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.
Based on a number of experiments it is concluded that the fluorescein labeled beta-heptapeptide fluoresceinyl-NH-CS-(S)-beta(3)hAla-(S)-beta(3)hArg-(R)-beta(3)hLeu-(S)-beta(3)hPhe-(S)-beta(3)hAla-(S)-beta(3)hAla-(S)-beta(3)hLys-OH translocates across lipid vesicle bilayers formed from DOPC (1,2-dioleoyl-sn-glycero-3-phosphocholine). The conclusion is based on the following observations: (i) addition of the peptide to the vicinity of micrometer-sized giant vesicles leads to an accumulation of the peptide inside the vesicles; (ii) if the peptide is injected inside individual giant vesicles, it is released from the vesicles in a time dependent manner; (iii) if the peptide is encapsulated within sub-micrometer-sized large unilamellar vesicles, it is released from the vesicles as a function of time; (iv) if the peptide is submitted to immobilized liposome chromatography, the peptide is retained by the immobilized DOPC vesicles. Furthermore, the addition of the peptide to calcein-containing DOPC vesicles does not lead to significant calcein leakage and vesicle fusion is not observed. The finding that derivatives of the beta-heptapeptide (S)-beta(3)hAla-(S)-beta(3)hArg-(R)-beta(3)hLeu-(S)-beta(3)hPhe-(S)-beta(3)hAla-(S)-beta(3)hAla-(S)-beta(3)hLys-OH can translocate across phospholipid bilayers is supported by independent measurements using Tb(3+)-containing large unilamellar vesicles prepared from egg phosphatidylcholine and wheat germ phosphatidylinositol (molar ratio of 9:1) and a corresponding peptide that is labeled with dipicolinic acid instead of fluorescein. The experiments show that this dipicolinic acid labeled beta-heptapeptide derivative also permeates across phospholipid bilayers. The possible mechanism of the translocation of the particular beta-heptapeptide derivatives across the membrane of phospholipid vesicles is discussed within the frame of the current understanding of the permeation of certain oligopeptides across simple phospholipid bilayers.  相似文献   

2.
We attempted to reconcile three microbial maintenance models (Herbert, Pirt, and Compromise) through a theoretical reassessment. We provided a rigorous proof that the true growth yield coefficient (Y(G) ) is the ratio of the specific maintenance rate (a in Herbert) to the maintenance coefficient (m in Pirt). Other findings from this study include: (1) the Compromise model is identical to the Herbert for computing microbial growth and substrate consumption, but it expresses the dependence of maintenance on both microbial biomass and substrate; (2) the maximum specific growth rate in the Herbert (μ(max,H) ) is higher than those in the other two models (μ(max,P) and μ(max,C) ), and the difference is the physiological maintenance factor (m(q) =?a); and (3) the overall maintenance coefficient (m(T) ) is more sensitive to m(q) than to the specific growth rate (μ(G) ) and Y(G) . Our critical reassessment of microbial maintenance provides a new approach for quantifying some important components in soil microbial ecology models.  相似文献   

3.
The essential Saccharomyces cerevisiae tRNA(His) guanylyltransferase (Thg1p) is responsible for the unusual G(-1) addition to the 5' end of cytoplasmic tRNA(His). We report here that tRNA(His) from Thg1p-depleted cells is uncharged, although histidyl tRNA synthetase is active and the 3' end of the tRNA is intact, suggesting that G(-1) is a critical determinant for aminoacylation of tRNA(His) in vivo. Thg1p depletion leads to activation of the GCN4 pathway, most, but not all, of which is Gcn2p dependent, and to the accumulation of tRNA(His) in the nucleus. Surprisingly, tRNA(His) in Thg1p-depleted cells accumulates additional m(5)C modifications, which are delayed relative to the loss of G(-1) and aminoacylation. The additional modification is likely due to tRNA m(5)C methyltransferase Trm4p. We developed a new method to map m(5)C residues in RNA and localized the additional m(5)C to positions 48 and 50. This is the first documented example of the accumulation of additional modifications in a eukaryotic tRNA species.  相似文献   

4.
Pterin-free inducible nitric oxide synthase (iNOS) was reconstituted with tetrahydrobiopterin (H(4)B) or tetrahydrobiopterin analogues (5-methyl-H(4)B and 4-amino-H(4)B), and the ability of bound 5-methyl-H(4)B and 4-amino-H(4)B to support catalysis by either full-length iNOS (FLiNOS) or the isolated heme domain (HDiNOS) was examined. In a single turnover with HDiNOS, 5-methyl-H(4)B forms a very stable radical, 5-methyl-H(3)B(*), that accumulates in the arginine reaction to approximately 60% of the HDiNOS concentration and decays approximately 400-fold more slowly than H(3)B(*) (0.0003 vs 0.12 s(-1)). The amount of radical (5-methyl-H(3)B(*) or H(3)B(*)) observed in the NHA reaction is very small (<3% of HDiNOS). The activity of 5-methyl-H(4)B-saturated FLiNOS and HDiNOS is similar to that when H(4)B is bound: arginine is hydroxylated to NHA, and NHA is oxidized exclusively to citrulline and (*)NO. A pterin radical was not observed with 4-amino-H(4)B- or pterin-free HDiNOS with either substrate. The catalytic activity of 4-amino-H(4)B-bound FLiNOS and HDiNOS resembles that of pterin-free iNOS: the hydroxylation of arginine is very unfavorable (<2% that of H(4)B-bound iNOS), and NHA is oxidized to a mixture of amino acid products (citrulline and cyanoornithine) and NO(-) rather than (*)NO. These results demonstrate that the bound pterin cofactor undergoes a one-electron oxidation (to form a pterin radical), which is essential to its ability to support normal NOS turnover. Although binding of H(4)B also stabilizes the NOS structure and active site, the most critical role of the pterin cofactor in NOS appears to be in electron transfer.  相似文献   

5.
1. Pig kidney enzyme resembles other non-specific alkaline phosphatases in its ability to hydrolyse inorganic pyrophosphate (PP(i)). 2. Studies of enzyme velocity as a function of PP(i) concentration show that Michaelis-Menten kinetics are obeyed when a constant PP(i)/Mg(2+) concentration ratio is maintained, but velocity-substrate concentration curves are sigmoid when the concentration of PP(i) is increased but that of Mg(2+) is kept constant. The enzyme is inhibited when the total PP(i) concentration is greater than the total concentration of Mg(2+). Pyrophosphatase activity is activated by Mg(2+), but if the concentration of the metal ion is increased to a value in excess of the total PP(i) concentration Mg(2+) is then strongly inhibitory. 4. It appears that the enzyme is most active towards the complex ion MgPP(i) (2-). The enzyme probably hydrolyses PP(i) (4-) also, but this is a poorer substrate and its competition with MgPP(i) (2-) leads to inhibition. At high Mg(2+) concentrations Mg(2)PP(i) is formed. This complex appears to be a potent inhibitor. 5. Sigmoid plots of v against s and of v against i result from interactions occurring between Mg(2+) and PP(i) (4-) leading to MgPP(i) (2-) and Mg(2)PP(i), and are not indicative of allosteric behaviour.  相似文献   

6.
The serine and cysteine proteases SspA and SspB of Staphylococcus aureus are secreted as inactive zymogens, zSspA and zSspB. Mature SspA is a trypsin-like glutamyl endopeptidase and is required to activate zSspB. Although a metalloprotease Aureolysin (Aur) is in turn thought to contribute to activation of zSspA, a specific role has not been demonstrated. We found that pre-zSspA is processed by signal peptidase at ANA(29) downward arrow, releasing a Leu(30) isoform that is first processed exclusively through autocatalytic intramolecular cleavage within a glutamine-rich propeptide segment, (40)QQTQSSKQQTPKIQ(53). The preferred site is Gln(43) with secondary processing at Gln(47) and Gln(53). This initial processing is necessary for optimal and subsequent Aur-dependent processing at Leu(58) and then Val(69) to release mature SspA. Although processing by Aur is rate-limiting in zSspA activation, the first active molecules of Val(69)SspA promote rapid intermolecular processing of remaining zSspA at Glu(65), producing an N-terminal (66)HANVILP isoform that is inactive until removal of the HAN tripeptide by Aur. Modeling indicated that His(66) of this penultimate isoform blocks the active site by hydrogen bonding to Ser(237) and occlusion of substrate. Binding of glutamate within the active site of zSspA is energetically unfavorable, but glutamine fits into the primary specificity pocket and is predicted to hydrogen bond to Thr(232) proximal to Ser(237), permitting autocatalytic cleavage of the glutamine-rich propeptide segment. These and other observations suggest that zSspA is activated through a trypsinogen-like mechanism where supplementary features of the propeptide must be sequentially processed in the correct order to allow efficient activation.  相似文献   

7.
Compound 48/80 (C48/80) is a synthetic condensation product of N-methyl-p-methoxyphenethylamine with formaldehyde and is an experimental drug used since the 1950s to induce anaphylactic shock through histamine release. This study was carried out to further elucidate the mechanism by which this drug induces nitric oxide (NO) release. Our specific goals were: (a) to verify if C48/80's relaxation occurs through the stimulation of histamine receptors; (b) to evaluate the endothelium-dependent relaxation induced by C48/80; (c) to identify NO as the endothelium-relaxing factor released by C48/80; (d) to identify the NO synthase (NOS) responsible for NO release; and (e) to verify if the relaxation induced by C48/80 is calcium and cyclic guanidine monophosphate (cGMP) dependent. Rabbit aorta segments, with and without endothelium, were suspended in organ chambers (25ml) filled with Krebs solution maintained at 37 degrees C, bubbled with 95% O(2)/5% CO(2) (pH 7.4). Phenylephrine was used to contract the segments. Other protocol drugs included H(1)- and H(2)-receptor antagonists, cyclooxygenase, NOS, guanylyl cyclase and phospholipase C (PLC) inhibitors. Endothelium-dependent relaxation induced by C48/80 was also studied in calcium-free Krebs solution associated with a calcium chelator. In summary, our investigation demonstrated that the C48/80 vasodilating action: (a) does not depend on H(1) and H(2) histamine receptors; (b) is NO endothelium-dependent; (c) is dependent on the endothelial constitutive NOS (NOS-3) isoform activation; (d) is cGMP-dependent; and that NOS-3 activation by C48/80: (a) is independent of PLC up to 25mug/ml and (b) is partially dependent of this lipase in higher doses.  相似文献   

8.
B Malfoy  B Hartmann    M Leng 《Nucleic acids research》1981,9(21):5659-5669
Poly(dG-dC) . poly(dG-dC) was modified by chlorodiethylenetriamino platinum (II) chloride, cis-dichlorodiammine platinum (II) and trans-dichlorodiammine platinum (II), respectively. The conformation of these modified poly(dG-dC) . poly(dG-dC) was studied by circular dichroism. In 4 M Na+, the circular dichroism spectra of poly(dG-dC)dien-Pt (0 less than or equal to rb less than or equal to 0.2) are similar (rb is the amount of bound platinum per base). It is concluded that the conformation of these polymers belongs to the Z-family. Dien-Pt complexes stabilize the Z-form. The midpoint of the Z goes to B transition of poly(dG-dC)dien-Pt(0.12) is at 0.2 M NaCl. Moreover another B goes to Z transition is observed at lower salt concentration (midpoint at 6 mM NaCl). In 1 mM phosphate buffer, the stability of Z-poly(dG-dC)dien-Pt(0.12) is greatly affected by the presence of small amounts of EDTA. Poly(dG-dC) . poly(dG-dC) modified by cis-Pt and trans-Pt complexes do not adopt the Z-form even in high salt concentration.  相似文献   

9.
Angiotensin (Ang) I-converting enzyme (ACE) is a Zn(2+) metalloprotease with two homologous catalytic domains. Both the N- and C-terminal domains are peptidyl dipeptidases. Hydrolysis by ACE of its decapeptide substrate Ang I is increased by Cl(-), but the molecular mechanism of this regulation is unclear. A search for single substitutions to Gln among all conserved basic residues (Lys/Arg) in human ACE C-domain identified R1098Q as the sole mutant that lacked Cl(-) dependence. Cl(-) dependence is also lost when the equivalent Arg in the N-domain, Arg(500), is substituted with Gln. The Arg(1098) to Lys substitution reduced Cl(-) binding affinity by approximately 100-fold. In the absence of Cl(-), substrate binding affinity (1/K(m)) of and catalytic efficiency (k(cat)/K(m)) for Ang I hydrolysis are increased 6.9- and 32-fold, respectively, by the Arg(1098) to Gln substitution, and are similar (<2-fold difference) to the respective wild-type C-domain catalytic constants in the presence of optimal [Cl(-)]. The Arg(1098) to Gln substitution also eliminates Cl(-) dependence for hydrolysis of tetrapeptide substrates, but activity toward these substrates is similar to that of the wild-type C-domain in the absence of Cl(-). These findings indicate that: 1) Arg(1098) is a critical residue of the C-domain Cl(-)-binding site and 2) a basic side chain is necessary for Cl(-) dependence. For tetrapeptide substrates, the inability of R1098Q to recreate the high affinity state generated by the Cl(-)-C-domain interaction suggests that substrate interactions with the enzyme-bound Cl(-) are much more important for the hydrolysis of short substrates than for Ang I. Since Cl(-) concentrations are saturating under physiological conditions and Arg(1098) is not critical for Ang I hydrolysis, we speculate that the evolutionary pressure for the maintenance of the Cl(-)-binding site is its ability to allow cleavage of short cognate peptide substrates at high catalytic efficiencies.  相似文献   

10.
Multidrug resistance protein 1 (MRP1) is capable of actively transporting a wide range of conjugated and unconjugated organic anions. The protein can also transport additional conjugated and unconjugated compounds in a GSH- or S-methyl GSH-stimulated manner. How MRP1 binds and transports such structurally diverse substrates is not known. We have used [(3)H]leukotriene C(4) (LTC(4)), a high affinity glutathione-conjugated physiological substrate, to photolabel intact MRP1, as well as fragments of the protein expressed in insect cells. These studies revealed that: (i) LTC(4) labels sites in the NH(2)- and COOH-proximal halves of MRP1, (ii) labeling of the NH(2)-half of MRP1 is localized to a region encompassing membrane-spanning domain (MSD) 2 and nucleotide binding domain (NBD) 1, (iii) labeling of this region is dependent on the presence of all or part of the cytoplasmic loop (CL3) linking MSD1 and MSD2, but not on the presence of MSD1, (iv) labeling of the NH(2)-proximal site is preferentially inhibited by S-methyl GSH, (v) labeling of the COOH-proximal half of the protein occurs in a region encompassing transmembrane helices 14-17 and appears not to require NBD2 or the cytoplasmic COOH-terminal region of the protein, (vi) labeling of intact MRP1 by LTC(4) is strongly attenuated in the presence of ATP and vanadate, and this decrease in labeling is attributable to a marked reduction in LTC(4) binding to the NH(2)-proximal site, and (vii) the attenuation of LTC(4) binding to the NH(2)-proximal site is a consequence of ATP hydrolysis and trapping of Vi-ADP exclusively at NBD2. These data suggest that MRP1-mediated transport involves a conformational change, driven by ATP hydrolysis at NBD2, that alters the affinity with which LTC(4) binds to one of two sites composed, at least in part, of elements in the NH(2)-proximal half of the protein.  相似文献   

11.
The standard redox potential of acrylyl-CoA/propionyl-CoA couple (C(3)) was determined to be 69 mV (vs. standard hydrogen electrode) at pH 7 and 25 degrees C. This value implies that the 2, 3-dehydrogenation of propionyl-CoA is thermodynamically much more unfavorable than that of longer acyl-CoAs because the standard redox potentials of crotonyl-CoA/butyryl-CoA (C(4)), octenoyl-CoA/octanoyl-CoA (C(8)), and hexadecenoyl-CoA/palmitoyl-CoA (C(16)) are all about -10 mV. The unusually high standard redox potential of the acrylyl-CoA/propionyl-CoA couple is thought to be one of the reasons that in mammals propionyl-CoA is not metabolized by beta-oxidation as in the case of longer acyl-CoAs, but by a methylmalonyl-CoA pathway. The obvious structural difference between C(3) and C(4) (and longer) is whether an H or the C(4) atom is connected to -C(3)H=C(2)H-C(1)O-S-CoA. The molecular orbital calculations (MOPAC) for the enoyl and acyl forms of C(3) and C(4) revealed that this structural feature is the main cause for the higher standard redox potential of the C(3) couple. That is, the C(4)-C(3) bond is stabilized by the dehydrogenation to a greater degree than the H-C(3) bond.  相似文献   

12.
13.
The interaction of thiourea (Tu) with auranofin (Et(3)PAuSATg) and its analogue, Et(3)PAuCl has been studied using (13)C, (31)P and (15)N NMR spectroscopy. It is observed that Tu is able to replace both the ligands, Et(3)P and SATg(-) simultaneously from gold(I) in auranofin, forming [Et(3)P-Au-Tu](+) and Tu-Au-SATg complexes. However, no separate resonances for these species were observed either due to their rapid exchange with auranofin and thus giving only the average resonances or because the chemical shifts of either two species are same so that they cannot be resolved. The displaced SATg(-) is oxidized to its disulfide, (SATg)(2). However, some of the displaced Et(3)P is oxidized to Et(3)PO while the remaining reacts with Tu to form Et(3)P-Tu species, characterized by delta 31P of 1.0 ppm, assigned after an independent reaction between Et(3)P and Tu. In an experiment using a 0.05 M solution of auranofin, the Et(3)PO resonance appeared in auranofin spectrum after 4 days of addition of 1.0 equivalent of Tu, showing that the reaction is slow. A resonance for free Et(3)P is also detected in 31P NMR on the addition of CN(-). It is also observed that Tu reacts with Et(3)PAuCl to form [Et(3)P-Au-Tu](+) via displacement of Cl(-), consistent with an upfield shift of 6.2 ppm in >C [double bond] S resonance of Tu in (13)C NMR. In (15)N NMR, a smaller downfield, instead of an upfield shift, in NH(2) resonance of Tu on its addition to auranofin and Et(3)PAuCl indicates that it is not binding to gold(I) through nitrogen.  相似文献   

14.
Yu CA  Cen X  Ma HW  Yin Y  Yu L  Esser L  Xia D 《Biochimica et biophysica acta》2008,1777(7-8):1038-1043
Intensive biochemical, biophysical and structural studies of the cytochrome (cyt) bc(1) complex in the past have led to the formulation of the "protonmotive Q-cycle" mechanism for electron and proton transfer in this vitally important complex. The key step of this mechanism is the separation of electrons during the oxidation of a substrate quinol at the Q(P) site with both electrons transferred simultaneously to ISP and cyt b(L) when the extrinsic domain of ISP (ISP-ED) is located at the b-position. Pre-steady state fast kinetic analysis of bc(1) demonstrates that the reduced ISP-ED moves to the c(1)-position to reduce cyt c(1) only after the reduced cyt b(L) is oxidized by cyt b(H). However, the question of how the conformational switch of ISP-ED is initiated remains unanswered. The results obtained from analysis of inhibitory efficacy and binding affinity of two types of Q(P) site inhibitors, Pm and Pf, under various redox states of the bc(1) complex, suggest that the electron transfer from heme b(L) to b(H) is the driving force for the releasing of the reduced ISP-ED from the b-position to c(1)-position to reduce cyt c(1).  相似文献   

15.
The antioxidant functions of cytochrome c   总被引:6,自引:0,他引:6  
Low (C(1/2) = 1.5 x 10(-7) M) concentrations of horse cytochrome c strongly inhibit H(2)O(2) production by rat heart mitochondria under conditions of reverse electron transfer from succinate to NAD(+). The effect is abolished by binding of cytochrome c with liposomes and is not prevented by SOD. Yeast cytochrome c is much less effective than the horse protein whereas acetylated horse cytochrome c is without effect. H(2)O(2) formation stimulated by antimycin A is resistant to added cytochrome c. In inside-out submitochondrial vesicles, H(2)O(2) production is suppressed by all three cytochrome c samples tested, but at higher concentrations (C(1/2) is about 5 x 10(-7) M). In vesicles, SOD abolishes the cytochrome c inhibition. We conclude that extramitochondrial cytochrome c is competent in down-regulation of the Complex I H(2)O(2) production linked to the reverse electron transfer. Such an effect is absent in the inside-out submitochondrial vesicles where another antioxidant cytochrome c function can be observed, i.e. the oxidation of O(2-*) to O(2). A possible role of cytochrome c in the antioxidant defence is discussed.  相似文献   

16.
Duodenal alkaline phosphatase of juvenile (11-day-old) mice, like other non-specific alkaline phosphatases, has the ability to hydrolyse PP(i). When a constant Mg(2+)/PP(i) concentration ratio is maintained, plots of velocity as a function of PP(i) concentration are consistent with Michaelis-Menten kinetics. Mg(2+) activates pyrophosphate hydrolysis and maximal activity is obtained at a constant Mg(2+)/PP(i) concentration ratio of 0.66. At higher ratios there is strong inhibition. At constant concentrations of Mg(2+) and increasing concentrations of PP(i), the velocity-substrate (PP(i)) concentration plots show sigmoidal dependence. By assuming that the true substrate is MgP(2)O(7) (2-) complex, and using complexity constants, the concentrations of free Mg(2+), Mg(2)P(2)O(7) and MgP(2)O(7) (2-) were calculated in assay mixtures ranging in PP(i) concentration from 0.1 to 2.5mm and in total Mg(2+) concentration from 0.6 to 2.6mm. From these data, the concentrations of added Mg(2+) and PP(i) in the assay mixtures were selected so that the velocity could be measured (1) at three fixed concentrations of free Mg(2+) ions with varied concentrations of MgP(2)O(7) (2-) and (2) at four fixed concentrations of Mg(2)P(2)O(7) with varied concentrations of MgP(2)O(7) (2-). Lineweaver-Burk and Hill plots from these data showed that the inhibition is caused by free Mg(2+) ions, of a mixed type and consistent with Michaelis-Menten kinetics. The sigmoidal dependence observed between velocity and PP(i) concentration at constant concentration of total Mg(2+) is therefore not due to allosteric inhibition. It is due to a combined effect of (1) inhibition by free Mg(2+) ions, (2) depletion of the true substrate, MgP(2)O(7) (2-), owing to the formation of Mg(2)P(2)O(7) and (3) the manner in which the concentrations of these three molecular or ionic species change when PP(i) concentration is increased maintaining the total Mg(2+) concentration constant.  相似文献   

17.
Louie TM  Xie XS  Xun L 《Biochemistry》2003,42(24):7509-7517
4-Hydroxyphenylacetate (4HPA) 3-monooxygenase (HpaB) is a reduced flavin adenine dinucleotide (FADH(2)) utilizing monooxygenase. Its cosubstrate, FADH(2), is supplied by HpaC, an NAD(P)H-flavin oxidoreductase. Because HpaB is the first enzyme for 4HPA metabolism, FADH(2) production and utilization become a major metabolic event when Escherichia coli W grows on 4HPA. An important question is how FADH(2) is produced and used, as FADH(2) is unstable in the presence of free O(2). One solution is metabolic channeling by forming a transitory HpaB-HpaC complex. However, our in vivo and in vitro data failed to support the interaction. Further investigation pointed to an alternative scheme for HpaB to sequester FADH(2). The intracellular HpaB concentration was about 122 microM in 4HPA-growing cells, much higher than the total intracellular FAD concentration, and HpaB had a high affinity for FADH(2) (K(d) of 70 nM), suggesting that most FADH(2) is bound to HpaB in vivo. The HpaB-bound FADH(2) was either used to rapidly oxidize 4HPA or slowly oxidized by O(2) to FAD and H(2)O(2) in the absence of 4HPA. Thus, HpaB's high intracellular concentration, its high affinity for FADH(2), its property of protecting bound FADH(2) in the absence of 4HPA, and its ability to rapidly use FADH(2) to oxidize 4HPA when 4HPA is available can coordinate FADH(2) production and utilization by HpaB and HpaC in vivo. This type of coordination, in responding to demand, for production and utilization of labile metabolites has not been reported to date.  相似文献   

18.
A series of heterobinuclear cofacial porphyrin-corrole dyads containing a Co(IV) corrole linked by one of four different spacers in a face-to-face arrangement with an Fe(III) or Mn(III) porphyrin have been examined as catalysts for the electroreduction of O(2) to H(2)O and/or H(2)O(2) when adsorbed on the surface of a graphite electrode in air-saturated aqueous solutions containing 1M HClO(4). The examined compounds are represented as (PCY)M(III)ClCo(IV)Cl where P is a porphyrin dianion, C is a corrole trianion and Y is a biphenylene (B), 9,9-dimethylxanthene (X), dibenzofuran (O) or anthracene (A) spacer. The catalytic behavior of the seven investigated dyads in the two heterobimetallic (PCY)MClCoCl series of catalysts is compared on one hand to what was previously reported for related dyads with a single Co(III) corrole macrocycle linked to a free-base porphyrin with the same set of linking bridges, (PCY)H(2)Co, and on the other hand to dicobalt porphyrin-corrole dyads of the form (PCY)Co(2) which were shown to efficiently electrocatalyze the four electron reduction of O(2) at a graphite electrode in acid media. Comparisons between the four series of porphyrin-corrole dyads, (PCY)Co(2), (PCY)H(2)Co, (PCY)FeClCoCl and (PCY)MnClCoCl, show that in all cases the biscobalt dyads catalyze O(2) electroreduction at potentials more positive by an average 110mV as compared to the related series of compounds containing a Co(III) or Co(IV) corrole macrocycle linked to a free-base metalloporphyrin or a metalloporphyrin with an Fe(III) or Mn(III) central metal ion. The data indicates that the E(1/2) values where electrocatalysis is initiated is related to the initial site of electron transfer, which is the Co(III)/Co(II) porphyrin reduction process in the case of (PCY)Co(2) and the Co(IV)/Co(III) corrole reduction in the case of (PCY)MnClCoCl, (PCY)FeClCoCl and (PCY)H(2)Co. The overall data also suggests that the catalytically active form of the biscobalt dyad in (PCY)Co(2) contains a Co(II) porphyrin and a Co(IV) corrole.  相似文献   

19.
20.
Glucose oxidase (GO) and copper amine oxidase (CAO) catalyze the reduction of molecular oxygen to hydrogen peroxide. If a closed-shell cofactor (like FADH(2) in GO and topaquinone (TPQ) in CAO) is electron donor in dioxygen reduction, the formation of a closed-shell species (H(2)O(2)) is a spin forbidden process. Both in GO and CAO, formation of a superoxide ion that leads to the creation of a radical pair is experimentally suggested to be the rate-limiting step in the dioxygen reduction process. The present density functional theory (DFT) studies suggest that in GO, the creation of the radical pair induces a spin transition by spin orbit coupling (SOC) in O(2)(-)(rad), whereas in CAO, it is induced by exchange interaction with the paramagnetic metal ion (Cu(II)). In the rate-limiting step, this spin-transition is suggested to transform the O(2)(-)(rad)-FADH(2)(+)(rad) radical pair in GO and the Cu(II)-TPQ (triplet) species in CAO, from a triplet (T) to a singlet (S) state. For CAO, a mechanism for the O[bond]O cleavage step in the biogenesis of TPQ is also suggested.  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