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 93 毫秒
1.
Four 9,10-anthraquinones (AQ) mono- or bis-substituted with the -NH(CH2)2 NH(CH2)2OH group were studied. 1-AQ, 1,5-AQ and 1,8-AQ but not 1,4-AQ (100°M) generated pBR322 plasmid DNA single strand breaks in the presence of purified NADPH dependent cytochrome P450 reductase. 1-AQ, 1,5-AQ and 1,8-AQ (at 100 °M) stimulated hydroxyl radical formation in MCF-7 S9 cell fraction (as measured by dimethyl pyrolline N-oxide spin trapping) and MCF-7 DNA strand breaks as measured by alkaline filter elution. In contrast 1,4-AQ did not stimulate hydroxyl radical formation and produced considerably less strand breaks in MCF-7 cells compared to the other AQ's. It would appear that the position of the -NH(CH2)2 NH(CH2)2OH groups on the chromophore is an important determinant in the metabolic activation of cytotoxic anthraquinones. This may contribute to the cytotoxicity (ID50 values) of 1-AQ (0.06 °M), 1-8-AQ (0.5 °M) and 1,5-AQ (12.3 °M) but not the 1,4-AQ (1.2 °M).  相似文献   

2.
The pincer ligands 2,6-H3C5N(CH2NR2)2, LR, have been studied in their reaction towards CuCl2 and CuCl. For CuCl2, the case R=Et gives square-pyramidal (η3-LEt)CuCl2 with an apical Cu---Cl distance 0.27 Å longer than the equatorial one. For R=iPr, the chloride-loss product (η3-LiPr)CuCl+ is established as its CuCl4 2− salt. The mer geometry of the ligand in these two compounds is intolerable for Cu(I), and a ligand-redistribution product from CuCl is (η2-LMe)2Cu+, together with linear CuCl2 −. Density functional theory (DFT) calculations of monomeric (LMe)Cu(I)Lq with L=MeCN, C2H4 or Cl show a distinct tendency for one or both NMe2 arms to dissociate from Cu(I), while the Cu(II) analogs adopt planar geometry.  相似文献   

3.
The isomers n- and iso-butyraldehyde (BuA) in combination with Cu(II) induced single and double strand breaks in PM2 DNA, whereas the aldehydes, or Cu(II) alone had only negligible effect. The DNA damage was the result of radical oxidations of the aldehydes under formation of Cu(I). Cu(I) formation was independent of molecular oxygen. Extensive DNA degradation was only observed in the presence of molecular oxygen. Characterization of DNA damage pointed to different ultimate DNA damaging species. While catalase and neocuproine inhibited strand break formation induced by iso-BuA/Cu(II) to a high degree, these inhibitors were less effective in the n-BuA/Cu(II) reaction. On the other hand, sodium azide showed a high strand break inhibition in the n-BuA/Cu(II) reaction, but low inhibition in the iso-BuA/Cu(II) reaction. 2-Deoxyguanosine was hydroxylated in the 8-position by iso-BuA/Cu(II) but little reaction occurred with n-BuA/Cu(II). Chemiluminescence was detected during both BuA/Cu(II) reactions, whereby the intensity of the luminescence signal was 3.5-fold higher for n-BuA/Cu(II) than for iso-BuA/Cu(II). We suppose that the copper(II)-driven oxidation of n- and iso-BuA proceeds via different pathways with different DNA damaging consequences. Whereas the oxidation of iso-BuA mainly results in damage by ·OH-radicals, the oxidation of n-BuA may lead to a radical reaction chain whereby excited states are involved and the resulting DNA-damaging species are not ·OH-radicals.  相似文献   

4.
In vitro copper (II) complex presents antimitotic effects. In this work, we have studied the in vivo seasonal toxic effects of copper (II), ligand (H2L) and the complex [Cu(H2L)(H2O)2]Cl2·4H2O in male Swiss mice. During spring, an i.p. injection of CuCl2 in aqueous NaCl (9 g·l-1) up to 0.05 µmol·kg-1 b.w. (body weight) killed 60% of the rodents after 6 days. LD100 was up to 0.3 µmol·kg-1; H2L was well tolerated, while the complex was 30% lethal with 50 µmol·kg-1. In autumn, mice were less sensitive to CuCl2, and both ligand and complex were equally tolerated and this leads to the conclusion that, in vivo, chronotoxicities of copper (II) and complex in NaCl aqueous solutions are quite different in spring and autumn seasons.  相似文献   

5.
Wu LT  Chu CC  Chung JG  Chen CH  Hsu LS  Liu JK  Chen SC 《Mutation research》2004,556(1-2):75-82
The effect of tannic acid (TA), gallic acid (GA), propyl gallate (PA) and ellagic acid (EA) on DNA damage in human lymphocytes induced by food mutagens [3-amino-1-methyl-5H-pyrido (4,3-b) indole (Trp-P-2) and 2-amino-1-methyl-6-phenylimadazo (4,5-b) pyridine (PhIP) or H2O2 was evaluated by using single-cell electrophoresis (comet assay). The toxicity of these tested compounds (0.1–100 μg/ml) on lymphocytes was not found. These compounds did not cause DNA strand breaks at lower concentrations of 0.1–10 μg/ml. At a concentration of 100 μg/ml, TA and GA exhibited slight DNA damage, whereas PA and EA showed no DNA strand breaks. TA and its related compounds decreased the DNA strand breaks induced by Trp-P-2, PhIP or H2O2 at concentrations of 0.1–10 μg/ml. DNA repair enzymes endonuclease III (Endo III) and formamidopyrimidine-DNA glycoslase (FPG)] were used to examine the levels of oxidised pyrimidines and purines in human lymphocytes induced by H2O2. All the compounds at 10 μg/ml can reduce the level of FPG sensitive sites. However, only EA inhibited the formation of EndoIII sensitive sites. The results indicated that these compounds can enhance lymphocytes resistance towards DNA strand breaks induced by food mutagens or H2O2 in vitro.  相似文献   

6.
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   

7.
A microporous coordination polymer formulated as {[Cu(L)2] · (DMF)2}n (1) has been prepared by the direct reaction of copper nitrate with 6-quinolinecarboxylic acid (HL) in DMF. X-ray single crystal diffraction of 1 reveals that the [Cu2(COO)4] secondary building units are interconnected by the bridging L ligands to generate a layered framework with the terminal L ligands as lateral pendants at both sides. Furthermore, the unusual inserted integration of the coordination layers, regulated and fixed by interlayer aromatic stacking interactions between the terminal ligands, leads to the formation of a novel 3-D microporous crystalline lattice with different 1-D channels along three directions. The gas adsorption and magnetic character of this crystalline material have also been investigated.  相似文献   

8.
Ceruloplasmin enhances DNA damage induced by hydrogen peroxide in vitro   总被引:3,自引:0,他引:3  
Ceruloplasmin (Cp) was found to promote the oxidative damage to DNA, as evidenced by the formation of 8-hydroxy-2'-deoxyguanosine and strand breaks, when incubated with H2O2 in vitro. The capacity of Cp to enhance oxidative damage to DNA was inhibited by hydroxyl radical scavengers such as sodium azide and mannitol, a metal chelator, diethylenetriaminepenta-acetic acid, and catalase. Although the oxidized protein resulted in an increase in the content of carbonyl groups, the ferroxidase activity and the proteolytic susceptibility were not significantly altered. The release of a portion of Cu from Cp was observed, and conformational alterations were indicated by the changes in fluorescence spectra. Based on these results, we suggest that damage to DNA is mediated in the H2O2/Cp system via the generation of ·OH by released Cu2+ and/or loosely bound Cu exposed from oxidatively damaged Cp through the conformational change. The release of Cu from Cp during oxidative stress could enhance the formation of reactive oxygen species and could also potentiate cellular damage.  相似文献   

9.
The preparation of N-, S- and O-donor ligand adducts with CuX+(HX=6-methyl-2-formylpyridinethiosemicarbazone (6HL); 2-formylpyridine-2-methylthiosemicarbazone (2′L); 2-formylpyridine-4′-methylthiosemicarbazone (4′HL)) is described. The N-donors, 2,2′-bipyridyl (bipy), 4-dimethylaminopyridine (dmap) give the complexes [Cu(6L)(bipy)]PF6, [Cu(6L)(bipy)]Cl·5H2O, [Cu(4′L)(bipy)]PF6, [Cu(6L)(dmap)2]PF6·2.5 H2O and [Cu(4′L)(dmap)2]PF6·H2O which have been characterized by physical and spectroscopic techniques. Pentafluorothiophenolate (pftp) gives S-donor complexes [CuX(pftp)] (X=6L and 4′L) and thiolato co-ordination is proposed on the basis of spectroscopic evidence. Paratritylphenolate (ptp) and HPO2−4 give O-donor complexes [Cu(6L)(ptp)], [Cu(4′L)(ptp)], [{Cu(6L)}2HPO4]·4H2O, and [{Cu(4L)}2HPO4]·5H2O which have been characterized by physical and spectroscopic techniques, as have the precursor complexes [Cu(6L)(CH3COO)]·H2O, [Cu(4′L)(CH3COO)], Cu(6HL)(CF3COO)](CF3COO)·0.5H2O, [Cu(4′HL)(CF3COO)](CF3COO), [Cu(2′L)Cl2] and [Cu(2′L)(NO3)2]. Protonation constants for the ligands and some of their complexes have been determined. 2-Formylpyridinethiosemicarbazone (HL) complexes of silver, gold, zinc, mercury, cadmium and lead are also discussed. Cytotoxicity against the human tumor cell line HCT-8 and antiviral data for selected compounds are presented.  相似文献   

10.
The kinetics and equilibria of complex formation by Ga(III) with NCS in aqueous solution have been measured over a range of acidities and temperatures, the contributing paths to the reaction resolved, and their rate constants and activation parameters determined. The hydrolysis equilibria required to carry out this resolution of kinetic behaviour have also been measured.

Unlike the other reported complexation reactions of Ga(III) in aqueous solution, the separate reaction pathways can be assigned with no ambiguity. At 25 °C and ionic strength 0.5 M, the observed forward rate constant for the complex formation is described by {k1 + k2K1h/[H+] + k3K1hK2h/[H+]2} M−1 s−1. For these conditions, the first and second successive hydrolysis constants of Ga(H2O)63+ are given by pK1h = 3.69 ± 0.01 and pK2h = 3.74 ± 0.04. The rate constants corresponding to the reactions of the species Ga(H2O)63+, Ga(H2O)5(OH)2+ and Ga(H2O)4(OH)2+ with NCS are k1 = 57 ± 4 M−1 −1, k2 = (1.08 ± 0.01) × 105 M−1 s−1 and k3 = 3 × 106 M−1 s−1 respectively. The complexation equilibrium quotient [GaNCS2+]/([Ga3+][NCS]) has been independently determined by spectrophotometric titration to be 20.8 ± 0.3 M−1 at 25 °C and ionic strength 0.5 M.

These kinetic results lead to an interpretation of the data, and a reinterpretation of other data for aquo-Ga(III) complex formation kinetics from the literature which support the assignment of a dissociative interchange mechanism for these reactions rather than the associative activation mode sometimes proposed.  相似文献   


11.
In vitro, 3-(2-furyl) prop-2-enal semicarbazone-copper (II) complex [CuCl2(FASC)2] presents antimitotic effects. In this work we studied the in vivo seasonal toxic effects in male Swiss mice of CuCl2, and FASC and the [CuCl2(FASC)2] complex. In spring, one injection of CuCl28.10-2 mmol killed 16% of animals after 24 h. Cupric chloride lethal dose was up to 64.10-2 mmol with 100% mice dead after 24 h. FASC was well tolerated from 0.65 to 1.3 mmol. The complex was 100% lethal with 48.10-2 mmol. In autumn, mice were more sensitive to CuCl2 and to the complex with lethal doses up to 32.10-2 mmol and 8.10-2 mmol, respectively. On the other hand, FASC was well tolerated. It is concluded that the in vivo toxic effects of CuCl2 and [CuCl2(FASC)2] complex are quite different in spring and autumn.  相似文献   

12.
The reaction of dilithiated o-carborane (closo-1,2-Li2-1,2-C2B10H10) with CuCl2 gives 1,1′-bis(o-carborane) (1), 1,3′-bis(o-carborane) (2) and 1,4′-bis(o-carborane) (3). Compound 2 (C4B20H22) crystallizes in the monoclinic space group P21/n with A = 6.9275(6), B = 9.7655(8), C = 12.356(1) Å, β = 90.028(2)° and Z = 2. The structure was solved by direct methods and refined to R = 0.048 and Rw = 0.074. Compound 3 (C4B20H22) crystallizes in the orthorhombic space group P212121 with A = 6.8854(5), B = 12.523(1), C = 19.847(1) Å and Z = 4. The structure was solved by direct methods and refined to R = 0.078 and Rw = 0.091. The coupling reaction of dilithiated m-carborane (closo-1,7-Li2-1,7-C2B10H10) with CuCl2 results in the formation of 1,1′-bis(m-carborane) (4) and tetra(m-carborane) (5).  相似文献   

13.
Hydroxyl radical (·OH) generation in the kidney of mice treated with ferric nitrilotriacetate (Fe-NTA) or potassium bromate (KBrO3) in vivo was estimated by the salicylate hydroxylation method, using the optimal experimental conditions we recently reported. Induction of DNA lesions and lipid peroxidation in the kidney by these nephrotoxic compounds was also examined. The salicylate hydroxylation method revealed significant increases in the ·OH generation after injection of Fe-NTA or KBrO3 in the kidneys. A significant increase in 8-hydroxy-2'-deoxyguanosine in nuclei of the kidney was detected only in the KBrO3 treated mice, while the comet assay showed that the Fe-NTA and KBrO3 treatments both resulted in significant increases in DNA breakage in the kidney. With respect to lipid peroxidation, the Fe-NTA treatment enhanced lipid peroxidation and ESR signals of the alkylperoxy radical adduct. These DNA breaks and lipid peroxidation mediated by ·OH were diminished by pre-treatment with salicylate in vivo. These results clearly demonstrated the usefulness of the salicylate hydroxylation method as well as the comet assay in estimating the involvement of ·OH generation in cellular injury induced by chemicals in vivo.  相似文献   

14.
The formation of complexes between copper(II) halides and 2,2′-dipyridylamine (dipyam) has been studied systematically. Only complexes with a 1:1 and 1:2 metal-to-ligand ratio are formed. Some mixed chloro–iodide and halide–PF6 compounds have also been isolated. The X-ray diffraction structures of the [Cu(dipyam)2Br2] · 2H2O (I) and the [Cu(dipyam)2Cl]2I2 · 2CH3CN (II) complexes are reported. I is a rare example of an octahedral coordination among the copper(II) halide complexes of dipyam. The two bromo atoms, which occupy the apical positions, are H-bonded to the water molecules of crystallization. II is a dimer, where each copper forms a cationic chloro-complex of approximately trigonal bipyramidal geometry, the dimerization being due to hydrogen bonds formed by the NH group of one of the two dipyams coordinated to each metal atom with the chlorine atom of the centrosymmetric cationic complex. The iodide anions are hydrogen-bonded to the NH groups of the dipyams not involved in the dimerization.  相似文献   

15.
4-Hydrazinobenzoic acid, an ingredient of mushroom Agaricus bisporus, is carcinogenic to rodents. To clarify the mechanism of carcinogenesis, we investigated DNA damage by 4-hydrazinobenzoic acid using 32P-labeled DNA fragments obtained from the human p53 and p16 tumor suppressor genes. 4-Hydrazinobenzoic acid induced Cu(II)-dependent DNA damage especially piperidine-labile formation at thymine and cytosine residues. Typical hydroxyl radical scavengers showed no inhibitory effects on Cu(II)-mediated DNA damage by 4-hydrazinobenzoic acid. Bathocuproine and catalase inhibited the DNA damage, indicating the participation of Cu(I) and H2O2 in the DNA damage. These findings suggest that H2O2 generated by the autoxidation of 4-hydrazinobenzoic acid reacts with Cu(I) to form reactive oxygen species, capable of causing DNA damage. Interestingly, catalase did not completely inhibit DNA damage caused by a high concentration of 4-hydrazinobenzoic acid (over 50 μM) in the presence of Cu(II). 4-Hydrazinobenzoic acid induced piperidine-labile sites frequently at adenine and guanine residues in the presence of catalase. 4-Hydrazinobenzoic acid increased formation of 8-oxo-7,8-dihydro-2'-deoxyguanosine (8-oxodG), a characteristic oxidative DNA lesion, in calf thymus DNA, whereas 4-hydrazinobenzoic acid did not increase the formation of 8-oxodG in the presence of catalase. ESR spin-trapping experiments showed that the phenyl radical was formed during the reaction of 4-hydrazinobenzoic acid in the presence of Cu(II) and catalase. Matrix-assisted laser desorption/ionization time-of-flight mass (MALDI-TOF/mass) spectrometry analysis showed that phenyl radical formed adduct with adenosine and guanosine. These results suggested that 4-hydrazinobenzoic acid induced DNA damage via not only H2O2 production but also phenyl radical production. This study suggests that both oxidative DNA damage and DNA adduct formation play important roles in the expression of carcinogenesis of 4-hydrazinobenzoic acid.  相似文献   

16.
Unlike ZrCl4, ZrBr4 is not ammonolysed in liquid ammonia at temperatures up to −33 °C. The existence of ammoniates ZrBr4nH3 (n = 17, 12 and 9) at −36 °C has been established; at room temperature, the hexammine ZrBr4 · 6NH3 is the stable species which becomes ZrBr4 · 2NH3 at 200 °C. When treated with an excess of NH4CN in liquid ammonia, complete replacement of bromide ions by cyanide occurs to give an inseparable mixture of Zr(CN)4 · 2NH3 and NH4Br. The chloride and bromide of zirconium(III) also undergo no ammonolysis in liquid ammonia; the ammoniates stable at room temperature are ZrCl3 · 2.5NH3 and ZrBr3 · 6NH3.  相似文献   

17.
The stepwise synthesis of mononuclear (4f) and heterodinuclear (3d–4f) Salen-like complexes has been investigated through structural determination of the intermediate and final products occurring in the process. In the first step, reactions of ligand H2L and Ln(NO3)3 · 6H2O give rise to three mononuclear lanthanide complexes Ln(H2L)(NO3)3 [H2L = N,N′-ethylene-bis(3-methoxysalicylideneimine), Ln = Nd (1), Eu (2) and Tb (3)], in which N,N′-ethylene-bis(3-methoxysalicylideneimine) acts as tetradentate ligands with the O2O2 set of donor atoms capable of effective coordination. These species are fairly stable and have been isolated. Then, addition of Cu(Ac)2 · H2O to the mononuclear lanthanide complex yields expected heterodinuclear (3d–4f) complexes Cu(L)Ln(NO3)3 · H2O [Ln = Nd (4) and Eu (5)] where the Cu(II) ion is inserted to the inner N2O2 cavity. Luminescent analysis reveals that complex 3 exhibits characteristic metal-centered fluorescence of Tb(III) ion. However, the characteristic luminescence of both Sm(III) and Eu(III) ions is not observed both in solution and solid state of the complexes.  相似文献   

18.
In this work we present the synthesis, structural and spectroscopic characterisation of a series of cobalt(II) and zinc(II) coordination compounds with benzimidazole (bz) and its 2-benzimidazole derivatives: 2-aminobenzimidazole (2ab), albendazole (abz) and tris(2-benzimidazolylmethyl)amine (ntb). The compounds were evaluated for their in vitro antimicrobial activity against Staphylococcus aureus, Micrococcus luteus, Salmonella typhi, Pseudomonas aeruginosa, Escherichia coli and Proteus vulgaris. Their cytotoxic activity was also evaluated using human cancer lines, HeLa, HCT-15 and SKLU-1. The halide tetrahedral compounds [Co(bz)2Br2] 3, [Zn(2ab)2Cl2] · 0.5H2O 11, [Co(abz)Cl2(H2O)] · 3H2O 14, [Co(abz)Br2(H2O)] 15, [Zn(abz)Cl2(H2O)] · 3H2O 17 and [Zn(abz)Br2(H2O)] · H2O 18 displayed similar minimal inhibition concentration (MIC) values against Micrococcus luteus and Escherichia coli, comparable to those of amoxicillin and chloramphenicol. Additionally, 11 showed a wide range of activity towards Gram(+) and Gram(−) microorganisms. The tetradentate ntb and its trigonal bipyramidal cobalt(II) and zinc(II) compounds were active, regardless of the anion present in the complex. Compound [Co(abz)Cl2(H2O)] · 3H2O 14 showed promising activity in HeLa cells, while [Co(ntb)Br]Br · H2O 21 inhibited Hela and HCT-15 cell lines.  相似文献   

19.
The genotoxic effects of 2,4-D and its commercial derivative 2,4-D DMA were studied by measuring sister chromatid exchange (SCE), cell-cycle progression and mitotic index in human whole blood (WBC) and plasma leukocyte cultures (PLC). Concentrations of 10, 25, 50 and 100 microg herbicide/ml were used during 72 h. In WBC, a significant increase in SCE frequency was observed within the 10-50 microg 2,4-D/ml and 25-100 microg 2,4-D DMA/ml dose range. Contrarily, in PLC, none of the concentrations employed affected the SCEs frequency. A significant delay in cell proliferation was observed in WBC after treatments with 25 and 50 microg 2,4-D/ml and 50 and 100 microg 2,4-D DMA/ml. In PLC, only 100.0 microg 2,4-D/ml altered cell-cycle progression. For both chemicals, a progressive dose-related inhibition of mitotic activity was observed. The results demonstrated that the presence of erythrocytes in the culture system modulated the DNA and cellular damage inflicted by 2,4-D and 2,4-D DMA into human lymphocytes in vitro as well as both 2,4-D and 2,4-D DMA were more potent genotoxic agents in the presence of human red cells.  相似文献   

20.
Two new copper(II) complexes, Cu(L1)(ClO4)2 (1), {[(μ-oxalate)Cu(L1)] · 5H2O}n (2), and a zinc(II) complex, {[(μ-oxalate)Zn(L2)] · 3H2O · 0.5DMF}n (3) (L = 3,14-dimethyl-2,6,13,17-tetraazatricyclo[14,4,01.18,07.12]docosane), have been synthesized and characterized by X-ray crystallography. In 1, the ligand conformation is planar, and the octahedral coordination about the copper(II) ion is completed by weakly interacting ions. In 2 and 3, bridging oxalate ligands coordinate to copper(II) or zinc(II) ions in an unusually twisted bis-monodentate (trans-1,1′-bicoordination) mode.

The rigidity and steric hindrance of macrocycles L1 and L2 by the introduction of two cyclohexane rings and methyl groups on a cyclam (1,4,8,11-tetraazacyclotetradecane) skeleton cause the bridging oxalate ligands to adopt such unusual geometries in 2 and 3.  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