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.
Previous work has shown that Y-family DNA polymerases tolerate large DNA adducts, but a substantial decrease in catalytic efficiency and fidelity occurs during bypass of N2,N2-dimethyl (Me2)-substituted guanine (N2,N2-Me2G), in contrast to a single methyl substitution. Therefore, it is unclear why the addition of two methyl groups is so disruptive. The presence of N2,N2-Me2G lowered the catalytic efficiency of the model enzyme Sulfolobus solfataricus Dpo4 16,000-fold. Dpo4 inserted dNTPs almost at random during bypass of N2,N2-Me2G, and much of the enzyme was kinetically trapped by an inactive ternary complex when N2,N2-Me2G was present, as judged by a reduced burst amplitude (5% of total enzyme) and kinetic modeling. One crystal structure of Dpo4 with a primer having a 3′-terminal dideoxycytosine (Cdd) opposite template N2,N2-Me2G in a post-insertion position showed Cdd folded back into the minor groove, as a catalytically incompetent complex. A second crystal had two unique orientations for the primer terminal Cdd as follows: (i) flipped into the minor groove and (ii) a long pairing with N2,N2-Me2G in which one hydrogen bond exists between the O-2 atom of Cdd and the N-1 atom of N2,N2-Me2G, with a second water-mediated hydrogen bond between the N-3 atom of Cdd and the O-6 atom of N2,N2-Me2G. A crystal structure of Dpo4 with dTTP opposite template N2,N2-Me2G revealed a wobble orientation. Collectively, these results explain, in a detailed manner, the basis for the reduced efficiency and fidelity of Dpo4-catalyzed bypass of N2,N2-Me2G compared with mono-substituted N2-alkyl G adducts.  相似文献   

2.
We have synthesized 35 N-phenyl-N′-(4-pyridyl)urea derivatives and tested their cytokinin activity in the tobacco callus bioassay. Among them, N-phenyl-N′- (2-chloro-4-pyridyl)urea is highly active, the optimum concentration of which is lower than 4 × 10?9 M (0.001 ppm), 3 compounds, i.e. N-(2-methylphenyl)-N′-(2-chloro-4-pyridyl)urea, N-(3-methylphenyl)-N′-(2-chloro-4-pyridyl)urea and N-(3-chlorophenyl)-N′-(2-chloro-4-pyridyl) urea are as active as N6-benzyladenine (concentration for optimum yield: 4.4 × 10?8 M or 0.01 ppm), and N-phenyl-N′-(2-methyl-4-pyridyl)urea and N-(2-chlorophenyl)-N′-(2-chloro-4-pyridyl)urea are as active as N-phenyl-N′-(4-pyridyl)urea (concentration for optimum yield: 4.7 × 10?7 M or 0.1 ppm), while the activity of the other 29 compounds are not so remarkable and 11 of them are almost or completely inactive.  相似文献   

3.
The annual shoots of apple and pear trees which accumulated a high concentration of arginine during the dormant stage also contained Nα-acylarginine derivatives. Nα-(2-Hydroxysuccinyl)arginine, Nα-(3-hydroxysuccinyl)arginine and Nα-oxalylarginine were found in apple trees, and Nα-succinylarginine and Nα-(2-carboxymethyl-2-hydroxysuccinyl)arginine, besides the former three, were found in pear trees. Nα-(3-Hydroxysuccinyl)arginine, Nα-oxalylarginine and Nα-succinylarginine are new arginine derivatives.  相似文献   

4.
Lung N-oxidase enzyme activity was about three times higher than liver N-oxidase at the pH optimum, about pH 8.9, whereas the activities were nearly the same at more physiological ranges of pH. The lung N-oxidase was also stimulated about 2-fold by 100 mM Mg2+ and by 0.1 mM Hg2+, whereas liver N-oxidase activity was inhibited by these concentrations of ions. The difference in response of liver and lung enzymes to Mg2+ and Hg2+ was not altered by preparing the microsomes in the presence of 50 mM ethylenediamine tetraacetic acid (EDTA) in 0.1 M Tris (hydroxymethyl) amino methane (Tris) buffer or 50 mM EDTA in 0.1 M KPO4 buffer, both at pH 7.6, indicating that the differences are probably not due to the presence of endogenous metals. The difference between the liver and lung N-oxidase systems may be due to the tissue environment rather than to the enzyme itself since mercury stimulation of lung N-oxidation began to disappear upon partial purification of the N-oxidase enzymes. In contrast to the effects of Hg2+ and Mg2+, 1 mM Ni2+ enhanced liver N-oxidase activity about 30% and 5 mM Ni2+ stimulated lung enzyme activity about 30% whereas concentrations above 10 mM were inhibitory to both N-oxidases. Both liver and lung demethylase activities were inhibited by these concentrations of Mg2+, Hg2+ and Ni2+.Various suifhydryl reagents were also tested for their effects on these enzymes. The mercurials, para-chloromercurybenzoate (pCMB) and phenylmercuryacetate (PMA) at concentrations of 0.1 mM had the same effect as HgCl2 inhibiting both demethylases and liver N-oxidase, but stimulating lung N-oxidase activity. However, 0.1 mM to 1 mMN-ethylmaleimide (NEM) and iodoacetamide had little if any effect on either liver or lung N-oxidase. It was also shown that Hg2+ effects on N-oxidase activity could be overcome by dilution.Changes in N,N-dimethyl aniline (DMA) metabolism with age were followed in rabbits from 4 days old to adult. There was a steady increase in lung demethylase activity and N-oxidase activity in the liver and lung to adult levels. However, the liver demethylase had a sharp increase in activity between 2 weeks and 1 month much like that seen with benzphetamine demethylase in rabbit liver.Activities of N-demethylase in liver and lung, and N-oxidr.se in liver from new-born rabbits were from 10 to 20 % of adult levels. However, in lung, N-oxidase activities in the newborn were about 50 % of adult levels. Microsomal N-oxidation in lungs from 2-day-old rabbits was stimulated by 0.1 mM mercury just as in the adult.  相似文献   

5.
Reported are four iron(II) complexes with N-benzyl-N,N′-bis(2-pyridylmethyl)-1,2-ethanediamine (LH) and three electronically modified derivatives: N-(4-methoxy)benzyl-N,N′-bis(2-pyridylmethyl)-1,2-ethanediamine (LOMe), N-(4-chloro)benzyl-N,N′-bis(2-pyridylmethyl)-1,2-ethanediamine (LCl), and N-(4-nitro)benzyl-N,N′-bis(2-pyridylmethyl)-1,2-ethanediamine (LNO2). The four ligands react with FeCl2 to form a series of mononuclear species with the general formula [Fe(LR)Cl2]. The cis-α conformation of the ligand places the amine N-donors trans to the Fe-Cl bonds. The identity of the 4-benzyl substituent has profound influences on the lengths of the iron-ligand bonds, the optical spectra, and the redox activities of the [Fe(LR)Cl2] compounds.  相似文献   

6.
The reduced one-carbon-substituted derivatives of folic acid can be grouped in three pools according to their response to acid treatment. Pool 1 is made up of N5,N10-methylene-tetrahydrofolic acid and unsubstituted dihydro- and tetrahydrofolic acid which at pH 1.0 and subsequent exposure to air cleave to p-aminobenzoylglutamic acid. Pool 2 is made up by the acid-stable N5-methyl-tetrahydrofolic acid, and pool 3 includes N5,N10-methenyl-tetrahydrofolic acid, N10-formyltetrahydrofolic acid, N5-formyltetrahydrofolic acid, and N5-formiminotetrahydrofolic acid, all of which convert to the stable N5,N10-methenyl-tetrahydro form when acid treated. Conditions are described to selectively cleave the C9-N10 bond of the folates of pool 1, pools 1 + 2, and pools 1 + 2 + 3. The cleaved pools are quantitated as the Bratton-Marshall azo dyes of p-aminobenzoylglutamate. The uncleaved pools are converted to Bratton-Marshall-negative products. Pool 1 is determined by converting pool 2 to 4a-hydroxy-5-methyltetrahydrofolic acid and pool 3 to N10-formylfolic acid, both Bratton-Marshall negative, by 10% hydrogen peroxide oxidation at pH 6.0. Pools 1 + 2 are cleaved with 0.015% hydrogen peroxide and 0.1% potassium permanganate at pH 9.0 which convert the N5-methyltetrahydrofolic acid to the acid-cleavable N5-methyl-dihydrofolic acid. Pool 3 oxidizes to the Bratton-Marshall-negative N10-formylfolic acid. Pools 1 + 2 + 3 are cleaved by first reducing pool 3 to N5-methyltetrahydrofolic acid with sodium borohydride followed by oxidation at pH 9.0 to its acid-labile dihydro form. Determination of the poly-γ-glutamyl chain length of each pool is possible by chromatographing the azo-p-aminobenzoylpolyglutamates with authentic synthetic markers.  相似文献   

7.
The N6-alkyladenosines and 2-methylthio-N6-alkyladenosines are the most common modified adenosine nucleosides and transfer ribonucleic acids (tRNA) are particularly rich in these modified nucleosides. They are present at position 37 of the anticodon arm and the contribution of these hypermodified nucleosides to codon–anticodon interactions, as well as translation, are significant, although not fully understood. Herein we described a new chemical synthesis method of the oligoribonucleotides containing N6-alkyladenosines and 2-methylthio-N6-alkyladenosines via post-synthetic modifications of precursor oligoribonucleotides. To obtain oligoribonucleotides containing N6-alkyladenosines, the precursor oligoribonucleotide carrying 6-methylthiopurine riboside residue was used, whereas for the synthesis of oligoribonucleotides containing 2-methylthio-N6-alkyladenosines the precursor oligoribonucleotide carrying the 2-methylthio-6-chloropurine riboside was applied. Among the modified oligoribonucleotides of different length and secondary structures, there were several containing naturally occurring modified nucleosides such as: N6-isopentenyladenosine (i6A), N6-methyladenosine (m6A), 2-methylthio-N6-isopentenyladenosine (ms2i6A), and 2-methylthio-N6-methyladenosine (ms2m6A), as well as several unnaturally modified adenosine derivatives.  相似文献   

8.
To evaluate the denitrification abilities of many Bradyrhizobium field isolates, we developed a new 15N-labeled N2 detection methodology, which is free from interference from atmospheric N2 contamination. 30N2 (15N15N) and 29N2 (15N14N) were detected as an apparent peak by a gas chromatograph equipped with a thermal conductivity detector with N2 gas having natural abundance of 15N (0.366 atom%) as a carrier gas. The detection limit was 0.04% 30N2, and the linearity extended at least to 40% 30N2. When Bradyrhizobium japonicum USDA110 was grown in cultures anaerobically with 15NO3, denitrification product (30N2) was detected stoichiometrically. A total of 65 isolates of soybean bradyrhizobia from two field sites in Japan were assayed by this method. The denitrification abilities were partly correlated with filed sites, Bradyrhizobium species, and the hup genotype.  相似文献   

9.
6-N-[3-3H]Trimethyl-dl-lysine was synthesized from 6-N-acetyl-l-lysine by the following chemical scheme: 6-N-acetyl-l-lysine → 2-keto-6-N-acetylcaproic acid → 2-[3-3H]keto-6-N-acetylcaproic acid → 2-[3-3H]keto-6-N-acetylcaproic acid oxime → 6-N-[3-3H]acetyl-dl-lysine → dl-[3-3H]lysine → 2-N-[3-3H]formyl-dl-lysine → 2-[3-3H]formyl-6-N-trimethyl-dl-lysine → 6-N-[3-3H]trimethyl-dl-lysine. Using a 70% ammonium sulfate fraction obtained from a high-speed rat kidney supernatant, the cosubstrate and cofactor requirements for 6-N-trimethyl-l-lysine hydroxylase activity as measured by tritium release from 6-N-[3-3H]trimethyl-dl-lysine were: α-ketoglutarate, ferrous ions, l-ascorbate, and oxygen, with added catalase showing a slight but distinct stimulatory effect. On incubation with the crude rat kidney preparation, the release of tritium from 6-N-[3-3H]trimethyl-dl-lysine was linear with both time of incubation and protein concentration. Hydroxylation of 6-N-trimethyl-l-lysine, as measured by tritium release from the labeled substrate, was examined in rat kidney, heart, liver, and skeletal muscle tissues, and found to be most active in the kidney.  相似文献   

10.
Sirtuins catalyze the NAD+ dependent deacetylation of Nε-acetyl lysine residues to nicotinamide, O′-acetyl-ADP-ribose (OAADPR) and Nε-deacetylated lysine. Here, an easy-to-synthesize Ac-Ala-Lys-Ala sequence has been used as a probe for the screening of novel Nε-modified lysine containing inhibitors against SIRT1 and SIRT2. Nε-Selenoacetyl and Nε-isothiovaleryl were the most potent moieties found in this study, comparable to the widely studied Nε-thioacetyl group. The Nε-3,3-dimethylacryl and Nε-isovaleryl moieties gave significant inhibition in comparison to the Nε-acetyl group present in the substrates. In addition, the studied Nε-alkanoyl, Nε-α,β-unsaturated carbonyl and Nε-aroyl moieties showed that the acetyl binding pocket can accept rather large groups, but is sensitive to even small changes in electronic and steric properties of the Nε-modification. These results are applicable for further screening of Nε-acetyl analogues.  相似文献   

11.
The N6-alkyladenosines and 2-methylthio-N6-alkyladenosines make up over half of the population of all naturally modified adenosines and they are present in the transfer ribonucleic acids (tRNA) at position 37. We measured effects of N6-alkyladenosines and 2-methylthio-N6-alkyladenosines on the thermodynamic stability of RNA duplexes containing a U-AMod base pair at internal and terminal duplex positions, as well as containing modified adenosines as a 3′-terminal unpaired nucleotide. Beside naturally modified adenosines such as N6-isopentenyladenosine (i6A), N6-methyladenosine (m6A), 2-methylthio-N6-isopentenyladenosine (ms2i6A) and 2-methylthio-N6-methyladenosine (ms2m6A), we studied several artificial modifications to evaluate the steric and electronic effects of N6-alkyl substituents. Moreover, some N6-alkyladenosines and 2-methylthio-N6-alkyladenosines were placed in hairpins at positions corresponding to nucleotide 37 of the tRNA anticodon arm, and the thermodynamic stability of those hairpins was studied. The stability of the modified RNA hairpins was measured in standard melting buffer containing 1 M sodium chloride as well as in physiological buffer containing 10 mM magnesium chloride and 150 mM potassium chloride. The results obtained indicate that the nature of the adenosine modification and the position of U-AMod base pairs within the duplex influence the thermodynamic stability of RNA duplexes. For most of the modification, the destabilization of duplexes was observed. Moreover, we found that the buffer composition and the structure of the modified adenosine very significantly affect the thermodynamic stability of RNA.  相似文献   

12.
This report describes synthesis and evaluation of cationic complexes, [99mTc(CO)3(L)]+ (L = N-methoxyethyl-N,N-bis[2-(bis(3-ethoxypropyl)phosphino)ethyl]amine (L1), N-[(15-crown-5)-2-yl]-N,N-bis[2-(bis(3-ethoxypropyl)phosphino)ethyl]amine (L2) and N-[(18-crown-6)-2-yl]-N,N-bis[2-(bis(3-ethoxypropyl)phosphino)ethyl]amine (L3)) as potential radiotracers for heart imaging. Preliminary results from biodistribution studies in female adult BALB-c mice indicated that the cationic 99mTc(I)-tricarbonyl complex, [99mTc(CO)3(L2)]+, has a significant localization in the heart at 60 min post-injection. To understand the coordination chemistry of these bisphosphine ligands with the 99mTc(I)-tricarbonyl core, we prepared [Re(CO)3(L4)]Br (L4: N,N-bis[(2-diphenylphosphino)ethyl]methoxyethylamine) as a model compound. [Re(CO)3(L4)]Br has been characterized by elemental analysis, IR, ESI-MS, NMR (1H, 13C, 1H-1H COSY, and 1H-13C HMQC) methods, and X-ray crystallography. In solid state, [Re(CO)3(L4)]+ has a distorted octahedron coordination geometry with PNP occupying one facial plane. The chelator backbone adopts a “chair” conformation with phosphine-P atoms at equatorial positions and the amine-N at the apical site. In solution, [Re(CO)3(L4)]+ is able to maintain its cationic nature with no dissociation of carbonyl ligands or any of the three PNP donors.  相似文献   

13.
N6-(Δ2-Isopentenyl) adenosine antibodies were used for the isolation of free cytokinins and cytokinin-containing tRNAs from parts of Cucumis sativus L. var. Guntur seedlings and for the estimation of cytokinins in them. Immobilized N6-(Δ2-isopentenyl) adenosine antibodies retained tRNAs containing N6-(Δ2-isopentenyl) adenosine and N6-(4-hydroxy-3-methylbut-2-enyl) adenosine with equal efficiencies. There were at least five cytokinins in the free form in cucumber seedlings. N6-(4-Hydroxy-3-methylbut-2-enyl) adenosine, N6-(Δ2-isopentenyl) adenosine, and N6-(Δ2-isopentenyl) adenine were present at least to the extent of 80, 23, and 9 nanograms, respectively, in the cotyledons and 40, 6, and 3 nanograms, respectively, in the decotyledonated seedlings per gram of tissue. Only two cytokinins were found in the tRNAs of cucumber cotyledons, namely N6-(Δ2-isopentenyl) adenosine and N6-(4-hydroxy-3-methylbut-2-enyl) adenosine in amounts of 12 and 318 nanograms, respectively, per gram of tissue. Immunoaffinity chromatographic analysis of radiolabeled aminoacyl tRNAs from cucumber cotyledons showed that tRNAPhe and tRNATyr contained cytokinins whereas tRNAAla did not.  相似文献   

14.
Novel, thermally stable, dark-red to orange Pt02-N,N′-diazadiene)(η2-alkene) compounds have been synthesized in good yields from Pt0(COD)2 or Pt0(NBE)3, by stepwise substitution of the respective dienes or alkenes by an electron-poor alkene (dimethyl fumarate, maleic anhydride or fumaronitrile), followed by the appropriate diazadiene ligand in dry diethyl ether at 20 °C (diazadiene=various N,N′-disubstitued-1,4-diaza-1,3-dienes). The complex Pt(DBA)2 is less suited as a precursor for the synthesis of Pt02-N,N′-diazadiene)(η2-alkene) compounds. These zerovalent Pt(diazadiene)(η2-alkene) compounds constitute a useful category of starting materials for synthetic organoplatinum chemistry and catalysis.  相似文献   

15.
N2O reductase activity in soybean nodules formed with Bradyrhizobium japonicum was evaluated from N2O uptake and conversion of 15N-N2O into 15N-N2. Free-living cells of USDA110 showed N2O reductase activity, whereas a nosZ mutant did not. Complementation of the nosZ mutant with two cosmids containing the nosRZDFYLX genes of B. japonicum USDA110 restored the N2O reductase activity. When detached soybean nodules formed with USDA110 were fed with 15N-N2O, they rapidly emitted 15N-N2 outside the nodules at a ratio of 98.5% of 15N-N2O uptake, but nodules inoculated with the nosZ mutant did not. Surprisingly, N2O uptake by soybean roots nodulated with USDA110 was observed even in ambient air containing a low concentration of N2O (0.34 ppm). These results indicate that the conversion of N2O to N2 depends exclusively on the respiratory N2O reductase and that soybean roots nodulated with B. japonicum carrying the nos genes are able to remove very low concentrations of N2O.  相似文献   

16.
Etheno DNA adducts are a prevalent type of DNA damage caused by vinyl chloride (VC) exposure and oxidative stress. Etheno adducts are mutagenic and may contribute to the initiation of several pathologies; thus, elucidating the pathways by which they induce cellular transformation is critical. Although N2,3-ethenoguanine (N2,3-εG) is the most abundant etheno adduct, its biological consequences have not been well characterized in cells due to its labile glycosidic bond. Here, a stabilized 2′-fluoro-2′-deoxyribose analog of N2,3-εG was used to quantify directly its genotoxicity and mutagenicity. A multiplex method involving next-generation sequencing enabled a large-scale in vivo analysis, in which both N2,3-εG and its isomer 1,N2-ethenoguanine (1,N2-εG) were evaluated in various repair and replication backgrounds. We found that N2,3-εG potently induces G to A transitions, the same mutation previously observed in VC-associated tumors. By contrast, 1,N2-εG induces various substitutions and frameshifts. We also found that N2,3-εG is the only etheno lesion that cannot be repaired by AlkB, which partially explains its persistence. Both εG lesions are strong replication blocks and DinB, a translesion polymerase, facilitates the mutagenic bypass of both lesions. Collectively, our results indicate that N2,3-εG is a biologically important lesion and may have a functional role in VC-induced or inflammation-driven carcinogenesis.  相似文献   

17.
Humans are exposed to both endogenous and exogenous N-nitroso compounds (NOCs), and many NOCs can be metabolically activated to generate a highly reactive species, diazoacetate, which is capable of inducing carboxymethylation of nucleobases in DNA. Here we report, for the first time, the chemical syntheses of authentic N6-carboxymethyl-2′-deoxyadenosine (N6-CMdA) and N4-carboxymethyl-2′-deoxycytidine (N4-CMdC), liquid chromatography–ESI tandem MS confirmation of their formation in calf thymus DNA upon diazoacetate exposure, and the preparation of oligodeoxyribonucleotides containing a site-specifically incorporated N6-CMdA or N4-CMdC. Additionally, thermodynamic studies showed that the substitutions of a dA with N6-CMdA and dC with N4-CMdC in a 12-mer duplex increased Gibbs free energy for duplex formation at 25°C by 5.3 and 6.8 kcal/mol, respectively. Moreover, primer extension assay revealed that N4-CMdC was a stronger blockade to Klenow fragment-mediated primer extension than N6-CMdA. The polymerase displayed substantial frequency of misincorporation of dAMP opposite N6-CMdA and, to a lesser extent, misinsertion of dAMP and dTMP opposite N4-CMdC. The formation and the mutagenic potential of N6-CMdA and N4-CMdC suggest that these lesions may bear important implications in the etiology of NOC-induced tumor development.  相似文献   

18.
Qualitative differences in the active center of rat trypsins 1 and 2 resulted in different ratios of Kcat for N1-tosyl-l-arginine methyl ester vs Kcat for N1-benzoyl-l-arginine ethyl ester. These ratios were 2.5 for trypsin 1 and 1.2 for trypsin 2.Substrate activation with N1-tosyl-l-arginine methyl ester enhanced the catalytic rate constant of rat trypsin 1 2.5-fold and that of rat trypsin 2 only 1.5-fold. The increase in the catalytic rate constant found with N1-benzoyl-l-arginine ethyl ester was the same (1.5-fold) for both trypsins. Consequently, at 20 mm substrate concentration, trypsin 1 catalyzed the esterolysis of N1-tosyl-l-arginine methyl ester 4.5 times faster than that of N1-benzoyl-l-arginine ethyl ester, while trypsin 2 was only 1.3 times more efficient with the first substrate.Furthermore, the activation of both rat enzymes by N-acetyl-l-tyrosine ethyl ester was even more effective than that obtained with the two cationic esters; the maximum rates of hydrolysis of this neutral substrate by trypsins 1 and 2 were enhanced 120- and 50-fold, respectively, by high concentrations of N-acetyl-l-tyrosine ethyl ester.  相似文献   

19.
β-Carbolines constitute a vast group of indole alkaloids and exhibit various biological actions. The objective of this study was to investigate the structure–activity relationships of β-carboline derivatives on in vitro inhibitory effects against clinically relevant microorganisms. A series of β-carboline dimers and their N2-alkylated analogues were therefore prepared and evaluated for their antimicrobial effects. Among these, a dimeric 6-chlorocarboline N2-benzylated salt exerted potent activity against Staphylococcus aureus at MICs of 0.01–0.05?μmol/mL. Our work highlights that N1-N1 dimerization and N2-benzylation significantly enhanced the antimicrobial effects of compounds.  相似文献   

20.
Nickel(II) complexes of N,N′-dimethyl-N,N′-bis(pyridyl-2yl-methyl)ethylene-diamine (L1), N,N′-dimethyl-N,N′-bis(pyridyl-2-ylmethyl)-1,2-diaminopropane (L2) and N,N′-dimethyl-N,N′-bis(pyridyl-2-ylmethyl)-1,3-diaminopropane (L3) were prepared and their spectroscopic and redox properties studied. The distorted octahedral structure was determined for [NiL3ClCH3OH](ClO4) by using X-ray crystallography. The electronic spectral behavior of the complexes at different pHs was analyzed; it is shown that a new band grew at the expense of the other band intensity in acid media. The redox properties of ligands and their complexes show the peaks of Ni(II) → Ni(III) and Ni(II) → Ni(0) as these were detected at low concentration while Ni(II) → Ni(I) process was detectable clearly at high concentration. Furthermore, the interaction studies of 2-mercaptoethanesulfonic acid as a simulator of coenzyme M reductase (CoM) with NiN4 chromophores are discussed.  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