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.
The temperature dependence of the enantioselectivity of Candida antarctica lipase B for 3-hexanol, 2-butanol, 3-methyl-2-butanol, 3,3-dimethyl-2-butanol, and 1-bromo-2-butanol revealed that the differential activation entropy, deltaR-SdeltaS, was as significant as the differential activation enthalpy, DeltaR-SdeltaH, to the enantiomeric ratio, E. 1-Bromo-2-butanol, with isosteric substituents, displayed the largest deltaR-SdeltaS. 3-Hexanol displayed, contrary to other sec-alcohols, a positive deltaR-SdeltaS. In other words, for 3-hexanol the preferred R-enantiomer is not only favored by enthalpy but also by entropy. Molecular dynamics (MD) simulations and systematic search calculations of the substrate accessible volume within the active site revealed that the (R)-3-hexanol transition state (TS) accessed a larger volume within the active site than the (S)-3-hexanol TS. This correlates well with the higher TS entropy of (R)-3-hexanol. In addition, this enantiomer did also yield a higher number of allowed conformations, N, from the systematic search routines, than did the S-enantiomer. The substrate accessible volume was greater for the enantiomer preferred by entropy also for 2-butanol. For 3,3-dimethyl-2-butanol, however, neither MD-simulations nor systematic search calculations yielded substrate accessible volumes that correlate to TS entropy. Ambiguous results were achieved for 3-methyl-2-butanol.  相似文献   

2.
Esters of dimethylcarbamic acid are known to be poor substrates of acetylcholinesterase. They carbamylate the active catalytic site of the enzyme and the subsequent decarbamylation is a slow but measurable process. Similarly, acetylcholinesterase can be phosphonylated, and the dephosphonylation is extremely slow. Rapid hydrolysis of phosphonylated acetylcholinesterase can be brought about by oximes, but dealkylation of the phosphonyl group on the enzyme (known as ageing) renders the inhibited enzyme insensitive to oximes.

In the present work, decarbamylation of dimethylcarbamyl-acetylcholinesterase and ageing of isopropylmethylphosphonyl-acetylcholinesterase were studied at a physiological ionic strength (154 mM). Gallamine, d-tubocurarine and alcuronium accelerated reactivation of dimethylcarbamyl-acetylcholinesterase. Gallamine and tubocurarine enhanced the effect of the nucleophile 3,3-dimethyl-1-butanol on decarbamylation, and the interaction was synergistic in the case of gallamine. Gallamine and tubocurarine retarded ageing of isopropylmethylphosphonyl-acetylcholinesterase, whereas 3,3-dimethyl-1-butanol had no effect. Nevertheless 3,3-dimethyl-1-butanol enhanced the retarding effects of gallamine and tubocurarine.

All these effects, except the effects of 3,3-dimethyl-1-butanol on ageing, had been previously observed at low ionic strength, in which case the effects were more marked and were observed at lower concentrations of the drugs. The effects at low ionic strength have been attributed to binding of the drugs to a peripheral site on the enzyme with a consequent change in conformation at the active site, leading to altered kinetic properties. The present results suggest that such allosteric effects may persist at physiological ionic strength. There have been few indications previously that this is so, particularly in the case of solubilised acetylcholinesterase.  相似文献   


3.
C-6-based green leaf volatiles (GLVs) are signal molecules to herbivorous insects and play an important role in plant–herbivore interactions. How isomerization of GLVs affects insect’s olfactory response has been rarely tested. In laboratory and field experiments, we examined the effect of hexanol isomers on olfactory orientation of the spiraling whitefly, Aleruodicus dispersus Russell, a highly polyphagous pest. In a Y-tube oflactometer, we found that (±)-2-hexanol, 3-methyl-3-pentanol and 3,3-dimethyl-1-butanol significantly attracted female A. dispersus. The trap captures of 3,3-dimethyl-1-butanol were significantly more than that of (±)-2-hexanol and 3-methyl-3-pentanol, and its optimum concentration was 1 μ1/ml. We suggest that the anthropogenic compound 3,3-dimethyl-1-butanol can be exploited as a parakairomone (synthetic analogues of kairomone) to monitor and control adult A. dispersus.  相似文献   

4.
Thiamin phosphate synthase catalyzes the formation of thiamin phosphate from 4-amino-5-(hydroxymethyl)-2-methylpyrimidine pyrophosphate and 5-(hydroxyethyl)-4-methylthiazole phosphate. Several lines of evidence suggest that the reaction proceeds via a dissociative mechanism. The previously determined crystal structure of thiamin phosphate synthase in complex with the reaction products, thiamin phosphate and magnesium pyrophosphate, provided a view of the active site and suggested a number of additional experiments. We report here seven new crystal structures primarily involving crystals of S130A thiamin phosphate synthase soaked in solutions containing substrates or products. We prepared S130A thiamin phosphate synthase with the intent of characterizing the enzyme-substrate complex. Surprisingly, in three thiamin phosphate synthase structures, the active site density cannot be modeled as either substrates or products. For these structures, the best fit to the electron density is provided by a model that consists of independent pyrimidine, pyrophosphate, and thiazole phosphate fragments, consistent with a carbenium ion intermediate. The resulting carbenium ion is likely to be further stabilized by proton transfer from the pyrimidine amino group to the pyrophosphate to give the pyrimidine iminemethide, which we believe is the species that is observed in the crystal structures.  相似文献   

5.
Although indirect evidence has suggested that liver microsomal cytochrome P-450 can reductively dehalogenate several compounds to carbene metabolites, there has been no direct proof for the formation of these reactive species. We report in this paper that carbenes can be chemically trapped and identified as metabolites. For example, 1,1-dichloro-2,2,3,3-tetramethylcyclopropane was identified as a metabolite by gas chromatography mass spectrometry when carbon tetrachloride (CCl4) was incubated anaerobically with rat liver microsomes, NADPH and 2,3-dimethyl-2-butene. The reaction required NADPH and was inhibited by carbon monoxide. These findings show that cytochrome P-450 in rat liver microsomes can reductively metabolize CCl4 to dichloromethyl carbene (:CCl2) which can be trapped with 2,3-dimethyl-2-butene to form 1,1-dichloro-2,2,3,3-tetramethylcyclopropane. A similar approach may be used for the identification of carbene metabolites of other compounds.  相似文献   

6.
Two new compounds (E)-2-(5,7-dibromo-3,3-dimethyl-3,4-dihydroacridin-1(2H)-ylidene)hydrazinecarbothiomide (3) and (E)-2-(5,7-dibromo-3,3-dimethyl-3,4-dhihydroacridin-1(2H)-ylidene)hydrazinecarboxamide (4) were synthesized and evaluated for their anticholinesterase activities. In vitro tests performed by NMR and Ellman’s tests, pointed to a mixed kinetic mechanism for the inhibition of acetylcholinesterase (AChE). This result was corroborated through further docking and molecular dynamics studies, suggesting that the new compounds can work as gorge-spanning ligands by interacting with two different binding sites inside AChE. Also, in silico toxicity evaluation suggested that these new compounds can be less toxic than tacrine.  相似文献   

7.
Three carboxylated cyanine dyes, 2-[(1-butyl-3,3-dimethyl-5-carboxylindoline-2-ylidene)propenyl]-[1-butyl-3,3-dimethyl-7-(1-ethyl-1H-1,2,3-triazole-4-yl]-1H-benz[e]indolium iodide (), 2-[(1-butyl-3,3-dimethyl-5-carboxyl-indoline-2-ylidene)propenyl]-{1-butyl-3,3-dimethyl-7-[(4-piperidine-N-ethyl-1,8-naphthalimide)-1H-1,2,3-triazole-4-yl]}-1H-benz[e]indolium iodide (Cy2) and 2-[(1-butyl-3,3-dimethyl-5-carboxyl-indoline-2-ylidene)propenyl)-[1-butyl-3,3-dimethyl-7-{(4-piperidine-N-butyl-1,8-naphthalimide)-1H-1,2,3-triazole-4-yl}]-1H-benz[e]indolium iodide (Cy3), have been synthesized and characterized with regard to their structures and electrochemical properties. Upon adsorption onto a TiO(2) electrode, the absorption spectra of the three cyanine dyes are all broadened to both red and blue sides compared with their respective spectra in an acetonitrile and ethanol mixture. Cy2 and Cy3, containing a naphthalimide group, have stronger absorption intensities and broader absorption spectra than , which consequently leads to better light-to-electricity conversion properties. Among the three cyanine dyes, generated the highest photoelectric conversion yield of 4.80% (J(sc) = 14.5 mA cm(-2), V(oc) = 500 mV, FF = 0.49) under illumination with 75 mW cm(-2) white light from a Xe lamp.  相似文献   

8.
A novel water-soluble 2-[2-(2-chloro-3-{2-[3,3-dimethyl-5-sulfo-1-(4-sulfo-butyl)-3H-indol-2-yl]-vinyl}-cyclohex-2-enylidene)-ethylidene]-3,3-dimethyl-1-(4-sulfo-butyl)-2,3-dihydro-1H-indole-5-carboxylic acid (dye 2) was developed via an asymmetric approach. With an additional sulfonate group, the near-infrared feature of this dye exhibited a 2-fold increase in quantum yield compared to the previous generation. The current synthetic strategy provided a single carboxylic group as a handle for conjugation, thus allowing selectivity for bioconjugation. The stability of this dye was demonstrated by labeling peptides via solid-phase peptide chemistry. The in vivo optical imaging showed potential and broad applications of this dye in developing molecular-based beacons for cancer detection.  相似文献   

9.
The chemical transfer of alkyl groups from alkylcobalamins to mercuric ion has been studied in detail by using ultraviolet-visible conventional and stopped-flow kinetics and, in the case of methyl group transfer, by 220 MHz NMR spectroscopy. These experiments show that heterolytic cleavage of the Co–C δ-bond occurs during electrophilic attack by mercuric ion to give alkylmercury and aquocobalamin as the reaction products. Equilibrium and kinetic experiments are consistent with the initial displacement of 5,6-dimethylbenzimidazole by mercuric ion which results in a deactivaion toward dealkylation by a second mercuric ion. Consequently the main dealkylation reaction at pH 5.0 occurs with uncomplexed alkylcobalamin with the overall rate kobd being controlled by the above equilibrium. Both the displacement of 5,6-dimethylbenzimidazole (“fast reaction”) and dealkylation (“slow reaction”) are first order in the active mercuric species.  相似文献   

10.
Methyl 2,3-O-isopropylidene-alpha-D-mannofuranosidurononitrile [alternative name: methyl (5R)-5-C-cyano-2,3-O-isopropylidene-alpha-D-lyxofuranoside] (2), methyl 2,3-O-isopropylidene-alpha-D-mannofuranosiduronamide [methyl (5S)-5-C-carbamoyl-2,3-O-isopropylidene-alpha-D-lyxofuranoside; methyl (5S)-2,3-O-isopropylidene-alpha-D-lyxo-hexofuranosiduronamide] (3), methyl 2,3-O-isopropylidene-alpha-D-mannofuranosiduronic acid [methyl (5S)-2,3-O-isopropylidene-alpha-D-lyxo-hexofuranosiduronic acid] (4), methyl 5-deoxy-2,3-O-isopropylidene-5-ureido-beta-L-gulofuranosiduronamide [methyl (5R)-5-deoxy-2,3-O-isopropylidene-5-ureido-alpha-D-lyxo-hexofuranosiduronamide (5), and (4S,5S,6R)-5,6-dihydro-6-hydroxy-4,5-isopropylidenedioxy-4H-pyrido[2,1-e]imidazolidine-2',4'-dione [IUPAC name: (3aS,4R,8aS)-4-hydroxy-2,2-dimethyl-3a,8a-dihydro-4H-1,3-dioxa-4a,6-diaza-s-indacene-5,7-dione] (6), instead of the expected hydantoin derivative, were obtained from the Bucherer-Bergs reaction of methyl 2,3-O-isopropylidene-alpha-D-lyxo-pentodialdo-1,4-furanoside (1). The structure of 6 was deduced from NMR and mass spectral data and confirmed by X-ray crystallography. The configuration at C-5 in 2-5 was confirmed by establishing the 5S configuration of 3 by X-ray crystallography. Conformations of the six- and five-membered rings in 3 and 6 are also discussed.  相似文献   

11.
Further 2,3-unsaturated acids are revealed which can be reduced by Clostridium kluyveri with crotonate or butyrate as hydrogen donors. Unsaturated and saturated 3-halogenated acids are transformed into the saturated halogen-free acids. The following reaction sequence is proposed: a) hydrogenation, b) elumination of HX and c) again hydrogenation. Tiglinate ((E)-2-methyl-2-butenoate) and (E)-2-methyl-2-pentenoate are stereospecifically reduced to the (S)-2-methyl substituted acids. C. kluyveri contains endogenous material; in the presence of hydrogen acceptors such as 2,3-unsaturated acids this is degraded to acetate, and the reducing equivalents liberated hydrogenate the unsaturated acid. In a transient phase the hydration products of the unsaturated acids are present in the non-activated form in appreciable amounts. Tiglinate as well as crotonate is partially converted to ethyl methyl ketone and aceton and/or propanol, respectively.  相似文献   

12.
1. The steady-state parameters kcat and Km and the rate constants of hydride transfer for the substrates isopropanol/acetone; (S)-2-butanol, (R)-2-butanol/2-butanone; (S)-2-pentanol, (R)-2-pentanol/2-pentanone; 3-pentanol/3-pentanone; (S)-2-octanol and (R)-2-octanol have been determined for the native Zn(II)-containing horse-liver alcohol dehydrogenase (LADH) and the specific active-site-substituted Co(II)LADH. 2. A combined evaluation of steady-state kinetic data and rate constants obtained from stopped-flow measurements, allowed the determination of all rate constants of the following ordered bi-bi mechanism: E in equilibrium E.NAD in equilibrium E.NAD.R1R2 CHOH in equilibrium E.NADH.R1R2CO in equilibrium E.NADH in equilibrium E. 3. On the basis of the different substrate specificities of LADH and yeast alcohol dehydrogenase (YADH), a procedure has been developed to evaluate the enantiomeric product composition of ketone reductions. 2-Butanone and 2-pentanone reductions revealed (S)-2-butanol (86%) and (S)-2-pentanol (95%) as the major products. 4. The observed enantioselectivity implies the existence of two productive ternary complexes; E.NADH.(pro-S) 2-butanone and E.NADH.(pro-R) 2-butanone. All rate constants describing the kinetic pathways of the system (S)-2-butanol, (R)-2-butanol/2-butanone have been determined. These data have been used to estimate the expected enantiomer product composition of 2-butanone reductions using apparent kcat/Km values for the two different ternary-complex configurations of 2-butanone. Additionally, these data have been used for computer simulations of the corresponding reaction cycles. Calculated, simulated and experimental data were found to be in good agreement. Thus, the system (S)-2-butanol, (R)-2-butanol/2-butanone is the first example of a LADH-catalyzed reaction for which the stereochemical course could be described in terms of rate constants of the underlying mechanism. 5. The effects of Co(II) substitution on the different steps of the kinetic pathway have been investigated. The free energy of activation is higher for alcohol oxidation and lower for ketone reduction when catalyzed by Co(II)LADH in comparison to Zn(II)LADH. However, the free energies of binding are affected by metal substitution in such a way that the enantioselectivity of ketone reduction is not significantly changed by the substitution of Co(II) for Zn(II). 6. Evaluation of the data shows that substrate specificity and stereoselectivity result from combination of the free energies of binding and activation, with differences in binding energies as the dominating factors. In this regard, the interactions of substrate molecules with the protein moiety are dominant over the interactions with the catalytic metal ion.  相似文献   

13.
The radical scavenging mechanisms for the 2-pyrone compound, 4-hydroxy-3,6-dimethyl-2H-pyrane-2-one (1), and the 1,1-diphenyl-2-picrylhydrazyl (DPPH) radical (4) in several solvent systems were evaluated by the quantitative change in compounds detected at 270 nm and subsequent HPLC analyses. The HPLC profile for each condition suggested that the reaction proceeded by a different mechanism in each solvent system. In organic solvents (CHCl3, iso-propanol, and EtOH), 1-[4-(3,4-dihydro-3,6-dimethyl-2,4-dioxo-2H-pyran-3-yl) phenyl]-1-phenyl-2-picrylhydrazine (2) was produced as an adduct of the DPPH radical and 1. On the other hand, the reaction in a buffer solution (an acetate buffer at pH 5.5) gave several degradation products with 1[4-(2,3-dihydro-2,5-dimethyl-3-oxo-fur-2-yl) phenyl]-1-phenyl-2-picrylhydrazine (5), this being structurally elucidated by spectroscopic analyses. The decrease of the DPPH radical in each reaction system suggests that compound 1 could scavenge about 1.5-1.8 equivalents of the radical in organic solvents and about 3.5-3.9 in the buffer solution.  相似文献   

14.
The NAD-dependent oxidation of ethanol, 2,3-butanediol, and other primary and secondary alcohols, catalyzed by alcohol dehydrogenases derived from Penicillium charlesii, was investigated. Alcohol dehydrogenase, ADH-I, was purified to homogeneity in a yield of 54%. The enzyme utilizes several primary alcohols as substrates, with Km values of the order of 10?4m. A Km value of 60 mm was obtained for R,R,-2,3-butanediol. The stereospecificity of the oxidation of 2-butanol was investigated, and S-(+)-2-butanol was found to be oxidized 2.4 times faster than was R-(?)-2-butanol. The reduction of 2-butanone was shown to produce S-(+)-2-butanol and R-(?)-butanol in a ratio of 7:3. ADH-I is the primary isozyme of alcohol dehydrogenase present in cultures utilizing glucose as the sole carbon source. The level of alcohol dehydrogenase activity increased 7.6-fold in mycelia from cultures grown with glucose and 2,3-butanediol (0.5%) as carbon sources compared with the activity in cultures grown on only glucose. Two additional forms of alcohol dehydrogenase, ADH-II and ADH-III, were present in the cultures supplemented with 2,3-butanediol. These forms of alcohol dehydrogenase catalyze the oxidation of ethanol and 2,3-butanediol. These data suggest that P. charlesii carries out an oxidation of 2,3-butanediol which may constitute the first reaction in the degradation of 2,3-butanediol as well as the last reaction in the mixed-acid fermentation. Alcohol dehydrogenase activities in P. charlesii may be encoded by multiple genes, one which is expressed constitutively and others whose expression is inducible by 2,3-butanediol.  相似文献   

15.
The Rieske nonheme mononuclear iron oxygenase MdpJ of the fuel oxygenate-degrading bacterial strain Aquincola tertiaricarbonis L108 has been described to attack short-chain tertiary alcohols via hydroxylation and desaturation reactions. Here, we demonstrate that also short-chain secondary alcohols can be transformed by MdpJ. Wild-type cells of strain L108 converted 2-propanol and 2-butanol to 1,2-propanediol and 3-buten-2-ol, respectively, whereas an mdpJ knockout mutant did not show such activity. In addition, wild-type cells converted 3-methyl-2-butanol and 3-pentanol to the corresponding desaturation products 3-methyl-3-buten-2-ol and 1-penten-3-ol, respectively. The enzymatic hydroxylation of 2-propanol resulted in an enantiomeric excess of about 70% for the (R)-enantiomer, indicating that this reaction was favored. Likewise, desaturation of (R)-2-butanol to 3-buten-2-ol was about 2.3-fold faster than conversion of the (S)-enantiomer. The biotechnological potential of MdpJ for the synthesis of enantiopure short-chain alcohols and diols as building block chemicals is discussed.  相似文献   

16.
Reinvestigation of the fermentation broth and mycelium of the coprophilous fungus Guanomyces polythrix, grown in static conditions, led to the isolation of several phytotoxic compounds, including two new naphthopyranone derivatives, namely (2S, 3R)-5-hydroxy-6,8-dimethoxy-2,3-dimethyl-2,3-dihydro-4H-naphtho[2,3-b]-pyran-4-one and (2S, 3R)-5-hydroxy-6,8,10-trimethoxy-2,3-dimethyl-2,3-dihydro-4H-naphtho[2,3-b]-pyran-4-one. The structures of the new compounds were established by spectral and chiroptical methods. In addition, the structure of 8-hydroxy-6-methyl-9-oxo-9H-xanthene-1-carboxylic acid methyl ester was unambiguously determined by X-ray analysis. The isolates caused significant inhibition of radicle growth of two weed seedlings (Amaranthus hypochondriacus and Echinochloa crusgalli) and interacted with both spinach and bovine brain calmodulins.  相似文献   

17.
F Haeffner  T Norin    K Hult 《Biophysical journal》1998,74(3):1251-1262
Two strategies based on the use of subsets for calculating the enantioselectivity in lipase-catalyzed transesterifications using the CHARMM force field were investigated. Molecular dynamics was used in our search for low energy conformations. Molecular mechanics was used for refining these low energy conformations. A tetrahedral intermediate with a rigid central part was used for mimicking the transition state. The energy differences between the transition states of the diastereomeric enzyme-substrate complexes were calculated. The way of defining the subsets was based on two fundamentally different strategies. The first strategy used predefined parts of the enzyme and the substrate as subsets. The second approach formed energy-based subsets, varying in size with the substrates studied. The selection of residues to be included in these energy-based subsets was based on the energy of the interaction between the specific residue or water molecule and the transition state. The reaction studied was the kinetic resolution of secondary alcohols in transesterifications using the Candida antarctica lipase B as chiral biocatalyst. The secondary alcohols used in the study were 2-butanol, 3-methyl-2-butanol, and 3,3-dimethyl-2-butanol.  相似文献   

18.
1,3-Butadiene and isoprene (2-methyl-1,3-butadiene) are chemically related substances that are carcinogenic to rodents. The overall aim of this work is to elucidate the role of the genotoxic action of diepoxide metabolites in the carcinogenesis of the dialkenes. In vivo doses of the diepoxide metabolites were measured through reaction products with hemoglobin (Hb adducts) in studies of induced micronuclei (MN) in rodents. In the reaction with N-terminal valine in Hb, diepoxybutane and isoprenediepoxide form ring-closed adducts, pyrrolidines [N,N-(2,3-dihydroxy-1,4-butadiyl)valine and N,N-(2,3-dihydroxy-2-methyl-1,4-butadiyl)valine, respectively]. The method applied for Hb-adduct measurement is based on tryptic degradation of the protein and liquid chromatography electrospray ionisation tandem mass spectrometry (LC-ESI-MS/MS) analysis. Mice were given single i.p. injections of the monoepoxides of butadiene and isoprene, 1,2-epoxy-3-butene or 1,2-epoxy-2-methyl-3-butene, respectively. Rats were treated in the same way with 1,2-epoxy-3-butene. In mice pyrrolidine adduct levels increased with increasing administered doses of the monoepoxides. The in vivo dose of diepoxybutane was on average twice as high (0.29+/-0.059 mMh) as the in vivo dose of isoprenediepoxide (0.15+/-0.053 mMh) per administered dose (mmol/kg body weight) of the monoepoxides. In mice the genotoxic effects of the two monoepoxides, measured as the increase in the frequencies of micronuclei (MN), were approximately linearly correlated to the in vivo doses of the diepoxides (except at the highest dose of diepoxybutane). In rats the pyrrolidine-adduct levels from diepoxybutane were below the limit of quantification at all administered doses of 1,2-epoxy-3-butene and no significant increase was observed in the frequency of MN. Measurement of the ring-closed adducts to N-termini in Hb by the applied method permits analysis of in vivo doses of diepoxybutane and isoprenediepoxide, which may be further used for the elucidation of the mechanisms of carcinogenesis of butadiene and isoprene.  相似文献   

19.
1,3-Butadiene (BD) is a major industrial chemical and a rodent carcinogen, with mice being much more susceptible than rats. Oxidative metabolism of BD, leading to the DNA-reactive epoxides 1,2-epoxy-3-butene (BMO), 1,2-epoxy-3,4-butanediol (EBD) and 1,2:3,4-diepoxybutane (DEB), is greater in mice than rats. In the present study the DNA adduct profiles in liver and lungs of rats and mice were determined following exposure to BMO and to BD since these profiles may provide qualitative and quantitative information on the DNA-reactive metabolites in target tissues. Adducts detected in vivo were identified by comparison with the products formed from the reaction of the individual epoxides with 2'-deoxyguanosine (dG). In rats and mice exposed to [4-14C]-BMO (1-50 mg/kg, i.p.), DNA adduct profiles were similar in liver and lung with N7-(2-hydroxy-3-butenyl)guanine (G1) and N7-(1-(hydroxymethyl)-2-propenyl)guanine (G2) as major adducts and N7-2,3,4-trihydroxybutylguanine (G4) as minor adduct. In rats and mice exposed to 200 ppm [2,3-14C]-BD by nose-only inhalation for 6 h, G4 was the major adduct in liver, lung and testes while G1 and G2 were only minor adducts. Another N7-trihydroxybutylguanine adduct (G3), which could not unambiguously be identified but is either another isomer of N7-2,3,4-trihydroxybutylguanine or, more likely, N7-(1-hydroxymethyl-2,3-dihydroxypropyl)guanine, was present at low concentrations in liver and lung DNA of mice, but absent in rats. The evidence indicates that the major DNA adduct formed in liver, lung and testes following in vivo exposure to BD is G4, which is formed from EBD, and not from DEB.  相似文献   

20.
The Salmonella/microsome assay with strains TA97, TA98, TA100, TA1535, TA1537 and TA1538 was used to examine the potential mutagenicity of 5 dipyridyls, 1 tripyridyl, 3 dipyridinium diiodides and 2 pyridinium monoiodides. The widely used herbicide paraquat (1,1-dimethyl-4,4'-dipyridinium diiodide) and its precursor 4,4'-dipyridyl gave weak and marginal mutagenic activity to Salmonella typhimurium TA1535 and TA1538 in the presence of S9-mix. Significantly high mutagenicity was obtained with 2,2'-, 3,3'-, 2,3'-, and 2,4'-dipyridyls, 2,2',2"-tripyridyl, and 5 pyridinium salts under the same conditions. The positive mutagenic response of 2,2',2"-tripyridyl suggests that higher polymers of pyridine contaminating paraquat preparations might be mutagenic. The dose-response curves of 1,1-dimethyl-3,3'-dipyridinium diiodide and 1,1'-dimethyl-2,2'-dipyridinium diiodide revealed an exponential relationship between the number of induced revertants and the compound concentrations. The results suggested that the mechanism of mutation induced by these two compounds might be attributed to the chain reactions of their free-radicals with molecular oxygen.  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