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 171 毫秒
1.
Addition of excess trimethylphosphine and a halide source to a solution of W(CO)(acac)2(η2-L) (L = NCPh and OCMe2) leads to displacement of L and one acetylacetonate chelate to produce electron-rich, seven-coordinate complexes of the formula W(CO)(acac)(X)(PMe3)3 (X = Cl, Br, and I). Use of NaN3 instead of a halide source leads primarily to loss of carbon monoxide and dinitrogen, and protonation from adventitious water yields the cationic imido complex [W(NH)(acac)(PMe3)3]+. Heating [W(NH)(acac)(PMe3)3]+ in aromatic isocyanates at high temperature results in isocyanate insertion into the NH imido bond to form new C-N bonds. An alternate route to related imido complexes involves heating [W(O)(acac)(PMe3)3]+ with phenyl isocyanate at high temperatures to yield the substituted imido complex [W(NPh)(acac)(PMe3)3]+.  相似文献   

2.
Metal-oxygen bonding complexes (M = MgII, MnII, NiII, MoVI, WVI, PdII, SbIII, BiIII, FeIII, TiIV, KI, BaII, ZrIV and HfIV) with a hinokitiol (Hhino; 2-hydroxy-4-isopropylcyclohepta-2,4,6-trienone or β-thujaplicin) ligand, which has two unequivalent oxygen donor atoms, were synthesized and characterized by elemental analysis, TG/DTA, FT-IR and solution (1H and 13C) NMR spectroscopy. Single-crystal X-ray structure analysis revealed various molecular structures for the complexes, which were classified into several families of family, i.e. type A [MII(hino)2(L)]2 (M = MgII, MnII, NiII; L = EtOH or MeOH), with a dimeric structure consisting of one bridging hino anion, one chelating hino anion and one alcohol or water molecule, type B, with the octahedral, cis-dioxo, bis-chelate complexes cis-[MVIO2(hino)2] (M = MoVI, WVI), type C, with square planar complex [MII(hino)2] (M = PdII), type D, with tris-chelate, 7-coordinate complexes with one inert electron pair [MIII(hino)3] (M = SbIII, BiIII), type D′, with the bis-chelate, pseudo-6-coordinate complexes with one inert electron pair [MIII(hino)2X] (M = SbIII, X = Br), type E, with tris-chelate, 6-coordinate complexes with Δ and Λ isomers [MIII(hino)3] (M = FeIII), type E′ of bis-chelate, 6-coordinate complex [MIV(hino)2X2] (M = TiIV, X = Cl), type F, with water-soluble alkali metal salts [MI(hino)] (M = KI), and type H, with tetrakis-chelate, 8-coordinate complexes [MIV(hino)4](M = ZrIV, HfIV). These structural features were compared with those of metal complexes with a related ligand, tropolone (Htrop). The antimicrobial activities of these complexes, evaluated in terms of minimum inhibitory concentration (MIC; μg mL−1) in two systems, were compared to elucidate the relationship between structure and antimicrobial activity.  相似文献   

3.
The present paper describes a new tripodal ligand containing imidazole and pyridine arms and its first cis-[RuIII(L)(Cl)2]ClO4 complex (1). The crystal structure of 1 shows RuIII in a distorted octahedral geometry, in which two chloride ions, cis-positioned to each other, are coordinated besides the four nitrogen atoms from the tetradentate ligand L. The cyclic voltammogram of 1 exhibits three redox processes at −67, +73 and +200 mV versus SCE, which are attributed to the RuIII/RuII couple in the cis-[RuIII(L)(Cl)2]+, cis-[RuII(L)(H2O)(Cl)]+ and cis-[RuII(L)(H2O)2]2+, respectively. After chemical reduction (Zn(Hg) or EuII) only the cis-[RuII(L)(H2O)2]2+ species is observed in the cyclic voltammetry. Complex 1 absorbs at 470 nm (ε=1.4×103 mol−1 L cm−1), 335 nm (ε=7.9×103 mol−1 L cm−1), 301 nm (ε=6.7×103 mol−1 L cm−1) and 264 nm (ε=9.9×103 mol−1 L cm−1), in water solution (CF3COOH, 0.01 mol L−1, μ=0.1 mol L−1 with CF3COONa). Spectroelectrochemical experiments show a decrease of the bands at 335 and 301 nm, which are attributed to LMCT transitions from the chloride to the RuIII center and the appearance of a broad band at 402 nm ascribed to MLCT transition from the RuII center to the pyridine ligand. The lability of the water ligands in the cis-[RuII(L)(H2O)2]2+ species has been investigated using the auxiliary ligand pyrazine. Reactions in the presence of stoichiometric and excess of pyrazine yield the same species, cis-[RuII(L)(H2O)(pz)]2+, which exhibits a reversible redox process at 493 mV versus SCE and absorbs at 438 nm (ε=5.1×103 mol−1 L cm−1) and 394 nm (ε=4.2×103 mol−1 L cm−1). Experiments performed with a large excess of pyrazine gave a specific rate constant k1=(2.8±0.5)×10−2 M−1 s−1, at 25 °C, in CF3COOH, 0.01 mol L−1, μ=0.1 mol L−1 (with CF3COONa).  相似文献   

4.
Two series of A-frame complexes, [Pd2(dppm)2(R)2(μ-X)]+ (R = Me and X = Cl, Br, I, H; R = Mes and X = Br, I), were investigated by cyclic voltammetry (CV). The 2-electron reduction potentials for the first series increase from I (−1.10), Br (−1.17), Cl (−1.25) to H (−1.65 V versus SCE, in CHCl3), as well as in the second series; Br (−1.35) and I (−1.38 V versus SCE, in THF). The nature of the LUMO where the electron reduction takes place is qualitatively addressed by DFT on the corresponding model complexes [Pd2(H2PCH2PH2)2(R)2(μ-X)]+. The LUMO and (LUMO + 1) of the halide derivatives exhibit the presence of Pd dx2-y2 atomic orbitals interacting in an anti-bonding fashion with the n-donor orbitals of X, P, and Me, explaining in part the observed reactivity upon reduction. The X-ray structure of [Pd2(dppm)2(Me)2(μ-Br)]+ compound exhibits the typical A-frame structure with a Pd?Pd non-bonding distance of 3.036(1) Å, and long Pd-Br bonds of 2.5623(5) and 2.5793(5) Å.  相似文献   

5.
Bis(azido)bis(phosphine)-Pd(II) and -Pt(II) complexes, [M(N3)2L2] {L = PMe3, PEt3, PMe2Ph, dppe = 1,2-bis(diphenylphosphino)ethane}, underwent 1,3-dipolar cycloaddition with organic chiral isothiocyanates (R-NCS: R = (S)-(+)-1-phenylethyl, (R)-(−)-1-phenylethyl, (±)-1-phenylethyl, (S)-(+)-1-indanyl) to give the corresponding tetrazole-thiolato Pd(II) and Pt(II) complexes, trans-[M{S[CN4(R)]}2L2] or [M{S[CN4(R)]}2(dppe)]. Spectroscopic (IR and NMR) and X-ray structural analyses of the products showed that the absolute configuration of the starting organic isothiocyanates is retained throughout the reaction. Further treatments of the isolated tetrazole-thiolato complexes with electrophiles such as HCl or benzoyl chloride produced heterocyclic compounds containing a tetrazole thione or a tetrazolyl sulfide group. In addition, organic tetrazole thiones, [S = {CN4H(R)}] containing a chiral moiety, were prepared from NaN3 and R-NCS in the presence of water.  相似文献   

6.
Cation effects are studied on the excitation energy transfer reaction between anionic complexes, i.e., [Tb(dpa)3]3− (dpa=2,6-pyridinedicarboxylate) quenched by [Cr(ox)3]3− (ox=oxalate ion), [Cr(mal)3]3− (mal=malonate ion) and [Nd(dpa)3]3− in aqueous solutions in the presence of alkali metal ions added for adjustments of ionic strengths. In the quenching reaction of [Cr(ox)3]3−, magnitudes of quenching rate constants (kq) and energy transfer rate constant in encounter complex (k1) are changed by the cations in the order of Li+ < Na+ < K+ ≈ Rb+ ≈ Cs+, that is quite contrary of the cation effect on energy transfer reaction between [Ru(N-N)3]2+ and [Cr(ox)3]3−, reported in the previous paper. On the other hand, the rate constants in quenching reactions by [Cr(mal)3]3− and [Nd(dpa)3]3− remain almost constant. This result indicates that more separated donor-acceptor pair is not sensitive to coexisting cations.  相似文献   

7.
The solution structures of the lanthanide complexes, [Ln(L)(NO3)3] and [Ln(L)2(NO3)3], where L = bis(diphenylphosphorylmethyl)mesitylene and Ln = La, Ce, Nd, Er, were investigated by 31P NMR and IR spectroscopy, conductivity and sedimentation analysis. Variable-temperature 31P{1H} NMR spectroscopy was used to identify species present in solution and to monitor their interconversions. The results indicate that equilibrium between molecular complexes [Ln(L)n(NO3)3]0 and cationic species (as ion pairs [Ln(L)n(NO3)2]+ · (NO3) and as free ions [Ln(L)n(NO3)2]+, throughout n = 1, 2) in solutions can be observed by 31P{1H} NMR spectroscopy due to separate detection of the molecular complexes and cationic species. The chelate coordination of the ligand and nitrate ions is retained in all complex species at ambient temperature except for [Er(L)2(NO3)3]. The crystal structure of [Nd(L)(NO3)3(MeCN)]MeCN was determined by X-ray diffraction.  相似文献   

8.
The interaction of [PdII(mida)(Cl)] (1) (mida2− = N-methyliminodiacetate) and [PdII(pydc)(Cl)] (2) (pydc2− = pyridyl-2,6-dicarboxylate) with adenosine-5′-monophosphate (AMP), inosine-5′-monophosphate (IMP) and glutathione (GSH) was studied kinetically as a function of [L] (L = AMP, IMP, GSH) and [Cl] and temperatures (10-35 °C) at pH 4.0. The kinetic results suggest that the reaction of 1 and 2 with the 5′-nucleotides (AMP, IMP) is characterized by the hydrolysis of chloro-complexes followed by the aquo-substitution with purine based 5′-nucleotides through its N7 atom. The reaction of 1 and 2 with GSH takes place through the direct chloride replacement with GSH. Kinetic data and activation parameters are interpreted in terms of an associative mechanism and discussed in reference to the data reported earlier. The [PtII(mida)(Cl)] (3) and [PtII(pydc)(Cl)] (4) complexes were prepared and allowed to interact with AMP and IMP and their reaction products were characterized by 1H NMR studies. The antitumor activity of 3 and 4 was examined against MCF-7 (breast cancer), NCI-H460 (lung cancer) and SF-268 (CNS) cell lines.  相似文献   

9.
Preliminary pharmacological studies of various nitric oxide (NO) photo-releasing agents are reported based on the flash-photolysis studies of the nitro ruthenium complexes cis-[RuII(NO2)L(bpy)2]+ (bpy = 2,2′-bipyridine and L = pyridine, 4-picoline and pyrazine) and [RuII(NO2)(bpy)(terpy)]+ (terpy = terpyridine) in physiological medium. The net photoreactions under these conditions are two primary photoproducts, in (I) there is RuII-NO2 photoaquation, where the photoproducts are RuII-H2O plus and (II) homolytic dissociation of NO from a coordinated nitrito to derive the RuII-OH2 specie and NO. Based on photochemical processes, the nitro ruthenium complexes were incorporated in water in oil (W/O) microemulsion and used in the vasorelaxation induced experiment. Denuded rat aortas were contracted with KCl and nitro ruthenium complexes in microemulsion were added. Perfusion pressures were recorded while arteries were irradiated at 355 nm The time to reach maximum relaxation was longer for [RuII(NO2)(bpy)(terpy)]+ complex (ca. 50 min, n = 6) than for cis-[Ru(NO2)L(bpy)2]+ with L = py and 4-pic complex (ca. 28 min, n = 6) and cis-[Ru(NO2)(bpy)2 (pz)]2+ complex (ca. 24 min, n = 5).  相似文献   

10.
A series of osmium(VI) nitrido complexes containing pyridine-carboxylato ligands OsVI(N)(L)2X (L = pyridine-2carboxylate (1), 2-quinaldinate (2) and X = Cl (a), Br (1b and 2c) or CH3O (2b)) and [OsVI(N)(L)X3] (L = pyridine-2,6-dicarboxylate (3) and X = Cl (a) or Br (b)) have been synthesised. Complexes 1 and 2 are electrophilic and react readily with various nucleophiles such as phosphine, sulfide and azide. Reaction of OsVI(N)(L)2X (1 and 2) with triphenylphosphine produces the osmium(IV) phosphiniminato complexes OsVI(NPPh3)(L)2X (4 and 5). The kinetics of nitrogen atom transfer from the complexes OsVI(N)(L)2Br (2c) (L = 2-quinaldinate) with triphenylphosphine have been studied in CH3CN at 25.0 °C by stopped-flow spectrophotometric method. The following rate law is obtained: −d[Os(VI)]/dt = k2[Os(VI)][PPh3]. OsVI(N)(L)2Cl (L = 2-quinaldinate) (2a) reacts also with [PPN](N3) to give an osmium(III) dichloro complex, trans-[PPN][OsIII(L)2Cl2] (6). Reaction of OsVI(N)(L)2Cl (L = 2-quinaldinate) (2a) with lithium sulfide produces an osmium(II) thionitrosyl complex OsII(NS)(L)2Cl (7). These complexes have been structurally characterised by X-ray crystallography.  相似文献   

11.
A series of new heterometallic CuIIZnII and NiIIZnII complexes with N- and N,O open-chain multidentate ligands (L1 = 4,6,6-trimethyl-1,9-diamino-3,7-diazanon-3-ene; L2 = 3,7-bis(2-aminoethyl)-1,3,5,7-tetraazabicyclo[3.3.1]nonane; L3 = 1,15-dihydroxy-7,9,9-trimethyl-3,6,10,13-tetraazapentadec-6-ene and L4 = 1-hydroxy-9-oxy-4,6,6-trimethyl-3,7-diazanon-3-ene) have been prepared through the “direct template synthesis” approach, which is a combination of classical template reactions of amines with acetone/formaldehyde and the “direct synthesis” method based on using elemental metals as starting materials. There is a significant decrease in the reaction time when the “direct synthesis” method is used compared to the conventional template condensation methods. X-ray crystallographic analyses of the complexes with the general formula M(L)ZnX4 and [CuL4ZnCl3]2 (M = Cu2+, Ni2+; L = L1-L3; X = Cl, NCS) reveal the presence of long intermolecular distance interactions, such as semi-coordination, S?S and H-bonding, in their crystal organization.  相似文献   

12.
A series of cationic palladium complexes [(4-Mebti)PdL]+ with 4-Mebti = anion of bis(4-methylthiazolylimino)isoindoline and L = neutral ligand with group 16 donor atom has been prepared from the chlorido derivative [(4-Mebti)PdCl] and NaBArF (BArF = tetrakis(3,5-bis(trifluoromethyl)phenyl)boranate) in the presence of the respective donor ligand. Crystallographic and spectroscopic analyses were achieved for species with L = SMe2, SeMe2, dmf, acetamide, diphenylurea, and formiate. The latter two complexes represent products from hydrolyses of phenyl isocyanate and dmf, respectively, which occur during the ligand exchange reactions. Several other O-donor ligands like thf, acetone, Me2O, water, and others are not bound to the palladium ion, and the dinuclear μ-chlorido derivative [{(4-Mebti)Pd}2Cl]+ is isolated in these cases instead. The crystallographic analyses prove the expected presence of distorted, pseudo-planar palladium chelates, and the degree of distortion correlates well with the chemical shifts observed for the proton nuclei of the terminal methyl groups in the 1H NMR experiment.  相似文献   

13.
The substitution behaviour of [PtCl(R)(COD)] (R = Me and Fc) complexes, by the stepwise addition of phosphine ligands, L (L = PPh3, PEt3 and P(NMe2)3), were investigated in situ by 1H and 31P NMR spectroscopy. Addition of less than two equivalents of the phosphine ligand results in the formation of dimeric molecules with the general formula trans-[Pt(R)(μ-Cl)(L)]2 for the sterically demanding systems where R = Me/L = P(NMe2)3 and R = Fc/L = PEt3, PPh3 and P(NMe2)3 while larger quantities resulted in cis- and trans mixtures of mononuclear complexes being formed. In the case of the relatively small steric demanding, strongly coordinating, PEt3 ligand the trans-[PtCl(R)(PEt3)2] mononuclear complexes were exclusively observed in both cases. The crystal structures of the two substrates, [PtCl(R)(COD)] (R = Me or Fc), as well as the cis-[PtCl(Fc)(PPh3)2] substitution product are reported.  相似文献   

14.
Cobalt(III) and rhodium(III) complexes of the series of [MIIICl3 − n(P)3 + n]n+ (M = Co or Rh; n = 0, 1, 2 or 3) have been prepared with the use of 1,1,1-tris(dimethylphosphinomethyl)ethane (tdmme) and mono- or didentate phosphines. The single-crystal X-ray analyses of both series of complexes revealed that the M-P and M-Cl bond lengths were dependent primarily on the strong trans influence of the phosphines, and secondarily on the steric congestion around the metal center resulting from the coordination of several phosphine groups. In fact, the M-P(tdmme) bonds became longer in the order of [MCl3(tdmme)] < [MCl2(tdmme)(PMe3)]+ < [MCl(tdmme)(dmpe)]2+ (dmpe = 1,2-bis(dimethylphosphino)ethane) < [M(tdmme)2]3+ for both CoIII and RhIII series of complexes, while the M-Cl bond lengths were shortened in this order (except for [M(tdmme)2]3+). Such a steric congestion around the metal center can also account for the structural and spectroscopic characteristics of the series of complexes, [MCl(tdmme)(dmpm, dmpe or dmpp)]2+ (dmpm = bis(dimethylphosphino)methane, dmpp = 1,3-bis(dimethylphosphino)propane). The X-ray analysis for [CoCl(tdmme)(dmpm or dmpe)](BF4)2 showed that all Co-P bonds in the dmpm complex were shorter by 0.03-0.04 Å than those in the dmpe complex. Furthermore, the first d-d transition energy of the CoIII complexes and the 1JRh-P(tdmme) coupling constants observed for the RhIII complexes indicated an unusual order in the coordination bond strengths of the didentate diphosphines, i.e., dmpm > dmpe > dmpp.  相似文献   

15.
Cytidine (cyt) and adenosine (ado) react with cis-[L2Pt(μ-OH)]2(NO3)2 (L = PMe3, PPh3) in various solvents to give the nucleoside complexes cis-[L2Pt{cyt(− H),N3N4}]3(NO3)3 (L = PMe3, 1),cis-[L2Pt{cyt(− H),N4}(cyt,N3)]NO3 (L = PPh3, 2), cis-[L2Pt{ado(− H),N1N6}]2(NO3)2 (L = PMe3, 3) and cis-[L2Pt{ado(− H),N6N7}]NO3 (L = PPh3, 4). When the condensation reaction is carried out in solution of nitriles (RCN, R = Me, Ph) the amidine derivatives cis-[(PPh3)2PtNH=C(R){cyt(− 2H)}]NO3 (R = Me, 5a; R = Ph, 5b) and cis-[(PPh3)2PtNH=C(R){ado(− 2H)}]NO3 (R = Me, 6a: R = Ph, 6b) are quantitatively formed. The coordination mode of these nucleosides, characterized in solution by multinuclear NMR spectroscopy and mass spectrometry, is similar to that previously observed for the nucleobases 1-methylcytosine (1-MeCy) and 9-methyladenine (9-MeAd). The cytotoxic properties of the new complexes, and those of the nucleobase analogs, cis-[(PPh3)2PtNH=C(R){1-MeCy(− 2H)}]NO3 (R = Me, 7a: R = Ph, 7b), cis-[(PPh3)2PtNH=C(R){9-MeAd(− 2H)}]NO3 (R = Me, 8a: R = Ph, 8b) have been investigated in a wide panel of human cancer cells. Interestingly, whereas the Pt(II) nucleoside complexes (1-4) did not show appreciable cytotoxicity, the corresponding amidine derivatives (7a, 7b, 8a, 8b, 5b, and 6b) exhibited a significant in vitro antitumor activity.  相似文献   

16.
Two closely related 1:1 salts are obtained upon electrocrystallization of BEDT-TTF (BEDT-TTF: bis(ethylenedithio)tetrathiafulvalene) in the presence of the isosteric [M(tfadt)2] dithiolene complexes (tfadt: 1-trifluoromethyl-2-cyano-1,2-dithiolato), which essentially differ by their spin state, S = 0 in [Au(tfadt)2], S = 1/2 in [Ni(tfadt)2]. In both [BEDT-TTF][M(tfadt)2] salts, the BEDT-TTF radical cations form chains with a strong lateral overlap and strong antiferromagnetic interactions while the paramagnetic anions in the nickel-containing salt [BEDT-TTF][Ni(tfadt)2] are essentially non-interacting. The structural differences between the nickel and gold complexes are analyzed and discussed.  相似文献   

17.
Solution and solid state 31P NMR studies were carried out on a series of [Pd2X2(dppm)2] (X = Cl (1a), Br (1b), I (1c)), or [Pd2XY(dppm)2] (X = Cl, (1d)) complexes and on methyl substituted derivatives such as [Pd2Cl2(dppm)(dppmMe)] (2), syn-[Pd2Cl2(dppmMe)2] (3), and anti-[Pd2Cl2(dppmMe)2] (4) (dppmMe = 1,1-bis(diphenylphosphino)ethane) in order to study and understand the conformational behaviour of the eight-membered Pd2P4C2 rings depending on the substituents and their stereochemistry. These complexes with metal-metal bonds and mutually trans-dppm ligands act as molecular pendulums. On the basis of temperature dependent spectra qualitative correlations have been found between the molecular conformations and the rate of a specific intramolecular motion called “swinging”. While for the extended-boat conformers (2 and 3) this exchange process is of intermediate energy (41-45 kJ mol−1), the barrier is definitely higher (∼54 kJ mol−1) for the extended-chair conformer 4. Changes of symmetry relations are reflected very vividly in the 31P NMR spectra.The observed different chemical shifts, “swinging” rates and activation free energies obtained for the boat and chair conformers are explained by the steric effects and low-temperature conformations of the axial phenyl groups.  相似文献   

18.
Syntheses, crystal structures and magnetic properties of two salts, [FBzTPP][Ni(mnt)2](1) and [FBzTPP]2[Cu(mnt)2](2) ([FBzTPP]+ = 1-(4′-fluorobenzyl)triphenylphosphinium, mnt2− = maleonitriledithiolate), are investigated. In 1, the anions and cations stack into well-segregated columns, and the Ni(III) ions form a 1D alternating chain in a [Ni(mnt)2] column through intermolecular Ni?S and π?π interactions with the Ni?Ni distances of 3.939, 4.057 and 4.101 Å, and the C-H?N hydrogen bonds are found between the [Ni(mnt)2] anion and the neighboring [FBzTPP]+ cation. The [Cu(mnt)2]2− anions in 2 do not form a column and no weak interactions exist between the anions duo to isolation of the [FBzTPP]+ cations, while C-H?F and C-H?S hydrogen bonds were found in 2. Magnetic susceptibility measurements in the temperature range 2-300 K show that 1 exhibits a spin-gap transition around 46 K, and antiferromagnetic interaction (θ = −49.0 K) in the high-temperature phase (HT) and spin gap (Δ/kb = 88.2 K) in the low-temperature phase (LT), while 2 shows a very weak antiferromagnetic coupling behavior with θ = −1.33 × 10−2 K.  相似文献   

19.
Cobalt(II), copper(II) and zinc(II) dipicolinato complexes having thiamine dication with compositions [HT][CoL2]·5H2O (1), [HT][CuL2]·5H2O (2) and [HT][ZnL2]·5H2O (3) (L = dipicolinato anion, T = thiamine cation) are synthesized, characterized by X-ray diffraction and other spectroscopic techniques. The thiamine part in these complexes exists as divalent cations. The second protonation of thiamine takes place at the less crowded nitrogen atom of the pyrimidine ring. These complexes are stabilized by electrostatic and hydrogen-bonding interactions of -N+-H and -COO groups along with crystallized water molecules. The complexes are stable in solution as determined by 1H NMR and visible study. The complex 1 has two medicinal components which can be easily separated by controlling pH in aqueous medium. It gets decomposed on treatment with sodium hydroxide at pH > 8 to form neutral complex {[Na2(μ-H2O)3(H2O)3][CoL2]·H2O}2.  相似文献   

20.
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