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.
The title compound, {[Cd2(C9HNO8)(H2O)4]·H2O}n, consists of two crystallographically independent CdII cations, one tetrabasic pyridine‐2,3,5,6‐tetracarboxylate (pdtc) anion, four coordinated water molecules and one solvent water molecule. The CdII cations have distorted square‐antiprismatic (one pyridine N, six carboxylate O and one water O atom) and octahedral (three carboxylate O and three water O atoms) coordination environments. Each pdtc ligand employs its pyridine and carboxylate groups to chelate and bridge seven CdII cations. The square‐antiprismatic coordinated CdII cations are linked by pdtc ligands into a lamellar framework structure, while the octahedral coordinated CdII cations are bridged by the μ2‐carboxylate O atoms and the pdtc ligands into a chain network that further joins neighbouring lamellae into a three‐dimensional porous network. The cavities are filled with solvent water molecules that are linked to the host through complex hydrogen bonding.  相似文献   

2.
The title compound, (C24H24N7)2[Cd5Cl16(H2O)4]·H2O, contains a [Cd5Cl16(H2O)4]6− anion, two triply protonated tris[(1H‐benzimidazol‐3‐ium‐2‐yl)methyl]amine cations and one solvent water molecule. The structure of the anion is a novel chloride‐bridged pentanuclear cluster. The five unique CdII centres have quite different coordination environments. Two of the central hexacoordinated CdII cations have a CdOCl5 chromophore, in which each CdII cation is ligated by four bridging chloride ligands, one terminal chloride ligand and one water molecule, adopting a distorted octahedral environment. The third central CdII cation is octahedrally coordinated by four bridging chloride ligands and two water molecules. Finally, the two terminal CdII cations are pentacoordinated by two bridging and three terminal chloride ligands and adopt a trigonal–bipyramidal geometry. A three‐dimensional supramolecular network is formed through intra‐ and intermolecular O—H...O, O—H...Cl, N—H...Cl and N—H...O hydrogen bonds and π–π interactions between the cations and anions.<!?tpb=20.6pt>  相似文献   

3.
Poly(ethylene oxide)-poly(methyl acrylate) diblock copolymers with narrow molecular weight distributions were synthesized using atom transfer radical polymerization. The copolymers were used as micellar templates for the synthesis of mesoporous silicas. The products were characterized using small-angle X-ray scattering, transmission electron microscopy (TEM) and nitrogen adsorption. The obtained silicas exhibited two-dimensional hexagonal structures of cylindrical mesopores, and thus can be classified as SBA-15 silicas. In some cases, the size of ordered domains was very small. The (100) interplanar spacings were 13–17 nm, depending on the size of the diblock copolymer used and on the synthesis conditions. Nitrogen adsorption showed that the silicas exhibited specific surface areas of 350–800 m2 g−1, pore volumes ∼1 cm3 g−1, and narrow pore size distributions. The BJH (nominal) pore diameters were up to ∼20 nm, but actual diameters of cylindrical pores are expected to be somewhat smaller. In many cases, the mesopores exhibited constrictions.  相似文献   

4.
Three novel complex salts containing the cation trans-[Rh(β-Pic)4Cl2]+ with the anions Cl (I), ReO4 (II), and ClO4 (III) were obtained and characterized by elemental analysis, X-ray diffraction, NMR spectroscopy, and IR spectroscopy. The complex trans-[Rh(β-Pic)4Cl2]ReO4 crystallizes from DMF as a solvate in which solvent molecules fill the channels formed by the cations and anions. The thermal properties of complexes I, II, and II · DMF were examined by DTA. Final and some intermediate products of the thermolysis were isolated and characterized by physicochemical methods.  相似文献   

5.
The complexation behavior of a bis‐strapped porphyrin ligand ( 1 ) towards CdII has been investigated by 1H and 113Cd NMR spectroscopy with the help of X‐ray diffraction structures. The presence of an overhanging carboxylic acid group on each side of the macrocycle is responsible for the instantaneous insertion of the metal ion(s) at room temperature, and allows the formation of bimetallic species with unusual coordination modes at the origin of unique dynamic behaviors. In the absence of base, a C2‐symmetric bimetallic complex ( 1Cd2 ) is readily formed, in which the porphyrin acts as a bridging ligand. Both CdII ions are bound to the N core and to a COO? group of a strap. In contrast, the presence of a base induces a two‐step binding process with the successive formation of mono and bimetallic species ( 1Cd and 1Cd?CdOAc ). Formally, a CdII ion is first inserted into the N core and experiences a strong out‐of‐plane (OOP) displacement due to the binding of an overhanging carbonyl group in an apical position. A second CdII ion then binds exclusively to the strap on the opposite side, in a so‐called hanging‐atop (HAT) coordination mode. These two complexes display a fluxional behavior that relies on intraligand migration processes of the metal ion(s). In 1Cd , the CdII ion exchanges between the two equivalent overhanging apical ligands by funneling through the porphyrin ring. In 1Cd?CdOAc , the two CdII ions exchange their coordination mode (HAT?OOP) in a concerted way while staying on their respective side of the macrocycle, in a so‐called Newton’s cradle‐like motion. The intramolecular pathway was notably evidenced by variable temperature 113Cd heteronuclear NMR experiments. This coupled motion of the CdII cations is under allosteric control; the addition of an acetate anion (the allosteric effector) to the “resting” C2‐symmetric complex 1Cd2 affords the dissymmetric complex 1Cd?CdOAc and triggers equilibrium between its two degenerate states. The rate of the swinging motion further depends on the concentration of AcO?, with a higher concentration leading to a slower motion. As compared with the related PbII and BiIII bimetallic complexes, the Newton’s cradle‐like motion proceeds faster with the smaller CdII ion. These results open the way to novel multistable devices and switches.  相似文献   

6.
The displacement adsorption enthalpies (ΔH) of denatured α-Amylase (by 1.8 mol L−1 GuHCl) adsorbed onto a moderately hydrophobic surface (PEG-600, the end-group of polyethylene glycol) from solutions (x mol L−1 (NH4)2SO4, 0.05 mol L−1 KH2PO4, pH 7.0) at 298 K are determined by microcalorimeter. Further, entropies (ΔS), Gibbs free energies (ΔG) and the fractions of ΔH, ΔS, and ΔG for net adsorption of protein and net desorption of water are calculated in combination with adsorption isotherms of α-Amylase based on the stoichiometric displacement theory for adsorption (SDT-A) and its thermodynamics. It is found that the displacement adsorptions of denatured α-Amylase onto PEG-600 surface are exothermic and enthalpy driven processes, and the processes of protein adsorption are accompanied with the hydration by which hydrogen bond form between the adsorbed protein molecules favor formation of β-sheet and β-turn structures. The Fourier transformation infrared spectroscopy (FTIR) analysis shows that the contents of ordered secondary structures of adsorbed α-Amylase increase with surface coverages and salt concentrations increment.  相似文献   

7.
In this study, the new vic-dioxime ligand (LH2) and its complexes with CoII, NiII, CuII, ZnII, CdII and UO2VI are described. The structures of these complexes were characterized by elemental analyses, i.r., 1H- and 13C-n.m.r. spectra, u.v.–vis. spectroscopy, magnetic susceptibility measurements, conductivity measurements and thermogravimetric analyses (t.g.a.).  相似文献   

8.
Adsorption isotherms of potential-determining H+ and OH ions and the pH dependences of the specific surface charge of detonation nanodiamond (DND) particles are obtained in a pH range of 3–10 by the acid-base titration of their hydrosols containing 0.001–1 M LiCl, NaCl, KCl, NaNO3, KNO3, and NaClO4 as background electrolytes. The data obtained attest to the chemical nonuniformity (heterogeneity) of a DND surface and different degrees of binding of background electrolyte cations and anions with ionized groups. It is revealed that the adsorption of OH-anions diminishes in the lyotropic series of cations Na+ > K+ > Li+ and increases with a decrease in the adsorbability of anions in the following series: NO3 ≊ ClO4 > Cl. The adsorption of potential-determining H+ and OH ions on a DND surface containing two types of functional groups, i.e., acidic carboxyl and amphoteric hydroxyl groups, is simulated by the Protofit software package. The optimal surface densities and ionization constants that correspond to minimal deviations of model adsorption isotherms from the experimental curves are found for these groups.  相似文献   

9.
3‐(Pyridin‐4‐yl)acetylacetone (HacacPy) acts as a pyridine‐type ligand towards CdII and HgII halides. With CdBr2, the one‐dimensional polymer [Cd(μ‐Br)2(HacacPy)Cd(μ‐Br)2(HacacPy)2] is obtained in which five‐ and six‐coordinated CdII cations alternate in the chain direction. Reaction of HacacPy with HgBr2 results in [Hg(μ‐Br)Br(HacacPy)], a polymer in which each HgII centre is tetracoordinated. In both compounds, each metal(II) cation is N‐coordinated by at least one HacacPy ligand. Equimolar reaction between these CdII and HgII derivatives, either conducted in ethanol as solvent or via grinding in the solid state, leads to ligand redistribution and the formation of the well‐ordered bimetallic polymer catena‐poly[[bromidomercury(II)]‐μ‐bromido‐[aquabis[4‐hydroxy‐3‐(pyridin‐4‐yl)pent‐3‐en‐2‐one]cadmium(II)]‐di‐μ‐bromido], [CdHgBr4(C10H11NO2)2(H2O)]n or [{HgBr}(μ‐Br){(HacacPy)2Cd(H2O)}(μ‐Br)2]. HgII and CdII cations alternate in the [100] direction. The HacacPy ligands do not bind to the HgII cations, which are tetracoordinated by three bridging and one terminal bromide ligand. The CdII centres adopt an only slightly distorted octahedral coordination. Three bromide ligands link them in a (2 + 1) pattern to neighbouring HgII atoms; two HacacPy ligands in a cis configuration, acting as N‐atom donors, and a terminal aqua ligand complete the coordination sphere. Classical O—H…Br hydrogen bonds stabilize the polymeric chain. O—H…O hydrogen bonds between aqua H atoms and the uncoordinated carbonyl group of an HacacPy ligand in a neighbouring strand in the c direction link the chains into layers in the (010) plane.  相似文献   

10.
Lewis acidic properties of transition aluminas whose surfaces have been doped with alkaline-earth metal cations (Ca2+ and Ba2+) were studied by means of the room temperature adsorption of carbon monoxide. The vibrational features of CO adsorbed at the surface of doped aluminas were investigated by IR spectroscopy in comparison with pure parent aluminas, while the quantitative and energetic features were studied by adsorption microcalorimetry. Various CO adspecies were found to form at the surface of both pure and doped-alumina, owing to the structural heterogeneity of the Al2O3 surface and to the presence of alkaline-earth metal cations. The surface heterogeneity was revealed by different vco stretching frequencies, namely vco≈2230, 2218 and 2205 cm−1 for coordinatively unsaturated tetrahedral Al3+ cations in different crystallographic configurations, and vco≈2186 and 2172 cm−1 for coordinatively unsaturated Ca2+ and Ba2+ cations, respectively. Heats of adsorption of ≈80, 70 and 55 kJ/mol were assigned to the formation of Al3+/CO complexes, ≈45 kJ/mol for Ca2+/CO and ≈30 kJ/mol for Ba2+/CO complexes. The latter value was estimated through a correlation curve existing between vco stretching frequencies and adsorption enthalpies. This correlation, already proposed in the past for CO adsorbed on non-d/d0/d10 metal cations, has been revisited and confirmed here, by including Al2O3 data for which an apparent lack of correlation between the two parameters was first observed. With respect to pure alumina, the population of Lewis acidic sites was found to be significantly depressed by the presence of alkaline-earth cus metal cations. These acidic sites are intrinsically weaker than tetrahedral cus Al3+ cations, as witnessed by smaller upward shifts of the vco stretching frequencies with respect to CO gas and lower heats of adsorption, in accordance with expectations from the charge/ionic radius ratios. Ca2+ cations were found to compete in adsorbing CO with Al3+ cations more efficiently than the larger Ba2+ cations. In the case of CaO/Al2O3 systems outgassed at 1023 K, a thin surface layer of calcium aluminate, not detected by XRD or HRTEM, was suggested to form.  相似文献   

11.
The title CdII compound, {[Cd2(C13H7NO4)2(H2O)4]·5H2O}n, was synthesized by the hydrothermal reaction of Cd(NO3)2·4H2O and 5‐(pyridin‐4‐yl)isophthalic acid (H2L). The asymmetric unit contains two crystallographically independent CdII cations, two deprotonated L2− ligands, four coordinated water molecules and five isolated water molecules. One of the CdII cations adopts a six‐coordinate octahedral coordination geometry involving three O atoms from one bidentate chelating and one monodentate carboxylate group of two different L2− ligands, one N atom of another L2− ligand and two coordinated water molecules. The second CdII cation adopts a seven‐coordinate pentagonal–bipyramidal coordination geometry involving four O atoms from two bidentate chelating carboxylate groups of two different L2− ligands, one N atom of another L2− ligand and two coordinated water molecules. Each L2− ligand bridges three CdII cations and, likewise, each CdII cation connects to three L2− ligands, giving rise to a two‐dimensional graphite‐like 63 layer structure. These two‐dimensional layers are further linked by O—H...O hydrogen‐bonding interactions to form a three‐dimensional supramolecular architecture. The photoluminescence properties of the title compound were also investigated.  相似文献   

12.
The adsorption isotherms of hemoglobin, peroxidase, and β-galactosidase on silochrome and mesoporous and biporous silicas were comparatively studied. Adsorption developed in two stages, including fast “reversible” protein adsorption (equilibrium was reached in t ≤ 1–2 h) and a “slow stage” of irreversible binding in t ≫ 24 h (multipoint adsorption). The corresponding equilibrium constants were determined. The mechanism of unlimited linear association of peroxidase in the adsorption layer on the surface of silochrome was established.  相似文献   

13.
It was established by X-ray diffraction, TPR, and EPR that microemulsion (m.e.) synthesis yields the binary oxides ZrO2(m.e.) and CeO2(m.e.) and the mixed oxide Zr0.5Ce0.5O2(m.e.) in the form of a tetragonal, cubic, and pseudocubic phase, respectively, having crystallite sizes of 5–6 nm. The bond energy of surface oxygen in the (m.e.) samples is lower than in their analogues prepared by pyrolysis. Hydrogen oxidation on the oxides under study occurs at higher temperatures than CO oxidation. ZrO2(m.e.) and CeO2(m.e.) are active in O2 formation during NO + O2 adsorption, while CeO2 is active during CO + O2 adsorption, too. However, its amount here is one-half to one-third its amount in the pyrolysis-prepared samples, signifying a reduced number of active sites, which are Zr4+ and Ce4+ coordinatively unsaturated cations and Me4+-O2− pairs. O2 radical anions are stabilized in the coordination sphere of Zr4+ coordinatively unsaturated cations via ionic bonding, and in the sphere of Ce4+ cations, via covalent bonding. Ionic bonds are stronger than ionic-covalent bonds and do not depend on the ZrO2 phase composition. Zr0.5Ce0.5O2 is inactive in these reactions because of the strong interaction of Zr and Ce cations. It is suggested that Ce(4 + β)+ coordinatively unsaturated cations exist on its surface, and their acid strength is lower than that of Zr4+ and Ce4+ cations in ZrO2 and CeO2, according to the order ZrO2 > CeO2 ≥ Zr0.5Ce0.5O2. Neither TPR nor adsorption of probe molecules revealed Zr cations on the surface of the mixed oxide.  相似文献   

14.
A relation was established between the composition of Cd x Zn1–x S nanoparticles and their ability to accumulate excess negative charge during irradiation. The rate of expenditure of the accumulated charge depends on the composition of the nanoparticles and is determined by their electric capacitance. A correlation was found between the photocatalytic activity of the Cd x Zn1–x S nanoparticles in the release of hydrogen from solutions of Na2SO3, their composition, and their capacity for photoinduced accumulation of excess charge. It was shown that Ni0 nanoparticles photodeposited on the surface of Cd x Zn1–x S are effective cocatalysts for the release of hydrogen. It was found that ZnII additions in photocatalytic systems based on Cd x Zn1–x S/Ni0 nanostructures have a promoting action on the release of hydrogen from water–ethanol mixtures. Translated from Teoreticheskaya i éksperimental’naya Khimiya, Vol. 45, No. 1, pp. 8–16, January-February, 2009.  相似文献   

15.
The novel title CdII coordination polymer, poly­[[di­chlorocad­mium(II)]‐di‐μ‐1,1′‐(1,4‐butane­diyl)­di­imidazole], [CdCl2(C10H14N4)2]n, (I), was obtained by reaction of CdCl2·2.5H2O and 1,1′‐(1,4‐butane­diyl)diimidazole (hereafter L). In (I), each L molecule coordinates to two CdII cations through its two aromatic N atoms, thus acting as a bridging bidentate ligand. The CdII cations, which lie on the inversion centre, are bridged by four L molecules to form a two‐dimensional (4,4)‐network. The two‐dimensional square‐grid sheets are superimposed in an offset fashion.  相似文献   

16.
Summary Complexes of CoII, NiII, CuII, ZnII, CdII, HgII and UO 2 II with benzil bis(4-phenylthiosemicarbazone), H2BPT, have been synthesized and their structures assigned based on elemental analysis, molar conductivity, magnetic susceptibility and spectroscopic measurements. The i.r. spectra suggest that the ligand behaves as a binegative quadridentate (NSSN) (CoII, CuII, HgII and UO 2 II complexes) or as a binegative quadridentate-neutral bidentate chelating agent (NiII, ZnII and CdII complexes). Octahedral structures for the CoII and NiII complexes and square-planar structure for the CuII complex are suggested on the basis of magnetic and spectral evidence. The crystal field parameters (Dq, B and B) for the CoII complex are calculated and agree fairly well with the values reported for known octahedral complexes. The ligand can be used for the microdetermination of NiII ions of concentration in the 0.4–6×10–4 mol l–1 range and the apparent formation constant for the species generated in solution has also been calculated.  相似文献   

17.
Summary New complexes of general formulae [Ni(HL)2], [ML]·H2O and [Cu(HL)X] (H2L = pyrrole-2-aldehyde Schiff bases ofS-methyl- andS-benzyldithiocarbazates; X = Cl or Br; M = NiII, CuII, ZnII or CdII) were prepared and characterized by a variety of physicochemical techniques. The Schiff bases coordinate as NS bidentate chelating agents in [Ni(HL)2] and [Cu(HL)X], and as tridentate NNS chelates in [ML] (M = NiII, CuII, ZnII or CdII). Both the [Ni(HL)2] and [NiL] complexes are diamagnetic and square-planar. Based on magnetic and spectroscopic evidence, thiolate sulphur-bridged dimeric square-planar structures are assigned to the [Cu(HL)X] and [ML] (M = NiII or CuII) complexes. The complexes ML (M = ZnII or CdII) are polymeric and octahedral.  相似文献   

18.
The title salt, catena‐poly[trimethylsulfonium [μ2‐chlorido‐di‐μ2‐thiocyanato‐cadmate(II)]] {(C3H9S)[CdCl(NCS)2]}n, consists of trimethylsulfonium cations sandwiched between layers of a two‐dimensional polyanion. The CdII centre displays a distorted octahedral environment coordinated by two bridging Cl atoms, two thiocyanate N atoms and two thiocyanate S atoms. The thiocyanate groups adopt the μ‐1,3‐coordination mode and bridge the CdII centres into a one‐dimensional zigzag chain extended along the [110] direction. The CdII centres of the zigzag chains are crosslinked by bridging Cl atoms, forming a two‐dimensional polyanion. The two‐dimensional anions are linked to layers of trimethylsulfonium cations by weak intermolecular C—H...Cl hydrogen bonds, forming the three‐dimensional structure.  相似文献   

19.
The dissociation of [CuII(L)His]•2+ complexes [L=diethylenetriamine (dien) or 1,4,7-triazacyclononane (9-aneN3)] bears a strong resemblance to the previously reported behavior of [CuII(L)GGH]•2+ complexes. We have used low-energy collision-induced dissociation experiments and density functional theory (DFT) calculations at the B3LYP/6-31+G(d) level to study the macrocyclic effect of the auxiliary ligands on the formation of His•+ from prototypical [CuII(L)His]•2+ systems. DFT revealed that the relative energy barriers of the same electron-transfer (ET) dissociation pathways of [CuII(9-aneN3)His]•2+ and [CuII(dien)His]•2+ are very similar, with the ET reactions of [CuII(9-aneN3)His]•2+ leading to the generation of two distinct His•+ species; in contrast, the proton transfer (PT) dissociation pathways of [CuII(9-aneN3)His]•2+ and [CuII(dien)His]•2+ differ considerably. The PT reactions of [CuII(9-aneN3)His]•2+ are associated with substantially higher barriers (>13 kcal/mol) than those of [CuII(dien)His]•2+. Thus, the sterically encumbered auxiliary 9-aneN3 ligand facilitates ET reactions while moderating PT reactions, allowing the formation of hitherto nonobservable histidine radical cations.  相似文献   

20.
Summary New metal complexes [M(NNNS)X] (M = NiII, CuII, ZnII and CdII; NNNS = anion of the quadridentate ligands formed from S-methyl--N-(2-aminophenyl)-methylenedithiocarbazate and pyridine-2-aldehyde or 6-methylpyridine-2-aldehyde; X = Cl, NCS, NO3 or I) and [Co(NNNS)Cl2]·2H2O have been prepared and characterized by elemental analysis and conductance measurements. Magnetic and spectroscopic evidence support a five-coordinate structure for [M(NNNS)X] (M = NiII, CuII, ZnII and CdII; X = Cl, NCS) and a squareplanar structure for [Ni(NNNS)]X (X = NO3 or I). The [Co(NNNS)Cl3]·2H2O complex is low-spin and octahedral. The Schiff bases and some of their metal complexes were tested against three pathogenic fungi, Alternaria alternata, Curvularia geniculata and Fusarium palidoroseum. The metal complexes are less fungitoxic than the free 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