共查询到20条相似文献,搜索用时 0 毫秒
1.
Patrick J. Collins Alan D. W. Dobson Jim A. Field 《Applied and environmental microbiology》1998,64(6):2026-2031
Laccase is a copper-containing phenoloxidase, involved in lignin degradation by white rot fungi. The laccase substrate range can be extended to include nonphenolic lignin subunits in the presence of a noncatalytic cooxidant such as 2,2′-azinobis(3-ethylbenzthiazoline-6-sulfonate) (ABTS), with ABTS being oxidized to the stable cation radical, ABTS·+, which accumulates. In this report, we demonstrate that the ABTS·+ can be efficiently reduced back to ABTS by physiologically occurring organic acids such as oxalate, glyoxylate, and malonate. The reduction of the radical by oxalate results in the formation of H2O2, indicating the formation of O2·− as an intermediate. O2·− itself was shown to act as an ABTS·+ reductant. ABTS·+ reduction and H2O2 formation are strongly stimulated by the presence of Mn2+, with accumulation of Mn3+ being observed. Additionally, 4-methyl-O-isoeugenol, an unsaturated lignin monomer model, is capable of directly reducing ABTS·+. These data suggest several mechanisms for the reduction of ABTS·+ which would permit the effective use of ABTS as a laccase cooxidant at catalytic concentrations.Lignin, the second most abundant renewable organic compound in the biosphere after cellulose, is highly recalcitrant, and therefore its biodegradation is a rate-limiting step in the global carbon cycle (9). White rot fungi have evolved a unique mechanism to accomplish this degradation, which utilizes extracellular enzymes to generate oxidative radical species (16). This degradative system is highly nonspecific, and as a consequence, these fungi can also oxidize a broad spectrum of structurally diverse environmental pollutants (4, 18). Three main groups of enzymes, i.e., lignin peroxidases (LiP), manganese peroxidases (MnP), and laccases, along with their low-molecular-weight cofactors, have been implicated in the lignin degradation process. LiP can oxidize the nonphenolic aromatic moieties that make up approximately 85% of the lignin polymer (21), while MnP uses the Mn2+/Mn3+ couple to oxidize phenolic subunits (19). Laccase, a copper-containing phenoloxidase, catalyzes the four-electron reduction of oxygen to water, and this is accompanied by the oxidation of a phenolic substrate (32).In recent years, however, the laccase substrate range has been extended to include nonphenolic lignin subunits in the presence of readily oxidizable primary substrates. These cooxidants have been denoted mediators because they were previously speculated (but not proven) to act as electron transfer mediators. The most extensively investigated laccase mediator is 2,2′-azinobis(3-ethylbenzthiazoline-6-sulfonate) (ABTS), a synthetic nitrogen-substituted aromatic compound which allows the oxidation of nonphenolic lignin model compounds (6) and the delignification of kraft pulp (8) by laccase. More recent work has also focused on an alternative compound, 1-hydroxybenzotriazole (7, 10). In the presence of these compounds, laccase can also catalyze the oxidation of polycyclic aromatic hydrocarbons (PAH) (12, 23), chemical synthesis (29), and textile dye bleaching (31). ABTS is oxidized by laccase to its corresponding cation radical. In the case of ABTS, the radical (ABTS·+) is highly stable, and it has been suggested that it may act as a diffusible oxidant of the enzyme (7). However, although the redox chemistry of ABTS (22) and its radical has been characterized, the mechanisms by which it interacts with laccase to “mediate” lignin oxidation are still unknown. Potthast et al. (28) have found evidence suggesting that ABTS acts as an activator or cooxidant of the enzyme. The observation that the laccase/ABTS couple can oxidize the nonphenolic veratryl alcohol, while ABTS·+ alone cannot (6), provides a further indication of this activator role for ABTS. If compounds such as ABTS do indeed act as cooxidants of the enzyme, it is necessary that some mechanism(s) exists for the recycling of their cation radicals back to their reduced forms so as to be available for subsequent catalytic cycles.A number of low-molecular-weight compounds have been implicated in the catalysis of MnP during the oxidation of lignin. The most important of these is manganese, which is present in virtually all woody tissues (17). Divalent manganese (Mn2+) is oxidized by the enzyme to the trivalent form (Mn3+), which is capable of oxidizing an extensive range of phenolic compounds (19). To catalyze lignin oxidation, Mn3+ is chelated and stabilized by organic acids, which facilitate its diffusion to act as an oxidant at a distance from the MnP active site (19, 33). A range of these acids are produced by ligninolytic fungi (25, 30, 33), but the most ubiquitous is oxalate, whose production at levels as high as 28 mM by cultures of Pleurotus ostreatus has been observed (1). Oxalate can itself be oxidized by Mn3+, producing the formate anion radical (CO2·−), which can then reduce molecular oxygen to produce superoxide (O2·−) (24), and a role for these radicals as reducing agents in lignin degradation has been suggested (24).In this report, evidence is presented indicating that physiologically occurring organic acids can directly reduce ABTS·+. The rate of reduction is highly stimulated by the presence of manganese, and the results indicate a mechanism involving O2·−. 相似文献
2.
《Free radical research》2013,47(3):159-168
N-phenylacetyl dehydroalanines are captodative olefins. They inhibit two processes mediated by superoxide anion (O2-) in a concentration dependent manner: reduction of NBT to blue formazan and oxidation of epinephrine to adrenochrome. They also inhibit in a dose related way the degradation of deoxyribose produced during either the Fenton reaction or the radiolysis of water, which are the two experimental sources of hydroxyl radical (HO?) production. Based on the results obtained with superoxide dismutase, mannitol, thiourea, and uric acid, we postulate that these competitive inhibitory effects suggest a reaction between the dehydroalanine derivatives and the two oxygen derived radicals. Hydroxyl free radical is scavenged more efficiently than superoxide anion. Substitution of the phenyl ring by methoxy groups does not modify significantly the activity. These molecules possess three target active sites which can react with free radicals. 相似文献
3.
Advanced oxidation processes are currently used for the treatment of different reactive dyes which involve use of toxic catalysts. Peroxidases are reported to be effective on such dyes and require hydrogen peroxide and/or metal ions. Cyathus bulleri laccase, expressed in Pichia pastoris, catalyzes efficient degradation (78 to 85%) of reactive azo dyes (reactive black 5, reactive orange 16, and reactive red 198) in the presence of synthetic mediator ABTS [2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid)]. This laccase was engineered to degrade effectively reactive blue 21 (RB21), a phthalocyanine dye reported to be decolorized only by peroxidases. The 816-bp segment (toward the C terminus) of the lcc gene was subjected to random mutagenesis and enzyme variants (Lcc35, Lcc61, and Lcc62) were selected based on increased ABTS oxidizing ability. Around 78 to 95% decolorization of RB21 was observed with the ABTS-supplemented Lcc variants in 30 min. Analysis of the degradation products by mass spectrometry indicated the formation of several low-molecular-weight compounds. Mapping the mutations on the modeled structure implicated residues both near and far from the T1 Cu site that affected the catalytic efficiency of the mutant enzymes on ABTS and, in turn, the rate of oxidation of RB21. Several inactive clones were also mapped. The importance of geometry as well as electronic changes on the reactivity of laccases was indicated. 相似文献
4.
《Inorganica chimica acta》1986,122(1):111-118
The title complex, prepared in 1 M NaOH, was crystallized from hot N,N-dimethylformamide/ ethanol solutions to give Na12[Ce(C6H2O2(SO3)2)4]· 9H2O·6DMF. The purple—brown crystals were examined by X-ray diffraction while inside quartz capillaries filled with DMF, (λmax 425 nm, ϵ 3664; λsh 520 nm, ϵ 2240) and belong to space group Pbca, Z=8 with a=21.846(4), b=17.348(2), c=43.103- (6) Å, V=16.335(7) Å3, Dc=1.693 gcmt−3, Do=1.725 g cmt−3. Diffractometer data were collected using Mo Kα radiation to 2θ=43o. For 7331 independent data with Fo2>3σ(Fo2) full matrix least squares refinement converged to unweighted and weighted R factors of 0.072 and 0.110, respectively, with a mixture of anisotropic and isotropic thermal parameters. The disordered DMF atom parameters were not refined. The structure consists of discrete monomeric Ce(C6H2S2O8)412− units with 12 Na+ counter cations and 10 H2O molecules (two with half occupancy), and 6 DMF molecules of solvation filling up spaces between cations and anions. Cerium(IV) is in a general position with a coordination polyhedron close to the trigonal-faced dodecahedron, D2d, with the angles between the two BAAB trapezoids of 2.3o and 3.7o. The average CeO(A) distance, 2.363(9) Å is longer than the average CeO(B) distance, 2.326(15)Å, with the reverse being true for one of the four tironato ligands. The average ring OCeO angle is 67.9(1)o. The cerium (IV) complex is found by cyclic voltammetry to undergo a quasi-reversible one-electron reduction (in strongly basic solution with excess tiron) with Ef=−497 mV vs. SCE, hence the ratio of the formation constants for tetrakis(tironato)cerate(IV) to that for tetrakis(tironato)cerate(III), KIV/KIII, is 1033. Characterization of other tiron salts is reported. 相似文献
5.
One of the radical species produced by the reaction of dehydro-l-ascorbic acid with an α-amino acid gave a very characteristic hyperfine structure in its electron spin resonance spectrum. The same spectrum was also obtained when l-scorbamic acid was oxidized with some oxidants, indicating the formation of the radical via the oxidation of l-scorbamic acid. From the results of deuterium exchange experiments, simulation spectra and the reduction of 2,2′-nitrilodi-2(2′)-deoxy-l-ascorbic acid monoammonium salt, the radical was concluded to be monodehydro-2,2′-iminodi-2(2′)-deoxy-l-ascorbic acid. Possible formation mechanism of the radical was also discussed. 相似文献
6.
7.
《Inorganica chimica acta》1986,117(1):45-48
The resonance Raman spectra of K2(Ti(O2)(SO4)2)·5H2O and K2(Ti(O2)(C2O4)2)·3H2O are recorded. The results are consistent with the triangular structure of the peroxotitanium unit, Ti(O2), with C2ν symmetry. The ν(OO), νs(TiO) and νas(TiO) are observed around 890, 610 and 535 cm−1, respectively. The resonance effects are shown to be associated with the 425 nm absorption band. This band is assigned to the O22− → Ti(IV) charge-transfer transition. The calculated force constant values for the O22− and TiO bonds are 320 and 275 N m−1, respectively. 相似文献
8.
《Inorganica chimica acta》1988,142(1):165-168
Some luminescent properties of single crystals of Eu(NO3)3·6H2O are reported at temperatures down to liquid helium. The water molecules are responsible for a considerable amount of quenching of the emission. The Eu3+ coordination changes upon lowering the temperature. The number of 5D0-7F2 lines observed is higher than expected. 相似文献
9.
《Inorganica chimica acta》1986,116(1):11-13
The emission spectra and excited state decay rates have been recorded for Cr(CN)6−n(H2O)nn−3 (n = 0-6) complexes. Both the transition energy and relaxation rates increase with n but the large changes in transition energies are not sufficient to account for the failure of the displaced coordinate to explain the relaxation rate results. 相似文献
10.
《Inorganica chimica acta》1988,154(2):215-219
By reacting 2,2′-biimidazole and copper(II) chloride in aqueous HCl we obtained the complex CuCl2(H2bim) as the main product and a compound with stoichiometry Cu1.5Cl3(H2bim)2 as a byproduct. The structure of the latter compound has been determined by X-ray analysis: monoclinic, a= 794.0(3), b=3146.8(6), c=722.9(4) pm, β= 114.2(1)°, space group P21/c. The compound actually contains two species, namely [Cu(H2bim)2]Cl2 and [CuCl2(H2bim)] in a 1:2 molar ratio. 相似文献
11.
12.
Hiroyoshi Yuki Qing-Li Zhao Yoshisada Fujiwara Kiyoshi Tanabe Ryohei Ogawa 《Free radical research》2013,47(6):631-643
Hyperthermia-induced apoptosis and its enhancement in the presence of a temperature-dependent free radical initiator, 2,2′-azobis (2-aminopropane) dihydrochloride (AAPH) were examined in human uterine cervical cancer cell lines, CaSki and HeLa. When both cell lines were treated with hyperthermia at 44°C for 60?min, minimal apoptosis was observed. When combined with nontoxic AAPH (50?mM), significant enhancement of apoptosis was observed, where the initial rate of free radical formation was about twice as high than that at 37°C. Augmentation of the growth delay, lipid peroxidation (LPO), activation of caspase-3 and increase in [Ca2+]i were also observed after the combined treatment. A water-soluble vitamin E, Trolox, blocked the increase in [Ca2+]i and an intracellular Ca2+ chelator, BAPTA-AM, prevented the DNA fragmentation induced by the combination. Cytochrome c release was also revealed by fluorescence microscopy. However, no significant change in mitochondrial membrane potential and expression of Bax and Bcl-2 was observed. A slight increase in Fas expression was observed only in CaSki cells after the combined treatment. These results indicate that hyperthermia and AAPH induce enhanced apoptosis and subsequent cell killing via two pathways; a pathway dependent on increase in LPO and [Ca2+]i, and a pathway associated with cytochrome c release and subsequent caspase activation without changes of mitochondrial membrane potential and Bax/Bcl-2 expression in these cell lines. Since it is known that cancer cells are generally resistant to physical and chemical stress-induced apoptosis, free radical generators like AAPH appear to be a useful thermosensitizer for hyperthermic cancer therapy. 相似文献
13.
Guduru Moulika Sannapureddy Sailaja Jillela Santhosh Vijitha Putluru Bayapu Reddy Kondala Shanthi Latha Gasthi Venkata Chalapathi Busireddy Sudhakar Reddy 《Luminescence》2022,37(7):1073-1077
Calcium boro fluoro zinc phosphate glasses modified using alkali oxide and doped with Nd3+ and Er3+ ions with the chemical composition of 69.5 (B2O3) + 10 (P2O5) + 10 (CaF2) + 5 (ZnO) + 5 (Na2O/Li2O/K2O) + 0.5 (Er2O3/Nd2O3) were prepared using a conventional melt quenching technique. The results of X-ray diffraction patterns indicated the amorphous nature of all the prepared glasses. The visible–near-infrared red (NIR) absorption spectra of these glasses were analyzed systematically. The NIR emission spectra of Er3+ and Nd3+:calcium boro fluoro zinc phosphate glasses showed prominent emission bands at 1536 nm (4I13/2→4I15/2) and 1069 nm (4F3/2→4I11/2) respectively with λexci = 514.5 nm (Ar+ laser) as the excitation source. 相似文献
14.
15.
《Bioscience, biotechnology, and biochemistry》2013,77(4):659-661
The effects of drinking deaerated water on serum biochemical values, and on the concentrations of short-chain fatty acids (SCFAs) derived from bacterial fermentation in the colon were examined in rats. Drinking deaearted water decreased the levels of serum alkaline phosphatase (SAP) and serum urea nitrogen (SUN), and increased the serum potassium (SK) and serum phosphorus (SP) levels. Although the concentration of propionic acid in the cecum was decreased by drinking deaerated water, the concentrations of isobutyric, valeric, and isovaleric acids in the cecum were increased. 相似文献
16.
An enzyme which catalyzes the oxidation of poly(vinyl alcohol) (PVA) has been purified from a fraction adsorbed to DEAE-Sephadex at pH 7.0 from PVA-degrading enzyme activities produced by a bacterial symbiotic mixed culture in a culture broth when the culture was grown in a minimal medium where PVA served as a sole source of carbon and energy. The enzyme was separated from a coexisting oxidized PVA hydrolase by dye-ligand chromatography on Matrex Gel Blue A. The purified enzyme was homogeneous as judged by polyacrylamide gel electrophoreses in the absence and presence of SDS.The enzyme is a single polypeptide with a molecular weight of about 40,000 and has an isoelectric point of 4.5. The amino acid composition of the enzyme has been determined and found to have no histidine. The N- and C-terminal amino acid residues are both alanine. The enzyme solution is pink and shows absorption maxima at 276, 364, and 469 nm. One atom of non-heme iron has been detected per molecule in the enzyme.The enzyme catalyzes the oxidation of PVA and also of various low molecular weight secondary alcohols to the corresponding ketones with the production of H202 and the consumption of 02. The molar ratio of these ketones, H202 and 02 is 1:1:1. The most effective electron acceptor is 02, while 2,6-dichlorophenolindophenol and nitro blue tetrazolium also serve as the acceptor with efficiencies to 02 of about 31 and 16%, respectively. The enzyme is, therefore, considered to be a secondary alcohol oxidase.The enzyme is most active at pH 7.0 and at 45°C and is stable between pH 5.0 and 9.0 and at temperatures below 45°C. The activity is inhibited by Hg2+ and is restored by the addition of reduced glutathione, although p-chloromercuribenzoate has no effect.The enzyme shows a common antigenicity in immunodiffusion and neutralization reactions with antisera to a secondary alcohol oxidase previously isolated from another fraction adsorbed on SP-Sephadex at pH 7.0 of the PVA-degrading enzyme activities [Agric. Biol. Chem., 43, 1225 (1979)]. The relations between these two secondary alcohol oxidases are discussed. 相似文献
17.
Zhuqing Wang Na Li Longmei Lv Yongqi Zhang Guang Chai Xiru Cao 《Molecular simulation》2013,39(16):1312-1321
ABSTRACTThe effect of (H2O)n (n?=?1–3) on the HNO2?+?HO → H2O?+?NO2 reaction has been investigated theoretically at the CCSD(T)/CBS//B3LYP/6-311?+?G(3df,2pd) level of theory, coupled with rate constant calculations by using variational transition state theory. Our results show that, when (H2O)n (n?=?1–3) was introduced into HNO2?+?HO → H2O?+?NO2 reaction, the product of the reaction did not change, but the potential energy surface became quite complex, yielding two kinds of reactions, namely HNO2···(H2O)n (n?=?1–3)?+?HO and HO···(H2O)n (n?=?1–3)?+?HNO2. In all catalysed reactions with (H2O)n (n?=?1–3), the former reaction type is favourable than the latter one with its effective rate constant respectively larger by 6–1 orders of magnitude than that of latter one. Within the temperature range of 240–320?K, the relative impacts on water monomer are much more obvious than dimer and trimer. However, the effective rate constant with water is larger by 658%–17% times of magnitude, showing that the positive water effect is obvious under atmospheric conditions. 相似文献
18.
《Journal of inorganic biochemistry》1986,26(1):1-21
Two structurally different phases of a uric acid salt of magnesium, Mg(hydrogenurate)2 · 8H2O, have been prepared by crystallization from solution at pH = 7.5–8.0 and were investigated by x-ray crystallography, thermal analysis, and ir spectroscopy. Both phases are monoclinic, space group P21/c with a = 9.573(2), b = 14.627(3), c = 7.170(1) Å, β = 101.91(1)° (phase I) and a = 10.397(2), b = 14.306(3), c = 6.732(1) Å, β = 104.64(2)° (phase II). The crystal structures of both phases (R = 0.053 and 0.051, respectively) contain isolated octahedral [Mg(H2O)6]2+ cations, hydrogenurate monoanions, and two molecules of water of crystallization per formula unit. The structural formula representing these facts is [Mg(H2O)6] (hydrogenura-te)2·2H2O. The tautomeric form of the hydrogenurate molecule corresponds to the tri-keto form of uric acid deprotonated at N(3). Differences in bond length between uric acid and the hydrogenurate molecule may be described in terms of three additional resonance structures distributing the formal negative charge at N(3) within the pyrimidine (but not the imidazole) ring. Deprotonation at N(3) significantly decreases the internal C-N-C angle at N(3). Alternating pairs of medium-strong intermolecular N-HO hydrogen bonds lead to infinite chains of hydrogenurate molecules extending along the b axis of the unit cells in both phases. The main difference between the two phases lies in their stacking pattern of the hydrogenurate molecules. Infrared data confirm the hydrogen bonding characteristics resulting from the crystal structure analysis. Thermogravimetric measurements and differential scanning calorimetry data show that the dehydration of both phases occurs in two distinct steps with Mg(hydrogenurate)2.6H2O as an intermediate phase. The first dehydration step (−2H2O) is a topotactic reaction with three-dimensional preservation of the main structure elements of the octahydrate in the structure of the hexahydrate. 相似文献
19.
20.
Mechanism of O2−- and H2O2-induced stimulation of sugar transport in mouse fibroblast BALB/3T3 cells
《Biochimica et Biophysica Acta (BBA)/Molecular Cell Research》1988,972(3):293-298
Xanthine/xanthine oxidase and H2O2 stimulated sugar transport. Application of superoxide dismutase and catalase to the cells showed an inhibitory effect on these agent-stimulated sugar transports. Addition of amiloride and 4-acetamide-4′-isothiocyanostilbene-2,2′-disulfonic acid (SITS), which abolish the cytoplasmic alkalinization, inhibited the stimulation of sugar transport by xanthine/xanthine oxidase in the presence of catalase. The calmodulin antagonists, N-(6-aminohexyl)-5-chloro-1-naphthalenesulfonamide (W-7) and trifluoperazine inhibited H2O2-stimulated sugar transport. These results suggest that O2− stimulates sugar transport in an intracellular pH-dependent manner and that H2O2 stimulates sugar transport in a calcium-calmodulin-dependent manner. These mechanisms may be involved in sugar-transport stimulation in mouse fibroblast BALB/3T3 cells by the tumor-promoting phorbol ester phorbol-12,13-dibutyrate and insulin, since the stimulatory effects of these agents were inhibited by scavengers of oxygen radicals. 相似文献