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 15 毫秒
1.
Amino‐ and halofunctional Siloxititanes Amino‐di‐tert‐butylsilanol reacts with tetrabutoxititane in a molar ratio of 2:1 to give di‐n‐butoxi(bis(di‐tert‐butyl‐n‐butoxi)siloxi)titane, (C4H9OSi(CMe3)2‐O)2Ti(OC4H9)2 ( 1 ), and lithium‐di‐tert‐butylchlorosilanolate in a molar ratio of 3:1 to give n‐butoxi(tris(di‐tert‐butyl‐n‐butoxi)siloxi)titane, (H9C4OSi(CMe3)2‐O)3TiOC4H9 ( 2 ). The amino‐di‐tert‐butylsilanol substitutes the four chloroatoms of TiCl4 in the presence of triethylamine as HCl‐acceptor. The tetrakis(amino‐di‐tert‐butyl)siloxititane ( 3 ) is formed. The lithium salt of di‐tert‐butylfluorosilanol reacts with TiCl4 in a molar ratio of 2:1 to give 1, 1, 3, 3‐tetra‐tert‐butyl‐1‐fluoro‐3‐trichlorotitoxi‐1, 3‐disiloxane, FSi(CMe3)2‐O‐Si(CMe3)2‐O‐TiCl3 ( 4 ). In the reaction of di‐tert‐butyl‐chlorosilanol and TiCl4, the anion [chlorosiloxi‐octa(tri‐μ2‐chlorotitanate)] ( 5 ) with protonated diethylether as counterion is obtained by using diethylether as HCl‐acceptor. The crystal structure determinations of 3 and 5 are reported.  相似文献   

2.
Diphosphapodands, [12]‐, [15]‐, and [18]Diphosphacoronands, Diphosphacryptand‐8, and Alkali‐Metal Complexes The cyclizing bis‐phosphonium‐salt formation of the open‐chain bis‐phosphine 17a (1,1,7,7‐tetrabenzyl〈P.O.P‐podand‐7〉) with diethylene glycol derived dibromide 13a yields the 12‐membered cyclic bis‐phosphonium salt 20 (4,4,10,10‐tetrabenzyl‐12〈O.P.O.P‐coronand‐4〉‐4,10‐diium dibromide) in yields as high as 50–60%. The 1,1,10,10‐tetrabenzyl〈P.O2.P‐podand‐10〉 17b forms with 13a the 15‐membered cyclic bis‐phosphonium salt 21 (7,7,13,13‐tetrabenzyl‐15〈O2.P.O.P‐coronand‐5〉‐7,13‐diium dibromide) with the same high yield. By quaternization of the bis‐phosphine 17b with triethylene glycol derived dibromide 13b , the 18‐membered 7,7,16,16‐tetrabenzyl‐18〈O2.P.O2.P‐coronand‐6〉‐7,16‐diium dibromide 24 is obtained in 50% yield, too. The Wittig reaction of the cyclic phosphonium salts with benzaldehyde yields the 12‐, 15‐, and 18‐membered cyclic bis‐benzylphosphine dioxides 9, 10 , and 11 as cis‐ and trans‐isomers beside trans‐stilbene. The 7,13‐dioxido‐7,13‐dibenzyl‐15〈O2.P.O2.P‐coronand‐5〉 10 forms a crystalline 1 : 1 Na‐complex 23 , which exists as a dimer. The structure of 23 was established by an X‐ray analysis and spectroscopic data. The 7,16‐dibenzyl‐18〈O2.P.O2.P‐coronand‐6〉 28 that is available by reduction of 11 with CeCl3/LiAlH4 reacts with triethylene glycol derived dibromide 13b under Ruggly Ziegler‐dilution conditions to give the bicyclic bis‐phosphonium salt 29 (1,10‐dibenzyl〈P[O2]3.P‐cryptand‐8〉‐1,10‐diium dibromide) in 18% yield. Again, by the Wittig procedure with benzaldehyde, the 7,16‐dioxido〈P[O2]3P‐cryptand‐8〉 12 is obtained as the first diphosphacryptand. The FD‐MS (CH2Cl2) of the cyclic bis‐phosphine dioxides 10 – 12 show that they exist as [2M+Na]+ complexes. The complex formation constants Ka of 9 – 11 with alkali‐metal cations are studied and compared with the complex formation of corresponding crown ethers.  相似文献   

3.
The tridentate organic ligand 4,4′,4′′‐(4,4,8,8,12,12‐hexamethyl‐8,12‐dihydro‐4H‐benzo[9,1]quinolizino[3,4,5,6,7‐defg]acridine‐2,6,10‐triyl)tribenzoic acid ( H3L ) has been synthesized (as the methanol 1.25‐solvate, C48H39NO6·1.25CH3OH). As a donor–acceptor motif molecule, H3L possess strong intramolecular charge transfer (ICT) fluorescence. Through hydrogen bonds, H3L molecules construct a two‐dimensional (2D) network, which pack together into three‐dimensional (3D) networks with an ABC stacking pattern in the crystalline state. Based on H3L and M(NO3)2 salts (M = Cd and Zn) under solvothermal conditions, two metal–organic frameworks (MOFs), namely, catena‐poly[[triaquacadmium(II)]‐μ‐10‐(4‐carboxyphenyl)‐4,4′‐(4,4,8,8,12,12‐hexamethyl‐8,12‐dihydro‐4H‐benzo[9,1]quinolizino[3,4,5,6,7‐defg]acridine‐2,6‐diyl)dibenzoato], [Cd(C48H37NO6)(H2O)3]n, I , and poly[[μ3‐4,4′,4′′‐(4,4,8,8,12,12‐hexamethyl‐8,12‐dihydro‐4H‐benzo[9,1]quinolizino[3,4,5,6,7‐defg]acridine‐2,6,10‐triyl)tribenzoato](μ3‐hydroxido)zinc(II)], [Zn2(C48H36NO6)(OH)]n, II , were synthesized. Single‐crystal analysis revealed that both MOFs adopt a 3D structure. In I , partly deprotonated HL 2? behaves as a bidentate ligand to link a CdII ion to form a one‐dimensional chain. In the solid state of I , the existence of weak interactions, such as O—H…O hydrogen bonds and π–π interactions, plays an essential role in aligning 2D nets and 3D networks with AB packing patterns for I . The deprotonated ligand L 3? in II is utilized as a tridentate building block to bind ZnII ions to construct 3D networks, where unusual Zn4O14 clusters act as connection nodes. As a donor–acceptor molecule, H3L exhibits fluorescence with a photoluminescence quantum yield (PLQY) of 70% in the solid state. In comparison, the PL of both MOFs is red‐shifted with even higher PLQYs of 79 and 85% for I and II , respectively.  相似文献   

4.
Lanthanum fluoborate modified by oleylamine [denoted as La(BF4)3‐OA] was synthesized as a potential lubricant additive by direct precipitation method with sodium tetrafluoroborate and lanthanum nitrate [La(NO3)3] as the staring materials and oleylamine (OA) as the surface‐modifying agent in distilled water‐ethanol mixed solvent. The effects of reaction temperature, OA to La(NO3)3 ratio, and surfactant cetyltrimethyl ammonium bromide on the size and shape of as‐synthesized La(BF4)3‐OA were investigated. The crystalline structure and morphology of as‐obtained La(BF4)3‐OA were characterized by X‐ray diffraction, transmission electron microscopy, Fourier transform infrared spectrometry, and X‐ray photoelectron spectroscopy. Moreover, the tribological properties of La(BF4)3‐OA as an additive in dioctyl sebacate, a synthetic ester, were evaluated with a four‐ball machine, and the worn surfaces of the steel balls were analyzed with a field emission scanning electron microscope equipped with an energy dispersive spectrometer accessory. It was found that the as‐synthesized La(BF4)3‐OA exhibits disk‐like nanoflake shape and have a diameter of 10–45 nm, depending on varying synthetic conditions. As‐synthesized La(BF4)3‐OA as an additive in DIOS possesses excellent antiwear and friction‐reduction performance for the steel–steel pair, which is because the as‐synthesized additive simultaneously contain tribologically active elements La, B, and F that facilitate the formation of a boundary lubricating and protecting film on sliding steel surfaces thereby avoiding direct contact of the steel–steel pair and significantly reducing the friction and wear. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

5.
Four crystal structures of 3‐cyano‐6‐hydroxy‐4‐methyl‐2‐pyridone (CMP), viz. the dimethyl sulfoxide monosolvate, C7H6N2O2·C2H6OS, (1), the N,N‐dimethylacetamide monosolvate, C7H6N2O2·C4H9NO, (2), a cocrystal with 2‐amino‐4‐dimethylamino‐6‐methylpyrimidine (as the salt 2‐amino‐4‐dimethylamino‐6‐methylpyrimidin‐1‐ium 5‐cyano‐4‐methyl‐6‐oxo‐1,6‐dihydropyridin‐2‐olate), C7H13N4+·C7H5N2O2, (3), and a cocrystal with N,N‐dimethylacetamide and 4,6‐diamino‐2‐dimethylamino‐1,3,5‐triazine [as the solvated salt 2,6‐diamino‐4‐dimethylamino‐1,3,5‐triazin‐1‐ium 5‐cyano‐4‐methyl‐6‐oxo‐1,6‐dihydropyridin‐2‐olate–N,N‐dimethylacetamide (1/1)], C5H11N6+·C7H5N2O2·C4H9NO, (4), are reported. Solvates (1) and (2) both contain the hydroxy group in a para position with respect to the cyano group of CMP, acting as a hydrogen‐bond donor and leading to rather similar packing motifs. In cocrystals (3) and (4), hydrolysis of the solvent molecules occurs and an in situ nucleophilic aromatic substitution of a Cl atom with a dimethylamino group has taken place. Within all four structures, an R22(8) N—H...O hydrogen‐bonding pattern is observed, connecting the CMP molecules, but the pattern differs depending on which O atom participates in the motif, either the ortho or para O atom with respect to the cyano group. Solvents and coformers are attached to these arrangements via single‐point O—H...O interactions in (1) and (2) or by additional R44(16) hydrogen‐bonding patterns in (3) and (4). Since the in situ nucleophilic aromatic substitution of the coformers occurs, the possible Watson–Crick C–G base‐pair‐like arrangement is inhibited, yet the cyano group of the CMP molecules participates in hydrogen bonds with their coformers, influencing the crystal packing to form chains.  相似文献   

6.
This article discusses a new borane chain transfer reaction in olefin polymerization that uses trialkylboranes as a chain transfer agent and thus can be realized in conventional single site polymerization processes under mild conditions. Commercially available triethylborane (TEB) and synthesized methyl‐B‐9‐borabicyclononane (Me‐B‐9‐BBN) were engaged in metallocene/MAO [depleted of trimethylaluminum (TMA)]‐catalyzed ethylene (Cp2ZrCl2 and rac‐Me2Si(2‐Me‐4‐Ph)2ZrCl2 as a catalyst) and styrene (Cp*Ti(OMe)3 as catalyst) polymerizations. The two trialkylboranes were found—in most cases—able to initiate an effective chain transfer reaction, which resulted in hydroxyl (OH)‐terminated PE and s‐PS polymers after an oxidative workup process, suggesting the formation of the B‐polymer bond at the polymer chain end. However, chain transfer efficiencies were influenced substantially by the steric hindrances of both the substituent on the trialkylborane and that on the catalyst ligand. TEB was more effective than TMA in ethylene polymerization with Cp2ZrCl2/MAO, whereas it became less effective when the catalyst changed to rac‐Me2Si(2‐Me‐4‐Ph)2ZrCl2. Both TEB and Me‐B‐9‐BBN caused an efficient chain transfer in the Cp2ZrCl2/MAO‐catalyzed ethylene polymerization; nevertheless, Me‐B‐9‐BBN failed in vain with rac‐Me2Si(2‐Me‐4‐Ph)2ZrCl2/MAO. In the case of styrene polymerization with Cp*Ti(OMe)3/MAO, thanks to the large steric openness of the catalyst, TEB exhibited a high efficiency of chain transfer. Overall, trialkylboranes as chain transfer agents perform as well as B? H‐bearing borane derivatives, and are additionally advantaged by a much milder reaction condition, which further boosts their applicability in the preparation of borane‐terminated polyolefins. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 3534–3541, 2010  相似文献   

7.
Addition reactions of thioamide dianions that were derived from N‐arylmethyl thioamides to imines and aziridines were carried out. The reactions of imines gave the addition products of N‐thioacyl‐1,2‐diamines in a highly diastereoselective manner in good‐to‐excellent yields. The diastereomeric purity of these N‐thioacyl‐1,2‐diamines could be enriched by simple recrystallization. The reduction of N‐thioacyl‐1,2‐diamines with LiAlH4 gave their corresponding 1,2‐diamines in moderate‐to‐good yields with retention of their stereochemistry. The oxidative‐desulfurization/cyclization of an N‐thioacyl‐1,2‐diamine in CuCl2/O2 and I2/pyridine systems gave the cyclized product in moderate yield and the trans isomer was obtained as the sole product. On the other hand, a similar cyclization reaction with antiformin (aq. NaClO) as an oxidant gave the cis isomer as the major product. The reactions of N‐tosylaziridines gave the addition products of N‐thioacyl‐1,3‐diamines with low diastereoselectivity but high regioselectivity and in good‐to‐excellent yields. The use of AlMe3 as an additive improved the efficiency and regioselectivity of the reaction. The stereochemistry of the obtained products was determined by X‐ray diffraction.  相似文献   

8.
Non‐metal nitrides such as BN, Si3N4, and P3N5 meet numerous demands on high‐performance materials, and their high‐pressure polymorphs exhibit outstanding mechanical properties. Herein, we present the silicon phosphorus nitride imide SiP2N4NH featuring sixfold coordinated Si. Using the multi‐anvil technique, SiP2N4NH was obtained by high‐pressure high‐temperature synthesis at 8 GPa and 1100 °C with in situ formed HCl acting as a mineralizer. Its structure was elucidated by a combination of single‐crystal X‐ray diffraction and solid‐state NMR measurements. Moreover, SiP2N4NH was characterized by energy‐dispersive X‐ray spectroscopy and (temperature‐dependent) powder X‐ray diffraction. The highly condensed Si/P/N framework features PN4 tetrahedra as well as the rare motif of SiN6 octahedra, and is discussed in the context of ambient‐pressure motifs competing with close‐packing of nitride anions, representing a missing link in the high‐pressure chemistry of non‐metal nitrides.  相似文献   

9.
The direct depolymerization of SiO2 to distillable alkoxysilanes has been explored repeatedly without success for 85 years as an alternative to carbothermal reduction (1900 °C) to Simet, followed by treatment with ROH. We report herein the base‐catalyzed depolymerization of SiO2 with diols to form distillable spirocyclic alkoxysilanes and Si(OEt)4. Thus, 2‐methyl‐2,4‐pentanediol, 2,2,4‐trimethyl‐1,3‐pentanediol, or ethylene glycol (EGH2) react with silica sources, such as rice hull ash, in the presence of NaOH (10 %) to form H2O and distillable spirocyclic alkoxysilanes [bis(2‐methyl‐2,4‐pentanediolato) silicate, bis(2,2,4‐trimethyl‐1,3‐pentanediolato) silicate or Si(eg)2 polymer with 5–98 % conversion, as governed by surface area/crystallinity. Si(eg)2 or bis(2‐methyl‐2,4‐pentanediolato) silicate reacted with EtOH and catalytic acid to give Si(OEt)4 in 60 % yield, thus providing inexpensive routes to high‐purity precipitated or fumed silica and compounds with single Si−C bonds.  相似文献   

10.
The direct depolymerization of SiO2 to distillable alkoxysilanes has been explored repeatedly without success for 85 years as an alternative to carbothermal reduction (1900 °C) to Simet, followed by treatment with ROH. We report herein the base‐catalyzed depolymerization of SiO2 with diols to form distillable spirocyclic alkoxysilanes and Si(OEt)4. Thus, 2‐methyl‐2,4‐pentanediol, 2,2,4‐trimethyl‐1,3‐pentanediol, or ethylene glycol (EGH2) react with silica sources, such as rice hull ash, in the presence of NaOH (10 %) to form H2O and distillable spirocyclic alkoxysilanes [bis(2‐methyl‐2,4‐pentanediolato) silicate, bis(2,2,4‐trimethyl‐1,3‐pentanediolato) silicate or Si(eg)2 polymer with 5–98 % conversion, as governed by surface area/crystallinity. Si(eg)2 or bis(2‐methyl‐2,4‐pentanediolato) silicate reacted with EtOH and catalytic acid to give Si(OEt)4 in 60 % yield, thus providing inexpensive routes to high‐purity precipitated or fumed silica and compounds with single Si?C bonds.  相似文献   

11.
Supramolecular ensembles adopting ring‐in‐ring structures are less developed compared with catenanes featuring interlocked rings. While catenanes with inter‐ring closed‐shell metallophilic interactions, such as d10–d10 AuI–AuI interactions, have been well‐documented, the ring‐in‐ring complexes featuring such metallophilic interactions remain underdeveloped. Herein is described an unprecedented ring‐in‐ring structure of a AuI‐thiolate Au12 cluster formed by recrystallization of a AuI‐thiolate Au10 [2]catenane from alkane solvents such as hexane, with use of a bulky dibutylfluorene‐2‐thiolate ligand. The ring‐in‐ring AuI‐thiolate Au12 cluster features inter‐ring AuI–AuI interactions and underwent cluster core change to form the thermodynamically more stable Au10 [2]catenane structure upon dissolving in, or recrystallization from, other solvents such as CH2Cl2, CHCl3, and CH2Cl2/MeCN. The cluster‐to‐cluster transformation process was monitored by 1H NMR and ESI‐MS measurements. Density functional theory (DFT) calculations were performed to provide insight into the mechanism of the “ring‐in‐ring? [2]catenane” interconversions.  相似文献   

12.
Lithium 8‐amidoquinoline ( 1 ) and lithium 8‐(trialkylsilylamido)quinoline [SiMe2tBu ( 2 ), SiiPr3 ( 3 )] react with dimethylgallium chloride to the metathesis products dimethylgallium 8‐amidoquinoline ( 4 ) as well as dimethylgallium 8‐(trialkylsilylamido)quinoline [SiMe2tBu ( 5 ), SiiPr3 ( 6 )]. The gallium atoms are in distorted tetrahedral environments. During the synthesis of 5 , orange dimethylgallium 2‐butyl‐8‐(tert‐butyldimethylsilylamido)quinoline ( 7 ) was found as by‐product. The metathesis reactions of Me2GaCl with LiN(R)CH2Py (Py = 2‐pyridyl) yield the corresponding 2‐pyridylmethylamides Me2Ga‐N(H)CH2Py ( 8 ), Me2Ga‐N(SiMe2tBu)CH2Py ( 9 ) and Me2Ga‐N(SiiPr3)CH2Py ( 10 ). In these complexes the gallium atoms show a distorted tetrahedral coordination sphere. However, derivative 8 crystallizes dimeric with bridging amido units whereas in 9 and 10 the 2‐pyridylmethylamido moieties act as bidentate ligands leading to monomeric molecules.  相似文献   

13.
4,4′‐{9,9′‐Spirobi[10H‐acridine]‐10,10′‐diyl}dibenzoic acid ( L , C29H26N2O4) was designed and synthesized as a new donor–acceptor motif molecule. Due to the large dihedral angle between the planes of the carboxyphenyl group and the spiroacridine moiety, L possess thermally activated delayed fluorescence (TADF). By applying L as a ligand and using Cd as a metal connector, we synthesized the coordination polymer catena‐poly[hemi‐μ‐aqua‐aqua(μ3‐4,4′‐{9,9′‐spirobi[10H‐acridine]‐10,10′‐diyl}dibenzoato)cadmium(II)], [Cd(C29H24N2O4)(H2O)1.5]n, ( I ). X‐ray crystallographic analysis revealed that this coordination polymer exhibits one‐dimensional chains constructed from molecular twist‐ring moieties, with Cd2O11 clusters as the connection nodes. The stacking pattern of the two‐dimensional network was formed by C—H…π interactions in the solid state. Similar to L , ( I ) presents a sky‐blue TADF emission, together with a photoluminescence quantum yield (PLQY) of 40%. It is worth noting that the photocatalytic activity toward the generation of singlet oxygen of this coordination polymer is confirmed.  相似文献   

14.
Two alkylimido derivatives of hexamolybdate, (Bu4N)2[Mo6O18(≡N‐o‐COOCH3C6H4)] ( 1 ) and (Bu4N)2[Mo6O18(≡N‐o‐COOCH2CH3C6H4)] ( 2 ), were synthesized in high purity and good yields by the reaction of [(C4H9)4N]4[α‐Mo8O26] and methyl anthranilate or ethyl‐o‐aminobenzoate hydrochloride with N,N′‐dicyclohexylcarbodiimide (DCC) as a dehydrating agent in dry acetonitrile solution, which were characterized by elemental analyses, IR, UV/Vis, and 1H NMR spectroscopy as well as ESI‐MS, and single‐crystal X‐ray diffraction study. Compound 1 crystallizes in the monoclinic space group P21/n with one‐dimensional chain structure via intramolecular hydrogen bond. Compound 2 also crystallizes in the monoclinic space group P21/n with dimer structure by intramolecular hydrogen bonds and π–π interactions between the pairs of cluster anions.  相似文献   

15.
Despite being widely used as electron acceptor in polymer solar cells, commercially available PC71BM (phenyl‐C71‐butyric acid methyl ester) usually has a “random” composition of mixed regioisomers or stereoisomers. Here PC71BM has been isolated into three typical isomers, α‐, β1‐ and β2‐PC71BM, to establish the isomer‐dependent photovoltaic performance on changing the ternary composition of α‐, β1‐ and β2‐PC71BM. Mixing the isomers in a ratio of α/β12=8:1:1 resulted in the best power conversion efficiency (PCE) of 7.67 % for the polymer solar cells with PTB7:PC71BM as photoactive layer (PTB7=poly[[4,8‐bis[(2‐ethylhexyl)oxy]benzo[1,2‐b:4,5‐b′]dithiophene‐2,6‐diyl][3‐fluoro‐2‐[(2‐ethylhexyl)carbonyl]thieno[3,4‐b]thiophenediyl]]). The three typical PC71BM isomers, even though sharing similar LUMO energy levels and light absorption, render starkly different photovoltaic performances with average‐performing PCE of 1.28–7.44 % due to diverse self‐aggregation of individual or mixed PC71BM isomers in the otherwise same polymer solar cells.  相似文献   

16.
The ternary systems of C2H4 (C2H2 or C6H6)‐MCN‐HF (M=Cu, Ag, Au) and the respective binary systems were investigated to study the interplay between metal???π interactions and hydrogen bonds. The metal???π interactions in C2H4‐MCN become stronger with the irregular order Ag<Cu<Au, while the hydrogen bonds in MCN‐HF become weaker following the same order. The metal???π interactions are weakened as the H atoms in the π system are replaced with electron‐withdrawing groups and enhanced by electron‐donating groups. Type 1 of these ternary systems, in which MCN acts as Lewis base and acid simultaneously, is more stable than type 2, in which C2H4 acts as a double Lewis base. Negative cooperativity is present in type 2 ternary systems with a weakening of the metal???π interactions and the hydrogen bonds. Positive cooperativity is found in type 1 ternary systems with an enhancement of the metal???π interactions and the hydrogen bonds, except for C2(CN)4‐AuCN‐HF‐1. The weaker metal???π interaction in C6H6‐AuCN has a greater enhancing effect on the hydrogen bond in AuCN‐HF than those in C2H4‐AuCN and C2H2‐AuCN. These synergetic effects were analyzed with the natural bond orbital and energy decomposition.  相似文献   

17.
Geometrical structures of three investigated molecules Sc3N@C80, Sc3N@C80‐Fc, and C60‐Fc were optimized by density functional theory (DFT) at the B3LYP/6‐31G* level. Then the time‐dependent DFT was employed to investigate the excited states of these molecules. After exohedral functionalization by ferrocene (Fc‐) group as the electron donor or replacing C60 with Sc3N@C80 as the electron acceptor, the wavelengths of the first one‐photon absorption peak and the strongest two‐photon absorption (2PA) and three‐photon absorption (3PA) peaks shift red. The corresponding cross sections of Sc3N@C80‐Fc in the 2PA and 3PA processes increase as compared with those of Sc3N@C80, which originate from the contributions of charge transfers from Fc‐ group to C80 cage and simultaneously the transfers from the C80 cage to the encapsulated Sc3N cluster. When compared with C60‐Fc, the 2PA and 3PA cross sections of Sc3N@C80‐Fc decrease, which may result from the more negative charge surface of C80 cage in Sc3N@C80‐Fc molecule which blocks the charge transfers from Fc‐ moiety to the C80 cage in the excitation processes by compared with C60‐Fc. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012  相似文献   

18.
Aluminium dihydroxyterephthalate [Al8(OH)4(OCH3)8(BDC(OH)2)6] ? x H2O (denoted CAU‐1‐(OH)2) was synthesized under solvothermal conditions and characterized by X‐ray powder diffraction, IR spectroscopy, sorption measurements, as well as thermogravimetric and elemental analysis. CAU‐1‐(OH)2 is isoreticular to CAU‐1 and its pores are lined with OH groups. It is stable under ambient conditions and in water, and it exhibits permanent porosity and two types of cavities with effective diameters of approximately 1 and 0.45 nm. The crystallization of CAU‐1‐(OH)2 was studied by in situ energy‐dispersive X‐ray diffraction (EDXRD) experiments in the 120–145 °C temperature range. Two heating methods—conventional and microwave—were investigated. The latter leads to shorter induction periods as well as shorter reaction times. Whereas CAU‐1‐(OH)2 is formed at all investigated temperatures using conventional heating, it is only observed below 130 °C using microwave heating. The calculation of the activation energy of the crystallization of CAU‐1‐(OH)2 exhibits similar values for microwave and conventional synthesis.  相似文献   

19.
1‐Benzofuran‐2,3‐dicarboxylic acid (C10H6O5) is a dicarboxylic acid ligand which can readily engage in organometallic complexes with various metal ions. This ligand is characterized by an intramolecular hydrogen bond between the two carboxyl residues, and, as a monoanionic species, readily forms supramolecular adducts with different organic and inorganic cations. These are a 1:1 adduct with the dimethylammonium cation, namely dimethylammonium 3‐carboxy‐1‐benzofuran‐2‐carboxylate, C2H8N+·C10H5O5, (I), a 2:1 complex with Cu2+ ions in which four neutral imidazole molecules also coordinate the metal atom, namely bis(3‐carboxy‐1‐benzofuran‐2‐carboxylato‐κO3)tetrakis(1H‐imidazole‐κN3)copper(II), [Cu(C10H5O5)2(C3H4N2)4], (II), and a 4:1 adduct with [La(H2O)7]3+ ions, namely heptaaquabis(3‐carboxy‐1‐benzofuran‐2‐carboxylato‐κO3)lanthanum 3‐carboxy‐1‐benzofuran‐2‐carboxylate 1‐benzofuran‐2,3‐dicarboxylic acid solvate tetrahydrate, [La(C10H5O5)2(H2O)7](C10H5O5)·C10H6O5·4H2O, (III). In the crystal structure, complex (II) resides on inversion centres, while complex (III) resides on axes of twofold rotation. The crystal packing in all three structures reveals π–π stacking interactions between the planar aromatic benzofuran residues, as well as hydrogen bonding between the components. The significance of this study lies in the first crystallographic characterization of the title framework, which consistently exhibits the presence of an intramolecular hydrogen bond and a consequent monoanionic‐only nature. It shows further that the anion can coordinate readily to metal cations as a ligand, as well as acting as a monovalent counter‐ion. Finally, the aromaticity of the flat benzofuran residue provides an additional supramolecular synthon that directs and facilitates the crystal packing of compounds (I)–(III).  相似文献   

20.
For the purpose of investigating the coordination behavior of sterically congested alkenes and exploring the possibility of cofacial complexation in the polycyclic aromatic system for the formation of extended polymeric networks, a new tetradentate ligand, 1,1,2,2‐tetrakis[4‐(1H‐1,2,4‐triazol‐1‐yl)phenyl]ethylene (TTPE), has been designed and synthesized. By using TTPE as a building block with regard to the self‐assembly with MnCl2 ? 4 H2O, a novel two‐dimensional coordination framework {[Mn(TTPE)Cl2] ? 4 CHCl3}n ( 1 ) can be isolated. Anion‐exchange and organic‐group‐functionalized aromatic guest TTPE‐loaded host–guest complex experimental results indicate that coordinated Cl? anions in the 2D framework of 1 can be completely replaced with dissociative ClO4? groups in an irreversible single‐crystal‐to‐single‐crystal transformation fashion, as evidenced by the anion‐exchange products of {[Mn(TTPE)(H2O)2](ClO4)2 ? 0.5 TTPE ? 5.25 H2O}n ( 2 ). Interestingly, TTPE, acting as an organic template, was encapsulated in the confined space of the 2D grid of 2 . To the best of our knowledge, such large organic molecules encapsulated in the reactive organic‐group‐functionalized aromatic‐guest‐loaded host–guest complex are unprecedented up to now. Luminescence measurements illustrate that 1 and 2 represent novel examples of sensing materials based on triazole derivatives. Further, 2 has been demonstrated by tuning the fluorescence response of porous metal–organic frameworks as a function of adsorbed small analytes.  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