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.
One group of ligands used in transition metal complexes is synthesized by derivatizing 1,10‐phenanthroline. These metal complexes are of interest for study in the field of photovoltaic devices and solar fuels. Previous strategies for obtaining the 5,6‐diones of substituted 1,10‐phenanthrolines do not work for 2,9‐diphenyl‐1,10‐phenanthroline due to undesired products resulting from oxidation of the phenyl substituents. However, 2,9‐diphenyl‐1,10‐phenanthroline‐5,6‐dione can be obtained in reasonable yield by oxidation with BrO3? in weak aqueous acid. The resulting dione can be converted directly to the 5,6‐dialkoxy product upon two electron reduction in aprotic solvent followed by treatment with appropriate alkylating agents.  相似文献   

2.
Treatment of 2‐chloro‐ or 2,9‐dichloro‐1,10‐phenanthroline with aqueous HBr or aqueous H2SO4 at 120°C yielded 1,10‐phenanthroline‐2(1H)‐one or 1,10‐dihydro‐1,10‐phenanthroline‐2,9‐dione, respectively. The hydrolysis of 2,9‐dichloro‐1,10‐phenanthroline with 37% aqueous HC1 led to the half hydrolyzed amide and the bis‐amide. Under comparable reactions conditions, using aqueous HBr, H2SO4 or HC1, 2‐chloropyridine was found to be hydrolytically stable. On the other hand, 2‐chloro‐ or 2,9‐dichloro‐1,10‐phenanthroline on heating with 57% aqueous HI afforded the HI salts of 2‐iodo‐ or 2,9‐diiodo‐1,10‐phenanthroline, which could be isolated. These salts on treatment with aqueous ammonium hydroxide led to good yields of 2‐iodo‐ and 2, 9‐diiodo‐1,10‐phenanthroline, respectively. Treatment of 2‐chloropyridine with 57% aqueous HI under similar reaction conditions led to 2‐iodopyridine in a 10% conversion.  相似文献   

3.
Efficient separation of minor actinides from spent nuclear fuel remains a formidable challenge. As part of ongoing efforts to identify effective ligands for separation of toxic radionuclides, a series of bis‐1,2,4‐triazines, three novel, have been prepared from [1,10]‐phenanthroline‐2,9‐dicarbonitrile in two‐telescoped steps without additives, complicated workups, prolonged reaction times, or additional purification.  相似文献   

4.
We synthesized self‐assembled coordination cages in which 1,10‐phenanthroline derivatives serve as capping ligands. Substituents at the 2,9‐positions of the phenanthroline ligand covered the outside of the cage but had an impact on the guest binding inside the cage. Introduction of methyl groups at the 2,9‐positions allowed the cage to accommodate tetraphenylsilane. Bulky mesityl groups overhanging the cage framework significantly shrunk the cage cavity through π–π interactions with the aromatic panels of the cage. The p‐methyl group of the mesityl substituent was a determinant of the restricted motion of 4,4′‐dimethoxybenzil inside the cage at high temperature. Thus, the presence or absence of one methyl group, which is far from the guest‐binding site, makes a significant difference in the guest species and motions inside the cage.  相似文献   

5.
Abstract

Poly(aryl ether)s were synthesized by reaction of 4,7‐dichloro‐3,8‐diphenyl‐1,10‐phenanthroline and 4,7‐dichloro‐2,9‐dimethyl‐1,10‐phenanthroline with bisphenol A (BPA) in the presence of potassium carbonate in N,N‐dimethylacetamide or N‐methylpyrrolidinone. High molecular weight homopolymers could not be prepared because of the insolubility of the polymers resulting in premature precipitation from the reaction mixture. Soluble, high molecular weight copolymers were readily prepared containing up to 70 mol% of the 1,10‐phenanthroline moieties. The copolymers were all highly fluorescent with blue emission.  相似文献   

6.
The structure of the title compound, [Co(C12H8N2)(H2O)4](NO3)2, consists of tetra­aqua­(1,10‐phenanthroline)cobalt(II) cations and nitrate anions. The Co atom is located on a twofold rotation axis and is coordinated by the two N atoms of a 1,10‐phenanthroline ligand and four O atoms of water mol­ecules. The cations and anions are linked by hydrogen‐bond inter­actions into a three‐dimensional supra­molecular network.  相似文献   

7.
The novel 1,10‐phenanthroline‐2,9‐dicarboxylate containing Chromium(III) complex, (pydaH)[Cr(phendc)2] · 5H2O, was synthesized using proton‐transfer compound LH2, (pydaH2)2+(phendc)2?, (pyda: 2,6‐pyridinediamine; phendcH2: 1,10‐phenanthroline‐2,9‐dicarboxylic acid) and thoroughly characterized by elemental analysis, IR spectroscopy, X‐ray crystallography and cyclic voltammetry. The complex crystallizes in the monoclinic space group P21/n with four formula units in the unit cell. The unit cell dimensions are: a = 13.962(3) Å, b = 14.529(3) Å, c = 16.381(3) Å and β = 106.691(4)°. In this complex, 1,10‐phenanthroline‐2,9‐dicarboxylate acts as a tridentate ligand and the lattice is composed of anionic hexacoordinated complex, [Cr(phendc)2]?, 2,6‐pyridiniumdiamine counter ion, (pydaH)+, and five lattice water molecules. Crystallographic characterization revealed that the resulting supramolecular structure is strongly stabilized by complicated network of hydrogen bonds between the crystallization water molecules, counter ion and both coordinated and uncoordinated carboxylate groups. There is no relevant π‐π interaction for this anionic complex between pyda or phendc moieties. The electrochemical studies indicated over potential for both the cathodic and anodic peaks of the complex with respect to the free Cr3+ ion, as a consequence of the energy requirement for rearrangement of the ligand at electrode surface.  相似文献   

8.
Enantiomerization of octahedral tris(α‐diimine)–transition metal complexes was investigated by enantioselective dynamic MEKC. Varying both the transition metal ion (Fe2+, Fe3+, and Ni2+) and the bidentate diimine ligand (1,10‐phenanthroline and 2,2′‐bipyridyl), the enantiomer separations were performed either in a 100 mM sodium tetraborate buffer (pH 9.3) or in a 100 mM sodium tetraborate/sodium dihydrogenphosphate buffer (pH 8.0) both containing sodium cholate as chiral surfactant. The unified equation of dynamic chromatography was employed to determine apparent reaction rate constants from the electropherograms showing distinct plateau formation. Apparent activation parameters ΔH? and ΔS? were calculated from temperature‐dependent measurements between 10.0 and 35.0°C in 2.5 K steps. It was found that the nature of the central metal ion and the ligand strongly influence the enantiomerization barrier. Surprisingly, complexes containing the 2,2′‐bipyridyl ligand show highly negative activation entropies between ?103 and ?116 J (K mol)?1 while the activation entropy of tris(1,10‐phenanthroline) complexes is positive indicating a different mechanism of interconversion. Furthermore, it was found that the Ni2+ complexes are stereostable under the conditions investigated here making them a lucent target as enantioselective catalysts.  相似文献   

9.
The use of selective metal chelating agents with preference for binding of a specific metal ion to investigate its biological role is becoming increasingly common. We found recently that a well‐known copper‐specific chelator 2,9‐dimethyl‐1,10‐phenanthroline (2,9‐Me2OP) could completely inhibit the synergistic toxicity induced by tetrachlorocatechol (TCC) and sodium azide (NaN3). However, its underlying molecular mechanism is still not clear. Here, we show that the protection by 2,9‐Me2OP is not due to its classic copper‐chelating property, but rather due to formation of a multiple hydrogen‐bonded complex between 2,9‐Me2OP and TCC, featuring an unusual perpendicular arrangement of the two binding partners. The two methyl groups at the 2,9 positions in 2,9‐Me2OP were found to be critical to stabilize the 2,9‐Me2OP/TCC complex due to steric hindrance, and therefore completely prevents the generation of the reactive and toxic semiquinone radicals by TCC/NaN3. This represents the first report showing that an unexpected new protective mode of action for the copper “specific” chelating agent 2,9‐Me2OP by using its steric hindrance effect of the two CH3 groups not only to chelate copper, but also to “chelate” a catechol through multiple H‐bonding. These findings may have broad biological implications for future research of this widely used copper‐chelating agent and the ubiquitous catecholic compounds.  相似文献   

10.
《Electroanalysis》2006,18(24):2475-2480
The thin films containing transition metal complex tris(1,10‐phenanthroline) ruthenium(II) Ru(phen)3Cl2 (abbr Ru(phen)3, phen=1,10‐phenanthroline), and 12‐molybdophosphoric acid [PMo12O40]3? (abbr PMo12) were fabricated on quartz, silicon and ITO substrates by layer‐by‐layer (LBL) method. The LBL films were characterized by the UV‐vis spectroscopy, X‐ray photoelectron spectroscopy, atomic force microscopy and cyclic voltammetry. The films can catalyze both the reduction of ClO , BrO , IO , and the oxidation of C2O due to the presence of bifunctional composite, and the redox potentials depend on pH as a result of protonation. The photoluminescence of films were also investigated. The films exhibited photoluminescence arising from π*–t2g ligand‐to‐metal transition of Ru(phen)3.  相似文献   

11.
The mononuclear structure of Cd(S2CNEt2)2(2,9‐Me2‐1,10‐phenanthroline) shows symmetric coordination of the dithiocarbamate ligands and a distorted octahedral geometry for cadmium, defined by an N2S4 donor set, results. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

12.
The decarboxylation of 2‐cyano‐2‐phenylpropanoic acid is fast and quantitative when carried out in the presence of 1 molar equivalent of a [2]catenane composed of two identical macrocycles incorporating a 1,10‐phenanthroline unit in their backbone. When decarboxylation is over, all of the catenane molecules have experienced large‐amplitude motions from neutral to protonated catenane, and back again to the neutral form, so that they are ready to perform another cycle. This study provides the first example of the cyclic operation of a molecular switch at the sole expenses of the energy supplied by the substrate undergoing chemical transformation, without recourse to additional stimuli.  相似文献   

13.
The synthesis is described of a set of new N‐heterocyclic compounds which are derivatives of 1,10‐phenanthroline. The compounds are designed to be general purpose chelating agents which could function as tri‐, tetra‐ or hexadentate ligands with transition metal ions. Fused‐ring molecular components have been included in the design of the compounds so that they may function as binding agents to DNA through intercalation. This includes the synthesis of substituted derivatives of pyrazino[2,3‐f][1,10]phenanthroline and dipyrido[3,2‐a:2′3′‐c]phenazine.  相似文献   

14.
A set of cyclo‐7,10‐tris‐triphenylenyl macrocycles have been prepared by a Yamamoto cyclotrimerization protocol. In these novel macrocycles, three triphenylene units are covalently linked to each other, resulting in the formation of triangular‐shaped molecules. The fully planar derivative revealed pronounced self‐assembly behavior. NMR spectroscopy was used to determine the association constant in solution. 2D wide‐angle X‐ray scattering was applied to the study of the liquid crystallinity of this new discotic mesogen in the bulk state. Furthermore, nonplanar, laterally substituted derivatives were successfully tested as blue emitters in organic light‐emitting diodes owing to their unique optoelectronic properties and their high stability. In this case, substitution with sterically demanding phenyl groups was efficiently used to suppress intermolecular packing, thus preventing undesired quenching effects.  相似文献   

15.
1,10‐Phenanthroline reacts with malonitrile and aldehydes in the presence of isocyanides as domino‐Knoevenagel‐nucleophilic cycloaddition for generation of a new class of 10‐(aryl)‐11‐(alkyl‐ or arylamino‐)pyrrolo[1,2‐a][1,10]phenanthroline‐9‐carbonitrile compounds in excellent yield. All compounds are fully characterized with one structurally authenticated by a single X‐ray diffraction study.  相似文献   

16.
The title compounds, tris(1,10‐phenanthroline‐κ2N,N′)iron(II) bis(2,4,5‐tricarboxybenzoate) monohydrate, [Fe(C12H8N2)3](C10H5O8)2·H2O, (I), and tris(2,2′‐bipyridine‐κ2N,N′)iron(II) 2,5‐dicarboxybenzene‐1,4‐dicarboxylate–benzene‐1,2,4,5‐tetracarboxylic acid–water (1/1/2), [Fe(C10H8N2)3](C10H4O8)·C10H6O8·2H2O, (II), were obtained during an attempt to synthesize a mixed‐ligand complex of FeII with an N‐containing ligand and benzene‐1,2,4,5‐tetracarboxylic acid via a solvothermal reaction. In both mononuclear complexes, each FeII metal ion is six‐coordinated in a distorted octahedral manner by six N atoms from three chelating 1,10‐phenanthroline or 2,2′‐bipyridine ligands. In compound (I), the FeII atom lies on a twofold axis in the space group C2/c, whereas (II) crystallizes in the space group P21/n. In both compounds, the uncoordinated carboxylate anions and water molecules are linked by typical O—H...O hydrogen bonds, generating extensive three‐dimensional hydrogen‐bond networks which surround the cations.  相似文献   

17.
The structure of bis(1,10‐phenanthroline‐κ2N,N′)(thio­sulfato‐κ2O:S)­manganese(II) methanol solvate, [Mn(S2O3)(C12H8N2)2]·CH3OH, is made up of Mn2+ centers coordinated to two bidentate phenanthroline (phen) groups and an S,O‐chelating thio­sulfate anion, forming monomeric entities. The structure of catena‐poly­[[di­aqua(2,9‐di­methyl‐1,10‐phen­anthro­line‐κ2N,N′)­manganese(II)]‐μ‐thio­sulfato‐κ2O:S], [Mn(S2O3)(C14H12N2)(H2O)2]n, is polymeric, consisting of Mn(dmph)(H2O)2 units (dmph is 2,9‐di­methyl‐1,10‐phenanthroline) linked by thio­sulfate anions acting in an S,O‐chelating manner.  相似文献   

18.
Two large rings, 66‐ (m‐66 ) and 78‐membered ( m‐78 ) rings, each one incorporating two pairs of transition‐metal‐complexing units, have been prepared. The coordinating fragments are alternating bi‐ and tridentate chelating groups, namely, 2,9‐diphenyl‐1,10‐phenanthroline (dpp) and 2,2′,2′,6′′‐terpyridine (terpy) respectively. Both macrocycles form molecular figures‐of‐eight in the presence of FeII, affording a classical bis‐terpy complex as the central core. The larger m‐78 ring can accommodate a four‐coordinate CuI center with the formation of a {Cu(dpp)2}+ central complex and a highly twisted figure‐of‐eight backbone, whereas m‐66 is too small to coordinate CuI. Macrocycle m‐78 thus affords stable complexes with both FeII and CuI; the ligand around the metal changes from (terpy)2 to (dpp)2. This bimodal coordination situation allows for a large amplitude rearrangement of the organic backbone. When coordinated to preferentially octahedrally coordinated FeII or CuII, the height of the molecule along the coordinating axis of the tridentate terpy ligands is only about 11 Å, whereas the height of the molecule along the same vertical axis is several times as large for the tetrahedral CuI complex. Chemically or electrochemically driven contraction and extension motions along a defined axis make this figure‐of‐eight particularly promising as a new class of molecular machine prototype for use as a constitutive element in muscle‐like dynamic systems.  相似文献   

19.
Two new 1,10‐phenanthroline (Phen) containing conjugated oligomers, oligo‐3,8‐bis(4‐octylthiophene‐2‐yl)‐1,10‐phenanthroline) (PDTPh) and oligo‐3,8‐bis‐(4‐octyl‐5‐(4‐ctylthiophene‐2‐yl)thiophene‐2‐yl)‐1,10‐phenanthroline) (PTTPh), as well as their corresponding monomers (OTPhOT and OTOTPhOTOT) were prepared and their metal ion‐sensing properties were investigated. The oligomers showed high thermal stability, good proccessibility, and gave varied color when reacted with different metal ions. Oligomers also showed distinct responses toward metal ions when compared with their corresponding monomers, suggesting that the ionochromic responses were determined by not only the coordinating ability of Phen unit but also the conformation of oligomer chains. Moreover, the differences in the ion‐sensing behaviors between OTPhOT and OTOTPhOTOT also suggested that the coordination ability of Phen depends on its substituents. The oligo‐alkylthiophene moieties in PDTPh and PTTPh acted as spacers to reorganize the conformation of the oligomer chains, as well as the electron donating groups to adjust the coordination ability of the Phen. These findings provide a clue for designing Phen‐containing ion‐sensors for specified ion‐sensing applications. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 1586–1597, 2008  相似文献   

20.
Organic semiconductors containing metal binding sites within their molecular backbones are of a general interest in organic materials chemistry. In this paper, we describe a straightforward synthetic procedure, which gives access to a series of 2-(oligothienyl)-[1,10]phenanthrolines (nT-phen), 2,9-bis(oligothienyl)-[1,10]phenanthrolines (nT-phen-nT) and 2,2'-(oligothienyl)bis-[1,10]phenanthrolines (phen-nT-phen). By a Negishi-type cross coupling of 2-iodo-[1,10]phenanthroline or 2,9-diiodo-[1,10]phenanthroline with in situ generated alpha-zinc derivatives of different mono-, ter-, and quinquethiophenes we were able to synthesize the corresponding oligothienyl-phenanthrolines in medium to excellent yields. Furthermore, characterization of the optical properties of the new materials indicated that the two subunits, oligothiophene and phenanthroline, are in pi-conjugation. Characterization of the redox properties revealed additional evidence for the role of [1,10]phenanthroline as a pi-bridging unit in the nT-phen-nT series.  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