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.
A series of new germylene compounds has been synthesized offering systematic variation in the σ‐ and π‐capabilities of the α‐substituent and differing levels of reactivity towards E?H bond activation (E=H, B, C, N, Si, Ge). Chloride metathesis utilizing [(terphenyl)GeCl] proves to be an effective synthetic route to complexes of the type [(terphenyl)Ge(ERn)] ( 1 – 6 : ERn=NHDipp, CH(SiMe3)2, P(SiMe3)2, Si(SiMe3)3 or B(NDippCH)2; terphenyl=C6H3Mes2‐2,6=ArMes or C6H3Dipp2‐2,6=ArDipp; Dipp=C6H3iPr2‐2,6, Mes=C6H2Me3‐2,4,6), while the related complex [{(Me3Si)2N}Ge{B(NDippCH)2}] ( 8 ) can be accessed by an amide/boryl exchange route. Metrical parameters have been probed by X‐ray crystallography, and are consistent with widening angles at the metal centre as more bulky and/or more electropositive substituents are employed. Thus, the widest germylene units (θ>110°) are found to be associated with strongly σ‐donating boryl or silyl ancillary donors. HOMO–LUMO gaps for the new germylene complexes have been appraised by DFT calculations. The aryl(boryl)‐germylene system [ArMesGe{B(NDippCH)2}] ( 6 ‐Mes), which features a wide C‐Ge‐B angle (110.4(1)°) and (albeit relatively weak) ancillary π‐acceptor capabilities, has the smallest HOMO–LUMO gap (119 kJ mol?1). These features result in 6 ‐Mes being remarkably reactive, undergoing facile intramolecular C?H activation involving one of the mesityl ortho‐methyl groups. The related aryl(silyl)‐germylene system, [ArMesGe{Si(SiMe3)3}] ( 5 ‐Mes) has a marginally wider HOMO–LUMO gap (134 kJ mol?1), rendering it less labile towards decomposition, yet reactive enough to oxidatively cleave H2 and NH3 to give the corresponding dihydride and (amido)hydride. Mixed aryl/alkyl, aryl/amido and aryl/phosphido complexes are unreactive, but amido/boryl complex 8 is competent for the activation of E?H bonds (E=H, B, Si) to give hydrido, boryl and silyl products. The results of these reactivity studies imply that the use of the very strongly σ‐donating boryl or silyl substituents is an effective strategy for rendering metallylene complexes competent for E?H bond activation.  相似文献   

2.
3.
The formation and the presence of hypervalent Si in the electron donor-acceptor complex H3N·SiH3Cl have been investigated by ab initio calculation. The results show that there is a 0.707eV decrease of energy when the complex H3N·SiH3Cl is formed from NH3 and H3SiCl, the interaction potential between the donor NH3 and the acceptor H3SiCl belongs to the Morse type, and the bond angle A(H-Si-Cl) versus bond length d(N-Si) presents a linear relation. The results also show that the interaction is mainly from giving the lone pair electrons in HOMO of NH3 to LUMO of H3SiCl, in which the 2p2 of N and the 3d0 of Si play important role. Bond N-Si is a weak n-σ* type dative one.  相似文献   

4.
Preparation and Crystal Structures of Dipyridiniomethane Monohalogenohydro-closo-Dodecaborates(2?), [(C5H5N)2CH2][B12H11X]; X = Cl, Br, I [B12H12]2? reacts with dihalogenomethanes CH2X2 in presence of trifluoro acetic acid, yielding the monohalogenododecaborates [B12H11X]2? (X = Cl, Br, I), which are separated by ion exchange chromatography on diethylaminoethyl(DEAE) cellulose from the starting compound and higher halogenated products. The X-ray structure determinations of [(C5H5N)2CH2][B12H11Cl] · 2(CH3)2SO (orthorhombic, space group Pnma, a = 17.351(6), b = 16.034(5), c = 9.659(2) Å, Z = 4) and of the isotypic bromo and iodo compounds [(C5H5N)2CH2][B12H11X] (monoclinic, space group P21/n, Z = 4; for X = Br: a = 7.339(2), b = 15.275(3), c = 16.761(4) Å, β = 96.80(2)°; for X = I: a = 7.4436(8), b = 15.3510(8), c = 16.9213(16) Å, ß = 97.326(7)°) exhibit crystal lattices build up by columns of substituted boron clusters and angular dications [(C5H5N)2CH2]2+ orientated along the shortest axis which are assembled to alternating layers.  相似文献   

5.
Preparation and Crystal Structure of the First Mixed Alkalimetal Hydrogencarbonates NaA2[H(CO3)2] · 2H2O with A = K, Rb The new hydrogencarbonates NaK2[H(CO3)2] · 2H2O (Pnma, a = 934.07(13) pm, b = 789.31(10) pm, c = 1142.1(5) pm, VEZ = 842.0(4) · 106 pm3, Z = 4, R1 (I ? 2σ(I)) = 0.023, wR2 = 0.066 for 989 reflections) and NaRb2[H(CO3)2] · 2H2O (Pnma, a = 948.24(11) pm, b = 811.37(9) pm, c = 1189.0(2) pm, VEZ = 914.8(2) · 106 pm3, Z = 4, R1 (I ≤ 2σ(I)) = 0.031, wR2 = 0.077 for 1063 reflections) were prepared from aqueous solutions. The crystal structures were determined. The isostructural compounds contain dimeric, non centrosymmetric [H(CO3)2]3? anions. In NaK2[H(CO3)2] · 2H2O a short hydrogen bond (d(O … O) = 246.1(2) pm) with an asymmetric potential was detected. In NaRb2[H(CO3)2] · 2H2O a hydrogen bond with symmetric potential (d(O … O) = 247.8(5) pm) can be assumed. The IR-spectra of NaK2[H(CO3)2] · 2H2O and Na3[H(CO3)2] · 2H2O are compared.  相似文献   

6.
Ab initio calculations were carried out to understand the effect of electron donating groups (EDG) and electron withdrawing groups (EWG) at the C5 position of cytosine (Cyt) and saturated cytosine (H2Cyt) of the deamination reaction. Geometries of the reactants, transition states, intermediates, and products were fully optimized at the B3LYP/6-31G(d,p) level in the gas phase as this level of theory has been found to agree very well with G3 theories. Activation energies, enthalpies, and Gibbs energies of activation along with the thermodynamic properties (ΔE, ΔH, and ΔG) of each reaction were calculated. A plot of the Gibbs energies of activation (ΔG) for C5 substituted Cyt and H2Cyt against the Hammett σ-constants reveal a good linear relationship. In general, both EDG and EWG substituents at the C5 position in Cyt results in higher ΔG and lower σ values compared to those of H2Cyt deamination reactions. C5 alkyl substituents ( H,  CH3,  CH2CH3,  CH2CH2CH3) increase ΔG values for Cyt, while the same substituents decrease ΔG values for H2Cyt which is likely due to steric effects. However, the Hammett σ-constants were found to decrease at the C5 position of cytosine (Cyt) and saturated cytosine (H2Cyt) on the deamination reaction. Both ΔG and σ values decrease for the substituents Cl and Br in the Cyt reaction, while ΔG values increase and σ decrease in the H2Cyt reaction. This may be due to high polarizability of bromine which results in a greater stabilization of the transition state in the case of bromine compared to chlorine. Regardless of the substituent at C5, the positive charge on C4 is greater in the TS compared to the reactant complex for both the Cyt and H2Cyt. Moreover, as the charges on C4 in the TS increase compared to reactant, ΔG also increase for the C5 alkyl substituents ( H,  CH3,  CH2CH3,  CH2CH2CH3) in Cyt, while ΔG decrease in H2Cyt. In addition, analysis of the frontier MO energies for the transition state structures shows that there is a correlation between the energy of the HOMO–LUMO gap and activation energies.  相似文献   

7.
Electron‐donating molecules play an important role in the development of organic solar cells. (Z )‐2‐(2‐Phenylhydrazinylidene)acenaphthen‐1(2H )‐one (PDAK), C18H12N2O, was synthesized by a Schiff base reaction. The crystal structure shows that the molecules are planar and are linked together forming `face‐to‐face' assemblies held together by intermolecular C—H…O, π–π and C—H…π interactions. PDAK exhibits a broadband UV–Vis absorption (200–648 nm) and a low HOMO–LUMO energy gap (1.91 eV; HOMO is the highest occupied molecular orbital and LUMO is the lowest unoccupied molecular orbital), while fluorescence quenching experiments provide evidence for electron transfer from the excited state of PDAK to C60. This suggests that the title molecule may be a suitable donor for use in organic solar cells.  相似文献   

8.
The first silicon analogues of carbonic (carboxylic) esters, the silanoic thio‐, seleno‐, and tellurosilylesters 3 (Si?S), 4 (Si?Se), and 5 (Si?Te), were prepared and isolated in crystalline form in high yield. These thermally robust compounds are easily accessible by direct reaction of the stable siloxysilylene L(Si:)OSi(H)L′ 2 (L=HC(CMe)2[N(aryl)2], L′=CH[(C?CH2)‐CMe][N(aryl)]2; aryl=2,6‐iPr2C6H3) with the respective elemental chalcogen. The novel compounds were fully characterized by methods including multinuclear NMR spectroscopy and single‐crystal X‐ray diffraction analysis. Owing to intramolecular N→Si donor–acceptor support of the Si?X moieties (X=S, Se, Te), these compounds have a classical valence‐bond N+–Si–X? resonance betaine structure. At the same time, they also display a relatively strong nonclassical Si?X π‐bonding interaction between the chalcogen lone‐pair electrons (nπ donor orbitals) and two antibonding Si? N orbitals (σ*π acceptor orbitals mainly located at silicon), which was shown by IR and UV/Vis spectroscopy. Accordingly, the Si?X bonds in the chalcogenoesters are 7.4 ( 3 ), 6.7 ( 4 ), and 6.9 % ( 5 ) shorter than the corresponding Si? X single bonds and, thus, only a little longer than those in electronically less disturbed Si?X systems (“heavier” ketones).  相似文献   

9.
Is2Si(Li)? P(Li)SiR3 (Is = 2,4,6-iPr3C6H2): The First Lithiumsilanidyl-lithiumphosphanides and their Transformation into Disilaphosphiranes and Novel Mercuriophosphanes Upon addition of two mol equivalents lithium metal to the Si?P bond of the silylidenephosphanes (“phosphasilenes”) Is2Si?P(SiiPr3) and Is2Si?P(SiPh2Me) (Is = 2,4,6-iPr3C6H2) in tetrahydrofurane, the corresponding lithiumsilanidyl-lithiumsilylphosphanides are formed, which have been isolated as pale yellow solids; each Li center is solvated by two tetrahydrofurane molecules. The constitution of these compounds is established by multi nuclei NMR spectroscopy and derivatization reactions with water, deuteriumoxide, dichlorodimethylsilane, and tert-butylmercuriochloride, respectively. The reaction with dichlorodimethylsilane yields, via cyclocondensation reaction and elimination of lithiumchloride, the first disilaphosphacyclopropane derivatives, and the reaction with tert-butylmercuriochloride gives, under loss of the triorganosilyl group at phosphorus and via rearrangement processes, an unusual P, P-dimercuriosilylphosphane; the latter reacts with tert-butylmercuriochloride and mercuriodichloride in the molar ratio of 2 : 1 : 1 to give an molecular aggregate bearing a P2Hg6Cl3 framework, which can be regared as an Lewis-acid base complex of a mercuriophosphane chelate ligand and mercuriodichloride.  相似文献   

10.
The First Hydrogencarbonates with a Trimeric [H2(CO3)3]4? Group: Preparation and Crystal Structure of Rb4H2(CO3)3 · H2O and K4H2(CO3)3 · 1.5 H2O Rb4H2(CO3)3 · H2O and K4H2(CO3)3 · 1,5 H2O were prepared by means of the reaction of (CH3)2CO3 with RbOH resp. KOH in aqueous methanole. Trimer [H2(CO3)3]4?-anions were found in the crystal structure of Rb4H2(CO3)3 · H2O (orthorhombic, Pnma (no. 62), a = 1 218.0(1) pm, b = 1 572.3(6) pm, c = 615.9(1) pm, VEZ = 1 179.5(5) · 106 pm3, Z = 4, R1(I ≥ 2σ(I)) = 0.027, wR2(I ≥ 2σ(I)) = 0.055). K4H2(CO3)3 · 1,5 H2O crystallizes in an OD-structure. The determined superposition structure (orthorhombic, Pbam (no. 55), a = 1 161.8(1) pm, b = 597.0(1) pm, c = 383.85(3) pm, VEZ = 266.3(1) · 106 pm3, Z = 1, R1(I ≥ 2σ(I)) = 0.035, wR2(I ≥ 2σ(I)) = 0.074) can be derived from the structure of the rubidium compound. The thermal decomposition of the substances is discussed.  相似文献   

11.
The structure and electronic parameters of ClZ(CH3)2X molecules (Z = C, Si, Ge, X = CH3, OCH3) were calculated by the RHF/6–31G(d) and RHF/6–311G(d,p) methods with full geometry optimization; calculations of ClZ(CH3)2OCH3 molecules were also performed by the RHF/6–31G(d) method with partial geometry optimization. The 35Cl NQR frequencies calculated from the populations of less diffuse 3p constituents of valence p orbitals of chlorine [RHF/6–31G(d)] were in agreement with the experimental values. The 35Cl NQR frequencies for molecules with X = OCH3 are lower than those for molecules with X = CH3 (the Z atom being the same), due mainly to direct through-field polarization of the Z-Cl bond, induced by the effect of unshared electron pair of the oxygen atom in the trans position with respect to that bond. The difference in the 35Cl NQR frequencies decreases in going from Z = C to Z = Si, Ge, in parallel with variation of the Z-Cl bond polarization as the size of Z increases.  相似文献   

12.
Syntheses and Crystal Structures of tBu‐substituted Disiloxanes tBu2SiX‐O‐SiYtBu2 (X = Y = OH, Br; X = OH, Y = H) and of the Adducts tBu3SiOH·(HO3SCF3)0.5·H2O and tBu3SiOLi·(LiO3SCF3)2·(H2O)2 The disiloxanes tBu2SiX‐O‐SiYtBu2 (X = Y = H, OH) are accessible from the reaction of CF3SO2Cl with tBu2SiHOH or tBu2Si(OH)2. By this reaction the disiloxane tBu2SiH‐O‐SiHtBu2 is formed together with tBu2SiH‐O‐SiOHtBu2. The disiloxanes tBu2SiX‐O‐SiYtBu2 (X = Y = Cl, Br) can be synthesized almost quantitatively from tBu2SiH‐O‐SiHtBu2 with Cl2 and Br2 in CH2Cl2. The structures of the disiloxanes tBu2SiX‐O‐SiYtBu2 (X = H, Y = OH; X = Y = OH, Br) show almost linear Si‐O‐Si units with short Si‐O bonds. Single crystals of the adducts tBu3SiOH·(HO3SCF3)0.5·H2O and tBu3SiOLi·(LiO3SCF3)2·(H2O)2 have been obtained from the reaction of tBu3SiOH with CF3SO3H and of tBu3SiO3SCF3 with LiOH. According to the result of the X‐ray structural analysis (hexagonal, P‐62c), tBu3SiOLi · (LiO3SCF3)2·(H2O)2 features the ion pair [(tBu3SiOLi)2(LiO3SCF3)3(H2O)3Li]+ [CF3SO3]?. The central framework of the cation forms a trigonal Li6 prism.  相似文献   

13.
The Crystal Structures of [Cu2Cl2(AA · H+)2](NO3)2 and [AA · H+]Picr? (AA · H+ = Allylammonium; Picr? = Picrat) By an alternating current electro synthesis the crystal-line π-complex [Cu2Cl2(AA · H+)2](NO3)2 has been obtained from CuCl2 · 2H2O, allylamine (AA), and HNO3 in ethanolic solution. X-ray structure analysis revealed that the compound crystallized in the monoclinic system, space group P21/a, a = 7.229(3), b = 7.824(3), c = 26.098(6) Å, γ = 94.46(5)°, Z = 4, R = 0.025 for 2 023 reflections. The crystal structure is built up of CunCln chains which are connected by π-bonding bidentate AA · H+ …? ON(O)O …? H+ · AA units. For comparision with the above complex the structure of [AA · H+]Picr? (Picr? = picrate anion) is also reported.  相似文献   

14.
Quantum chemical calculations of reaction mechanisms for the formal [2+2] addition of ethylene and acetylene to the amido‐substituted digermyne and distannyne Ph2N?EE?NPh2 (E=Ge, Sn) have been carried out by using density functional theory at the BP86/def2‐TZVPP level. The nature and bonding situations were studied with the NBO method and with the charge and energy decomposition analysis EDA‐NOCV. The addition of ethylene to Ph2N?EE?NPh2 takes place through an initial [2+1] addition to one metal atom and consecutive rearrangement to four‐membered cyclic species, which feature a weak E?E bond. Rotation about the C?C bond with concomitant rupture of the E?E bond leads to the 1,2‐disubstituted ethanes, which have terminal E(NPh2) groups. The overall reaction Ph2N?EE?NPh2+C2H4→(Ph2N)E?C2H4?E(NPh2) has very low activation barriers and is slightly exergonic for E=Ge but slightly endergonic for E=Sn. The analysis of the electronic structure shows that there is charge donation of nearly one electron to the ethylene moiety already in the first part of the reaction. The energy partitioning analysis suggests that the HOMO(Ph2N?EE?NPh2)→LUMO(C2H4) interaction has a similar strength as the HOMO(C2H4)→LUMO(Ph2N?EE?NPh2) interaction. The [2+2] addition of acetylene to Ph2N?EE?NPh2 also takes place through an initial [2+1] approach, which eventually leads to 1,2‐disubstituted olefins (Ph2N)E?C2H2?E(NPh2). The formation of the energetically lowest lying conformations of cis‐(Ph2N)E?C2H2?E(NPh2), which occurs with very low activation barriers, is clearly exergonic for the germanium and the tin compound. The trans‐coordinated isomers of (Ph2N)E?C2H2?E(NPh2) are slightly lower in energy than the cis form but they are separated by a substantial energy barrier for the rotation about the C?C bond. The energy decomposition analysis indicates that the initial reaction takes place under formation of electron‐sharing bonds between triplet fragments rather than HOMO–LUMO interactions.  相似文献   

15.
The layered acid solids of formula H3OUO2XO4· 3 H2O (X = As, P) intercalate aniline and benzidine arylamines, by protonation of the guest molecules. The intercalates maintain the original laminar structure.The insertion of aniline and benzidinein the metal derivativesM(UO2XO4)2·n H2O (M = Cu, VO, Fe) requires rather drastic conditions. Near and medium infrared spectra of intercalates in which M = VO2+ and Cu2+, indicate that the polymerization and/or oxidation of sorbed amines occurs; however, the guest-host reactions for Fe2+-derivatives are of the acid-base type.  相似文献   

16.
During phase formation experiments under hydrothermal conditions (250 °C, 5d) in the systems HgO/MXO4/H2O (M = Co, Zn, Cd; X = S, Se), single crystals of the mercuric compounds (CdSO4)2(HgO)2H2O (I), (CdSeO4)2(HgO)2H2O (II), (CdSeO4)Hg(OH)2 (III), (CoSO4)2(HgO)2H2O (IV), (ZnSO4)2(HgO)2H2O (V), (ZnSeO4)2(HgO)2H2O (VI), and the mixed‐valent (ZnSeIVO3)(ZnSeVIO4)HgI2(OH)2 (VII) were obtained. The crystal structure determinations from X‐ray diffraction data revealed four unique structure types for these compounds. I and II crystallise isotypically in space group P2/n (a ≈ 7.85, b ≈ 6.28, c ≈ 10.5Å, β ≈ 102°), compound III crystallises in space group C2/m (a = 10.540(2), b = 9.0120(8), c = 6.1330(12)Å, β = 100.45(3)°), and the isotypic compounds IV, V and VI crystallise in space group Pbcm (a ≈ 6.15, b ≈ 11.3, c ≈ 13.1Å). Common with these three structure types are distorted octahedral [MO6] and tetrahedral XO4 building units which are organised in a layered assembly. Within the layers H bonding of OH groups or H2O molecules of the [MO6] octahedra leads to an additional stabilisation. Adjacent layers are separated by mercury atoms which are linearly bonded to two O atoms at short distances, forming either interconnecting [O‐Hg‐O] units which are part of [O‐Hg‐O] zig‐zag chains, or single [HO‐Hg‐OH] units (realised in compound III). VII is the only compound with mercury in oxidation state +1. It crystallises in space group C2/m (a = 17.342(3), b = 6.1939(10), c = 4.4713(8)Å, β = 90.154(3)°) and is made up of Hg22+ dumbbells, [ZnO4(OH)2] octahedra, and statistically distributed SeVIO4 and SeIVO3 groups as the main building units.  相似文献   

17.
Reaction of CuCl2 · 2H2O, phenanthroline, maleic acid and NaOH in CH3OH/H2O (1:1 v/v) at pH = 7.0 yielded blue {[Cu(phen)]2(C4H2O4)2} · 4.5H2O, which crystallizes in the monoclinic space group C2/c (no. 15) with cell dimensions: a = 18.127(2)Å, b = 12.482(2)Å, c = 14.602(2)Å, β = 103.43(1)°, U = 3213.5(8)Å3, Z = 4. The crystal structure consists of the centrosymmetric dinuclear {[Cu(phen)]2(C4H2O4)2} complex molecules and hydrogen bonded H2O molecules. The Cu atoms are each square‐pyramidally coordinated by two N atoms of one phen ligand and three carboxyl O atoms of two maleato ligands with one carboxyl O atom at the apical position (d(Cu‐N) = 2.008, 2.012Å, equatorial d(Cu‐O) = 1.933, 1.969Å, axial d(Cu‐O) = 2.306Å). Two square‐pyramids are condensed via two apical carboxyl O atoms with a relatively larger Cu···Cu separation of 3.346(1)Å. The dinuclear complex molecules are assembled via the intermolecular π—π stacking interactions into 1D ribbons. Crossover of the resulting ribbons via interribbon π—π stacking interactions forms a 3D network with the tunnels occupied by H2O molecules. The title complex behaves paramagnetically between 5—300 K, following the Curie‐Weiss law χm(T—θ) = 0.435 cm3 · mol—1 · K with θ = 1.59 K.  相似文献   

18.
Single crystals of HgII(H4TeVIO6) (colourless to light‐yellow, rectangular plates) and HgI2(H4TeVIO6)(H6TeVIO6)·2H2O (colourless, irregular) were grown from concentrated solutions of orthotelluric acid, H6TeO6, and respective solutions of Hg(NO3)2 and Hg2(NO3)2. The crystal structures were solved and refined from single crystal diffractometer data sets (HgII(H4TeVIO6): space group Pna21, Z = 4, a =10.5491(17), b = 6.0706(9), c = 8.0654(13)Å, 1430 structure factors, 87 parameters, R[F2 > 2σ(F2)] = 0.0180; HgI2(H4TeVIO6)(H6TeVIO6)·2H2O: space group P1¯, Z = 1, a = 5.7522(6), b = 6.8941(10), c = 8.5785(10)Å, α = 90.394(8), β = 103.532(11), γ = 93.289(8)°, 2875 structure factors, 108 parameters, R[F2 > 2σ(F2)] = 0.0184). The structure of HgII(H4TeVIO6) is composed of ribbons parallel to the b axis which are built of [H4TeO6]2— anions and Hg2+ cations held together by two short Hg—O bonds with a mean distance of 2.037Å. Interpolyhedral hydrogen bonding between neighbouring [H4TeO6]2— groups, as well as longer Hg—O bonds between Hg atoms of one ribbon to O atoms of adjacent ribbons lead, to an additional stabilization of the framework structure. HgI2(H4TeVIO6)(H6TeVIO6)·2H2O is characterized by a distorted hexagonal array made up of [H4TeO6]2— and [H6TeO6] octahedra which spread parallel to the bc plane. Interpolyhedral hydrogen bonding between both building units stabilizes this arrangement. Adjacent planes are stacked along the a axis and are connected by Hg22+ dumbbells (d(Hg—Hg) = 2.5043(4)Å) situated in‐between the planes. Additional stabilization of the three‐dimensional network is provided by extensive hydrogen bonding between interstitial water molecules and O and OH‐groups of the [H4TeO6]2— and [H6TeO6] octahedra. Upon heating HgI2(H4TeVIO6)(H6TeVIO6)·2H2O decomposes into TeO2 under formation of the intermediate phases HgII3TeVIO6 and the mixed‐valent HgIITeIV/VI2O6.  相似文献   

19.
Reaction of a fresh Cu(OH)2x(CO3)1—x · yH2O precipitate with adipic acid (H2L) and 2, 2'—bipyridine (bpy) in ethanolic aqueous solution at room temperature afforded the hydrogen adipato bridged CuII coordination polymer [Cu(bpy)(HL)]2L · 6H2O consisting of double chains according to {[Cu(bpy)(HL)2/2]2L} and hydrogen bonded H2O molecules. The chains result from [Cu(bpy)]2+ units bridged by bis—monodentate hydrogen adipato ligands and further crosslinked by bis—monodentate adipato ligands. Through the interchain π—π stacking interactions and interchain C(bpy)—H···O(adipato) hydrogen bonding interactions, the double chains are assembled into layers, between which the crystal H2O molecules are located. The Cu atoms are square pyramidally coordinated by two N atoms of one bpy ligand and three O atoms of one adipato ligand and two hydrogen adipato ligands. The doubly bonded oxygen atom of the protonated carboxyl group occupies the apical position (Cu—N: 1.997, 2.005 Å; equatorial Cu—O: 1.925, 1.942 Å; apical Cu—O: 2.354 Å). Furthermore, the thermal behavior of the compound will be discussed. Crystal data: triclinic, P1¯ (no. 2), a = 9.618(1) Å, b = 9.933(1) Å, c = 12.782(2) Å, α = 70.88(1)°, β = 73.70(1)°, γ = 75.60(1)°, V = 1090.7(2) Å3, Z = 1, R = 0.0453 and wR2 = 0.1265 for 4643 observed reflections (Fo2 > 2σ(Fo2)) out of 4985 unique reflections.  相似文献   

20.
Two CoII complexes, Co(phen)(HL)2 ( 1 ) and [Co2(phen)2(H2O)4L2]·H2O ( 2 ) (H2L = HOOC‐(CH2)5‐COOH), were synthesized and structurally characterized on the basis of single crystal X‐ray diffraction data. In complex 1 the Co atoms are tetrahedrally coordinated by two N atoms of one phen ligand and two O atoms of different hydrogenpimelato ligands. Through π—π stacking interactions between carboxyl group and phen ligand, the complex molecules are assembled into 1D columnar chains, which are connected by intermolecular hydrogen bonds. Complex 2 consists of the centrosymmetric dinuclear [Co2(phen)2(H2O)4L2] molecules and hydrogen bonded H2O molecules. The Co atoms are each octahedrally surrounded by two N atoms of one phen ligand and four O atoms from two bis‐monodentate pimelato ligands and two H2O molecules at the trans positions. The results about thermal analyses, which were performed in flowing N2 atmosphere, on both complexes were discussed. Crystal data: ( 1 ) C2/c (no. 15), a = 13.491(1)Å, b = 9.828(1)Å, c = 19.392(2)Å, β = 100.648(1)°, U = 2526.9(4)Å3, Z = 4; ( 2 ) P1 (no. 2), a = 11.558(1)Å, b = 11.947(3)Å, c = 15.211(1)Å, α = 86.17(1)°, β = 75.55(1)°, γ = 69.95(1)°, U = 1910.3(3)Å3, Z = 2.  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