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.
2,3-Dihydroxybiphenyl-1,2-dioxygenase plays an important role in the degradation of polychlorinated biphenyls. The gene (BsbphCI) encoding a 2,3-DHBP dioxygenase from Bacillus sp. JF8 is 960 bp. We synthesized a 960 bp BsbphCI gene encoding a 2,3-DHBP dioxygenase derived from Bacillus sp. JF8 and expressed it in Escherichiacoli. The recombinant protein was about 36 kDa, confirmed by SDS-PAGE. The concentration of the purified protein was about 1.8 mg/mL. With 2,3-DHBP as a substrate, the optimal temperature for enzyme activity at pH 8.5 was 50 °C. The optimal pH for the 2,3-DHBP dioxygenase was 8.5. The enzyme retained 33% activity after heating at 60 °C for 60 min. We found that Cu(2+), K(+), Zn(2+), Mg(2+), Ni(2+), Co(2+), and Cd(2+) activated the enzyme. However, Ca(2+), Fe(2+), Li(+), and Cr(3+) inhibited it. Enzyme activity was reduced by exposure to H(2)O(2), SDS, and KI. The results of HPLC indicated that the transgenic E. coli strain with the BsbphCI gene degraded 2,3-DHBP more quickly than the wild type strain.  相似文献   

2.
The exposure of various human populations to Cd(2+) is of increasing health concern. After its gastrointestinal absorption into the bloodstream, Cd(2+) binds to α(2)-macroglobulin and serum albumin. Although animal studies have demonstrated that meso-2,3-dimercaptosuccinic acid (DMSA) and diethylenetriamine pentaacetic acid (DTPA) can effectively mobilize Cd(2+) to urine and decrease the Cd concentrations of the kidneys, the liver and the brain, not much is known about the abstraction of Cd(2+) from blood plasma proteins. We prepared a stock of Cd(2+) spiked rabbit plasma (2.0 μg of Cd(2+)/mL) and analyzed aliquots by size exclusion chromatography coupled on-line to an inductively coupled plasma atomic emission spectrometer (SEC-ICP-AES) while simultaneously monitoring the emission lines of Ca, Cd, Cu, Fe, and Zn. After the addition of 0.33 mM, 0.66 mM or 0.99 mM of DMSA, DTPA, 2,3-dimercapto-1-propanesulfonic acid (DMPS) or N-acetyl-l-cysteine (NAC) to plasma aliquots, the obtained mixtures were analyzed by SEC-ICP-AES after 5 min and 30 min. None of the investigated compounds adversely affected the plasma distribution of Fe at all investigated doses. At 0.33 mM, DTPA was most effective at mobilizing plasma protein bound Cd(2+) to a ~5 kDa Cd-species (100% removal), followed by DMPS (94%), DMSA (83%) and NAC (3%). All investigated compounds also mobilized Zn(2+) from plasma proteins to ~5 kDa Zn-species (DTPA: 80% removal; DMPS: 63%; DMSA: 29% and NAC: 3%). The addition of DTPA resulted in the dose-dependent elution of a [Ca-DTPA](3-) complex. Based on these results, 0.33 mM DMSA represents the best compromise that can be achieved between maximizing the abstraction of Cd(2+) from plasma proteins (83%), while minimizing the mobilization of Zn(2+) from plasma proteins (29%), and avoiding the complexation of Ca(2+).  相似文献   

3.
Inhibition by Zn(2+) of iron uptake by apoferritin at very low substrate concentrations is shown to be competitive. It is proposed that Zn(2+) competes with Fe(2+) for sites on the protein at which the oxidation of Fe(2+) is catalysed. Interpretation of titration data suggests there are two independent classes of binding site for Zn(2+) and several other cations. Sites in one such class are probably on the external surface of the apoferritin molecule. The catalytic binding sites are presumed to be internal and may involve histidine or possibly cysteine as ligands.  相似文献   

4.
5.
Abnormally high concentrations of Zn(2+), Cu(2+), and Fe(3+) are present along with amyloid-β (Aβ) in the senile plaques in Alzheimer disease, where Al(3+) is also detected. Aβ aggregation is the key pathogenic event in Alzheimer disease, where Aβ oligomers are the major culprits. The fundamental mechanism of these metal ions on Aβ remains elusive. Here, we employ 4,4'-Bis(1-anilinonaphthalene 8-sulfonate) and tyrosine fluorescence, CD, stopped flow fluorescence, guanidine hydrochloride denaturation, and photo-induced cross-linking to elucidate the effect of Zn(2+), Cu(2+), Fe(3+), and Al(3+) on Aβ at the early stage of the aggregation. Furthermore, thioflavin T assay, dot blotting, and transmission electron microscopy are utilized to examine Aβ aggregation. Our results show that Al(3+) and Zn(2+), but not Cu(2+) and Fe(3+), induce larger hydrophobic exposures of Aβ conformation, resulting in its significant destabilization at the early stage. The metal ion binding induces Aβ conformational changes with micromolar binding affinities and millisecond binding kinetics. Cu(2+) and Zn(2+) induce similar assembly of transiently appearing Aβ oligomers at the early state. During the aggregation, we found that Zn(2+) exclusively promotes the annular protofibril formation without undergoing a nucleation process, whereas Cu(2+) and Fe(3+) inhibit fibril formation by prolonging the nucleation phases. Al(3+) also inhibits fibril formation; however, the annular oligomers co-exist in the aggregation pathway. In conclusion, Zn(2+), Cu(2+), Fe(3+), and Al(3+) adopt distinct folding and aggregation mechanisms to affect Aβ, where Aβ destabilization promotes annular protofibril formation. Our study facilitates the understanding of annular Aβ oligomer formation upon metal ion binding.  相似文献   

6.
The high-resolution structure of the non-haem ferritin from Escherichia coli (EcFtnA) is presented together with those of its Fe(3+) and Zn(2+) derivatives, this being the first high-resolution X-ray analysis of the iron centres in any ferritin.The binding of both metals is accompanied by small changes in the amino acid ligand positions. Mean Fe(A)(3+)-Fe(B)(3+) and Zn(A)(2+)-Zn(B)(2+) distances are 3.24 A and 3.43 A, respectively. In both derivatives, metal ions at sites A and B are bridged by a glutamate side-chain (Glu50) in a syn-syn conformation. The Fe(3+) derivative alone shows a third metal site (Fe( C)( 3+)) joined to Fe(B)(3+) by a long anti-anti bidentate bridge through Glu130 (mean Fe(B)(3+)-Fe(C)(3+) distance 5.79 A). The third metal site is unique to the non-haem bacterial ferritins.The dinuclear site lies at the inner end of a hydrophobic channel connecting it to the outside surface of the protein shell, which may provide access for dioxygen and possibly for metal ions shielded by water. Models representing the possible binding mode of dioxygen to the dinuclear Fe(3+) pair suggest that a gauche micro-1,2 mode may be preferred stereochemically.Like those of other ferritins, the 24 subunits of EcFtnA are folded as four-helix bundles that assemble into hollow shells and both metals bind at dinuclear centres in the middle of the bundles. The structural similarity of EcFtnA to the human H chain ferritin (HuHF) is remarkable (r.m.s. deviation of main-chain atoms 0.66 A) given the low amino acid sequence identity (22 %). Many of the conserved residues are clustered at the dinuclear centre but there is very little conservation of residues making inter-subunit interactions.  相似文献   

7.
Thermal denaturation of the mesophilic rubredoxin from Clostridium pasteurianum occurs through a number of temperature-dependent steps, the last and irreversible one being release of iron from the [Fe(2+)(SCys)(4)] site. We show here that thermally induced [Fe(2+)(SCys)(4)] site destruction is largely determined by the local environment, and not directly connected to thermostability of the native polypeptide fold of rubredoxin. Hydrophobic residues on the protein surface, V8 and L41, that shield the [Fe(SCys)(4)] site from solvent and form N-H(.)S hydrogen bonds to the metal-coordinating sulfurs, were mutated to residues with both uncharged and charged side chains. On these mutated rubredoxins the temperature dependence was measured for: (1) global unfolding of the protein by NMR, (2) loss of Fe(2+)at various ionic strengths and pH values, (3) the rates of non-denaturing displacement of Fe(2+) by Cd(2+) or Zn(2+). For reversible temperature-dependent changes in the global protein folding that occur prior to loss of iron, no thermostability differences were found among the wild-type, V8A, V8D, L41R, and L41D rubredoxins. However, for irreversible loss of iron from the [Fe(2+)(SCys)(4)] site, relative to the wild-type protein, L41R was more thermostable, V8A was somewhat less thermostable, and the acidic mutants L41D, V8D and [V8D, L41D] showed dramatically lowered thermostability. Lower pH facilitated - both kinetically and thermodynamically - thermally induced iron release, likely through protonation of ligand cysteines' thiols. For all of the rubredoxins a direct correlation was found between the midpoint temperature for thermally induced Fe(2+) loss and the rate of non-denaturing Fe(2+) displacement by Cd(2+) or Zn(2+) at room temperature. A mechanism is proposed involving transient movement of residue-8 and -41 side chains, allowing, and, in the case of negatively charged side chains, also facilitating, attack of a ligand cysteine by the incoming positively charged species (H(+), Cd(2+), or Zn(2+)). Thus, localized charge density and solvent accessibility modulate the stability of Fe(2+) ligation in rubredoxin. However, the reduced [Fe(SCys)(4)] site does not control the thermostability of the native polypeptide fold of rubredoxin.  相似文献   

8.
Recombinant Escherichia coli cytosine deaminase is purified as a mixture of Zn(2+) and Fe(2+) forms of the enzyme. Fe(2+) is removed readily by o-phenanthroline to yield apoenzyme (apoCDase) that contains <0.2 mol of Zn(2+)per mol of subunit. ApoCDase was efficiently reconstituted to Zn(2+)CDase by treatment with ZnCl(2). The interaction of cytosine with apoCDase and Zn(2+)CDase was investigated at pH 7.5 and 25 degrees C by monitoring changes in intrinsic protein fluorescence. The values for the kinetic data K(1), k(2), and k(3) for Zn(2+)CDase were 0.25 mM, 80 s(-1), and 38 s(-1), respectively. The value for k(-2) was statistically indistinguishable from zero. The analogous values for K(1), k(2), and k(-2), (k(3)=0) for apoCDase were 0.157 mM, 186 s(-1) and approximately 0.8 s(-1), respectively. The overall dissociation constant of apoCDase for cytosine was 0.00069 mM, whereas the K(m) of Zn(2+)CDase for cytosine was 0.20 mM. The pre-steady state phase of the reaction was associated with an absorbance increase at 280 nm that was attributed to solvent perturbation of the spectrum of cytosine or enzyme. Formation of the Fe(2+)CDase-cytosine complex was too rapid to monitor by these techniques.  相似文献   

9.
1. Incorporation of [(3)H]thymidine into DNA was inhibited by EDTA and diethylenetriamine-NNN'N'N'-penta-acetate but not by nitrilotriacetate even when Ca(2+) was present at more than twice the concentration of the chelators. 2. The inhibition caused by EDTA was most effectively reversed by Zn(2+) but also to a lesser extent by Cd(2+). Very little if any activation of the EDTA-inhibited system was obtained with Co(2+), Cu(2+), Fe(3+), Mn(2+) or Ni(2+) added alone. 3. Fe(3+) added to the Zn(2+)-activated lymphocytes in the presence of EDTA markedly increased thymidine incorporation. Addition of Cd(2+) prevented the inhibition of incorporation which occurred at high Zn(2+) concentrations. 4. If EDTA was added more than 15h after phytohaemagglutinin, the inhibition of incorporation was less than that obtained by its addition at zero time. If Zn(2+) was added later than 12h after EDTA and phytohaemagglutinin, the inhibition of incorporation by EDTA was not fully reversed. A study of the time-course of the effects of delayed additions of EDTA and Zn(2+) suggested that, on average, the cells required Zn(2+) between 20 and 30h after phytohaemagglutinin addition to allow the full rate of thymidine incorporation at 37h. 5. The increase in the rate of protein synthesis caused by phytohaemagglutinin was not inhibited by EDTA until about 8h. The further increase after this was totally inhibited by EDTA but this inhibition was fully reversible with Zn(2+). The rate of protein synthesis in EDTA-inhibited cultures at 40h was the same as that at 10h. 6. There was a close similarity between the effects of EDTA on lymphocyte stimulation and those reported by Kay et al. (1969) with low doses of actinomycin D.  相似文献   

10.
Bacterioferritins are members of a class of spherical shell-like iron storage proteins that catalyze the oxidation and hydrolysis of iron at specific sites inside the protein shell, resulting in formation of a mineral core of hydrated ferric oxide within the protein cavity. Electrode oximetry/pH stat was used to study iron oxidation and hydrolysis chemistry in E. coli bacterioferritin. Consistent with previous UV-visible absorbance measurements, three distinct kinetic phases were detected, and the stoichiometric equations corresponding to each have been determined. The rapid phase 1 reaction corresponds to pairwise binding of 2 Fe(2+) ions at a dinuclear site, called the ferroxidase site, located within each of the 24 subunits, viz., 2Fe(2+) + P(Z) --> [Fe(2)-P](Z) + 4H(+), where P(Z) is the apoprotein of net charge Z and [Fe(2)-P](Z) represents a diferrous ferroxidase complex. The slower phase 2 reaction corresponds to the oxidation of this complex by molecular oxygen according to the net equation: [Fe(2)-P](Z) + (1)/(2)O(2) --> [Fe(2)O-P](Z) where [Fe(2)O-P](Z) represents an oxidized diferric ferroxidase complex, probably a mu-oxo-bridged species as suggested by UV-visible and EPR spectrometric titration data. The third phase corresponds to mineral core formation according to the net reaction: 4Fe(2+) + O(2) + 6H(2)O --> 4FeO(OH)((core)) + 8H(+). Iron oxidation is inhibited by the presence of Zn(2+) ions. The patterns of phase 2 and phase 3 inhibition are different, though inhibition of both phases is complete at 48 Zn(2+)per 24mer, i.e., 2 Zn(2+) per ferroxidase center.  相似文献   

11.
Five conserved histidine residues are found in the human endothelial nitric-oxide synthase (NOS) heme domain: His-420, His-421, and His-461 are close to the heme, whereas His-146 and His-214 are some distance away. To investigate whether the histidines form a non-heme iron-binding site, we have expressed the H146A, H214A, H420A, H421A, and H461A mutants. The H420A mutant could not be isolated, and the H146A and H421A mutants were inactive. The H214A mutant resembled the wild-type enzyme in all respects. The H461A mutant had a low-spin heme, but high concentrations of L-Arg and tetrahydrobiopterin led to partial recovery of activity. Laser atomic emission showed that the only significant metal in NOS other than calcium and iron is zinc. The activities of the NOS isoforms were not increased by incubation with Fe(2+), but were inhibited by high Fe(2+) or Zn(2+) concentrations. The histidine mutations altered the ability of the protein to dimerize and to bind heme. However, the protein metal content, the inability of exogenous Fe(2+) to increase catalytic activity, and the absence of evidence that the conserved histidines form a metal site provide no support for a catalytic role for a non-heme redox-active metal.  相似文献   

12.
Recent studies have shown that overexpression of the transmembrane protein Zrt- and Irt-like protein 14 (Zip14) stimulates the cellular uptake of zinc and nontransferrin-bound iron (NTBI). Here, we directly tested the hypothesis that Zip14 transports free zinc, iron, and other metal ions by using the Xenopus laevis oocyte heterologous expression system, and use of this approach also allowed us to characterize the functional properties of Zip14. Expression of mouse Zip14 in RNA-injected oocytes stimulated the uptake of (55)Fe in the presence of l-ascorbate but not nitrilotriacetic acid, indicating that Zip14 is an iron transporter specific for ferrous ion (Fe(2+)) over ferric ion (Fe(3+)). Zip14-mediated (55)Fe(2+) uptake was saturable (K(0.5) ≈ 2 μM), temperature-dependent (apparent activation energy, E(a) = 15 kcal/mol), pH-sensitive, Ca(2+)-dependent, and inhibited by Co(2+), Mn(2+), and Zn(2+). HCO(3)(-) stimulated (55)Fe(2+) transport. These properties are in close agreement with those of NTBI uptake in the perfused rat liver and in isolated hepatocytes reported in the literature. Zip14 also mediated the uptake of (109)Cd(2+), (54)Mn(2+), and (65)Zn(2+) but not (64)Cu (I or II). (65)Zn(2+) uptake also was saturable (K(0.5) ≈ 2 μM) but, notably, the metal-ion inhibition profile and Ca(2+) dependence of Zn(2+) transport differed from those of Fe(2+) transport, and we propose a model to account for these observations. Our data reveal that Zip14 is a complex, broad-scope metal-ion transporter. Whereas zinc appears to be a preferred substrate under normal conditions, we found that Zip14 is capable of mediating cellular uptake of NTBI characteristic of iron-overload conditions.  相似文献   

13.
The kinetics and spectroscopic properties of the single polypeptide and proteolytically cleaved form of recombinant Fe(3+)Fe(2+) human purple acid phosphatase (recHPAP) exhibit significant differences, primarily due to a difference in pK(es,1) (the value of an acid dissociation constant of the ES complex). These differences are due to the presence or absence, respectively, of an interaction between an aspartate residue in an exposed loop of the protein and one or more active site residues. To further explore the origin of these differences, the ferrous ion of recHPAP has been replaced by zinc. Analysis of the reconstituted Fe(3+)Zn(2+)recHPAP reveals an unexpected catalytic activity versus pH profile, in that the optimal pH is 6.3, similar to that of the proteolytically cleaved form (6.5). Moreover, replacement of the ferrous ion by zinc increases the turnover number more than 10-fold; the pK(es) values are also shifted as expected for the change in the divalent metal ion. Although the EPR spectra of both single polypeptide and proteolytically cleaved Fe(3+)Zn(2+)-recHPAP are independent of pH over the range 4.5-6.2, the visible spectrum of Fe(3+)Zn(2+)-recHPAP is pH dependent. These results suggest that the properties and environment of the divalent metal are important in determining the catalytic properties of mammalian PAPs, and in particular that a solvent molecule coordinated to the divalent metal ion may play a critical role in the catalytic cycle of these enzymes.  相似文献   

14.
Actinomycetes, such as Mycobacterium species, are Gram-positive bacteria that utilize the small molecule mycothiol (MSH) as their primary reducing agent. Consequently, the enzymes involved in MSH biosynthesis are targets for drug development. The metal-dependent enzyme N-acetyl-1-D-myo-inosityl-2-amino-2-deoxy-α-D-glucopyranoside deacetylase (MshB) catalyzes the hydrolysis of N-acetyl-1-D-myo-inosityl-2-amino-2-deoxy-α-D-glucopyranoside to form 1-D-myo-inosityl-2-amino-2-deoxy-α-D-glucopyranoside and acetate, the fourth overall step in MSH biosynthesis. Inhibitors of metalloenzymes typically contain a group that binds to the active site metal ion; thus, a comprehensive understanding of the native cofactor(s) of metalloenzymes is critical for the development of biologically effective inhibitors. Herein, we examined the effect of metal ions on the overall activity of MshB and probed the identity of the native cofactor. We found that the activity of MshB follows the trend Fe(2+) > Co(2+) > Zn(2+) > Mn(2+) and Ni(2+). Additionally, our results show that the identity of the cofactor bound to purified MshB is highly dependent on the purification conditions used (aerobic versus anaerobic), as well as the metal ion content of the medium during protein expression. MshB prefers Fe(2+) under anaerobic conditions regardless of the metal ion content of the medium and switches between Fe(2+) and Zn(2+) under aerobic conditions as the metal content of the medium is altered. These results indicate that the cofactor bound to MshB under biological conditions is dependent on environmental conditions, suggesting that MshB may be a cambialistic metallohydrolase that contains a dynamic cofactor. Consequently, biologically effective inhibitors will likely need to dually target Fe(2+)-MshB and Zn(2+)-MshB.  相似文献   

15.
Kinetics and inhibition of Na(+)/K(+)-ATPase and Mg(2+)-ATPase activity from rat synaptic plasma membrane (SPM), by separate and simultaneous exposure to transition (Cu(2+), Zn(2+), Fe(2+) and Co(2+)) and heavy metals (Hg(2+) and Pb(2+)) ions were studied. All investigated metals produced a larger maximum inhibition of Na(+)/K(+)-ATPase than Mg(2+)-ATPase activity. The free concentrations of the key species (inhibitor, MgATP(2-), MeATP(2-)) in the medium assay were calculated and discussed. Simultaneous exposure to the combinations Cu(2+)/Fe(2+) or Hg(2+)/Pb(2+) caused additive inhibition, while Cu(2+)/Zn(2+) or Fe(2+)/Zn(2+) inhibited Na(+)/K(+)-ATPase activity synergistically (i.e., greater than the sum metal-induced inhibition assayed separately). Simultaneous exposure to Cu(2+)/Fe(2+) or Cu(2+)/Zn(2+) inhibited Mg(2+)-ATPase activity synergistically, while Hg(2+)/Pb(2+) or Fe(2+)/Zn(2+) induced antagonistic inhibition of this enzyme. Kinetic analysis showed that all investigated metals inhibited Na(+)/K(+)-ATPase activity by reducing the maximum velocities (V(max)) rather than the apparent affinity (Km) for substrate MgATP(2-), implying the noncompetitive nature of the inhibition. The incomplete inhibition of Mg(2+)-ATPase activity by Zn(2+), Fe(2+) and Co(2+) as well as kinetic analysis indicated two distinct Mg(2+)-ATPase subtypes activated in the presence of low and high MgATP(2-) concentration. EDTA, L-cysteine and gluthathione (GSH) prevented metal ion-induced inhibition of Na(+)/K(+)-ATPase with various potencies. Furthermore, these ligands also reversed Na(+)/K(+)-ATPase activity inhibited by transition metals in a concentration-dependent manner, but a recovery effect by any ligand on Hg(2+)-induced inhibition was not obtained.  相似文献   

16.
Particle beam/hollow cathode-optical emission spectroscopy (PB/HC-OES) is presented as a tool for the determination of metal ion loading in transferrin (Tf). The elemental specificity of optical emission spectroscopy provides a means of assessing metal ion concentrations as well as the relative amounts of metal per unit protein concentration (up to 2 moles of Fe per mole of protein). The PB/HC-OES method allows for the simultaneous detection of metal content (Fe (I) 371.99, Ni (I) 341.41 nm, Zn (I) 213.86 nm, and Ag (I) 338.28 nm in this case), as well as elemental carbon and sulfur (C (I) 156.14 nm and S (I) 180.73 nm) that are reflective of the protein composition and concentration. Quantification for the metal species is based on calibration functions derived from aqueous solutions, with limits of detection for the entire suite being less than 1.0 μM. Determinations in this manner eliminate much of the ambiguity inherent in UV-VIS absorbance determinations of Tf metal binding. Validation of this method is obtained by analyzing loading response of Fe(3+) into Tf using the PB/HC-OES method and comparing the results with those of the standard UV-VIS absorbance method. Maximum Fe(3+) loading of Tf (based on the number of available binding sites) was determined to be 71.2 ± 4.7% by the PB/HC-OES method and 67.5 ± 2.5% for the UV-VIS absorbance method. Element emission ratios between the dopant metals and the carbon and sulfur protein constituents allow for concentration independent determinations of metal binding into Tf. Loading percentages were determined for Ni(2+), Zn(2+), and Ag(+) into Tf with maximum loading values of 19.5 ± 0.4%, 41.0 ± 4.4%, and 141.2 ± 4.3%, respectively. While of no apparent biological significance, Ag(+) presents an interesting case as a surrogate for Pt(2+), whose binding with Tf has shown to be quite different from the other metals. A different mode from the others is indeed observed, and is consistent with conjecture on the Pt(2+) mechanisms. Competitive binding studies not easily performed using absorbance spectroscopy are easily performed by simultaneous, multielement analysis, reflective of the metals and the protein content. In this work, there is clear competition between and Fe(3+) and Zn(2+) for binding in the C-terminus lobe of Tf, while Ni(2+) binds within the N-terminus lobe. Addition of Ag(+) to this mixture does not affect the other metals' distributions, but reflects binding at other protein sites.  相似文献   

17.
To achieve cellular iron deprivation by chelation, it is important to develop chelators with selective metal-binding properties. Selectivity for iron has long been the province of certain oxygen-donor chelators such as desferrioxamine, which target Fe(III) and exploit the strength of a relatively ionic Fe(III)-O interaction. We have been studying novel chelators that possess mechanisms to selectively chelate +2 biometals, particularly tachpyr [N,N',N"-tris(2-pyridylmethyl)-1,3,5-cis,cis-triaminocyclohexane] and derivatives from N,N',N"-trialkylation and pyridine ring alkylation. Metal-exchange and metal-binding competition reactions have been conducted at pH 7.4, 37 degrees C and time periods until no further change was observed (generally 24-48 h). Under anaerobic conditions, tachpyr is strongly selective for iron, binding 95+/-5% Fe(II) versus 5+/-5% Zn(II) in the forms [Fe(tachpyr)](2+) and [Zn(tachpyr)](2+) respectively. Under aerobic conditions, tachpyr complexes Fe(II) more effectively than Fe(III), forming iminopyridyl complexes [Fe(tachpyr-ox-n)](2+) (n=2, 4) by O(2)-induced and iron-mediated oxidative dehydrogenation. Complexes [Fe(tachpyr-ox-n)](2+) are also strongly bound forms of iron that are unaffected by an excess of Zn(II) (75 mol zinc:1 mol iron complex). The preference of tachpyr for iron over zinc under aerobic conditions appears to be hindered by oxidation of Fe(II) to Fe(III), such that the proportions bound are 44+/-10% Fe(II) versus 56+/-10% Zn(II), in the respective forms [Fe(tachpyr-ox-n)](2+) and [Zn(tachpyr)](2+). However, upon addition of the reducing agent Na(2)S(2)O(4) that converts Fe(III) to Fe(II), the binding proportions shift to 76+/-10% Fe(II) versus 24+/-10% Zn(II), demonstrating a clear preference of tachpyr for Fe(II) over Zn(II). Iron(II) is in the low-spin state in [Fe(tachpyr)](2+) and [Fe(tachpyr-ox-n)](2+) (n=2, 4), which is a likely cause of the observed selectivity. N-methylation of tachpyr [giving (N-methyl)(3)tachpyr] results in the loss of selectivity for Fe(II), which is attributed to the steric effect of the methyl groups and a resulting high-spin state of Fe(II) in [Fe(N-methyl)(3)tachpyr)](2+). The relationship of chelator selectivity to cytotoxicity in the tach family will be discussed.  相似文献   

18.
Ferritins are nearly ubiquitous iron storage proteins playing a fundamental role in iron metabolism. They are composed of 24 subunits forming a spherical protein shell encompassing a central iron storage cavity. The iron storage mechanism involves the initial binding and subsequent O2-dependent oxidation of two Fe2+ ions located at sites A and B within the highly conserved dinuclear "ferroxidase center" in individual subunits. Unlike animal ferritins and the heme-containing bacterioferritins, the Escherichia coli ferritin possesses an additional iron-binding site (site C) located on the inner surface of the protein shell close to the ferroxidase center. We report the structures of five E. coli ferritin variants and their Fe3+ and Zn2+ (a redox-stable alternative for Fe2+) derivatives. Single carboxyl ligand replacements in sites A, B, and C gave unique effects on metal binding, which explain the observed changes in Fe2+ oxidation rates. Binding of Fe2+ at both A and B sites is clearly essential for rapid Fe2+ oxidation, and the linking of FeB2+ to FeC2+ enables the oxidation of three Fe2+ ions. The transient binding of Fe2+ at one of three newly observed Zn2+ sites may allow the oxidation of four Fe2+ by one dioxygen molecule.  相似文献   

19.
Zinc ions are required to maintain the biological activity of numerous proteins. However, when mislocalized or accumulated in excess, Zn(2+) ions are toxic because of adventitious binding to proteins and displacement of other metal ions, among them Fe(2+), from their binding sites. Heterologous expression of a previously uncharacterized Arabidopsis thaliana metal tolerance protein, MTP3, in the zrc1 cot1 mutant of budding yeast restores tolerance to, and cellular accumulation of, zinc and cobalt. An MTP3-GFP fusion protein localizes to the vacuolar membrane when expressed in Arabidopsis. Ectopic over-expression of MTP3 increases Zn accumulation in both roots and rosette leaves of A. thaliana, and enhances Zn tolerance. Exposure of wild-type plants to high but non-toxic concentrations of Zn or Co, or Fe deficiency, strongly induce MTP3 expression specifically in epidermal and cortex cells of the root hair zone. Silencing of MTP3 by RNA interference causes Zn hypersensitivity and enhances Zn accumulation in above-ground organs of soil-grown plants and of seedlings exposed to excess Zn or to Fe deficiency. Our data indicate that, in wild-type A. thaliana, the AtMTP3 protein contributes to basic cellular Zn tolerance and controls Zn partitioning, particularly under conditions of high rates of Zn influx into the root symplasm.  相似文献   

20.
The effect of Zn(2+) or Cu(2+) ions on Mn-depleted photosystem II (PS II) has been investigated using EPR spectroscopy. In Zn(2+)-treated and Cu(2+)-treated PS II, chemical reduction with sodium dithionite gives rise to a signal attributed to the plastosemiquinone, Q(A)(*)(-), the usual interaction with the non-heme iron being lost. The signal was identified by Q-band EPR spectroscopy which partially resolves the typical g-anisotropy of the semiquinone anion radical. Illumination at 200 K of the unreduced samples gives rise to a single organic free radical in Cu(2+)-treated PS II, and this is assigned to a monomeric chlorophyll cation radical, Chl a(*)(+), based on its (1)H-ENDOR spectrum. The Zn(2+)-treated PS II under the same conditions gives rise to two radical signals present in equal amounts and attributed to the Chl a(*)(+) and the Q(A)(*)(-) formed by light-induced charge separation. When the Cu(2+)-treated PS II is reduced by sodium ascorbate, at >/=77 K electron donation eliminates the donor-side radical leaving the Q(A)(*)(-) EPR signal. The data are explained as follows: (1) Cu(2+) and Zn(2+) have similar effects on PS II (although higher concentrations of Zn(2+) are required) causing the displacement of the non-heme Fe(2+). (2) In both cases chlorophyll is the electron donor at 200 K. It is proposed that the lack of a light-induced Q(A)(*)(-) signal in the unreduced Cu(2+)-treated sample is due to Cu(2+) acting as an electron acceptor from Q(A)(*)(-) at low temperature, forming the Cu(+) state and leaving the electron donor radical Chl a(*)(+) detectable by EPR. (3) The Cu(2+) in PS II is chemically reducible by ascorbate prior to illumination, and the metal can therefore no longer act as an electron acceptor; thus Q(A)(*)(-) is generated by illumination in such samples. (4) With dithionite, both the Cu(2+) and the quinone are reduced resulting in the presence of Q(A)(*)(-) in the dark. The suggested high redox potential of Cu(2+) when in the Fe(2+) site in PS II is in contrast to the situation in the bacterial reaction center where it has been shown in earlier work that the Cu(2+) is unreduced by dithionite. It cannot be ruled out however that Q(A)-Cu(2+) is formed and a magnetic interaction is responsible for the lack of the Q(A)(-) signal when no exogenous reductant is present. With this alternative possibility, the effects of reductants would be explained as the loss of Cu(2+) (due to formation of Cu(+)) leading to loss of the Cu(2+) from the Fe(2+) site due to the binding equilibrium. The quite different binding and redox behavior of the metal in the iron site in PS II compared to that of the bacterial reaction center is presumably a further reflection of the differences in the coordination of the iron in the two systems.  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