the redox cycling of quinones other than MD was still required.In the present paper, we describe the induction of extracellular OH production by P. eryngii upon its incubation with BQ, 2-methoxy-1,4-benzoquinone (MBQ), 2,6-dimethoxy-1,4-benzoquinone (DBQ), and MD in the presence of Fe3+-EDTA. The three benzoquinones were selected because they are oxidation products of p-hydroxyphenyl, guaiacyl, and syringyl units of lignin (MD was included as a positive control). Along with laccase, the involvement of P. eryngii VP in the production of O2 and H2O2 from hydroquinone oxidation has also been reported (13). Since hydroquinones are substrates of all known ligninolytic enzymes, quinone redox cycling catalysis could involve any of them. Here, we demonstrate OH production by P. eryngii under two different culture conditions, leading to the production of laccase or laccase and VP. We also show that quinone redox cycling is widespread among white-rot fungi by using a series of well-studied species that produce different combinations of ligninolytic enzymes.  相似文献   

8.
Reaction of sodium picolinate with FeIII oxo-centered carboxylate triangles in MeCN in the presence of PPh4Cl yields (PPh4)[Fe4O2(O2CR)7(pic)2] (R = Ph (1), But (2)). Omitting the phosphonium cation produces [Fe8Na4O4(O2CPh)16(pic)4(H2O)4] (3), which contains two Fe4Na2 units bridged by two picolinate ligands. X-ray crystal structures of 1 and 3 are reported.Voltammetric profiles in MeCN show four one-electron reduction steps for complexes 1 and 2. Variable-temperature magnetic susceptibility measurements in polycrystalline samples of 1 and 3 reveal strong antiferromagnetic couplings leading to = 0 ground states.  相似文献   

9.
Two iron(III) complexes, [Fe4OCl(O2CMe)3(O3PC6H9)3(py)5] (1) and [Fe7O2(O2CPh)9(O3PC6H9)4(py)6] (2), have been prepared through solution reactions of [Fe3O(O2CR)6(H2O)3]Cl (R = Me, Ph) with cyclohexenephosphonic acid. Both compounds contain triangular oxo-centered [Fe33-O)]7+ units. In complex 1, the fourth iron atom is capped on this triangular unit through O-P-O bridges, forming a tetranuclear cluster with a tetrahedral arrangement of iron atoms. In complex 2, two equivalent [Fe33-O)]7+ units are connected by the fourth iron atom through four phosphonate ligands, forming a heptanuclear cluster. Variable temperature susceptibility measurements were performed for 1 and 2. Both exhibit dominant antiferromagnetic interactions between the Fe(III) centers.  相似文献   

10.
Summary Fibrin-enrobed, commercially produced glycogen was treated, without prior glutaraldehyde fixation, to a form of post-fixation with solutions of OsVIIIO4 or with a mixture of either OsVIIIO4 plus K3FeIII(CN)6 or K2OsVIO4 plus K4FeII(CN)6.Only the last mixture gave constrast staining of the glycogen in unstained ultrathin sections. The first mixture rendered the glycogen just barely visible but the glycogen contrast was increased by lead staining. The glycogen treated with the OsVIIIO4 solution was not contrast stained and was just observable after lead staining.Qualitative X-ray microanalysis of the glycogen in the ultrathin sections confirmed the presence of osmium and iron in the glycogen treated with both mixtures. The glycogen treated with OsVIIIO4 alone was difficult to analyse.Quantitative X-ray microanalysis showed that, in the glycogen treated with the OsVIIIO4 mixture plus K3FeIII(CN)6, the mean atomic osmium to iron ratio was 15. In the glycogen treated with K2OsVIO4 plus K4FeII(CN)6 this ratio was 117. However, the mean net osmium intensity in the latter case was 15 times higher than in the former case and for the iron even 40 times higher.The Unit for Analytical Electron Microscopy was established by collaboration between the Erasmus University of Rotterdam (W. C. de Bruijn), the University of Leiden and the Organization for Health Research TNO. The analytical microscope was purchased with funds from the Netherlands Organization for Pure Scientific Research (ZWO).  相似文献   

11.
Interaction of the hexa-lacunary polyanion precursor [α-H2P2W12O48]12− and the FeIII in aqueous solution results in the formation of an equatorial tri-iron substituted Wells-Dawson type compound, K4Cs2Fe2[P2W15(FeOH)3O59]·22H2O (1). Compound 1 was characterized by IR, elemental, single-crystal X-ray diffraction, thermogravimetric, magnetic, as well as electrochemical analysis. The polyoxoanion [P2W15(FeOH)3O59]12− can be viewed as a derivative of the parent polyoxoanion [α-P2W18O62]6− by removal of three belt WO groups and then inhabited by three FeOH groups. The compound 1-modified carbon paste electrode (1-CPE) presents good electrocatalytic activity not only toward the reduction of nitrite which is attributed to the function of tungstophosphate, but also toward the oxidation of ascorbic acid which is primarily attributed to the function of FeIII. The magnetic properties of 1 have been studied by magnetic susceptibility and fitted according to an isotropic exchange model. Compound 1 exhibits strong antiferromagnetic spin exchange interactions between the FeIII centers.  相似文献   

12.
The extrinsic PsbU and PsbV proteins are known to play a critical role in stabilizing the Mn4CaO5 cluster of the PSII oxygen-evolving complex (OEC). However, most isolates of the marine cyanobacterium Prochlorococcus naturally miss these proteins, even though they have kept the main OEC protein, PsbO. A structural homology model of the PSII of such a natural deletion mutant strain (P. marinus MED4) did not reveal any obvious compensation mechanism for this lack. To assess the physiological consequences of this unusual OEC, we compared oxygen evolution between Prochlorococcus strains missing psbU and psbV (PCC 9511 and SS120) and two marine strains possessing these genes (Prochlorococcus sp. MIT9313 and Synechococcus sp. WH7803). While the low light-adapted strain SS120 exhibited the lowest maximal O2 evolution rates (Pmax per divinyl-chlorophyll a, per cell or per photosystem II) of all four strains, the high light-adapted strain PCC 9511 displayed even higher PChlmax and PPSIImax at high irradiance than Synechococcus sp. WH7803. Furthermore, thermoluminescence glow curves did not show any alteration in the B-band shape or peak position that could be related to the lack of these extrinsic proteins. This suggests an efficient functional adaptation of the OEC in these natural deletion mutants, in which PsbO alone is seemingly sufficient to ensure proper oxygen evolution. Our study also showed that Prochlorococcus strains exhibit negative net O2 evolution rates at the low irradiances encountered in minimum oxygen zones, possibly explaining the very low O2 concentrations measured in these environments, where Prochlorococcus is the dominant oxyphototroph.  相似文献   

13.
A mass spectrometric 16O2/18O2-isotope technique was used to analyse the rates of gross O2 evolution, net O2 evolution and gross O2 uptake in relation to photon fluence rate by Dunaliella tertiolecta adapted to 0.5, 1.0, 1.5, 2.0 and 2.5 M NaCl at 25°C and pH 7.0.At concentrations of dissolved inorganic carbon saturating for photosynthesis (200 M) gross O2 evolution and net O2 evolution increased with increasing salinity as well as with photon fluence rate. Light compensation was also enhanced with increased salinities. Light saturation of net O2 evolution was reached at about 1000 mol m-2s-1 for all salt concentrations tested. Gross O2 uptake in the light was increased in relation to the NaCl concentration but it was decreased with increasing photon fluence rate for almost all salinities, although an enhanced flow of light generated electrons was simultaneously observed. In addition, a comparison between gross O2 uptake at 1000 mol photons m-2s-1, dark respiration before illumination and immediately after darkening of each experiment showed that gross O2 uptake in the light paralleled but was lower than mitochondrial O2 consumption in the dark.From these results it is suggested that O2 uptake by Dunaliella tertiolecta in the light is mainly influenced by mitochondrial O2 uptake. Therefore, it appears that the light dependent inhibition of gross O2 uptake is caused by a reduction in mitochondrial O2 consumption by light.Abbreviations DCMU 3-(3, 4-dichlorophenyl)-1, 1-dimethylurea - DHAP dihydroxy-acetonephosphate - DIC dissolved inorganic carbon - DRa rate of dark respiration immediately after illumination - DRb rate of dark respiration before illumination - E0 rate of gross oxygen evolution in the light - NET rate of net oxygen evolution in the light - PFR photon fluence rate - RubP rubulose-1,5-bisphosphate - SHAM salicyl hydroxamic acid - U0 rate of gross oxygen uptake in the light  相似文献   

14.
To reveal the mechanisms of sedimental H2S accumulation, annual investigations on sedimental environments were conducted in two temperate estuarine lagoons. The lagoons, Gamo and Idoura (Japan), have similar shapes, locations, and topographical properties but different degrees of H2S accumulation. Water stagnation causes a high phytoplankton biomass (Chl. a; 26–52 g l–1) in the inner Gamo Lagoon. Gamo Lagoon sediment was characterized by high bounded sulfides (bounded Smainly FeS) and H2S contents, and low C/N ratios (mean = 10.4) and iron (reactive Fe2+ and total Fe) contents. H2S was not detected in Idoura Lagoon where phytoplankton biomass was much lower (Chl. a; 0.6–4 g l–1). Idoura Lagoon sediment had high C/N ratios (mean=17.9) and high iron contents. The C/N ratio difference implies that organic matter in Gamo Lagoon originates mainly from more decomposable phytoplankton, while organic matter in Idoura Lagoon derives mainly from terrestrial vascular plants with lower decomposability. The excess loading of phytoplanktonic detritus in Gamo accelerates sedimentary microbial activity, including sulfate reduction (i.e., H2S production). High Fe2+and low bounded S contents in Idoura sediment indicate a high chemical buffering capacity toward H2S. In contrast, almost all Fe2+ in Gamo Lagoon had already reacted with H2S as FeS. H2S accumulation in Gamo Lagoon is caused by low sedimentary chemical buffering capacity toward H2S, as well as higher microbial H2S production, caused by the excess loading of phytoplanktonic detritus.  相似文献   

15.
1. Arbuscular mycorrhizal fungi (AMF) commonly colonise isoetid species inhabiting oxygenated sediments in oligotrophic lakes but are usually absent in other submerged plants. We hypothesised that organic enrichment of oligotrophic lake sediments reduces AMF colonisation and hyphal growth because of sediment O2 depletion and low carbon supply from stressed host plants. 2. We added organic matter to sediments inhabited by isoetids and measured pore‐water chemistry (dissolved O2, inorganic carbon, Fe2+ and ), colonisation intensity of roots and hyphal density after 135 days of exposure. 3. Addition of organic matter reduced AMF colonisation of roots of both Lobelia dortmanna and Littorella uniflora, and high additions stressed the plants. Even small additions of organic matter almost stopped AMF colonisation of initially un‐colonised L. uniflora, though without reducing plant growth. Mean hyphal density in sediments was high (6 and 15 m cm?3) and comparable with that in terrestrial soils (2–40 m cm?3). Hyphal density was low in the upper 1 cm of isoetid sediments, high in the main root zone between 1 and 8 cm and positively related to root density. Hyphal surface area exceeded root surface area by 1.7–3.2 times. 4. We conclude that AMF efficiently colonise isoetids in oligotrophic sediments and form extensive hyphal networks. Small additions of organic matter to sediments induce sediment anoxia and reduce AMF colonisation of roots but cause no apparent plant stress. High organic addition induces night‐time anoxia in both the sediment and the plant tissue. Tissue anoxia reduces root growth and AMF colonisation, probably because of restricted translocation of nutrient ions and organic solutes between roots and leaves. Isoetids should rely on AMF for P uptake on nutrient‐poor mineral sediments but are capable of growing without AMF on organic sediments.  相似文献   

16.
Yeast cytochrome c peroxidase was used to construct a model for the reactions catalyzed by the second cycle of nitric oxide synthase. The R48A/W191F mutant introduced a binding site for N-hydroxyguanidine near the distal heme face and removed the redox active Trp-191 radical site. Both the R48A and R48A/W191F mutants catalyzed the H2O2 dependent conversion of N-hydroxyguanidine to N-nitrosoguanidine. It is proposed that these reactions proceed by direct one-electron oxidation of NHG by the Fe+4O center of either Compound I (Fe+4O, porph+) or Compound ES (Fe+4O, Trp+). R48A/W191F formed a Fe+2O2 complex upon photolysis of Fe+2CO in the presence of O2, and N-hydroxyguanidine was observed to react with this species to produce products, distinct from N-nitrosoguanidine, that gave a positive Griess reaction for nitrate + nitrite, a positive Berthelot reaction for urea, and no evidence for formation of NO. It is proposed that HNO and urea are produced in analogy with reactions of nitric oxide synthase in the pterin-free state.  相似文献   

17.
Mitochondria from respiring cells were isolated under anaerobic conditions. Microscopic images were largely devoid of contaminants, and samples consumed O2 in an NADH-dependent manner. Protein and metal concentrations of packed mitochondria were determined, as was the percentage of external void volume. Samples were similarly packed into electron paramagnetic resonance tubes, either in the as-isolated state or after exposure to various reagents. Analyses revealed two signals originating from species that could be removed by chelation, including rhombic Fe3+ (g = 4.3) and aqueous Mn2+ ions (g = 2.00 with Mn-based hyperfine). Three S = 5/2 signals from Fe3+ hemes were observed, probably arising from cytochrome c peroxidase and the a3:Cub site of cytochrome c oxidase. Three Fe/S-based signals were observed, with averaged g values of 1.94, 1.90 and 2.01. These probably arise, respectively, from the [Fe2S2]+ cluster of succinate dehydrogenase, the [Fe2S2]+ cluster of the Rieske protein of cytochrome bc 1, and the [Fe3S4]+ cluster of aconitase, homoaconitase or succinate dehydrogenase. Also observed was a low-intensity isotropic g = 2.00 signal arising from organic-based radicals, and a broad signal with g ave = 2.02. Mössbauer spectra of intact mitochondria were dominated by signals from Fe4S4 clusters (60–85% of Fe). The major feature in as-isolated samples, and in samples treated with ethylenebis(oxyethylenenitrilo)tetraacetic acid, dithionite or O2, was a quadrupole doublet with ΔE Q = 1.15 mm/s and δ = 0.45 mm/s, assigned to [Fe4S4]2+ clusters. Substantial high-spin non-heme Fe2+ (up to 20%) and Fe3+ (up to 15%) species were observed. The distribution of Fe was qualitatively similar to that suggested by the mitochondrial proteome.  相似文献   

18.
Chloroplasts with high rates of photosynthetic O2 evolution (up to 120 mol O2· (mg Chl)-1·h-1 compared with 130 mol O2· (mg Chl)-1·h-1 of whole cells) were isolated from Chlamydomonas reinhardtii cells grown in high and low CO2 concentrations using autolysine-digitonin treatment. At 25° C and pH=7.8, no O2 uptake could be observed in the dark by high- and low-CO2 adapted chloroplasts. Light saturation of photosynthetic net oxygen evolution was reached at 800 mol photons·m-2·s-1 for high- and low-CO2 adapted chloroplasts, a value which was almost identical to that observed for whole cells. Dissolved inorganic carbon (DIC) saturation of photosynthesis was reached between 200–300 M for low-CO2 adapted chloroplasts, whereas high-CO2 adapted chloroplasts were not saturated even at 700 M DIC. The concentrations of DIC required to reach half-saturated rates of net O2 evolution (Km(DIC)) was 31.1 and 156 M DIC for low- and high-CO2 adapted chloroplasts, respectively. These results demonstrate that the CO2 concentration provided during growth influenced the photosynthetic characteristics at the whole cell as well as at the chloroplast level.Abbreviations Chl chlorophyll - DIC dissolved inorganic carbon - Km(DIC) coneentration of dissolved inorganic carbon required for the rate of half maximal net O2 evolution - PFR photon fluence rate - SPGM silicasol-PVP-gradient medium  相似文献   

19.
The ability of paraquat radicals (PQ+.) generated by xanthine oxidase and glutathione reductase to give H2O2-dependent hydroxyl radical production was investigated. Under anaerobic conditions, paraquat radicals from each source caused chain oxidation of formate to CO2, and oxidation of deoxyribose to thiobarbituric acid-reactive products that was inhibited by hydroxyl radical scavengers. This is in accordance with the following mechanism derived for radicals generated by γ-irradiation [H. C. Sutton and C. C. Winterbourn (1984) Arch. Biochem. Biophys.235, 106–115] PQ+. + Fe3+ (chelate) → Fe2+ (chelate) + PQ++ H2O2 + Fe2+ (chelate) → Fe3+ (chelate) + OH? + OH.. Iron-(EDTA) and iron-(diethylenetriaminepentaacetic acid) (DTPA) were good catalysts of the reaction; iron complexed with desferrioxamine or transferrin was not. Extremely low concentrations of iron (0.03 μm) gave near-maximum yields of hydroxyl radicals. In the absence of added chelator, no formate oxidation occurred. Paraquat radicals generated from xanthine oxidase (but not by the other methods) caused H2O2-dependent deoxyribose oxidation. However, inhibition by scavengers was much less than expected for a reaction of hydroxyl radicals, and this deoxyribose oxidation with xanthine oxidase does not appear to be mediated by free hydroxyl radicals. With O2 present, no hydroxyl radical production from H2O2 and paraquat radicals generated by radiation was detected. However, with paraquat radicals continuously generated by either enzyme, oxidation of both formate and deoxyribose was measured. Product yields decreased with increasing O2 concentration and increased with increasing iron(DTPA). These results imply a major difference in reactivity between free and enzymatically generated paraquat radicals, and suggest that the latter could react as an enzyme-paraquat radical complex, for which the relative rate of reaction with Fe3+ (chelate) compared with O2 is greater than is the case with free paraquat radicals.  相似文献   

20.
Oxygen and CO2 exchange were measured concurrently in leaves of shade-grownAlocasia macrorrhiza (L.) G. Don during lightflecks consisting of short periods of high photon flux density (PFD) superimposed on a low-PFD background illumination. Oxygen exchange was measured with a zirconium-oxide ceramic cell in an atmosphere containing 1 600 bar O2 and 350 bar CO2. Following an increase in PFD from 10 to 500 mol photons·m-2·s-1, O2 evolution immediately increased to a maximum rate that was about twice as high as the highest CO2-exchange rates that were observed. Oxygen evolution then decreased over the next 5–10 s to rates equal to the much more slowly increasing rates of CO2 uptake. When the PFD was decreased at the end of a lightfleck, O2 evolution decreased nearly instantaneously to the low-PFD rate while CO2 fixation continued at an elevated rate for about 20 s. When PFD during the lightfleck was at a level that was limiting for steady-state CO2 exchange, then the O2-evolution rate was constant during the lightfleck. This observed pattern of O2 evolution during lightflecks indicated that the maximum rate of electron transport exceeded the maximum rate of CO2 fixation in these leaves. In noninduced leaves, rates of O2 evolution for the first fraction of a second were about as high as rates in fully induced leaves, indicating that O2 evolution and the electron-transport chain are not directly affected by the leaf's induction state. Severalfold differences between induced and noninduced leaves in O2 evolution during a lightfleck were seen for lightflecks longer than a few seconds where the rate of O2 evolution appeared to be limited by the utilization of reducing power in CO2 fixation.Abbreviation PFD photon flux density (of photosynthetically active radiation)  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