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.
The green alga Chlamydomonas reinhardtii Dangeard CW-15 exhibited very low rates of plasma-membrane Fe(III) reductase activity when grown under Fe-sufficient conditions. After switching the medium to an Fe-free formulation, both ferricyanide reductase and ferric chelate reductase activities rapidly increased, reaching a maximum after 3 d under iron-free conditions. Both of the Fe(III) reductase activities increased in parallel over time, they exhibited similar K m values (approximately 10 μM) with respect to Fe(III), displayed the same pH profile of activity, and both exhibited the same degree of light stimulation which could be inhibited by 3-(3′,4′-dichlorophenyl)-1,1-dimethylurea (DCMU). Furthermore, ferricyanide competitively inhibited ferric chelate reduction by iron-limited cells. These results indicate that both Fe(III) reductase activities were mediated by the same iron-limitation-induced plasma-membrane reductase. No evidence was found for the presence of Fe(III)-reducing substances in the culture medium, or for the involvement of active oxygen species in the process of Fe(III) reduction. Chlamydomonas reinhardtii appears to respond to iron limitation in a manner similar to Strategy I higher plants. Received: 24 June 1997 / Accepted: 2 August 1997  相似文献   

2.
3.
The properties of NADH-dependent Fe3+-EDTA reductase in plasma membranes (PM) from roots of iron-deficient and -sufficient tomato plants [Lycopersicon esculentum L. (Mill.) cv. Abunda] were examined. Iron deficiency resulted in a 3-fold increase of in vivo root iron-chelate reductase activity with a Km (Fe3+-EDTA) of 230 μM. In purified root PM, average specific activities of ferric chelate reductase of 410 and 254 nmol Fe (mg protein)?1 min?1 were obtained for iron-deficient and -sufficient plants, respectively. In both cases, the PM-bound activity showed a pH optimum at pH 6.8. Activity depended on NADH and not on NADPH and on the presence of detergent. The activity was inhibited 40-50% by superoxide dismutase (EC 1.15.1.1) and ca 30% by oxygen. Kinetic analysis of the membrane-bound enzyme revealed a Km (Fe3+-EDTA) of ca 200 μM for both iron-stressed and -sufficient plants. For NADH, Km values around 230 μM were obtained. The ferric chelate reductase could be solubilised from salt-washed PM with Triton X-100 at a protein:detergent ratio of 1:2.8 (w/w). The Triton-soluble fraction revealed one enzyme-stained band in native polyacrylamide electrophoresis. Although the membranes showed no nitrate reductase (NR; EC 1.6.6.1) activity, anti-spinach NR immunoglobulin G (IgG) recognized a 54 kDa band both in the PM and the Triton-soluble fraction, but not in the enzymatically active material obtained from the native gel. No evidence could be found for the synthesis of a new, biochemically distinct PM-bound ferric chelate reductase under iron deficiency, which might be identified as the so-called Turbo reductase. It is concluded that iron deficiency in tomato induces increased expression of a ferric chelate reductase in root PM, which is already present in iron-sufficient plants and probably also in plants, which do not contain the Turbo reductase, like the grasses. The iron reductase is not identical with the recently reported PM-associated nitrate reductase.  相似文献   

4.
Under iron limitation, bacteria scavenge ferric (Fe3+) iron bound to siderophores or other chelates from the environment to fulfill their nutritional requirement. In gram‐negative bacteria, the siderophore uptake system prototype consists of an outer membrane transporter, a periplasmic binding protein and a cytoplasmic membrane transporter, each specific for a single ferric siderophore or siderophore family. Here, we show that spontaneous single gain‐of‐function missense mutations in outer membrane transporter genes of Bradyrhizobium japonicum were sufficient to confer on cells the ability to use synthetic or natural iron siderophores, suggesting that selectivity is limited primarily to the outer membrane and can be readily modified. Moreover, growth on natural or synthetic chelators required the cytoplasmic membrane ferrous (Fe2+) iron transporter FeoB, suggesting that iron is both dissociated from the chelate and reduced to the ferrous form within the periplasm prior to cytoplasmic entry. The data suggest rapid adaptation to environmental iron by facile mutation of selective outer membrane transporter genes and by non‐selective uptake components that do not require mutation to accommodate new iron sources.  相似文献   

5.
Since the recognition of iron‐limited high nitrate (or nutrient) low chlorophyll (HNLC) regions of the ocean, low iron availability has been hypothesized to limit the assimilation of nitrate by diatoms. To determine the influence of non‐steady‐state iron availability on nitrogen assimilatory enzymes, cultures of Thalassiosira weissflogii (Grunow) Fryxell et Hasle were grown under iron‐limited and iron‐replete conditions using artificial seawater medium. Iron‐limited cultures suffered from decreased efficiency of PSII as indicated by the DCMU‐induced variable fluorescence signal (Fv/Fm). Under iron‐replete conditions, in vitro nitrate reductase (NR) activity was rate limiting to nitrogen assimilation and in vitro nitrite reductase (NiR) activity was 50‐fold higher. Under iron limitation, cultures excreted up to 100 fmol NO2?·cell?1·d?1 (about 10% of incorporated N) and NiR activities declined by 50‐fold while internal NO2? pools remained relatively constant. Activities of both NR and NiR remained in excess of nitrogen incorporation rates throughout iron‐limited growth. One possible explanation is that the supply of photosynthetically derived reductant to NiR may be responsible for the limitation of nitrogen assimilation at the NO2? reduction step. Urease activity showed no response to iron limitation. Carbon:nitrogen ratios were equivalent in both iron conditions, indicating that, relative to carbon, nitrogen was assimilated at similar rates whether iron was limiting growth or not. We hypothesize that, diatoms in HNLC regions are not deficient in their ability to assimilate nitrate when they are iron limited. Rather, it appears that diatoms are limited in their ability to process photons within the photosynthetic electron transport chain which results in nitrite reduction becoming the rate‐limiting step in nitrogenassimilation.  相似文献   

6.
The response to iron deficiency was investigated in 16 ecotypes of Arabidopsis thaliana (L.) Heynh. and in Arabidopsis griffithiana. An increase in root ferric reductase activity was observed under conditions of iron deficiency in these ecotypes and in both species. This observation is consistent with a Strategy I response which is typical for dicot plants. A. griffithiana, however, showed a lower induction of ferric reductase activity in response to iron deficiency than that of the commonly studied A. thaliana Columbia ecotypes.  相似文献   

7.
Reduction of Fe(III) to Fe(II) by Fe(III) chelate reductase is thought to be an obligatory step in iron uptake as well as the primary factor in making iron available for absorption by all plants except grasses. Fe(III) chelate reductase has also been suggested to play a more general role in the regulation of cation absorption. In order to experimentally address the importance of Fe(III) chelate reductase activity in the mineral nutrition of plants, three Arabidopsis thaliana mutants (frd1-1, frd1-2 and frd1-3), that do not show induction of Fe(III) chelate reductase activity under iron-deficient growth conditions, have been isolated and characterized. These mutants are still capable of acidifying the rhizosphere under iron-deficiency and accumulate more Zn and Mn in their shoots relative to wild-type plants regardless of iron status. frd1 mutants do not translocate radiolabeled iron to the shoots when roots are presented with a tightly chelated form of Fe(III). These results: (1) confirm that iron must be reduced before it can be transported, (2) show that Fe(III) reduction can be uncoupled from proton release, the other major iron-deficiency response, and (3) demonstrate that Fe(III) chelate reductase activity per se is not necessarily responsible for accumulation of cations previously observed in pea and tomato mutants with constitutively high levels of Fe(III) chelate reductase activity.  相似文献   

8.
Harold G. Weger 《Planta》1999,207(3):377-384
Cells of the green alga Chlamydomonas reinhardtii Dangeard were grown in Fe-limited chemostat culture over a range of growth rates (0.15–1.5 d−1). Greater cell densities and culture chlorophyll levels were achieved using an excess of chelator [ethylenediamine di-(o-hydroxyphenylacetic acid)] relative to FeCl3 (80:1), compared to growth using a 1:1 chelator:FeCl3 ratio. The C. reinhardtii cells reduced extracellular ferric chelates, and ferric chelate reductase activity increased with increasing Fe-limited growth rates. However Fe-sufficient cells exhibited a low rate of ferric chelate reductase activity, similar to severely Fe-limited cells. Iron-limited cells were capable of reducing a wide variety of ferric chelates, representing a wide range of stability constants, at similar rates, suggesting that the stability constants of ferric complexes are not important determinants of ferric reducing activity. Cupric reductase activity also increased with increasing Fe-limited growth rates, and Cu(II) was preferentially reduced compared to Fe(III). These results suggest that both reductase activities may represent the same plasma-membrane enzyme. The rate of cupric reduction was a function of the free [Cu2+], not the total [Cu(II)], suggesting that free Cu2+ is the actual substrate for cupric reductase activity. Received: 8 July 1998 / Accepted: 5 August 1998  相似文献   

9.
Two free flavin-independent enzymes were purified by detecting the NAD(P)H oxidation in the presence of Fe(III)-EDTA and t-butyl hydroperoxide from E. coli. The enzyme that requires NADH or NADPH as an electron donor was a 28 kDa protein, and N-terminal sequencing revealed it to be oxygen-insensitive nitroreductase (NfnB). The second enzyme that requires NADPH as an electron donor was a 30 kDa protein, and N-terminal sequencing revealed it to be ferredoxin-NADP+ reductase (Fpr). The chemical stoichiometry of the Fenton activities of both NfnB and Fpr in the presence of Fe(III)-EDTA, NAD(P)H and hydrogen peroxide was investigated. Both enzymes showed a one-electron reduction in the reaction forming hydroxyl radical from hydrogen peroxide. Also, the observed Fenton activities of both enzymes in the presence of synthetic chelate iron compounds were higher than their activities in the presence of natural chelate iron compounds. When the Fenton reaction occurs, the ferric iron must be reduced to ferrous iron. The ferric reductase activities of both NfnB and Fpr occurred with synthetic chelate iron compounds. Unlike NfnB, Fpr also showed the ferric reductase activity on an iron storage protein, ferritin, and various natural iron chelate compounds including siderophore. The Fenton and ferric reductase reactions of both NfnB and Fpr occurred in the absence of free flavin. Although the k cat/K m value of NfnB for Fe(III)-EDTA was not affected by free flavin, the k cat/K m value of Fpr for Fe(III)-EDTA was 12-times greater in the presence of free FAD than in the absence of free FAD.  相似文献   

10.
Arabidopsis thaliana (L.) Heynh. Columbia wild type and a root hair-less mutant RM57 were grown on iron-containing and iron-deficient nutrient solutions. In both genotypes, ferric chelate reductase (FCR) of intact roots was induced upon iron deficiency and followed a Michaelis-Menten kinetic with a K m of 45 and 54 M FeIII-EDTA and a V max of 42 and 33 nmol Fe2+·(g FW)–1·min–1 for the wild type and the mutant, respectively. The pH optimum for the reaction was around pH 5.5. The approximately four fold stimulation of FCR activity was independent of formation of root hairs and/or transfer cells induced by iron deficiency. Iron-deficiency-induced chlorosis and the development of a rigid root habit disappeared when ferric chelate was applied to the leaves, while FCR activity remained unchanged. The time course of the responses to iron deficiency showed that morphological and physiological responses were controlled separately.Abbreviations FCR ferric chelate reductase - FW fresh weight Thanks are due to Klaas Sjollema (Department of Electronmicroscopy, University of Groningen, The Netherlands) for help with the electron microscopy sample preparation and especially to Dr. Uwe Santore (Heinrich-Heine-University for electron microscopy. This work was supported by the SCIENCE programm of the European community; P.R.M.) and a Personal Research Grant by the Ministerium für Wissenschaft und Forschung of Nordrhein-Westfalen (P.R.M.) and last, not least by the productive discussions in ECOTRANS B.V.  相似文献   

11.
Iron acquisition from various ferric chelates and colloids was studied using iron‐limited cells of Anabaena flos‐aquae (Lyng.) Brèb UTEX 1444, a cyanobacterial strain that produces high levels of siderophores under iron limitation. Various chelators of greatly varying affinity for Fe3+ (HEDTA, EDDHA, desferrioxamine mesylate, HBED, 8‐hydroxyquinoline) were assayed for the degree of iron acquisition by iron‐limited cyanobacterial cells. Iron uptake rates (measured by graphite furnace atomic absorption spectrometry) varied approximately inversely with calculated [Fe3+] (calculated as pFe) and decreased with increasing chelator‐to‐iron ratio. No iron uptake was observed when Fe3+ was chelated with HBED, the strongest of the tested chelators. Iron‐limited Anabaena cells were able to take up iron from 8‐hydroxyquinoline (oxine or 8HQ), a compound sometimes used to quantify aquatic iron bioavailability. Iron bound to purified humic acid was poorly available but did support some growth at high humic acid concentrations. These results suggest that for cyanobacteria, even tightly bound iron is biologically available, including to a limited extent iron bound to humic acids. However, iron bound to some extremely strong chelators (e.g. HBED) is likely to be biologically unavailable.  相似文献   

12.
The Arabidopsis gene FRO6(AtFRO6) encodes ferric chelate reductase and highly expressed in green tissues of plants. We have expressed the gene AtFRO6 under the control of a 35S promoter in transgenic tobacco plants. High-level expression of AtFRO6 in transgenic plants was confirmed by northern blot analysis. Ferric reductase activity in leaves of transgenic plants grown under iron-sufficient or iron-deficient conditions is 2.13 and 1.26 fold higher than in control plants respectively. The enhanced ferric reductase activity led to increased concentrations of ferrous iron and chlorophyll, and reduced the iron deficiency chlorosis in the transgenic plants, compared to the control plants. In roots, the concentration of ferrous iron and ferric reductase activity were not significantly different in the transgenic plants compared to the control plants. These results suggest that FRO6 functions as a ferric chelate reductase for iron uptake by leaf cells, and overexpression of AtFRO6 in transgenic plants can reduce iron deficiency chlorosis.  相似文献   

13.
A novel catechol-substituted cephalosporin, S-9096, showed potent antibacterial activity against Pseudomonas aeruginosa under both iron-deficient and aerobic conditions. S-9096 and ferric iron formed a chelate complex at the molar ratio of 3 to 1, which could be incorporated into P. aeruginosa cells grown under such conditions. Incorporation decreased when the cells were grown under either iron-sufficient or anaerobic conditions, with a concomitant disappearance of iron-regulated outer membrane proteins that were considered to function as receptors for ferric siderophores. These results indicated that the ferric chelate of S-9096 was incorporated into P. aeruginosa cells via a ferric iron transport pathway, which caused the high antibacterial potency of S-9096. All of the S-9096-resistant mutants that were able to grow even under iron-deficient conditions lacked an iron-regulated outer membrane protein having an apparent molecular mass of 66 kDa, suggesting the role of this protein as a receptor for the ferric chelate of S-9096. Correspondence to: Y. Yamano  相似文献   

14.
Redox enzymes in the plant plasma membrane and their possible roles   总被引:1,自引:0,他引:1  
Purified plasma membrane (PM) vesicles from higher plants contain redox proteins with low‐molecular‐mass prosthetic groups such as flavins (both FMN and FAD), hemes, metals (Cu, Fe and Mn), thiol groups and possibly naphthoquinone (vitamin K1), all of which are likely to participate in redox processes. A few enzymes have already been identified: Monodehydroascorbate reductase (EC 1.6.5.4) is firmly bound to the cytosolic surface of the PM where it might be involved in keeping both cytosolic and, together with a b‐type cytochrome, apoplastic ascorbate reduced. A malate dehydrogenase (EC 1.1.1.37) is localized on the inner side of the PM. Several NAD(P)H‐quinone oxidoreductases have been purified from the cytocolic surface of the PM, but their function is still unknown. Different forms of nitrate reductase (EC 1.6.6.1–3) are found attached to, as well as anchored in, the PM where they may act as a nitrate sensor and/or contribute to blue‐light perception, although both functions are speculative. Ferric‐chelate‐reducing enzymes (EC 1.6.99.13) are localized and partially characterized on the inner surface of the PM but they may participate only in the reduction of ferric‐chelates in the cytosol. Very recently a ferric‐chelate‐reducing enzyme containing binding sites for FAD, NADPH and hemes has been identified and suggested to be a trans‐PM protein. This enzyme is involved in the reduction of apoplastic iron prior to uptake of Fe2+ and is induced by iron deficiency. The presence of an NADPH oxidase, similar to the so‐called respiratory burst oxidase in mammals, is still an open question. An auxin‐stimulated and cyanide‐insensitive NADH oxidase (possibly a protein disulphide reductase) has been characterized but its identity is still awaiting independent confirmation. Finally, the only trans‐PM redox protein which has been partially purified from plant PM so far is a high‐potential and ascorbate‐reducible b‐type cytochrome. In co‐operation with vitamin K1 and an NAD(P)H‐quinone oxidoreductase, it may participate in trans‐PM electron transport.  相似文献   

15.
16.
17.
18.
The development of plasma membrane-associated iron(III) reductase activity was characterized in root systems of Pisum sativum during the first 2 wk of growth, as plants were challenged with iron-deficiency stress. Plants of a parental genotype (cv. Sparkle) and a functional iron-deficiency mutant genotype (E107) were grown hydroponically with or without supplemental iron. Iron(III) reductase activity was visualized by placing the roots in an agarose matrix containing 0.2 idm Fe(III)-ethylenediaminetetraacetic acid and 0.3 mM Na2-bathophenanthrolinedisulfonic acid (BPDS). Red staining patterns, resulting from the formation of Fe(II)-BPDS, were used to identify iron(III)-reducing regions. Iron(III) reduction was extensive on roots of E107 as early as d 7, but not until d 11 for -Fe-treated Sparkle. Roots of +Fe-treated Sparkle showed limited regions of reductase activity throughout the period of study. For secondary lateral roots, iron(III) reduction was found for all growth types except + Fe-treated Sparkle. Treating Sparkle plants alternately to a cycle of iron deficiency, iron sufficiency, and iron deficiency revealed that reductase activity at a given root zone could be alternatively present, absent, and again present. Our results suggest that for Pisum roots grown under the present conditions, iron-deficiency stress induces the activation of iron(III) reductase capacity within 2 d.  相似文献   

19.
Stichococcus, a genus of green algae, distributes in ice-free areas throughout Antarctica. To understand adaptive strategies of Stichococcus to permanently cold environments, the physiological responses to temperature of two psychrotolerants, S. bacillaris NJ-10 and S. minutus NJ-17, isolated from rock surfaces in Antarctica were compared with that of one temperate S. bacillaris FACHB753. Two Antarctic Stichococcus strains grew at temperature from 4 to 25°C, while the temperate strain could grow above 30°C but could not survive at 4°C. The photosynthetic activity of FACHB753 at lower than 10°C was less than that of Antarctic algae. Nitrate reductase in NJ-10 and NJ-17 had its optimal temperature at 20°C, in comparison, the maximal activity of nitrate reductase in FACHB753 was found at 25°C. When cultured at 4–15°C a large portion of unsaturated fatty acids in the two Antarctic species was detected and the regulation of the degree of unsaturation of fatty acids by temperature was observed only above 15°C, though the content of the major unsaturated fatty acid αC18:3 in FACHB753 decreased with the temperatures elevated from 10 to 25°C. Elevated nitrate reductase activity and photosynthetic rates at low temperatures together with the high proportion of unsaturated fatty acids contribute to the ability of the Antarctic Stichococcus to thrive.  相似文献   

20.
The distributions of invasive Neogobius species were investigated in the Slovak section of the River Danube from Bratislava downstream to the village of Chl'aba. During October 2004, the main channel of the Danube was sampled, including by‐pass, head‐race and tail‐race canals of the Gab?íkovo dam, backwaters and the lower‐most sections of the tributaries Malý Dunaj, Hron, Váh and Ipel’. Three Neogobius species already documented in Slovakia were captured (monkey goby Neogobius fluviatilis, bighead goby N. kessleri, round goby N. melanostomus), with the latter two species being found in almost all stretches of the Slovak Danube. Monkey goby had a most limited distribution, and no racer goby N. gymnotrachelus were observed. The abundance of particular Neogobius species appeared to depend on the character of the shoreline habitat, and a possible association between larger towns and the abundance of bighead and round gobies requires further investigation.  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