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 31 毫秒
1.
To reveal the mechanisms of previously reported lung-specific DNA strand scissions in murine after oral administration of dimethylarsinic acid (DMAA), a main metabolite of inorganic arsenics in mammals, the ultimate substance causing DNA lesion was investigated using dimethylarsine which was a further metabolite of DMAA. The alkaline elution assay using 3H-labeled DNA showed that a major portion of the strand breaks was not suppressed by SOD and catalase, suggesting an ultimate substance other than active oxygen participated in the DNA damage. By ESR analysis, a radical estimated to be (CH3)2AsOO. was detected as a reaction product of dimethylarsine and molecular oxygen. This peroxyl radical, rather than active oxygen, was assumed to play a major role in DNA damage.  相似文献   

2.
We have proposed that oral administration of dimethylarsinic acid (DMA), a metabolite of inorganic arsenics in mammals, rather than inorganic arsenics themselves, promotes lung and skin tumors by way of the metabolic production of free radicals such as dimethylarsenic peroxy radical [(CH(3))(2)AsOO*]. The purpose of the present study was to examine if dimethylarsenic has the ability to induce oxidative damage. 8-oxo-2'-deoxyguanosine (8-oxodG) was used as a biomarker of DNA oxidation. The oral administration of DMA enhanced significantly the amounts of 8-oxodG specifically in the target organs (skin, lung, liver, and urinary bladder) of arsenic carcinogenesis and also in urine, whereas arsenite did not. The dimethylarsenics thus may play an important role in arsenic carcinogenesis through the induction of oxidative damage, particularly of base oxidation.  相似文献   

3.
Hanaoka  Ken'ichi  Tagawa  Shoji  Kaise  Toshikazu 《Hydrobiologia》1992,235(1):623-628
Two growth media containing arsenobetaine [(CH3)3 As+ CH2COO] were mixed with coastal marine sediments, the latter providing a source of microorganisms. The mixtures were kept at 25 °C in the dark and shaken for several weeks under an atmosphere of air. The disappearance of arsenobetaine and the appearance of two metabolites were followed by HPLC. The HPLC-retention time of the first metabolite agreed with that of trimethylarsine oxide [(CH3)3AsO]. The second metabolite was identified as arsenate (As(V)) using hydride generation/cold trap/GC MS analysis and thin layer chromatography. This is the first scientific evidence showing that arsenobetaine is degraded by microorganisms to inorganic arsenic via trimethylarsine oxide. The degradation of arsenobetaine to inorganic arsenic completes the marine arsenic cycle that begins with the methylation of inorganic arsenic on the way to arsenobetaine.  相似文献   

4.
A detailed doublet potential energy surface for the reaction of CH with CH3CCH is investigated at the B3LYP/6-311G(d,p) and G3B3 (single-point) levels. Various possible reaction pathways are probed. It is shown that the reaction is initiated by the addition of CH to the terminal C atom of CH3CCH, forming CH3CCHCH 1 (1a,1b). Starting from 1 (1a,1b), the most feasible pathway is the ring closure of 1a to CH3–cCCHCH 2 followed by dissociation to P 3 (CH3–cCCCH+H), or a 2,3 H shift in 1a to form CH3CHCCH 3 followed by C–H bond cleavage to form P 5 (CH2CHCCH+H), or a 1,2 H-shift in 1 (1a, 1b) to form CH3CCCH2 4 followed by C–H bond fission to form P 6 (CH2CCCH2+H). Much less competitively, 1 (1a,1b) can undergo 3,4 H shift to form CH2CHCHCH 5. Subsequently, 5 can undergo either C–H bond cleavage to form P 5 (CH2CHCCH+H) or C–C bond cleavage to generate P 7 (C2H2+C2H3). Our calculated results may represent the first mechanistic study of the CH + CH3CCH reaction, and may thus lead to a deeper understanding of the title reaction.  相似文献   

5.
DNA damage induced by administration of dimethylarsinic acid (DMAA) to rats and mice was investigated. At 12 h after administration of DMAA, DNA single-strand breaks were induced markedly in lung. The majority of dimethylarsine, one of the main metabolites, in the expired air was excreted within 6–18 h after administration of DMAA to rats. In vitro experiments using nuclei isolated from lung of mice indicated that DNA strand breaks were caused by dimethylarsine. Furthermore, the strand breaks after exposure to dimethylarsine were reduced in the presence of catalase and/or superoxide dismutase. These results strongly suggest that the strand breaks are induced not by dimethylarsine itself but by active oxygen, e.g., O 2 ? and ·OH, produced both by dimethylarsine and molecular oxygen. When DNA was exposed to dimethylarsine, thiobarbituric acid (TBA)-reactive intermediates andcis-thymine glycol were produced. Dimethylarsine appears to induce DNA damage by the mechanism similar to the damage produced by ionizing radiation.  相似文献   

6.
A cobalt(II) complex [Co((bnzim)2CO)Cl2] (1) (where (bnzim)2CO; bis(benzimidazol-2-yl)methanone), was obtained by reacting cobalt(II) chloride with bis(benzimidazol-2-yl)methane ((bnzim)2CH2) in CH3OH/CH3CN and characterized using EA, IR, XRD and X-ray single crystal diffraction. The process of ketonization of methylene of (bnzim)2CH2 was investigated using spectroscopy and electrochemistry method. The experimental results reveal that there were new species forming in the reaction system and the formation of new species was related to the dissolving oxygen in the mixture. Oxygenation mechanism was thus proposed, in which oxygen atom transfer from a superoxo intermediate species {((bnzim)2CH2)Co(III)-O-O} to the methylenic carbon of (bnzim)2CH2, and then goes O-O hemolytic cleavage and hydrogen migration to give intermediate species [Co((bnzim)2CHOH)]2+. The ketonization was achieved after [Co((bnzim)2CHOH)]2+repeating the process.  相似文献   

7.
Hiroshi Maeda  Kenta Ooi 《Biopolymers》1981,20(7):1549-1563
The β-coil transition of poly(S-carboxymethyl-L -cysteine) (poly[Cys(CH2CO2H)]) and poly(S-carboxyethyl-L -cysteine) (poly[Cys((CH2)2CO2H)]) was followed by CD, potentiometric titration, and viscosity in the absence of added salt. These different properties give consistent results for poly[Cys((CH2)2CO2H)]. The CD spectra of poly[Cys(CH2CO2H)] change considerably with the degree of neutralization α even for a low-molecular-weight sample incapable of forming the β-structure. Because of the superposition of this additional effect, the dependence of CD on α is inconsistent with titration data for the case of poly[Cys(CH2CO2H)], particularly when the nπ transition is used to follow the β-coil transition. The change of CD inherent to the β-coil transition is characterized by an isodichroic point: 215 nm for poly[Cys((CH2)2CO2H)] and 218 nm for poly[Cys(CH2CO2H)]. A criterion supporting the stacking of the pleated sheet is suggested based on the isodichroic point.  相似文献   

8.
H Inoue  T Izumi 《Biopolymers》1976,15(4):797-812
The preferential binding of solvent components with a nonionic homopolypeptide, poly(N5-(3-hydroxypropyl)-L -glutamine), ([Gln((CH2)3OH)]n), has been determined in water/dioxane mixtures using differential refractometry. The degree of preferential binding was calculated from the difference between the refractive index increments of [Gln((CH2)3OH)]n obtained from experiments carried out under two conditions: experiments where the molality of dioxane was kept identical in both compartments of the differential cell, and experiments where the chemical potential was kept identical. The polypeptide was preferentially hydrated between 10 and 70 wt % of dioxane; the amount of preferential hydration per gram of the mixed solvent increases monotonically (with a plateau region between 40 and 60 wt %) with the dioxane concentration. A monotonic increase was also observed in the degree of helicity of the polypeptide. The absolute amounts of water and dioxane bound by [Gln((CH2)3OH)]n were investigated in the frozen state by the method of nuclear magnetic resonance. Hydration was measured using a mixed solvent, water/dioxane-d8; dioxane solvation was measured using a mixed solvent, dioxane/D2O. The polypeptide binds about 0.35 g of water per g of the polymer in aqueous solution, and hydration decreases gradually with an increase in dioxane concentration. On the other hand, the amount of dioxane solvation increases to 0.04 g per g of the polymer in the dioxane concentration range between 0 and 20 wt %, and then levels off. The rapid increase in solvation is observed before the conformational transition from random coil to α-helix occurs in [Gln((CH2)3OH)]n. The dependence of the preferential and absolute binding of solvent components to [Gln((CH2)3OH)]n on dioxane concentration and the conformational change in the homopolypeptide suggest that addition of dioxane to aqueous solutions induces lowering of water activity and that the helical structure of the polypeptide is enhanced by the formation of intrachain hydrogen bonds. The validity of the frozen method is also discussed.  相似文献   

9.
The International Agency for Research on Cancer (IARC) has concluded that dimethylarsinic acid [(CH3)2AsO(OH), DMAV], a main metabolite of inorganic arsenic, is responsible for carcinogenesis in urinary bladder and lung in rodents, and various modes of carcinogenic action have been proposed. One theory concerning the mode of action is that the biotransformation of dimethylarsinous acid [(CH3)2AsOH, DMAIII] from DMAV plays an important role in the carcinogenesis by way of reactive oxygen species (ROS) production. Furthermore, dimethylmonothioarsinic acid [(CH3)2AsS(OH), DMMTAV], a metabolite of DMAV, has also been noted because of its higher toxicity. However, the metabolic mechanisms of formation and disappearance of DMAIII and DMMTAV, and their toxicity are not fully understood. Thus, the purpose of the present study was to clarify the mechanism of metabolic formation of DMMTAV and DMAV from DMAIII. The in vitro transformation of arsenicals by treatment with liver homogenate from rodents and sulfur transferase was detected by HPLC-ICP-MS and HPLC-tandem MS. DMMTAV is produced from DMAIII but not DMAV by cellular fractions from mouse liver homogenates and by rhodanese from bovine liver in the presence of thiosulfate, a sulfur donor. Not only DMMTAV thus produced but also DMAIII are re-converted into DMAV by an in vitro addition of S9 mix. These findings indicate that the metabolic process not only of DMAIII to DMAV or DMMTAV but also of DMMTAV to DMAV consists of a complicated mode of interaction between monooxygenase including cytochrome P450 (CYP) and/or sulfur transferase.  相似文献   

10.
《Free radical research》2013,47(5):453-463
The question whether hydroxyl free radicals are formed in the reactions of divalent iron complexes Fe(II)L; L = edta; hedta; tcma (tcma = l-acetato-l,4,7-triazacyclononane) with hydrogen peroxide in neutral and slightly acidic solutions was studied by using the β elimination reaction as an assay for the formation of hydroxyl free radicals, OH. The results show that at pH<5.5 the iron(II)peroxide intermediate complex decomposes rapidly to yield free hydroxyl radicals for L=edta and hedta. This is in contrast to the mechanism of the corresponding Fe(II)nta peroxide complex, which probably decomposes to form Fe(IV)nta which then reacts with organic substrates to yield aliphatic free radicals. Thus, the non-participating ligand L has an appreciable effect on the mechanism of reaction of the metal center with hydrogen peroxide. Blank experiments using ionizing radiation as the source of CH2CR(CH3)OH, R = H or CH3 radicals indicate that when L = tcma intermediates of the type LFeIII-CH2CR(CH3)OHaq are formed, but their major mode of decomposition is not the β elimination reaction. Thus, the present assay for the formation of hydroxyl free radicals by the Fenton Reaction does not fit the latter system.  相似文献   

11.
Dimethylarsine and trivalent dimethylated arsenic, metabolites of inorganic arsenics, have received considerable attention in current research because of their biological activities. We attempted to determine the appearance of micronucleated reticulocytes (MNRETs) in mouse peripheral blood following intraperitoneal administration of dimethylarsinous iodide (DMI) and trimethylarsine (TMA), model compounds of trivalent dimethylated arsenic and dimethylarsine, respectively. A significant increase in the number of MNRETs was observed with TMA, but not with DMI. Furthermore, MNRETs only appeared with 10.6 mg/kg of dimethylarsinic acid (DMA) following its co-injection with reduced glutathione (GSH). These results suggest that micronucleus formation may need further metabolic reduction of trivalent dimethylated arsenic, i.e. the production of dimethylarsine, by an excess amount of GSH. Meanwhile, the increase in MNRETs by administration of arsenite at 7.6 mg/kg, an equivalent dose to DMA as As, was remarkably diminished by co-administration with GSH. These results indicate that GSH plays an important role in the genotoxic process of arsenics, particularly by dimethylated arsenic.  相似文献   

12.
Using dihydrogendisulphide (H2S2), dimethyl‐ ((CH3)2S2), and diethyldisulphide ((CH3CH2)2S2)as model molecules, theoretical ECD, VCD, and ROA spectra of nonplanar disulphides were calculated by DFT methods. Most of the calculated electronic and vibrational chiroptical features suffer an equivocal relation between calculatedsigns of ECD, VCD, or ROA and the sense of disulphide nonplanarity as noted earlier for low‐lying ECD bands. This is a consequence of local C2 symmetry of a disulphide group causing most electronic and vibrational transitions to occur as pairs falling to alternative A, B symmetry species, which become degenerate and switch their succession (and consequently the observed chiroptical sign pattern) at the energetically most favorable perpendicular conformation. According to present calculations, the key to resolving this ambiguity may involve the S? S stretching vibrational mode at ~500 cm?1. The relation of signs of the relevant VCD and ROA features to sense of disulphide chirality seems simpler and less ambiguous. The right‐handed arrangement of the S? S group (0 < χS? S < 180°) results in mostly negative VCD signals. Although relation to ROA still suffers some ambiguity, it gets clearer along the series H2S2–(CH3)2S2–(CH3CH2)2S2. ROA is also attractive for the analysis of disulphide‐containing peptides and proteins, because applying it to aqueous solutions is not problematic. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

13.
This article compiles available data and presents an approach for predicting human intakes of inorganic arsenic (Asi), monomethylarsonic acid (MMA), and dimethylarsinic acid (DMA) from marine, estuarine, and freshwater seafood when only total arsenic (Astot) concentrations are reported. Twenty studies provided data on total arsenic (Astot) and Asi. Mean Asi concentrations were approximately 10 to 20 ng/g wet weight (ww) in freshwater, anadromous, and marine fish, whereas crustaceans and molluscs had mean Asi concentrations of 40 to 50 ng/g ww. Thirteen studies provided data for MMA and DMA. MMA was seldom detected, whereas DMA averaged 10 ng/g ww in freshwater fish, and 45 to 95 ng/g ww in anadromous fish, marine fish, crustaceans, and molluscs. There was little correlation between Astot concentrations and Asi concentrations; however, when only Astot data are available to assess health risks from arsenic in seafood, these data could support conservative, upper end estimates of the percent of Astot likely to be Asi. For marine and estuarine fish, and crustaceans and molluscs 2–3% of Astot was Asi at the 75th percentile of the dataset. For freshwater fish Asi was 10% of Astot at the 75th percentile. Due to the nonlinearity and low carcinogenic potency of DMA, the reported DMA concentrations should not contribute substantially to potential health risks from arsenic in seafood.  相似文献   

14.
Abstract

As a series of X-ray structural studies of Z-DNA Polyamine complex, the crystal structure of Z-DNA hexamer, d(CG)3, duplex complexed with a synthetic polyamine, N,N′-bis(2-aminoethyl)-1,2-ethanediamine, NH2-(CH2)2-NH-(CH2)2-NH-(CH2)2-NH2 [PA(222)], has been determined.  相似文献   

15.
Complex (S,S)-[Pd{C6H4(CH2CHNH2CO2CH2CH3)}(μ-Br)]2 (3) was prepared following the method by Vicente and Saura-Llamas (Organometallics 26:2768–2776, 2007), by the reaction of l-ethylphenylalanate and Pd(OAc)2 in 1:1 molar ratio under acetonitrile heating conditions and subsequently treating with NaBr. In addition, the cleavage of halogeno-bridge of the complex 3 via nucleophilic attack of some neutral ligands such as triphenylphosphine, pyridine, 2,4,6-trimethylpyridine and piperidine were investigated and the corresponding complexes (S)-[Pd{C6H4(CH2CHNH2CO2CH2CH3)(Y)(Br)}] (4a–f) were obtained in moderate yields. The six-member orthopalladated complexes were characterized by 1H-NMR, 31P-NMR, FT-IR and elemental analysis techniques.  相似文献   

16.
As we reported previously, hypochlorite interacting with organic hydroperoxides causes their decomposition ((1995) Biochemistry (Moscow), 60, 1079-1086). This interaction was supposed to be a free-radical process and serve as a source of free radicals initiating lipid peroxidation (LP). The present study is the first attempt to detect and identify free radicals produced in the reaction of hypochlorite with tert-butyl hydroperoxide, (CH3)3COOH, which we have used as an example of organic hydroperoxides. We have used a direct method for free radical detection, EPR of spin trapping, and the following spin traps: N-tert-butyl--phenylnitrone (PBN) and -(4-pyridyl-1-oxyl)-N-tert-butylnitrone (4-POBN). When hypochlorite was added to (CH3)3COOH in the presence of a spin trap, an EPR spectrum appeared representing a superposition of two signals. One of them belonged to a spin adduct formed as a result of direct interaction of hypochlorite with the spin trap (hyperfine splitting constants were: H H = 0.148 mT; aN = 1.537 mT; and HPP = 0.042 mT for 4-POBN and H = 0.190 mT; aN = 1.558 mT; and HPP = 0.074 mT for PBN). The other signal was produced by hypochlorite interactions with (CH3)3COOH itself (hyperfine splitting constants were: H = 0.233 mT; aN = 1.484 mT; HPP = 0.063 mT and H = 0.360 mT; aN = 1.547 mT; HPP = 0.063 mT for 4-POBN and PBN, respectively). Comparison of spectral characteristics of this spin adduct with those of tert-butoxyl or tert-butyl peroxyl radicals produced in known reactions of (CH3)3COOH with Fe2+ and Ce4+, respectively, showed that the radical (CH3)3COO. is produced from the interaction of hypochlorite with (CH3)3COOH. Like Ce4+ but not Fe2+, hypochlorite addition to (CH3)3COOH was accompanied by a bright flash of chemiluminescence characteristic of the reactions in which peroxyl radicals are produced. Thus, all these results suggest peroxyl radical production in the reaction of hypochlorite with hydroperoxide. This reaction is one of the most possible ways for the initiation of free-radical LP that occurs in vivo, when hypochlorite interacts with unsaturated lipids comprising natural protein–lipid complexes, such as lipoproteins and biological membranes.  相似文献   

17.
The 16-mer peptide nucleic acid sequence H-A GAT CAT GCC CGG CAT-Lys-NH2 (1), which is complementary to the translation start region of the N-myc oncogene messenger RNA, was synthesized and conjugated to a pyrazolyl diamine bifunctional chelator (pz). The novel conjugate pz-A GAT CAT GCC CGG CAT-Lys-NH2 (2) was labeled with technetium tricarbonyl, yielding quantitatively the complex fac-[99mTc(CO)33-pz-A GAT CAT GCC CGG CAT-Lys-NH2)]2+ (4). Complex 4 was obtained with high radiochemical purity and high specific activity, revealing high stability in human serum and in cell culture medium. The identity of 4 was confirmed by comparing its reversed-phase high performance liquid chromatography profile with that of the rhenium analog fac-[Re(CO)33-pz-A GAT CAT GCC CGG CAT-Lys-NH2)]2+ (3), prepared by conjugation of fac-[Re(CO)3(3,5-Me2pz(CH2)2N((CH2)3COOH)(CH2)2NH2)]+ to 1, using solid-phase techniques. UV melting experiments of 1 and 3 with the complementary DNA sequence led to the formation of stable duplexes, indicating that the conjugation of 1 to the pyrazolyl chelator and to the metal fragment fac-[M(CO)3]+ did not affect the recognition of the complementary sequence as well as the duplex stability. For a first screening, SH-SY5Y human neuroblastoma cells, which express N-myc, were treated with 4. The results show that 4 internalizes (7% of the activity goes into the cells, after 4 h at 37 °C), presenting also a relatively high cellular retention (only 40% of internalized activity is released from the cells after 5 h). An erratum to this article can be found at  相似文献   

18.
In this study, the reactions of N-acetyl-L-methionine (AcMet) with [{trans-PtCl(NH3)2}2-μ-H2N(CH2)6NH2](NO3)2 (BBR3005: 1,1/t,t 1) and its cis analog [{cis-PtCl(NH3)2}2-μ-{H2N(CH2)6NH2}]Cl2 (1,1/c,c 2) were analyzed to determine the rate and reaction profile of chloride substitution by methionine sulfur. The reactions were studied in PBS buffer at 37°C by a combination of multinuclear (195Pt, {1H-15N} HSQC) magnetic resonance (NMR) spectroscopy and electrospray ionization time of flight mass spectrometry (ESITOFMS). The diamine linker of the 1,1/t,t trans complex was released as a result of the trans influence of the coordinated sulfur atom, producing trans-[PtCl(AcMet)(NH3)2]+ (III) and trans-[Pt(AcMet)2(NH3)2]2+ (IV). In contrast the cis geometry of the dinuclear compound maintained the diamine bridge intact and a number of novel dinuclear platinum compounds obtained by stepwise substitution of sulfur on both platinum centers were identified. These include (charges omitted for clarity): [{cis-PtCl(NH3)2}-μ-NH2(CH2)6NH2-{cis-Pt(AcMet)(NH3)2}] (V); [{cis-Pt(AcMet)(NH3)2}2-μ-NH2(CH2)6NH2] (VI); [{cis-PtCl(NH3)2}-μ-NH2(CH2)6NH2-{PtCl(AcMet)NH3] (VII); [{PtCl(AcMet)(NH3)}2-μ-NH2(CH2)6NH2] (VIII); [{trans-Pt(AcMet)2(NH3)}-μ-NH2(CH2)6NH2-{PtCl(AcMet)(NH3)] (IX) and the fully substituted [{trans-Pt(AcMet)2(NH3)}2-μ-{NH2(CH2)6NH2] (X). For both compounds the reactions with methionine were slower than those with glutathione (Inorg Chem 2003, 42:5498–5506). Further, the 1,1/c,c geometry resulted in slower reaction than the trans isomer, because of steric hindrance of the bridge, as observed previously in reactions with DNA and model nucleotides.  相似文献   

19.
In order to study the genotoxicity of arsenics, we focused our attention on dimethylarsinic acid (DMAA) which was a main metabolite of inorganic arsenics in mammals. ICR mice were orally administered DMAA-Na (1500mg/kg). DNA single-strand breaks occurred specifically in lung at 12h after administration. An in vitro experiment indicated that the breaks were not caused directly by DMAA but by dimethylarsine, a further metabolite of DMAA. Furthermore, the dimethylarsine-induced breaks were diminished by the addition of SOD and catalase, suggesting that active oxygen produced by dimethylarsine was involved in the induction of DNA damage.  相似文献   

20.

Background and aims

Efficient accumulation of arsenic (As) in rice (Oryza sativa L.) poses a potential health risk to rice consumers. The aim of this study was to investigate the mechanisms of uptake, transport and distribution of inorganic arsenic (Asi) and dimethylarsinic acid (DMA) in rice plants.

Methods

Rice was exposed to Asi (As(V)) and DMA in hydroponics. High-performance liquid chromatography inductively coupled plasma mass spectrometry (HPLC-ICP-MS) and synchrotron X-ray fluorescence (SXRF) microprobe were used to determine As concentration and the in situ As distribution.

Results

DMA induced abnormal florets before flowering and caused a sharp decline in the seed setting rate after flowering compared to Asi. Rice grains accumulated 2-fold higher DMA than Asi. The distribution of Asi concentration (root?>?leaf?>?husk?>?caryopsis) in As(V) treatments was different from that of the DMA concentration (caryopsis?>?husk?>?root?≥?leaf) in DMA treatments. SXRF showed that Asi mainly accumulated in the vascular trace of caryopsis with limited distribution to the endosperm, whereas DMA was observed in both tissues.

Conclusions

DMA tended to accumulate in caryopsis and induced higher toxicity to the reproductive tissues resulting in markedly reduced grain yield, whereas Asi mainly remained in the vegetative tissues and had no significant effect on yield. DMA is more toxic than Asi to the reproductive tissues when both of them are at similar concentrations in nutrient solution.  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