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 471 毫秒
1.
The synthesis and characterization of new transition metal complexes of Mn(II), Co(II), Ni(II), Cu(II) and Zn(II) with 3‐(2‐hydroxynaph‐1‐ylazo)‐1,2,4‐triazole ( HL1 ) and 3‐(2‐hydroxy‐3‐carboxynaph‐1‐ylazo)‐1,2,4‐triazole ( HL2 ) have been carried out. Their structures were confirmed by elemental analyses, thermal analyses, spectral and magnetic data. The IR and 1H NMR spectra indicated that HL1 and HL2 coordinated to the metal ions as bidentate monobasic ligands via the hydroxyl O and azo N atoms. The UV‐Vis, ESR spectra and magnetic moment data revealed the formation of octahedral complexes [Mn L1 (AcO)(H2O)3] ( 1 ), [Co L1 (AcO)(H2O)3]·H2O ( 2 ), [Mn L2 (AcO)(H2O)3] ( 6 ) and [Co L2 (AcO)(H2O)3] ( 7 ), [Ni L1 (AcO)(H2O)] ( 3 ), [Zn L1 (AcO)(H2O)]·H2O ( 5 ), [Ni L2 (AcO)(H2O)] ( 8 ), [Zn L2 (AcO)(H2O)]·10H2O ( 10 ) have tetrahedral geometry, whereas [Cu L1 (AcO)(H2O)2] ( 4 ) and [Cu L2 (AcO)(H2O)2]·5H2O ( 9 ) have square pyramidal geometry.. The mass spectra of the complexes under EI‐con‐ ditions showed the highest peaks corresponding to their molecular weights, based on the atomic weights of 55Mn, 59Co, 58Ni, 63Cu and 64Zn isotopes; besides, other peaks containing other isotopes distribution of the metal. Kinetic and thermodynamic parameters of the thermal decomposition stages were computed from the thermal data using Coats‐Redfern method. HL2 and complexes 6 – 10 were found to have moderate antimicrobial activities against Staphylococcus aureus (gram positive), Escherichia coli (gram negative) and Salmonella sp bacteria, and antifungal activity against Fusarium oxysporum, Aspergillus niger and Candida albicans. Also, in most cases, metallation increased the activity compared with the free ligand.  相似文献   

2.
The influence of the potentially chelating imino group of imine‐functionalized Ir and Rh imidazole complexes on the formation of functionalized protic N‐heterocyclic carbene (pNHC) complexes by tautomerization/metallotropism sequences was investigated. Chloride abstraction in [Ir(cod)Cl{C3H3N2(DippN=CMe)‐κN3}] ( 1 a ) (cod=1,5‐cyclooctadiene, Dipp=2,6‐diisopropylphenyl) with TlPF6 gave [Ir(cod){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 3 a +[PF6]?). Plausible mechanisms for the tautomerization of complex 1 a to 3 a +[PF6]? involving C2?H bond activation either in 1 a or in [Ir(cod){C3H3N2(DippN=CMe)‐κN3}2]+[PF6]? ( 6 a +[PF6]?) were postulated. Addition of PR3 to complex 3 a +[PF6]? afforded the eighteen‐valence‐electron complexes [Ir(cod)(PR3){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 7 a +[PF6]? (R=Ph) and 7 b +[PF6]? (R=Me)). In contrast to Ir, chloride abstraction from [Rh(cod)Cl{C3H3N2(DippN=CMe)‐κN3}] ( 1 b ) at room temperature afforded [Rh(cod){C3H3N2(DippN=CMe)‐κN3}2]+[PF6]? ( 6 b +[PF6]?) and [Rh(cod){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 3 b +[PF6]?) (minor); the reaction yielded exclusively the latter product in toluene at 110 °C. Double metallation of the azole ring (at both the C2 and the N3 atom) was also achieved: [Ir2(cod)2Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 10 ) and the heterodinuclear complex [IrRh(cod)2Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 12 ) were fully characterized. The structures of complexes 1 b , 3 b +[PF6]?, 6 a +[PF6]?, 7 a +[PF6]?, [Ir(cod){C3HN2(DippN=CMe)(DippN=CH)(Me)‐κ2(N3,Nimine)}]+[PF6]? ( 9 +[PF6]?), 10? Et2O ? toluene, [Ir2(CO)4Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 11 ), and 12? 2 THF were determined by X‐ray diffraction.  相似文献   

3.
The structures of the title compounds, [Ho(C5H7O2)3(H2O)2]·H2O and [Ho(C5H7O2)3(H2O)2]·C5H8O2·2H2O, both show an eight‐coordinate holmium(III) ion in a square antiprismatic configuration. The packing of these structures consists of an infinite two‐dimensional network of hydrogen‐bonded mol­ecules. In both structures, the same hydrogen‐bonded chain of HoIII complexes is found.  相似文献   

4.
Palladacyclic compounds [Pd(C6H4(C6H5C?O)C?N? R)(N? N)] [X] (R = Et, iPr, 2,6‐iPr2C6H3; N? N = bpy = 2,2′‐bipyridine, or 1,4‐(o,o′‐dialkylaryl)‐1,4‐diazabuta‐1,3‐dienes; [X]? = [BF4]? or [PF6]?) were synthesized from the dimers [{Pd(C6H4(C6H5C?O)C?N? R)(μ‐Cl)}2] and N? N ligands. Their interionic structure in CD2Cl2 was determined by means of 19F,1H‐HOESY experiments and compared with that in the solid state derived from X‐ray single‐crystal studies. [Pd(C6H4(C6H5C?O)C?N? R)(N? N)] [X] complexes were found to copolymerize CO and p‐methylstyrene affording syndiotactic or isotactic copolymers when bpy or 1,4‐(o,o′‐dimethylaryl)‐1,4‐diazabuta‐1,3‐dienes were used, respectively. The reactions with CO and p‐methylstyrene of the bpy derivatives were investigated. Two intermediates derived from a single and a double insertion of CO into the Pd? C bonds were isolated and completely characterized in solution.  相似文献   

5.
In the title compound, C6H16N22+·2C2H4O5P?, the cations lie across centres of inversion; in the anions, two of the H‐atom sites have 0.50 occupancy. The anions are linked by short O—H?O hydrogen bonds [O?O 2.465 (3)–2.612 (3) Å and O—H?O 165–171°] into sheets of alternating R(12) and R(28) rings, both of which are centrosymmetric; the cations lie at the centres of the larger rings linked to the anion sheet by N—H?O hydrogen bonds [N?O 2.642 (2) Å and N—H?O 176°].  相似文献   

6.
7.
The interaction of bovine serum albumin (BSA) with Fe(III)?Ccitrate complexes ([FeIII(cit)(H2O)3]? and [FeIII(cit)2]5?) and the sonocatalytic damage of BSA under ultrasonic irradiation were studied. Additionally, the various factors influencing the sonocatalytic damage of BSA were also studied by means of UV?CVis and fluorescence spectra. The experimental results indicate that the probable fluorescence quenching mechanisms of BSA by Fe(III)?Ccitrate complexes ([FeIII(cit)(H2O)3]? and [FeIII(cit)2]5?) are both static quenching. Under certain conditions, the degree of damage to BSA is aggravated with increases of ultrasonic irradiation time, Fe(III)?Ccitrate complex concentration, pH value and ionic strength. Moreover, all of the results demonstrate that [FeIII(cit)2]5? displays higher sonocatalytic activity than [FeIII(cit)(H2O)3]? under the same experimental conditions during the damage process of BSA. Finally, the generation of ·O2 ? and ·OH during sonocatalytic processes was estimated using scavengers. Perhaps, the results will be significant for promoting sonodynamic treatment to treat tumors at the molecular level.  相似文献   

8.
The closo‐dodecaborate [B12H12]2? is degraded at room temperature by oxygen in an acidic aqueous solution in the course of several weeks to give B(OH)3. The degradation is induced by Ag2+ ions, generated from Ag+ by the action of H2S2O8. Oxa‐nido‐dodecaborate(1?) is an intermediate anion, that can be separated from the reaction mixture as [NBzlEt3][OB11H12] after five days in a yield of 18 %. The action of FeCl3 on the closo‐undecaborate [B11H11]2? in an aqueous solution gives either [B22H22]2? (by fusion) or nido‐B11H13(OH)? (by protonation and hydration), depending on the concentration of FeCl3. In acetonitrile, however, [B11H11]2? is transformed into [OB11H12]? by Fe3+ and oxygen. The radical anions [B12H12] ˙ ? and [B11H11] ˙ ? are assumed to be the primary products of the oxidation with the one‐electron oxidants Ag2+ and Fe3+, respectively. These radical anions are subsequently transformed into [OB11H12]? by oxygen. The crystal structure analysis shows that the structure of [OB11H12]? is derived from the hypothetical closo‐oxaborane OB12H12 by removal of the B3 vertex, leaving a non‐planar pentagonal aperture with a three‐coordinate O vertex, as predicted by NMR spectra and theory.  相似文献   

9.
The kinetics of oxidation of the chromium(III) complexes, [Cr(Ino)(H2O)5]3+ and [Cr(Ino)(Gly)(H2O)3]2+ (Ino?=?Inosine and Gly?=?Glycine) involving a ligands of biological significance by N-bromosuccinimide (NBS) in aqueous solution to chromium(VI) have been studied spectrophotometrically over the 25–45°C range. The reaction is first order with respect to both [NBS] and [Cr], and increases with pH over the 6.64–7.73 range in both cases. The experimental rate law is consistent with a mechanism in which the hydroxy complexes [Cr(Ino)(H2O)4(OH)]2+ and [Cr(Ino)(Gly)(H2O)2(OH)]+ are significantly more reactive than their conjugate acids. The value of the intramolecular electron transfer rate constant, k 1, for the oxidation of the [Cr(Ino)(H2O)5]3+ (6.90?×?10?4?s?1) is lower than the value of k 2 (9.66?×?10?2?s?1) for the oxidation of [Cr(Ino)(Gly)(H2O)2]2+ at 35°C and I?=?0.2?mol?dm?3. The activation parameters have been calculated. Electron transfer apparently takes place via an inner-sphere mechanism.  相似文献   

10.
Reactions of vanadyl sulfate (VOSO4?3H2O) with dimethylmalonic (H2Me2mal = = C3H6(CO2H)2), cyclobutane-1,1-dicarboxylic (H2cbdc = C4H6(CO2H)2), or butylmalonic acid (H2Bumal = C4H10(CO2H)2) and Li2CO3 in a ratio of 1: 2: 2 afforded novel coordination polymers of similar compositions ([Li2(VO)(Me2mal)2]n (1), [Li2(VO)(Me2mal)2(EtOH)-(H2O)]n (2), [Li4(VO)2(cbdc)4(H2O)7]n (3), and [Li2(VO)(Bumal)2(H2O)5.5]n (4)) but different structures. The crystal structures of the compounds containing the Me2mal2– and Bumal2– anions depend on the solvent nature.  相似文献   

11.
Solvothermal reaction of Zn(NO3)2 ? 4 H2O, 1,4‐bis[2‐(4‐pyridyl)ethenyl]benzene (bpeb) and 4,4′‐oxybisbenzoic acid (H2obc) in the presence of dimethylacetamide (DMA) as one of the solvents yielded a threefold interpenetrated pillared‐layer porous coordination polymer with pcu topology, [Zn2(bpeb)(obc)2] ? 5 H2O ( 1 ), which comprised an unusual isomer of the well‐known paddle‐wheel building block and the transtranstrans isomer of the bpeb pillar ligand. When dimethylformamide (DMF) was used instead of DMA, a supramolecular isomer [Zn2(bpeb)(obc)2] ? 2 DMF ? H2O ( 2 ), with the transcistrans isomer of the bpeb ligand with a slightly different variation of the paddle‐wheel repeating unit, was isolated. In MeOH, single crystals of 2 were transformed by solvent exchange in a single‐crystal‐to‐single‐crystal (SCSC) manner to yield [Zn2(bpeb)(obc)2] ? 2 H2O ( 3 ), which is a polymorph of 1 . SCSC conversion of 3 to 2 was achieved by soaking 3 in DMF. Compounds 1 and 2 as well as 2 and 3 are supramolecular isomers.  相似文献   

12.
The kinetics of stripping of Ni2+ from a Ni‐BTMPPA complex, dissolved in a kerosene solution of BTMPPA (H2A2, Cyanex 272), by acidic sulfate‐acetato solution, was studied using the single (falling) drop technique and flux (F) method of data treatment. The empirical flux equation at 303 K is Fb (kmol/m2s) = 10?4.35 [Ni2+] (1+10?3.42 [H+]?1)?1 ([H2A2](o)0.5+2.50 [H2A2](o))?1 (1+6[SO42?]) (1+3.20 [Ac?]). Activation energy (Ea), entropy change in activation (ΔS±), and enthalpy change in activation (ΔH±) were measured under different experimental conditions. Based on the empirical flux equation, Ea and ΔS±, the mechanism of Ni2+ stripping is provided. In a low [H+] region, the stripping reaction steps appear as [NiA+] → Ni2+ + A? and [Ni(HA2)2](int) → [NiHA2]+(int) + HA2(int)? in lower and higher concentration regions of free BTMPPA, respectively, provided [SO42?] and [Ac?] are kept low. However, at higher [H+] concentrations, the stripping is under diffusion control. With increasing [SO42?] and [Ac?], the enhancement of the rate is attributed to the attack of the Ni(II) complex by SO42? or HSO4? and Ac? to form NiSO4 or NiHSO4+ and NiAc+ complexes. Negative ΔS± values indicate that the rate‐determining stripping reaction steps occur via an substitution nucleophilic, bimolecular (SN2) mechanism.  相似文献   

13.
Examined in this study is the kinetics of a net 2e transfer between [Fe2(μ‐O)(phen)4(H2O)2]4+ ( 1 ) and its hydrolytic derivatives [Fe2(μ‐O)(phen)4(H2O)(OH)]3+ ( 2 ) and [Fe2(μ‐O)(phen)4(OH)2]2+ ( 3 ) with in aqueous media and in presence of excess 1,10‐phenanthroline (phen). The reaction is quantitative with a 1 : 1 stoichiometry between the oxidant and reductant to produce ferroin ([Fe(phen)3]2+) and . The order of reactivity of the oxidant species is 1 > 2 > 3 , in agreement with the progressive cationic charge reduction. The reactions appear to be inner‐sphere where the initial one‐electron proton‐coupled redox (1e, 1H+; electroprotic) seems to be rate‐determining.  相似文献   

14.
The 1:1 organic salt of the title compound, C7H6ClN2O+·C8H5Cl2O3? or [(2‐ABOX)(3,4‐D)], comprises the two constituent mol­ecules associated by an R22(8) graph‐set interaction through the carboxyl­ate group of 3,4‐D across the protonated N/N sites of 2‐ABOX [N?O 2.546 (3) and 2.795 (3) Å]. Cation/anion pairs associate across an inversion centre forming discrete tetramers via an additional three‐centre hydrogen‐bonding association from the latter N amino proton to a phenoxy O atom [N?O 3.176 (3) Å] and a carboxyl­ate O atom [N?O 2.841 (3) Å]. This formation differs from the polymeric hydrogen‐bonded chains previously observed for adduct structures of 2‐ABOX with carboxyl­ic acids.  相似文献   

15.
The structure of the title compound, [Co4(C9H3O6)2(OH)2(C8H6N4)(H2O)2]·2H2O, contains three separate species, namely the μ5‐bridging C9H3O63? anion, the doubly chelating and therefore μ2‐bridging C8H6N4 ligand (bi­pyrimidine, BPM), and the dihydrated di­aqua­di­hydroxy tetranuclear cationic cluster, [Co4(OH?)2(H2O)2]6+·2H2O, which lies on a crystallographic centre of symmetry, as does the BPM ligand with, in this case, the centre of symmetry coincident with the midpoint of the C—C bond joining the six‐membered rings. Within the cation cluster, the Co atoms of one pair are five‐coordinate and those of the other six‐coordinate.  相似文献   

16.
The two‐step one‐pot oxidative decarbonylation of [Fe2(S2C2H4)(CO)4(PMe3)2] ( 1 ) with [FeCp2]PF6, followed by addition of phosphane ligands, led to a series of diferrous dithiolato carbonyls 2 – 6 , containing three or four phosphane ligands. In situ measurements indicate efficient formation of 1 2+ as the initial intermediate of the oxidation of 1 , even when a deficiency of the oxidant was employed. Subsequent addition of PR3 gave rise to [Fe2(S2C2H4)(μ‐CO)(CO)3(PMe3)3]2+ ( 2 ) and [Fe2(S2C2H4)(μ‐CO)(CO)2(PMe3)2(PR3)2]2+ (R=Me 3 , OMe 4 ) as principal products. One terminal CO ligand in these complexes was readily substituted by MeCN, and [Fe2(S2C2H4)(μ‐CO)(CO)2(PMe3)3(MeCN)]2+ ( 5 ) and [Fe2(S2C2H4)(μ‐CO)(CO)(PMe3)4(MeCN)]2+ ( 6 ) were fully characterized. Relevant to the Hred state of the active site of Fe‐only hydrogenases, the unsymmetrical derivatives 5 and 6 feature a semibridging CO ligand trans to a labile coordination site.  相似文献   

17.
The title compound, [Zn(C2H3O2)(C6H18N4)][B5O6(OH)4], contains mixed‐ligand [Zn(CH3COO)(teta)]+ complex cations (teta is triethylenetetramine) and pentaborate [B5O6(OH)4] anions. The [B5O6(OH)4] anions are connected to one another through hydrogen bonds, forming a three‐dimensional supramolecular network, in which the [Zn(CH3COO)(teta)]+ cations are located.  相似文献   

18.
The primary geometry about the TeIV atom in the title compound, [TeCl2(C8H6Cl)(C3H5O)] or C11H11Cl3OTe, is a pseudo‐trigonal‐bipyramidal arrangement, with two Cl atoms in apical positions, and the lone pair of electrons and C atoms in the equatorial plane. The TeIV atom is involved in three secondary interactions, two intramolecular [Te?O = 2.842 (3) Å and Te?Cl3 = 3.209 (1) Å] and one intermolecular [Te?Cl = 3.637 (1) Å], the latter giving rise to a helical chain. These helices are linked by C—H?O interchain interactions.  相似文献   

19.
The dicarbollide ion, nido‐C2B9H112? is isoelectronic with cyclopentadienyl. Herein, we make dysprosiacarboranes, namely [(C2B9H11)2Ln(THF)2][Na(THF)5] (Ln=Dy, 1Dy ) and [(THF)3(μ‐H)3Li]2[{η5‐C6H4(CH2)2C2B9H9}Dy{η25‐C6H4(CH2)2C2B9H9}2Li] 3Dy and show that dicarbollide ligands impose strong magnetic axiality on the central DyIII ion. The effective energy barrier (Ueff) for the loss of magnetization can be varied by the substitution pattern on the dicarbollide. This finding is demonstrated by comparing complexes of nido‐C2B9H112? and nido‐[o‐xylylene‐C2B9H9]2?, which show a Ueff of 430(5) K and 804(7) K, respectively. The blocking temperature defined by the open hysteresis temperature of 3Dy reaches 6.8 K. Moreover, the linear complex [Dy(C2B9H11)2]? is predicted to have comparable properties with the linear [Dy(CpMe3)2]+ complex. As such, carboranyl ligands and their derivatives may provide a new type of organometallic ligand for high‐performance single‐molecule magnets.  相似文献   

20.
The one‐dimensional chain catena‐poly­[[aqua(2,2′:6′,2′′‐terpyridyl‐κ3N)­nickel(II)]‐μ‐cyano‐κ2N:C‐[bis­(cyano‐κC)nickelate(II)]‐μ‐cyano‐κ2C:N], [Ni(terpy)(H2O)]‐trans‐[Ni‐μ‐(CN)2‐(CN)2]n or [Ni2­(CN)4­(C15H11N3)(H2O)], consists of infinite linear chains along the crystallographic [10] direction. The chains are composed of two distinct types of nickel ions, paramagnetic octahedral [Ni(terpy)(H2O)]2+ cations (with twofold crystallographic symmetry) and diamagnetic planar [Ni(CN)4]2? anions (with the Ni atom on an inversion center). The [Ni(CN)4]2? units act as bidentate ligands bridging through two trans cyano groups thus giving rise to a new example of a transtrans chain among planar tetra­cyano­nickelate complexes. The coordination geometry of the planar nickel unit is typical of slightly distorted octahedral nickel(II) complexes, but for the [Ni(CN)4]2? units, the geometry deviates from a planar configuration due to steric interactions with the ter­pyridine 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