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 437 毫秒
1.
Cyanuric acid (C3H3N3O3) is widely used as cross‐linker in basic polymers (often in combination with other crosslinking agents like melamine) but also finds application in more sophisticated materials such as in supramolecular assemblies and molecular sheets. The unknown phosphorus analogue of cyanuric acid, P3C3(OH)3, may become an equally useful building block for phosphorus‐based polymers or materials which have unique properties. 1 Herein we describe a straightforward synthesis of 2,4,6‐tri(hydroxy)‐1,3,5‐triphosphinine and its derivatives P3C3(OR)3 which have been applied as strong π‐acceptor η6‐ligands in piano stool Mo(CO)3 complexes.  相似文献   

2.
The colorimetric detection of anionic species has been studied for α‐amino acid‐conjugated poly(phenylacetylene)s, which were prepared by the polymerization of the ethyl esters of N‐(4‐ethynylphenylsulfonyl)‐L ‐alanine, L ‐isoleucine, L ‐valine, L ‐phenylalanine, L ‐aspartic acid, and L ‐glutamic acid using Rh+(2,5‐norbornadiene)[(η6‐C6H5)B?(C6H5)3] as the catalyst in CHCl3. The one‐handed helical conformations of all the sulfonamide‐functionalized polymers were characterized by Cotton effects in the circular dichroism spectra. The addition of anions with a relatively high basicity, such as tetra‐n‐butylammonium acetate and fluoride, induced drastic changes in both the optical and chiroptical properties. On the other hand, anions with a relatively low basicity, such as tetra‐n‐butylammonium nitrate, azide, and bromide, had essentially no effects on the helical conformation of all the sulfonamide‐functionalized polymers. The anion signaling property of the sulfonamide‐functionalized polymers possessing α‐amino acid moieties was significantly affected by the installed residual amino acid structures. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 1683–1689, 2010  相似文献   

3.
[CpR(OC)Mo(μ‐η2:2‐P2)2FeCpR′] as Educt for Heterobimetallic Dinuclear Clusters with P2 and CnRnP4‐n Ligands (n = 1, 2) The cothermolysis of [CpR(OC)Mo(μ‐η2:2‐P2)2FeCpR′] ( 1 ) and tBuC≡P ( 2 ) as well as PhC≡CPh ( 3 ) affords the heterobimetallic triple‐decker like dinuclear clusters [(Cp'''Mo)(Cp*′Fe)(P3CtBu)(P2)] ( 4 ), Cp''' = C5H2tBu3‐1,2,4, Cp*′ = C5Me4Et, and [(Cp*Mo)(Cp*Fe)(P2C2Ph2)(P2)] ( 5 ) with a bridging tri‐ and diphosphabutadiendiyl ligand. 4 and 5 have been characterized additionally by X‐ray crystallography.  相似文献   

4.
The title two‐dimensional hydrogen‐bonded coordination compounds, [Cu(C8H5O4)2(C4H6N2)2], (I), and [Cu(C8H7O2)2(C4H6N2)2]·H2O, (II), have been synthesized and structurally characterized. The molecule of complex (I) lies across an inversion centre, and the Cu2+ ion is coordinated by two N atoms from two 4‐methyl‐1H‐imidazole (4‐MeIM) molecules and two O atoms from two 3‐carboxybenzoate (HBDC) anions in a square‐planar geometry. Adjacent molecules are linked through intermolecular N—H...O and O—H...O hydrogen bonds into a two‐dimensional sheet with (4,4) topology. In the asymmetric part of the unit cell of (II) there are two symmetry‐independent molecules, in which each Cu2+ ion is also coordinated by two N atoms from two 4‐MeIM molecules and two O atoms from two 3‐methylbenzoate (3‐MeBC) anions in a square‐planar coordination. Two neutral complex molecules are held together via N—H...O(carboxylate) hydrogen bonds to generate a dimeric pair, which is further linked via discrete water molecules into a two‐dimensional network with the Schläfli symbol (43)2(46,66,83). In both compounds, as well as the strong intermolecular hydrogen bonds, π–π interactions also stabilize the crystal stacking.  相似文献   

5.
The synthesis and characterization of the first supramolecular aggregates incorporating the organometallic cyclo‐P3 ligand complexes [CpRMo(CO)23‐P3)] (CpR=Cp (C5H5; 1a ), Cp* (C5(CH3)5; 1b )) as linking units is described. The reaction of the Cp derivative 1a with AgX (X=CF3SO3, Al{OC(CF3)3}4) yields the one‐dimensional (1D) coordination polymers [Ag{CpMo(CO)2(μ,η311‐P3)}2]n[Al{OC(CF3)3}4]n ( 2 ) and [Ag{CpMo(CO)2(μ,η311‐P3)}3]n[X]n (X=CF3SO3 ( 3a ), Al{OC(CF3)3}4 ( 3b )). The solid‐state structures of these polymers were revealed by X‐ray crystallography and shown to comprise polycationic chains well‐separated from the weakly coordinating anions. If AgCF3SO3 is used, polymer 3a is obtained regardless of reactant stoichiometry whereas in the case of Ag[Al{OC(CF3)3}4], reactant stoichiometry plays a decisive role in determining the structure and composition of the resulting product. Moreover, polymers 3a, b are the first examples of homoleptic silver complexes in which AgI centers are found octahedrally coordinated to six phosphorus atoms. The Cp* derivative 1b reacts with Ag[Al{OC(CF3)3}4] to yield the 1D polymer [Ag{Cp*Mo(CO)2(μ,η321‐P3)}2]n[Al{OC(CF3)3}4]n ( 4 ), the crystal structure of which differs from that of polymer 2 in the coordination mode of the cyclo‐P3 ligands: in 2 , the Ag+ cations are bridged by the cyclo‐P3 ligands in a η11 (edge bridging) fashion whereas in 4 , they are bridged exclusively in a η21 mode (face bridging). Thus, one third of the phosphorus atoms in 2 are not coordinated to silver while in 4 , all phosphorus atoms are engaged in coordination with silver. Comprehensive spectroscopic and analytical measurements revealed that the polymers 2 , 3a , b , and 4 depolymerize extensively upon dissolution and display dynamic behavior in solution, as evidenced in particular by variable temperature 31P NMR spectroscopy. Solid‐state 31P magic angle spinning (MAS) NMR measurements, performed on the polymers 2 , 3b , and 4 , demonstrated that the polymers 2 and 3b also display dynamic behavior in the solid state at room temperature. The X‐ray crystallographic characterisation of 1b is also reported.  相似文献   

6.
The metal‐directed self‐assembly of biphenylantimony trichloride and homocarboxylic acids LH [L = 2‐CHO‐C6H4COO ( 1 ), 2, 3‐2F‐C6H4COO ( 2 ), 4‐CF3–C6H4COO ( 3 )] provided three novel tetranuclear organoantimony(V) complexes, which were characterized by elemental analysis, FT‐IR, 1H, and 13C NMR spectroscopy as well as melting point, and X‐ray single crystal analysis. In the molecular structure, four hexacoordinate antimony atoms are linked into a [Sb2(μ‐O)2]2(μ‐O)2 “cage” architecture by oxo‐bridges which are terminally bridged by two carboxyl groups.  相似文献   

7.
The title compound, {[Zn4(C8H4O4)3(OH)2(C12H6N2O2)2]·2H2O}n, has been prepared hydrothermally by the reaction of Zn(NO3)2·6H2O with benzene‐1,4‐dicarboxylic acid (H2bdc) and 1,10‐phenanthroline‐5,6‐dione (pdon) in H2O. In the crystal structure, a tetranuclear Zn4(OH)2 fragment is located on a crystallographic inversion centre which relates two subunits, each containing a [ZnN2O4] octahedron and a [ZnO4] tetrahedron bridged by a μ3‐OH group. The pdon ligand chelates to zinc through its two N atoms to form part of the [ZnN2O4] octahedron. The two crystallographically independent bdc2− ligands are fully deprotonated and adopt μ3‐κOO′:κO′′ and μ4‐κOO′:κO′′:κO′′′ coordination modes, bridging three or four ZnII cations, respectively, from two Zn4(OH)2 units. The Zn4(OH)2 fragment connects six neighbouring tetranuclear units through four μ3‐bdc2− and two μ4‐bdc2− ligands, forming a three‐dimensional framework with uninodal 6‐connected α‐Po topology, in which the tetranuclear Zn4(OH)2 units are considered as 6‐connected nodes and the bdc2− ligands act as linkers. The uncoordinated water molecules are located on opposite sides of the Zn4(OH)2 unit and are connected to it through hydrogen‐bonding interactions involving hydroxide and carboxylate groups. The structure is further stabilized by extensive π–π interactions between the pdon and μ4‐bdc2− ligands.  相似文献   

8.
Colourless crystals of the title compound, [Cd2(C7H4IO2)4(C12H10N2)(H2O)2]n, were obtained by the self‐assembly of Cd(NO3)2·4H2O, 1,2‐bis(pyridin‐4‐yl)ethene (bpe) and 4‐iodobenzoic acid (4‐IBA). Each CdII atom is seven‐coordinated in a pentagonal–bipyramidal coordination environment by four carboxylate O atoms from two different 4‐IBA ligands, two O atoms from two water molecules and one N atom from a bpe ligand. The CdII centres are bridged by the aqua molecules and bpe ligands, which lie across centres of inversion, to give a two‐dimensional net. Topologically, taking the CdII atoms as nodes and the μ‐aqua and μ‐bpe ligands as linkers, the two‐dimensional structure can be simplified as a (6,3) network.  相似文献   

9.
The two isomorphous lanthanide coordination polymers, {[Ln2(C6H4NO2)2(C8H4O4)(OH)2(H2O)]·H2O}n (Ln = Er and Tm), contain two crystallographically independent Ln ions which are both eight‐coordinated by O atoms, but with quite different coordination environments. In both crystal structures, adjacent Ln atoms are bridged by μ3‐OH groups and carboxylate groups of isonicotinate and benzene‐1,2‐dicarboxylate ligands, forming infinite chains in which the Er...Er and Tm...Tm distances are in the ranges 3.622 (3)–3.894 (4) and 3.599 (7)–3.873 (1) Å, respectively. Adjacent chains are further connected through hydrogen bonds and π–π interactions into a three‐dimensional supramolecular framework.  相似文献   

10.
With regard to crystal engineering, building block or modular assembly methodologies have shown great success in the design and construction of metal–organic coordination polymers. The critical factor for the construction of coordination polymers is the rational choice of the organic building blocks and the metal centre. The reaction of Zn(OAc)2·2H2O (OAc is acetate) with 3‐nitrobenzoic acid (HNBA) and 4,4′‐bipyridine (4,4′‐bipy) under hydrothermal conditions produced a two‐dimensional zinc(II) supramolecular architecture, catena‐poly[[bis(3‐nitrobenzoato‐κ2O,O′)zinc(II)]‐μ‐4,4′‐bipyridine‐κ2N:N′], [Zn(C7H4NO4)2(C10H8N2)]n or [Zn(NBA)2(4,4′‐bipy)]n, which was characterized by elemental analysis, IR spectroscopy, thermogravimetric analysis and single‐crystal X‐ray diffraction analysis. The ZnII ions are connected by the 4,4′‐bipy ligands to form a one‐dimensional zigzag chain and the chains are decorated with anionic NBA ligands which interact further through aromatic π–π stacking interactions, expanding the structure into a threefold interpenetrated two‐dimensional supramolecular architecture. The solid‐state fluorescence analysis indicates a slight blue shift compared with pure 4,4′‐bipyridine and HNBA.  相似文献   

11.
A large cationic triangular metallo‐prism, [Ru6(p‐PriC6H4Me)6(tpt)2(dhbq)3]6+ ( 1 )6+, incorporating p‐cymene ruthenium building blocks, bridged by 2,5‐dihydroxy‐1,4‐benzoquinonato (dhbq) ligands, and connected by two 2,4,6‐tri(pyridin‐4‐yl)‐1,3,5‐triazine (tpt) subunits, allows the permanent encapsulation of the triphenylene derivatives hexahydroxytriphenylene, C18H6(OH)6 and hexamethoxytriphenylene, C18H6(OMe)6. These two cationic carceplex systems [C18H6(OH)6⊂ 1 ]6+ and [C18H6(OMe)6⊂ 1 ]6+ have been isolated as their triflate salts. The molecular structure of these systems has been established by one‐dimensional 1H ROESY NMR experiments as well as by the single‐crystal structure analysis of [C18H6(OMe)6⊂ 1 ][O3SCF3]6.  相似文献   

12.
The α,ω‐end‐capped poly(2‐methyl‐2‐oxazoline) (Cn‐POXZ‐Cn) have been synthesized by a one‐pot process using cationic ring‐opening polymerization with an appropriate initiator and terminating agent. The polymers bearing different alkyl groups C12 and C18 have molecular weight in the range of 2.4 × 103 to 14 × 103 with a small polydispersity index. The solution behavior of the free chains has been analyzed in a nonselective solvent, dichloromethane, by small‐angle neutron scattering and dynamic light scattering. These amphiphilic polymers associate in water to form flower‐like micellar structures. Critical micelle concentrations, investigated by fluorescence technique, are in the range of 0.03–0.5 g L?1 and are dependent on the hydrophilic/lipophilic balance. The structural properties of the aggregates have also been investigated by viscometry. Intrinsic viscosities of these polymers are in the same range as that of the precursors poly(2‐methyl‐2‐oxazoline) (POXZ) and mono‐functionalized polymers. Large viscosity increase corresponding to intermicellar bridging was observed in the vicinity of the micelle overlap concentration. Addition of hydroxypropyl β‐cyclodextrin (HβCD) has dissociated the aggregates and the intrinsic viscosities of the HβCD‐end‐capped chains have become comparable with the ones of POXZ precursor chains. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 2477–2485, 2010  相似文献   

13.
The reaction between fluorenyllithium and Mo(η3‐C3H5)Cl(NCMe)2(CO)2 led to the isolation of di‐μ3‐chlorido‐di‐μ3‐hydroxido‐tetrakis[(η3‐allyl)dicarbonylmolybdenum(II)]–9‐fluorenone–tetrahydrofuran (1/1/1), [Mo4(C3H5)4Cl2(OH)2(CO)8]·C4H8O·C13H8O. The tetrametallic Mo4 unit constitutes the first example of a complex containing simultaneously two μ3‐OH groups and two μ3‐Cl anions capping the metallic trigonal prism. The four crystallographically independent Mo2+ centres exhibit distorted octahedral geometry with the η3‐allyl groups being trans‐coordinated to a μ3‐OH group and the carbonyl groups occupying the equatorial plane. Space‐filling tetrahydrofuran and 9‐fluorenone molecules are engaged in strong O—H...O hydrogen‐bonding interactions with Mo43‐allyl)4Cl2(OH)2(CO)8 complexes.  相似文献   

14.
The dinuclear AuI complex containing the 4,5‐bis(diphenylphosphino)‐9,9‐dimethylxanthene (xantphos) ligand and trifluoroacetate anions exists in a solvent‐free form, [μ‐4,5‐bis(diphenylphosphino)‐9,9‐dimethylxanthene]bis[(trifluoroacetato)gold(I)], [Au2(C2F3O2)2(C39H32OP2)], (I), and as a dichloromethane solvate, [Au2(C2F3O2)2(C39H32OP2)]·0.58CH2Cl2, (II). The trifluoroacetate anions are coordinated to the AuI centres bridged by the xantphos ligand in both compounds. The AuI atoms are in distorted linear coordination environments in both compounds. The phosphine substituents are in a syn arrangement in the xantphos ligand, which facilitates the formation of short aurophilic Au...Au interactions of 2.8966 (8) Å in (I) and 2.9439 (6) Å in (II).  相似文献   

15.
The solid‐state structure of the title compound, [Na2Mn2(C32H56N2OSi2)2O2] or [1,8‐C10H6(NSiiPr3)2Mn(μ3‐O)Na(THF)]2, which lies across a crystallographic twofold axis, exhibits a central [Mn2O2Na2]4+ core, with two oxide groups, each triply bridging between the two MnIII ions and an Na+ ion. Additional coordination is provided to each MnIII centre by a 1,8‐C10H6(NSiiPr3)2 [1,8‐bis(triisopropylsilylamido)naphthalene] ligand and to the Na+ centres by a tetrahydrofuran molecule. The presence of an additional Na...H—C agostic interaction potentially contributes to the distortion around the bridging oxide group.  相似文献   

16.
In the title compounds, 4‐carboxyanilinium (2R,3R)‐tartrate, C7H8NO2+·C4H5O6, (I), and 4‐aminobenzoic acid, C7H7NO2, (II), the carboxyl planes of the 4‐carboxyanilinium cations/4‐aminobenzoic acid are twisted from the aromatic plane. In (I), the characteristic head‐to‐tail interactions are observed through the tartrate anions, forming two C22(7) chain motifs propagating parallel to the a and c axes of the unit cell. Also, the tartrate anions are connected through two primary C11(6) and C11(7) chain motifs, leading to a secondary R44(22) ring motif. In (II), head‐to‐tail interaction is seen through a discrete D11(2) motif and carboxyl group dimerization is observed through centrosymmetrically related R22(8) motifs around the inversion centres of the unit cell. The crystal structures of both compounds are stabilized by intricate three‐dimensional hydrogen‐bonding networks. Alternate hydrophobic and hydrophilic layers are observed in (I) as a result of a column‐like arrangement of the anions and the aromatic rings of the cations.  相似文献   

17.
Reacting stoichiometric amounts of 1‐(diphenylphosphino)ferrocene­carboxylic acid and [Ti(η5‐C5HMe4)22‐Me3SiC[triple‐bond]CSiMe3)] produced the title carboxyl­atotitanocene complex, [{μ‐1κ2O,O′:2(η5)‐C5H4CO2}{2(η5)‐C5H4P(C6H5)2}{1(η5)‐C5H(CH3)4}2FeIITiIII] or [FeTi(C9H13)2(C6H4O2)(C17H14P)]. The angle subtended by the Ti/O/O′ plane, where O and O′ are the donor atoms of the κ2‐carboxy­late group, and the plane of the carboxyl‐substituted ferrocene cyclo­penta­dienyl is 24.93 (6)°.  相似文献   

18.
The structure of the title compound, hexa­carbonyl‐1κ3C,2κ3C‐[3(η5)‐cyclo­penta­dienyl]­bis(μ3‐selenido)­diiron(II)­cobalt(II),[CoFe23‐Se)2(C5H5)(CO)6], was redetermined at room temperature and the correct C2/c space group was assumed instead of the previously reported P space group [Mathur et al. (1995). Organometallics, 14 , 2115–2118]. Analysis of the literature data showed that the previously reported triclinic parameters correspond to a primitive subcell of the actual monoclinic C‐centred cell with cell dimensions close to those found by us. The title compound appeared to be isostructural with the sulfur–selenium analogue.  相似文献   

19.
The title compound, {[Cd4(C5H2N2O4)(C5HN2O4)2(C10H8N2)2(H2O)]·2H2O}n, crystallized in the monoclinic space group P21/n and displays a three‐dimensional architecture. The asymmetric unit is composed of four crystallographically independent CdII centres, two triply deprotonated pyrazole‐3,5‐dicarboxylic acid molecules, one doubly deprotonated pyrazole‐3,5‐dicarboxylic acid molecule, two 2,2′‐bipyridine ligands, one coordinated water molecule and two interstitial water molecules. Interestingly, the CdII centers exhibit two different coordination numbers. Two CdII centres adopt a distorted octahedral arrangement and a third a trigonal–prismatic geometry, though they are all hexacoordinated. However, the fourth CdII center is heptacoordinated and displays a pentagonal–bipyramidal geometry. The three anionic ligands adopt μ3‐, μ4‐ and μ5‐bridging modes, first linking CdII centers into a one‐dimensional wave‐like band, then into a wave‐like layer and finally into a three‐dimensional coordination framework, which is stabilized by hydrogen bonds.  相似文献   

20.
The 1:1 proton‐transfer compounds of l ‐tartaric acid with 3‐aminopyridine [3‐aminopyridinium hydrogen (2R,3R)‐tartrate dihydrate, C5H7N2+·C4H5O6·2H2O, (I)], pyridine‐3‐carboxylic acid (nicotinic acid) [anhydrous 3‐carboxypyridinium hydrogen (2R,3R)‐tartrate, C6H6NO2+·C4H5O6, (II)] and pyridine‐2‐carboxylic acid [2‐carboxypyridinium hydrogen (2R,3R)‐tartrate monohydrate, C6H6NO2+·C4H5O6·H2O, (III)] have been determined. In (I) and (II), there is a direct pyridinium–carboxyl N+—H...O hydrogen‐bonding interaction, four‐centred in (II), giving conjoint cyclic R12(5) associations. In contrast, the N—H...O association in (III) is with a water O‐atom acceptor, which provides links to separate tartrate anions through Ohydroxy acceptors. All three compounds have the head‐to‐tail C(7) hydrogen‐bonded chain substructures commonly associated with 1:1 proton‐transfer hydrogen tartrate salts. These chains are extended into two‐dimensional sheets which, in hydrates (I) and (III) additionally involve the solvent water molecules. Three‐dimensional hydrogen‐bonded structures are generated via crosslinking through the associative functional groups of the substituted pyridinium cations. In the sheet struture of (I), both water molecules act as donors and acceptors in interactions with separate carboxyl and hydroxy O‐atom acceptors of the primary tartrate chains, closing conjoint cyclic R44(8), R34(11) and R33(12) associations. Also, in (II) and (III) there are strong cation carboxyl–carboxyl O—H...O hydrogen bonds [O...O = 2.5387 (17) Å in (II) and 2.441 (3) Å in (III)], which in (II) form part of a cyclic R22(6) inter‐sheet association. This series of heteroaromatic Lewis base–hydrogen l ‐tartrate salts provides further examples of molecular assembly facilitated by the presence of the classical two‐dimensional hydrogen‐bonded hydrogen tartrate or hydrogen tartrate–water sheet substructures which are expanded into three‐dimensional frameworks via peripheral cation bifunctional substituent‐group crosslinking interactions.  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