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 62 毫秒
1.
Electrochemistry of microperoxidase-11 (MPx-11) anchored on the mixed self-assembled monolayer (SAM) of 2-(2-mercaptoethylpyrazine) (PET) and 4,4'-dithiodibutyric acid (DTB) on gold (Au) electrode and the biosensing of uric acid (UA) is described. MPx-11 has been covalently anchored on the mixed SAM of PET and DTB on Au electrode. MPx-11 on the mixed self-assembly exhibits reversible redox response characteristic of a surface confined species. The heterocyclic ring of PET promotes the electron transfer between the electrode and the redox protein. The apparent standard rate constant kapps obtained for the redox reaction of MPx-11 on the mixed monolayer is approximately 2.15 times higher than that on the single monolayer of DTB modified electrode. MPx-11 efficiently mediates the electrocatalytic reduction of H2O2. MPx-11 electrode is highly sensitive to H2O2 and it shows linear response for a wide concentration range. The electrocatalytic activity of the MPx-11 electrode is combined with the enzymatic activity of uricase (UOx) to fabricate uric acid biosensor. The bienzyme assembly is highly sensitive towards UA and it could detect UA as low as 2 microM at the potential of -0.1 V. The biosensor shows linear response with a sensitivity of 3.4+/-0.08 nA cm(-2) microM(-1). Ascorbate (AA) and paracetamol (PA) do not significantly interfere in the amperometric sensing of UA.  相似文献   

2.
The electrochemical characteristics of kanamycin onto self-assembled monolayer (SAM) modified gold electrode (SAM/Au) is investigated by cyclic voltammetry. In the potential range 0-0.6 V, Cu(II) yields a pair of stable redox waves at the bare gold electrode. E(pa) is located at 0.189 V and E(pc) at 0.254 V. In contrast, Cu(II) is reduced at a more positive potential and a decreasing current at the kanamycin SAM/Au electrode. Cu(II) and kanamycin can form a stoichiometry complex with chemical ratio of 2:1. The interaction of Cu(II)-kanamycin complex with calf thymus DNA is also studied in the solution. And the interactive mode between Cu(II)-kanamycin complex and DNA is verified by the fluorescence method. Binding constants (K) of the Cu(II)-kanamycin complex to DNA and binding site size (s) are calculated from voltammetric data and equal to 1.5 x 10(7) l/mol and 4 bp, respectively.  相似文献   

3.
A surface modification procedure for the creation of self-assembled monolayers (SAMs) that can be used as a scaffold for double-stranded DNA (dsDNA) incorporation onto the gold surfaces is described. The SAMs of an azidohexane thiol derivative were prepared on the Au electrode and then used for the immobilization of dsDNA. The electrochemical characteristics of dsDNA onto the SAM-modified gold electrode were investigated by cyclic voltammetry and electrochemical impedance spectroscopy, and the surface concentration of dsDNA onto the SAMs surface was estimated. The interaction of dsDNA with the anticancer drug, taxol (paclitaxel), was also studied on the surface of DNA/SAM/Au electrode. The observed decrease in the guanine oxidation peak current was used to monitor the interaction of taxol with DNA. The resulting Langmuir isotherm for taxol binding to DNA at the modified electrode was used to evaluate the binding constant of taxol-DNA. The results obtained supported the groove binding interaction of taxol with DNA. The modified electrode was used as a sensitive sensor for quantification of taxol in human serum sample.  相似文献   

4.
MutY and endonuclease III, two DNA glycosylases from Escherichia coli, and AfUDG, a uracil DNA glycosylase from Archeoglobus fulgidus, are all base excision repair enzymes that contain the [4Fe-4S](2+) cofactor. Here we demonstrate that, when bound to DNA, these repair enzymes become redox-active; binding to DNA shifts the redox potential of the [4Fe-4S](3+/2+) couple to the range characteristic of high-potential iron proteins and activates the proteins toward oxidation. Electrochemistry on DNA-modified electrodes reveals potentials for Endo III and AfUDG of 58 and 95 mV versus NHE, respectively, comparable to 90 mV for MutY bound to DNA. In the absence of DNA modification of the electrode, no redox activity can be detected, and on electrodes modified with DNA containing an abasic site, the redox signals are dramatically attenuated; these observations show that the DNA base pair stack mediates electron transfer to the protein, and the potentials determined are for the DNA-bound protein. In EPR experiments at 10 K, redox activation upon DNA binding is also evident to yield the oxidized [4Fe-4S](3+) cluster and the partially degraded [3Fe-4S](1+) cluster. EPR signals at g = 2.02 and 1.99 for MutY and g = 2.03 and 2.01 for Endo III are seen upon oxidation of these proteins by Co(phen)(3)(3+) in the presence of DNA and are characteristic of [3Fe-4S](1+) clusters, while oxidation of AfUDG bound to DNA yields EPR signals at g = 2.13, 2.04, and 2.02, indicative of both [4Fe-4S](3+) and [3Fe-4S](1+) clusters. On the basis of this DNA-dependent redox activity, we propose a model for the rapid detection of DNA lesions using DNA-mediated electron transfer among these repair enzymes; redox activation upon DNA binding and charge transfer through well-matched DNA to an alternate bound repair protein can lead to the rapid redistribution of proteins onto genome sites in the vicinity of DNA lesions. This redox activation furthermore establishes a functional role for the ubiquitous [4Fe-4S] clusters in DNA repair enzymes that involves redox chemistry and provides a means to consider DNA-mediated signaling within the cell.  相似文献   

5.
This article describes the selective determination of guanine (G) using the self-assembled monolayer (SAM) of 1,8,15,22-tetraaminophthalocyanatonickel(II) (4α-Ni(II)TAPc) modified glassy carbon electrode (GCE) in 0.2 M acetate buffer solution (pH 4.0). The SAM of 4α-Ni(II)TAPc was formed on GCE by spontaneous adsorption of 1 mM 4α-Ni(II)TAPc in dimethylformamide (DMF). It shows two pairs of redox waves corresponding to Ni(III)/Ni(II) and Ni(III)Pc(-1)/Ni(III)Pc(-2) in 0.2 M acetate buffer solution. The SAM modified electrode exhibits excellent electrocatalytic activity toward the oxidation of G by enhancing its oxidation current with 150 mV less positive potential shift in contrast to bare GCE. Furthermore, the SAM modified electrode selectively determines G in the presence of high concentration of adenine (A). In differential pulse voltammetry measurements, the oxidation current response of G was increased linearly in the concentration range of 10 to 100 μM, and a detection limit was found to be 3×10(-8)M (signal/noise=3).  相似文献   

6.
The 12p13 ETV6 (TEL) gene is frequently targeted by chromosomal translocations in human malignancies, resulting in the formation of oncogenic ETV6 gene fusions. Many of the known partner genes encode protein tyrosine kinases (PTKs), generating fusion proteins that function as chimeric PTKs. ETV6-NTRK3 (EN), comprised of the ETV6 SAM domain fused to the NTRK3 PTK, is unique among ETV6 chimeric oncoproteins, as it is expressed in cancers of multiple lineages. We initially hypothesized that, similar to other ETV6-PTK chimeras, SAM-mediated dimerization of EN leads to constitutive activation of the PTK and downstream signaling cascades. However, when the EN SAM domain was replaced with an inducible FK506 binding protein (FKBP) dimerization system, resulting FKBP-NTRK3 chimeras failed to transform NIH 3T3 cells even though PTK activation was preserved. It was recently shown that the ETV6 SAM domain has two potential interacting surfaces, raising the possibility that this domain can mediate protein polymerization. We therefore mutated each EN SAM binding interface in a manner shown previously to abolish self-association of wild-type ETV6. Each mutation completely blocked the ability of EN to polymerize, to activate its PTK, and to transform NIH 3T3 cells. Furthermore, EN itself formed large polymeric structures within cells while mutant EN proteins were present only as monomers. Finally, we observed a dominant negative effect on the transformation of isolated SAM domains coexpressed in EN-transformed cells. Taken together, our results suggest that higher-order polymerization may be a critical requirement for the transformation activity of EN and possibly other ETV6-PTK fusion proteins.  相似文献   

7.
To determine the effect of the redox state of the Rieske protein on ligand binding to the quinol oxidation site of the bc(1) complex, we measured the binding rate constants (k(1)) for stigmatellin and myxothiazol, at different concentrations of decylbenzoquinone or decylbenzoquinol, in the bovine bc(1) complex with the Rieske protein in the oxidized or reduced state. Stigmatellin and myxothiazol bound tightly and competitively with respect to quinone or quinol, independently of the redox state of the Rieske protein. In the oxidized bc(1) complex, the k(1) values for stigmatellin ( approximately 2.6 x 10(6) m(-1)s(-1)) and myxothiazol ( approximately 8 x 10(5) m(-1)s(-1)), and the dissociation constant (K(d)) for quinone, were similar between pH 6.5 and 9, indicating that ligand binding is independent of the protonation state of histidine 161 of the Rieske protein (pK(a) approximately 7.6). Reduction of the Rieske protein increased the k(1) value for stigmatellin and decreased the K(d) value for quinone by 50%, without modifying the k(1) for myxothiazol. These results indicate that reduction of the Rieske protein and protonation of histidine 161 do not induce a strong stabilization of ligand binding to the quinol oxidation site, as assumed in models that propose the existence of a highly stabilized semiquinone as a reaction intermediate during quinol oxidation.  相似文献   

8.
Clarkson J  Smith DA 《FEBS letters》2001,503(1):30-34
The binding of sulphate to human serum apo-transferrin has been examined by ultraviolet absorption and ultraviolet resonance Raman difference spectroscopies between pH 6.0 and 9.0. The ultraviolet absorption data reveals a negative feature at 245 nm that increases in magnitude with pH, with an apparent pK(a) of 7.57, which the Raman difference data reveals to be due to tyrosine. The pK(a) of this tyrosine is unusually low and is measured at 7.84 by the Raman difference method and is elevated to greater than 9.0 upon addition of sulphate. Previous studies on the N-lobe imply that Tyr 188 is the tyrosine with a low pK(a) and also that Arg 124 is the primary binding site for the sulphate. The functional relevance may be that with sulphate bound, both carbonate binding and the deprotonation of Tyr will be disfavoured, and as a result so is iron binding.  相似文献   

9.
Luzhkov VB  Almlöf M  Nervall M  Aqvist J 《Biochemistry》2006,45(36):10807-10814
We report results from microscopic molecular dynamics and free energy perturbation simulations of substrate binding and selectivity for the Escherichia coli high-affinity ammonium transporter AmtB. The simulation system consists of the protein embedded in a model membrane/water surrounding. The calculated absolute binding free energies for the external NH(4)(+) ions are between -5.8 and -7.3 kcal/mol and are in close agreement with experimental data. The apparent pK(a) of the bound NH(4)(+) increases by more than 4 units, indicating a preference for binding ammonium ion and not neutral ammonia. The external binding site is also selective for NH(4)(+) toward monovalent metal cations by 2.4-4.4 kcal/mol. The externally bound NH(4)(+) shows strong electrostatic interactions with the proximal buried Asp160, stabilized in the anionic form, whereas the interactions with the aromatic rings of Phe107 and Trp148, lining the binding cavity, are less pronounced. Simulated mutation of the highly conserved Asp160 to Asn reduces the pK(a) of the bound ammonium ion by approximately 7 units and causes loss of its binding. The calculations further predict that the substrate affinity of E. coli AmtB depends on the ionization state of external histidines. The computed free energies of hypothetical intermediate states related to transfer of NH(3), NH(4)(+), or H(2)O from the external binding site to the first position inside the internal channel pore favor permeation of the neutral species through the channel interior. However, the predicted change in the apparent pK(a) of NH(4)(+) upon translocation from the external site, Am1, to the first internal site, Am2, indicates that ammonium ion becomes deprotonated only when it enters the channel interior.  相似文献   

10.
The redox potential of the Rieske [2Fe-2S] cluster of the bc1 complex from bovine heart mitochondria was determined by cyclic voltammetry of a water-soluble fragment of the iron/sulfur protein. At the nitric-acid-treated bare glassy-carbon electrode, the fragment gave an immediate and stable quasireversible response. The midpoint potential at pH 7.2, 25 degrees C and I of 0.01 M was Em = +312 +/- 3 mV. This value corresponds within 20 mV to results of an EPR-monitored dye-mediated redox titration. With increasing ionic strength, the midpoint potential decreased linearly with square root of I up to I = 2.5 M. From the cathodic-to-anodic peak separation, the heterogeneous rate constant, k degrees, was calculated to be approximately 2 x 10(-3) cm/s at low ionic strength; the rate constant increased with increasing ionic strength. From the temperature dependence of the midpoint potential, the standard reaction entropy was calculated as delta S degrees = -155 J.K-1.mol-1. The pH dependence of the midpoint potential was followed over pH 5.5-10. Above pH 7, redox-state-dependent pK changes were observed. The slope of the curve, -120 mV/pH above pH9, indicated two deprotonations of the oxidized protein. The pKa values of the oxidized protein, obtained by curve fitting, were 7.6 and 9.2, respectively. A group with a pKa,ox of approximately 7.5 could also be observed in the optical spectrum of the oxidized protein. Redox-dependent pK values of the iron/sulfur protein are considered to be essential for semiquinone oxidation at the Qo center of the bc1 complex.  相似文献   

11.
Photosystem II (PSII) of photosynthesis has the unique ability to photochemically oxidize water. Recently an engineered bacterioferritin photochemical ‘reaction centre’ (BFR-RC) using a zinc chlorin pigment (ZnCe6) in place of its native heme has been shown to photo-oxidize bound manganese ions through a tyrosine residue, thus mimicking two of the key reactions on the electron donor side of PSII. To understand the mechanism of tyrosine oxidation in BFR-RCs, and explore the possibility of water oxidation in such a system we have built an atomic-level model of the BFR-RC using ONIOM methodology. We studied the influence of axial ligands and carboxyl groups on the oxidation potential of ZnCe6 using DFT theory, and finally calculated the shift of the redox potential of ZnCe6 in the BFR-RC protein using the multi-conformational molecular mechanics–Poisson-Boltzmann approach. According to our calculations, the redox potential for the first oxidation of ZnCe6 in the BRF-RC protein is only 0.57 V, too low to oxidize tyrosine. We suggest that the observed tyrosine oxidation in BRF-RC could be driven by the ZnCe6 di-cation. In order to increase the efficiency of tyrosine oxidation, and ultimately oxidize water, the first potential of ZnCe6 would have to attain a value in excess of 0.8 V. We discuss the possibilities for modifying the BFR-RC to achieve this goal.  相似文献   

12.
Efimov I  McIntire WS 《Biochemistry》2004,43(32):10532-10546
The spectral and redox properties are described for the wild-type and Y384F mutant forms of the flavoprotein component (PchF) of flavocytochrome, p-cresol methylhydroxylase (PCMH), and cytochrome-free PchF that harbor FAD analogues. The analogues are iso-FAD (8-demethyl-6-methyl-FAD), 6-amino-FAD (6-NH(2)-FAD), 6-bromo-FAD (6-Br-FAD), 8-nor-8-chloro-FAD (8-Cl-FAD), and 5-deaza-5-carba-FAD (5-deaza-FAD). All of the analogues bound noncovalently and stoichiometrically to cytochrome-free apo-PchF, and the resulting holoproteins had high affinity for the cytochrome subunit, PchC. Noncovalently bound FAD, 6-Br-FAD, or 6-NH(2)-FAD can be induced to bind covalently by exposing holo-PchF to PchC. The rate of this process and the redox potential of the noncovalently bound flavin may be correlated. In addition, the redox potential of each FAD analogue was higher when it was covalently bound than when noncovalently bound to PchF. Furthermore, the potential of a covalently bound or noncovalently bound FAD analogue increased on association of the corresponding holo-PchF with PchC, and the activity increased as the flavin's redox potential increased. It was discovered also that 4-hydroxybenzaldehyde, the final p-cresol oxidation product, is an efficient competitive inhibitor for substrate oxidation by PchF since it binds tightly to this protein when the flavin is oxidized, although it binds more loosely to the enzyme with reduced flavin. Finally, the energies of the charge-transfer bands for the interaction of bound flavin analogues with 4-Br-phenol (a substrate mimic) increased as the potential decreases, although a simple global correlation was not seen. This is the case because the energy is also a function of the redox properties of the bound mimic. The implications of these findings to covalent flavinylation and catalysis are discussed.  相似文献   

13.
A refinement of the protonmotive Q cycle mechanism is proposed in which oxidation of ubiquinol is a concerted reaction and occurs by an alternating, half-of-the-sites mechanism. A concerted mechanism of ubiquinol oxidation is inferred from the finding that there is reciprocal control between the high potential and low potential redox components involved in ubiquinol oxidation. The potential of the Rieske iron-sulfur protein controls the rate of reduction of the b cytochromes, and the potential of the b cytochromes controls the rate of reduction of the Rieske protein and cytochrome c(1). A concerted mechanism of ubiquinol oxidation reconciles the findings that the ubiquinol-cytochrome c reductase kinetics of the bc(1) complex include both a pH dependence and a dependence on Rieske iron-sulfur protein midpoint potential.An alternating, half-of-the-sites mechanism for ubiquinol oxidation is inferred from the finding that some inhibitory analogs of ubiquinol that block ubiquinol oxidation by binding to the ubiquinol oxidation site in the bc(1) complex inhibit the yeast enzyme with a stoichiometry of 0.5 per bc(1) complex. One molecule of inhibitor is sufficient to fully inhibit the dimeric enzyme, and the binding is anti-cooperative, in that a second molecule of inhibitor binds with much lower affinity to a dimer in which an inhibitor molecule is already bound. An alternating, half-of-the-sites mechanism implies that, at least under some conditions, only half of the sites in the dimeric enzyme are reactive at any one time. This provides a raison d'être for the dimeric structure of the enzyme, in that bc(1) activity may be regulated and capable of switching between a half-of-the-sites active and a fully active enzyme.  相似文献   

14.
The pH dependence of the affinity of a 11-mer phosphotyrosine (pY) peptide (EPQpYEEIPIYL-NH2) for the SH2 domain of the tyrosine kinase p56(lck) was investigated with surface plasmon resonance (SPR). From SPR competition experiments the affinity in solution was obtained. The pH dependence of the affinity in solution can be well described by a proton linkage model with a single pK(a) shift upon binding, from 6.1 to 4.7. This shift is ascribed to the transition from the -2 to the -1 ionisation state of the tyrosine phosphate group. Based on the X-ray structure for the complex with Lck SH2, a pK(a) value of 5.3 for the bound pY peptide was computed, modelling the solvated protein as a system of point charges in a continuum. With the phosphate in the -2 state the binding energy is 1.8 kcal/mol more favourable than for the -1 state, corresponding to a 20-fold higher affinity. A proper charge is relevant in the design of potential therapeutic Lck SH2 ligands with mimics for the metabolically unstable tyrosine phosphate group.  相似文献   

15.
The side chain of aspartate 95 in flavodoxin from Desulfovibrio vulgaris provides the closest negative charge to N(1) of the bound FMN in the protein. Site-directed mutagenesis was used to substitute alanine, asparagine, or glutamate for this amino acid to assess the effect of this charge on the semiquinone/hydroquinone redox potential (E(1)) of the FMN cofactor. The D95A mutation shifts the E(1) redox potential positively by 16 mV, while a negative shift of 23 mV occurs in the oxidized/semiquinone midpoint redox potential (E(2)). The crystal structures of the oxidized and semiquinone forms of this mutant are similar to the corresponding states of the wild-type protein. In contrast to the wild-type protein, a further change in structure occurs in the D95A mutant in the hydroquinone form. The side chain of Y98 flips into an energetically more favorable edge-to-face interaction with the bound FMN. Analysis of the structural changes in the D95A mutant, taking into account electrostatic interactions at the FMN binding site, suggests that the pi-pi electrostatic repulsions have only a minor contribution to the very low E(1) redox potential of the FMN cofactor when bound to apoflavodoxin. Substitution of D95 with glutamate causes only a slight perturbation of the two one-electron redox potentials of the FMN cofactor. The structure of the D95E mutant reveals a large movement of the 60-loop (residues 60-64) away from the flavin in the oxidized structure. Reduction of this mutant to the hydroquinone causes the conformation of the 60-loop to revert back to that occurring in the structures of the wild-type protein. The crystal structures of the D95E mutant imply that electrostatic repulsion between a carboxylate on the side chain at position 95 and the phenol ring of Y98 prevents rotation of the Y98 side chain to a more energetically favorable conformation as occurs in the D95A mutant. Replacement of D95 with asparagine has no effect on E(2) but causes E(1) to change by 45 mV. The D95N mutant failed to crystallize. The K(d) values of the protein FMN complex in all three oxidation-reduction states differ from those of the wild-type complexes. Molecular modeling showed that the conformational energy of the protein changes with the redox state, in qualitative agreement with the observed changes in K(d), and allowed the electrostatic interactions between the FMN and the surrounding groups on the protein to be quantified.  相似文献   

16.
The direct electrochemical redox reaction of bovine erythrocyte copper--zinc superoxide dismutase (Cu(2)Zn(2)SOD) was clearly observed at a gold electrode modified with a self-assembled monolayer (SAM) of cysteine in phosphate buffer solution containing SOD, although its reaction could not be observed at the bare electrode. In this case, SOD was found to be stably confined on the SAM of cysteine and the redox response could be observed even when the cysteine-SAM electrode used in the SOD solution was transferred to the pure electrolyte solution containing no SOD, suggesting the permanent binding of SOD via the SAM of cysteine on the electrode surface. The electrode reaction of the SOD confined on the cysteine-SAM electrode was found to be quasi-reversible with the formal potential of 65 +/- 3 mV vs. Ag/AgCl and its kinetic parameters were estimated: the electron transfer rate constant k(s) is 1.2 +/- 0.2 s(-1) and the anodic (alpha(a)) and cathodic (alpha(c)) transfer coefficients are 0.39 +/- 0.02 and 0.61 +/- 0.02, respectively. The assignment of the redox peak of SOD at the cysteine-SAM modified electrode could be sufficiently carried out using the native SOD (Cu(2)Zn(2)SOD), its Cu- or Zn-free derivatives (E(2)Zn(2)SOD and Cu(2)E(2)SOD, E designates an empty site) and the SOD reconstituted from E(2)Zn(2)SOD and Cu(2+). The Cu complex moiety, the active site for the enzymatic dismutation of the superoxide ion, was characterized to be also the electroactive site of SOD. In addition, we found that the SOD confined on the electrode can be expected to possess its inherent enzymatic activity for dismutation of the superoxide ion.  相似文献   

17.
Leggate EJ  Hirst J 《Biochemistry》2005,44(18):7048-7058
Rieske [2Fe-2S] clusters have reduction potentials which vary by over 500 mV, and which are pH dependent. In the cytochrome bc(1) complex, the high-potential and low-pK values of the cluster may be important in the mechanism of quinol oxidation. Hydrogen bonds, from both side-chain and mainchain groups, are crucial for these properties, but solvent accessibility and a disulfide bond (present in only high-potential Rieske proteins) have been suggested to be important determinants also. Previous studies have addressed the hydrogen bonds, disulfide bond, and a leucine residue which may restrict solvent access, by mutations in the cytochrome bc(1) complex. However, influences on the complex (disruption of quinol binding and displacement of the Rieske domain) are difficult to deconvolute from intrinsic effects on the Rieske cluster. Here, the effects of similar mutations on cluster potential, pK values, and stability are characterized comprehensively in the isolated Rieske domain of the bovine protein. Hydrogen bonds from Ser163 and Tyr165 are important in increasing the reduction potential and decreasing the pK values. The disulfide has a limited effect on the redox properties, but is crucial for cluster stability, particularly in the oxidized state. Mutations of Leu142 had little effect on cluster potential, pK values, or stability, in contrast to the significant effects which were observed in the complex. The sum of the effects of all the mutated residues accounts for most of the differences between high- and low-potential Rieske proteins.  相似文献   

18.
大肠杆菌拓扑异构酶 I(E. coli TopA)属于 I 型拓扑异构酶,在DNA复制、转录、重组和基因表达调控等过程中发挥关键作用。E. coli TopA 不仅能结合锌,还可以结合铁。细胞内过量铁可与锌竞争,通过与锌指结构域结合减弱其 DNA 结合能力和改变蛋白质空间构象,从而抑制TopA拓扑异构酶活性。然而,铁结合形式TopA的氧化还原特性以及氧化还原条件对其活性的影响仍不清楚。本研究通过紫外分光光谱和体外DNA拓扑异构酶活性分析,发现体外纯化得到的铁结合形式的 TopA 呈氧化状态,能够被二硫苏糖醇和连二亚硫酸钠还原,原本氧化状态下无活性的TopA在还原条件下,可恢复其拓扑异构酶活性。当还原剂被去除后,铁结合的TopA在空气中能够重新被氧化,且其活性重新受到抑制。这说明,氧化还原条件对铁结合的 TopA 功能具有可逆调节作用。通过金属 蛋白体外结合实验进一步发现,无金属结合的TopA蛋白(apo-TopA)在无氧条件下,与 Fe2+ 和 Fe3+ 均能结合,但与Fe2+ 结合能力较弱,并且TopA结合的Fe3+ 被还原成Fe2+ 后,结合力显著下降,能够被铁螯合指示剂菲咯嗪快速捕获。此外,蛋白质内源性荧光光谱分析实验表明,铁结合的TopA在氧化还原的不同状态时,其在330 nm左右的荧光值有显著差异。这提示,氧化还原条件可能通过影响铁离子与TopA的结合状态,引起蛋白质空间构象改变,从而对TopA的拓扑异构酶活性进行调节。此研究表明,铁结合TopA的拓扑异构酶活性会受到细胞内氧化还原信号的可逆调控,也提示I型拓扑异构酶可能是细胞铁超载通过氧化损伤引起细胞功能障碍(或铁死亡)的靶点之一。  相似文献   

19.
The redox potential of deazariboflavin has been determined for pH values from 5.5 to 9.2 by equilibration with riboflavin and lumiflavin 3-acetate. The position of the equilibrium with riboflavin was measured spectrophotometrically and fluorimetrically; the equilibrium potential with lumiflavin 3-acetate was measured spectrophotometrically and potentiometrically. The Em7 for deazariboflavin was found to be--0.273 +/- 0.003 V against the standard hydrogen electrode. Equilibrium with flavodoxin at pH 9.5 and 10.0 was also used to determine the redox potential of deazariboflavin at high pH values. The pK of dihydrodeazariboflavin was found from the break in the potential vs. pH diagram and from spectrophotometric pH titration. The pK value obtained by both methods is 7.00 +/- 0.05. We found that borate, a product of the reducing agent borohydride, complexed with the ribityl sidechain of deazariboflavin, causing a shift in the pK for the reduced form to values of about 8.  相似文献   

20.
The similar dimensions of biomolecules such as enzymes, antibodies or DNA, and metallic or semiconductor nanoparticles (NPs) enable the synthesis of biomolecule-NP hybrid systems where the unique electronic, photonic and catalytic properties of NPs are combined with the specific recognition and biocatalytic properties of biomolecules. The unique functions of biomolecule-NP hybrid systems are discussed with several examples: (i) the electrical contacting of redox enzymes with electrodes is the basis for the development of enzymatic electrodes for amperometric biosensors or biofuel cell elements. The reconstitution of the apo-glucose oxidase or apo-glucose dehydrogenase on flavin adenine dinucleotide (FAD)-functionalized Au NPs (1.4 nm) associated with electrodes, or on pyrroloquinoline quinone (PQQ)-functionalized Au NPs (1.4 nm) associated with electrodes, respectively, yields electrically contacted enzyme electrodes. The aligned, reconstituted enzymes on the electrode surfaces reveal effective electrical contacting, and the glucose oxidase and glucose dehydrogenase reveal turnover rates of 5000 and 11,800 s(-1), respectively. (ii) The photoexcitation of semiconductor nanoparticles yields fluorescence with a wavelength controlled by the size of the NPs. The fluorescence functions of semiconductor NPs are used to develop a fluorescence resonance energy transfer (FRET) assay for nucleic acids, and specifically, for analyzing telomerase activity in cancer cells. CdSe-ZnS NPs are functionalized by a primer recognized by telomerase, and this is elongated by telomerase extracted from HeLa cancer cells in the presence of dNTPs and Texas-red-functionalized dUTP. The dye integrated into the telomers allows the FRET process that is intensified as telomerization proceeds. Also, the photoexcited electron-hole pair generated in semiconductor NPs is used to generate photocurrents in a CdS-DNA hybrid system associated with an electrode. A redox-active intercalator, methylene blue, was incorporated into a CdS-duplex DNA monolayer associated with a Au electrode, and this facilitated the electron transfer between the electrode and the CdS NPs. The direction of the photocurrent was controlled by the oxidation state of the intercalator. (iii) Biocatalysts grow metallic NPs, and the absorbance of the NPs provides a means to assay the biocatalytic transformations. This is exemplified with the glucose oxidase-induced growth of Au NPs and with the tyrosinase-stimulated growth of Au NPs, in the presence of glucose or tyrosine, respectively. The biocatalytic growth of the metallic NPs is used to grow nanowires on surfaces. Glucose oxidase or alkaline phosphatase functionalized with Au NPs (1.4 nm) acted as 'biocatalytic inks' for the synthesis of metallic nanowires. The deposition of the Au NP-modified glucose oxidase, or the Au NP-modified alkaline phosphatase on Si surfaces by dip-pen nanolithography led to biocatalytic templates, that after interaction with glucose/AuCl4- or p-aminophenolphosphate/Ag+, allowed the synthesis of Au nanowires or Ag nanowires, respectively.  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