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.
The alpha-ketoglutate (alpha-KG)-dependent dioxygenases are a large class of mononuclear non-heme iron enzymes that require Fe(II), alpha-KG and dioxygen for catalysis, with the alpha-KG cosubstrate supplying the two additional electrons required for dioxygen activation. A sub-class of these enzymes exists in which the alpha-keto acid is covalently attached to the substrate, including (4-hydroxy)mandelate synthase (HmaS) and (4-hydroxyphenyl)pyruvate dioxygenase (HPPD) which utilize the same substrate but exhibit two different general reactivities (H-atom abstraction and electrophilic attack). Previous kinetic studies of Streptomyces avermitilis HPPD have shown that the substrate analog phenylpyruvate (PPA), which only differs from the normal substrate (4-hydroxyphenyl)pyruvate (HPP) by the absence of a para-hydroxyl group on the aromatic ring, does not induce a reaction with dioxygen. While an Fe(IV)O intermediate is proposed to be the reactive species in converting substrate to product, the key step utilizing O(2) to generate this species is the decarboxylation of the alpha-keto acid. It has been generally proposed that the two requirements for decarboxylation are bidentate coordination of the alpha-keto acid to Fe(II) and the presence of a 5C Fe(II) site for the O(2) reaction. Circular dichroism and magnetic circular dichroism studies have been performed and indicate that both enzyme complexes with PPA are similar with bidentate alpha-KG coordination and a 5C Fe(II) site. However, kinetic studies indicate that while HmaS reacts with PPA in a coupled reaction similar to the reaction with HPP, HPPD reacts with PPA in an uncoupled reaction at an approximately 10(5)-fold decreased rate compared to the reaction with HPP. A key difference is spectroscopically observed in the n-->pi( *) transition of the HPPD/Fe(II)/PPA complex which, based upon correlation to density functional theory calculations, is suggested to result from H-bonding between a nearby residue and the carboxylate group of the alpha-keto acid. Such an interaction would disfavor the decarboxylation reaction by stabilizing electron density on the carboxylate group such that the oxidative cleavage to yield CO(2) is disfavored.  相似文献   

2.
Glu-69 belongs to a proposed active-site consensus motif His62-X-His64-X4-Glu69 (where X is any amino acid) that acetylacetone dioxygenase Dke1 from Acinetobacter johnsonii shares with structurally related non-heme metal enzymes of the cupin protein superfamily. We report functional consequences of the site-directed replacement Glu-69 → Gln based on a detailed biochemical and kinetic characterization of the purified Dke1 mutant. Perturbations of the free energy profile of the wild-type caused by the mutation were surprisingly small, with key points of the reaction pathway such as β-diketone substrate binding, the rate-limiting reduction of dioxygen, and CC bond cleavage essentially left unaltered. Release of Fe2+ from the mutant active site occurred at twice the wild-type rate, and the thermal stability of β-sheet secondary structure in Fe2+-depleted apo-proteins was lower in the mutant. The substitution Glu-69 → Gln is thus remarkably silent regarding Dke1 function. These results do not support a unified catalytic or metal-coordinating role of Glu-69 (and its positional homologues) in O2-dependent cupin-fold enzymes.  相似文献   

3.
Hydroxymandelate synthase (HMS) catalyzes the committed step in the formation of para-hydroxyphenylglycine, a recurrent substructure of polycyclic non-ribosomal peptide antibiotics such as vancomycin. HMS uses the same substrates as 4-hydroxyphenylpyruvate dioxygenase (HPPD), 4-hydroxyphenylpyruvate (HPP) and O2, and also conducts a dioxygenation reaction. The difference between the two lies in the insertion of the second oxygen atom, HMS directing this atom onto the benzylic carbon of the substrate while HPPD hydroxylates the aromatic C1 carbon. We have shown that HMS will bind NTBC, a herbicide/therapeutic whose mode of action is based on the inhibition of HPPD. This occurs despite residue differences at the active site of HMS from those known to contact the inhibitor in HPPD. Moreover, the minimal kinetic mechanism for association of NTBC to HMS differs only slightly from that observed with HPPD. The primary difference is that three charge-transfer species are observed to accumulate during association. The first reversible complex forms with a weak dissociation constant of 520 μM, the subsequent two charge-transfer complexes form with rate constants of 2.7 s−1 and 0.67 s−1. As was the case for HPPD, the final complex has the most intense charge-transfer, is not observed to dissociate, and is unreactive towards dioxygen.  相似文献   

4.
(4-hydroxyphenyl)pyruvate dioxygenase (HPPD) catalyzes the second step in the pathway for the catabolism of tyrosine, the conversion of (4-hydroxyphenyl)pyruvate (HPP) to homogentisate (HG). This reaction involves decarboxylation, substituent migration, and aromatic oxygenation. HPPD is a member of the alpha-keto acid dependent oxygenases that require Fe(II) and an alpha-keto acid substrate to oxygenate an organic molecule. We have examined the binding of ligands to HPPD from Streptomyces avermitilis. Our data show that HPP binds to the apoenzyme and that the apo-HPPD.HPP complex does not bind Fe(II) to generate active holoenzyme. The binding of HPP, phenylpyruvate (PPA), and pyruvate to the holoenzyme produces a weak ligand charge-transfer band at approximately 500 nm that is indicative of bidentate binding of the 1-carboxylate and 2-keto pyruvate oxygen atoms to the active site metal ion. For HPPD from this organism the 4-hydroxyl group of (4-hydroxyphenyl)pyruvate is a requirement for catalysis; no turnover is observed in the presence of phenylpyruvate. The rate constant for the dissociation of Fe(II) from the holoenzyme is 0.0006 s(-)(1) and indicates that this phenomenon is not significantly relevant in steady-state turnover. The addition of HPP and molecular oxygen to the holoenzyme is formally random. The basis of the ordered bi bi steady-state kinetic mechanism previously observed by Rundgren (Rundgren, M. (1977) J. Biol. Chem. 252, 5094-9) is the 3600-fold increase in oxygen reactivity when holo-HPPD is in complex with HPP. This complex reacts with molecular oxygen with a second-order rate constant of 1.4 x 10(5) M(-)(1) s(-)(1) inducing the formation of an intermediate that decays at the catalytically relevant rate of 7.8 s(-)(1).  相似文献   

5.
Hydrogen bonding networks proximal to metal centers are emerging as a viable means for controlling secondary coordination spheres. This has led to the regulation of reactivity and isolation of complexes with new structural motifs. We have used the tridenate ligand bis[(N′-tert-butylureido)-N-ethyl]-N-methylaminato ([H21]2−) that contains two hydrogen bond donors to examine the oxidation of the FeII-acetate complex, [FeIIH212-OAc)] with dioxygen, amine N-oxides, and xylyl azide. A complex with FeIII-O-FeIII core results from the oxidation with dioxygen and amine N-oxides, in which the oxo ligand is involved in hydrogen bonding to the [H21]2− ligand. A distinctly different hydrogen bonding network was found in FeIII dimer isolated from the reaction with the xylyl azide: a rare FeIII-N(R)-FeIII core was observed that does not have hydrogen bonds to the bridging nitrogen atom. The intramolecular H-bond networks within these dimers appear to adjust to the presence of the bridging species and rearrange to its size and electron density.  相似文献   

6.
In this work we report on the synthesis, crystal structure, and physicochemical characterization of the novel dinuclear [FeIIICdII(L)(μ-OAc)2]ClO4·0.5H2O (1) complex containing the unsymmetrical ligand H2L = 2-bis[{(2-pyridyl-methyl)-aminomethyl}-6-{(2-hydroxy-benzyl)-(2-pyridyl-methyl)}-aminomethyl]-4-methylphenol. Also, with this ligand, the tetranuclear [Fe2IIIHg2II(L)2(OH)2](ClO4)2·2CH3OH (2) and [FeIIIHgII(L)(μ-CO3)FeIIIHgII(L)](ClO4)2·H2O (3) complexes were synthesized and fully characterized. It is demonstrated that the precursor [FeIII2HgII2(L)2(OH)2](ClO4)2·2CH3OH (2) can be converted to (3) by the fixation of atmospheric CO2 since the crystal structure of the tetranuclear organometallic complex [FeIIIHgII(L)(μ-CO3)FeIIIHgII(L)](ClO4)2·H2O (3) with an unprecedented {FeIII(μ-Ophenoxo)2(μ-CO3)FeIII} core was obtained through X-ray crystallography. In the reaction 2 → 3 a nucleophilic attack of a FeIII-bound hydroxo group on the CO2 molecule is proposed. In addition, it is also demonstrated that complex (3) can regenerate complex (2) in aqueous/MeOH/NaOH solution. Magnetochemical studies reveal that the FeIII centers in 3 are antiferromagnetically coupled (J = − 7.2 cm− 1) and that the FeIII-OR-FeIII angle has no noticeable influence in the exchange coupling. Phosphatase-like activity studies in the hydrolysis of the model substrate bis(2,4-dinitrophenyl) phosphate (2,4-bdnpp) by 1 and 2 show Michaelis-Menten behavior with 1 being ~ 2.5 times more active than 2. In combination with kH/kD isotope effects, the kinetic studies suggest a mechanism in which a terminal FeIII-bound hydroxide is the hydrolysis-initiating nucleophilic catalyst for 1 and 2. Based on the crystal structures of 1 and 3, it is assumed that the relatively long FeIII…HgII distance could be responsible for the lower catalytic effectiveness of 2.  相似文献   

7.
Two biomimetic iron(II) benzoylformate complexes, [LFeII(BF)2] (2) and [LFeII(NO3)(BF)] (3) (L is 2,9-dimethyl-1,10-phenanthroline and BF is monoanionic benzoylformate), have been synthesized from an iron(II)–dichloro complex [LFeIICl2] (1). All the iron(II) complexes have been structurally and spectroscopically characterized. The iron(II) center in 2 is coordinated by a bidentate NN ligand (2,9-dimethyl-1,10-phenanthroline) and two monoanionic benzoylformates to form a distorted octahedral coordination geometry. One of the benzoylformates binds to the iron in 2 via both carboxylate oxygens but the other one binds in a chelating bidentate fashion via one carboxylate oxygen and the keto oxygen. On the other hand, the iron(II) center in 3 is ligated by one NN ligand, one bidentate nitrate, and one monoanionic chelating benzoylformate. Both iron(II) benzoylformate complexes exhibit the facial NNO donor environment in their solid-state structures. Complexes 2 and 3 are stable in noncoordinating solvents under an inert atmosphere, but react with dioxygen under ambient conditions to undergo oxidative decarboxylation of benzoylformate to benzoate in high yields. Evidence for the formation of an iron(IV)–oxo intermediate upon oxidative decarboxylation of benzoylformate was obtained by interception and labeling experiments. The iron(II) benzoylformate complexes represent the functional models of α-keto acid dependent oxygenases.  相似文献   

8.
Although a large number of key odorants of Swiss-type cheese result from amino acid catabolism, the amino acid catabolic pathways in the bacteria present in these cheeses are not well known. In this study, we compared the in vitro abilities of Lactobacillus delbrueckii subsp. lactis, Lactobacillus helveticus, and Streptococcus thermophilus to produce aroma compounds from three amino acids, leucine, phenylalanine, and methionine, under mid-pH conditions of cheese ripening (pH 5.5), and we investigated the catabolic pathways used by these bacteria. In the three lactic acid bacterial species, amino acid catabolism was initiated by a transamination step, which requires the presence of an α-keto acid such as α-ketoglutarate (α-KG) as the amino group acceptor, and produced α-keto acids. Only S. thermophilus exhibited glutamate dehydrogenase activity, which produces α-KG from glutamate, and consequently only S. thermophilus was capable of catabolizing amino acids in the reaction medium without α-KG addition. In the presence of α-KG, lactobacilli produced much more varied aroma compounds such as acids, aldehydes, and alcohols than S. thermophilus, which mainly produced α-keto acids and a small amount of hydroxy acids and acids. L. helveticus mainly produced acids from phenylalanine and leucine, while L. delbrueckii subsp. lactis produced larger amounts of alcohols and/or aldehydes. Formation of aldehydes, alcohols, and acids from α-keto acids by L. delbrueckii subsp. lactis mainly results from the action of an α-keto acid decarboxylase, which produces aldehydes that are then oxidized or reduced to acids or alcohols. In contrast, the enzyme involved in the α-keto acid conversion to acids in L. helveticus and S. thermophilus is an α-keto acid dehydrogenase that produces acyl coenzymes A.  相似文献   

9.
The induction of hydroxyl radical (OH) production via quinone redox cycling in white-rot fungi was investigated to improve pollutant degradation. In particular, we examined the influence of 4-methoxybenzaldehyde (anisaldehyde), Mn2+, and oxalate on Pleurotus eryngii OH generation. Our standard quinone redox cycling conditions combined mycelium from laccase-producing cultures with 2,6-dimethoxy-1,4-benzoquinone (DBQ) and Fe3+-EDTA. The main reactions involved in OH production under these conditions have been shown to be (i) DBQ reduction to hydroquinone (DBQH2) by cell-bound dehydrogenase activities; (ii) DBQH2 oxidation to semiquinone (DBQ) by laccase; (iii) DBQ autoxidation, catalyzed by Fe3+-EDTA, producing superoxide (O2) and Fe2+-EDTA; (iv) O2 dismutation, generating H2O2; and (v) the Fenton reaction. Compared to standard quinone redox cycling conditions, OH production was increased 1.2- and 3.0-fold by the presence of anisaldehyde and Mn2+, respectively, and 3.1-fold by substituting Fe3+-EDTA with Fe3+-oxalate. A 6.3-fold increase was obtained by combining Mn2+ and Fe3+-oxalate. These increases were due to enhanced production of H2O2 via anisaldehyde redox cycling and O2 reduction by Mn2+. They were also caused by the acceleration of the DBQ redox cycle as a consequence of DBQH2 oxidation by both Fe3+-oxalate and the Mn3+ generated during O2 reduction. Finally, induction of OH production through quinone redox cycling enabled P. eryngii to oxidize phenol and the dye reactive black 5, obtaining a high correlation between the rates of OH production and pollutant oxidation.The degradation of lignin and pollutants by white-rot fungi is an oxidative and rather nonspecific process based on the production of substrate free radicals (36). These radicals are produced by ligninolytic enzymes, including laccase and three kinds of peroxidases: lignin peroxidase, manganese peroxidase, and versatile peroxidase (VP) (23). The H2O2 required for peroxidase activities is provided by several oxidases, such as glyoxal oxidase and aryl-alcohol oxidase (AAO) (9, 18). This free-radical-based degradative mechanism leads to the production of a broad variety of oxidized compounds. Common lignin depolymerization products are aromatic aldehydes and acids, and quinones (34). In addition to their high extracellular oxidation potential, white-rot fungi show strong ability to reduce these lignin depolymerization products, using different intracellular and membrane-bound systems (4, 25, 39). Since reduced electron acceptors of oxidized compounds are donor substrates for the above-mentioned oxidative enzymes, the simultaneous actions of both systems lead to the establishment of redox cycles (35). Although the function of these redox cycles is not fully understood, they have been hypothesized to be related to further metabolism of lignin depolymerization products that require reduction to be converted in substrates of the ligninolytic enzymes (34). A second function attributed to these redox cycles is the production of reactive oxygen species, i.e., superoxide anion radicals (O2), H2O2, and hydroxyl radicals (OH), where lignin depolymerization products and fungal metabolites act as electron carriers between intracellular reducing equivalents and extracellular oxygen. This function has been studied in Pleurotus eryngii, whose ligninolytic system is composed of laccase (26), VP (24), and AAO (9). Incubation of this fungus with different aromatic aldehydes has been shown to provide extracellular H2O2 on a constant basis, due to the establishment of a redox cycle catalyzed by an intracellular aryl-alcohol dehydrogenase (AAD) and the extracellular AAO (7, 10). The process was termed aromatic aldehyde redox cycling, and 4-methoxybenzaldehyde (anisaldehyde) serves as the main Pleurotus metabolite acting as a cycle electron carrier (13). A second cyclic system, involving a cell-bound quinone reductase activity (QR) and laccase, was found to produce O2 and H2O2 during incubation of P. eryngii with different quinones (11). The process was described as the cell-bound divalent reduction of quinones (Q) by QR, followed by extracellular laccase oxidation of hydroquinones (QH2) into semiquinones (Q), which autoxidized to some extent, producing O2 (Q + O2 ⇆ Q + O2). H2O2 was formed by O2 dismutation (O2 + HO2 + H+ → O2 + H2O2). In an accompanying paper, we describe the extension of this O2 and H2O2 generation mechanism to OH radical production by the addition of Fe3+-EDTA to incubation mixtures of several white-rot fungi with different quinones (6). Among them, those derived from 4-hydroxyphenyl, guaiacyl, and syringyl lignin units were used: 1,4-benzoquinone (BQ), 2-methoxy-1,4-benzoquinone (MBQ), and 2,6-dimethoxy-1,4-benzoquinone (DBQ), respectively. Semiquinone autoxidation under these conditions was catalyzed by Fe3+-EDTA instead of being a direct electron transfer to O2. The intermediate Fe2+-EDTA reduced not only O2, but also H2O2, leading to OH radical production by the Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+).Although OH radicals are the strongest oxidants produced by white-rot fungi (2, 14), studies of their involvement in pollutant degradation are quite scarce. In this context, the objectives of this study were to (i) determine possible factors enhancing the production of OH radicals by P. eryngii via quinone redox cycling and (ii) test the validity of this inducible OH production mechanism as a strategy for pollutant degradation. Our selection of possible OH production promoters was guided by two observations (6). First, the redox cycle of benzoquinones working with washed P. eryngii mycelium is rate limited by hydroquinone oxidation, since the amounts of the ligninolytic enzymes that remained bound to the fungus under these conditions were not large. Second, H2O2 is the limiting reagent for OH production by the Fenton reaction.With these considerations in mind, anisaldehyde and Mn2+ were selected to increase H2O2 production. As mentioned above, anisaldehyde induces H2O2 production in P. eryngii via aromatic aldehyde redox cycling (7). Mn2+ has been shown to enhance H2O2 production during the oxidation of QH2 by P. eryngii laccase by reducing the O2 produced in the semiquinone autoxidation reaction (Mn2+ + O2 → Mn3+ + H2O2 + 2 H+) (26). Mn2+ was also selected to increase the hydroquinone oxidation rate, since this reaction has been shown to be propagated by the Mn3+ generated in the latter reaction (QH2 + Mn3+ → Q + Mn2+ + 2 H+). The replacement of Fe3+-EDTA by Fe3+-oxalate was also planned in order to increase the QH2 oxidation rate above that resulting from the action of laccase. Oxalate is a common extracellular metabolite of wood-rotting fungi to which the function of chelating iron and manganese has been attributed (16, 45). The use of Fe3+-oxalate and nonchelated Fe3+, both QH2 oxidants, has been proven to enable quinone redox cycling in fungi that do not produce ligninolytic enzymes, such as the brown-rot fungus Gloeophyllum trabeum (17, 40, 41). Finally, phenol and the azo dye reactive black 5 (RB5) were selected as model pollutants.  相似文献   

10.
A simple strategy for the induction of extracellular hydroxyl radical (OH) production by white-rot fungi is presented. It involves the incubation of mycelium with quinones and Fe3+-EDTA. Succinctly, it is based on the establishment of a quinone redox cycle catalyzed by cell-bound dehydrogenase activities and the ligninolytic enzymes (laccase and peroxidases). The semiquinone intermediate produced by the ligninolytic enzymes drives OH production by a Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+). H2O2 production, Fe3+ reduction, and OH generation were initially demonstrated with two Pleurotus eryngii mycelia (one producing laccase and versatile peroxidase and the other producing just laccase) and four quinones, 1,4-benzoquinone (BQ), 2-methoxy-1,4-benzoquinone (MBQ), 2,6-dimethoxy-1,4-benzoquinone (DBQ), and 2-methyl-1,4-naphthoquinone (menadione [MD]). In all cases, OH radicals were linearly produced, with the highest rate obtained with MD, followed by DBQ, MBQ, and BQ. These rates correlated with both H2O2 levels and Fe3+ reduction rates observed with the four quinones. Between the two P. eryngii mycelia used, the best results were obtained with the one producing only laccase, showing higher OH production rates with added purified enzyme. The strategy was then validated in Bjerkandera adusta, Phanerochaete chrysosporium, Phlebia radiata, Pycnoporus cinnabarinus, and Trametes versicolor, also showing good correlation between OH production rates and the kinds and levels of the ligninolytic enzymes expressed by these fungi. We propose this strategy as a useful tool to study the effects of OH radicals on lignin and organopollutant degradation, as well as to improve the bioremediation potential of white-rot fungi.White-rot fungi are unique in their ability to degrade a wide variety of organopollutants (36, 47), mainly due to the secretion of a low-specificity enzyme system whose natural function is the degradation of lignin (11). Components of this system include laccase and/or one or two types of peroxidase, such as lignin peroxidase (LiP), manganese peroxidase (MnP), and versatile peroxidase (VP) (31). Besides acting directly, the ligninolytic enzymes can bring about lignin and pollutant degradation through the generation of low-molecular-weight extracellular oxidants, including (i) Mn3+, (ii) free radicals from some fungal metabolites and lignin depolymerization products (7, 22), and (iii) oxygen free radicals, mainly hydroxyl radicals (OH) and lipid peroxidation radicals (21). Although OH radicals are the strongest oxidants found in cultures of white-rot fungi (1), studies of their involvement in pollutant degradation are scarce. One of the reasons is that the mechanisms proposed for OH production still await in vivo validation.Several potential sources of extracellular OH based on the Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+) have been postulated for white-rot fungi. In one case, an extracellular fungal glycopeptide has been shown to reduce O2 and Fe3+ to H2O2 and Fe2+ (45). Enzymatic sources include cellobiose dehydrogenase, LiP, and laccase. Among these, only cellobiose dehydrogenase is able to directly catalyze the formation of Fenton''s reagent (33). The ligninolytic enzymes, however, act as an indirect source of OH through the generation of Fe3+ and O2 reductants, such as formate (CO2) and semiquinone (Q) radicals. The first time evidence was provided that a ligninolytic enzyme was involved in OH production, oxalate was used to generate CO2 in a LiP reaction mediated by veratryl alcohol (4). The proposed mechanism consisted of the following cascade of reactions: production of veratryl alcohol cation radical (Valc+) by LiP, oxidation of oxalate to CO2 by Valc+, reduction of O2 to O2 by CO2, and a superoxide-driven Fenton reaction (Haber-Weiss reaction) in which Fe3+ was reduced by O2. The OH production mechanism assisted by Q was inferred from the oxidation of 2-methoxy-1,4-benzohydroquinone (MBQH2) and 2,6-dimethoxy-1,4-benzohydroquinone (DBQH2) by Pleurotus eryngii laccase in the presence of Fe3+-EDTA. The ability of Q radicals to reduce both Fe3+ to Fe2+ and O2 to O2, which dismutated to H2O2, was demonstrated (14). In this case, OH radicals were generated by a semiquinone-driven Fenton reaction, as Q radicals were the main agents accomplishing Fe3+ reduction. The first evidence of the likelihood of this OH production mechanism being operative in vivo had been obtained from incubations of P. eryngii with 2-methyl-1,4-naphthoquinone (menadione [MD]) and Fe3+-EDTA (15). Extracellular OH radicals were produced on a constant basis through quinone redox cycling, consisting of the reduction of MD by a cell-bound quinone reductase (QR) system, followed by the extracellular oxidation of the resulting hydroquinone (MDH2) to its semiquinone radical (MD). The production of extracellular O2 and H2O2 by P. eryngii via redox cycling involving laccase was subsequently confirmed using 1,4-benzoquinone (BQ), 2-methyl-1,4-benzoquinone, and 2,3,5,6-tetramethyl-1,4-benzoquinone (duroquinone), in addition to MD (16). However, the demonstration of OH production based on the redox cycling of quinones other than MD was still required.In the present paper, we describe the induction of extracellular OH production by P. eryngii upon its incubation with BQ, 2-methoxy-1,4-benzoquinone (MBQ), 2,6-dimethoxy-1,4-benzoquinone (DBQ), and MD in the presence of Fe3+-EDTA. The three benzoquinones were selected because they are oxidation products of p-hydroxyphenyl, guaiacyl, and syringyl units of lignin (MD was included as a positive control). Along with laccase, the involvement of P. eryngii VP in the production of O2 and H2O2 from hydroquinone oxidation has also been reported (13). Since hydroquinones are substrates of all known ligninolytic enzymes, quinone redox cycling catalysis could involve any of them. Here, we demonstrate OH production by P. eryngii under two different culture conditions, leading to the production of laccase or laccase and VP. We also show that quinone redox cycling is widespread among white-rot fungi by using a series of well-studied species that produce different combinations of ligninolytic enzymes.  相似文献   

11.
A novel heterobinuclear mixed valence complex [FeIIICuII(BPBPMP)(OAc)2]ClO4, 1, with the unsymmetrical N5O2 donor ligand 2-bis[{(2-pyridylmethyl)aminomethyl}-6-{(2-hydroxybenzyl)(2-pyridylmethyl)}aminomethyl]-4-methylphenol (H2BPBPMP) has been synthesized and characterized. A combination of data from mass spectrometry, potentiometric titrations, X-ray absorption and electron paramagnetic resonance spectroscopy, as well as kinetics measurements indicates that in ethanol/water solutions an [FeIII–()OH–CuIIOH2]+ species is generated which is the likely catalyst for 2,4-bis(dinitrophenyl)phosphate and DNA hydrolysis. Insofar as the data are consistent with the presence of an FeIII-bound hydroxide acting as a nucleophile during catalysis, 1 presents a suitable mimic for the hydrolytic enzyme purple acid phosphatase. Notably, 1 is significantly more reactive than its isostructural homologues with different metal composition (FeIIIMII, where MII is ZnII, MnII, NiII, or FeII). Of particular interest is the observation that cleavage of double-stranded plasmid DNA occurs even at very low concentrations of 1 (2.5 M), under physiological conditions (optimum pH of 7.0), with a rate enhancement of 2.7×107 over the uncatalyzed reaction. Thus, 1 is one of the most effective model complexes to date, mimicking the function of nucleases.Electronic Supplementary Material Supplementary material is available for this article at .  相似文献   

12.
Iron (II) and iron (III) complexes, [FeII(DEDTC)2(dppe)] · CH2Cl2 (1), [FeII(ETXANT)2(dppe)] (2) (DEDTC = diethyldithiocarbamate, ETXANT = ethyl xanthate, dppe = 1,2-bis (diphenylphosphino) ethane), and [FeIII(DEDTC)2(dppe)] [FeIIICl4] (3) have been synthesized and characterized. Since 3 contains two magnetic centers, an anion metathesis reaction has been conducted to replace the tetrahedral FeCl4 by a non-magnetic BPh4 ion producing [FeIII(DEDTC)2(dppe)]BPh4 (4) for the sake of unequivocal understanding of the magnetic behavior of the cation of 3. With the similar end in view, the well-known FeCl4 ion, the counter anion of 3, is trapped as PPh4[FeIIICl4] (5) and its magnetic property from 298 to 2 K has been studied. Besides the spectroscopic (IR, UV-Vis, NMR, EPR, Mass and XPS) characterization of the appropriate compounds, especially 2, others viz. 1, 3 and 4 have been structurally characterized by X-ray crystallography. While FeII complexes, 1 and 2, are diamagnetic, the FeIII systems, namely the cations of 3, and 4 behave as low-spin (S = 1/2) paramagnetic species from 298 to 50 K. Below 50 K 3 shows gradual increase of χMT up to 2 K suggesting ferromagnetic behavior while 4 exhibits gradual decrease of magnetic moment from 60 to 2 K, indicating the occurrence of weak antiferromagnetic interaction. These conclusions are supported by the Mössbauer studies of 3 and 4. The Mössbauer pattern of 1 exhibits a doublet site for diamagnetic (2-400 K) FeII. The compounds 1, 2 and 4 encompass interesting cyclic voltammetric responses involving FeII, FeIII and FeIV.  相似文献   

13.
Isopenicillin N synthase (IPNS) catalyzes formation of the β-lactam and thiazolidine rings of isopenicillin N from its linear tripeptide l-δ-(α-aminoadipoyl)-l-cysteinyl-d-valine (ACV) substrate in an iron- and dioxygen (O2)-dependent four-electron oxidation without precedent in current synthetic chemistry. Recent X-ray free-electron laser studies including time-resolved serial femtosecond crystallography show that binding of O2 to the IPNS–Fe(II)–ACV complex induces unexpected conformational changes in α-helices on the surface of IPNS, in particular in α3 and α10. However, how substrate binding leads to conformational changes away from the active site is unknown. Here, using detailed 19F NMR and electron paramagnetic resonance experiments with labeled IPNS variants, we investigated motions in α3 and α10 induced by binding of ferrous iron, ACV, and the O2 analog nitric oxide, using the less mobile α6 for comparison. 19F NMR studies were carried out on singly and doubly labeled α3, α6, and α10 variants at different temperatures. In addition, double electron–electron resonance electron paramagnetic resonance analysis was carried out on doubly spin-labeled variants. The combined spectroscopic and crystallographic results reveal that substantial conformational changes in regions of IPNS including α3 and α10 are induced by binding of ACV and nitric oxide. Since IPNS is a member of the structural superfamily of 2-oxoglutarate-dependent oxygenases and related enzymes, related conformational changes may be of general importance in nonheme oxygenase catalysis.  相似文献   

14.
Reactions of FeII, CoII, NiII, and ZnII salts with 6-quinolinecarboxylic acid (HL) under the hydrothermal conditions afford three monomeric complexes [M(L)2(H2O)4] (M = FeII for 1, CoII for 2, and NiII for 3) and a 1-D polymeric species {[Zn(L)2(H2O)] · H2O}n (4). The crystal structures of the ligand HL and these four complexes have been determined by using the X-ray single-crystal diffraction technique. The results suggest that complexes 1-3 are isostructural, displaying novel 3-D pillar-layered networks through multiple intermolecular hydrogen bonds, whereas in coordination polymer 4, the 1-D comb-like coordination chains are extended to generate a hydrogen-bonded layer, which is further reinforced via aromatic stacking interactions. Solid-state properties such as thermal stability and fluorescence emission of the polymeric ZnII complex 4 have also been investigated.  相似文献   

15.
Decarboxylation rates for a series of C-3 to C-6 α-keto acids were determined in the presence of resting cells and cell-free extracts of Streptococcus lactis var. maltigenes. The C-5 and C-6 acids branched at the penultimate carbon atom were converted most rapidly to the respective aldehydes in the manner described for α-carboxylases. Pyruvate and α-ketobutyrate did not behave as α-carboxylase substrates, in that O2 was absorbed when they were reacted with resting cells. The same effect with pyruvate was noted in a nonmalty S. lactis, accounting for CO2 produced by some “homofermentative” streptococci. Mixed substrate reactions indicated that the same enzyme was responsible for decarboxylation of α-ketoisocaproate and α-ketoisovalerate, but it appeared unlikely that this enzyme was responsible for the decarboxylation of pyruvate. Ultrasonic disruption of cells of the malty culture resulted in an extract inactive for decarboxylation of pyruvate in the absence of ferricyanide. Dialyzed cell-free extracts were inactive against all keto acids and could not be reactivated.  相似文献   

16.
Brown rot basidiomycetes initiate wood decay by producing extracellular reactive oxygen species that depolymerize the structural polysaccharides of lignocellulose. Secreted fungal hydroquinones are considered one contributor because they have been shown to reduce Fe3+, thus generating perhydroxyl radicals and Fe2+, which subsequently react further to produce biodegradative hydroxyl radicals. However, many brown rot fungi also secrete high levels of oxalate, which chelates Fe3+ tightly, making it unreactive with hydroquinones. For hydroquinone-driven hydroxyl radical production to contribute in this environment, an alternative mechanism to oxidize hydroquinones is required. We show here that aspen wood undergoing decay by the oxalate producer Postia placenta contained both 2,5-dimethoxyhydroquinone and laccase activity. Mass spectrometric analysis of proteins extracted from the wood identified a putative laccase (Joint Genome Institute P. placenta protein identification number 111314), and heterologous expression of the corresponding gene confirmed this assignment. Ultrafiltration experiments with liquid pressed from the biodegrading wood showed that a high-molecular-weight component was required for it to oxidize 2,5-dimethoxyhydroquinone rapidly and that this component was replaceable by P. placenta laccase. The purified laccase oxidized 2,5-dimethoxyhydroquinone with a second-order rate constant near 104 M−1 s−1, and measurements of the H2O2 produced indicated that approximately one perhydroxyl radical was generated per hydroquinone supplied. Using these values and a previously developed computer model, we estimate that the quantity of reactive oxygen species produced by P. placenta laccase in wood is large enough that it likely contributes to incipient decay.Brown rot basidiomycetes are the principal recyclers of woody biomass in coniferous forest ecosystems and also the chief cause of decay in wooden structures (8, 41). Unlike the closely related white rot fungi, they degrade the cellulose and hemicellulose in wood while mineralizing little of the lignin that shields these structural polysaccharides from enzymatic attack. As a result, extensively brown-rotted wood consists primarily of an oxidized, partially cleaved residue derived from the original lignin (7, 16, 21, 22, 40). This failure to remove lignin efficiently suggests that brown rot systems contain fewer components than white rot systems and, in agreement, the recently published genome sequence of Postia placenta shows that this brown rot fungus lacks the ligninolytic peroxidases generally thought important for white rot (26). Instead, brown rot fungi appear to rely, at least during incipient decay, on small agents that can penetrate the lignin to access the polysaccharides (10, 14).A better understanding of the biodegradative agents produced by P. placenta may provide clues about what constitutes a minimally effective system for the microbial deconstruction of lignocellulose. One potential low-molecular-weight contributor is an extracellular metabolite, 2,5-dimethoxyhydroquinone (2,5-DMHQ), which has been found in cultures of P. placenta and other brown rot fungi and also shown to reduce Fe3+ with concomitant H2O2 production, thus producing hydroxyl radicals (·OH) via the Fenton reaction (Fig. (Fig.1,1, reaction 6) (4, 17, 20, 29, 34, 36). Past work has shown that the chemical changes introduced by brown rot fungi into wood, cellulose, and other polymers are consistent with attack by reactive oxygen species (ROS) such as ·OH (4, 7, 19, 21-23). A second small agent with a proposed role is extracellular oxalic acid, which P. placenta produces in sufficient quantity to acidify colonized wood to pH 2 to 4. Assays in vitro have shown that cellulose is slowly hydrolyzed at these acidities (11).Open in a separate windowFIG. 1.Chemical reactions discussed in the text. For simplicity, the HOO·/O2·− acid/base pair is shown only as HOO·. H2Q, hydroquinone; HQ·, semiquinone; Q, quinone.However, there is an apparent contradiction between these two mechanisms: oxalate is a strong chelator of Fe3+, and the resulting Fe3+ trioxalate complex has too negative a reduction potential to react readily with methoxyhydroquinones such as 2,5-DMHQ (28, 37). In considering this problem, we noted the surprising finding that the P. placenta genome encodes two putative laccases, enzymes that are considered atypical of brown rot fungi (26). Laccases oxidize methoxyhydroquinones to semiquinone radicals, which generally have more negative reduction potentials than their parent hydroquinones (38), and are therefore expected to be better reductants of Fe3+. In addition, methoxysemiquinones reduce O2 to generate perhydroxyl radicals (HOO·) and their conjugate base superoxide (O2·−), which dismutate to produce H2O2. Furthermore, HOO·/O2·− can reduce some Fe3+ chelates to generate additional Fe2+ and can oxidize some Fe2+ chelates to generate additional H2O2 (9, 13, 33, 38). By these routes, a P. placenta laccase could bypass the requirement for the hydroquinone to react directly with Fe3+ and could thus generate a complete Fenton system (Fig. (Fig.1,1, reactions 1 to 7).Here we have expressed one of the P. placenta putative laccase genes heterologously and thus demonstrate that it encodes a typical laccase. In addition, we show that laccase activity and this particular enzyme are present in wood undergoing decay by P. placenta. Furthermore, we report that 2,5-DMHQ is present in the biodegrading wood, that it is a substrate for the P. placenta laccase, and that its oxidation during incipient wood decay requires a macromolecular component that is replaceable by P. placenta laccase. Finally, we show that the oxidation of 2,5-DMHQ by the P. placenta laccase results in significant H2O2 production, and we estimate that the quantity of ROS produced by this route is large enough that it probably contributes to incipient brown rot.  相似文献   

17.
The in vivo metabolism of 6-keto PGF was investigated in rats. Following continuous intravenous infusion for 14 days the urinary metabolites were isolated and identified. A substantial amount of unchanged 6-keto PGF was recovered in the urine. The metabolic pattern very closely resembles that of PGI2 in rats. Metabolites were found which represented 15-dehydrogenation, β-oxidation, ω and ω-1-hydroxylation and oxidation.Previous work showed that 6-keto PGF is very poorly oxidized by 15-PGDH. We administered 15-[H3]-PGI2 and 15-[H3]-6-keto PGF to rats and measured urinary tritiated water as an index for in vivo 15-PGDH activity. The results showed that PGI2 and 6-keto PGF were both oxidized to the 15-keto product, although the rate of oxidation of PGI2 was greater than that of 6-keto PGF. We concluded that the administered PGI2 was oxidized by 15-PGDH before hydrolysis to 6-keto PGF. A portion of the dose is probably hydrolyzed before 15-dehydrogenation.  相似文献   

18.
Ray K  Lee SM  Que L 《Inorganica chimica acta》2008,361(4):1066-1069
The mechanism of formation of [FeIV(O)(N4Py)]2+ (2, N4Py = N,N-bis(2-pyridylmethyl)-N-bis(2-pyridyl)methylamine) from the reaction of [FeII(N4Py)(CH3CN)]2+ (1) with m-chloroperbenzoic acid (mCPBA) in CH2Cl2 at −30 °C has been studied on the basis of the visible spectral changes observed and the reaction stoichiometry. It is shown that the conversion of 1 to 2 in 90% yield requires 1.5 equiv. peracid and takes place in two successive one-electron steps via an [FeIII(N4Py)OH]2+(3) intermediate. The first oxidation step uses 0.5 equiv. peracid and produces 0.5 equiv. 3-chlorobenzoic acid, while the second step uses 1 equiv. peracid and affords byproducts derived from chlorophenyl radical. We conclude that the FeII(N4Py) center promotes O-O bond heterolysis, while the FeIII(N4Py) center favors O-O bond homolysis, so the nature of O-O bond cleavage is dependent on the iron oxidation state.  相似文献   

19.
The purpose of our study was to explore the effect and intrinsic mechanism of wild-type IDH1 and its substrate α-KG on renal cell carcinoma (RCC). IDH1 was observed lower expression in RCC cell lines. Phenotype experiment was carried out in the wild-type IDH1 and mutant IDH1R132H plasmid treated cell line. The results showed that the wild-type IDH1 could significantly inhibit the proliferation, migration and promote the apoptosis of RCC cell lines, which were consistent with the IDH1''s substrate α-KG. The mutant IDH1R132H was found to lose this biological function of IDH1. Moreover, we verified the proliferation inhibition of IDH1 in vivo. In addition, we verified the correlation between IDH1 and hypoxia signal-related proteins in vitro and in vivo, specifically, IDH1 overexpression could significantly reduce the expression of HIF-1α and HIF-2α proteins and its downstream proteins (VEGF, TGF-α). Furthermore, we preliminarily verified the possibility of α-KG in the RCC''s treatment by injecting α-KG into the xenograft model. α-KG significantly reduced tumor size and weight in tumor-bearing mice. This study provided a new therapeutic target and small molecule for the study of the treatment and mechanism of RCC.  相似文献   

20.
Analyses of chlorophyll fluorescence induction kinetics from DCMU-poisoned thylakoids were used to examine the contribution of the light-harvesting chlorophyll protein complex (LHCP) to Photosystem II (PS II) heterogeneity. Thylakoids excited with 450 nm radiation exhibited fluorescence induction kinetics characteristic of major contributions from both PS IIα and PS IIβ centres. On excitation at 550 nm the major contribution was from PS IIβ centres, that from PS IIα centres was only minimal. Mg2+ depletion had negligible effect on the induction kinetics of thylakoids excited with 550 nm radiation, however, as expected, with 450 nm excitation a loss of the PS IIα component was observed. Thylakoids from a chlorophyll-b-less barley mutant exhibited similar induction kinetics with 450 and 550 nm excitation, which were characteristic of PS IIβ centres being the major contributors; the PS IIα contribution was minimal. The fluorescence induction kinetics of wheat thylakoids at two different developmental stages, which exhibited different amounts of thylakoid appression but similar chlorophyll ratios and thus similar PS II:LHCP ratios, showed no appreciable differences in the relative contributions of PS IIα and PS IIβ centres. Mg2+ depletion had similar effects on the two thylakoid preparations. These data lead to the conclusion that it is the PS II:LHCP ratio, and probably not thylakoid appression, that is the major determinant of the relative contributions of PS IIα and PS IIβ to the fluorescence induction kinetics. PS IIα characteristics are produced by LHCP association with PS II, whereas PS IIβ characteristic can be generated by either disconnecting LHCP from PS II or by preferentially exciting PS II relative to LHCP.  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