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 62 毫秒
1.
The chloroimide 3,3-dichloro-4-(dichloromethylene)-2,5-pyrrolidinedione, a tetrachloroitaconimide, is the principal mutagen produced by chlorination of simulated poultry chiller water. It is the second most potent mutagenic disinfection by-product of chlorination ever reported. Six of seven new synthetic analogs of this compound are direct-acting mutagens in Ames tester strain TA-100. Computed energies of the lowest unoccupied molecular orbital (E(LUMO)) and of the radical anion stability (DeltaH(f)(rad)-DeltaH(f)) from MNDO-PM3 for the chloroimides show a quantitative correlation with the Ames TA-100 bacterial mutagenicity values. The molar mutagenicities of these direct acting mutagenic imides having an exocyclic double bond fit the same linear correlation (lnM(m) vs. E(LUMO); lnM(m) vs. DeltaH(f)(rad)--DeltaH(f)) as the chlorinated 2(5H)-furanones, including the potent mutagen MX, 3-chloro-4-(dichloro-methyl)-5-hydroxy-2(5H)-furanone, a by-product of water chlorination and paper bleaching with chlorine. Mutagenicity data for related haloimides having endocyclic double bonds are also given. For the same number of chlorine atoms, the imides with endocyclic double bonds have significantly higher Ames mutagenicity compared to their structural analogs with exocyclic double bonds, but do not follow the same E(LUMO) or DeltaH(f)(rad)-DeltaH(f) correlation as the exocyclic chloroimides and the chlorinated 2(5H)-furanones.  相似文献   

2.
The kinetics of the decomposition reactions of the CO(py)3(CO3)(H2O)+ ion have been investigated in aqueous perchloric acid solutions over a range of hydrogen ion concentrations (0.10 to 5.0 M) and at two ionic strengths (I = 1.0 and 5.0 M). At the lower ionic strength, plots of ln (AtA versus time show a nonlinearity that is consistent with that expected for consecutive first-order reactions. The rates of the faster reaction are similar to those reported for the spontaneous reduction of aquopyridine-cobalt(III) cations. At the higher ionic strength, the above noted curvature is not apparent and the decarboxylation kinetics of the title complex may be described by a pseudo-first-order rate constant: kobs = k[H3O+]. At 20°C, k = (1.75−+0.09) s−1 M−1 with activation parameters ofΔH = (97 −+ 4) kJ mol−1 and ΔS = −(54 −+ 32) J deg−1 mol−1. These kinetic parameters are compared with those previously reported for the similar complexes, Co(py)4CO3+ and Co(py)2(CO3)(H2O)2+.  相似文献   

3.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

4.
Association constants for N---H+O hydrogen bond formation between substituted ammonium dications and phenolate ion were measured in water and deuterium oxide at 25°C and 2.0 ionic strength. In combination with isotopic fractionation factors for phenol and the conjugate diacid of 1,2-ethanediamine determined by 13C NMR spectroscopy, these yield isotopic fractionation factors for amine dication-phenolate ion hydrogen bonds in water: φAB = 0.69 for 1,2-propanediamine dication with a pK difference between donor and acceptor, ΔpKa = −0.45, φAB = 0.88 for 1,2-ethanediamine dication (ΔpKa = −2.1), and φAB = 1.1 for piperizine dication (ΔpKa = −3.5). The hydrogen bond association constants follow Brønsted correlations α = 0.19 in water and α = 0.27 in deuterium oxide. The results are consistent with a double-minimum potential with a significant barrier for motion across the hydrogen bond.  相似文献   

5.
We investigate the effect of the prostaglandin D2 metabolite Δ12−PGD2 (9−Deoxy−Δ9, Δ12−13,14-dihydroprostaglandin D2) on collagen synthesis in human osteoblast. Δ12-PGJ2 at 10−5M enhanced collagen synthesis in the presence of 2 mM α-glycerophosphate-2Na. The stimulative effect appeared as early as 3 days after addition and continued until 22 days. The enhancement of type I collagen synthesis was confirmed by polyacrylamide gel electrophoresis. The potency was the same as 101t-8M 1 α, 25 dihydroxy vitamine D3 (1,25(OH)2D3). Northern blot analysis showed that 10−5M Δ 12-PGD2 and 10−8M 1,25(OH)2D3 enhanced the transcribtion of type 1 procollagen (α1) mRNA levels in osteoblasts.  相似文献   

6.
7.
Hepatitis B surface antibody (HBsAb) was immobilized to the surface of a gold electrode modified with cysteamine and colloidal gold as matrices to detect hepatitis B surface antigen (HBsAg). Differential pulse voltammetry (DPV) method was used for the investigation of the specific interaction between the immobilized HBsAb and HBsAg in solution, which was followed as a change of peak current in DPV with time. With the modified gold electrode, the differences in affinity of HBsAb with HBsAg at the temperatures of 37 and 40 °C were easily distinguished and the kinetic rate constants (kass and kdiss) and kinetic affinity constant K were determined from the curves of current versus time. In addition, the thermodynamic constants, ΔG, ΔH and ΔS, of the interaction at 37 °C were calculated, which were −56.65, −64.54 and −25.45 kJ mol−1, respectively.  相似文献   

8.
  1. In US Pacific Northwest ponderosa pine forests the primary attraction order shown previously for red turpentine beetle, Dendroctonus valens (Coleoptera: Curculionidae: Scolytinae), is (−)-β-pinene+ethanol > (+)-3-carene+ethanol > (+)-α-pinene+ethanol. The monoterpenes are bicyclic C10H16 isomers containing one 6-carbon ring with one double bond. Both pinenes have a 4-carbon second ring and differ only by their endocyclic or exocyclic double bond. The (+)-3-carene second ring has 3-carbons; its double bond is endocyclic like (+)-α-pinene.
  2. Ring system and double bond influences on primary attraction were evaluated by hydrogenating (+)-3-carene and (+)-α-pinene to cis-carane and cis-pinane, respectively. Field test primary attraction strengths were (−)-β-pinene+ethanol > cis-carane+ethanol > cis-pinane+ethanol > ethanol.
  3. In combination with ethanol (i) a double bond is not required in either ring system to attract D. valens, (ii) the cis-carane bicyclic 3, 6-carbon ring system provides stronger beetle attraction than the cis-pinane 4, 6-carbon bicyclic ring system, and likely structural basis for stronger (+)-3-carene attraction over (+)-α-pinene, (iii) adding an exocyclic double bond to the 4, 6-carbon ring system elevates attraction above the 3, 6-carbon ring system with no double bond, and (iv) the 4, 6-carbon ring system is a much stronger attractant with an exocyclic rather than endocyclic double bond.
  相似文献   

9.
The kinetics of formation of the complex ion, μ-carbonato-di-μ-hydroxo-bis((1,5-diamino-3-aza-pentane) cobalt(III), from the tri-μ-hydroxo-bis((1,5-diamino-3-aza-pentane(III)cobalt(III)) ion in aqueous buffered carbonate solution have been studied spectrophotometrically at 295 nm over the ranges 20.0θ°C34.8, 8.03pH9.44, 5 mM [CO32−35 mM and at an ionic strength of 0.1 M (LiClO4). On the basis of the kinetic results a mechanism, involving rapid cleavage of an hydroxo bridge followed by carbon dioxide uptake with subsequent bridge formation, has been proposed. At 25 °C, the rate of the carbon dioxide uptake is 0.58 M−1 s−1 with ΔH≠ = (13.2±0.7) kcal mol−1 and ΔS≠ = (−15.1 ± 0.7) cal deg−1 mol−1. The results are composed with those obtained for several mononuclear cobalt(III) and one dinuclear cobalt(III) complexes.  相似文献   

10.
The synthesis and pharmacology of 15 1-deoxy-Δ8-THC analogues, several of which have high affinity for the CB2 receptor, are described. The deoxy cannabinoids include 1-deoxy-11-hydroxy-Δ8-THC (5), 1-deoxy-Δ8-THC (6), 1-deoxy-3-butyl-Δ8-THC (7), 1-deoxy-3-hexyl-Δ8-THC (8) and a series of 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=0–4, 6, 7, where n=the number of carbon atoms in the side chain−2). Three derivatives (1719) of deoxynabilone (16) were also prepared. The affinities of each compound for the CB1 and CB2 receptors were determined employing previously described procedures. Five of the 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=1–5) have high affinity (Ki=<20 nM) for the CB2 receptor. Four of them (2, n=1–4) also have little affinity for the CB1 receptor (Ki=>295 nM). 3-(1′,1′-Dimethylbutyl)-1-deoxy-Δ8-THC (2, n=2) has very high affinity for the CB2 receptor (Ki=3.4±1.0 nM) and little affinity for the CB1 receptor (Ki=677±132 nM).
Scheme 3. (a) (C6H5)3PCH3+ Br, n-BuLi/THF, 65°C; (b) LiAlH4/THF, 25°C; (c) KBH(sec-Bu)3/THF, −78 to 25°C then H2O2/NaOH.  相似文献   

11.
We investigated the adaptative response of S. cerevisiae in sod mutants (sod1Δ, sod2Δ and sod1Δsod2Δ) after H2O2 treatment in the stationary phase. sod2Δ and sod1Δsod2Δ demonstrated the highest levels of GSH in the control, suggesting that pathways which include GSH protect these double mutants against oxidative stress. In addition, sod1Δ and sod1Δsod2Δ had higher iron levels than the wild-type, independently of H2O2 stress. Fe levels were increased in sod2Δ following H2O2 In addition, the sod2Δ mutant was more sensitive to H2O2 treatment than the wild-type. These results suggest that sod2Δ sensibility may be associated with •OH production by the Fenton reaction. This increased iron demand in the sod2Δ mutant may be a reflection of the cells’ efforts to reconstitute proteins that are inactivated in conditions of excess superoxide. MDA levels were assayed by HPLC in these mutants. The highest MDA levels could be observed after 10mM H2O2 treatment in the sod1Δsod2Δ double mutant. After treatment with a GSH inhibitor, the MDA level was still higher in the same strain. Thus, both direct and indirect GSH pathways are involved in the protection of lipid membranes and proteins in these mutants and may constitute an adaptative response to enhanced basal oxidative damage produced by superoxide.  相似文献   

12.
A sensitive, selective, and rapid enzymatic method is proposed for the quantification of hydrogen peroxide (H2O2) using 3-methyl-2-benzothiazolinonehydrazone hydrochloride (MBTH) and 10,11-dihydro-5H-benz(b,f)azepine (DBZ) as chromogenic cosubstrates catalyzed by horseradish peroxidase (HRP) enzyme. MBTH traps free radical released during oxidation of H2O2 by HRP and gets oxidized to electrophilic cation, which couples with DBZ to give an intense blue-colored product with maximum absorbance at 620 nm. The linear response for H2O2 is found between 5 × 10−6 and 45 × 10−6 mol L−1 at pH 4.0 and a temperature of 25 °C. Catalytic efficiency and catalytic power of the commercial peroxidase were found to be 0.415 × 106 M−1 min−1 and 9.81 × 10−4 min−1, respectively. The catalytic constant (kcat) and specificity constant (kcat/Km) at saturated concentration of the cosubstrates were 163.2 min−1 and 4.156 × 106 L mol−1 min−1, respectively. This method can be incorporated into biochemical analysis where H2O2 undergoes catalytic oxidation by oxidase. Its applicability in the biological samples was tested for glucose quantification in human serum.  相似文献   

13.
To gain the structure–activity relationship of Δ1-androstenediones (Δ1-ADs) as mechanism-based inactivator of aromatase, series of 2-alkyl- and 2-alkoxy-substitiuted Δ1-ADs (6 and 9) as well as 2-bromo-Δ1-AD (14) were synthesized and tested. All of the inhibitors examined blocked aromatase in human placental microsomes in a competitive manner. In a series of 2-alkyl-Δ1-ADs (6), n-hexyl compound 6f was the most powerful inhibitor with an apparent Ki value of 31 nM. The inhibitory activities of 2-alkoxy steroids 9 decreased in relation to length of the alkyl chain up to n-hexyloxy group (Ki: 95 nM for methoxy 9a). All of the alkyl steroids 6 along with the alkoxy steroid 9, except for the ethyl and n-propyl compounds 6b and 6c, caused a time-dependent inactivation of aromatase. The inactivation rates (kinact: 0.020–0.084 min−1) were comparable to that of the parent compound Δ1-AD. The inactivation was prevented by the substrate AD, and no significant effect of l-cysteine on the inactivation was observed in each case. The results indicate that the 2-hexyl compound 6f act as the most powerful mechanism-based inactivator of aromatase among Δ1-AD analogs and may be submitted to the preclinical study in estrogen-dependent breast cancer.  相似文献   

14.
The aim of this study was to determine whether hyperreninemia in the adrenalectomized (ADX) rat is dependent on renal prostaglandin synthesis, as has been suggested for two other hyperreninemic conditions, Bartter's syndrome and chronic liver disease.Plasma renin concentration (PRC) in anesthetized, ADX rats was significantly increased (Δ +480%; p < 0.001) compared to sham-operated controls. , indomethacin (10 mg/kg i.v.) significantly reduced PRC of anesthetized, ADX rats after both 45 min (Δ −34%; p < 0.05) and 90 min (Δ −47%; p < 0.05). renin release from renal cortical slices of ADX rats was also significantly greater (Δ +130%; p < 0.05) than from sham-operated control cortical slices. Renin release from cortical slices of ADX rats given dexamethasone (10 μg/kg/day) for 4 days prior to sacrifice did not differ from sham-operated control values.Prostaglandin E2 (PGE2) release from cortical slices of ADX rats did not differ significantly from controls. However, PGE2 synthesis in glomeruli microdissected from ADX rats was significantly increased (Δ +110%; p < 0.001) compared to controls. PGE2 synthesis in glomeruli of dexamethasone-treated ADX rats remained significantly elevated compared to controls. Ibuprofen (10−6 M) decreased PGE2 synthesis in cortical slices by 80%. However, prostaglandin synthesis inhibition had no effect on renin release from either ADX or control renal cortical slices.These results suggest that despite increased glomerular synthesis, prostaglandins do not directly influence renin release in the ADX rat.  相似文献   

15.
The seasonal variability of specific growth rate and the carbon stable isotope ratio (δ13C) of leaf blades (δ13Cleaf) of a temperate seagrass, Zostera marina (within 10 days old) were measured simultaneously, together with the δ13C of dissolved inorganic carbon (δ13CDIC) at three sites in the semi-closed Akkeshi estuary system, northeastern Japan, in June, September, and November 2004. The δ13Cleaf ranged from −16.2 to −6.3‰ and decreased from summer to winter. The simultaneous measurement of the δ13Cleaf, growth rate, and morphological parameters (mean leaf length and width, mean number of leaves per shoot, and sheath length) of the seagrass and δ13CDIC in the surrounding water allowed us to compare directly the δ13Cleaf and specific growth rate of seagrass. The difference in the δ13C of seagrass leaves relative to the source DIC (Δδ13Cleaf − DIC) was the least negative (−11 to −7‰) in June at all three sites and became more negative (−17 to −8‰) as the specific growth rate decreased. This positive correlation between Δδ13Cleaf − DIC and specific growth rate can be used to diagnose the growth of seagrasses. Δδ13Cleaf − DIC changed by −1.7 ± 0.2‰ when the leaf specific growth rate decreased by 1% d−1.  相似文献   

16.
17.
Synthetic peptide S3Δ has high affinity for bacterial endotoxin or lipopolysaccharide (LPS). Under tested conditions of pH 5–9 and 0–0.4 M NaCl, the affinity constant, KD ranged from 2·10−6 to 2·10−9 M−1. A novel affinity matrix based on peptide S3Δ was developed for removal of LPS from solutions such as: water; buffers with a wide range of ionic strength and pH; medium for cell culture; and protein solutions under optimized conditions. At a starting LPS of ≈100 EU/ml, a post-purification level below 0.005 EU/ml was achieved.  相似文献   

18.
OCP1 and OCP2, the most abundant proteins in the cochlea, are evidently subunits of an SCF E3 ubiquitin ligase. Although transcribed from a distinct gene, OCP2 is identical to Skp1. OCP1 is equivalent to the F-box protein known as Fbs1, Fbx2, or NFB42 — previously shown to bind N-glycosylated proteins and believed to function in the retrieval and recycling of misfolded proteins. The high concentrations of OCP1 and OCP2 in the cochlea suggest that the OCP1–OCP2 heterodimer may serve an additional function independent of its role in a canonical SCF complex. At 25 °C, urea-induced denaturation of OCP1 is slow, but reversible. The data suggest that the protein possesses one or more disordered regions, a conclusion supported by analysis of the far-UV circular dichroism spectrum and the appearance of the 1H, 15N-HSQC spectrum. Thermal denaturation of OCP1 is irreversible, evidently due to formation of high molecular weight aggregates. Analysis with a kinetic model yields an estimate for the activation energy for unfolding of 49 kcal/mol. Urea denaturation data for OCP2 returns ΔGo and m values of 6.2 kcal/mol and 1.5 kcal mol− 1 M− 1, respectively. In contrast to OCP1, thermal denaturation of OCP2 is reversible. In phosphate-buffered saline, at pH 7.40, the protein exhibits a ΔHvHHcal ratio of 1.69, suggesting that denaturation proceeds largely from the native dimer directly to the unfolded state. OCP1 and OCP2 associate tightly at room temperature. However, DSC data for the complex suggest that they denature independently, consistent with the highly exothermic enthalpy of complex formation reported previously.  相似文献   

19.
Yeast cytochrome c peroxidase (CCP) efficiently catalyzes the reduction of H2O2 to H2O by ferrocytochrome c in vitro. The physiological function of CCP, a heme peroxidase that is targeted to the mitochondrial intermembrane space of Saccharomyces cerevisiae, is not known. CCP1-null-mutant cells in the W303-1B genetic background (ccp1Δ) grew as well as wild-type cells with glucose, ethanol, glycerol or lactate as carbon sources but with a shorter initial doubling time. Monitoring growth over 10 days demonstrated that CCP1 does not enhance mitochondrial function in unstressed cells. No role for CCP1 was apparent in cells exposed to heat stress under aerobic or anaerobic conditions. However, the detoxification function of CCP protected respiring mitochondria when cells were challenged with H2O2. Transformation of ccp1Δ with ccp1W191F, which encodes the CCPW191F mutant enzyme lacking CCP activity, significantly increased the sensitivity to H2O2 of exponential-phase fermenting cells. In contrast, stationary-phase (7-day) ccp1Δ-ccp1W191F exhibited wild-type tolerance to H2O2, which exceeded that of ccp1Δ. Challenge with H2O2 caused increased CCP, superoxide dismutase and catalase antioxidant enzyme activities (but not glutathione reductase activity) in exponentially growing cells and decreased antioxidant activities in stationary-phase cells. Although unstressed stationary-phase ccp1Δ exhibited the highest catalase and glutathione reductase activities, a greater loss of these antioxidant activities was observed on H2O2 exposure in ccp1Δ than in ccp1Δ-ccp1W191F and wild-type cells. The phenotypic differences reported here between the ccp1Δ and ccp1Δ-ccp1W191F strains lacking CCP activity provide strong evidence that CCP has separate antioxidant and signaling functions in yeast.  相似文献   

20.
We explored the effects of compounds known or proposed to affect microtubule functions on superoxide (O2) production in human polymorphonuclear leukocytes stimulated by N-formyl-methionyl-phenylalanine (f-Met-Phe), calcium ionophore A23187 and phorbol myristate acetate. F-Met-Phe-induced O2 production was markedly potentiated not only by microtubule-disrupting agents, including colchicine, vincristine, vinblastine, nocodazole, podophyllotoxin and griseofulvin, but also deuterium oxide (2H2O), which is proposed to stabilize microtubules, and not affected by lumicolchicine. Ionophore A23187-induced O2 production was not influenced by colchicine, and markedly enhanced by 2H2O, whereas phorbol myristate acetate-induced O2 production was not influenced by colchicine, and slightly inhibited by 2H2O. 2H2O did not counteract the effects of colchicine and vice versa. Dibutyryl cyclic AMP and prostaglandin E1 inhibited O2 production stimulated by f-Met-Phe and ionophore A23187, whereas phorbol myristate acetate-induced O2 production was strongly resistant to the inhibitory effect of these agents. The enhancing effect of colchicine and 2H2O on f-Met-Phe-induced O2 production was abolished by dibutyryl cyclic AMP. Colchicine promoted concanavalin A cap formation, and 2H2O produced cancanavalin A patch formation, whereas dibutyryl cyclic AMP did not affect the distribution of concanavalin A receptors. In addition, 2H2O and dibutyryl cyclic AMP did not interfere with the colchicine-induced concanavalin A cap formation. These findings suggest that f-Met-Phe, ionophore A23187 and phorbol myristate acetate may activate the oxidative metabolism of human polymorphonuclear leukocytes through different mechanisms, and that microtubule-disrupting agents, 2H2O and cyclic AMP agonists may affect the different steps of the activating system of NAD(P)H oxidase.  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