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.
A novel azo dye ligand, namely 1‐[(5‐mercapto‐1H‐1,2,4‐triazole‐3‐yl)diazenyl]naphthalen‐2‐ol (HL), was synthesized. Mn2+, Co2+, Ni2+, Cu2+ and UO22+ complexes were also prepared by the treatment of HL with Mn(CH3COO)2?4H2O, Co(CH3COO)2?4H2O, Ni(CH3COO)2?4H2O, Cu(CH3COO)2?H2O, CuCl2?2H2O, Cu(NO3)2?6H2O and UO2(NO3)2?6H2O. The structures of these metal chelates were confirmed using elemental, spectral, magnetic moment, molar conductance and thermal analyses. The analytical data confirmed the formation of the chelates in 1:1 (metal‐to‐ligand) ratio having the formula [ML(H2O)X]Y?H2O, where M is Mn2+, Co2+, Ni2+, Cu2+ or UO22+; X is Cl?, NO3? or CH3COO?; and Y is H2O. The azo compound acts in a monobasic bidentate manner via the nitrogen and oxygen atoms of azo and hydroxyl groups, respectively. All complexes were found to have tetrahedral structures, except the UO22+ complex that showed octahedral geometry. The mode of interaction between the synthesized complexes and calf thymus DNA was explored by the aid of absorption spectroscopy and viscosity measurements. The azo dye and its chelates were evaluated against the growth of various bacterial and fungal strains (Escherichia coli, Staphylococcus aureus, Aspergillus flavus and Candida albicans) with insight gained into the effect of type of metal centre, type of coordinated anion and position of the metal in the periodic table on the activity of the complexes. The geometric structure of the complexes was optimized using molecular modelling. The in vitro cytotoxicity of the synthesized compounds was tested against HEPG2 cell line.  相似文献   

2.
Bis(N, N′‐dialkyldithiocarbamato)antimony(III) alkylenedithiophosphates of the type [R2NCS2]2 SbS(S)POGO [where NR2 = N(CH3)2, N(C2H5)2 and N(CH2)4; G = ? CH2? C(C2H5)2? CH2? , ? CH2? C(CH3)2? CH2? , ? CH(CH3)? CH(CH3)? and ? C(CH3)2? C(CH3)2? ] were synthesized and characterized by physico‐chemical, spectral [UV, IR and NMR (1H, 13C and 31P)] and thermal (TG, DTA and DSC) analysis. The TG decomposition analysis step of the complex indicated the formation of Sb2S3 as the final product. The first endothermic peak in DSC indicated the melting point of the complexes. These complexes were screened for their antimicrobial activities using the disk diffusion method. All the complexes showed good activity as antibacterial and antifungal agents on some selected bacterial and fungal strains, which increased on increasing the concentration. Chloroamphenicol and terbinafin were used as standards for comparison. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

3.
The hydrogen-bonded complexes between CH3NH2 and (CH3)2NH with HCl have been studied by the ab initio molecular orbital method using the 4–31G basis set. Calculations show that the proton potential curve has a single minimum near to the nitrogen atom in both complexes. This means that the proton has been transferred from HCl to the amine. ΔE and the dipole moment of the complexes studied are as follows: ?18.2 kcal mol?1, 10.3 D for methylamine ·HCl, and ?21.7 kcal mol?1 11.1 D for the corresponding dimethylamine complex. Other properties of the hydrogen-bonded ion pairs are discussed.  相似文献   

4.
Ligands of the type H3-nN(CH2CH2OCH2CH2OCH2CH2)nNH3?n with n values from 1 to 3 have been investigated. The stability constants and the heat evolved by formation of the 1:1 complexes of Na+, K+, Rb+, Tl+, Ca2+, Sr2+, Ba2+, Cd2+, Ag+, Hg2+ and Pb2+ have been determined. The complex formation is discussed in terms of ΔH and ΔS taking into consideration the radii of the cations. In contrast with the normal trend, for the A cations, complex formation is exothermic and almost exclusively favoured by the reaction enthalpy.  相似文献   

5.
The syntheses of a number of N-substituted α-amidinium thiolsulfates, CH2(S2O3?)-C(=NH2+)NH(CH2)nR are described, where n was varied from 1 to 3 and R represents such heteroaryl groups as 2-furyl, 2-thienyl, 3-indolyl and 2-, 3- and 4-pyridyl. The preparation of S-(2-imidazolinemethyl)thiolsulfuric acid, as an example of an N, N'-disubstituted α-amidinium thiolsulfate, is also reported.  相似文献   

6.
Two lanthanide complexes with 2-fluorobenzoate (2-FBA) and 1,10-phenanthroline (phen) were synthesized and characterized by X-ray diffraction. The structure of each complex contains two non-equivalent binuclear molecules, [Ln(2-FBA)3?·?phen?·?CH3CH2OH]2 and [Ln(2-FBA)3?·?phen]2 (Ln?=?Eu (1) and Sm (2)). In [Ln(2-FBA)3?·?phen?·?CH3CH2OH]2, the Ln3+ is surrounded by eight atoms, five O atoms from five 2-FBA groups, one O atom from ethanol and two N atoms from phen ligand; 2-FBA groups coordinate Ln3+ with monodentate and bridging coordination modes. The polyhedron around Ln3+ is a distorted square-antiprism. In [Ln(2-FBA)3?·?phen]2, the Ln3+ is coordinated by nine atoms, seven O atoms from five 2-FBA groups and two N atoms of phen ligand; 2-FBA groups coordinate Ln3+ ion with chelating, bridging and chelating-bridging three coordination modes. The polyhedron around Ln3+ ion is a distorted, monocapped square-antiprism. The europium complex exhibits strong red fluorescence from 5D0?→?7F j ( j?=?1–4) transition emission of Eu3+.  相似文献   

7.
An account is given of the development of the proposal that ion–neutral complexes are involved in the unimolecular reactions of onium ions (R1R2C?Z+R3; Z = O, S, NR4; R1, R2, R3, R4 = H, CnH2n + 1), with particular emphasis on the informative C4H9O+ oxonium ion system (Z = O; R1, R2 = H; R3 = C3H7). Current ideas on the role of ion-neutral complexes in cation rearrangements, hydrogen transfer processes and more complex isomerizations are illustrated by considering the behaviour of isomeric CH3CH2CH2X+ and (CH3)2CHX+ species [X = CH2O, CH3CHO, H2O, CH3OH, NH3, NH2CH3, NH(CH3)2, CH2?NH, CH2?NCH3, CO, CH3˙, Br˙ and I˙]. Attention is focused on the importance of four energetic factors (the stabilization energy of the ion–neutral complex, the energy released by rearrangement of the cationic component, the enthalpy change for proton transfer between the partners of the ion neutral complex and the ergicity of recombination of the components) which influence the reactivity of the complexes. The nature and extent of the chemistry involving ion-neutral complexes depend on the relative magnitudes of these parameters. Thus, when the magnitude of the stabilization energy exceeds the energy released by cation rearrangement, the ergicity of proton transfer is small, and recombination of the components in a new way is energetically favourable, extensive complex-mediated isomerizations tend to occur. Loss of H2O from metastable CH2?O+C3H7 ions is an example of such a reaction. Conversely, if the stabilization energy is small compared with the magnitude of the energy released by eation rearrangement, the opportunities for complex-mediated processes to become manifest are decreased, especially if proton transfer is endoergic. Thus, CH3CH2CH2CO+ expels CO, with an increased kinetic energy release, after rate-limiting isomerization of CH3CH2CH2+? CO to (CH3)2CH+? CO has taken place. When proton transfer between the components of the complex is strongly exoergic, fragmentation corresponding to single hydrogen transfer occurs readily. The proton-transfer step is often preceded by cation rearrangement for CH3CH2CH2X+ species. In such circumstances, the involvement of ion–neutral complexes can be detected by the observation of unusual site selectivity in the hydrogen-transfer step. Thus, C3H6 loss from CH2?N+(R1)CH2CH2CH3 (R1 = H, CH3, C3H7) immonium ions is found by 2H-labelling experiments to proceed via preferential α-and γ-hydrogen transfer; this finding is explained if the incipient +CH2CH2CH3 ion isomerizes to CH3CH+CH3 prior to proton abstraction. In contrast, the isomeric CH2?N+(R1)CH(CH3)2 species undergo specific β-hydrogen transfer because the developing CH3CH+CH3 cation is stable with respect to rearrangements involving a 1,2-H shift.  相似文献   

8.
The complex formation of silver(I) has been studied with the anions of simple mercaptans RSH which have been rendered soluble by replacing some H in the substituent R by OH. All equilibria constants refer to a solvent of ionic strength μ = 0,1 and 20°C. Monothioglycol HO? CH2? CH2? SH (pK = 9.48) forms an amorphous insoluble mercaptide {AgSR} (s), ionic product [Ag+] [SR?] = 10?19.7. The solution in equilibrium with the solid contains the molecule AgSR at a constant concentration of 10?6.7 M which furnishes the formation constant of the 1:1-complex: K1 = 1013. 0. The solid is soluble in excess of mercaptide (AgSR+SR? → Ag(SR)2?: K2 = 104. 8) as well as in an excess of silver ion (AgSR + Ag+ → Ag2SR+K ≈? 106). With the bulky monothiopentaerythrite (HO? CH2? )3C? CH2? SH (pK = 9.89) no precipitation occurs with silver when the mercaptan concentration is below 10?3. 2M. A single polynuclear Ag10(SR)9+10.9 = 10175) is formed in acidic solutions which breaks up with the formation of Ag2SR+2.1 = 1019.0) when an excess of silver ion is added. Below the mononuclear wall ([RS]total < 10?6) Ag2SR+ is formed via the mononuclear AgSR (K1 = 1013). At higher mercaptan concentrations ([RS]tot > 10?3.2) an amorphous precipitate is formed which has almost the same solubility product as silver thioglycolate ([Ag+] [SR?] = 10?19.1). Apparently silver(I) forms with mercaptans always the complexes Ag2SR+, AgSR and Ag(SR)2?. Above the mononuclear wall, these species condense to chain-like polynuclears which are cations Ag(SRAg)n+ in presence of an excess of Ag+, and anions SR (AgSR)n? when the concentration [RS?] is larger than [Ag+]. Usually n becomes rapidly very large as soon as the condensation starts (n → ∞: precipitate). The decanuclear Ag(SRAg)9+ formed with thiopentaerythrit is somewhat more stable than the shorter chains (n < 9) and larger chains (n > 9), because it can tangle up to a ball by coordination of bridging mercapto-sulfur to the terminal silver ions (figure 12, page 2179). This ball seems to be further stabilized by hydrogen bonds between the many alcoholic OH groups of the substituent R = (HO? CH2)3C? CH2? . The stability of the bonds Ag? S, however, is little influenced by the substituent R which carries the mercaptide sulfure.  相似文献   

9.
The kinetics of the base catalysed racemization of [Co(EN3A)H2O]
  • 1 Abbreviations: EN3A3?=(?OOCCH2)2N(CH2)2NHCH2COO?; ME3A3?=(?OOCCH2)2N(CH2)2 N(CH3)CH2COO?; EDDA2?=?OOCCH2NH(CH2)2NHCH2COO?; EDTA4?=(?OOCCH2)2N(CH2)2N(CH2COO?)2;TNTA4?=(?OOCCH2)2N(CH2)3N(CH3COO?)2; HETA3?=(?OOCCH2)2N(CH2)2N(CH2COO?)CH2CH2OH; en=H2N(CH2)2NH2; Meen=H2N(CH2)2NHCH3; sar?=?OOCCH2NHCH3.
  • were studied polarimetrically in aqueous buffer solution. The reaction rate is first order in OH? and in complex, in weakly acidic medium. Activation parameters are ΔH≠=22 kcal · mol?1, ΔS≠=26 cal · K?1. The results are discussed in terms of an SN1CB mechanism involving exchange of the ligand water molecule. The N-methylated analogue [Co(ME3A)H2O] does not racemize in the pH-range investigated. Loss of optical activity occurs at a rate which is about 1,000 times slower than the racemization of [Co(EN3A)H2O](60°) and coincides with the decomposition of the complex.  相似文献   

    10.
    The stability of the complexes of AgI with sulphur-containing aminopyridines of general formula Py? (CH2)n?1? S? (CH2)m? NH2 (where: n ? 1, m = 1,2; 1,3; 2,2; 2,3) have been determined at 25°C and for an ionic strength of 0.5 M (K)NO3 by a combined pH–pM metric method. The study of complex formation and the determination of the approximative values for the constants proceeded by graphical means. The constants were refined by a least squares computer method. In acid medium (pH < 3) the protonated species AgLH23+ and AgL2H45+ are formed and coordination occurs through the thioether group. At higher pH values (pH > 3) additional chelation occurs through the pyridine-nitrogen donor in the complexes AgLH2+, AgL2H34+, and AgL2H23+. The former complex extensively transforms into a dimer Ag2L2H24+, which may also arise on the addition of a protonated ligand to the dinuclear Ag2LH3+ species. At still higher PH values (pH > 6) also the aminogroup participates in chelation in the species Ag2L2H3+, Ag2L22+ and Ag2L2+. With an excess of ligand however a white precipitate is formed.  相似文献   

    11.
    [o-, m- and p-Bis(alkylamino or alkyloxy)benzene] (cyclopentadienyl)iron(1+) hexafluorophosphates {2 and 4; [(CnH2n+1X)2C6H4](C5H5)Fe+PF6?, X?NH or O} were prepared by aromatic nucleophilic substitution of the (dichlorobenzene)iron cationic complexes (1). Critical micelle concentrations of the complex chlorides (3), prepared from 2 (n=8, X?NH) by anion exchange and soluble in water, gave much smaller values than those of bis(long-chain alkyl)dimethylammonium surfactants. Furthermore, the substitution positions scarcely affected their surface activites. However, the surface pressure-molecular area isotherm of the hexafluorophosphates (2 and 4, n=18, X?NH; insoluble in water) were severely transformed by change in the substitution position of the long-chain alkyl groups on the benzene ligand in the iron cationic complexes: the o-substituted complex gave a molecularly assembled film by the Langmuir-Blodgett (LB) method, but the P-substituted one did not.  相似文献   

    12.
    The complex formation of PdII with tris[2-(dimethylamino)ethyl]amine (N(CH2CH2N(CH3)2)3, Me6tren) was investigated at 25° and ionic strength I = 1, using UV/VIS, potentiometric, and NMR measurements. Chloride, bromide, and thiocyanate were used as auxiliary ligands. The stability constant of [Pd(Me6tren)]2+ in various ionic media was obtained: log β([Pd(Me6tren)] = 30.5 (I = 1(NaCl)) and 30.8 (I = 1(NaBr)), as well as the formation constants of the mixed complexes [Pd(HMe6tren)X]2+ from [Pd(HMe6tren)(H2O)]3+:log K = 3.50 = Cl?) and 3.64 (X? = Br?) and [Pd(Me6tren)X]+ from [Pd(Me6tren)(H2O)]2+: log K = 2.6 (X? = Cl?), 2.8(Br?) and 5.57 (SCN?) at I = 1 (NaClO3). The above data, as well as the NMR measurements do not provide any evidence for the penta-coordination of PdII, proposed in some papers.  相似文献   

    13.
    A potentiometric study of complex formation between gallium and L-serine, L-tyrosine, L-methionine shows the existence of the complexes Ga(L0)3+ and Ga(L?)2+. In the case of L-cysteine, the complex Ga(L2?)+ exists also. L0, L?, L2? designate respectively the zwitterion and the mono- and divalent anions of amino acid. The stability constants of the complexes have been determined.  相似文献   

    14.
    Chemistry of Polyfunctional Molecules. 119 [1]. Tetracarbonyl-dicobalt-tetrahedrane Complexes with the Ligands Bis(diphenylphosphanyl)-amine, 2-Butin-1,4-diol, and tert.-Butylphosphaacetylene — Crystal Structure of the Phosphaalkyne Derivative Co2(μ-CO)2(CO)4(μ-Ph2P? NH? PPh2P,P′) · 1/2C6H5CH3 ( 4 · 1/2C6H5CH3) reacts with 2-butine-1,4-diol, HOCH2? C?C? CH2OH ( 5 ), to the dark-red tetrahedrane complex Co2(CO)4(μ-η22-HOCH2? C?C? CH2OH? C2, C3) · (μ-Ph2P? NH? PPh2? P,P′) · THF (6 · THF). With t-butyl-phosphaacetylene, tBu? C?P ( 7 ), 4 · THF forms Co2(CO)4(μ-η22-tBu? C?P)(μ-Ph2P? NH? PPh2? P,P′) ( 8 ), which also belongs to the tetrahydrane type. The compounds were characterized by their mass, IR, 31P{1H} NMR, 13C{1H} NMR, and1H NMR spectra. Crystals suitable for X-ray structure analyses have been obtained for 8 from dioxane. The dark red blocks crystallize in the monoclinic P21/c space group with the lattice constants a = 1404,1(5), b = 1330,0(7), c = 2578,8(10)pm; β = 90,82(3)°.  相似文献   

    15.
    Two new Ni(II) complexes of 2,6-bis[1-(2,6-diethylphenylimino)ethyl]pyridine (L1), 2,6-bis[1-(4-methylphenylimino)ethyl]pyridine (L2 ) have been synthesized and structurally characterized. Complex Ni(L1)Cl2?·?CH3CN (1), exhibits a distorted trigonal bipyramidal geometry, whereas complex Ni(L1)(CH3CN)Cl2 (2), is six-coordinate with a geometry that can best be described as distorted octahedral. The catalytic activities of complexes 1, 2, Ni{2,6-bis[1-(2,6-diisopropyl-phenylimino)ethyl]pyridine} Cl2?·?CH3CN (3), and Ni{2,6-bis[1-(2,6-dimethylphenylimino) ethyl]pyridine}Cl2?·?CH3CN (4), for ethylene polymerization were studied under activation with MAO.  相似文献   

    16.
    Abstract

    Cobalt(III) complexes of the type [Co(en)2(chel)]X.nH2O where en = ethylenediamine, chel = phthalato = C6H4CO2)2? 2, maleato = (O2CCH = CHCO2)2?, succinato = (O2CCH2CH2CO2)2?, homophthalato = (O2CC6H4(CH2)CO2)2?, citraconato = (O2CC(CH3) = CHCO2)2?, itaconato = (CH2 = C(CO2)CH2CO2)2?, X = NO? 3, Br?, (O2CC6H4CO2H)?, (O2CHC = CHCO2H)?, (O2C(CH2)2CO2H)?, (O2CC6H4(CH2)CO2H)?, (O2CHC = C(CH2)-CO2H)?, and (O2C-CH2?C(= CH2)-CO2H)?, [Co(en)2(malonato)]X.2H2O (where malonato = (O2CCH2CO2)2?, X = Cl?, Br?, and NO? 3) and [Co(en)2CO3]Cl.2H2O have been investigated for their bacterial activity against Escherichia coli B growing on EMB agar and in minimal glucose media both in lag and log phases. Among the most active are where chel = phthalato and homophthalato. The effects are distinct from those known for compounds of Pt, e.g., cis?[Pt(NH3)2Cl2] and rhodium, e.g., trans?[Rh(C5H5N)4,Cl2].6H2O. Antagonisms are reported.  相似文献   

    17.
    The syntheses of the linear tetraamines H2N? (CH2)2? NH? (CH2)2? NH? (CH2)2? NH2 (n = 4 and 5) are described. The protonation of the homologous tetraamines for n = 2, 3, 4, 5, 6 and 8, as well as of N-methylethylenediamine, was investigated using potentiometric and calorimetric measurements. The results obtained are discussed taking into consideration the substituent effect on the basicity of the aminic N-atoms.  相似文献   

    18.
    Intramolecular H‐atom transfer in model peptide‐type radicals was investigated with high‐level quantum‐chemistry calculations. Examination of 1,2‐, 1,3‐, 1,5‐, and 1,6[C ? N]‐H shifts, 1,4‐ and 1,7[C ? C]‐H shifts, and 1,4[N ? N]‐H shifts (Scheme 1), was carried out with a number of theoretical methods. In the first place, the performance of UB3‐LYP (with the 6‐31G(d), 6‐31G(2df,p), and 6‐311+G(d,p) basis sets) and UMP2 (with the 6‐31G(d) basis set) was assessed for the determination of radical geometries. We found that there is only a small basis‐set dependence for the UB3‐LYP structures, and geometries optimized with UB3‐LYP/6‐31G(d) are generally sufficient for use in conjunction with high‐level composite methods in the determination of improved H‐transfer thermochemistry. Methods assessed in this regard include the high‐level composite methods, G3(MP2)‐RAD, CBS‐QB3, and G3//B3‐LYP, as well as the density‐functional methods B3‐LYP, MPWB1K, and BMK in association with the 6‐31+G(d,p) and 6‐311++G(3df,3pd) basis sets. The high‐level methods give results that are close to one another, while the recently developed functionals MPWB1K and BMK provide cost‐effective alternatives. For the systems considered, the transformation of an N‐centered radical to a C‐centered radical is always exothermic (by 25 kJ ? mol?1 or more), and this can lead to quite modest barrier heights of less than 60 kJ ? mol?1 (specifically for 1,5[C ? N]‐H and 1,6[C ? N]‐H shifts). H‐Migration barriers appear to decrease as the ring size in the transition structure (TS) increases, with a lowering of the barrier being found, for example when moving from a rearrangement proceeding via a four‐membered‐ring TS (e.g., the 1,3[C ? N]‐H shift, CH3? C(O)? NH..CH2? C(O)? NH2) to a rearrangement proceeding via a six‐membered‐ring TS (e.g., the 1,5[C ? N]‐H shift, .NH? CH2? C(O)? NH? CH3 → NH2? CH2? C(O)? NH? CH2.).  相似文献   

    19.
    Ab initio SCF and CI calculations on the cationic and neutral complexes of formaldehyde and lithium are reported. For the cationic complex CH2O/Li+, the stabilization energy of 41.7 kcal/mol obtained from the SCF calculation increases to 51.6 kcal/mol if a configuration interaction is introduced. For the neutral complex CH2O?/Li+, the C2v-conformer of the 2A1-state with the equilibrium bond distances of d(C? O) = 1.23 Å and d (O? Li) = 1.90 Å is calculated to be more stable than the 2B1-state with d (C? O) = 1.34 Å, and d (O? Li) = 1.65 Å. Charge transfer and polarization effects upon complex formation are discussed.  相似文献   

    20.
    Reactions of the platinum(IV) nitrile complexes [PtCl4(RCN)2] (R = Me, CH2Ph, Ph) with 1,2- and 1,4-PhS(=NH)C6H4SPh in CH2Cl2 afforded addition products of sulfimides and coordinated nitriles, viz., the [PtCl4{NH=C(R)N=S(Ph)(C6H4SPh)}2] complexes. The latter were isolated in 75—90% yields and characterized by elemental analysis, positive-ion FAB mass spectrometry, IR spectroscopy, and 1H and 13C1H NMR spectroscopy. The temperature dependence of the 1H NMR spectra of the model [PtCl4{NH=C(R)N=SPh2}2] complexes (R = Me, Et) in CD2Cl2 studied in a temperature range from +40 to -70 °C demonstrated that EZ isomerization of the ligands is a dynamic process in a range from +40 to -10 °C. The activation free energy of this process was calculated.Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1618–1622, August, 2004.  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