首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The molecular mechanisms of iron trafficking in neurons have not been elucidated. In this study, we characterized the expression and localization of ferrous iron transporters Zip8, Zip14 and divalent metal transporter 1 (DMT1), and ferrireductases Steap2 and stromal cell‐derived receptor 2 in primary rat hippocampal neurons. Steap2 and Zip8 partially co‐localize, indicating these two proteins may function in Fe3+ reduction prior to Fe2+ permeation. Zip8, DMT1, and Steap2 co‐localize with the transferrin receptor/transferrin complex, suggesting they may be involved in transferrin receptor/transferrin‐mediated iron assimilation. In brain interstitial fluid, transferring‐bound iron (TBI) and non‐transferrin‐bound iron (NTBI) exist as potential iron sources. Primary hippocampal neurons exhibit significant iron uptake from TBI (Transferrin‐59Fe3+) and NTBI, whether presented as 59Fe2+‐citrate or 59Fe3+‐citrate; reductase‐independent 59Fe2+ uptake was the most efficient uptake pathway of the three. Kinetic analysis of Zn2+ inhibition of Fe2+ uptake indicated that DMT1 plays only a minor role in the uptake of NTBI. In contrast, localization and knockdown data indicate that Zip8 makes a major contribution. Data suggest also that cell accumulation of 59Fe from TBI relies at least in part on an endocytosis‐independent pathway. These data suggest that Zip8 and Steap2 play a major role in iron accumulation from NTBI and TBI by hippocampal neurons.

  相似文献   


2.
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   

3.
The thermostabilities of Fe2+ ligation in rubredoxins (Rds) from the hyperthermophile Pyrococcus furiosus (Pf) and the mesophiles Clostridium pasteurianum (Cp) and Desulfovibrio vulgaris (Dv) were compared. Residue 44 forms an NH...S(Cys) hydrogen bond to one of the cysteine ligands to the [Fe(SCys)4] site, and substitutions at this location affect the redox properties of the [Fe(SCys)4] site. Both Pf Rd and Dv Rd have an alanine residue at position 44, whereas Cp Fd has a valine residue. Wild-type proteins were examined along with V44A and A44V exchange mutants of Cp and Pf Rds, respectively, in order to assess the effects of the residue at position 44 on the stability of the [Fe(SCys)4] site. Stability of iron ligation was measured by temperature-ramp and fixed-temperature time course experiments, monitoring iron release in both the absence and presence of more thiophilic metals (Zn2+, Cd2+) and over a range of pH values. The thermostability of the polypeptide fold was concomitantly measured by fluorescence, circular dichroism, and 1H NMR spectroscopies. The A44V mutation strongly lowered the stability of the [FeII(SCys)4] site in Pf Rd, whereas the converse V44A mutation of Cp Rd significantly raised the stability of the [FeII(SCys)4] site, but not to the levels measured for wild-type Dv Rd. The region around residue 44 is thus a significant contributor to stability of iron coordination in reduced Rds. This region, however, made only a minor contribution to the thermostability of the protein folding, which was found to be higher for hyperthermophilic versus mesophilic Rds, and largely independent of the residue at position 44. These results, together with our previous studies, show that localized charge density, solvent accessibility, and iron site/backbone interactions control the thermostability of the [Fe(SCys)4] site. The iron site thermostability does make a minor contribution to the overall Rd thermostability. From a mechanistic standpoint, we also found that attack of displacing ions (H+, Cd2+) on the Cys42 sulfur ligand at the [Fe(SCys)4] site occurs through the V8 side and not the V44 side of the iron site.Electronic Supplementary Material Supplementary material is available in the online version of this article at http://dx.doi.org/10.1007/s00775-004-0525-4Abbreviations BPS bathophenanthroline sulfonate, sodium salt - Cp Rd (Pf Rd, Dv Rd) recombinant rubredoxin from Clostridium pasteurianum (Pyrococcus furiosus, Desulfovibrio vulgaris) - HEPES hydroxyethylpiperazineethanesulfonic acid - MES morpholinoethanesulfonic acid - Tris tris(hydroxymethyl)aminomethane - wt wild-type - ZnRd recombinant rubredoxin containing a [Zn(SCys)4] site  相似文献   

4.
Iron-efficient (WF9 corn and Coker 227 oat) and Fe-inefficient (ys1 corn and TAM 0–312 oat) cultivars were comparatively tested for their response to Fe-deficiency stress induced by the use of either ferrous or ferric chelators. Corn and oats were grown in 20 M Fe with 0, 60, and 120 M BPDS and 40 M Fe with 0, 120, and 240 M BPDS and 20 M Fe with 0 and 40 M EDDHA. All four cultivars tested, both Fe-efficient and Fe-inefficient, continuously reduced Fe3+ to Fe2+ at a low level as evidenced by the production of Fe2+ (BPDS)3 in test nutrient solutions over time. Severity of chlorosis increased as more BPDS was added to the nutrient solutions for both WF9 and ys1 corn, but unlike corn, Coker 227 and TAM 0-312 oats were both able to obtain Fe from the Fe2+ (BPDS)3 complex and were less chlorotic as a result. In short-term (4-hour) in vivo measurements, iron-stressed WF9 (Fe-efficient) corn reduced more Fe3+ to Fe2+ than similarly stressed ys1 corn, Coker 227 oat or TAM 0-312 oat. Thus, at the same time that Fe-efficient WF9 corn reduces more Fe than the other cultivars, it is also unable to compete with BPDS for that Fe in the nutrient solution. These differences coupled with the observation that only Coker 227 oat produced measureable iron solubilizing substances (phytosiderophores) suggest that these two species differ in their mechanisms for obtaining Fe during Fe-deficiency stress.  相似文献   

5.
The uptake of 59Fe from FeCl3, ferric (Fe3+) citrate (FeCitr) and Fe3+-EDTA (FeEDTA) was studied in leaf mesophyll of Vigna unguiculata (L.) Walp. Uptake rates decreased in the order FeCl3>FeCitrFeEDTA, and uptake depended on an obligatory reduction step of Fe3+ to Fe2+, after which the ion could be taken up independently of the chelator, citrate. Uptake was strongly increased by photosynthetically active light (>630 nm), and kinetic analysis revealed saturation kinetics with a K m (FeCitr) of 80–110 M. In the presence of an external Fe2+ scavenger, bathophenanthroline disulfonate, the mesophyll also reduced external FeCitr with a K m of approx. 50–60 M. The reduction rates for FeCitr were five-to eightfold higher than necessary for uptake. Purified plasma membranes from leaves revealed an NADH-dependent FeCitr- and FeEDTA-reductase activity, which had a pH optimum of 6.5–6.8 and a K m of approx. 20 M for NADH. Under anaerobic conditions, a K m of 130–170 M for ferric chelates was obtained, while in the presence of oxygen a K m (FeCitr) of approx. 100 M was found. It is concluded that the leaf plasma membrane provides a ferric-chelate-reductase activity, which plays a crucial role in iron uptake of leaf cells. Under in-vivo conditions, however, reactive oxygen species or strong (blue) light may also contribute to the obligatory reduction of Fe3+ prior to uptake.Abbreviations BPDS bathophenanthroline disulfonate - DCMU 3-(3,4 dichlorophenyl)-1,1-dimethyl urea - FCR ferricchelate reductase - FeCitr Fe3+-citrate - FeEDTA Fe3+-EDTA - PM plasma membrane This work was supported by the SCIENCE program of the European Community (contract no. SC1000344; P.R.M.). We wish to thank P. Siersma and C. Winter for their cooperation at the Central Isotope Laboratory of the Biological Centre of the University of Groningen.  相似文献   

6.
Summary Geotrichum candidum (isolate 1–9) pathogenic on citrus fruits, appears to lack siderophore production. Iron uptake byG. candidum is mediated by two distinct iron-regulated, energy-and temperature-dependent transport systems that require sulfhydryl groups. One system exhibits specificity for either ferric or ferrous iron, whereas the other exhibits specificity for ferrioxamine-B-mediated iron uptake and presumably other hydroxamate siderophores. Radioactive iron uptake from59FeCl3 showed an optimum at pH 6 and 35° C, and Michaelis-Menten kinetics (apparentK m = 3 m,V max = 0.054 nmol · mg–1 · min–1). The maximal rate of Fe2+ uptake was higher than Fe3+ (V max = 0.25 nmol · mg–1 · min–1) but theK m was identical. Reduction of ferric to ferrous iron prior to transport could not be detected. The ferrioxamine B system exhibits an optimum at pH 6 and 40° C and saturation kinetics (K m = 2 M,V max = 0.22 nmol · mg–1 · min–1). The two systems were distinguished as two separate entities by negative reciprocal competition, and on the basis of differential response to temperature and phenazine methosulfate. Mössbauer studies revealed that cells fed with either57FeCl3 or57FeCl2 accumulated unknown ferric and ferrous binding metabolites.  相似文献   

7.
M. Skerrett  S. D. Tyerman 《Planta》1994,192(3):295-305
An anion channel that only allows outward current flow (anion influx) has been identified in protoplasts derived from wheat (Triticum aestivum L., Triticum turgidum L.) roots. The anion outward rectifier (anion OR) measured by patch-clamp of whole cells activated very quickly, usually reaching a steady-state level in less than 100 ms and was easily distinguished from the cation outward rectifier (cation OR) which activated more slowly during membrane depolarisation. The anion OR is permeable to NO 3 and Cl, moderately permeable to I, and relatively impermeable to H2PO4/ and ClO4/. An anomalous mole-fraction effect between ClO4/ and Cl was observed on the outward current, indicating that the channel is a multi-ion pore. The anion OR is gated by both voltage and external anion concentration such that it activates near to the equilibrium potential for the permeant anion. It activated at more negative membrane potentials when NO 3 was substituted for Cl in the external medium, indicating that the channel may function to allow NO 3 influx under luxuriant external NO 3 concentrations. For most experiments, K+ and Cl were the main cation and anion in solution, and under these conditions it appeared likely that the anion OR functioned in membrane-potential regulation by facilitating a Cl influx at membrane potentials more positive than the chloride reversal potential (ECl). If ECl was more negative than the K+ reversal potential (EK) then the anion OR dominated but both the anion and cation ORs occurred together when the membrane potential difference (Vm) was positive of both ECl and EK. The cation OR was inhibited by increasing external Cl concentrations, but the anion OR was not affected by external K+ or Na+ concentration. The anion-transport inhibitors, zinc and phenylglyoxal were ineffective in blocking the anion OR. 4,4-Di-isothiocyanostilbene-2, 2-disulfonic acid (DIDS) irreversibly blocked about 34% of the current when applied extracellularly at a concentration of 25 M, and about 69% at a concentration of 200 M. However, DIDS (200 M) also occasionally acted as an irreversible blocker of the cation OR. Perchlorate blocked irreversibly 75% of the current at an external concentration of 10 mM and did not block the cation OR. Whole-cell currents also indicated that the anion OR was insensitive to external pH (pH=5–7) and calcium concentration ([Ca2+]=0.1–10 mM). Increasing intracellular calcium concentration significantly increased the occurrence of the fast outward current in whole cells (P < 0.005, X2 test). With approximately 10 nM calcium inside the cell the anion outward current was observed in 64% (n = 45) of cells and with 50 nM calcium inside the cell the anion current was observed in 88% (n = 69) of cells. Single-anion OR channels observed in outside-out patches had a conductance in 300 mM KCl (external) of about 4 pS. When voltage pulses were applied to outside-out patches the average currents were similar to those observed in whole cells. The significance of the anion OR as a likely route for Cl uptake in high salinities is discussed.Abbreviations Bath solution bathing the extracellular face of the membrane - DIDS (4,4-diisothiocyanostilbene-2,2-disulfonic acid) - Ex reversal potential for ion x - OR outward rectifier - Pip solution inside the pipette - TEACl (tetraethyl-ammonium chloride) - Vm membrane potential difference We thank the Australian Research Council for financial support, G.P. Findlay and A. Garrill for helpful discussions, and K. Morris and D. Mackenzie for expert technical assistance. M.S. was supported by an Australian Postgraduate Research Award.  相似文献   

8.
Summary Positive and negative ion FAB mass spectrometry were found to be useful for the structural analysis of phosphorylated peptides containing multiple O-phosphoseryl residues. The positive ion FAB mass spectra obtained for Ac-Ser(P)-Ser(P)-NHMe and Ac-Ser(P)-Ser(P)-Ser(P)-NHMe showed that -eliminative loss of H3PO4 from the Ser(P)-residue was a major event in the fragmentation of the two phosphopeptides and that successive losses of H3PO4 from the [M+H]+ ion occurred when the Ser(P)-cluster was located at the N-terminus. In contrast, the FAB mass spectrum of Ac-Glu-Ser(P)-Leu-Ser(P)-Ser(P)-Ser(P)-Glu-Glu-NHMe showed only a single loss of H3PO4 from the [M+H]+ ion, with further losses of H3PO4 from internal Ser(P)-residues only occurring when fragmentation of the parent phosphopeptide generated daughter fragments that contained (part of) an N-terminal Ser(P)-residue. Negative ion FAB mass spectrometry also proved useful for the structural analysis of the three Ser(P)-peptides and showed high-intensity [M-H]- ions along with minor [M-H-80]- fragment ions.Abbreviations Ac acetyl - Ala dehydroalanyl - FAB-MS fast atom bombardment mass spectrometry - LSIMS liquid secondary ion mass spectrometry - NHMe N-methylamide - Ser(P) O-phosphoseryl - Thr(P) O-phosphothreonyl  相似文献   

9.
This study demonstrates a pH-dependent inhibition of Mg2+- and Ca2+- ATPase activities ofNostoc linckia andChlorella vulgaris exposed to AlCl3, AlF3, NaF and AlCl3 + NaF together. AlF3 and the combination of AlCl3 + NaF were more inhibitory to both the enzymes as compared with AlCl3 and NaF. Toxicity of the test compounds increased with increasing acidity. Interaction of AlCl3 + NaF was additive onN. linckia andC. vulgaris, respectively, at pH 7.5 and 6.8, and synergistic at pH 6.0 and 4.5. In the presence of 60 and 100 m PO4 3- an increased NaF concentration (in the AlCl3 + NaF combination) was required to produce the same degree of inhibition in ATP synthesis and ATPase activity. Toxicity of fluoroaluminate was reduced in the presence of EDTA and citrate. Except for beryllium to some extent, combinations of cadmium, cobalt, iron, manganese, tin and zinc with fluoride were not as effective as aluminium in inhibiting the ATPase activity. The presence of a 100 kDa protein band in SDS-PAGE of both control as well as AlCl3 + NaF-treated samples suggested that AlF4 inhibits the ATPase activity by acting as a functional barrier without affecting the structure of the enzyme.  相似文献   

10.
Iron has a central role in bioleaching and biooxidation processes. Fe2+ produced in the dissolution of sulfidic minerals is re-oxidized to Fe3+ mostly by biological action in acid bioleaching processes. To control the concentration of iron in solution, it is important to precipitate the excess as part of the process circuit. In this study, a bioprocess was developed based on a fluidized-bed reactor (FBR) for Fe2+ oxidation coupled with a gravity settler for precipitative removal of ferric iron. Biological iron oxidation and partial removal of iron by precipitation from a barren heap leaching solution was optimized in relation to the performance and retention time (τFBR) of the FBR. The biofilm in the FBR was dominated by Leptospirillum ferriphilum and “Ferromicrobium acidiphilum.” The FBR was operated at pH 2.0 ± 0.2 and at 37 °C. The feed was a barren leach solution following metal recovery, with all iron in the ferrous form. 98–99% of the Fe2+ in the barren heap leaching solution was oxidized in the FBR at loading rates below 10 g Fe2+/L h (τFBR of 1 h). The optimal performance with the oxidation rate of 8.2 g Fe2+/L h was achieved at τFBR of 1 h. Below the τFBR of 1 h the oxygen mass transfer from air to liquid limited the iron oxidation rate. The precipitation of ferric iron ranged from 5% to 40%. The concurrent Fe2+ oxidation and partial precipitative iron removal was maximized at τFBR of 1.5 h, with Fe2+ oxidation rate of 5.1 g Fe2+/L h and Fe3+ precipitation rate of 25 mg Fe3+/L h, which corresponded to 37% iron removal. The precipitates had good settling properties as indicated by the sludge volume indices of 3–15 mL/g but this step needs additional characterization of the properties of the solids and optimization to maximize the precipitation and to manage sludge disposal.  相似文献   

11.
Competitive binding of Fe3+, Cr3+, and Ni2+ to transferrin (Tf) was investigated at various physiological iron to Tf concentration ratios. Loading percentages for these metal ions are based on a two M n+ to one Tf (i.e., 100% loading) stoichiometry and were determined using a particle beam/hollow cathode–optical emission spectroscopy (PB/HC-OES) method. Serum iron concentrations typically found in normal, iron-deficient, iron-deficient from chronic disease, iron-deficient from inflammation, and iron-overload conditions were used to determine the effects of iron concentration on iron loading into Tf. The PB/HC-OES method allows the monitoring of metal ions in competition with Fe3+ for Tf binding. Iron-overload concentrations impeded the ability of chromium (15.0 μM) or nickel (10.3 μM) to load completely into Tf. Low Fe3+ uptake by Tf under iron-deficient or chronic disease iron concentrations limited Ni2+ loading into Tf. Competitive binding kinetic studies were performed with Fe3+, Cr3+, and Ni2+ to determine percentages of metal ion uptake into Tf as a function of time. The initial rates of Fe3+ loading increased in the presence of nickel or chromium, with maximal Fe3+ loading into Tf in all cases reaching approximately 24%. Addition of Cr3+ to 50% preloaded Fe3+–Tf showed that excess chromium (15.0 μM) displaced roughly 13% of Fe3+ from Tf, resulting in 7.6 ± 1.3% Cr3+ loading of Tf. The PB/HC-OES method provides the ability to monitor multiple metal ions competing for Tf binding and will help to understand metal competition for Tf binding.  相似文献   

12.
Summary The kinetics of the separate processes of Fe2(III)-transferrin binding to the transferrin receptor, transferrin-receptor internalization, iron dissociation from transferrin, iron passage through the membrane, and iron mobilization into the cytoplasm were studied by pulse-chase experiments using rabbit reticulocytes and59Fe,125I-labeled rabbit transferrin. The binding of59Fe-transferrin to transferrin receptors was rapid with an apparent rate constant of 2×105 m –1 sec–1. The rate of internalization of59Fe-transferrin was directly measured at 520±100 molecules of Fe2(III)-transferrin internalized/sec/cell with 250±43 sec needed to internalize the entire complement of reticulocyte transferrin receptors. Subsequent to Fe2(III)-transferrin internalization the flux of59Fe was followed through three compartments: internalized transferrin, membrane, and cytosol.A process preceding iron dissociation from transferrin and a reaction involving membrane-associated iron required 17±2 sec and 34±5 sec, respectively. Apparent rate constants of 0.0075±0.002 sec–1 and 0.0343±0.0118 sec–1 were obtained for iron dissociation from transferrin and iron mobilization into the cytosol, respectively. Iron dissociation from transferrin is the rate-limiting step. An apparent rate constant of 0.0112±0.0025 sec–1 was obtained for processes involving iron transport through the membrane although at least two reactions are likely to be involved. Based on mechanistic considerations, iron transport through the membrane may be attributed to an iron reduction step followed by a translocation step. These data indicate that the uptake of iron in reticulocytes is a sequential process, with steps after the internalization of Fe2(III)-transferrin that are distinct from the handling of transferrin.  相似文献   

13.
Freshly-added iron only slightly affected the growth of iron-sufficient cells of the green alga Scenedesmus incrassatulus Bohl, strain R-83, but induced accumulation of malondialdehyde (MDA) in cells and excretion of MDA in the medium. These effects were stronger in response to Fe2+ as compared to Fe3+, but Fe3+ induced the release of more iron-binding chelators from these cells than Fe2+. Fe3+ added either in dark or in light induced release of equal concentrations of iron-complexing agents, part of which formed strong chelates with iron in the medium. Exogenously added hydrogen peroxide inhibited iron-induced release of chelators but the effect was removed by addition of the hydroxyl radical scavenger dimethylsulfoxide (DMSO). Malondialdehyde also inhibited the release of chelators. Release of chelators was induced in the absence of iron salts by photoexcited chlorophyll (Chl). The Chl-induced release was efficiently inhibited by singlet oxygen scavengers such as dimethylfuran, -carotene, sodium azide and vitamin B6, and stimulated in D2O or DMSO. Exogenously added catalase inhibited the release more than added superoxide dismutase. The Fe3-induced release of chelators was also inhibited by scavengers of singlet oxygen, but was not affected by sodium azide and by ethanol. Hence both H2O2 and singlet oxygen were involved in induction of chelator release in the absence of iron in light. The induction of chelator release by iron in dark involved H2O2, but not singlet oxygen.  相似文献   

14.
Balnokin YV  Popova LG  Pagis LY  Andreev IM 《Planta》2004,219(2):332-337
Our previous investigations have established that Na+ translocation across the Tetraselmis viridis plasma membrane (PM) mediated by the primary ATP-driven Na+-pump, Na+-ATPase, is accompanied by H+ counter-transport [Y.V. Balnokin et al. (1999) FEBS Lett 462:402–406]. The hypothesis that the Na+-ATPase of T. viridis operates as an Na+/H+ exchanger is tested in the present work. The study of Na+ and H+ transport in PM vesicles isolated from T. viridis demonstrated that the membrane-permeant anion NO3 caused (i) an increase in ATP-driven Na+ uptake by the vesicles, (ii) an increase in (Na++ATP)-dependent vesicle lumen alkalization resulting from H+ efflux out of the vesicles and (iii) dissipation of electrical potential, , generated across the vesicle membrane by the Na+-ATPase. The (Na++ATP)-dependent lumen alkalization was not significantly affected by valinomycin, addition of which in the presence of K+ abolished at the vesicle membrane. The fact that the Na+-ATPase-mediated alkalization of the vesicle lumen is sustained in the absence of the transmembrane is consistent with a primary role of the Na+-ATPase in driving H+ outside the vesicles. The findings allowed us to conclude that the Na+-ATPase of T. viridis directly performs an exchange of Na+ for H+. Since the Na+-ATPase generates electric potential across the vesicle membrane, the transport stoichiometry is mNa+/nH+, where m>n.Abbreviations BTP Bis-Tris-Propane, 1,3-bis[tris(hydroxymethyl)methylamino]-propane - CCCP Carbonyl cyanide m-chlorophenylhydrazone - DTT Dithiothreitol - NCDC 2-Nitro-4-carboxyphenyl N,N-diphenylcarbamate - PMSF Phenylmethylsulfonyl fluoride - PM Plasma membrane  相似文献   

15.
Serum transferrin (sTf) is a bi-lobal protein. Each lobe of sTf binds one Fe3+ ion in the presence of a synergistic anion. Physiologically, carbonate is the main synergistic anion but other anions such as oxalate, malonate, glycolate, maleate, glycine, etc. can substitute for carbonate in vitro. The present work provides the possible pathways by which the substitution of carbonate with oxalate affects the structural, kinetic, thermodynamic, and functional properties of blood plasma sTf. Analysis of equilibrium experiments measuring iron release and structural unfolding of carbonate and oxalate bound diferric-sTf (Fe2sTf) as a function of pH, urea concentration, and temperature reveal that the structural and iron-centers stability of Fe2sTf increase by substitution of carbonate with oxalate. Analysis of isothermal titration calorimetry (ITC) scans showed that the affinity of Fe3+ with apo-sTf is enhanced by substituting carbonate with oxalate. Analysis of kinetic and thermodynamic parameters measured for the iron release from the carbonate and oxalate bound monoferric-N-lobe of sTf (FeNsTf) and Fe2sTf at pH 7.4 and pH 5.6 reveals that the substitution of carbonate with oxalate inhibits/retards the iron release via increasing the enthalpic barriers.  相似文献   

16.
Two extracellular tannin acyl hydrolases (TAH I and TAH II) produced by an Antarctic filamentous fungus Verticillium sp. P9 were purified to homogeneity (7.9- and 10.5-fold with a yield of 1.6 and 0.9%, respectively) and characterized. TAH I and TAH II are multimeric (each consisting of approximately 40 and 46 kDa sub-units) glycoproteins containing 11 and 26% carbohydrates, respectively, and their molecular mass is approximately 155 kDa. TAH I and TAH II are optimally active at pH of 5.5 and 25 and 20°C, respectively. Both the enzymes were activated by Mg2+and Br ions and 0.5–2.0 M urea and inhibited by other metal ions (Zn2+, Cu2+, K+, Cd2+, Ag+, Fe3+, Mn2+, Co2+, Hg2+, Pb2+ and Sn2+), anions, Tween 20, Tween 60, Tween 80, Triton X-100, sodium dodecyl sulphate, β-mercaptoethanol, α-glutathione and 4-chloromercuribenzoate. Both tannases more efficiently hydrolyzed tannic acid than methyl gallate. E a of these reactions and temperature dependence (at 0–30°C) of k cat, k cat/K m, ΔG*, ΔH* and ΔS* for both the enzymes and substrates were determined. The k cat and k cat/K m values (for both the substrates) were considerably higher for the combined preparation of TAH I and TAH II.  相似文献   

17.
To model the mononuclear FeIII-OOH species identified in the catalytic cycle of the anticancer drug bleomycin, the iron chemistry of the pentadentate ligand N-[bis(2-pyridylmethyl)aminoethyl]pyridine-2-carboxamide (H-PaPy3) has been investigated. The complex [FeIII(PaPy3)OCH3](ClO4) was reacted with H2O2 to form a red species (max=480 nm, =1800 M–1 cm–1) with an S=1/2 EPR signal at g=2.25, 2.17, and 1.95. This species has been identified by electrospray ionization mass spectrometry as [FeIII(PaPy3)OOH](ClO4) and further characterized by resonance Raman and EXAFS analysis. The decomposition of this intermediate leads to the modification of the ligand, as revealed by 1H NMR. One hydrogen atom is substituted by a solvent-derived methoxy group. The substitution at this site is a result of the two-electron oxidation of the ligand following the heterolytic cleavage of the O–O bond of the FeIII-OOH species. This is a plausible mechanism to rationalize related ligand modifications that have been proposed in the decay of activated bleomycin.Abbreviations ABLM activated bleomycin - BLM bleomycin - ESI-MS electrospray ionization mass spectrometry - EXAFS extended X-ray absorption fine structure - H-PaPy3 N-[bis(2-pyridylmethyl)aminoethyl]pyridine-2-carboxamide  相似文献   

18.
Cohen CK  Garvin DF  Kochian LV 《Planta》2004,218(5):784-792
Fe uptake in dicotyledonous plants is mediated by a root plasma membrane-bound ferric reductase that reduces extracellular Fe(III)-chelates, releasing Fe2+ ions, which are then absorbed via a metal ion transporter. We previously showed that Fe deficiency induces an increased capacity to absorb Fe and other micronutrient and heavy metals such as Zn2+ and Cd2+ into pea (Pisum sativum L.) roots [Cohen et al. (1998) Plant Physiol 116:1063–1072). To investigate the molecular basis for this phenomenon, an Fe-regulated transporter that is a homologue of the Arabidopsis IRT1 micronutrient transporter was isolated from pea seedlings. This cDNA clone, designated RIT1 for root iron transporter, encodes a 348 amino acid polypeptide with eight putative membrane-spanning domains that is induced under Fe deficiency and can functionally complement yeast mutants defective in high- and low-affinity Fe transport. Chelate buffer techniques were used to control Fe2+ in the uptake solution at nanomolar activities representative of those found in the rhizosphere, and radiotracer methodologies were employed to show that RIT1 is a very high-affinity 59Fe2+ uptake system (K m =54–93 nM). Additionally, radiotracer (65Zn, 109Cd) flux techniques were used to show that RIT can also mediate a lower affinity Zn and Cd influx (K m of 4 and 100 M, for Zn2+ and Cd2+, respectively). These findings suggest that, in typical agricultural soils, RIT1 functions primarily as a high-affinity Fe2+ transporter that mediates root Fe acquisition. This is consistent with recent findings with Arabidopsis IRT1 knockout mutants that strongly suggest that this transporter plays a key role in root Fe uptake and nutrition. However, the ability of RIT1 to facilitate Zn and Cd uptake when these metals are present at elevated concentrations suggests that RIT1 may be one pathway for the entry of toxic metals into the food chain. Furthermore, the finding that plant Fe deficiency status may promote heavy metal uptake via increased expression of this transporter could have implications both for human nutrition and also for phytoremediation, the use of terrestrial plants to sequester toxic metals from contaminated soil.  相似文献   

19.
Saliva was collected from the mandibular glands of anaesthetized common wombats (Vombatus ursinus) to ascertain maximal flow rates, salivary compostion and possible adaptations, particularly PO4 3- secretion, to assist digestion. After temporary catheterization of the main duct through its oral opening, salivary secretion was evoked at flow rates ranging from 0.02±0.002 (±SEM) ml·min-1 (0.7±0.07 l·min-1·kg body weight-1) to 0.4±0.05 ml·min-1(14±1.9 l·min-1·kg body weight-1) by ipsilateral intracarotid infusion of acetylcholine. The [Na+] (15±5.1 to 58±8.6 mmol·l-1) and [HCO3 -] (35±1.9 to 60±1.9 mmol·l-1) were positively correlated with salivary flow rate. The [K+] (58±5.2 to 30±2.4 mmol·l-1), [Ca2+] (10.4±1.67 to 4.1±0.44 mmol·l-1), [Mg2+] (0.94±0.137 to 0.17±0.032 mmol·l-1), [Cl-] (71±9.2 to 45±6.0 mmol·l-1), [urea] (9.3±0.79 to 5.1±0.54 mmol·l-1), H+ activity (29±1.6 to 17±1.6 nEq·l-1) and amylase activity (251±57.4 to 92±23.3 kat·l-1) were negatively correlated with flow. Both concentration and osmolality fell with increasing flow at the lower end of the flow range but osmolality always increased again by maximal flow whereas the relation between protein and flow was not consistent at the higher levels of flow and stimulation. Salivary [PO4 3+] was not correlated with flow and at 3–14% of the plasma concentration was extremely low. Thus, in contrast to its nearest relative, the koala (Phascolarctos cinereus), the wombat secretes little PO4 3+ presumably because it does not need high levels of PO4 3+ in its saliva to facilitate microbial digestion of plant fibre.Abbreviations bw body weight - ww wet weight  相似文献   

20.
The crude extracellular cellulase from Clostridium thermocellum was oxidatively inactivated by air and inhibited by sulfhydryl reagents. Activity-loss was prevented and reversed by the addition of a high concentration (10 mM) dithiothreitol (DDT) at zero time and up to 24 h respectively. In the presence of a low concentration (0.4 mM) of DTT, the enzyme was more rapidly inactivated than in air alone. This was probably due to autoxidation of the low DTT concentration to H2O2 as shown by its prevention by a high DTT concentration, exclusion of air, or catalase; and by the oxidative inactivation of the enzyme by H2O2. The inactivation by H2O2 could be prevented by a high concentration of DTT but not by air exclusion. EDTA protected the enzyme from inactivation in air by a low concentration of DTT or by H2O2. This is presumably due to the role of metals in oxidation of SH groups. Furthermore, copper (5 M) also caused inactivation and this was prevented by the presence of a high DTT concentration. Even in the protective atmosphere of a high DTT concentration, cellulase was inactivated by certain apolar chelating agents such as o-phenanthroline and -1-dipyridyl, such inactivation being preventable by the prior incubation of the chelator with a mixture of Fe2+ and Fe3+. These data suggest that the clostridial cellulase, unlike the enzyme from aerobic fungi, contains essential sulfhydryl groups and is stimulated by iron. The endo--glucanase component of the cellulase complex was not susceptible to oxidative inactivation.Abbreviations DTT dithiothreitol - CMC carboxymethylcellulose - DTNB 5,5-dithiobis-(2-nitrobenzoic acid) - NEM N-ethylmaleimide - p-CMB p-chloromercuribenzoic acid  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