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 375 毫秒
1.
Salen type complexes, CuL, the corresponding tetrahydrosalen type complexes, Cu[H4]L, and N,N′-dimethylated tetrahydrosalen type complexes, Cu[H2Me2]L, were investigated using cyclic voltammetry, and electronic and ESR spectroscopy. In addition, the analogous copper(II) complexes with a derivative of the tetradentate ligand ‘salphen’ [salphen=H2salphen=N,N′-disalicylidene-1,2-diaminobenzene] were studied. Solutions of CuL, Cu[H4]L and Cu[H2Me2]L are air-stable at ambient temperature, except for the complex Cu(tBu, Me)[H4]salphen [H2(tBu, Me)[H4]salphen=N,N′-bis(2-hydroxy-3-tert-butyl-5-methylbenzyl)-1,2-diaminobenzene]. Cu(tBu, Me)[H4]salphen interacts with dioxygen and the ligand is oxidatively dehydrogenated (–CH2–NH–→–C=N–) to form Cu(tBu, Me)[H2]salphen and finally, in the presence of base, Cu(tBu, Me)salphen. X-ray structure analysis of Cu(tBu, Me)[H2Me2]salen confirms a slightly tetrahedrally distorted planar geometry of the CuN2O2 coordination core. The complexes were subjected to spectrophotometric titration with pyridine, to determine the equilibrium constants for adduct formation. It was found that the metal center in the complexes studied is only of weak Lewis acidity. In dichlormethane, the oxidation Cu(II)/Cu(III) is quasireversible for the CuL type complexes, but irreversible for the Cu[H4]L and Cu[H2Me2]L type. A poorly defined wave was observed for the irreversible reduction Cu(II)/Cu(I) at potentials less than −1.0 V. The ESR spectra of CuL at both 77 K and room temperature reveal that very well resolved lines can be attributed to the interaction of an unpaired electron spin with the copper nuclear spin, 14N donor nuclei and to a distant interaction with two equivalent protons [ACu(iso)≈253 MHz, AN(iso)≈43 MHz, AN(iso)≈20 MHz]. These protons are attached to the carbon atoms adjacent to the 14N nuclei. In contrast to CuL, the number of lines in the spectra of the complexes Cu[H4]L and Cu[H2Me2]L is greatly reduced. At room temperature, only a quintet with a considerably smaller nitrogen shf splitting constant [AN(iso)≈27 MHz] is observed. Both factors, planarity and conjugation, are thus essential for the observation of distant hydrogen shf splitting in CuL. Due to the C=N bond hydrogenation, the coordination polyhedra of the complexes Cu[H4]L and Cu[H2Me2]L is more flexible and more sensitive to ligand modification than that of CuL. The electron-withdrawing effect of the phenyl ring of the phenylenediamine bridge is reflected in a reduction of the copper hyperfine coupling constants in Cu(tBu, Me)[H4]salphen and Cu(tBu, Me)[H2Me2]salphen complexes [ACu(iso)≈215 MHz].  相似文献   

2.
Trigonal copper sulfide nanoparticles were synthesized from symmetrical [(Bu)2NC(S)NC(O)C6H3(3,5-NO2)2]2Cu(II) and [(Bu)2NC(S)NC(O)C6H4(4-NO2)]2Cu(II) complexes by thermolysis in the presence of surfactant oleylamine. The symmetrical copper complexes were synthesized by reaction of copper(II) acetate with N-(3,5-dinitrobenzoyl)-N′,N′-dibutylthiourea and N-(4-nitrobenzoyl)-N′,N′-dibutylthiourea. The symmetrical copper complexes were characterized by FT-IR spectroscopy, elemental analysis, and mass spectrometry (MS-APCI). The single-crystal X-ray structure of [(Bu)2NC(S)NC(O)C6H4(4-NO2)]2Cu(II) has been determined from single-crystal X-ray diffraction data. These metal complexes have been used as single source precursors for the preparation of copper sulfide nanoparticles. The deposited copper sulfide nanoparticles were characterized by X-ray powder diffraction and transmission electron microscopy.  相似文献   

3.
Two new volatile complexes of Cu(II) containing methoxy groups were studied in a combined X-ray diffraction investigation of the mono- (KUMA automatic diffractometer, MoKα radiation) and polycrystals (DRON-3M, CuKα radiation). The structures are molecular and consist of the trans-complexes of bis-(5-methyl-5-butoxy-hexanedionato-2,4)copper(II), C22H38CuO6, and bis-(2,2-dimethyl-5-cyclohexyl-pentanedionato-3,5) Copper(II), C28H46CuO6. The crystal data of C22H38CuO6: a = 8.939(1), b = 8.887(2), c = 8.326(1) å, α = 107.92(2), Β = 108.15(1), γ = 85.52(2)?, space group P1,V= 597.9(2) å3, Z = 1, dx = 1.283 g/cm3. The crystal data of C28H46CuO6: a = 18.625(3), b = 16.126(2), c= 13.613(3) å, Β = 132.19?, space group P21/n, V= 3029.3(9) å3, Z = 4, dx= 1.189 g/cm3. In both cases, the square planar environment of the Cu atom with Cu...O distances of 1.90 å is completed by interactions with two carbon atoms of the terminal groups of the two neighboring molecules at Cu...C distances of 3.66 å (average).  相似文献   

4.
A series of solvent-free heteroleptic terminal rare-earth-metal alkyl complexes stabilized by a superbulky tris(pyrazolyl)borato ligand with the general formula [TptBu,MeLnMeR] have been synthesized and fully characterized. Treatment of the heterobimetallic mixed methyl/tetramethylaluminate compounds [TptBu,MeLnMe(AlMe4)] (Ln=Y, Lu) with two equivalents of the mild halogenido transfer reagents SiMe3X (X=Cl, I) gave [TptBu,MeLnX2] in high yields. The addition of only one equivalent of SiMe3Cl to [TptBu,MeLuMe(AlMe4)] selectively afforded the desired mixed methyl/chloride complex [TptBu,MeLuMeCl]. Further reactivity studies of [TptBu,MeLuMeCl] with LiR or KR (R=CH2Ph, CH2SiMe3) through salt metathesis led to the monomeric mixed-alkyl derivatives [TptBu,MeLuMe(CH2SiMe3)] and [TptBu,MeLuMe(CH2Ph)], respectively, in good yields. The SiMe4 elimination protocols were also applicable when using SiMe3X featuring more weakly coordinating moieties (here X=OTf, NTf2). X-ray structure analyses of this diverse set of new [TptBu,MeLnMeR/X] compounds were performed to reveal any electronic and steric effects of the varying monoanionic ligands R and X, including exact cone-angle calculations of the tridentate tris(pyrazolyl)borato ligand. Deeper insights into the reactivity of these potential precursors for terminal alkylidene rare-earth-metal complexes were gained through NMR spectroscopic studies.  相似文献   

5.
Bis(hinokitiolato)copper(II), Cu(hino)2, exhibits both antibacterial and antiviral properties, and has been previously shown to exist in two modifications. A third modification has now been confirmed, namely tetrakis(μ2‐3‐isopropyl‐7‐oxocyclohepta‐1,3,5‐trien‐1‐olato)bis(3‐isopropyl‐7‐oxocyclohepta‐1,3,5‐trien‐1‐olato)tricopper(II)–bis(μ2‐3‐isopropyl‐7‐oxocyclohepta‐1,3,5‐trien‐1‐olato)bis[(3‐isopropyl‐7‐oxocyclohepta‐1,3,5‐trien‐1‐olato)copper(II)] (1/1), [Cu(C10H11O2)2]3·[Cu(C10H11O2)2]2, where 3‐isopropyl‐7‐oxocyclohepta‐1,3,5‐trien‐1‐olate is the systematic name for the hinokitiolate anion. This new modification is composed of discrete [cis‐Cu(hino)2]2[trans‐Cu(hino)2] trimers and [cis‐Cu(hino)2]2 dimers. The Cu atoms are bridged by μ2‐O atoms from the hinokitiolate ligands to give distorted square‐pyramidal and distorted octahedral CuII coordination environments. Hence, the CuII environments are CuO5/CuO6/CuO5 for the trimer and CuO5/CuO5 for the dimer. Each trimer and dimer has crystallographically imposed inversion symmetry. The trimer has never been observed before, the dimer has been seen only once before, and the combination of the two together in the same lattice is unprecedented. The CuO5 cores exhibit four strong basal Cu—O bonds [1.915 (2)–1.931 (2) Å] and one weak apical Cu—O bond [2.652 (2)–2.658 (2) Å]. The CuO6 core exhibits four strong equatorial Cu—O bonds [1.922 (2)–1.929 (2) Å] and two very weak axial Cu—O bonds [2.911 (3) Å]. The bite angles for the chelating hinokitiolate ligands range from 83.13 (11) to 83.90 (10)°.  相似文献   

6.
The two hypersilylcuprates LiCu2Hyp3 ( 2 ) and [Li7(OtBu)6][Cu2Hyp3] ( 3 ) (Hyp = Si(SiMe3)3) were synthesized by reactions of unsolvated lithium hypersilanide, LiHyp with hypersilylcopper and CuOtBu, respectively. Both contain the novel A‐frame trihypersilyldicuprate anion [Cu2Hyp3]. In the former case a molecular compound is produced containing intimate ion pairs. In the latter case the cuprate anion and the unique large [Li7(OtBu)6]+ cation form a salt‐like compound, only sparingly soluble in unpolar solvents. According to NBO analyses the bonding within the trihypersilyldicuprate moiety is best described by interaction of a bridging lewis‐basic hypersilanide anion with two lewis‐acidic hypersilyl copper fragments.  相似文献   

7.
α-Methoxypolyethylene oxide methacrylate was polymerized by copper(I)-mediated living radical polymerization in aqueous solution to give polymers with controlled number-average molecular masses and narrow polydispersities. When equimolar quantities of initiator with respect to copper(I) bromide were used, the reaction was extremely fast with quantitative conversion achieved in less than 5 min at ambient temperature. However, the molecular weight distribution was broad, and control over the number-average molecular weight (Mn) growth was extremely poor; this is ascribed to an increase in termination because of the increased rate as a result of the coordination of water at the copper center. The complex formed between copper(I) bromide and N-(n-propyl)-2-pyridylmethanimine, bis[N-(n-propyl)-2-pyridylmethanimine]copper(I), was demonstrated to be stable in aqueous solution by 1H NMR over 10 h at 25 °C. However, on increasing the temperature to 50 °C, decomposition occurred rapidly. Thus, polymerization temperatures were maintained at ambient temperature. When longer alkyl chains were utilized in the ligand, that is, pentyl and octyl, the complex acted as a surfactant leading to heterogeneous solutions. When the catalyst concentration was reduced by two orders of magnitude, the rate of polymerization was reduced with 100% conversion achieved after 60 min with the Mn of the final product being higher than that predicted and the polydispersity equal to 1.43. Copper(II) was added as an inhibitor to circumvent these problems. When 10% of Cu(I) was replaced by Cu(II) {[Cu(I)] + [Cu(II)]/[I] = 1/100}, the mass distribution showed a bimodal distribution, and the rate of polymerization decreased significantly. With a catalyst composition [Cu(I)]/[Cu(II)] = 0.5/0.5 {[Cu(I)] + [Cu(II)]}/[I] = 1/100, polymerization proceeded slowly with 80% conversion reached after 22 h. Thus, the concentration of Cu(I) was further reduced with [Cu(I)]/[Cu(II)] = 10/90, {[Cu(I)] + [Cu(II)]}/[I] = 1/100. The system then contained [Initiator]/[Cu(I)] = 1000/1 and [I]/[Cu(II)] = 1000/9. Under these conditions, the reaction reached 50% after 5 h with the polymer having both an Mn close to the theoretical value and a narrow polydispersity of PDi = 1.15. Optimum results were obtained by increasing the amount of catalyst. When a ratio of [Cu(I)]/[Cu(II)] = 10/90 with a ratio of [Cu]/[I] = 1/1, a conversion of 100% was achieved after less than 20 h, leading to a product having Mn = 8500 and PDi = 1.15. Decreasing the amount of Cu(II) relative to Cu(I) to [Cu(I)]/[Cu(II)] = 0.5/0.5 (maintaining the overall amount of copper) led to 100% conversion after 75 min: Mn = 9500, PDi = 1.10. Block copolymers have been demonstrated by sequential monomer addition with excellent control over Mn and PDi. © 2001 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 39: 1696–1707, 2001  相似文献   

8.
The complexes [Cu(dpp)Br2] ( 1 ) and [Cu(dpp)2][CuBr2] ( 2 ) (dpp = 2,9‐diphenyl‐1,10‐phenanthroline) were synthesized and characterized by single‐crystal X‐ray diffraction methods. Reaction of copper(II) bromide with the dpp ligand in dichloromethane at room temperature afforded 1 , which is a rare example of non‐square planar four‐coordinate copper(II) complexes. Complex 1 crystallizes in the monoclinic space group C2/c with a = 15.352(3), b = 13.192(3), c = 11.358(2) Å, β = 120.61(3)°, V = 1979.6(7) Å3, Z = 4, Dcalc = 1.865 g cm?3. The coordination geometry about the copper center is distorted about halfway between square planar and tetrahedral. The Cu‐N distance is 2.032(2) Å and the Cu‐Br distance 2.3521(5) Å. Heating a CH2Cl2 or acetone solution of 1 resulted in complex 2 , which consists of a slightly distorted tetrahedral [Cu(dpp)2]+ cation and a linear two‐coordinate [CuBr2]? anion. 2 crystallizes in the triclinic space group with a = 10.445(2), b = 11.009(2), c = 18.458(4) Å, α = 104.72(3), β = 94.71(3), γ = 103.50(3)°, V = 1973.3(7) Å3, Z = 2, Dcalc = 1.602 g cm?3. The four Cu(1)‐N distances are between 2.042(3) and 2.067(3) Å, the distance of Cu(2)‐Br(1) 2.2268(8) Å, and the disordered Cu(3)‐Br(2) distances are 2.139(7) and 2.237(4) Å, respectively. Complex 2 could also be prepared by directly reacting CuBr with dpp in CH2Cl2.  相似文献   

9.
New monoanionic CNC pincer ligands, [N{SiMe2CH2(RIm)}2] (R = tBu, iPr, Ph) featuring three different N-heterocyclic carbenes and a disilylamido moiety is reported. Treatment of the lithium salt of [N{SiMe2CH2(RIm)}2] with CuIOTf afforded the corresponding copper complexes [N{SiMe2CH2(RIm)}2]Cu in 41–56 % yield. X-ray crystal structures of [N{SiMe2CH2(RIm)}2]Cu show that they are monomeric and feature three-coordinate, pseudo T-shaped copper(I) sites. The X-ray crystal structure of one of the precursor lithium complexes, [N{SiMe2CH2(tBuIm)}2]Li is also presented.  相似文献   

10.
The synthesis and structures of the two CuI halide complexes [Cu5(dppm)(dppm?)2(OtBu)Cl2] and [Cu3(dppm)3Br2][CuBr2] (dppm = Ph2PCH2PPh2, dppm? = [Ph2PCHPPh2]?) are reported. The compounds were obtained by treating reaction mixtures of [CuOtBu] and dppm with dichloromethane or dibromomethane.  相似文献   

11.
The mononuclear N‐heterocyclic carbene (NHC) copper alkoxide complexes [(6‐NHC)CuOtBu] (6‐NHC=6‐MesDAC ( 1 ), 6‐Mes ( 2 )) have been prepared by addition of the free carbenes to the tetrameric tert‐butoxide precursor [Cu(OtBu)]4, or by protonolysis of [(6‐NHC)CuMes] (6‐NHC=6‐MesDAC ( 3 ), 6‐Mes ( 4 )) with tBuOH. In contrast to the relatively stable diaminocarbene complex 2 , the diamidocarbene derivative 1 proved susceptible to both thermal and hydrolytic ring‐opening reactions, the latter affording [(6‐MesDAC)Cu(OC(O)CMe2C(O)N(H)Mes)(CNMes)] ( 6 ). The intermediacy of [(6‐MesDAC)Cu(OH)] in this reaction was supported by the generation of Cu2O as an additional product. Attempts to generate an isolable copper hydride complex of the type [(6‐MesDAC)CuH] by reaction of 1 with Et3SiH resulted instead in migratory insertion to generate [(6‐MesDAC‐H)Cu(P(p‐tolyl)3)] ( 9 ) upon trapping by P(p‐tolyl)3. Migratory insertion was also observed during attempts to prepare [(6‐Mes)CuH], with [(6‐Mes‐H)Cu(6‐Mes)] ( 10 ) isolated, following a reaction that was significantly slower than in the 6‐MesDAC case. The longer lifetime of [(6‐Mes)CuH] allowed it to be trapped stoichiometrically by alkyne, and also employed in the catalytic semi‐reduction of alkynes and hydrosilylation of ketones.  相似文献   

12.
《Polyhedron》1999,18(8-9):1163-1169
The coordination of Cu(II) to the Keggin type anions α-undecatungstophosphocuprate(II) and α-undecatungstoborocuprate(II) was investigated in different environments by EPR and electronic spectroscopy. This study has shown that the coordination geometry around Cu(II) in the tetrabutylammonium (TBA) salts, (TBA)4Hx[XW11CuO39], with X=P or B, is square pyramidal, with copper bound to the five oxygen donor atoms of the polyoxometalate, whereas for the [XW11Cu(H2O)O39]n anion, on the corresponding potassium salt, a tetragonally elongated pseudo-octahedral geometry was found. For the potassium salts, in aqueous solution, six-coordinated copper is the only form found. For the TBA salts, in nonaqueous solvents, we can observe either the presence of only one form (the six-coordinated Cu(II) species, with a solvent molecule bound to copper), or of two forms: the solvent coordinated copper anions and the five-coordinated copper [XW11CuO39]n anions.  相似文献   

13.
New Amido and Imido Bridged Complexes of Copper – Syntheses and Structures of [{Li(OEt2)}2][Cu(NPh2)3], [ClCuN(SnMe3)3], [{CuN(SnMe3)2}4], [Cu16(NH2tBu)12Cl16], [{CuNHtBu}8], [Li(dme)3][Cu6(NHMes)3(NMes)2], [PPh3(C6H4)CuNHMes], [{[Li(dme)][Cu(NHMes)(NHPh)]}2], and [{Li(dme)3}3][Li(dme)2][Cu12(NPh)8] The reactions of stannylated and lithiated amines with coppersalts (halogenides, thiocyanates) lead to amido and imido bridged complexes which contain one to twelve metal atoms. [{Li(OEt2)}2][Cu(NPh2)3] ( 1 ) results from the reaction of CuCl with LiNPh2 in the presence of trimethylphosphine. With N(SnMe3)3, CuCl reacts to the donor‐acceptor complex [ClCuN(SnMe3)3] ( 2 ) that is transformed into the tetrameric complex [{CuN(SnMe3)2}4] ( 3 ) by thermolysis. 3 can also be obtained by the reaction of LiN(SnMe3)2 with Cu(SCN)2. While terminally bound in 1 , the amido ligand is μ2‐bridging between copper atoms in compound 3 . The influence of the alkyl amide's leaving group can be seen from a comparison of the reactivity of Me3SnNHtBu and LiNHtBu, respectively. With Me3SnNHtBu, CuCl2 forms the polymeric compound [Cu16(NH2tBu)12Cl16] ( 4 ) whereas in the case of LiNHtBu with both CuCl and CuSCN, the complex [{CuNHtBu}8] ( 5 ) is obtained. The latter contains two planar Cu4N4‐rings similar to those in 3 . If a mesityl group is introduced at the lithium amide, different products are accessible. Both, CuBr and CuSCN, lead to the formation of [Li(dme)3][Cu6(NHMes)3(NMes)2] ( 6 ) whose anion consists of a prismatic copper core with μ2‐bridging amido and μ3‐bridging imido ligands. In the presence of PPh4Cl, a mixture of Cu(SCN)2 and LiNHMes enables an ortho‐metallation reaction that produces [PPh3(C6H4)CuNHMes] ( 7 ). From the reaction of CuSCN with LiNHMes and LiNHPh either the dimeric complex [{[Li(dme)][Cu(NHMes)(NHPh)]}2] ( 8 ) or the cluster [{Li(dme)3}3][Li(dme)2][Cu12(NPh)8] ( 9 ) results. The anion in 9 exhibits a cubo‐octahedron of copper atoms μ3‐bridged by (NPh)2–‐ligands. The solid state structures of compounds 1 – 9 have been determined by single crystal X‐ray diffraction.  相似文献   

14.
A new copper carboxylate polymer with cyanoacetate anion as a ligand was synthesized and studied using X-ray diffraction, IR, and EPR spectroscopy. The crystal is tetragonal: a= 14.702(2) Å, c= 13.470(3) Å, Z= 8, space group I41/a, and R= 0.0634. The copper atoms in the centrosymmetric dimeric fragment have a square-pyramidal surrounding with the CuO4N coordination core and are joined through four bidentate bridging anions of cyanoacetic acid Cu(1)"–O(1A) 1.931(4) Å, Cu(1)"–O(1B) 1.926(4) Å, Cu(1)–O(2B) 2.018(3) Å, Cu(1)–O(2A) 2.036(4) Å, and Cu(1)–N(1A)" 2.206(5) Å). The Cu···Cu" distance in the dimer is 2.709 Å. The copper atom is extended from the mean equatorial plane toward the axial nitrogen atom by 0.23 Å. EPR data confirm strong antiferromagnetic interaction (2J –275 cm–1) between the copper(II) ions of the dimeric fragment, whereas the interaction between the dimers is significantly weaker (J< 0.3 cm–1).  相似文献   

15.
We develop a simple semiempirical model that correlates the Auger parameter to the ground state valence charge of the core-ionized atom with closed shell electron configuration. Until now, the Auger parameter was employed to separate initial and final state effects that influence the core electron binding energy. The model is applied to Cu(I) and Cu (II) compounds with the Auger parameter defined as α' = EbFL (2p3/2) + EkFL (L3M45M45;1G). The Auger parameter shift for Cu(I) ion in CuI, CuBr, CuFeS2, Cu2S, and Cu2O compounds—with respect to the copper free atom—increases with the electronic polarizability of the nearest-neighbour ligands suggesting a nonlocal screening mechanism. This relaxation process is interpreted as due to an electron transfer from the nearest-neighbour ligands toward the spatially extended 4sp valence orbitals of the core-ionized Cu(I) ion. In agreement with our model, a linear relationship is found between the Auger parameter shift and the ground state Bader valence charge obtained by density functional theory calculations. The Auger parameter shift for the Cu (II) ion in CuF2, CuCl2, CuBr2, CuSO4, Cu (NO3)2•3H2O, Cu3(PO4)2, Cu (OH)2, and CuO compounds is very close to the Auger parameter of metallic copper, and therefore, it is not related to the calculated ground state Bader valence charge. The relaxation process in the final state is dominated by the local screening mechanism, which involves an electron transfer from the nearest-neighbour ligands toward the spatially contracted 3d orbitals of the core-ionized Cu (II) ion.  相似文献   

16.
A copper(II) complex based on a V-shaped ligand, 2,6-bis(2-benzimidazolyl)pyridine (bbp), has been synthesized and characterized by elemental analysis, molecular conductivity, 1H NMR, IR, UV-Vis spectra, and X-ray single-crystal diffraction. The crystal structure of [Cu(bbp)2](pic)2?·?2DMF (pic?=?picrate) shows copper is six-coordinate forming a distorted octahedron. The interaction between Cu(II) complex and DNA was investigated by spectrophotometric methods and viscosity measurement. The experimental results suggest that the Cu(II) complex binds to DNA via intercalation. Antioxidant assay in vitro also shows that the Cu(II) complex possesses significant antioxidant activities.  相似文献   

17.
The synthesis, structure, substitution chemistry, and optical properties of the gold‐centered cubic monocationic cluster [Au@Ag8@Au6(C≡CtBu)12]+ are reported. The metal framework of this cluster can be described as a fragment of a body‐centered cubic (bcc) lattice with the silver and gold atoms occupying the vertices and the body center of the cube, respectively. The incorporation of alkali metal atoms gave rise to [MnAg8?nAu7(C≡CtBu)12]+ clusters (n=1 for M=Na, K, Rb, Cs and n=2 for M=K, Rb), with the alkali metal ion(s) presumably occupying the vertex site(s), whereas the incorporation of copper atoms produced [CunAg8Au7?n(C≡CtBu)12]+ clusters (n=1–6), with the Cu atom(s) presumably occupying the capping site(s). The parent cluster exhibited strong emission in the near‐IR region (λmax=818 nm) with a quantum yield of 2 % upon excitation at λ=482 nm. Its photoluminescence was quenched upon substitution with a Na+ ion. DFT calculations confirmed the superatom characteristics of the title compound and the sodium‐substituted derivatives.  相似文献   

18.
A new family of 14‐electron, four‐coordinate iron(II) complexes of the general formula [TptBu,MeFeX] (TptBu,Me is the sterically hindered hydrotris(3‐tert‐butyl‐5‐methyl‐pyrazolyl) borate ligand and X=Cl ( 1 ), Br, I) were synthesized by salt metathesis of FeX2 with TptBu,MeK. The related fluoride complex was prepared by reaction of 1 with AgBF4. Chloride 1 proved to be a good precursor for ligand substitution reactions, generating a series of four‐coordinate iron(II) complexes with carbon, oxygen, and sulphur ligands. All of these complexes were fully characterized by conventional spectroscopic methods and most were characterized by single‐crystal X‐ray crystallographic analysis. Magnetic measurements for all complexes agreed with a high‐spin (d6, S=2) electronic configuration. The halide series enabled the estimation of the covalent radius of iron in these complexes as 1.24 Å.  相似文献   

19.
The treatment of N,C,N‐chelated antimony(III) and bismuth(III) chlorides [C6H3‐2,6‐(CH=NR)2]MCl2 [R = tBu and M = Sb ( 1 ) or Bi ( 2 ); R = Dmp and M = Sb ( 3 ) or Bi ( 4 )] (Dmp = 2,6‐Me2C6H3) with one molar equivalent of Ag[CB11H12] led to a smooth formation of corresponding ionic pairs {[C6H3‐2,6‐(CH=NR)2]MCl}+[CB11H12] [R = tBu and M = Sb ( 7 ) or Bi ( 8 ), R = Dmp and M = Sb ( 9 ) or Bi ( 10 )]. Similarly, the reaction of C,N‐chelated analogues [C6H2‐2‐(CH=NDip)‐4,6‐(tBu)2]MCl2 [M = Sb ( 5 ) or Bi ( 6 ), Dip = 2′,6′‐iPr2C6H3] gave compounds {[C6H2‐2‐(CH=NDip)‐4,6‐(tBu)2]MCl}+[CB11H12] [M = Sb ( 11 ) or Bi ( 12 )]. All compounds 7 – 12 were characterized with 1H, 11B and 13C{1H} NMR spectroscopy, ESI‐mass spectrometry, IR spectroscopy, and molecular structures of 7 – 9 and 12 were determined by the help of single‐crystal X‐ray diffraction analysis. In contrast, all attempts to cleave also the second M–Cl bond in 7 – 12 using another molar equivalent Ag[CB11H12] remained unsuccessful. Nevertheless, the reaction between 7 (or 8 ) and Ag[CB11H12] produced unprecedented adducts of both reagents namely {[C6H3‐2,6‐(CH=NtBu)2]SbCl}22+[Ag2(CB11H12)4]2– ( 13 ) and {[C6H3‐2,6‐(CH=NtBu)2]BiCl}+[Ag(CB11H12)2] ( 14 ) in a reproducible manner. The molecular structures of these sparingly soluble compounds were determined by single‐crystal X‐ray diffraction analysis.  相似文献   

20.
New adducts of ethylenediamine (en), N,N-dimethylethylenediamine (ndmen) and N,N′-dimethylethylenediamine (dmen) with squarate as counter-ions were synthesized and characterized by physico-chemical methods (IR and UV/vis spectroscopy, magnetic susceptibility and thermoanalytical measurements). The crystal structure of tris(ethylenediamine)cobalt(III) 1.5 squarate hexahydrate, [Co(en)3](sq)1.5 · 6H2O, was determined by single crystal X-ray diffraction. Co(III), Ni(II) and Cu(II) ions in the monomeric octahedral tris(ethylenediamine)cobalt(III) 1.5 squarate hexahydrate (1), tris(ethylenediamine)nickel(II) squarate 0.5 hydrate (2) and diaquabis(ethylenediamine)copper(II) squarate dihydrate (3) are chelated by ethylenediamines through two amine nitrogen atoms. Cu(II) atoms in the diaquabis(ndmen)copper(II) squarate (4) and diaquabis(dmen)copper(II) squarate (5) monomeric octahedral complexes are coordinated by ndmen and dmen molecules through two amine nitrogen atoms in a bidentate chelating manner. Water molecules complete the octahedral coordination. The orange (1), violet (4) and violet (5) complexes upon heating transform to claret, green and green species on dehydration, respectively, which revert immediately after cooling in the open atmosphere. The violet (3) complex upon heating loses water molecules yielding a deep blue dehydrated species, which on further heating undergoes an exothermic phase transition accompanied by thermochromism, deep blue to brown in the solid state. The decomposition mechanism and thermal stability of the solid complexes are interpreted in terms of their structures. The final decomposition products – the respective metal oxides – were identified by IR spectroscopy.  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