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.
Parameters of the solvation equilibria \({\left[ {Fe{{\left( {{H_2}O} \right)}_6}} \right]^{3 + }} + nDMSO \rightleftarrows {\left[ {Fe{{\left( {{H_2}O} \right)}_{6 - n}}{{\left( {DMSO} \right)}_n}} \right]^{3 + }} + n{H_2}O\) have been determined in aqueous-dimethyl sulfoxide solutions (0–90 vol% DMSO) by means of spectrophotometry and mathematical modeling of equilibria. Iron(III) is not involved in the complex formation with derivatives of sym-triazine: 2,4-diamino-6-(carbamoylmethylsulfinylmethyl)-1,3,5-triazine and 2,4-diamino-6-(acetohydrazidomethylsulfinylmethyl)-1,3,5-triazine in aqueous DMSO medium (40 vol % DMSO). Bis(hydrazinocarbonylmethyl) sulfoxide forms two complexes with iron(III), with 1: 1 and 1: 2 compositions; in contrast to the Cu(II) and Ni(II) complexes, in the iron complexes the ligand exists in the amide form. The most probable structures of the complexes have been revealed by molecular mechanics simulation and (in selected cases) using the DFT/B3LYP/6-31++G(d,p) density functional theory method.  相似文献   

2.
The Hg2+aq- and HgCl+aq-assisted aquations of [PtCl4]2- (1), [PtCl3(H2O)]- (2), cis-[PtCl2(H2O)2] (3), trans-[PtCl2(H2O)2] (4), [PtCl(H2O)3]+ (5), [PtCl3Me2SO]- (6), trans-[PtCl2(H2O)Me2SO] (7), cis-[PtCl(H2O)2Me2SO]+ (8), trans-[PtCl(H2O)2M32SO]+ (9), trans-[PtCl2(NH3)2] (10), and cis-[PtCl2(NH3)2] (11) have been studied at 25.0 degrees C in a 1.00 M HClO4 medium buffered with chloride, using stopped-flow and conventional spectrophotometry. Saturation kinetics and instantaneous, large UV/vis spectral changes on mixing solutions of platinum complex and mercury are ascribed to formation of transient adducts between Hg2+ and several of the platinum complexes. Depending on the limiting rate constants, these adducts are observed for a few milliseconds to a few minutes. Thermodynamic and kinetics data together with the UV/vis spectral changes and DFT calculations indicate that their structures are characterized by axial coordination of Hg to Pt with remarkably short metal-metal bonds. Stability constants for the Hg2+ adducts with complexes 1-6, 10, and 11 are (2.1 +/- 0.4) x 10(4), (8 +/- 1) x 10(2), 94 +/- 6, 13 +/- 2, 5 +/- 2, 60 +/- 6, 387 +/- 2, and 190 +/- 3 M-1, respectively, whereas adduct formation with the sulfoxide complexes 7-9 is too weak to be observed. For analogous platinum(II) complexes, the stabilities of the Pt-Hg adducts increase in the order sulfoxide < aqua < ammine complex, reflecting a sensitivity to the pi-acid strength of the Pt ligands. Rate constants for chloride transfer from HgCl+ and HgCl2 to complexes 1-11 have been determined. Second-order rate constants for activation by Hg2+ are practically the same as those for activation by HgCl+ for each of the platinum complexes studied, yet resolved contributions for Hg2+ and HgCl+ reveal that the latter does not form dinuclear adducts of any significant stability. The overall experimental evidence is consistent with a mechanism in which the accumulated Pt(II)-Hg2+ adducts are not reactive intermediates along the reaction coordinate. The aquation process occurs via weaker Pt-Cl-Hg or Pt-Cl-HgCl bridged complexes.  相似文献   

3.
In this paper the synthesis, characterization and photoluminescent behavior of the [RE(DBM)3L2] complexes (RE=Gd and Eu) with a variety of sulfoxide ligands; L=benzyl sulfoxide (DBSO), methyl sulfoxide (DMSO), phenyl sulfoxide (DPSO) and p-tolyl sulfoxide (PTSO) have been investigated in solid state. The emission spectra of the Eu(3+)-beta-diketonate complexes show characteristics narrow bands arising from the 5D0-->7F(J) (J=0-4) transitions, which are split according to the selection rule for C(n), C(nv) or C(s) site symmetries. The experimental Judd-Ofelt intensity parameters (Omega2 and Omega4), radiative (A(rad)) and non-radiative (A(nrad)) decay rates, and R02 for the europium complexes have been determined and compared. The highest value of Omega2 (61.9x10(-20)cm2) was obtained to the complex with PTSO ligand, indicating that Eu3+ ion is in the highly polarizable chemical environment. The higher values of the experimental quantum yield (q) and emission quantum efficiency of the emitter 5D0 level (eta) for the Eu-complexes with DMSO, DBSO and PTSO sulfoxides suggest that these complexes are promising Light Conversion Molecular Devices (LCMDs). The lower value of quantum yield (q=1%), for the hydrated complex [Eu(DBM)3H2O], indicates that the luminescence quenching occurs via multiphonon relaxation by coupling with the OH-oscillators from water molecule coordinated to rare earth ion. The pure red emission of the Eu-complexes has been confirmed by (x, y) color coordinates.  相似文献   

4.
The potentials of mean force (PMFs) were determined for systems forming cationic and anionic homocomplexes composed of acetic acid, phenol, isopropylamine, n-butylamine, imidazole, and 4(5)-methylimidazole, and their conjugated bases or acids, respectively, in three solvents with different polarity and hydrogen-bonding propensity: acetonitrile (AN), dimethyl sulfoxide (DMSO), and water (H(2)O). For each pair and each solvent a series of umbrella-sampling molecular dynamics simulations with the AMBER force field, explicit solvent, and counterions added to maintain a zero net charge of a system were carried out and the PMF was calculated by using the Weighted Histogram Analysis Method (WHAM). Subsequently, homoconjugation-equilibrium constants were calculated by numerical integration of the respective PMF profiles. In all cases but imidazole stable homocomplexes were found to form in solution, which was manifested as the presence of contact minima corresponding to hydrogen-bonded species in the PMF curves. The calculated homoconjugation constants were found to be greater for complexes with the OHO bridge (acetic acid and phenol) than with the NHN bridge and they were found to decrease with increasing polarity and hydrogen-bonding propensity of the solvent (i.e., in the series AN > DMSO > H(2)O), both facts being in agreement with the available experimental data. It was also found that interactions with counterions are manifested as the broadening of the contact minimum or appearance of additional minima in the PMF profiles of the acetic acid-acetate, phenol/phenolate system in acetonitrile, and the 4(5)-methylimidazole/4(5)-methylimidzole cation conjugated base system in dimethyl sulfoxide.  相似文献   

5.
A new class of symmetric, end-off, N-methyl piperazine armed binucleating ligands 2,6-bis(4-methyl piperazin-1-yl-methyl)-4-acetyl phenol (HL1) and 2,6-bis[(4-methyl piperazin-1-yl-methyl)]-(4-methylcarboxy) phenol (HL2) were synthesized by the Mannich reaction. Their mononuclear and binuclear Cu(II), Ni(II) and Zn(II) complexes have been synthesized. These complexes were characterized by elemental analysis, infra-red and electronic spectral analysis. In the electronic spectra, the lower electron withdrawing nature of the C(O)CH3p-substituent (HL1) compared with the C(O)OCH3p-substituent (HL2) of the phenolic ring causes a red shift in the LMCT-charge transfer band. The mononuclear Cu(II) complexes 1 and 7 have a magnetic moment value close to the spin only value with four hyperfine EPR signals. The binuclear Cu(II) complexes 4 and 10 illustrate an antiferromagnetic interaction (μeff 1.56 and 1.55 BM) at 298 K with a broad EPR signal. A variable temperature magnetic moment study of the binuclear copper(II) complexes shows that the extent of antiferromagnetic coupling increases in the order: CHO [K. Shanmuga Bharathi, A. Kalilur Rahiman, K. Rajesh, S. Sreedaran, P.G. Aravindan, D. Velmurugan, V. Narayanan, Polyhedron 25 (2006) 2859] < C(O)CH3 < C(O)OCH3 (−2J values 134 [Shanmuga Bharathi et al., mentioned above], 149 and 158 cm−1, respectively). The mononuclear Ni(II) complexes 2 and 8 are square planar and diamagnetic. The six coordinated binuclear Ni(II) complexes 5 and 11 show a magnetic moment value of 2.96 and 2.95 BM, respectively. Electrochemical studies of the complexes reveal that all the mononuclear complexes show a single irreversible one-electron transfer reduction wave and the binuclear complexes show two irreversible one-electron transfer reduction waves in the cathodic region. There is an anodic shift in the reduction of the metal centres when the electron withdrawing nature of the p-substituent of the phenolic ring increases. The catecholase activity of the mono and binuclear copper(II) complexes, using pyrocatechol as a model substrate, and the hydrolysis of 4-nitrophenyl phosphate using the mono and binuclear copper(II), nickel(II) and zinc(II) complexes as catalysts showed that the binuclear complexes have higher rate constant values than those of the corresponding mononuclear complexes. A comparison of the spectral, electrochemical and magnetic behaviour of the complexes derived from the ligands is discussed on the basis of the substituent at the para position of the phenolic ring.  相似文献   

6.
Four water soluble azo dyes, 4-(isopropyl)-2-[(E)-(4-chlorophenyl)diazenyl]phenol (L 1), 4-(isopropyl)-2-[(E)-(2,4-dichlorophenyl)diazenyl]phenol (L2), 4-(sec-butyl)-2-[(E)-(4-chlorophenyl) diazenyl]phenol (L 3), 4-(sec-butyl)-2-[(E)-(2,4-dichlorophenyl)diazenyl]phenol (L 4), and their Cu(II) and Ni(II) complexes were synthesized and characterized using spectroscopic methods. Examination of their thermal stability revealed similar decomposition temperature of approximately 260–300°C and that they were more thermally stable than their metal complexes. Ni(II) complexes of ligands L2 and L4 were more stable than the other coordination compounds. Among the synthesized ligands, L2 and the complexes Cu(L3)2 and Ni(L4)2 showed both antimicrobial and antifungal activity. However, the other ligands and the complexes were poorly active against selected microorganisms.  相似文献   

7.
运用G94W量子化学程序包,在HF/6-31G基组水平上对酰胺(DMF,DMA,HCONH2,HCONHCH3andCH3CONH2)与苯酚形成的系列氢键复合物(看作超分子)进行从头计算研究。根据计算结果探讨复合物的稳定性、施体和受体间的电荷转移及几何参数变化等规律。结果表明苯酚与上述一系列酰胺都可形成稳定的氢键复合物,其稳定性次序为CH3CONH2~HCONHCH3>HCONH2>DMA>DMF。结果还表明形成氢键复合物的过程包含着电荷转移,电荷由供体酰胺转移到受体苯酚中,酰胺中C=O键长和苯酚中的O-H键长都明显有规律性地变长。计算结果与实验规律相符。  相似文献   

8.
Five oxovanadium(iv) complexes, which were divided into two groups, [V(IV)O(bhbb, nhbb)(H(2)O)(2)] (tridentate ligands: H(2)bhbb = 2-(5-bromo-2-hydroxylbenzylideneamino)benzoic acid, ; H(2)nhbb = 2-(5-nitro-2-hydroxylbenzylideneamino)benzoic acid, ) and [V(IV)O(cpmp, bpmp, npmp)(2)] (bidentate ligands: Hcpmp = 4-chloro-2-((phenylimino)methyl)phenol, ; Hbpmp = 4-bromo-2-((phenylimino)methyl)phenol, ; Hnpmp = 4-nitro-2-((phenylimino)methyl) phenol, ) have been prepared and characterized by elemental analysis, infrared, UV-visible and electrospray ionization mass spectrometry. The coordination in [V(IV)O(bhbb)(H(2)O)(2)] () was confirmed by X-ray crystal structure analysis. The oxidation state of V(iv) with d(1) configuration in was confirmed by EPR. The speciation of VO/H(2)bhbb in methanol-aqueous solution was investigated by potentiometric pH titrations. The result indicated that the main species were [V(IV)O(bhbb)(OH)](-) and [V(IV)O(bhbb)(OH)(2)](2-) at the pH range 7.0-7.4. The structure-activity relationship of the vanadium complexes in inhibiting protein tyrosine phosphatases (protein tyrosine phosphatase 1B, PTP1B; T-cell protein tyrosine phosphatase, TCPTP; megakaryocyte protein-tyrosine phosphatase, PTP-MEG2; Src homology phosphatase 1, SHP-1 and Src homology phosphatase 2, SHP-2) was investigated. The oxovanadium(iv) complexes were potent inhibitors of PTP1B, TCPTP, PTP-MEG2, SHP-1 and SHP-2, but exhibited different inhibitory abilities over different PTPs. Complexes and displayed better selectivity to PTP1B over the other four PTPs. Kinetic data showed that complex inhibited PTP1B, TCPTP and SHP-1 with a noncompetitive inhibition mode, but a classical competitive inhibition mode for PTP-MEG2 and SHP-2. The results demonstrated that both the structures of vanadium complexes and the conformations of PTPs influenced PTP inhibition activity. The proper modification of the organic ligand moieties may result in screening potent and selective vanadium-based PTP1B inhibitors.  相似文献   

9.
《印度化学会志》2021,98(4):100049
The new azo-imine ligands 2,4-di-tert-butyl-6-((2-((2-hydroxyphenyl)diazenyl) phenylimino)methyl)phenol, H2L1, 1a, and 2,4-di-tert-butyl-6-((2-((2-hydroxyphenyl) diazenyl)p-chlorophenylimino)phenol, H2L2, 1b, were prepared. Reaction of H2L1;1a, and H2L2;1b, with uranyl nitrate hexahydrate afforded the mononuclear complexes of compositions [U(O)2(L1)(H2O)]; 2a, and [U(O)2(L2)(H2O)]; 2b, complexes respectively. The newly synthesised ligands (1a and 1b) and the complexes (2a and 2b) were characterised unequivocally. The x-ray structure of 2a was determined. The tetradentate dianionic ligand (L1)2- coordinated the uranium ion equatorially with a water molecule in the same plane. Two axially coordinated oxo ligands completed the overall pentagonal bipyramid geometry around U(VI) ion. Structural pattern, electron transfer properties (oxidation near 1.32 ​V vs Ag/AgCl) and electronic transitions of [U(O)2(L1)(H2O)]; 2a, and [U(O)2(L2)(H2O)]; 2b have been rationalized by DFT calculations.  相似文献   

10.
The Schiff bases 2-bromo-6-[(3-cyclohexylaminopropylimino)methyl]phenol (HCMP) and 2-bromo-6-[(3-dimethylaminopropylimino)methyl]phenol (HDMP) derived from 3-bromosalicylaldehyde with N-cyclohexylpropane-1,3-diamine and N,N-dimethylpropane-1,3-diamine, respectively, and their nickel(II) and zinc(II) complexes [Ni(CMP)2] (I) and [ZnCl2(HDMP)] (II) have been prepared and characterized by elemental analyses, IR, and single crystal X-ray crystallographic determination. The crystal of I is monoclinic: space group P21/c, a = 12.0304(6), b = 13.1594(6), c = 10.2445(5) Å, β = 101.019(1)°, V = 1591.9(1) Å3, Z = 2. The crystal of II is monoclinic: space group C2/c, a = 22.286(5), b = 12.210(3), c = 14.513(3) Å, β = 124.118(3)°, V = 3269.5(13) Å3, Z = 8. The Schiff base HCMP coordinates to the Ni atom through the phenolate O, imine N, and amine N atoms, while the Schiff base HDMP coordinates to the Zn atom through the phenolate O and imine N atoms. The effect of these complexes on the antimicrobial activity against Staphylococcus aureus, Escherichia coli, and Candida albicans were studied.  相似文献   

11.
Half-sandwich titanium salicylbenzoxazole complexes CpTiLCl(2) 2a-2c [L = R-2-(benzo[d]xazol-2-yl)phenol (R = H (2a), R = 6-CH(3) (2b), R = 4-CH(3)-6-(t)Bu (2c)] and salicylbenzothiazole complexes CpTiLCl(2) 2d-2g [L = R-2-(benzo[d]thiazol-2-yl)phenol (R = H (2d), R = 6-CH(3) (2e), R = 6-(t)Bu (2f), R = 4-Cl (2g)] were synthesized by the reaction of CpTiCl(3) with the sodium salts of their corresponding precursors. Complexes 2a-2g were fully characterized by (1)H and (13)C NMR spectra and elemental analyses. The molecular structures of 2a and 2b were determined by single crystal X-ray diffraction methods. When activated by excess methylaluminoxane (MAO) these half-sandwich titanium complexes showed moderate to high activities for ethylene polymerization and produced high molecular weight polyethylenes. The half-sandwich titanium salicylbenzoxazole complexes (2a-2c) exhibited higher activities, of up to 1.23 × 10(6) g PE mol Ti(-1) h(-1) for the 2b/MAO system, than those of their analogues, half-sandwich titanium salicylbenzothiazole complexes (2d-2g).  相似文献   

12.
Synthesis of seven complexes containing oxazoline ([(L(1))(2)V=O] (4), [(L(1))(2)MoO(2)] (5), [(L(1))(2)UO(2)] (6); HL(1) (1) [HL(1) = 2-(4',4'-dimethyl-3'-4'-dihydroxazol-2'-yl)phenol]), chiral oxazoline ([(L(2))(2)UO(2)] (7); HL(2) (2) [HL(2) = (4'R)-2-(4'-ethyl-3'4'-dihyroxazol-2'-yl)phenol]), and oxazine ([(L(3))(2)V=O] (8), [(L(3))(2)Mn(CH(3)COO(-))] (9), [(L(3))(2)Co] (10); HL(3) (3) [HL(3) = 2-(5,6-dihydro-4H-1,3-oxazolinyl)phenol]) and their characterization by various techniques such as UV-vis, IR, and EPR spectroscopy, mass spectrometry, cyclic voltammetry, and elemental analysis are reported. The novel oxazine (3) and complexes 4, 5, 8 and 9 were also characterized by X-ray crystallography. Oxazine 3 crystallizes in the monoclinic system with the P2(1)/n space group, complexes 4 and 9 crystallize in the monoclinic system with the P2(1)/c space group, and complexes 5 and 8 crystallize in the orthorhombic system with the C222(1) space group and the P2(1)2(1)2(1) chiral space group, respectively. The representative synthetic procedure involves the reaction of metal acetate or acetylacetonate derivatives with corresponding ligand in ethanol. Addition of Mn(OAc)(2).4H(2)O to an ethanol solution of 3 gave the unexpected complex Mn(L(3))(2).(CH(3)COO(-)) (9) where the acetate group is coordinated with the metal center in a bidentate fashion. The catalytic activity of complexes 4-9 for oxidation of styrene with tert-butyl hydroperoxide was tested. In all cases, benzaldehyde formed exclusively as the oxidation product.  相似文献   

13.
《Polyhedron》1987,6(6):1313-1317
Nickel(II) perchlorate and nitrate complexes containing dimethyl, di-n-propyl, di-n-butyl, di-i-butyl and di-t-butyl sulfoxides have been synthesized and characterized by IR and electronic spectroscopies, magnetic-susceptibility and electrolytic-conductance measurements. In the complexes containing perchlorate, the metal: sulfoxide molar ratio is 1:6 and the perchlorate groups are ionic. In the nitrate compounds, the molar ratio decreases from 1:6 to 1:2 according to the increase in the steric bulk of the alkyl group from methyl to t-butyl. The nitrate group may either be non-coordinating or behave as a monodentate or bidentate ligand. All the complexes contain O-bonded sulfoxide molecules and are characterized as high-spin, with an octahedral or distorted octahedral geometry. The dialkyl sulfoxides studied in this work fall in the same position as dimethyl sulfoxide in the spectrochemical and nephelauxetic series. Electrolytic conductivities suggest that the compounds containing ionic nitrate exhibit sulfoxide—nitrate exchange in nitromethane solutions.  相似文献   

14.
The anilic acids, 2,5-dihydroxy-1,4-benzoquinone (1a), 2,5-dibromo-3,6-dihydroxy-1,4-benzoquinone (bromanilic acid; 1b), 2,5-dichloro-3,6-dihydroxy-1,4-benzoquinone (chloranilic acid; 1c), and 2,5-dicyano-3,6-dihydroxy-1,4-benzoquinone (cyananilic acid; 1d), were cocrystallized with rigid organic ligands containing two pyridine rings, 2,4-bipyridine (2a), 4,4'-bipyridine (2b), 1,2-bis(2-pyridyl)ethylene (3a), 1,2-bis(4-pyridyl)ethylene (3b), 2,2'-dipyridylacetylene (4a), 3,3'-dipyridylacetylene (4b), and 4,4'-dipyridylacetylene (4c). Fourteen complexes 5-18 were obtained as single crystals, and their crystal structures were successfully determined by X-ray analysis. All complexes except those with 2a are 1:1 and are composed of an infinite linear or zigzag tape structure, the formation of which is ascribed to intermolecular O-H...N, N(+)-H...O, or N(+)-H...O(-) hydrogen bonds or a combination of these between the anilic acids and the dipyridyl compounds. In the complexes 5 and 6, no infinite tape structure is observed although the molecular units connected by a similar hydrogen-bonding pattern are formed. For the 1:1 complexes, we have found two types of stacking arrangements, segregated stacks (7, 9, 12-15, 18) and alternated ones (8, 10, 11, 16, 17). In the complexes of 1c with the series of dipyridylacetylenes 4 (14, 15, 17), the neutral, dication, and monocaction states are formed depending on the nitrogen positions, which can be attributed to the different basicity of the pyridyl groups.  相似文献   

15.
《印度化学会志》2021,98(4):100048
The reaction of Ru(κ2C,O-RL)(PPh3)2(CO)Cl, 1 with excess sodium salt of pyridine-2-carboxylic acid (Napic) furnishes the complexes of the type Ru(κ1C-RL)(PPh3)2(CO) (pic), 2(R) with excellent yield (κ2C,O-RL is C6H2O-2-CHNHC6H4R(p)-3-Me-5, κ1C-RL is C6H2OH-2-CHNC6H4R(p)-3-Me-5 and R is Me, OMe, Cl). The chelation of pic is attended with the cleavage of Ru–O and Ru–Cl bonds and iminium–phenolato→imine–phenol prototropic shift. The 1 ​→ ​2 conversion is irreversible and the type 2 species are thermodynamically more stable than the acetate, nitrite and nitrate complexes of 1. The spectral (UV–vis, IR, 1H NMR) and electrochemical data of the complexes are reported. In dichloromethane solution the complexes display one quasi–reversible RuIII/RuII cyclic voltammetric response with E1/2 in the range 0.72–0.80 ​V vs. Ag/AgCl. The crystal and molecular structure of Ru(κ1C-MeOL)(PPh3)2(CO)(pic)∙CH3CN is reported which revealed distorted octahedral RuC2P2NO coordination sphere. The pairs (P, P), (C, O) and (C, N) define the three trans directions. The electronic structures of the complexes are also scrutinized by density functional theory (DFT) and time–dependent density functional theory (TD–DFT) calculations.  相似文献   

16.
《印度化学会志》2021,98(2):100006
The new cis-dioxomolybdenum (VI) complexes [MoO2(L2)(H2O)] (2) and [MoO2(L3)(H2O)] (3) containing the tridentate hydrazone-based ligands (H2L2 = N'-(3,5-di-tert-butyl-2-hydroxybenzylidene)-4-methylbenzohydrazide and H2L3 = N'-(2-hydroxybenzylidene)-2-(hydroxyimino)propanehydrazide) have been synthesized and characterized via IR, 1H and 13C NMR spectroscopy, mass spectrometry, and single crystal X-ray diffraction analysis. The catalytic activities of complexes 2 and 3, and the analogous known complex [MoO2(L1)(H2O)] (1) (H2L1 = N'-(2-hydroxybenzylidene)-4-methylbenzohydrazide) have been evaluated for various oxidation reactions, viz. oxygen atom transfer from dimethyl sulfoxide to triphenylphosphine, sulfoxidation of methyl-p-tolylsulfide or epoxidation of different alkenes using tert-butyl hydroperoxide as terminal oxidant. The catalytic activities were found to be comparable for all three complexes, but complexes 1 and 3 showed better catalytic performances than complex 2, which contains a more sterically demanding ligand than the other two complexes.  相似文献   

17.
Titanium porphyrin hydrazido complexes (TTP)Ti = NNR2 (TTP = meso-tetra-p-tolylporphyrinato dianion; R = Me (1), Ph (2)) were synthesized by treatment of (TTP)TiCl2 with 1,1-disubstituted hydrazines H2NNR2 (R = Me, Ph) in the presence of piperdine. The nucleophilic character of the hydrazido moiety was demonstrated in the reactions of complexes 1 and 2 with p-chlorobenzaldehyde, which yielded the titanium oxo complex (TTP)-Ti = O and the respective hydrazones. Protonation of complexes 1 and 2 with phenol or water produced the 1,1-disubstituted hydrazine along with (TTP)Ti(OPh)2 or (TTP)Ti = O, respectively. Similar reactivity of p-chlorobenzaldehyde and phenol with (TTP)Ti = NiPr, 3, was observed. The reaction of complex 3 with nitrosobenzene cleanly formed the azo compound iPrN = NPh and the terminal oxo product (TTP)Ti = O.  相似文献   

18.
Reaction of [Ru(arene)Cl(2)](2) (arene = benzene, toluene, p-cymene, hexamethylbenzene) with K(7)[PW(11)O(39)].14H(2)O provided two series of organometallic derivatives of heteropolytungstates: type-1 and type-2 complexes of general formulas [PW(11)O(39){Ru(arene)(H(2)O)}](5-) and [{PW(11)O(39){Ru(arene)}}(2){WO(2)}](8-), respectively. All compounds were characterized by infrared and multinuclear NMR ((1)H, (31)P, (183)W) spectroscopies. The crystal structures of Na(4)K(4)[{PW(11)O(39){Ru(benzene)}}(2){WO(2)}].6H(2)O (NaK-2a.6H(2)O), K(7)H[{PW(11)O(39){Ru(toluene)}}(2){WO(2)}].4H(2)O (K-2b.4H(2)O), and Cs(3)K(2)[PW(11)O(39){Ru(p-cymene)(H(2)O)}].4H(2)O (CsK-1c.4H(2)O) were obtained and revealed that the {Ru(arene)} fragment is supported on the oxometallic framework. Photochemical reactivity of [PW(11)O(39){Ru(arene)(H(2)O)}](5-) (arene = toluene, p-cymene) in the presence of various ligands L (L = H(2)O, dimethyl sulfoxide, tetramethylene sulfoxide, and diphenyl sulfoxide) was investigated, and led to the formation of [PW(11)O(39){Ru(L)}](5-), in which the ruthenium is incorporated into the lacunary [PW(11)O(39)](7-) anion.  相似文献   

19.
The complexes [(C5Me5)Ir(η6-arene)][BF4]2 (arene = toluene, toluene-d8, t-butylbenzene, methoxybenzene, chlorobenzene, o-xylene, p-xylene, tetralin and phenol) were prepared from the arene and reduced with NaBH4 to the η5-cyclohexadienyl complexes. Attack was exo at the arene and, with one exception, never at the substituent. Toluene showed no site preference but t-butylbenzene was attacked preferentially para, and chlorobenzene, ortho. Methoxybenzene was attacked ipso as well as ortho, meta (predominant), and para, and phenol gave only the meta-isomer. p-Xylene gave one isomer and o-xylene and tetralin gave two. Further reduction occurred on reaction with stronger hydride reducers (e.g., sodium bis(methoxyethoxy)dihydroaluminate) to give mixtures of 1- and 2-substituted cyclohexa-1,3-diene complexes (t-Bu, 2- ( > 95%); Me, 1- (25%), 2- (75%); Cl, 1- ( > 95%); and OMe, 1- (33%), 2- (67%)). The p-xylene complex gave a mixture of the η4-1,4-dimethylcyclohexa-1,3- and 1,4-diene complexes. Reaction of the cyclohexadiene complexes with HCl gas gave the free substituted cyclohexenes and [(C5Me5Ir)2Cl4]. The product from t-butylbenzene was predominantly (92%) the 3-substituted cyclohexene; that isomer (65%) and the 1-isomer (34%) were formed from toluene and the 1- (34%) and the 4-isomer (58%) were formed from chlorobenzene. Phenol gave only cyclohexanone. Overall these reactions yield the cyclohexene from the substituted benzene by addition of two hydrides and two protons and the iridium can be recycled.  相似文献   

20.
A series of metal-ligand complexes were prepared by the reaction of various metal ions, namely, Cu(II), Mn(II), or Fe(II) with phenolic derivatives of [catechol, chlorogenic acid (CGA), n-propyl gallate (nPG), 3-hydroxy anthranilic acid, resveratrol, and rutin] and characterized by UV-vis spectroscopy. The metal/ligand complexing ratio and complexation constants have been determined. The complexes were probed for their reactivity toward various free radicals (e aq-, CO2*-, and O2*-). Pulse radiolysis studies showed that the one-electron reduction of metal/phenol complexes by CO2*- radicals was metal-centered, and this was confirmed by the formation of an initial adduct with CO2*- radicals. Rate constants for the scavenging of superoxide anions with metal complexes ranged between 10(7)-10(9) dm3 mol(-1) s(-1) and those for the reaction of e aq- with the metal complexes were in the range of (1-5) x 10(9) dm3 mol(-1) s(-1), depending on the pH of the solution. Cyclic and differential pulse voltammetric studies showed that the reduction potential of the complexes are found to range between -0.022 to 0.45 V vs normal hydrogen electrode.  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