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.
Silver(I) halides react with tri(p-tolyl)phosphine (tptp, C21H21P) in MeOH/MeCN solutions in 1:1 or 1:3 molar ratios to give complexes of formulae {[AgCl(tptp)]4} (1) or [AgX(tptp)3] (X = Cl (2), Br (3), I (4)), respectively. The complexes were characterized by elemental analyses, and FT-IR far-IR, FT-Raman, TG and 1H, 13C, 31P NMR spectroscopic techniques. Crystal structures of complexes 2-4 were determined by X-ray diffraction at room temperature (rt). The crystal structure of 1 and 4 was also determined at 100(1) and 140(2) K (lt), respectively. In complex 1 four μ3-Cl ions are bonded with four Ag(I) ions forming a cubane while the coordination sphere of silver(I) ions is completed by one P atom from a terminal tri(p-tolyl)phosphine ligand. In complexes 2-3 one terminal halogen and three P atoms from phosphine ligands form a tetrahedral arrangement around the metal ion. Complexes 1-4 were tested for in vitro cytostatic activity against sarcoma cancer cells (mesenchymal tissue) from the Wistar rat, polycyclic aromatic hydrocarbons (PAH, benzo[a]pyrene) carcinogenesis and against murine leukemia (L1210) and human T-lymphocyte (Molt4/C8 and CEM) cells. The silver(I) complexes 1-4 show strong activity.  相似文献   

2.
The rhodium(I) complexes TpmsRh(CO)2 (1) and TpmsRh(cod) (2) of the tripodal nitrogen ligand tris(pyrazolyl)methanesulfonate, Tpms=[(pz)3CSO3], catalyze the hydroformylation of 1-hexene. Addition of phosphine has a negative effect on the activity. The hydroformylation activity reaches a maximum at about 60 °C. At temperatures above 80 °C hydrogenation becomes an important secondary reaction. When the catalysis is performed at 60 °C in acetone with 1 or 2 as catalyst precursor all of the rhodium is recovered in the form of the rhodium(III) bis(acyl) complex TpmsRh(CO)(COC6H13)2 (9). A similar behaviour is observed with rhodium(I) complexes bearing the tripodal oxygen ligand LOMe=[(cyclopentadienyl)tris(dimethylphosphito-P) cobalt O,O,O″]. In this case all of the rhodium is transformed into LOMeRh(CO)(COC6H13)2 (10). These hitherto unknown bis(acyl) rhodium(III) complexes show the same catalytic activity as the rhodium(I) starting compounds.  相似文献   

3.
Two new mixed ligand silver(I) complexes of formulae {[Ag(tpp)3(asp)](dmf)} (1) (aspH = o-acetylsalicylic acid and tpp = triphenylphosphine) and [Ag(tpp)2(o-Hbza)] (2) (o-HbzaH = o-hydroxy-benzoic acid) were synthesized and characterized by elemental analyses, spectroscopic techniques and X-ray crystallography at ambient conditions. Three phosphorus and one carboxylic oxygen atoms from a de-protonated aspirin ligand in complex 1 and two phosphorus and two carboxylic oxygen atoms from a chelating o-Hbza anion in complex 2 form a tetrahedral geometry around Ag(I) ions in both complexes.Complexes 1 and 2 and the silver(I) nitrate, tpp, aspNa and o-HbzaH were tested for their in vitro cytotoxic activity against leiomyosarcoma cells (LMS), human breast adenocarcinoma cells (MCF-7) and normal human fetal lung fibroblasts (MRC-5) cells with Thiazolyl Blue Tetrazolium Bromide (MTT) assay. For both cell lines 1 and 2 were found to be more active than cisplatin. Additionally, 1 and 2 exhibit lower activity on cell growth proliferation of MRC-5 cells. The type of LMS cell death caused by 1 and 2 were evaluated in vitro by use of flow cytometry assay. The results show that at concentrations of 1.5 and 1.9 μΜ of complex 1, 44.1% and 69.4%, respectively of LMS cells undergo programmed cell death (apoptosis). When LMS cells were treated with 1.6 and 2.3 μM of 2, LMS cells death was by 29.6% and 81.3%, respectively apoptotic. Finally, the influence of the complexes 1 and 2, upon the catalytic peroxidation of linoleic acid to hydroperoxylinoleic acid by the enzyme lipoxygenase (LOX) was kinetically and theoretically studied. The binding of 1 and 2 towards LOX was also investigated by Saturation Transfer Difference (STD) 1H NMR experiments.  相似文献   

4.
Formation of three Cd(II)-ethylenediamine (en) complexes ([Cd(en)n]2+, n = 1-3) in aqueous solution and in DMSO solvent has been established by means of 113Cd NMR spectroscopy. It is clearly shown that Cd(II)-en complexes form primarily in basic solutions. A correlation between the 113Cd NMR chemical shifts and the ethylenediamine (en) coordination number has been observed and discussed. Two single crystals with the composition [Cd2(en)5](ClO4)4 (1) and [Cd(en)3](ClO4)2 (2) were prepared from aqueous solution, and their structures were determined by single crystal X-ray diffraction. Cd(II) ions are coordinated by six atoms in both compounds, 1 and 2: via five N-donor atoms and one O-donor atom forming a bimetallic complex 1, and via six N-donor atoms forming a distorted octahedral monometallic complex 2. Raman spectra of complexes 1 and 2 also provide additional evidence that the cis-form of the bridging en is present in complex 1.  相似文献   

5.
Three new heteroleptic Cu(I) complexes containing one phenanthroline and one diphosphine type ligand ([Cu(N-N)(P-P)]+) have been prepared. In particular, one ligand is constituted by 1,10-phenanthroline (1), 2,9-dimethyl-1,10-phenanthroline (2) and 2,9-diphenethyl-1,10-phenanthroline (3) and the other ligand is in all cases 1,1′-bis(diphenylphosphino)ferrocene (dppf). Therefore, copper and iron metal centres are quite close one another, as evidenced by X-ray crystal diffraction. The structure together with the electrochemical and photophysical properties of these complexes have been compared to that of the corresponding complexes where dppf has been replaced by bis[2-(diphenylphosphino)-phenyl]ether (POP). Cyclic voltammetric experiments evidenced that the first oxidation process is located on the ferrocene moiety and that oxidation of Cu(I) is moved to more positive potential values and a chemical reaction is coupled to the electron transfer process. The absorption spectra show a metal-to-ligand charge transfer (MLCT) band, typical of Cu(I) phenanthroline complexes, at a higher energy compared to the homoleptic [Cu(N-N)2]+ species. No emission at either room temperature or 77 K has been observed for compounds 2 and 3, contrary to the high luminescence observed for the corresponding POP complexes. This result is consistent with a photoinduced energy transfer from the Cu(I) complex to the ferrocene moiety.  相似文献   

6.
The bulky phosphine ligands di-tert-butyl(1-naphthyl)phosphine (1) or di-tert-butyl(N-indolyl)phosphine (2) react at room temperature with [(μ-SMe2)PtMe2]2. Coordination of the phosphine and C-H bond activation at an sp2 carbon of the ligand with the release of methane takes place to form the PC cyclometalated products [(PC)PtMe(SMe2)] (3 or 4, respectively). The cyclometalated complexes 3 and 4 have both been characterized by X-ray crystallography. Complexes 3 and 4 were each observed to undergo intermolecular activation of arene C-H bonds. Upon thermolysis in benzene, complexes 3 and 4 react to eliminate methane and yield isolable platinum(II)-phenyl complexes.  相似文献   

7.
Reaction of the ligands diphenylphosphinylacetic acid Ph2P(O)CH2COOH (1) and 2-(tert-butylthio)phenyldiphenylphosphine oxide Ph2P(O)C6H4tBuS (2) with “MoO2Cl2”, resulted in two complexes MoO2Cl2Ph2P(O)CH2COOH (3) and MoO2Cl2Ph2P(O)C6H4tBuS(O) (4). Complexes 3 and 4 were isolated and analysed by 1H NMR, 31P NMR and X-ray crystallography. Complex 3 crystallised with a molecule of the free ligand in a 1:1 ratio (3·1) and complex 4 crystallised with molecules of the solvent CH2Cl2 within the unit cell in a 2:1 ratio (4·0.5CH2Cl2). Tetrameric arrangements comprised of hydrogen bonds were observed in complexes 1 and 3. Complex 4 exhibited a seven-membered ring structure owing to the oxidation of the sulphide in 2 to sulphoxide and coordination of this ligand via the oxygen atoms to the molybdenum atom.  相似文献   

8.
The distorted square-planar complexes [Pd(PNHP)Cl]Cl (1) (PNHP = bis[2-(diphenylphosphino)ethyl]amine), [M(P3)Cl]Cl [P3 = bis[2-(diphenylphosphino)ethyl]phenylphosphine; M = Pd (2), Pt (3)] and [Pt(NP3)Cl]Cl (5) (NP3 = tris[2-(diphenylphosphino)ethyl]amine), coexisting in the later case with a square-pyramidal arrangement, react with one equivalent of CuCl to give the mononuclear heteroionic systems [M(L)Cl](CuCl2) [L = PNHP, M = Pd (1a); L = P3, M = Pd (2a), Pt (3a); L = NP3, M = Pt (5a)]. The crystal structure of 3a confirms that Pt(II) retains the distorted square-planar geometry of 3 in the cation with P3 acting as tridentate chelating ligand, the central P atom being trans to one chloride. The counter anion is a nearly linear dichlorocuprate(I) ion. However, the five-coordinate complexes [Pd(NP3)Cl]Cl (4), [M(PP3)Cl]Cl (M = Pd (6), Pt (7); PP3 = tris[2-(diphenylphosphino)ethyl] phosphine) containing three fused five-membered chelate rings undergo a ring-opening by interaction with one (4, 6, 7) and two (6, 7) equivalents of CuCl with formation of neutral MCu(L)Cl3 [L = NP3, M = Pd (4a); L = PP3, M = Pd (6a), Pt (7a)] and ionic [MCu(PP3)Cl2](CuCl2) [M = Pd (6b), Pt (7b)] compounds, respectively. The heteronuclear systems were shown by 31P NMR to have structures where the phosphines are acting as tridentate chelating ligands to M(II) and monodentate bridging to Cu(I). Further additions of CuCl to the neutral species 6a and 7a in a 1:1 ratio resulted in the achievement of the ionic complexes 6b and 7b with ions as counter anions. It was demonstrated that the formation of heterobimetallic or just mononuclear mixed salt complexes was clearly influenced by the polyphosphine arrangement with the tripodal ligands giving the former compounds. However, complexes [M(NP3)Cl]Cl constitute one exception and the type of reaction undergone versus CuCl is a function of the d8 metal centre.  相似文献   

9.
Two platinum(II) complexes, [Pt(dpa)Cl2] (1) and [Pt(dpa)CBDCA] (2), where DPA=2,2-dipyridylamine and CBDCA=1,1-cyclobutanedicarboxylate, were synthesized and characterized by elemental analysis, IR spectroscopy, ES-MS and X-ray diffraction. Intermolecular hydrogen bonds were observed in both complexes (N-H?Cl for complex 1 and N-H?O for complex 2), which may play a role in formation of hydrogen bonding in metal-DNA adducts. Complex 2 adopts a boat conformation so that the cyclobutane ring and bipyridyl groups are on the same side of the platinum square. The interactions of complexes 1 and 2 with DNA were studied by UV and Fluorescence Spectroscopy, which indicated that both complexes could interact with DNA through groove binding or intercalation. The in vitro cytotoxic activity against melanoma B16-BL6 cells and human Jurkat T-cells was also reported.  相似文献   

10.
Aryloxide rhodium(I) complexes Rh(OAr)(PPh3)3 (1a: Ar=C6Cl5, 1b: Ar=C6F5, 1c: Ar=C6H4-NO2-4) react with CO in toluene solutions to produce Vaska-type complexes trans-Rh(OAr)(CO)(PPh3)2 (2a: Ar=C6Cl5, 2b: Ar=C6F5, 2c: Ar=C6H4-NO2-4). Carbonylation of a similar complex with PMe3 ligands, Rh(OC6H4-NO2-4)(PMe3)3 (3c), also forms trans-Rh(OC6H4-NO2-4)(CO)(PMe3)2 (4c). Molecular structures of the complexes are determined by X-ray crystallography and NMR spectroscopy. Complex 1a reacts with CO in the absence of solvent to produce a mixture of 2a and complex A, the latter of which shows the IR and 13C{1H} signals due to the carbonyl ligand at different positions from those of 2a. Addition of Et2O to the above mixture turns it into analytically pure 2a. Carbonylation of 1b and 1c under the solvent-free conditions produces complexes B and C as the respective products of the solid-gas reaction. Recrystallization of B and C turns them into 2b and 2c, respectively. Complex 3c also reacts with CO in the solid state to form a mixture of 4c and complex D, although the latter complex is converted slowly into 4c even in the solid state.  相似文献   

11.
Four new Ag(I) complexes with three different modes of structures were obtained by varying the counteranions , and their structures characterized by single-crystal X-ray diffraction analysis. Compounds 1, 2, and 3 crystalize in the C-centered monoclinic space group C2/m. Compound 4 crystalizes in the monoclinic space group P21/c. The crystal structures of these complexes show that the complexes 1, 2, and 3 form ligand-supported dinuclear rings, and the dinuclear units of 1 and 3 are further linked by anions to form one-dimensional polymer, while the complex 4 forms an one-dimensional zigzag chain. The structural differences between 1, 2, 3, and 4 show the influences of the counteranions on the structures of the complexes.  相似文献   

12.
The first FeIII complexes 1-6 with cyclin-dependent kinase (CDK) inhibitors of the type [Fe(Ln)Cl3nH2O (n = 0 for 1, 1 for 2, 2 for 3-6; L1-L6 = C2- and phenyl-substituted CDK inhibitors derived from 6-benzylamino-9-isopropylpurine), have been synthesized and characterized by elemental analysis, IR, 57Fe Mössbauer, 1H and 13C NMR, and ES+ mass spectroscopies, conductivity and magnetic susceptibility measurements, and thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC). The study revealed that the compounds are mononuclear, tetrahedral high-spin (S = 5/2) FeIII complexes with an admixture of an S = 3/2 spin state originating probably from five-coordinated FeIII ions either connecting with a bidentate coordination mode of the CDK inhibitor ligand or relating to the possibility that one crystal water molecule enters the coordination sphere of the central atom in a portion of molecules of the appropriate complex. Nearly spin-only value of the effective magnetic moment (5.82 μeff/μB) was determined for compound 1 due to absence of crystal water molecule(s) in the structure of the complex. Based on NMR data and DFT calculations, we assume that the appropriate organic ligand is coordinated to the FeIII ion through the N7 atom of a purine moiety. The cytotoxicity of the complexes was tested in vitro against selected human cancer cell lines (G-361, HOS, K-562 and MCF-7) along with the ability to inhibit the CDK2/cyclinE kinase. The best cytotoxicity (IC50: 4-23 μM) and inhibition activity (IC50: 0.02-0.09 μM) results have been achieved in the case of complexes 2-4, and complexes 3, 4 and 6, respectively. In addition, the X-ray structure of 2-chloro-6-benzylamino-9-isopropylpurine, i.e. a precursor for the preparation of L1, L4 and L5, is also described.  相似文献   

13.
Reaction of CdCl2 with N-alkylaminopyrazole ligands 1-[(2-ethylamino)ethyl]-3,5-dimethylpyrazole (deae), 1-[(2-(tert-butylamino)ethyl)]-3,5-dimethylpyrazole (deat), bis-[(3,5-dimethylpyrazolyl)methyl]ethylamine (bdmae), and bis-[(3,5-dimethylpyrazolyl)ethyl]ethylamine (ddae) in absolute ethanol yields [CdCl2(NN′)] (NN′ = deae (1), deat (2)), [CdCl2(bdmae)] (3), and [CdCl(ddae)]2[CdCl4] (4). The Cd(II) complexes have been characterised by elemental analyses, conductivity measurements, IR, 1H, 13C{1H} and 113Cd NMR spectroscopies, and X-ray diffraction methods. 1H and 113Cd NMR experiments at variable temperature for 3 and 4 show that dynamic processes are taking place in solution. We report the measurements of 113Cd NMR chemical shift data for complexes 1-4 in solution. X-ray crystal structures for complexes 2 and 3 have been determined. The Cd(II) is coordinated to the deat ligand, in 2, by one nitrogen atom of the pyrazolyl group and one nitrogen atom of the amine. It finishes a tetrahedral geometry with two chlorine atoms. The bdmae ligand is linked to Cd(II), in 3, by two nitrogens atoms of the pyrazolyl groups and one amine nitrogen, along with two chlorine atoms, in a distorted trigonal bipyramidal geometry.  相似文献   

14.
A series of para-substituted triaryltin(pentacarbonyl)manganese(I) compounds [(p-XC6H4)3SnMn(CO)5: II, X=CH3; III, X=CH3O; IV, X=CH3S; V, X=F; VI, X=Cl; VII, X=CH3S(O2)] is reported for comparison with the known phenyl analogue I. IR data [ν(CO)] as well as complete 119Sn/55Mn/13C solution NMR results are given for I-VII. Chemical shifts, 119Sn versus 55Mn, except I, correlate well, but have differing single parameter (SP) correlations, 119Sn versus σI and 55Mn versus σ°p. These results are compared with previous SP studies of the 119Sn solution NMR spectra of the series, (p-XC6H4)4Sn and (p-XC6H4)3SnY (Y=Cl, Br, I). Full crystal structures are reported for compounds II-VI. All are similar to that of I, with the Mn(CO)5 moiety being a distorted tetragonal pyramid, and having a quasi-mirror plane through the central C4MnSnC3 skeleton. The Ar3Sn are distorted trigonal propellers with ring torsion angles in the range 30-80°, the exception being IV with one torsion angle of 22°.  相似文献   

15.
Palladium complexes with a salicylaldiminato ligand bearing a hydroxyl group (1a and 1b) have been synthesized and characterized. The structures of these complexes were confirmed by X-ray crystallography. A reversible deprotonation/protonation of the hydroxyl moiety on 1b was observed, while such behaviour was impossible with a related palladium complex (1c) bearing a methoxyl group in place of the hydroxyl group. The deprotonation affected its catalytic behaviour: the activity for polymerization of methyl acrylate catalyzed by 1b considerably decreased in the presence of 1 equiv. of tBuOK.  相似文献   

16.
The reaction of 2-(2-aminophenyl)benzothiazole (Habt) with [Re(CO)5Br] led to the isolation of the rhenium(I) complex fac-[Re(Habt)(CO)3Br] (1). With trans-[ReOCl3(PPh3)2], the ligand Habt decomposed to form the oxofree rhenium(V) complex [Re(itp)2Cl(PPh3)] (2) (itp = 2-amidophenylthiolate). From the reaction of trans-[ReOBr3(PPh3)2] with 2-(2-hydroxyphenyl)benzothiazole (Hhpd) the complex [ReVOBr2(hpd)(PPh3)] (3) was obtained. Complexes 1-3 are stable and lipophilic. 1H NMR and infrared assignments, as well as the X-ray crystal structures, of the complexes are reported.  相似文献   

17.
The objective was to prepare mononuclear, binuclear, and insoluble polynuclear oxovanadium(IV)-Schiff base complexes and to use them for sulfoxidation and epoxidation of organic substrates. [VO(salen)] (complex 1) with tetradentate salen(salicylideneethylenediamine) being coordinated in the equatorial plane of oxovanadium(IV), [VO(salap)] (complex 2), and [(VO)2(sal2-dhdabp)] (complex 3) with tridentate salap(salicylideneorthoaminophenol) and sal2-dhdabp(salicylidene-3,3-dihydroxy-4,4-diaminobiphenyl) being bound, respectively, in the equatorial plane, of which polynuclear complexes were constituted as monomer units, were prepared and spectroscopically characterized. A sulfide and olefins were oxidized by use of complexes 1 and 2 (mononuclear), complex 3 (binuclear), and the polynuclear complexes (poly-1 and poly-3) synthesized with 1 and 3, respectively. The reaction rates for poly-1 and -3 were a little lower than those of the corresponding 1 and 3. On oxidation of sulfides, poly-3 exhibited lowering of activity by about 15% in three cycles, while poly-1 showed significant lose of activity with each use. Poly-3 was efficient for the oxidation of the olefins only in the first cycle. It was suggested that the loss of activity depends not only on the coordination geometry of the oxovanadium complex, but also on the kind of the substrate.  相似文献   

18.
Nickel(0) complexes 1-4 containing π-coordinated olefin and triphenylphosphine (tricyclohexylphosphine) (starting from Ni(cod)2) were prepared and the X-ray structures of 1 and 2 were resolved. The complexes appeared as efficient catalysts in dehydrogenative silylation of styrene and vinyltris(trimethylsiloxy)silane, but only after prior oxygenation of phosphine ligand. Stoichiometric studies of Ni(0) complexes with substrates showed that the bis(silyl)nickel(II) complex was a key intermediate in both reactions examined. A scheme of catalysis by Ni(0) complex involving olefin insertion into Ni-Si bond, as a crucial step, is presented.  相似文献   

19.
The reactions of 4-(p-dimethylaminophenyl)-6-phenyl-2,2′-bipyridine (HL) with three metal salts of platinum(II), copper(I) and zinc(II) provide the new complexes [Pt(L)(PPh3)]ClO4 (1), [Cu(HL)2]BF4 (2), [Cu(HL)(PPh3)]BF4 (3) and [Zn(HL)2](ClO4)2 (4). All the structures of these four complexes have been characterized by single crystal X-ray diffraction, and their spectroscopic properties were investigated. Especially for complex 1, upon protonation, the excited state can be tuned from the intraligand charge transfer (ILCT) to the metal-to-ligand charge transfer (MLCT), and such switching in the excited state is acid/base reversible. The time-dependent density functional theory (TD-DFT) calculation was used to interpret the absorption spectra of complex 1, and the calculated result is consistent with those of experiments results. In contrast with 1, the lowest energy absorption at 410-650 nm of complexes 2 and 3 can be assigned to MLCT excited state. In solid state or solution complex 4 exhibits intense photoluminescence attributed to a ILCT transition in nature.  相似文献   

20.
Protonation and Zn(II) complexation of N-substituted amino acids, valine (H2L1) and glycylglycine (H2L2), with 4-amino-1,6-dihydro-1-methyl-5-nitroso-6-oxopyrimidin-2-yl as substituent, were studied by potentiometric and UV-Vis measurements. Bianions L1 and L2 suffer three protonation steps in aqueous medium corresponding to the amide and carboxylate groups of the amino acidic moiety, and the nitrogen atom of the nitroso group of the pyrimidine fragment. Both ligands form mononuclear Zn(II) complexes in aqueous solutions. The binding donor groups are the nitroso and/or the oxo groups of the pyrimidinic moiety or the carboxylate group, depending on whether the ligands are neutral or anionic, respectively. Weak metal-to-ligand interactions were observed independently of the functionality used by the corresponding ligand on bonding to Zn(II). The reaction of ZnCl2 with the monodeprotonated ligands (1:1) yields a polynuclear 2D {[Zn(HL1)2] · 2H2O}n and a mononuclear [Zn(HL2)2(H2O)4] complexes, showing the influence of the susbtituent on the amino acids fragment as well as the versatility of this class of compounds when acting as ligands.  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