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 synthesis, characterization and crystal structures of substituted imidazolate bridged binuclear copper(II) complexes, [Cu2(dien)2(L)](ClO4)3, where dien = diethylenetriamine, L = imidazolate (im) ( 1 ), 2‐methylimidazolate (mim) ( 2 ) and benzimidazolate (bim) ( 3 ), are reported. The copper(II) ions of 1 — 3 posses a square planar coordination environment with dien coordinating as a tridentate ligand and the fourth position being occupied by a nitrogen atom of the bridging μ‐imidazolato group. In all three compounds the tendency to form additional long apical bonds at the copper(II) ions to the oxygen atoms of the perchlorate anions is observed. Temperature depended susceptibility data of polycrystalline samples reveal an antiferromagnetic coupling of the copper(II) atoms in 1 — 3 with J = —63.8, —75.4 and —36.8 cm—1, respectively. Significant changes for these coupling constants could not be observed for measurements on frozen aqueous solutions. ESR spectra for solids and frozen solutions are consistent with intramolecular antiferromagnetic exchange interaction between the metal ions. From the data reported it can be concluded that the predominate mechanism for transmitting exchange coupling through the imidazolate bridge is due to a σ exchange pathways.  相似文献   

2.
Multiply charged ions from electrospray ionization (ESI) were observed for ruthenium-bidentate ligand complexes, such as [RuL2B]X2 and [(RuL2)2B]X4, where L is 2,2′-bipyridine, B are tetradentate ligands of 2,2′-bis(2′-pyridyl)bibenzimidazole and 2,6-bis(2′-pyridyl)benzodiimidazole, bidentate ligand of 2-(2′-pyridyl)benzimidazole and related compounds and X is CIO4- or CI-. ESI mass spectra showed a simple mass pattern for easy structural assignment and detecting impurities. The mass spectra for binuclear complexes provide a charge state distribution ranging from 4+ to 2+ for Ru(II)—Ru(II) compounds and 5+ to 2+ for Ru(II)—Rh(III) compounds. It was found that different multiply charged ions are generated by loss of counterions and by protonation/deprotonation at the proton site of ligands B. The abundances of these ions are qualitatively explained in terms of the acidity of metal complexes depending on the bridging ligand structures and the charge of the metal ions. Ions produced by removal of ligands were hardly observed.  相似文献   

3.
In the structure of the novel zinc complex catena‐poly[[diaqua(4‐hydroxybenzohydrazide)zinc(II)]‐μ‐sulfato], [Zn(SO4)(C7H8N2O2)(H2O)2]n, the complex cations are linked by sulfate counter‐ions into helical polymeric chains extending along the b axis. Each helix is stabilized by six intrachain hydrogen bonds involving stronger O—H...O (1.83–2.06 Å) and weaker N—H...O (2.20–2.49 Å) interactions. The ZnII atom displays a distorted octahedral geometry formed by the 4‐hydroxybenzohydrazide ligand, two water molecules and two SO42− ions, which is very similar to the metal‐atom environment in a previously reported CoII complex [Zasłona, Drożdżewski & Kubiak (2010). J. Mol. Struct. 982 , 1–8], especially the Zn—O and Zn—N bond lengths of 2.0453 (12)–2.1602 (9) and 2.1118 (12) Å, respectively.  相似文献   

4.
Sehoon Park 《中国化学》2019,37(10):1057-1071
Transition metal‐catalyzed hydrosilylation is one of the most widely utilized reduction methods as an alternative to hydrogenation in academia and industry. One feature distinct from hydrogenation would be able to install sp3 C—Si bond(s) onto substrates skeleton via hydrosilylation of alkenes. Recently, B(C6F5)3 with hydrosilanes has been demonstrated to be an efficient, metal‐free catalyst system for the consecutive transformation of heteroatom‐containing substrates accompanied by the formation of sp3 C—Si bond(s), which has not been realized thus far under the transition metal‐catalyzed hydrosilylative conditions. In this review, I outline the B(C6F5)3‐mediated consecutive hydrosilylations of heteroarenes containing quinolines, pyridines, and furans, and of conjugated nitriles/imines to provide a new family of compounds having sp3 C—Si bond(s) with high chemo‐, regio‐ and/or stereoselectivities. The silylative cascade conversion of unactivated N‐aryl piperidines to sila‐N‐heterocycles catalyzed by B(C6F5)3 involving consecutive dehydrogenation, hydrosilylation, and intramolecular C(sp2)—H silylation, is presented in another section. Chemical selectivity and mechanism of the boron catalysis focused on the sp3 C—Si bond formation are highlighted.  相似文献   

5.
The model reactions CH3X + (NH—CH=O)M ➔ CH3—NH—NH═O or NH═CH—O—CH3 + MX (M = none, Li, Na, K, Ag, Cu; X = F, Cl, Br) are investigated to demonstrate the feasibility of Marcus theory and the hard and soft acids and bases (HSAB) principle in predicting the reactivity of ambident nucleophiles. The delocalization indices (DI) are defined in the framework of the quantum theory of atoms in molecules (QT-AIM), and are used as the scale of softness in the HSAB principle. To react with the ambident nucleophile NH═CH—O, the carbocation H3C+ from CH3X (F, Cl, Br) is actually a borderline acid according to the DI values of the forming C…N and C…O bonds in the transition states (between 0.25 and 0.49), while the counter ions are divided into three groups according to the DI values of weak interactions involving M (M…X, M…N, and M…O): group I (M = none, and Me4N) basically show zero DI values; group II species (M = Li, Na, and K) have noticeable DI values but the magnitudes are usually less than 0.15; and group III species (M = Ag and Cu(I)) have significant DI values (0.30–0.61). On a relative basis, H3C+ is a soft acid with respect to group I and group II counter ions, and a hard acid with respect to group III counter ions. Therefore, N-regioselectivity is found in the presence of group I and group II counter ions (M = Me4N, Li, Na, K), while O-regioselectivity is observed in the presence of the group III counter ions (M = Ag, and Cu(I)). The hardness of atoms, groups, and molecules is also calculated with new functions that depend on ionization potential (I) and electron affinity (A) and use the atomic charges obtained from localization indices (LI), so that the regioselectivity is explained by the atomic hardness of reactive nitrogen atoms in the transition states according to the maximum hardness principle (MHP). The exact Marcus equation is derived from the simple harmonic potential energy parabola, so that the concepts of activation free energy, intrinsic activation barrier, and reaction energy are completely connected. The required intrinsic activation barriers can be either estimated from ab initio calculations on reactant, transition state, and product of the model reactions, or calculated from identity reactions. The counter ions stabilize the reactant through bridging N- and O-site of reactant of identity reactions, so that the intrinsic barriers for the salts are higher than those for free ambident anions, which is explained by the increased reorganization parameter Δr. The proper application of Marcus theory should quantitatively consider all three terms of Marcus equation, and reliably represent the results with potential energy parabolas for reactants and all products. For the model reactions, both Marcus theory and HSAB principle/MHP principle predict the N-regioselectivity when M = none, Me4N, Li, Na, K, and the O-regioselectivity when M = Ag and Cu(I). © 2019 Wiley Periodicals, Inc.  相似文献   

6.
Abstract

Several polymetallic anionic clusters display a strong tendency to form robust ion-pairs with appropriate counter cations. We have isolated decavanadate cluster [H2V10O28]4– with, hitherto unreported, dimethylammonium ions, [N(CH3)2H2]4+ as counter cations from vanadyl sulfate and N,N’-dimethylformamide (DMF). It is proposed, for the first time, that dimethylammonium ions are formed by reduction and simultaneous decarboxylation of DMF with oxidation of vanadyl sulfate occurring concomitantly under specified reaction conditions. The different colors observed during the course of the reaction are monitored by UV-Visible spectrophotometry. The structure of isolated cluster compound with four counter cations is confirmed by single crystal X-ray diffraction techniques. The compound is also tested for its DNA/BSA binding efficacy and cytotoxicity on human epitheloid cervix carcinoma (HeLa) cells along with vanadates—vanadium pentoxide and ammonium metavanadate. Based on the results, the cluster shows better IC50 and binding values than vanadates.  相似文献   

7.
Summary Hydroxamic acids show a degree of selectivity towards transition metal ions having symmetrical d-electron configuration, e.g. vanadium(V) (d0) and iron(III) (d65). Hydroxamato complexes of metal ions having unsymmetrical d-electron distribution are rare. Thus for manganese(III) (d4) only some thiohydroxamato complexes(1) have been characterised so far. In this communication we report on the first synthesis of a salicylhydroxamato complex of manganese(III). Such investigations are of interest because these higher valent manganese complexes are potentially models for the water-splitting complex present in photosystem II(2).  相似文献   

8.
Thermally Induced Mobility in Crystalline Guanidinium Hexafluorometallates, (C(NH2)3)3MF6 (M = Al, Ga, In) — an in situ E.S.R. Study . The e.s.r. spectroscopic investigation of the title compounds was carried out most successfully by doping them with CrIII ions. The metal ions in the centres of the MF6 octahedra are substituted by the CrIII ions. The zero field splitting parameter | D |, being a measure of the axial distortion of the CrF6 octahedra, has been determined by fitting the calculated to the experimental resonant field values using a computer program. | D | decreases with increasing temperature. This process is completely reversible in the investigated temperature range of ? 190°C to 270°C. It is explained in terms of the thermally induced movability of the N? H … F network.  相似文献   

9.

A procedure was developed for the synthesis of mononuclear first row transition metal coordination compounds with adamantane-1-carboxylate (AdCO2), which allowed the synthesis of complexes NBu4[M(AdCO2)3], where MII = Mn, Ni, Co, Zn. The X-ray diffraction study showed that all AdCO2 act as chelating ligands, which is the main distinguishing structural feature of the synthesized compounds. The formation of four-membered metallocycles leads to a small O—M—O angle (58.4–63.0°), resulting in the distorted trigonal-prismatic environment of the central atom in the coordination anions [M(AdCO2)3]?. Under similar conditions of the synthesis, CuII forms the complex (NBu4)2[Cu(AdCO2)2(SO4)] existing in two differently colored modifications (green and light blue), which is due to a small difference in the environment of CuII in the solid state. Upon heating above 150°C, the light blue modification is transformed into the green modification. Using 1,3-diphenyltriazene (HL), it was demonstrated that the developed synthetic approach is applicable to the preparation of compounds, in which the coordination anions [ML3]? also contain only four-membered chelate metallocycles with the N—M—N angle tightened at the metal atom. The compounds NBu4[ML3], where MII = Co and Ni, were isolated in individual state and were structurally characterized.

  相似文献   

10.
(Extraction of alkali on alkaline earth metal ions with (sym-dibenzo-14-crown-4-oxy)- and (sym-dibenzo-16-crown-5-oxy)-carboxylic acids.)The extraction of lithium, sodium, potassium, calcium and some other metal ions with dibenzo-4-crown-4-oxy- and dibenzo-16-crown-5-oxycarboxylic acids containing the groups -CH2COOH, -(CH2)2COOH, -(CH2)3COOH, -CH(C2H5)COOH and -CH(C4H9)COOH was studied. The extraction increases as a function of the lipophilic character of the carboxylic acid group. Calcium, barium and strontium ions are better extracted than Li+, Na+ and K+; there are only small differences among the alkaline earth metal ions. Evaluated from the extraction data, the composition of the extracted species was 1:1 (metal/ligand) for Li+, and 1:2 for CaCa2+; Na+ and K+ favour the formation of 1:2 complexes with dibenzo-14-crown-4-derivatives bbut 1:1 complexes with dibenzo-16-crown-5-oxy-carboxylic acids. The dependence of the distribution ratio on pH does not provide unequivocal evidence for the composition of the extracted compounds.  相似文献   

11.
A study of the far infrared spectra (700—10 cm?1) of the divalent transition metal HgO complexes has confirmed the earlier conclusion that the compounds with formula MIIS(e)O4 · 2 HgO · (H2O) resemble HgSO4 · 2 HgO closely in structure, whereas a lower HgO content indicates a breaking of the HgO framework and coordination of the M2+-ions to the oxo anions.  相似文献   

12.
Zinc thiocyanate complexes have been found to be biologically active compounds. Zinc is also an essential element for the normal function of most organisms and is the main constituent in a number of metalloenzyme proteins. Pyrimidine and aminopyrimidine derivatives are biologically very important as they are components of nucleic acids. Thiocyanate ions can bridge metal ions by employing both their N and S atoms for coordination. They can play an important role in assembling different coordination structures and yield an interesting variety of one‐, two‐ and three‐dimensional polymeric metal–thiocyanate supramolecular frameworks. The structure of a new zinc thiocyanate–aminopyrimidine organic–inorganic compound, (C6H9ClN3)2[Zn(NCS)4]·2C6H8ClN3·2H2O, is reported. The asymmetric unit consist of half a tetrathiocyanatozinc(II) dianion, an uncoordinated 4‐amino‐5‐chloro‐2,6‐dimethylpyrimidinium cation, a 4‐amino‐5‐chloro‐2,6‐dimethylpyrimidine molecule and a water molecule. The ZnII atom adopts a distorted tetrahedral coordination geometry and is coordinated by four N atoms from the thiocyanate anions. The ZnII atom is located on a special position (twofold axis of symmetry). The pyrimidinium cation and the pyrimidine molecule are not coordinated to the ZnII atom, but are hydrogen bonded to the uncoordinated water molecules and the metal‐coordinated thiocyanate ligands. The pyrimidine molecules and pyrimidinium cations also form base‐pair‐like structures with an R22(8) ring motif via N—H…N hydrogen bonds. The crystal structure is further stabilized by intermolecular N—H…O, O—H…S, N—H…S and O—H…N hydrogen bonds, by intramolecular N—H…Cl and C—H…Cl hydrogen bonds, and also by π–π stacking interactions.  相似文献   

13.
La2O(CN2)2 was synthesized from a 1:1:2 molar reaction mixture of LaCl3, LaOCl, and Li2(CN2) at 650 °C. Well developed single crystals were grown from a LiCl‐KCl flux. The crystal structure was refined as monoclinic (space group C2/c, Z = 2, a = 13.530(2) Å, b = 6.250(1) Å, c = 6.1017(9) Å, β = 104.81(2)°) from single crystal X‐ray diffraction data. The La3+ and (CN2)2— ions in the crystal structure of La2O(CN2)2 can be compared to Fe3+ and S22— ions in the cubic pyrite structure, being arranged like in a distorted NaCl type structure with their centers of gravity. In addition, the O2— ions in La2O(CN2)2 are occupying 1/4 of the tetrahedral voids formed by the arrangement of metal ions.  相似文献   

14.
Liquid xenon difluoride at 140°C does not react with zirconium or hafnium tetrafluorides, neither does liquid xenon hexafluoride at 60°C. Therefore reactions between the corresponding hydrazinium fluorometalates or ammonium fluorometalates and xenon difluoride and xenon hexafluoride, respectively, were carried out. N2H6ZrF6 and N2H6HfF6 react with xenon difluoride at 60°C again yielding only the corresponding tetrafluorides, while the analogous reaction with (NH4)2ZrF6 and (NH4)2HfF6 proceeds at 170°C yielding the corresponding ammonium pentafluorometalates, which are stable and do not react further with excessive xenon difluoride up to 200°C.The reaction between N2H6ZrF6 or N2H6HfF6 and xenon hexafluoride proceeds at room temperature yielding a series of thermally unstable compounds of the type mXeF6.MF4 (M = Zr, Hf) where m ? 6. The final products which are stable at room temperature are XeF6.MF4 (M = Zr,Hf). Spectroscopic evidence suggests that these compounds are salts of a XeF+5 cation squashed between a polymeric anion of the type (MF5)x-x.  相似文献   

15.
In both title compounds, C18H24N2O2, (Ia), and C18H26N2O22+·2ClO4, (II), respectively, the two aryl rings are strictly parallel, with an inversion centre lying at the mid‐point of each central CH2—CH2 bond. Molecules in (Ia) are linked into two‐dimensional layers by N—H...O hydrogen bonds. The component ions in (II) are joined together by a combination of N/O/C—H...O hydrogen bonds and C—H...π and anion...π interactions, forming a three‐dimensional network. A structural understanding of (Ia) and (II) may provide some useful information about how and why their metal–organic complexes display various biological activities and function in catalytic processes.  相似文献   

16.
Three new thiostannates [M(en)3]2Sn2S6 (en = ethylenediamine, M = Mn( 1 ), Co( 2 ) and Zn( 3 )) were synthesized by solvothermal method. The crystals were grown up in a Teflon‐lined steel autoclave at temperature about 180 °C. All the three compounds consist of discrete [Sn2S6]4— anions, which are dimer of two tetrahedral SnS4 sharing a common edge. The transition metal cations are six‐coordinated by three ethylenediamine molecules forming octahedral complex ions. Although the synthetic procedures, the mole ratio of the reactants and the solvent are essentially the same, the compound of MnII is quite different in structure from that of compounds of CoII and ZnII. Compound 1 crystallizes in monoclinic crystal system, C2/c, whereas compounds 2 and 3 crystallize in the orthorhombic crystal system, Pbca. Unlike compound 1 , the [M(en)3]2+ cations in 2 and 3 are disordered. The difference of molecular packing between 1 and 2 ‐ 3 is considered due to the influence of the entities of the metal ions, such as radii and the coordination properties. The thermal chemical behaviors of the compounds 1 ‐ 3 were discussed and the results are also related to the property of the metal ions.  相似文献   

17.
The earth‐metal olefin complex [Ga I (COD)2]+[Al(ORF)4]? (COD=1,5‐cyclooctadiene; RF=C(CF3)3) constitutes the first homoleptic olefin complex of any main‐group metal accessible as a bulk compound. It is straight forward to prepare in good yield and constitutes an olefin complex of a main‐group metal that—similar to many transition‐metals—may adopt the +1 and +3 oxidation states opening potential applications. Crystallographic‐, vibrational‐ and computational investigations give an insight to the atypical bonding between an olefin and a main‐group metal. They are compared to classical transition‐metal relatives.  相似文献   

18.
The isomorphous title compounds, [Ni{(NH2)2CO}4(H2O)2](NO3)2 and [Co{(NH2)2CO}4(H2O)2](NO3)2, feature discrete centrosymmetric cations in octahedral coordinations, formed by four urea molecules linked via their O atoms to the central ion in equatorial positions and two water molecules in trans positions. The complexes are packed in a pseudo‐hexagonal manner separated by the nitrate counter‐ions. All H atoms are involved in moderate hydrogen bonds of four types: N—H...O=C, N—H...O—N, O—H...O—N and N—H...O—H. Graph‐set analysis was applied to distinguish the hydrogen‐bond patterns at unitary and higher level graph sets.  相似文献   

19.
Calcium and Strontium amide are ionic compounds crystallising in a tetragonally distorted anatase structure-type at ambient temperatures. The amide ions (NH2/ND2) resemble water molecules in structure and in charge distribution. By means of temperature dependent neutron diffraction investigations weak super-structure reflections were observed at temperatures below 90 K (Ca(ND2)2) and 60 K (Sr(ND2)2), respectively, indicating the existence of a so far unknown low-temperature (LT) phase. Using high resolution neutron powder diffraction at temperatures below 10 K the structure was determined for both compounds. The LT-phases are isotypic and crystallise monoclinic in the space group P21/c with four formula units within the unit cell: Ca(ND2)2 at 10 K a = 7.257(2) Å, b = 7.2434(2) Å, c = 6.300(1) Å, β = 124.73(1)° Sr(ND2)2 at 5 K a = 7.6950(1) Å, b = 7.68374(9) Å, c = 6.6324(3) Å, β = 124.917(2)°. Their structure is closely related to the tetragonal HT-phase, but an ordering of the amide ions occurs due to freezing of a lattice mode which is dominated by the librational motion of the amide ions in the {1 0 0} planes of the HT-phase.  相似文献   

20.
The crystal structure of aripiprazole nitrate (systematic name: 4‐(2,3‐dichlorophenyl)‐1‐{4‐[(2‐oxo‐1,2,3,4‐tetrahydroquinolin‐7‐yl)oxy]butyl}piperazin‐1‐ium nitrate), C23H28Cl2N3O2+·NO3 or AripH+·NO3, is presented and the molecule compared with the aripiprazole molecules reported so far in the literature. Bond distances and angles appear very similar, except for a slight lengthening of the C—NH distances involving the protonated N atom, and the main differences are to be found in the molecular spatial arrangement (revealed by the sequence of torsion angles) and the intermolecular interactions (resulting from structural elements specific to this structure, viz. the nitrate counter‐ions on one hand and the extra protons on the other hand as hydrogen‐bond acceptors and donors, respectively). The result is the formation of [100] strips, laterally linked by weak π–π and C—Cl...π interactions, leading to a family of undulating sheets parallel to (010).  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