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.
The reaction of H2[PtCl6] · 6H2O and (H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O (18C6 = 18-crown-6) with 9-methylguanine (MeGua) proceeded with the protonation of MeGua forming 9-methylguaninium hexachloroplatinate(IV) dihydrate (MeGuaH)2[PtCl6] · 2H2O (1).The same compound was obtained from the reaction of Na2[PtCl6] with (MeGuaH)Cl.On the other hand, the reaction of guanosine (Guo) with (H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O in methanol at 60 °C proceeded with the cleavage of the glycosidic linkage and with ligand substitution to give a guaninium complex of platinum(IV), [PtCl5(GuaH)] · 1.5(18C6) · H2O (2).Within several weeks in aqueous solution a slow reduction took place yielding the analogous guaninium platinum(II) complex, [PtCl3(GuaH)] · (18C6) · 2Me2CO (3).H2[PtCl6] · 6H2O and guanosine was found to react in water, yielding (GuoH)2[PtCl6] (4) and in ethanol at 50 °C, yielding [PtCl5(GuoH)] · 3H2O (5).Dissolution of complexes 2 and 5 in DMSO resulted in the substitution of the guaninium and guanosinium ligands, respectively, by DMSO forming [PtCl5(DMSO)].Reactions of 1-methylcytosine (MeCyt) and cytidine (Cyd) with H2[PtCl6] · 6H2O and(H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O resulted in the formation of hexachloroplatinates with N3 protonated pyrimidine bases as cation (MeCytH)2[PtCl6] · 2H2O (6) and (CydH)2[PtCl6] (7), respectively. Identities of all complexes were confirmed by 1H, 13C and 195Pt NMR spectroscopic investigations, revealing coordination of GuoH+ in complex 5 through N7 whereas GuaH+ in complex 3 may be coordinated through N7 or through N9. Solid state structure of hexachloroplatinate 1 exhibited base pairing of the cations yielding (MeGuaH+)2, whereas in complex 6 non-base-paired MeCytH+ cations were found. In both complexes, a network of hydrogen bonds including the water molecules was found. X-ray diffraction analysis of complex 3 exhibited a guaninium ligand that is coordinated through N9 to platinum and protonated at N1, N3 and N7. In the crystal, these NH groups form hydrogen bonds N–HO to oxygen atoms of crown ether molecules.  相似文献   

2.
Epidemiologic studies indicate that prolonged exposure to particulate air pollution may be associated with increased risk of cardiovascular diseases and cancer in general population. These effects may be attributable to polycyclic aromatic hydrocarbons (PAHs) adsorbed to respirable air particles. It is expected that metabolic and DNA repair gene polymorphisms may modulate individual susceptibility to PAH exposure. This study investigates relationships between exposure to PAHs, polymorphisms of these genes and DNA adducts in group of occupationally exposed policemen (EXP, N = 53, males, aged 22–50 years) working outdoors in the downtown area of Prague and in matched “unexposed” controls (CON, N = 52). Personal exposure to eight carcinogenic PAHs (c-PAHs) was evaluated by personal samplers during working shift prior to collection of biological samples. Bulky-aromatic DNA adducts were analyzed in lymphocytes by 32P-postlabeling assay. Polymorphisms of metabolizing (GSTM1, GSTP1, GSTT1, EPHX1, CYP1A1-MspI) and DNA repair (XRCC1, XPD) genes were determined by PCR-based RFLP assays. As potential modifiers and/or cofounders, urinary cotinine levels were analyzed by radioimmunoassay, plasma levels of vitamins A, C, E and folates by HPLC, cholesterol and triglycerides using commercial kits. During the sampling period ambient particulate air pollution was as follows: PM10 32–55 μg/m3, PM2.5 27–38 μg/m3, c-PAHs 18–22 ng/m3; personal exposure to c-PAHs: 9.7 ng/m3 versus 5.8 ng/m3 (P < 0.01) for EXP and CON groups, respectively. The total DNA adduct levels did not significantly differ between EXP and CON groups (0.92 ± 0.28 adducts/108 nucleotides versus 0.82 ± 0.23 adducts/108 nucleotides, P = 0.065), whereas the level of the B[a]P-“like” adduct was significantly higher in exposed group (0.122 ± 0.036 adducts/108 nucleotides versus 0.099 ± 0.035 adducts/108 nucleotides, P = 0.003). A significant difference in both the total (P < 0.05) and the B[a]P-“like” DNA adducts (P < 0.01) between smokers and nonsmokers within both groups was observed. A significant positive association between DNA adduct and cotinine levels (r = 0.368, P < 0.001) and negative association between DNA adduct and vitamin C levels (r = −0.290, P = 0.004) was found. The results of multivariate regression analysis showed smoking, vitamin C, polymorphisms of XPD repair gene in exon 23 and GSTM1 gene as significant predictors for total DNA adduct levels. Exposure to ambient air pollution, smoking, and polymorphisms of XPD repair gene in exon 6 were significant predictors for B[a]P-“like” DNA adduct. To sum up, this study suggests that polymorphisms of DNA repair genes involved in nucleotide excision repair may modify aromatic DNA adduct levels and may be useful biomarkers to identify individuals susceptible to DNA damage resulting from c-PAHs exposure.  相似文献   

3.
DNA-based stable isotope probing (SIP) is a novel technique for the identification of organisms actively assimilating isotopically labeled compounds. Herein, we define the limitations to using 15N-labeled substrates for SIP and propose modifications to compensate for these shortcomings. Changes in DNA buoyant density (BD) resulting from 15N incorporation were determined using cultures of disparate GC content (Escherichia coli and Micrococcus luteus). Incorporation of 15N into DNA increased BD by 0.015±0.002 g mL−1 for E. coli and 0.013±0.002 g mL−1 for M. luteus. The DNA BD shift was greatly increased (0.045 g mL−1) when dual isotope (13C plus 15N) labeling was employed. Despite the limited DNA BD shift following 15N enrichment, we found the use of gradient fractionation, followed by a comparison of T-RFLP profiles from fractions of labeled and control treatments, facilitated detection of enrichment in DNA samples from either cultures or soil.  相似文献   

4.
The formation of three [Tl(en)n]3+ complexes (n=1–3) in a pyridine solvent has been established by means of 205Tl and 1H NMR. Their stepwise stability constants based on concentrations, Kn=[Tl(en)n 3+]/{[Tl(en)n−1 3+]·[en]}, at 298 K in 0.5 M NaClO4 ionic medium in pyridine, were calculated from 205Tl NMR integrals: log K1=7.6±0.7; log K2=5.2±0.5 and log K3=2.64±0.05. Linear correlation between both the 205Tl NMR shifts and spin–spin coupling 205Tl–1H versus the stability constants has been found and discussed. A single crystal with the composition [Tl(en)3](ClO4)3 was synthesized and its structure determined by X-ray diffraction. The Tl3+ ion is coordinated by three ethylenediamine ligands via six N-donor atoms in a distorted octahedral fashion.  相似文献   

5.
Using the pulse radiolysis technique, we have demonstrated that bleomycin-Fe(III) is stoichiometrically reduced by CO2- to bleomycin-Fe(II) with a rate of (1.9 ± 0.2) × 108M-1s-1. In the presence of calf thymus DNA, the reduction proceeds through free bleomycin-Fe(III) and the binding constant of bleomycin-Fe(III) to DNA has been determined to be (3.8 ± 0.5) x 104 M-1. It has also been demonstrated that in the absence of DNA O2-1 reacts with bleomycin-Fe(III) to yield bleomycin-Fe(II)O2, which is in rapid equilibrium with molecular oxygen, and decomposes at room temperature with a rate of (700 ± 200) s-1. The resulting product of the decomposition reaction is Fe(III) which is bound to a modified bleomycin molecule. We have demonstrated that during the reaction of bleomycin-Fe(II) with O2, modification or self-destruction of the drug occurs, while in the presence of DNA no destruction occurs, possibly because the reaction causes degradation of DNA.  相似文献   

6.
Five heterometallic compounds with formulae [Ba(H2O)4Cr2(μ-OH)2(nta)2] · 3H2O (I), [M(bpy)2(H2O)2] [Cr2(OH)2(nta)2] · 7H2O, where M2+ = Zn, (II); Ni, (III); Co, (IV) and [Mn(H2O)3(bpy)Cr2(OH)2(nta)2] · (bpy) · 5H2O (V); bpy = 2,2′-bipyridine, (nta = nitrilotriacetate ion) have been prepared by reaction of I with the corresponding MII-sulfates in the presence of 2,2′-bipyridine. Substances I–V have been characterized by magnetic susceptibility measurements, EPR and X-ray determinations. I represents a 2D coordination polymer formed by coordination of centrosymmetrical dimeric chromium(III) units and Barium cations. The 10-coordinate Ba polyhedron is completed by four water molecules. Compounds II–IV are isostructural and consist of non-centrosymmetric dimeric anions [Cr2(μ-OH)2(nta)2]2−, complex cations [MII(bpy)2(H2O)2]2+ and solvate water molecules. The octahedral coordination of chromium atoms implies four donor atoms of the nta3− ligands and two bridging OH groups. Multiple hydrogen bonds of coordinated and solvate water molecules link anions and cations in a 3D network. A similar [Cr2(μ-OH)2(nta)2]2− unit is found in V. The bridging function is performed by a carboxylate oxygen atom of the nta ligand that leads to the formation of a trinuclear complex [Mn(bpy)(H2O)2Cr2(μ-OH)2(nta)2]. Experimental and calculated frequency and temperature dependences of EPR spectra of these compounds are presented. The fine structure appearing on the EPR spectra of compound V is analyzed in detail at different temperatures. It is established that the main part of the EPR signals is due to the transitions in the spin states of a spin multiplet with S = 2. Analyses of experimental and calculated spectra confirm the absence of interaction between metal ions (MII) and Cr-dimers in complexes III and IV and the presence of weak Mn–Cr interactions in V. The temperature dependence of magnetic susceptibilities for I–V was fitted on the basis of the expression derived from isotropic Hamiltonian including a bi-quadratic exchange term.  相似文献   

7.
The complexes [(bpy)2Ru(dpp)]Cl2, [(phen)2Ru(dpp)]Cl2, and [(Ph2phen)2Ru(dpp)]Cl2 (where dpp = 2,3-bis(2-pyridyl)pyrazine, bpy = 2,2′-bipyridine, phen = 1,10-phenanthroline, Ph2phen = 4,7-diphenyl-1,10-phenanthroline) have been investigated and found to photocleave DNA via an oxygen-mediated pathway. These light absorbing complexes possess intense metal-to-ligand charge transfer (MLCT) transitions in the visible region of the spectrum. The [(TL)2Ru(dpp)]2+ systems populate 3MLCT states after visible light excitation, giving rise to emissions in aqueous solution centered at 692, 690, and 698 nm for TL = bpy, phen, and Ph2phen respectively. The 3MLCT states and emissions are quenched by O2, producing a reactive oxygen species. These complexes photocleave DNA with varying efficiencies, [(Ph2phen)2Ru(dpp)]2+ > [(phen)2Ru(dpp)]2+ > [(bpy)2Ru(dpp)]2+. The presence of the polyazine bridging ligand will allow these chromophores to be incorporated into larger supramolecular assemblies.  相似文献   

8.
To examine associations between two different classes of DNA damage that can occur through endogenous processes or exogenous exposures such as smoking, N7-methyldeoxyguanosine (N7-MedG) and 8-oxodeoxyguanosine (8-oxodG) levels were measured in lymphocyte DNA from 22 bronchoscopy patients. 8-OxodG and N7-MedG was detected in 100% and 91% of samples, respectively with 8-oxodG levels being approx 20 times higher (mean 8.39 ± 3.57 8-oxodG/106dG versus 0.41 ± 0.33 N7-MedG/106 dG) which provides an indication of the relative importance of the agents that induce oxidative DNA damage or alkylation damage. The sources of these genotoxic lesions remain to be established but N7-MedG and 8-oxodG levels were not correlated (r2 < 0.01) suggesting that there is no association between alkylating agent and reactive oxygen species exposure, their metabolism and/or the DNA repair processes that can remove this DNA damage.  相似文献   

9.
Acellular assay of calf thymus DNA ± rat liver microsomal S9 fraction coupled with 32P-postlabelling was used to study the genotoxic potential of organic compounds bound onto PM10 particles collected in three European cities—Prague (CZ), Kosice (SK) and Sofia (BG) during summer and winter periods. B[a]P alone induced DNA adduct levels ranging from 4.8 to 768 adducts/108 nucleotides in the concentration dependent manner. However, a mixture of 8 c-PAHs with equimolar doses of B[a]P induced 3.7–757 adducts/108 nucleotides, thus suggesting the inhibition of DNA adduct forming activity by interaction among various PAHs. Comparison of DNA adduct levels induced by various EOMs indicates higher variability among seasons than among localities. DNA adduct levels for Prague collection site varied from 19 to 166 adducts/108 nucleotides, for Kosice from 22 to 85 and for Sofia from 6 to 144 adducts/108 nucleotides. Bioactivation with S9 microsomal fraction caused 2- to 7-fold increase in DNA adduct levels compared to −S9 samples, suggesting a crucial role of indirectly acting genotoxic EOM components, such as PAHs. We have demonstrated for the first time a significant positive correlation between B[a]P content in EOMs and total DNA adduct levels detected in the EOM treated samples (R = 0.83; p = 0.04). These results suggest that B[a]P content in EOM is an important factor for the total genotoxic potential of EOM and/or B[a]P is a good indicator of the presence of other genotoxic compounds causing DNA adducts. Even stronger correlation between the content of genotoxic compounds in EOMs and total DNA adduct levels detected (R = 0.94; p = 0.005) was found when eight c-PAHs were taken into the consideration. Our findings support a hypothesis that a relatively limited number of EOM components is responsible for a major part of its genotoxicity detectable as DNA adducts by 32P-postlabelling.  相似文献   

10.
It is well accepted that estradiol (E2) plays an important role in the genesis and evolution of breast cancer. Quantitative evaluation indicates that in human breast tumor, estrone sulfate (E1S) ‘via sulfatase’ is a much more likely precursor for E2 than is androstenedione ‘via aromatase’. In previous studies, it was demonstrated that in isolated MCF-7 and T-47D breast cancer cell lines, estradiol can block estrone sulfatase activity. In the present study, the effect of E2 was explored using total normal and cancerous breast tissues. This study was carried out with post-menopausal patients with breast cancer. None of the patients had a history of endocrine, metabolic or hepatic diseases or had received treatment in the previous 2 months. Each patient received local anaesthetic (lidocaine 1%) and two regions of the mammary tissue were selected: (A) the tumoral tissue and (B) the distant zone (glandular tissue) which was considered as normal. Samples were placed in liquid nitrogen and stored at –80 °C until enzyme activity analysis. Breast cancer histotypes were ductal and post-menopausal stages were T2. Homogenates of tumoral or normal breast tissues (45–75 mg) were incubated in 20 mM Tris–HCl, pH 7.2 with physiological concentrations of [3H]-E1S (5 × 10−9 M) alone or in the presence of E2 (5 × 10−5 to 5 × 10−7 M) during 30 min or 3 h. E1S, E1 and E2 were characterized by thin layer chromatography and quantified using the corresponding standard. The sulfatase activity is significantly more intense with the breast cancer tissue than normal tissue, since the concentration of E1 was 3.20 ± 0.15 and 0.42 ± 0.07 pmol/mg protein, respectively after 30 min incubation. The values were 27.8 ± 1.8 and 3.5 ± 0.21 pmol/mg protein, respectively after 3 h incubation. Estradiol at the concentration of 5 × 10−7 M inhibits this conversion by 33% and 31% in cancerous and normal breast tissues, respectively and by 53% and 88% at the concentration of 5 × 10−5 M after 30 min incubation. The values were 24% and 18% for 5 × 10−7 M and 49% and 42% for 5 × 10−5 M, respectively after 3 h incubation. It was observed that [3H]-E1S is only converted to [3H]-E1 and not to [3H]-E2 in normal or cancerous breast tissues, which suggests a low or no 17β-hydroxysteroid dehydrogenase (17β-HSD) Type 1 reductive activity in these experimental conditions. In conclusion, estradiol is a strong anti-sulfatase agent in cancerous and normal breast tissues. This data can open attractive perspectives in clinical trials using this hormone.  相似文献   

11.
The detection of DNA adducts is an important component in assessing the mutagenic potential of exogenous and endogenous compounds. Here, we report an in vitro quantitative long PCR (XL-PCR) assay to measure DNA adducts in human genomic DNA based on their ability to block and inhibit PCR amplification. Human genomic DNA was exposed to test compounds and then a target sequence was amplified by XL-PCR. The amplified sequence was then quantified using fluorogenic 5′ nuclease PCR (TaqMan®) and normalized to a solvent-treated control. The extent of DNA adduction was determined based on the reduction in amplification of the target sequence in the treated sample. A 17.7 kb β-globin fragment was chosen as the target sequence for these studies, since preliminary experiments revealed a two-fold increased sensitivity of this target compared to a 10.4 kb HPRT fragment for detecting hydrogen peroxide-induced DNA damage. Validation of the XL-PCR assay with various compounds demonstrated the versatility of the assay for detecting a wide range of adducts formed by direct acting or S9-activated mutagens. The same DNA samples were also analyzed using 32P-postlabeling techniques (thin-layer chromatography or high-performance liquid chromatography) to confirm the presence of DNA adducts and estimate their levels. Whereas 32P-postlabeling with nuclease P1 enrichment was more sensitive for detecting bulky adducts induced by the compounds benzo[a]pyrene, dimethylbenzanthracene, 3-methylindole, indole 3-carbinol, or 2-acetylaminofluorene, the XL-PCR procedure was more sensitive for detecting smaller or labile DNA adducts formed by the compounds methyl methanesulfonate, diethyl nitrosamine, ethylnitrosourea, diepoxybutane, ICR-191, styrene oxide, or aflatoxin B1. Compounds not expected to form adducts in DNA, such as clofibrate, phenobarbital, chloroform or acetone, did not produce a positive response in the XL-PCR assay. Thus, quantitative XL-PCR provides a rapid, high-throughput assay for detecting DNA damage that complements the existing 32P-postlabeling assay with nuclease P1 enrichment.  相似文献   

12.
A comparison study of biochemical parameters of semen from Muscovy drakes diluted and stored at 4 °C in three buffers—IMV-buffer (France), HIA-1 and AU (Bulgaria) was carried out. The ejaculates were collected twice a week from ten 1-year-old Muscovy drakes using laying Muscovy females as teaser. Semen was diluted immediately, respectively, with IMV-buffer, HIA-1 and AU, and cold-stored (4 °C) for 1, 3 and 6 h. The intensities of oxygen uptake at the third hour in semen diluted, respectively, with IMV-buffer (200 ± 1.6 nAO/109 sperm cells min), with HIA (224 ± 44 nAO/109 sperm cells min) and with AU (238 ± 48 nAO/109 sperm cells min) were highly significant in comparison with neat semen (75 ± 0.7 nAO/109 sperm cells min).

The observed intensity of fructolysis was highest when using AU, followed by HIA-1 and IMV-buffer. During the first hour of storage the level of pyruvic acid was significantly lower in semen diluted with Bulgarian extenders, and this stability for AU referred to the entire period. For lactic acid, the differences were not statistical significant. Our investigations do not show significant differences concerning the dynamics of inorganic phosphate and total lipids after dilution with all tested extenders. On the contrary, high increase of cholesterol efflux from spermatozoa to seminal plasma–diluents were obtained after 6 h of storage.

All extenders, IMV-buffer (France), HIA-1 and AU (Bulgaria) for diluting and short time storage of semen from Muscovy drakes at 4 °C maintain the necessary comfort of energy metabolism of the spermatozoa.  相似文献   


13.
Crystal structures of Co2(CO)6(dppm) (1) and Co2(CO)5(CHCO2Et)(dppm) (2) (dppm = Ph2PCH2PPh2) show asymmetry with respect to the orientation of the phenyl groups in 1 and owing to the bridging ethoxycarbonylcarbene ligand in 2. The effect of this asymmetry was recognized in the solid-state 31P NMR spectra of 1 and 2 and in the solid-state and solution 13C NMR spectra of 2 as well, but not in the solid-state and solution 13C NMR spectra of 1. In CH2Cl2 solution under an atmosphere of 13CO, the CO ligands of both complexes exchange with 13CO. The overall rate of 13CO exchange at 10 °C was found to be kobs = 0.107 × 10−3 s−1 for 1 and kobs = 0.243 × 10−3 s−1 for 2. Two-layered ONIOM(B3LYP/6-31G(d):LSDA/LANL2MB) studies revealed fluxional behavior of 1 with rather small barriers of activation of the rearrangements. Four possible isomers have been computed for 2, close to each other energetically.  相似文献   

14.
The self-complementary oligonucleotide CGCATATATGCG was used as a model to establish the binding interactions of antitumor molybdenocene dichloride and DNA. The free dodecamer was first characterized using 1H, NOESY, and DQF-COSY NMR experiments, which enable to pinpoint the guanines and adenines as well as the cytosines and thymines signals in the aromatic region. Molybdenocene dichloride was characterized in saline and buffer solutions as function of pH by 1H NMR spectroscopy. In 10 mM NaCl/D2O solution at pH of 6.5 and above, Cp2Mo(OD)(D2O)+ is in equilibrium with its dimeric species, [Cp2Mo(μ-OH)2MoCp2]2+. In 25 mM Tris/4 mM NaCl/D2O at physiological pH, a new stable species is formed, coordinated by the buffer, Tris(hydroxymethyl)aminomethane. The interactions of molybdenocene dichloride species with CGCATATATGCG were studied at different pH. At pH 6.5, in 4 mM NaCl/D2O solution, 1H NMR spectra of CGCATATATGCG exhibit downfield shifts in the signals associated mainly to adenines and guanines, upon addition of molybdenocene dichloride. At pH 7.4, in 25 mM Tris/4 mM NaCl/D2O, molybdenocene species causes broadening and small downfield shifts to the purines and pyrimidine signals, suggesting that molybdenocene dichloride can get engaged in binding interactions with the oligonucleotide in a weak manner. 31P NMR spectra of these interactions at pH 7.4 showed no changes associated to Mo(IV)-OP coordination, indicating that molybdenocene–oligonucleotide binding interactions are centered, most likely, on the bases. Cyclic voltammetry titration showed a 4.9% of molybdenocene–oligonucleotide interaction. This implicates that possible binding interactions with DNA are weak.  相似文献   

15.
1,25-Dihydroxyvitamin D3, an endogenous ligand with the highest affinity for the vitamin D receptor (VDR), was labeled with 11C for use in biological experiments. The radionuclide was incorporated via the reaction of [11C]methyllithium on a methyl ketone precursor in tetrahydrofuran at −10 °C. Deprotection of the labeled intermediate yielded 2.5–3 GBq [26,27-11C]1,25-dihydroxyvitamin D3 [11C-1,25(OH)2 D3] with specific radioactivity averaging 100 GBq/μmol at the end of synthesis and HPLC purification. The entire process took 48 min from the end of radionuclide production. In vitro binding experiments in rachitic chick purified VDR demonstrated the high affinity binding of this novel tracer. Thus; 11C-1,25(OH)2 D3 is available for in vivo distribution studies and may be suitable for the positron emission tomography (PET) determination of VDR levels and occupancy in animals and humans.  相似文献   

16.
The reaction of [Re(NMe)Cl3(PPh3)2] with the pentadentate [N3S2] ligand pyN2H2S2---H2 [2,6-bis(2-mercaptophenylamino)dimethylpyridine] (1) in the presence of triethylamine did not yield the anticipated six-coordinate complex [Re(NMe)(η5-pyN2HS2)] (2), but rather resulted in cleavage of the Re(V)=NMe bond. A novel six-coordinate Re(IV) [N3S]/[NS] complex [Re(η4-SC6H4---2-NCH2---C5H3N---C=NC6H4---2-S)(η2-NHC6H4---2-S)] (4) was thus obtained with the simultaneous coordination of 2-aminothiophenol, a dianionic bidentate [NS] donor resulting from the decomposition of the parent ligand and ligand 3, a dianionic tetradentate [N3S] donor formed by partial self-condensation and subsequent oxidation of the parent ligand 1. Crystal data for 4: C25H18N4S3Re·CH2Cl2, monoclinic, space group P21/n, a=9.255(2) Å, b=11.181(2) Å, c=25.316(4) Å, β=97.434(3)°, V=2587.8(7) Å3 and Z=4.  相似文献   

17.
The effect of vitamin C (ascorbate) on oxidative DNA damage was examined by first incubating cells with dehydroascorbate, which boosts the intracellular concentration of ascorbate, and then exposing cells to H2O2. Oxidative DNA damage was estimated by the analysis of 5-hydroxy-2′-deoxycytidine (oh5dCyd) and 8-oxo-7,8-dihydro-2′-deoxyguanosine (oxo8dGuo). The presence of a high concentration of ascorbate (30 mM), compared to the absence of ascorbate in cells, when exposed to H2O2 (200 μM), resulted in a remarkable sensitization of oh5dCyd from 2.7 ± 0.6 to 40.8 ± 6.1 lesions /106 dCyd (15-fold). In contrast, the level of oxo8dGuo increased from 8.4 ± 0.4 to 12.1 ± 0.5 lesions/106 dGuo (50%). The formation of oh5dCyd was also observed at lower concentrations of intracellular ascorbate and exogenous H2O2. Additional studies showed that replacement of H2O2 with tert-butyl hydroperoxide completely abolished damage, and that preincubation with iron and desferroxamine increased and decreased this damage, respectively. The latter studies suggest that a Fenton reaction is involved in the mechanism of damage. In conclusion, we report a novel model system in which ascorbate sensitizes H2O2-induced oxidative DNA damage in cells, leading to elevated levels of oh5dCyd and oxo8dGuo, with a strong bias toward the formation of oh5dCyd.  相似文献   

18.
Three series of new cannabinoids were prepared and their affinities for the CB1 and CB2 cannabinoid recptors were determined. These are the 1-methoxy-3-(1′,1′-dimethylalkyl)-, 1-deoxy-11-hydroxy-3-(1′,1′-dimethylalkyl)- and 11-hydroxy-1-methoxy-3-(1′,1′-dimethylalkyl)-Δ8-tetrahydrocannabinols, which contain alkyl chains from dimethylethyl to dimethylheptyl appended to C-3 of the cannabinoid. All of these compounds have greater affinity for the CB2 receptor than for the CB1 receptor, however only 1-methoxy-3-(1′,1′-dimethylhexyl)-Δ8-THC (JWH-229, 6e) has effectively no affinity for the CB1 receptor (Ki=3134±110 nM) and high affinity for CB2 (Ki=18±2 nM).  相似文献   

19.
Seeking insight into the possible role of estrogens in prostate cancer (PCa) evolution, we assayed serum E2, estrone (E1), and estrone sulfate (E1S) in 349 PCa and 100 benign prostatic hyperplasia (BPH) patients, and in 208 control subjects in the same age range (50–74 years).

E1 (pmol/L ± S.D.) and E1S (nmol/L ± S.D.) in the PCa and BPH patients (respectively 126.1 ± 66.1 and 2.82 ± 1.78, and 127.8 ± 56.4 and 2.78 ± 2.12) were significantly higher than in the controls (113.8 ± 47.6 and 2.11 ± 0.96). E2 was not significantly different among the PCa, BPH, and control groups. These assays were also carried out in PCa patients after partition by prognosis (PSA, Gleason score (GS), histological stage, and surgical margins (SM)). Significantly higher E1S levels were found in PCa with: PSA > 10 ng/L (3.05 ± 1.92) versus PSA ≤ 10 ng/mL (2.60 ± 1.55), stage pT3-T4 (2.99 ± 1.80) versus pT2 (2.58 ± 1.58), and positive (3.26 ± 1.95) versus negative margins (2.52 ± 1.48). E1 was higher in poor- than in better-prognosis PCa. E2 was significantly higher in PCa with GS ≥ 4 + 3 (109.5 ± 43.8) versus GS ≤ 3 + 4 (100.6 ± 36.5) and increased significantly when GS increased from 3 + 3 to 4 + 4. Estrogens, especially E1S appeared to be possible markers of PCa progression.

Attempting to identify potential sources of E2 in PCa according to prognosis, as well as in BPH, we found a significant correlation coefficient between E1S and E2 (0.266–0.347) in poor-prognosis PCa and no correlation in BPH (0.026) and better-prognosis PCa (0.013–0.104).

It is as though during progression of PCa from good to poor prognosis there were a shift in the E1 to E2 metabolic pathway from predominantly oxidative to predominantly reductive.  相似文献   


20.
The methanothermal reactions of M(CO)6 (M = Mo, W) with Na2S2 gave a series of homonuclear clusters [{M(CO)4}n(MS4)]2− (M=Mo, W; N=1, 2), i.e. (Ph4P)2[(CO)4Mo(MoS4)] (I), (Ph4P)2[(CO)4W(WS4)] (II), (Ph4P)2[(CO)4Mo(MoS4)Mo(CO)4] (III) and (Ph4P)2[(CO)4W(WS4)W(CO)4] (IV). The two dimers, I and II, as well as the two trimers, III and IV, are isostructural to each other, respectively. All compounds crystallize in the triclinic space group with Z=2. The cell dimensions are: a=12.393(8), b=19.303(9), c=11.909(6) Å, =102.39(5), β=111.54(5), γ=73.61(5)°, V=2522(3) Å3 at T=23 °C for I; a=12.390(3), b=19.314(4), c=11.866(2) Å, =102.66(2), β=111.49(1), γ=73.40(2)°, V=2511(1) Å3 at T=23 °C for II; a=11.416(3), b=22.524(4), c=10.815(4) Å, =91.03(2), β=100.57(3), γ=88.96(2)°, V=2733(1) Å3 at T=−100 °C for III, a=11.498(1), b=22.600(4), c=10.864(3) Å, =90.92(2), β=100.85(1), γ=88.58(1)°, V=2771(2) Å3 at T=23 °C for IV. The dimers are each formed by the coordination of the tetrathiometalate as a bidentate chelating ligand to an M(CO)4 fragment while addition of another M(CO)4 fragment to the dimers results in the trimers. All compounds contain both tetrahedral and octahedral metal centers with the formal 6+ and 0 oxidation states, respectively.  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