首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Electron transfer reactions between Clostridum pasteurianum flavodoxin semiquinone and various oxidants [horse heart cytochrome c, ferricyanide, and ferric ethylenediaminetetraacetic [horse heart cytochrome c, ferricyanide, and ferric ethylenediaminetetraacetic acid (EDTA)] have been studied as a function of ionic strength by using stopped-flow spectrophotometry. The cytochrome c reaction is complicated by the existence of two cytochrome species which react at different rates and whose relative concentrations are ionic strength dependent. Only the faster of these two reactions is considered here. At low ionic strength, complex formation between cytochrome c and flavodoxin is indicated by a leveling off of the pseudo-first-order rate constant at high cytochrome c concentration. This is not observed for either ferricyanide or ferric EDTA. For cytochrome c, the rate and association constants for complex formation were found to increase with decreasing ionic strength, consistent with negative charges on flavodoxin interacting with the positively charged cytochrome electron transfer site. Both ferricyanide and ferric EDTA are negatively charged oxidants, and the rate data respond to ionic strength changes as would be predicted for reactants of the same charge sign. These results demonstrate that electrostatic interactions involving negatively charged groups are important in orienting flavodoxin with respect to oxidants during electron transfer. We have also carried out computer modeling studies of putative complexes of flavodoxin with cytochrome c and ferricyanide, which relate their structural properties to both the observed kinetic behavior and some more general features of physiological electron transfer processes. The results of this study are consistent with the ionic strength behavior described above.  相似文献   

2.
We have measured the ionic strength dependence of the rate constants for the electron-transfer reactions of flavin mononucleotide (FMN) and flavodoxin semiquinones with 10 high redox potential ferredoxins (HiPIP's). The rate constants were extrapolated to infinite ionic strength by using a theoretical model of electrostatic interactions developed in our laboratory. In all cases, the sign of the electrostatic interaction was the same as the protein net charge, but the magnitudes were much smaller. The results are consistent with a model in which the electrical charges are approximately uniformly distributed over the HiPIP surface and in which there are both short- and long-range electrostatic interactions. An electrostatic field calculation for Chromatium vinosum HiPIP is consistent with this. The presumed site of electron transfer includes that region of the protein surface to which the iron-sulfur cluster is nearest and appears to be relatively hydrophobic. The principal short-range electrostatic interaction would involve the negative charge on the iron-sulfur cluster. For some net negatively charged proteins, this effect is magnified, and for net positively charged HiPIP's, it is counterbalanced. The rate constants extrapolated to infinite ionic strength can be correlated with redox potential differences between the reactants, as has previously been shown for cytochrome-flavin semiquinone reactions. Both electrostatic and redox potential effects are magnified for the flavodoxin semiquinone as compared to the FMN semiquinone-HiPIP reactions. This was also observed previously for the flavin semiquinone-cytochrome reactions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The effect of ionic strength on the rate constant for electron transfer has been used to determine the magnitude and charge sign of the net electrostatic potential which exists in close proximity to the sites of electron transfer on various c-type cytochromes. The negatively charged ferricyanide ion preferentially reacts at the positively charged exposed heme edge region on the front side of horse cytochrome c and Paracoccus cytochrome c2. In contrast, at low ionic strength, the positively charged cobalt phenanthroline ion interacts with the negatively charged back side of cytochrome c2, and at high ionic strength at a positively charged site on the front side of the cytochrome. With horse cytochrome c, over the ionic strength range studied, cobalt phenanthroline reacts only at a positively charged site which is probably not at the heme edge. These inorganic oxidants do not react at the relatively uncharged exposed heme edge sites on Azotobacter cytochrome c5 and Pseudomonas cytochrome c-551, but rather at a negatively charged site which is away from the heme edge. The results demonstrate that at least two electron-transferring sites on a single cytochrome can be functional, depending on the redox reactant used and the ionic strength. Electrostatic interactions between charge distributions on the cytochrome surface and the other reactant, or interactions involving uncharged regions on the protein(s), are critical in determining the preferred sites of electron transfer and reaction rate constants. When unfavorable electrostatic effects occur at a site near the redox center, less optimal sites at a greater distance can become kinetically important.  相似文献   

4.
As an extension of previous work from this laboratory using Clostridium pasteurianum flavodoxin [Tollin, G., Cheddar, G., Watkins, J. A., Meyer, T. E., & Cusanovich, M. A. (1984) Biochemistry 23, 6345-6349], we have measured the rate constants as a function of ionic strength for electron transfer from the semiquinones of Clostridium MP, Anacystis nidulans, and Azotobacter vinelandii flavodoxins to the following oxidants: cytochrome c from tuna and horse, Paracoccus denitrificans cytochrome c2, Pseudomonas aeruginosa cytochrome c-551, and ferricyanide. The rate constants extrapolated to infinite ionic strength (k infinity) for the C. MP flavodoxin are all slightly smaller than for the C. pasteurianum flavodoxin, as would be predicted on the basis of the higher redox potential of the C. MP protein. This indicates that there is a close similarity between the surface topographies of the two proteins in the vicinity of the coenzyme binding site. Moreover, the electrostatic interactions between the two flavodoxins and the various oxidants are also approximately the same. These studies justify our previous use of the crystallographic structure of the C. MP flavodoxin to interpret kinetic results obtained with the structurally uncharacterized C. pasteurianum flavodoxin. Despite their lower redox potentials, both Anacystis and Azotobacter flavodoxins are appreciably less reactive toward all of these oxidants (as much as 2 orders of magnitude in some cases) than are the Clostridium flavodoxins.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
We have measured the ionic strength dependence of the rate constants for electron transfer from the semiquinone of Clostridium pasteurianum flavodoxin to 12 c-type cytochromes and several inorganic oxidants using stopped-flow methodology. The experimental data were fit quite well by an electrostatic model that represents the interaction domains as parallel disks with a point charge equal to the charge within this region of the protein. The analysis provides an evaluation of the electrostatic interaction energy and the rate constant at infinite ionic strength (k affinity). The electrostatic charge on the oxidant within the interaction site can be obtained from the electrostatic energy, and for most of those reactants for which structures are available, the results are in good agreement with expectation. The k affinity values were found to correlate with redox potential differences, as expected from the theory of adiabatic (or nonadiabatic) outer-sphere electron-transfer reactions. Deviations from the theoretical curves are interpreted in terms of the influence of surface topology on reaction rate constants. In general, we find that electrostatic effects, steric influences, and redox potential all exert a much larger effect on reaction rate constants for the flavodoxin-cytochrome system than has been previously observed for free flavin-cytochrome interactions. The implications of this for determining biological specificity are discussed.  相似文献   

6.
A hypothetical model of the flavodoxin-tetraheme cytochrome c3 electron-transfer complex from the sulfate-reducing bacterium Desulfovibrio vulgaris has been constructed by using interactive computer graphics based on electrostatic potential field calculations and previous NMR experiments. Features of the proposed complex are (1) van der Waals contact between the flavin mononucleotide prosthetic group of flavodoxin and one heme of the cytochrome, (2) unique complementarity of electrostatic fields between the region surrounding this heme and the region surrounding the exposed portion of the flavin mononucleotide group of flavodoxin, and (3) no steric interferences between the two polypeptide chains in the complex. This complex is consistent with all structural and spectroscopic data available.  相似文献   

7.
The two proteins ferredoxin and flavodoxin can replace each other in the photosynthetic electron transfer chain of cyanobacteria and algae. However, structure, size, and composition of ferredoxin and flavodoxin are completely different. Ferredoxin is a small iron-sulfur protein (approximately 100 amino acids), whereas flavodoxin is a flavin-containing protein (approximately 170 amino acids). The crystal structure of both proteins from the cyanobacteria Anabeana PCC 7120 is known. We used these two protein structures to investigate the structural basis of their functional equivalence. We apply the Hodgkin index to quantify the similarity of their electrostatic potentials. The technique has been applied successfully in indirect drug design for the alignment of small molecule and bioisosterism elucidation. It requires no predefined atom-atom correspondences. As is known from experiments, electrostatic interactions are most important for the association of ferredoxin and flavodoxin with their reaction partners photosystem I and ferredoxin-NADP reductase. Therefore, use of electrostatic potentials for the structural alignment is well justified. Our extensive search of the alignment space reveals two alignments with a high degree of similarity in the electrostatic potential. In both alignments, ferredoxin overlaps completely with flavodoxin. The active sites of ferredoxin and flavodoxin rather than their centers of mass coincide in both alignments. This is in agreement with electron microscopy investigations on photosystem I cross-linked to ferredoxin or flavodoxin. We identify residues that may have the same function in both proteins and relate our results to previous experimental data.  相似文献   

8.
Rate constants have been determined for the electron-transfer reactions between reduced free flavins and flavodoxin semiquinone and several blue copper proteins. Correlations between these values and redox potentials demonstrate that spinach plastocyanin, Pseudomonas aeruginosa azurin, Alcaligenes sp. azurin, and Alcaligenes sp. nitrite reductase have the same intrinsic reactivities toward free flavins, whereas stellacyanin is more reactive (3.3 times) and laccase considerably less reactive (approximately 12 times). Electrostatic interactions between the negatively charged flavin mononucleotide (FMN) and the copper proteins show that the interaction site charges for laccase and nitrite reductase are opposite in sign to the net protein charge and that the signs and magnitudes of the charges are consistent with the known three-dimensional structures for plastocyanin and the azurins and with amino acid sequence homologies for stellacyanin. The results demonstrate that the apparent interaction site charge with flavodoxin is larger than that with FMN for plastocyanin, nitrite reductase, and stellacyanin but smaller for Pseudomonas azurin. This is interpreted in terms of a larger interaction domain for the flavodoxin reaction, which allows charged groups more distant from the actual electron-transfer site to become involved. The intrinsic reactivities of plastocyanin and azurin toward flavodoxin are the same, as was the case with FMN, but both stellacyanin and nitrite reductase are considerably less reactive than expected (approximately 2 orders of magnitude). This result suggests the involvement of steric factors with these latter two proteins which discriminate against large reactants such as flavodoxin.  相似文献   

9.
The effect of binding reduced tuna mitochondrial cytochrome c to negatively charged lipid bilayer vesicles at low ionic strength on the kinetics of electron transfer to various oxidants was studied by stopped-flow spectrophotometry. Binding strongly stimulated (up to 100-fold) the rate of reaction with the positively charged cobalt phenanthroline ion, whereas the rate of reaction with the negatively charged ferricyanide ion was greatly inhibited (up to 60-fold), as compared with the same systems either at high ionic strength or at low ionic strength either in the presence of electrically neutral vesicles or in the absence of vesicles. Reactions of tuna cytochrome c with uncharged or electrically neutral oxidants such as benzoquinone and Rhodospirillum rubrum cytochrome c2 were unaffected by binding to vesicles, suggesting little or no effect of membrane association on cytochrome structure or accessibility of the heme center. The kinetic effects were largest at lower cytochrome c to vesicle ratios, where there was a greater degree of exposure of negatively charged regions on the membrane. The reduction of cobalt phenanthroline and ferricyanide by bound cytochrome c proceeded by nonexponential kinetics, as compared with the monophasic kinetics observed in the absence of vesicles. This was probably due to the heterogeneous distribution of vesicle sizes which exists at a given lipid to protein ratio. Nonlinear oxidant concentration dependencies were observed for cobalt phenanthroline oxidation of membrane-bound cytochrome c, consistent with a (minimal) two-step kinetic mechanism involving association of the oxidant with the membrane followed by electron transfer. Based on a comparison of second-order rate constants as a function of lipid to protein mole ratio, binding of cytochrome c to the bilayer increased the efficiency of the cobalt phenanthroline reaction by a factor of approximately 500 at the highest lipid:protein ratio used. The results suggest a mechanism involving attractive and repulsive electrostatic interactions between the negatively charged bilayer and the electrically charged oxidants, which increase or decrease their effective concentrations at the membrane surface.  相似文献   

10.
The interaction of three different c-type cytochromes with flavodoxin has been studied by computer graphics modelling and computational methods. Flavodoxin and each cytochrome can make similar hypothetical electron transfer complexes that are characterized by nearly coplanar arrangement of the prosthetic groups, close intermolecular contacts at the protein-protein interface, and complementary intermolecular salt linkages. Computation of the electrostatic free energy of each complex showed that all were electrostatically stable. However, both the magnitude and behavior of the electrostatic stabilization as a function of solution ionic strength differed for the three cytochrome c-flavodoxin complexes. Variation in the computed electrostatic stabilization appears to reflect differences in the surface distribution of all charged groups in the complex, rather than differences localized at the site of intermolecular contact. The computed electrostatic association constants for the complexes and the measured kinetic rates of electron transfer in solution show a remarkable similarity in their ionic strength dependence. This correlation suggests electrostatic interactions influence electron transfer rates between protein molecules at the intermolecular association step. Comparative calculations for the three cytochrome c-flavodoxin complexes show that these ionic strength effects also involve all charged groups in both redox partners.  相似文献   

11.
The gene for the electron-transfer protein flavodoxin has been cloned from Megasphaera elsdenii using the polymerase chain reaction. The recombinant gene was sequenced, expressed in an Escherichia coli expression system, and the recombinant protein purified and characterized. With the exception of an additional methionine residue at the N-terminus, the physico-chemical properties of the protein, including its optical spectrum and oxidation-reduction properties, are very similar to those of native flavodoxin. A site-directed mutant, E60Q, was made to investigate the effects of removing the negatively charged group that is nearest to N(1) of the bound FMN. The absorbance maximum in the visible region of the bound flavin moves from 446 to 453 nm. The midpoint oxidation-reduction potential at pH 7 for reduction of oxidized flavodoxin to the semiquinone E2 becomes more negative, decreasing from -114 to -242 mV; E1, the potential for reduction of semiquinone to the hydroquinone, becomes less negative, increasing from -373 mV to -271 mV. A redox-linked pKa associated with the hydroquinone is decreased from 5.8 to < or = 4.3. The spectra of the hydroquinones of wild-type and mutant proteins depend on pH (apparent pKa values of 5.8 and < or = 5.2, respectively). The complexes of apoprotein and all three redox forms of FMN are much weaker for the mutant, with the greatest effect occurring when the flavin is in the semiquinone form. These results suggest that glutamate 60 plays a major role in control of the redox properties of M. elsdenii flavodoxin, and they provide experimental support to an earlier proposal that the carboxylate on its side-chain is associated with the redox-linked pKa of 5.8 in the hydroquinone.  相似文献   

12.
The oxidation-reduction potentials of ferredoxin-NADP+ reductase and flavodoxin from the cyanobacterium Anabaena PCC 7119 were determined by potentiometry. The potentials at pH 7 for the oxidized flavodoxin/flavodoxin semiquinone couple (E2) and the flavodoxin semiquinone/hydroquinone couple (E1) were -212 mV and -436 mV, respectively. E1 was independent of pH above about pH 7, but changed by approximately -60 mV/pH below about pH 6, suggesting that the fully reduced protein has a redox-linked pKa at about 6.1, similar to those of certain other flavodoxins. E2 varied by -50 mV/pH in the range pH 5-8. The redox potential for the two-electron reduction of ferredoxin-NADP+ reductase was -344 mV at pH 7 (delta Em = -30 mV/pH). In the 1:1 electrostatic complex of the two proteins titrated at pH 7, E2 was shifted by +8 mV and E1 was shifted by -25 mV; the shift in potential for the reductase was +4 mV. The potentials again shifted following treatment of the electrostatic complex with a carbodiimide, to covalently link the two proteins. By comparison with the separate proteins at pH 7, E2 for flavodoxin shifted by -21 mV and E1 shifted by +20 mV; the reductase potential shifted by +2 mV. The potentials of the proteins in the electrostatic and covalent complexes showed similar pH dependencies to those of the individual proteins. Qualitatively similar changes occurred when ferredoxin-NADP+ reductase from Anabaena variabilis was complexed with flavodoxin from Azotobacter vinelandii. The shifts in redox potential for the complexes were used with previously determined values for the dissociation constant (Kd) of the electrostatic complex of the two oxidised proteins, in order to estimate Kd values for the interaction of the different redox forms of the proteins. The calculations showed that the electrostatic complexes, formed when the proteins differ in their redox states, are stronger than those formed when both proteins are fully oxidized or fully reduced.  相似文献   

13.
The oxidation-reduction potentials for the riboflavin complex of the Desulfovibrio vulgaris flavodoxin are substantially different from those of the flavin mononucleotide (FMN) containing native protein, with the midpoint potential for the semiquinone-hydroquinone couple for the riboflavin complex being 180 mV less negative. This increase has been attributed to the absence in the riboflavin complex of unfavorable electrostatic effects of the dianionic 5'-phosphate of the FMN on the stability of the flavin hydroquinone anion. In this study, 15N and 1H-15N heteronuclear single-quantum coherence nuclear magnetic resonance spectroscopic studies demonstrate that when bound to the flavodoxin, (1) the N1 of the riboflavin hydroquinone remains anionic at pH 7.0 so the protonation of the hydroquinone is not responsible for this increase, (2) the N5 position is much more exposed and may be hydrogen bonded to solvent, and (3) that while the hydrogen bonding interaction at the N3H appears stronger, that at the N5H in the reduced riboflavin is substantially weaker than for the native FMN complex. Thus, the higher reduction potential of the riboflavin complex is primarily the consequence of altered interactions with the flavin ring that affect hydrogen bonding with the N5H that disproportionately destabilize the semiquinone state of the riboflavin rather than through the absence of the electrostatic effects of the 5'-phosphate on the hydroquinone state.  相似文献   

14.
The interaction of the local anesthetic tetracaine (TTC) with anionic sodium lauryl sulfate (SLS) and zwitterionic 3-(N-hexadecyl-N,N-dimethylammonio)propanesulfonate (HPS) micelles was investigated by fluorescence, spin labeling EPR and small angle X-ray scattering (SAXS). Fluorescence pH titrations allowed the choice of adequate pHs for the EPR and SAXS experiments, where either charged or uncharged TTC would be present. The data also indicated that the anesthetic is located in a less polar environment than its charged counterpart in both micellar systems. EPR spectra evidenced that both anesthetic forms increased molecular organization within the SLS micelle, the cationic form exerting a more pronounced effect. The SAXS data showed that protonated TTC causes an increase in the SLS polar shell thickness, hydration number, and aggregation number, whereas the micellar features are not altered upon incorporation of the uncharged drug. The combined results suggest that the electrostatic interaction between charged TTC and SLS, and the intercalation of the drug in the micellar polar region induce a change in molecular packing with a decrease in the mean cross-sectional area, not observed when the neutral drug sinks more deeply into the micellar hydrophobic domain. In the case of HPS micelles, the EPR spectral changes were small for the charged anesthetic and the SAXS data did not evidence any change in micellar structure, suggesting that this species protrudes more into the aqueous phase due to the lack of electrostatic attractive forces in this system.  相似文献   

15.
Flavodoxin from the nitrogen-fixing cyanobacteria Anabaena PCC 7119 forms an electron-transfer complex with ferredoxin--NADP+ reductase (FNR) from the same organism. The complex is mainly governed by electrostatic interactions between side-chain amino groups of the reductase and carboxyl residues of flavodoxin. In order to localize the binding site on flavodoxin, chemical modification of its carboxyl groups has been carried out. Treatment of flavodoxin with a water-soluble carbodiimide, N-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC), in the presence of a nucleophile, glycine ethyl ester, caused a time-dependent modification of the protein that is responsible for the loss of its ability to participate as electron carrier in the photoreduction of NADP+ by chloroplast membranes, and also in NADPH--cytochrome-c reductase activity, by about 85%. Nevertheless, the ability of flavodoxin to receive electrons from the reducing side of photosystem I was much less affected. The inhibition was enhanced at low pH, suggesting that carboxylic acid groups were the target of chemical modification. Treated flavodoxin failed to form covalent complexes with FNR and the dissociation constant for the non-covalent complex with FNR was fourfold higher. After tryptic digestion of a sample of flavodoxin modified by EDC in the presence of [1-14C]glycine ethyl ester, two major radioactive peptides were isolated. The first protein fragment contained three carboxylic residues (Asp123, Asp126 and Asp129), corresponding to the region where long-chain flavodoxins show an insert compared to short-chain flavodoxins. The second peptide corresponded to a similar region, either in the amino acid sequence or in the three-dimensional structure of the protein and also containing three carboxyl groups (Asp144, Glu145 and Asp146). Four of these carboxyl groups (Asp123, Asp126, Asp144 and Asp146) are highly conserved in all long-chain flavodoxins, suggesting that they could play an essential role in substrate recognition.  相似文献   

16.
To investigate the charge effect of the endothelial surface glycocalyx on microvessel permeability, we extended the three-dimensional model developed by Fu et al. (J Biomech Eng 116: 502-513, 1994) for the interendothelial cleft to include a negatively charged glycocalyx layer at the entrance of the cleft. Both electrostatic and steric exclusions on charged solutes were considered within the glycocalyx layer and at the interfaces. Four charge-density profiles were assumed for the glycocalyx layer. Our model indicates that the overall solute permeability across the microvessel wall including the surface glycocalyx layer and the cleft region is independent of the charge-density profiles as long as they have the same maximum value and the same total charge. On the basis of experimental data, this model predicts that the charge density would be 25-35 meq/l in the glycolcalyx of frog mesenteric capillaries. An intriguing prediction of this model is that when the concentrations of cations and anions are unequal in the lumen due to the presence of negatively charged proteins, the negatively charged glycocalyx would provide more resistance to positively charged solutes than to negatively charged ones.  相似文献   

17.
The flavodoxins are flavin mononucleotide-containing electron transferases. Flavodoxin I has been presumed to be the only flavodoxin of Escherichia coli, and its gene, fldA, is known to belong to the soxRS (superoxide response) oxidative stress regulon. An insertion mutation of fldA was constructed and was lethal under both aerobic and anaerobic conditions; only cells that also had an intact (fldA(+)) allele could carry it. A second flavodoxin, flavodoxin II, was postulated, based on the sequence of its gene, fldB. Unlike the fldA mutant, an fldB insertion mutant is a viable prototroph in the presence or absence of oxygen. A high-copy-number fldB(+) plasmid did not complement the fldA mutation. Therefore, there must be a vital function for which FldB cannot substitute for flavodoxin I. An fldB-lacZ fusion was not induced by H(2)O(2) and is therefore not a member of the oxyR regulon. However, it displayed a soxS-dependent induction by paraquat (methyl viologen), and the fldB gene is preceded by two overlapping regions that resemble known soxS binding sites. The fldB insertion mutant did not have an increased sensitivity to the effects of paraquat on either cellular viability or the expression of a soxS-lacZ fusion. Therefore, fldB is a new member of the soxRS (superoxide response) regulon, a group of genes that is induced primarily by univalent oxidants and redox cycling compounds. However, the reactions in which flavodoxin II participates and its role during oxidative stress are unknown.  相似文献   

18.
V Massey  S Ghisla  K Yagi 《Biochemistry》1986,25(24):8103-8112
6-Thiocyanatoflavins have been found to be susceptible to nucleophilic displacement reactions with sulfite and thiols, yielding respectively the 6-S-SO3--flavin and 6-mercaptoflavin, with rate constants at pH 7.0, 20 degrees C, of 55 M-1 min-1 for sulfite and 1000 M-1 min-1 for dithiothreitol. The 6-SCN-flavin binds tightly to riboflavin-binding protein as the riboflavin derivative, to apoflavodoxin, apo-lactate oxidase, and apo-Old Yellow Enzyme as the FMN derivative, and to apo-D-amino acid oxidase as the FAD derivative. The riboflavin-binding protein derivative is inaccessible to dithiothreitol attack, and the lactate oxidase and D-amino acid oxidase derivatives show only limited accessibility. However, the flavodoxin and Old Yellow Enzyme derivatives react readily with dithiothreitol, indicating that the flavin 6-position is exposed to solvent in these proteins. The lactate oxidase and D-amino acid oxidase derivatives convert slowly but spontaneously to the 6-mercaptoflavin enzyme forms in the absence of any added thiol, indicating the presence of a thiol residue in the flavin binding site of these proteins. The reaction rates have been investigated of 6-mercaptoflavins with iodoacetamide, N-ethylmaleimide, methyl methanethiosulfonate, H2O2, and m-chloroperbenzoate, in both the free and protein-bound state. The results confirm the conclusions drawn from the studies with 6-SCN-flavins described above and from 6-N3-flavins [Massey, V., Ghisla, S., & Yagi, K. (1986) Biochemistry (preceding paper in this issue)]. The spectral properties of the protein-bound 6-mercaptoflavin vary widely among the five proteins studied and show stabilization of the neutral flavin with flavodoxin and riboflavin-binding protein and of the anionic species by Old Yellow Enzyme, lactate oxidase, and D-amino acid oxidase. In the case of the latter two enzymes, the stabilization appears to be due to interaction of the negatively charged flavin with a positively charged protein residue located near the flavin pyrimidine ring. This positively charged residue appears to be responsible also for the strong stabilization of the two-electron oxidation state of the mercaptoflavin as the 6-S-oxide. With the other flavoproteins studied this oxidation level is stabilized as the 6-sulfenic acid or 6-sulfenate.  相似文献   

19.
Inactivation of potassium channels plays an important role in shaping the electrical signaling properties of nerve and muscle cells. The rapid inactivation of Kv1.4 has been assumed to be controlled by a "ball and chain" inactivation mechanism. Besides hydrophobic interaction between inactivation ball and the channel's inner pore, the electrostatic interaction has also been proved to participate in the "ball and chain" inactivation process of Kv1.4 channel. Based on the crystal structure of Kv1.2 channel, the acidic T1-S1 linker is indicated to be a candidate interacting with the positively charged hydrophilic region of the inactivation domain. In this study, through mutating the charged residues to amino acids of opposite polar, we identified the electrostatic interaction between the inactivation ball and the T1-S1 linker region of Kv1.4 channel. Inserting negatively charged peptide at the amino terminal of Kv1.4 channel further confirmed the electrostatic interaction between the two regions.  相似文献   

20.
Inactivation of potassium channels plays an important role in shaping the electrical signaling properties of nerve and muscle cells. The rapid inactivation of Kv1.4 has been assumed to be controlled by a “ball and chain” inactivation mechanism. Besides hydrophobic interaction between inactivation ball and the channel's inner pore, the electrostatic interaction has also been proved to participate in the “ball and chain” inactivation process of Kv1.4 channel. Based on the crystal structure of Kv1.2 channel, the acidic T1–S1 linker is indicated to be a candidate interacting with the positively charged hydrophilic region of the inactivation domain. In this study, through mutating the charged residues to amino acids of opposite polar, we identified the electrostatic interaction between the inactivation ball and the T1–S1 linker region of Kv1.4 channel. Inserting negatively charged peptide at the amino terminal of Kv1.4 channel further confirmed the electrostatic interaction between the two regions.  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