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 29 毫秒
1.
In contrast to the well‐known reaction of phosphonic acids RP(O)(OH)2 with divalent transition‐metal ions that yields layered metal phosphonates [RPO3M(H2O)]n, the 2,6‐diisopropylphenyl ester of phosphoric acid, dippH2, reacts with zinc acetate in methanol under ambient conditions to afford tetrameric zinc phosphate [(ArO)PO3Zn(MeOH)]4 ( 1 ). The coordinated methanol in 1 can be readily exchanged by stronger Lewis basic ligands at room temperature. This strategy opens up a new avenue for building double‐four‐ring (D4R) cubane‐based supramolecular assemblies through strong intercubane hydrogen‐bonding interactions. Seventeen pyridinic ligands have been used to synthesize as many D4R cubanes [(ArO)PO3Zn(L)]4 ( 2 – 18 ) from 1 . The ligands have been chosen in such a way that the majority of them contain an additional functional group that could be used for noncovalent synthesis of extended structures. When the ligand does not contain any other hydrogen‐bonding donor–acceptor sites (e.g., 2,4,6‐trimethylpyridine (collidine)), zero‐dimensional D4R cubanes have been obtained. The use of pyridine, lutidine, 2‐aminopyridine, and 2,6‐diaminopyridine, however, results in the formation of linear or zigzag one‐dimensional assemblies of D4R cubanes through strong intermolecular C? H???O or N? H???O interactions. Construction of two‐dimensional assemblies of zinc phosphates has been achieved by employing 2‐hydroxypyridine or 2‐methylimidazole as the exo‐cubane ligand on zinc centers. The introduction of an alcohol side chain on the pyridinic ligand in such a way that the ? CH2OH group cannot participate in intracubane hydrogen bonding (e.g., pyridine‐3‐methanol, pyridine‐4‐methanol, and 3,5‐dimethylpyrazole‐N‐ethanol) leads to the facile noncovalent synthesis of three‐dimensional framework structures. Apart from being useful as building blocks for noncovalent synthesis of zeolite‐like materials, compounds 1 – 18 can also be thermolyzed at approximately 500 °C to yield high‐purity zinc pyrophosphate (Zn2P2O7) ceramic material.  相似文献   

2.
Four coordination polymers, namely, [Zn2(TIYM)(2,6‐PYDC)2]n · n(CH3OH) · 3n(H2O) ( 1 ), [Cu(TIYM)(2,6‐PYDC)]n · 3n(H2O) ( 2 ), [Co(TIYM)(2,6‐PYDC)]n · n(CH3OH) · 3n(H2O) ( 3 ), and [Cd2(TIYM)(2,6‐PYDC)2(H2O)]n · n(H2O) ( 4 ) with the flexible N‐containing ligand [tetrakis(imidazol‐1‐ylmethyl)methane (TIYM)] and the N‐containing dicarboxylic acid [2,6‐pyridinedicarboxylic acid (2,6‐PYDC)] were prepared. Compounds 1 – 4 show various structures because of different N–Ccenter–N angles (θ) of TIYM ligands and changing coordination modes of 2,6‐PYDC. Compounds 1 , 2 , and 3 display a similar 1D ladder‐like chain, whereas 4 gives a 1D quad‐core lifting platform shaped belt. The structural diversities in 1 – 4 suggest that the multiple coordination modes or the different freely twist angles of ligands and the presence of different metal atoms play important roles in the resulting structures of the coordination polymers. Furthermore, the solid‐state luminescence properties of 1 and 4 , and the magnetic properties of 3 were investigated.  相似文献   

3.
A new dinuclear cobalt(II) complex [Co2L2Cl2(CH3OH)2] ( 1 ), where HL = 3‐[(furan‐2‐ylmethylimino)methyl]‐2‐hydroxy‐5‐methylbenzaldehyde, derived from the in situ condensation of 2,6‐diformyl‐4‐methylphenol with furfurylamine, was prepared and structurally and magnetically characterized. Single crystal X‐ray structural determination reveals that the structure consists of centrosymmetric dinuclear units with each CoII ion in a slightly distorted octahedral environment. Lines’ model, which in principle can theoretically separate in spin‐only and orbital contribution, was used to fit the variable temperature susceptibility (2–300 K), suggesting an intramolecular antiferromagnetic interaction between the cobalt(II) ions.  相似文献   

4.
Hydrothermal reaction of Cd(NO3)2·4H2O with bbp and p-PDOAH2 at 140 ℃ yielded a novel 1D cadmium(Ⅱ) coordination polymer, [Cd(bbp)(p-PDOA)]n (bbp=2,6-bis(benzimidazol-2-yl)pyridine, p-PDOA=p-phenylenedioxydiacetate dianion), in which CdN3O4 pentagonal bipyramids were linked by p-PDOA ligands in a bis-bidentate mode to construct a zigzag chain with the adjacent Cd…Cd distance of 1.14(1) nm, There exists a 2D supramolecular network linked by π-π stacking with a face-to-face distance of 0.35(1) nm between the 2,6-bis(benzimidazol-2-yl) pyridine ligands and hydrogen-bonding interactions (0.27(4) nm). A 3D supramolecular network was further constructed by these non-covalent interactions between the zippers. The TG/DTG showed that its chain skeleton was thermally stable up to 389 ℃ and the blue fluorescent emission of the complex was determined at 428 nm in a solid state with its long decay lifetime of 7.24 ns.  相似文献   

5.
Two closely related oximes, namely 1‐chloroacetyl‐3‐ethyl‐2,6‐diphenylpiperidin‐4‐one oxime, C21H23ClN2O2, (I), and 1‐chloroacetyl‐2,6‐diphenyl‐3‐(propan‐2‐yl)piperidin‐4‐one oxime, C22H25ClN2O2, (II), despite their identical sets of hydrogen‐bond donors and acceptors, display basically different hydrogen‐bonding patterns in their crystal structures. While the molecules of (I) are organized into typical centrosymmetric dimers, created by oxime–oxime O—H...N hydrogen bonds, in the structure of (II) there are infinite chains of molecules connected by O—H...O hydrogen bonds, in which the carbonyl O atom from the chloroacetyl group acts as the hydrogen‐bond acceptor. Despite the differences in the hydrogen‐bond schemes, the –OH groups are always in typical anti positions (C—N—O—H torsion angles of ca 180°). The oxime group in (I) is disordered, with the hydroxy groups occupying two distinct positions and C—C—N—O torsion angles of approximately 0 and 180° for the two alternatives. This disorder, even though the site‐occupancy factor of the less occupied position is as low as ca 0.06, is also observed at lower temperatures, which seems to favour the statistical and not the dynamic nature of this phenomenon.  相似文献   

6.
We have prepared two chiral Schiff base ligands, H2L1 and H2L2, and one achiral Schiff base ligand, H2L3, by treating 2,6‐diformyl‐4‐methylphenol separately with (R )‐1,2‐diaminopropane, (R )‐1,2‐diaminocyclohexane and 1,1′‐dimethylethylenediamine, in ethanolic medium, respectively. The complexes MnL1ClO4 ( 1 ), MnL2ClO4 ( 2 ), MnL3ClO4 ( 3 ), FeL1ClO4 ( 4 ), FeL2ClO4 ( 5 ) and FeL3ClO4 ( 6 ) have been obtained by reacting the ligands H2L1, H2L2 and H2L3 with manganese(III) perchlorate or iron(III) perchlorate in methanol. Circular dichroism studies suggest that ligands H2L1 and H2L2 and their corresponding complexes have asymmetric character. Complexes 1 – 6 have been used as homogeneous catalysts for epoxidation of alkenes. Manganese systems have been found to be much better than iron counterparts for alkene epoxidation, with 3 as the best catalyst among manganese systems and 6 as the best among iron systems. The order of their experimental catalytic efficiency has also been rationalized by theoretical calculations. We have observed higher enantiomeric excess product with catalysts 1 and 4 , so they were attached to surface‐modified magnetic nanoparticles to obtain two new magnetically separable nanocatalysts, Fe3O4@dopa@MnL1 and Fe3O4@dopa@FeL4. They have been characterized and their alkene epoxidation ability has been investigated. These catalysts can be easily recovered by magnetic separation and recycled several times without significant loss of catalytic activity. Hence our study focuses on the synthesis of a magnetically recoverable asymmetric nanocatalyst that finds applications in epoxidation of alkenes and at the same time can be recycled and reused.  相似文献   

7.
Two cadmium(II) entangled frameworks, Cd(BIPA)(bpe)1.5 ( 1 ) and Cd(BIPA)(bpp)(H2O) ( 2 ), were prepared by hydrothermal reactions based on rigid 5‐bromoisophthalic acid (H2BIPA) and two flexible bipyridyl ligands 1, 2‐bis(4‐pyridyl)ethane (bpe) and 1, 3‐bis(4‐pyridyl)propane (bpp). The complexes were characterized by elemental analysis, IR spectroscopy, thermogravimetric analysis, and single‐crystal X‐ray diffraction. Complex 1 is a rare example of a polycatenated array of 1D nanotubes, whereas complex 2 exhibits a three‐dimensional twofold interpenetrating diamondoid network. The analysis results reveal that the flexibility of the bipyridyl ligands plays a significant role in the structure of the final products. Moreover, the luminescent properties of the complexes were investigated.  相似文献   

8.
Complexes with Macrocyclic Ligands. V Dinuclear Copper(II) Complexes with Chiral Macrocyclic Ligands of Schiff‐Base Type: Syntheses and Structures The synthesis and properties of four chiral, dinuclear, macrocyclic, cationic copper(II) complexes, [Cu2(Lm,n)]2+ ( 1 – 4 ), are described. The two symmetrical compounds [Cu2(L2,2)][ClO4]2 ( 1 and 2 ) were synthesized in a one‐step reaction from 2,6‐diformyl‐4‐tert.‐butylphenol, copper(II)‐perchlorate and the chiral diamine (1S,2S)‐1,2‐diphenylethylenediamine (synthesis of 1 ) and (1R,2R)‐1,2‐diaminocyclohexane (synthesis of 2 ), respectively. For the synthesis of the two unsymmetrical compounds [Cu2(LPh,n)][ClO4]2 ( 3 and 4 ) the mononuclear, neutral copper(II) complex [CuLPh] ( 5 ) [synthesized from 2,6‐diformyl‐4‐tert.‐butylphenol, copper(II)‐acetate and 1,2‐phenylenediamine] was reacted with (1R,2R)‐1,2‐diaminocyclohexane (synthesis of 3 ) and (S)‐1,1′‐binaphthyl‐2,2′‐diamine (synthesis of 4 ), respectively. The structures of the two unsymmetrical copper(II) compounds ( 3 and 4 ) were determined by X‐ray diffraction.  相似文献   

9.
This paper is focused on investigation of coordination polymers constructed by Cu(II) and rigid pyridyl ligands, such as 4,4′‐bipyridyl‐1,2,4‐triazole (Hpytz) and 1,10‐phenanthroline (phen) from a mononuclear precursor [Cu(DMF)4(NCS)2] ( 1 ). As expected, the complex was self‐assembled with phen to form a 1D double‐stranded chain of [Cu(phen)(µ‐SCN)2] ( 2 ), with Hpytz to form 1D zigzag chain of [Cu2(µ‐Hpytz)2(NCS)2(DMF)2(µ‐SCN)2] ( 3 ) in which thiocyanate anion linked the Cu‐µ‐Hpytz‐Cu chain into an infinite 2D network via weak Cu···S interaction. To treat 3 with the bridged anion dca, a novel 3D framework [Cu(µ‐Hpytz)(µ‐dca)(µ‐SCN)] ( 4 ) was obtained in which Cu‐µ‐Hpytz‐Cu chain is preserved and both thiocyanate anion and dicyanamide (dca) act as bridging ligands. In addition, complex 3 was applied as a metal catalyst in polymerization of MMA in aqueous solution at room temperature.  相似文献   

10.
A novel complex [Cd(Bipy)2(L)] · 10H2O (I) (H2L = 2,6-pyridinedicarboxylic acid), has been synthesized by the reaction of bipyridyl and H2L with cadmium(II) salt. Elemental analysis, IR spectra, and X-ray single-crystal diffraction were carried out to determine the composition and crystal structure of complex I. The cadmium atom is seven-coordinated in a distorted pentagonal bipyramidal configuration. Ten water molecules formed a large water cluster with the oxygen atoms of the H2L ligand. The complexes form a 2D supramolecular framework and a 3D reticulate structure by hydrogen bonding and π-π-stacking of the neighboring bipyridyl ligands coming from the neighboring complexes.  相似文献   

11.
The dimethyl aryloxide complexes [(PNP)M(CH3)2(OAr)] (M=Zr or Hf; PNP?=N[2‐P(CHMe2)2‐4‐methylphenyl]2); Ar=2,6‐iPr2C6H3), which were readily prepared from [(PNP)M(CH3)3] by alcoholysis with HOAr, undergo photolytically induced α‐hydrogen abstraction to cleanly produce complexes [(PNP)M=CH2(OAr)] with terminal methylidene ligands. These unique systems have been fully characterized, including the determination of a solid‐state structure in the case of M=Zr.  相似文献   

12.
The sodium salt of a complex anion formed between gadolinium(III) and three variously deprotonated chelidamic acid (4‐hydroxy­pyridine‐2,6‐di­carboxyl­ic acid) ligand moi­eties, assigned as Na5[Gd(C7H2NO5)2(C7H3NO5)]·16H2O, i.e. pentasodium (4‐hydroxy­pyridine‐2,6‐di­carboxyl­ate)­bis(4‐oxido­pyridine‐2,6‐di­carboxyl­ate)­gadolinium(III) hexadecahydrate, forms as colourless monoclinic crystals upon vapour diffusion of ethanol into its aqueous solution. The ligand moieties, assigned as two trianionic and one dianionic chelidamate species, are all tridentate in the complex anion of tricapped trigonal prismatic donor‐atom geometry. The geometry of the ligands and that of the primary coordination sphere is very similar to that of the analogous anionic tris­(ligand)–rare earth complexes of the pyridine‐2,6‐di­carboxyl­ate (dipicolinate) dianion.  相似文献   

13.
The anionic lanthanide‐sodium‐2,6‐di‐tert‐butyl‐phenoxide complexes [Ln(OAr)4][Na(DME)3]·DME (Ln = Nd 1 (neodymium), Sm 2 (samarium), or Gd 3 (gadolium); DME = dimethoxyethane) were synthesized by the reaction of anhydrous LnCl3 with 4 equiv of sodium‐2,6‐di‐tert‐butyl‐phenoxide NaOAr in high yields and structurally characterized. These complexes showed high catalytic activity in the ring‐opening polymerizations of ?‐caprolactone (?‐CL) and trimethylene carbonate (TMC). The catalytic activity profoundly depended on the lanthanide metals. The active order of Gd < Sm < Nd for the polymerization of ?‐CL and TMC was observed. The polymers obtained with these initiators all showed a unimodal molecular weight distribution, indicating that the [Ln(OAr)4][Na(DME)3]·DME anionic complexes could be used as single‐component initiators. The anionic complex was more efficient than the corresponding neutral complex, Ln(OAr)3(THF)2. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 1210–1218, 2007  相似文献   

14.
Ion mobility coupled with mass spectrometry provides a fast and repeatable method to separate catechin epimers by previous complexation with selected chiral modifiers and transition metals. Several combinations with chiral ligands such as D‐ and L‐amino acids and/or additional metal cations, chiral crown ethers, tartaric acid and heptakis(2,6‐di‐O‐methyl)‐β‐cyclodextrin were screened for their ability to affect the separation efficiency. The clusters having the form of [2M + D‐amino acid + Cu2+ ? 3H]? (M stands for (?)‐epicatechin or (+)‐catechin) showed improvement in stereodifferentiation between two epimeric catechins in comparison to the analysis of pure epimers, where no separation was observed or the separation was hampered by the formation of mixed dimer complexes. Among various examined D‐amino acids only those possessing hydrophobic side chains induced the improvement of separation efficiency. The best peak‐to‐peak resolution (Rp–p) was determined to be 0.71 for [2M + D‐Leucine + Cu2+ ? 3H]? clusters. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

15.
Two new CdII complexes, [Cd( ces )(phen)] ( 1 ) and {[Cd( ces )(bpy)(H2O)](H2O)}2 ( 2 ), were prepared by slow solvent evaporation methods from mixtures of cis‐epoxysuccinic acid and Cd(ClO4)2 · 6H2O in the presence of phen or bpy co‐ligand ( ces = cis‐epoxysuccinate, phen = 1,10‐phenanthroline, and bpy = 2,2′‐bipyridine). Single‐crystal X‐ray diffraction analyses show that complex 1 has a one‐dimensional (1D) helical chain that is further assembled into a two‐dimensional (2D) sheet, and then an overall three‐dimensional (3D) network by the interchain C–H ··· O hydrogen bonds. Complex 2 features a dinuclear structure, which is further interlinked into a 3D supramolecular network by the co‐effects of intermolecular C–H ··· O and C–H ··· π hydrogen bonds as well as π ··· π stacking interactions. The structural differences between 1 and 2 are attributable to the intervention of different 2,2′‐bipyridyl‐like co‐ligands. Moreover, 1 and 2 exhibit intense solid‐state luminescence at room temperature, which mainly originates from the intraligand π→π* transitions of aromatic co‐ligands.  相似文献   

16.
手性高分子P–1由(R)-5,5′-二溴-6,6′-二(4-三氟甲基苯基)-2,2′-二正辛氧基-1,1′-联萘(R–M–1)和5,5′-二乙烯基-2,2′-联吡啶(M–2)通过Pd催化的Heck偶合反应合成得到,高分子配合物P-2和P-3由高分子P-1与Eu(TTA)3·2H2O和Gd(TTA)3·2H2O (TTA– = 2-噻吩甲酰三氟丙酮)反应生成。手性高分子P-1能发射强的蓝色荧光,这是由于手性重复单元(R)-6,6′-二(4-三氟甲基苯基)-2,2′-二正辛氧基-1,1′-联萘和单元2,2′-联吡啶通过亚乙烯基桥连形成共轭高分子结构造成的。在不同的激发波长激发下,含Eu(III)的高分子配合物P–2不仅显示高分子荧光,还可显示Eu(III) (5D0→7F2)特征荧光。含Gd(III)的高分子配合物P–3仅发射高分子荧光。基于高分子及含RE(III)的高分子配合物的荧光性质研究发现,共轭高分子并没有把能量转移到Eu(III)或Gd(III) 配合物部分,只发射它自身的荧光,含Eu(III)的高分子配合物P–2发射Eu(III) (5D0→7F2)特征荧光能量主要来源于配阴离子TTA–。  相似文献   

17.
Deoximation in metal chloride ionic liquids based on 1‐alkyl‐3‐methylimidazolium and triethylene ammonium cations, such as AmimBr(Cl)‐MClx(A=ethyl, butyl, benzyl; M=Al, Fe, Cu, Sn and Zn; x=2, 3) and Et3NHCl‐FeCl3 were investigated under mild conditions. Ferrate chloride ionic liquid was proved to be an effective catalyst for deoximation of cyclohexanone oxime, exhibiting high conversion of oximes and selectivity to cyclohexanone. Good performance for the deoximation of other oximes such as salicylald oxime, acetone oxime, benzophenone oxime, 4‐nitrobenzald oxime, acetophenone oxime, 2‐chlorobenzaldehyde oxime, Acetald oxime, 2‐butanone oxime and (1R)‐camphor oxime was also achieved with bmimBr‐FeCl3 as catalyst and solvent. The deoximation was determined to carry out via acid‐catalytic hydrolysis and the reaction mechanism was proposed.  相似文献   

18.
C? H activation by acetate‐assisted cyclometallation of a phenyl group with half‐sandwich complexes [{MCl2Cp*}2] (M=Ir, Rh) and [{RuCl2(p‐cymene)}2] can be directed by a wide range of nitrogen donor ligands including pyrazole, oxazoline, oxime, imidazole and triazole, and X‐ray structures of a number of complexes are reported. All the ligands tested cyclometallated at iridium, however ruthenium and rhodium fail to cause cyclometallation in some cases. As a result, the nitrogen donors have been categorised based on their reactivity with the three metals used. The relevance of these cyclometallation reactions to catalytic synthesis of carbocycles and heterocycles is discussed.  相似文献   

19.
The series of binuclear Cu(II) and Ni(II) complexes with an asymmetrical exchange fragment based on 2,6‐diformyl‐4‐methylphenol bishydrazone has been synthesized for the first time. The compositions and structures of both ligands and its complexes have been established with the data of IR, 1H NMR, and extended X‐ray absorption fine structure (EXAFS) spectroscopical studies as well as magnetic measurements. The structure of [Ni2L3(μ‐Pz)] · 2CH3OH (L = triply deprotonated form of bishydrazone, Pz = pyrazol) was confirmed by X‐ray crystallographic analysis. In this complex, the coordination environment of two nickel ions is quite different, one nickel atom is square‐planar and the other is distorted octahedral coordinated. The values of exchange parameter calculated in terms of HDVV theory have been compared with the features of an asymmetrical exchange fragment's electronic and geometrical structure.  相似文献   

20.
A new family of t‐butyl substituted chromium(III) chloride complexes ( Cr1 – Cr6 ), bearing 2‐(1‐(2,6‐dibenzhydryl‐4‐t‐butylphenylimino)ethyl)‐6‐(1‐(arylimino)ethyl)pyridine (aryl = 2,6‐Me2C6H3 Cr1 , 2,6‐Et2C6H3 Cr2 , 2,6‐i‐Pr2C6H3 Cr3 , 2,4,6‐Me3C6H2 Cr4 and 2,6‐Et2‐4‐MeC6H2 Cr5 ) or 2,6‐bis(1‐(2,6‐dibenzhydryl‐4‐t‐butylphenylimino)ethyl)pyridine ( Cr6 ), has been synthesized by the reaction of CrCl3·6H2O in good yield with the corresponding ligands ( L1 – L6 ), respectively. The molecular structures of Cr2 and Cr6 were characterized by X‐ray diffraction highlighted a distorted octahedral geometry with the coordinated N,N,N ligand and three bonded chlorides around the metal center. On activation with modified methylaluminoxane or triisobutyl aluminum, most of the chromium precatalysts exhibit good activities toward ethylene polymerization and produce linear polyethylenes with high‐molecular weight. In addition, an in‐depth catalytic evaluation of Cr2 was conducted to investigate how cocatalyst type and amount, reaction temperature, and run time affect the catalytic activities and polymer properties. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2019 , 57, 1049–1058  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